Skip to main content

Full text of "Inventory and monitoring of wildlife habitat"

See other formats


BLM    LIBRARY 


88048148 


^.---.   is     ■ ' 


'wJPJPP"*^!??*! 


,  I,:''    - 


:  ■&%&& \ 


mg&mMm 


■ 


I   WM 


■  *■ 


INVENarORY  fiy^D  MOMTORIflllf  i 


~;  PJ 


a.;      '  < 


.  $  I     _\> 


X 


T7/* < !  J^ 


h0      v^ 


^•."^"i 


iaffia&'J^  •- 


>:  S# 


¥*'£** 

•■/>" 


if/m 


Jf  s 


9TC& 


^S 


-  J/^r 


Iff 


**m£ 


i&M 


£&j£& 


^ 


!|  |^ 


•«'-V 


• 


sV-.Tn 


...>. 


.  j    •  *f/^tei' 


*****:* 


MK&S 


INVENTORY  AND  MONITORING 
OF  WILDLIFE  HABITAT 


This  report  should  be  cited  as: 

Cooperrider,  A.  Y.,  R.J.  Boyd,  and  H.  R.  Stuart,  eds.  1986.  Inventory  and  monitoring  of  wildlife  habitat.  U.S.  Dept.  Inter.. 
Bur.  Land  Manage.  Service  Center.  Denver,  Co.    xviii.  858  pp. 


BLM/YA/PT-87/001  +6600 


V 


Lead  Editor:  •*>•.'""" 

Janet  Poorman  j 

Copy  Editor:  '1 

Margaret  McGinnis 
Contract  Editor: 

William  Dryer 
Section  paintings,  chapter  and  cartoon  illustrations 

Peter  Doran  and  Shirley  McCulloch 
Technical  illustrations  and  cover: 

Peter  Doran 
Graphs: 

Geri  McBeth  and  Laura  Cuplin 
Computer  graphics 

Rose  Maruska 
Chapter  pasteups  and  chapter  illustrations: 

Laura  Cuplin 
Drawings  for  Amphibians  and  Reptiles  chapter: 

Lauren  Porzer  Keener 
Graphics  assistance: 

Keith  Francis  and  Herman  Weiss 
Editorial  assistance-  and  word  processing: 

Marilyn  Chatterton,  June  Johnson,  Geri  McBeth,  and  Margaret  Trujillo 
Computer  assistance 

Phyllis  Ellu >tt  and  Surma  Reitsma 


For  sale  by  the  Superintendent  of  Documents.  U.S.  Government  Printing  Office,  Washington,  D.C.  20402 


&  ■*  °  Q^ 


INVENTORY  AND  MONITORING 
OF  WILDLIFE  HABITAT 


Compiled  and  Edited  by 

Allen  Y.  Cooperrider,  Raymond  J.  Boyd, 
and  Hanson  R.  Stuart 

Design  by 

Shirley  L.  McCulloch 


U.S.  Department  of  the  Interior  •  Bureau  of  Land  Management  •  September  1986 


TABLE  OF  CONTENTS 


Preface  a 

Acknowledgments  xi 

Reviewers  xni 

Introduction  XVH 

I  PLANNING 

Chapter     1       Inventory  and  Monitoring  Process  1 

K  Bruce  Jones 

Chapter     2      Data  Types  1 1 

K  Bruce  Jones 

Chapter     3      Literature  Review  29 

Ora  Wagoner 

Chapter    4      Habitat  Mapping  49 

Richard  M.  Kerr 

II  MAJOR  HABITATS 

Chapter    5      Forests  73 

Jack  Ward  Thomas  and  Jared  Venter 

Chapter    6      Rangelands  93 

Henry  L.  Short 

Chapter     7      Deserts  123 

A]  Bruce  Jones 

Chapter    8      Tundra  149 

Peter  C.  Lent 

Chapter    9      Riparian  Habitats  169 

Robert  D.  Ohmart  and  Bertin  W.  Anderson 

Chapter  10      Marshes  201 

Milton  W.  Welter 

Chapter  1 1       Streams  225 

Paul  Cuplin 

Chapter  12      Lakes  237 

James  F.  LaBounty 

III      SPECIES  GROUPS 

Chapter  13      Fish  257 

Paul  Cuplin 

Chapter  14      Amphibians  and  Reptiles  267 

A.'  Bruce  Jones 


Chapter  15      Songbirds  291 

Ronald  A  Ryder 

Chapter  16      Raptors  31 3 

Michael  Kochert 

Chapter  17      Marsh  and  Shorebirds  35 1 

Peter  G.  Connors 

Chapter  18      Waterfowl  37 1 

Robert  L.  Eng 

Chapter  19      Colonial  Waterbirds  387 

Steven  M.  Speich 

Chapter  20      Upland  Game  Birds  407 

Robert  L.  Eng 

Chapter  21      Rodents  and  Insectivores  429 

Mayo  Call 

Chapter  22      Lagomorphs  453 

Joseph  A  Chapman  and  Gale  R  Willner 

Chapter  23      Carnivores  475 

Richard  A  Spowart  and  Fred  B.  Samson 

Chapter  24      Bats  497 

Stephen  P.  Cross 

Chapter  25      Ungulates  519 

Raymond  J.  Boyd,  Allen  Y.  Cooperrider, 
Peter  C  Lent,  and  James  A  Bailey 

IV  HABITAT  MEASUREMENTS 

Chapter  26      Soils  567 

James  E.  Stone 

Chapter  27      Terrestrial  Physical  Features  587 

Allen  Y.  Cooperrider 

Chapter  28      Aquatic  Physical  Features  603 

Paul  Cuplin 

Chapter  29      Hydrologic  Properties  613 

Bruce  Van  Haveren 

Chapter  30      Water  Quality  633 

Paul  Cuplin 

Chapter  31      Vegetation  639 

Bertin  W.  Anderson  and  Robert  D.  Ohmart 

Chapter  32      Macroinvertebrates  661 

Fred  Mangum 

V  SPECIAL  STUDIES 

Chapter  33      Radiotelemetry  679 

Paul  L.  Hegdal  and  Bruce  A  Colvin 


Chapter  34      Food  Habits  699 

Allen  Y.  Cooperrider 

Chapter  35      Weather  and  Climate  711 

James  A  Bailey 

VI      ANALYSIS  AND  PRESENTATION 

Chapter  36      Data  Management  727 

Larry  Peterson  and  Iris  Matney 

Chapter  37      Statistical  Analysis  741 

William  H.  West 

Chapter  38      Habitat  Evaluation  Systems  757 

Allen  Y.  Cooperrider 

Chapter  39      Evaluation  and  Interpretation 
Allen  Y  Cooperrider 

Chapter  40      Economic  Analysis  785 

John  Loom  is 

Chapter  41       Written  Communications  805 

Donald  E.  Zimmerman 

Chapter  42      Verbal  Presentations  829 

Eugene  Decker 

Glossary  841 

Index  853 


PREFACE 


This  book  is  intended  to  guide  field  biologists 
and  managers  in  planning,  organizing,  and  adminis- 
tering wildlife  inventory  and  monitoring  projects. 
Although  primarily  designed  for  the  professional 
wildlife  biologists  working  for  the  U.S.  Bureau  of 
Land  Management,  we  believe  much  of  the  material 
will  interest  those  working  in  other  agencies  and 
institutions  as  well. 

The  book  reviews  current  general  procedures 
and  specific  techniques.  It  is  not  a  "cookbook,"  how- 
ever, nor  an  attempt  to  standardize  techniques 
among  biologists.  We  believe  that  providing  a  set  of 
rigid  rules  for  inventorying  and  monitoring  wildlife 
habitat  would  be  presumptuous  since  nature  itself  is 
complex,  diverse,  and  dynamic.  We  firmly  believe 
the  professional  judgment  of  trained  biologists  is 
critical  in  land  management.  Their  expertise  cannot 
be  replaced  by  modern  technology.  Although  mod- 
ern technology  can  make  some  of  their  work  easier, 
only  the  biologist  can — 

•  identify  biological  problems; 

•  ensure  the  correct  problem  has  been  identified; 

•  collect  the  information  needed  to  address  a 
problem;  and 

•  analyze,  interpret,  evaluate,  and  effectively 
present  information  to  influence  management 
decisions. 

Management  biologists  have  demanding,  com- 
plex, and  challenging  tasks ;  few  have  the  luxury  of 
specializing.  Each  must  be  knowledgeable  of  many 
diverse  wildlife  groups,  other  natural  resources,  and 
resource  conflicts.  Management  biologists  also  need 
good  communication  skills  to  effectively  deal  with 
concerned  citizens,  ranchers,  miners,  administrators, 
and  scientists.  This  book  is  therefore  intended  to 
help  these  biologists  become  more  knowledgeable  of 
all  facets  of  wildlife  habitat  management. 

An  overriding  concern  in  designing  this  book 
was  to  cover  the  process  of  inventory  and  monitor- 
ing in  its  entirety,  i.e.,  from  initial  problem  identifica- 
tion through  presentation  of  results.  We  have 
observed  that  many  biologists  do  not  allocate  their 
time  and  energy  well  among  the  various  tasks  neces- 
sary to  complete  or  conduct  an  inventory  or  moni- 
toring project.  Good  data  are  often  collected  to 
solve  problems  that  are  trivial  compared  to  the  real 
priority  concerns.  Similarly,  data  are  sometimes  col- 
lected at  great  expense  but  are  not  analyzed,  or  the 


sampling  design  is  so  poor  that  the  data  cannot  be 
used.  Another  common  problem  is  that  good  data 
are  collected  and  analyzed,  but  the  results  are  never 
effectively  communicated  to  the  public  and  the  deci- 
sionmakers. Therefore,  the  entire  process  of  con- 
ducting an  inventory  or  monitoring  a  project  is 
presented.  Data  collection  and  analysis  are  obviously 
important  and  central  tasks  of  such  projects,  but 
they  tend  to  be  tasks  that  biologists  perform  well. 
Other  aspects  such  as  design,  planning,  and  commu- 
nication, are  included  to  help  biologists  improve 
their  skills  in  these  areas. 

The  book  is  modular;  chapters  can  be  read 
either  alone  or  in  combination  for  general  guidance 
or  as  a  reference  source.  As  a  reference  source,  a 
chapter  may  contain  a  detailed  description  of  a  tech- 
nique or  refer  the  reader  to  another  source  contain- 
ing a  description.  This  depends  on  the  amount  of 
work  available  on  the  subject,  the  detail  involved, 
and  the  availability  of  good  descriptions  in  readily 
accessible  publications.  For  instance,  field  techniques 
for  lizard  inventory  are  described  in  detail  since 
few  good  references  are  available  on  the  subject. 
However,  a  subject  such  as  mark-recapture  methods, 
to  which  biostatisticians  have  devoted  whole  careers 
and  about  which  numerous  books  have  been  written, 
has  brief  coverage  supplemented  by  relevant 
references. 

The  book  is  divided  into  six  major  sections. 
Section  I  covers  general  procedures  for  planning, 
designing,  and  organizing  wildlife  habitat  inventory 
and  monitoring  programs.  These  chapters  cover  such 
topics  as  determining  objectives;  design;  use  of  exist- 
ing literature;  habitat  mapping;  and  choice  of  areas, 
measurements,  and  techniques. 

Section  II  provides  guidelines  for  inventorying 
and  monitoring  particular  habitats.  Each  chapter 
includes  classification  systems,  species  groups  and 
habitat  features  that  are  particularly  important,  major 
impacts  on  such  habitats,  particular  problems  of 
inventory  or  monitoring,  and  habitat  analysis  sys- 
tems. The  chapters  in  this  section  are  designed  for 
biologists  who  are  planning  inventory  or  monitoring 
studies  in  habitats  they  have  not  worked  in 
previously. 

Section  III  provides  guidance  on  inventorying 
and  monitoring  habitat  for  particular  animal  groups. 
The  coverage  of  species  groups  is  not  organized 
along  taxonomic  lines,  but  rather  taxa  for  which  sim- 


ilar  techniques  are  used  are  lumped  together.  For 
each  species  group,  the  chapter  includes  a  descrip- 
tion of  ( 1 )  habitat  components,  e.g.,  vegetation 
structure,  physical  features,  that  correlate  with  spe- 
cies in  this  group  and  (2)  current  techniques  for 
population  inventories  and  monitoring. 

Section  IV  describes  techniques  for  measuring 
habitat  variables.  These  chapters  include  detailed 
descriptions  of  actual  measurement  techniques  and 
identify  advantages  and  limitations. 

Section  V  covers  special  monitoring  studies. 
These  include  methods  of  food  habits  determination, 
climatological  studies,  and  studies  of  movement  and 
habitat  use.  Such  information  is  generally  used  with 


other  habitat  or  population  measurements  to  deter- 
mine the  cause  of  observed  changes. 

The  final  section,  Section  VI,  covers  techniques 
and  procedures  for  analysis,  evaluation,  interpreta- 
tion, and  presentation  of  data  and  results. 

To  try  to  obtain  a  publication  that  is  as  readable 
and  accessible  as  possible,  we  have  at  times  edited 
or  rewritten  chapter  material.  In  some  cases,  in  the 
interest  of  time,  we  have  not  been  able  to  return 
materials  to  authors  for  review.  We  have  made  every 
effort  not  to  alter  meaning  or  emphasis;  nevertheless 
some  errors  may  have  occurred.  For  these  we  apolo- 
gize to  authors  and  readers  and  hope  that  such  cases 
are  few. 


ACKNOWLEDGMENTS 


A  work  such  as  this  could  not  be  completed 
without  the  assistance  and  support  of  many  individu- 
als too  numerous  to  mention.  In  addition  to  the  staff 
members  listed  and  the  authors  themselves,  other 
reviewers  have  contributed  substantially  to  the  pub- 
lication. We  sincerely  appreciate  their  efforts  to  help 
make  the  publication  more  complete,  accurate,  and 
readable.  They  are  listed  separately  in  the  following 
section. 

Certain  other  individuals  deserve  particular  rec- 
ognition. Mayo  Call,  formerly  of  the  U.S.  Bureau  of 
Land  Management  (BLM),  originally  recognized  the 
need  for  such  a  publication  and,  before  he  retired, 
convinced  BLM  management  to  support  this  effort. 
The  editors  left  with  the  task  have  alternately  cursed 
and  thanked  him  for  this  effort.  In  the  long  run,  he 
deserves  the  credit  for  the  idea,  but  in  the  short  run, 
we  have  often  wished  that  he  was  here  to  share  the 
frustrations. 

John  Crawford,  Chief  of  Wildlife  for  the  BLM 
from  1974  to  1985,  was  largely  responsible  for  pro- 
viding the  moral  and  financial  support  for  this  effort. 
During  his  tenure  as  Chief  of  Wildlife,  the  BLM  wild- 
life program  developed  from  infancy  to  a  substantial 
force  in  multiple-use  management.  John's  commit- 
ment and  effort  was  largely  responsible  for  this  de- 
velopment. He  recognized  the  need  for  sound 
inventory  and  monitoring  of  wildlife  habitat  on  pub- 
lic lands  and  for  providing  technical  support  to  field 
biologists  through  publications  of  this  sort. 


The  original  format  for  the  publication  was  de- 
signed with  the  assistance  of  Paul  Cuplin,  Kniffy 
Hamilton,  Bruce  Jones,  Richard  Kerr,  and  Chris 
Maser. 

The  continued  support  of  members  of  the  BLM 
wildlife  staff  in  Washington,  David  Almand,  Nancy 
Green,  Kniffy  Hamilton,  Alan  Kesterke,  Neal  Middle- 
brook,  and  William  Radtkey,  is  greatly  appreciated. 
The  ongoing  support  of  members  of  the  Denver 
Service  Center — Directors  Delmar  Vail,  Neil  Morck, 
and  Robert  Moore,  and  Chief  of  the  Office  of  Tech- 
nology Transfer,  Mary  Gaylord — is  appreciated.  The 
senior  editors  (Cooperrider  and  Boyd)  would  like  to 
express  particular  appreciation  to  their  supervisors, 
John  Baker  and  Allan  Strobel,  for  their  consistent 
support  and  encouragement  throughout  the  entire 
project,  and  also  extend  a  special  thank  you  to  Rob- 
ert Ader,  who  delayed  several  major  projects  so 
some  of  his  staff  could  help  with  this  book. 

Faye  Leonard  assisted  with  the  contracting  ef- 
forts competently  and  cheerfully.  Carl  Anderson 
provided  much  assistance  with  the  typesetting  and 
printing.  His  knowledge  of  the  technical  aspects 
of  printing  was  invaluable  and  is  much  appreciated. 

Finally,  the  editors  gratefully  acknowledge  the 
assistance  and  contributions  of  the  staff  of  the  BLM 
library,  particularly  Sandra  Bowers,  Teresa  Day,  and 
Judy  Moisey. 


xi 


Use  of  trade  or  firm  names  in  this  publication  is  for  the  information  and  convenience  of  the  reader.  Such  use  does  not 
constitute  an  official  endorsement  or  approval  by  the  U.S.  Department  of  the  Interior  of  any  product  or  service  to  the 
exclusion  of  others  which  may  be  suitable.  The  U.S.  Department  of  the  Interior  does  not  assume  any  liability  or  responsi- 
bility for  damages  caused  by  reliance  on  the  materials  presented  in  this  publication. 


All  scientific  and  common  names  of  animals  and  plants  used  in  this  publication  were  verified  according  to  the  following 
authorities: 

Plants — Checklist  of  North  American  Plants  for  Wildlife  Biologists.  Prepared  by  T.G.  Scott  and  C.H.  Wasser.  1980.  The 
Wildlife  Society,  Washington,  DC. 

Macroinvertebrates — Fresh-water  Invertebrates  of  the  United  States.  Prepared  by  R.W.  Pennak.  1953  The  Ronald  Press 
Company,  New  York,  NY. 

Fish — A  List  of  Common  and  Scientific  Names  of  Fishes  from  the  United  States  and  Canada.  Prepared  by  C.R.  Robbins,  R.M. 
Bailey,  C.E.  Bond,  JR.  Brooker,  E.A.  Lachner,  R.N.  Lea,  and  W.B.  Scott.  Special  Publication  12,  4th  edition.  1980.  Ameri- 
can Fisheries  Society,  Bethesda,  MD. 

Amphibians  and  Reptiles — Standard  Common  and  Current  Scientific  Names  for  North  American  Amphibians  and  Rep- 
tiles. Prepared  by  J.T.  Collins,  R.  Conant,  J.E.  Huheey,  J.L.  Knight,  EM.  Rundquist,  and  H.M.  Smith.  Herpetological  Circu- 
lar 12,  2nd  edition.  1982.  Society  for  the  Study  of  Amphibians  and  Reptiles,  Ohio  University,  Athens. 

Birds — Checklist  of  North  American  Birds.  Prepared  by  The  Committee  on  Classification  and  Nomenclature  of  the  Ameri- 
can Ornithologists'  Union.  6th  edition.  1983-  Allen  Press,  Inc.,  Lawrence,  KS. 

Mammals — Revised  Checklist  of  North  American  Mammals  North  of  Mexico.  1982.  Prepared  by  J.K.  Jones,  Jr.,  D.C.  Carter, 
H.H.  Genoways,  R.S.  Hoffman,  and  D.W.  Rice.  Occasional  Papers,  the  Museum,  Texas  Tech  University.  80.  Texas  Tech 
Press,  Lubbock. 


Unless  otherwise  credited,  all  photographs  were  obtained  from  BLM  photo  libraries. 
xu 


REVIEWERS 


The  accuracy,  thoroughness,  and  readability  of  the  entire  book  has  been  enhanced  greatly  by  the  follow- 
ing reviewers  who  have  read  and  critiqued  one  or  more  chapters.  Their  assistance  is  greatly  appreciated. 


Bruce  B.  Ackerman 

Idaho  Cooperative  Wildlife  Research  Unit 
University  of  Idaho 
Moscow,  ID  83843 

Layne  Adams,  Wildlife  Biologist 
U.S.  Bureau  of  Land  Management 
Fairbanks  District  Office 
Fairbanks,  AK  99797 

William  F.  Andelt,  Wildlife  Extension  Agent 
Kansas  State  University 
Cooperative  Extension  Service 
Garden  City,  KS  67846 

Stanley  Anderson,  Wildlife  Biologist 

U.S.  Fish  and  Wildlife  Service 

Wyoming  Cooperative  Fish  and  Wildlife  Research  Unit 

Laramie,  WY  82071 

Gary  N.  Back,  Research  Associate 
University  of  Nevada — Reno 
Range,  Wildlife,  and  Forestry 
Reno,  NY  89512 

James  A.  Bailey,  Professor  of  Wildlife  Biology 
Colorado  State  University 
Department  of  Fishery  and  Wildlife  Biology 
Fort  Collins,  CO  80523 

Robert  G.  Bailey,  Geographer 
U.S.  Forest  Service 
Land  Management  Planning  Systems 
Fort  Collins,  CO  80524 

Gary  Bateman 

Northern  Arizona  University 
Department  of  Biological  Sciences 
Flagstaff,  AZ  86011 

Fred  T.  Batson,  Natural  Resource  Specialist 

U.S.  Bureau  of  Land  Management 

Service  Center 

Denver,  CO  80225-0047 

Richard  G.  Beidleman,  Professor  of  Biology 
The  Colorado  College 
Department  of  Biology 
Colorado  Springs,  CO  80903 

John  Bosworth,  Environmental  Coordinator 
U.S.  Bureau  of  Land  Management 
Susanville  District  Office 
Susanville,  CA  96130-3730 


Sandra  L.  Bowers,  Librarian 
U.S.  Bureau  of  Land  Management 
Service  Center  Library 
Denver,  CO  80225-0047 

R.  Bruce  Bury,  Zoologist 
U.S.  Fish  and  Wildlife  Service 
Denver  Wildlife  Research  Center 
Fort  Collins,  CO  80524 

Jack  Chugg,  Soil  Scientist 
U.S.  Bureau  of  Land  Management 
Washington  Office  Range  Staff 
Washington,  DC  20240 

Lewis  M.  Cowardin,  Wildlife  Biologist 
U.S.  Fish  and  Wildlife  Service 
Northern  Prairie  Wildlife  Research  Center 
Jamestown,  ND  58401 

Paul  E.  Cuplin,  Fishery  Biologist 
U.S.  Bureau  of  Land  Management 
Service  Center 
Denver,  CO  80225-0047 

Robert  K.  Davis,  Wildlife  Economist 
U.S.  Bureau  of  Land  Management 
Washington  Office  Wildlife  Staff 
Washington,  DC  20006 

Russell  Davis 

The  University  of  Arizona 

Department  of  Ecology  and  Evolutionary  Biology 

Tucson,  AZ  85721 

Stephen  Destefano 

Idaho  Cooperative  Wildlife  Research  Unit 
University  of  Idaho 
Moscow,  ID  83843 

Richard  W.  Dierking,  Soil  Scientist 
U.S.  Soil  Conservation  Service 
West  National  Technical  Center 
Portland,  OR  97209 

Mark  A.  Dimmitt,  Curator  of  Plants 
The  Arizona-Sonora  Desert  Museum 
Tucson,  AZ  85704 


R.  Michael  Erwin,  Wildlife  Biologist 
U.S.  Fish  and  Wildlife  Service 
Patuxent  Wildlife  Research  Center 
Laurel,  MD  20811 


Claire  Farrell 

Southern  Oregon  State  College 
Department  of  Biology 
Ashland,  OR  97520 

Stephen  A.  Flickinger,  Professor  of  Fishery  Biology 

Colorado  State  University 

Department  of  Fishery  and  Wildlife  Biology 

Fort  Collins,  CO  80523 

Michael  Garratt,  Statistician 
U.S.  Bureau  of  Land  Management 
Service  Center 
Denver,  CO  80225 

Kenneth  M.  Giesen,  Wildlife  Researcher 
Colorado  Division  of  Wildlife 
Wildlife  Research  Center 
Fort  Collins,  CO  80526 

Nancy  F.  Green,  Wildlife  Biologist 
U.S.  Bureau  of  Land  Management 
Washington  Office  Wildlife  Staff 
Washington,  DC  20006 

Carole  K.  Hamilton,  Wildlife  Biologist 
U.S.  Bureau  of  Land  Management 
Washington  Office  Wildlife  Staff 
Washington,  DC  20006 

John  Haugh,  Vertebrate  Ecologist 
U.S.  Bureau  of  Land  Management 
Washington  Office  Resource  Sciences  Staff 
Washington,  DC  20006 

Steve  Hawks,  Wildlife  Biologist 
U.S.  Bureau  of  Land  Management 
Susanville  District  Office 
SusanvUle,  CA  96130-3730 

Dale  Hein,  Professor  of  Wildlife  Biology 
Colorado  State  University 
Department  of  Fishery  and  Wildlife  Biology 
Fort  Collins,  CO  80523 

Marian  Hershcopf,  Librarian 
Colorado  Division  of  Wildlife 
Research  Center  Library 
Fort  Collins,  CO  80526 

Dale  Hoffman,  Limnologist 
U.S.  Fish  and  Wildlife  Service 
Region  6 
Denver,  CO  80525 

Rick  Hoffman,  Wildlife  Researcher 
Colorado  Division  of  Wildlife 
Wildlife  Research  Center 
Fort  Collins,  CO  80526 

Jon  K.  Hooper,  Coordinator,  Parks  and  Natural  Resources 

Management  Option 
California  State  University,  Chico 
Department  of  Recreation  and  Parks  Management 
Chico,  CA  95929-0560 


Marshall  A.  Howe,  Section  Leader,  Migratory  Nongame 

Birds 
U.S.  Fish  and  Wildlife  Service 
Patuxent  Wildlife  Research  Center 
Laurel,  MD  20811 

Jack  D.  Jones,  Wildlife  Biologist 
U.S.  Bureau  of  Land  Management 
Butte  District  Office 
Butte,  MT  59702-3388 

K.  Bruce  Jones,  Wildlife  Biologist 
U.S.  Bureau  of  Land  Management 
Phoenix  Training  Center 
Phoenix,  AZ  85016 

William  Kepner,  Fishery  Biologist 
U.S.  Bureau  of  Land  Management 
Phoenix  District  Office 
Phoenix,  AZ  85017 

William  L.  Kirk,  Habitat  Specialist 
U.S.  Fish  and  Wildlife  Service 
Anchorage,  AK  99503 

James  F.  LaBounty,  Fishery  Biologist 
U.S.  Bureau  of  Reclamation 
Denver,  CO  80225 

Joe  Lint,  Wildlife  Biologist 
U.S.  Bureau  of  Land  Management 
Roseburg  District  Office 
Roseburg,  OR  97470 

CD.  1  ittk  field 
Malheur  Field  Station 
Princeton,  OR  97721 

Larry  S.  Mangan,  Wildlife  Biologist 
U.S.  Bureau  of  Land  Management 
Shoshone  District  Office 
Shoshone,  ID  83352 

Chris  Maser,  Research  Wildlife  Biologist 
U.S.  Bureau  of  Land  Management 
Corvallis,  OR  97331 

Ken  McGinty,  Technical  Editor 
U.S.  Bureau  of  Land  Management 
Service  Center 
Denver,  CO  80225-0047 

Gary  McVicker,  Natural  Resource  Specialist 
U.S.  Bureau  of  Land  Management 
Washington  Office  Renewable  Resources  Staff 
Washington,  DC  20240 

Brian  A.  Millsap,  Raptor  Biologist 
National  Wildlife  Federation 
Raptor  Information  Center 
Washington,  DC  20036 

Sam  Montgomery,  Wildlife  Biologist 
U.S.  Bureau  of  Land  Management 
Colorado  State  Office 
Denver,  CO  80205 


James  A.  Mosher,  Consultant 
Savage  River  Consulting 
Frostburg,  MD  21532 

Lewis  Nelson,  Jr.,  Professor 

University  of  Idaho 

College  of  Forestry,  Wildlife,  and  Range  Sciences 

Moscow,  ID  83843 

Nick  Nydegger,  Wildlife  Biologist 
U.S.  Bureau  of  Land  Management 
Boise  District  Office 
Boise,  ID  83705 

Richard  R.  Olendorflf,  Wildlife  Biologist 
U.S.  Bureau  of  Land  Management 
California  State  Office 
Sacramento,  CA  85825 

Lewis  W.  Oring,  Professor 
The  University  of  North  Dakota 
Department  of  Biology 
Grand  Forks,  ND  58202 

Ed  Parsons,  Natural  Resource  Specialist 
U.S.  Bureau  of  Land  Management 
Washington  Office  Range  Staff 
Washington,  DC  20006 

Howard  Quigley 

Idaho  Cooperative  Wildlife  Research  Unit 
University  of  Idaho 
Moscow,  ID  83843 

William  Radtkey,  Wildlife  Biologist 
U.S.  Bureau  of  Land  Management 
Washington  Office  Wildlife  Staff 
Washington,  DC  20006 

Harry  V.  Reynolds,  Game  Biologist 
Alaska  Department  of  Fish  and  Game 
Division  of  Game 
Fairbanks,  AK  99701 

Richard  T.  Reynolds,  Research  Wildlife  Biologist 
U.S.  Forest  Service 

Rocky  Mountain  Forest  and  Range  Experiment  Station 
Fort  Collins,  CO  80526-2098 

Chandler  S.  Robbins,  Research  Biologist 
U.S.  Fish  and  Wildlife  Service 
Patuxent  Wildlife  Research  Center 
Laurel,  MD  20708 

John  T.  Rotenberry,  Assistant  Professor 
Bowling  Green  State  University 
Department  of  Biological  Sciences 
Bowling  Green,  OH  43403 

Robert  Roudabush,  Soil  Scientist 
U.S.  Bureau  of  Land  Management 
Arizona  State  Office 
Phoenix,  AZ  85011 

Hal  Salwasser,  National  Wildlife  Ecologist 
U.S.  Forest  Service 


Wildlife  and  Fish  Ecology  Unit 
Fort  Collins,  CO  80524 

Henry  L.  Short,  Terrestrial  Ecologist 
U.S.  Fish  and  Wildlife  Service 
Western  Energy  and  Land  Use  Team 
Fort  Collins,  CO  80526-2899 

Ronnie  Sidner 

The  University  of  Arizona 

Department  of  Ecology  and  Evolutionary  Biology 

Tucson,  AZ  8572 1 

Karen  Steenhof,  Wildlife  Biologist 
U.S.  Bureau  of  Land  Management 
Boise  District  Office 
Boise,  ID  83705 

George  A.  Swanson,  Wildlife  Biologist 
U.S.  Fish  and  Wildlife  Service 
Northern  Prairie  Wildlife  Research  Center 
Jamestown,  ND  58401 

Charlene  Swibas,  Technical  Information  Specialist 
U.S.  Environmental  Protection  Agency 
National  Enforcement  Investigations  Center 
Denver,  CO  80225 

Michael  Tewes 

Idaho  Cooperative  Wildlife  Research  Unit 
University  of  Idaho 
Moscow,  ID  83843 

S.P.  Thompson 

U.S.  Fish  and  Wildlife  Service 
Stillwater  Habitat  Management  Area 
Fallon,  NV  89406 

Roy  E.  Tomlinson,  Southwest  Dove  Coordinator 
U.S.  Fish  and  Wildlife  Service 
Office  of  Migratory  Bird  Management 
Albuquerque,  NM  87103 

Jared  Verner,  Research  Ecologist 
U.S.  Forest  Service 
Forestry  Sciences  Laboratory 
Fresno,  CA  93710 

Lee  F.  Werth,  Natural  Resource  Specialist 
U.S.  Bureau  of  Land  Management 
Service  Center 
Denver,  CO  80225-0047 

U.W.  Wilson 

U.S.  Fish  and  Wildlife  Service 
Olympia,  WA  98504 

Jerry  O.  Wolff,  Assistant  Professor  of  Biology 
University  of  Virginia 
Department  of  Biology 
Charlottesville  VA  22901 

Leonard  S.  Young,  Wildlife  Biologist 
U.S.  Bureau  of  Land  Management 
Boise  District  Office 
Boise,  ID  83705 


INTRODUCTION 


Inventory  and  monitoring  of  wildlife  and  wild- 
life habitat  is  often  considered  mundane  and  uninter- 
esting work,  yet  many  great  discoveries  in  both 
general  biology  and  more  specifically  wildlife  biology 
were  made  while  biologists  and  naturalists  per- 
formed these  "mundane"  tasks.  Charles  Darwin  and 
Alfred  Wallace  were  both  conducting  "inventories" 
when  they  developed  their  concepts  of  evolution 
and  natural  selection.  In  more  recent  times,  mem- 
bers of  the  U.S.  Biological  Survey,  the  forerunner  of 
the  U.S.  Fish  and  Wildlife  Service,  inventoried  wild- 
life and  wildlife  habitat  of  western  North  America, 
contributing  significantly  to  our  understanding  of  the 
ecology  and  zoogeography  of  these  regions.  Al- 
though much  of  their  work  involved  collecting  and 
documenting  evidence  of  fauna  in  various  regions, 
their  contributions  went  far  beyond  merely  listing 
the  animals  present. 

Consider  the  contributions  of  C.  Hart  Merriam 
(Merriam's  Life  Zones),  Olaus  Murie  (The  Elk 
of  North  America  and  Alaska- Yukon  Caribou ),  or 

Aldo  Leopold,  who  developed  many  revolutionary 
ideas  about  such  subjects  as  predation  and  wildlife 
habitat  management  while  performing  duties  as  a 
forester  for  the  U.S.  Forest  Service  in  the  Southwest 
and  later  directing  game  surveys  in  the  Midwest. 
One  can,  of  course,  argue  that  these  men  were  doing 
work  more  akin  to  research,  particularly  later  in 
their  careers.  This  may  be  partly  true,  but  the  dis- 
tinction between  research  and  the  activities  of  inven- 
tory and  monitoring  can  be  nebulous.  In  any  case, 
the  skills  required  for  successfully  conducting  an 
inventory  or  monitoring  study  are  similar  to  those 
needed  for  successful  research.  Both  require  an  abil- 
ity to  systematically  make  and  record  observations, 
conceptually  and  quantitatively  synthesize  large 
amounts  of  data,  and  effectively  communicate,  both 
verbally  and  in  writing. 

Inventory  and  monitoring  of  wildlife  habitat  is  a 
statutory  responsibility  of  many  land  management 
agencies  including  the  U.S.  Bureau  of  Land  Manage- 
ment and  the  U.S.  Forest  Service. 

Inventory  and  monitoring  of  wildlife  habitat  is 
based  on  the  assumption  that  measurements  of  a  set 
of  habitat  attributes  can  be  used  to  predict  presence 
or  abundance  of  wildlife  species.  The  same  relation- 
ship between  habitat  variables  and  wildlife  popula- 
tions is  used  for  both  habitat  inventory  and 
monitoring.  This  relationship  has  long  been  recog- 
nized by  naturalists  and  is  the  basis  of  several  recent 
habitat  evaluation  systems.  Habitat  variables  may 


consist  of  vegetative  structure,  plant  species  compo- 
sition, presence  of  physical  features,  or  other  factors. 
Predictions  resulting  from  such  measurements  need 
to  be  verified  by  observing  presence  or  measuring 
relative  abundance  of  the  wildlife  species  of  interest. 
In  some  cases  these  measurements  need  to  be  taken 
concurrently  to  determine  the  most  relevant  habitat 
variables.  Thus  species  or  population  measurements 
are  a  necessary  part  of  development  and  use  of  any 
system  for  inventory  and  monitoring  of  wildlife 
habitat. 

Wildlife  habitat  inventory  consists  of  measuring 
selected  habitat  variables  on  a  piece  of  land  to  infer 
presence  or  abundance  of  wildlife  species.  The  pur- 
pose is  to  determine  the  wildlife  resources  currently 
supported  by  the  area. 

Wildlife  habitat  monitoring  consists  of  repeat- 
edly measuring  habitat  or  population  variables  to 
infer  changes  in  the  capability  of  the  land  to  support 
wildlife.  The  purpose  is  usually  issue-oriented,  i.e., 
it  determines  how  some  activity  such  as  mining, 
livestock  grazing,  or  recreational  activity  is  affecting 
wildlife  habitat  and  ultimately  wildlife  populations. 
Monitoring  is  also  used  to  determine  the  effective- 
ness of  habitat  management  practices  such  as  seed- 
ing, burning,  or  water  development.  Although 
monitoring  is  designed  to  detect  changes  from  hu- 
man activities,  it  may  also  detect  changes  caused  by 
climatic  conditions  or  other  factors  not  under  hu- 
man control.  Therefore,  the  purpose  of  monitoring  is 
not  only  to  measure  change  but  also  to  determine 
the  cause(s)  of  change.  In  particular,  programs  for 
monitoring  wildlife  habitat  must  be  designed  to  sepa- 
rate the  effects  of  human  activity  from  those  caused 
by  variations  in  weather. 

Typically,  a  monitoring  program  consists  of 
measuring  habitat  variables  that  are  required  by  key 
wildlife  species  or  correlate  with  presence  or  abun- 
dance of  key  wildlife  species.  For  example,  the  num- 
ber of  pileated  woodpeckers  highly  correlates  with 
the  number  of  suitable  nesting  snags.  A  reduction  in 
the  number  of  such  snags  can  thus  be  assumed  to 
cause  a  reduction  in  the  number  of  breeding 
woodpeckers. 

Habitat  monitoring,  like  inventory,  may  be  ac- 
companied by  population  measurements  to  confirm 
the  habitat  relationships.  If  the  purpose  is  to  measure 
the  effectiveness  of  a  habitat  management  practice, 
then  species  or  population  measurements  must  be 
taken.  For  example,  if  water  is  being  developed  to 


improve  habitat  for  Gambel's  quail,  then  monitor- 
ing should  be  designed  to  measure  changes  in  the 
number  of  quail,  not  changes  in  the  number  of 
"guzzlers." 

Over  50  years  ago,  Aldo  Leopold,  in  a  chapter 
on  "Game  Range"  in  his  classic  volume  on  Game 
Management  wrote: 

"What  is  Game  Range?  When  the  game  man- 
ager asks  himself  whether  a  given  piece  of  land 
is  suitable  for  a  given  species  of  game,  he  must 
realize  that  he  is  asking  no  simple  question, 
but  rather  he  is  facing  one  of  the  great  enigmas 
of  animate  nature.  An  answer  good  enough  for 
practical  purposes  is  usually  easy  to  get  by 
the  simple  process  of  noting  whether  the  spe- 
cies is  there  and  ready,  or  whether  it  occurs  on 
'similar'  range  nearby.  But  let  him  not  be  cock- 
sure about  what  is  similar,'  for  this  involves  the 
deeper  question  of  why  a  species  occurs  in 
one  place  and  not  in  another,  which  is  proba- 
bly the  same  as  why  it  persists  at  all.  No  living 
man  can  answer  that  question  fully  in  even  one 
single  instance." 

When  Leopold  referred  to  a  biologist  determin- 
ing "similar"  ranges,  he  introduced  the  concept  of 
a  "habitat  model"  and  described  the  reasons  moni- 
toring is  necessary. 

Habitat  models  can  of  course  be  of  many  types. 
Hunters  often  use  conceptual  models  for  determin- 
ing, for  example,  high  density  autumn  pheasant  habi- 
tat when  passing  some  fields  and  hunting  in  others. 
Hunter  perceptions  of  good  habitat  may  be  accurate 
even  though  they  are  unable  to  describe  specific 
qualities  when  choosing  sites.  After  a  day  in  the  field, 
a  hunter  may  reflect  on  the  day  afield  and  suggest 
that  "millet  fields  adjacent  to  uncultivated  swales  and 
breaks  are  the  best  places  to  hunt."  This  is  a  verbal 
model.  A  biologist  may  go  further  and  describe  opti- 
mum habitat  as  "grain  fields  of  40  acres  or  more 
within  Vz  mile  of  water  and  within  'A  mile  of  uncul- 
tivated areas,  having  brushy  cover  at  least  4  feet 
high."  This  is  a  quantitative  model  which  can  be  sys- 
tematically tested  and  evaluated. 


Two  points  deserve  noting  about  habitat  models 
and  their  use  by  biologists  and  others.  First,  we  all 
use  models  in  our  everyday  life,  including  habitat 
models.  Many  are  good,  although  based  only  on 
"professional  judgment,"  i.e.,  conceptual  models. 
Second,  with  the  controversy  surrounding  many  cur- 
rent biological  assessments,  biologists  are  being 
asked,  encouraged,  or  even  required  to  make  their 
models  more  explicit  by  various  parties,  interest 
groups,  publics,  and  agencies  involved  with  or  af- 
fected by  project  decisions.  Explicit  and  quantitative 
models  not  only  provide  a  mechanism  for  improving 
predictions,  but  also  provide  an  important  communi- 
cation tool  by  taking  some  of  the  mystery  out  of 
"professional  judgment." 


But  models  have  their  limits.  Models  are  approx- 
imations at  best.  As  Aldo  Leopold  suggested,  no  liv- 
ing person  can  develop  a  "perfect"  habitat  model. 
However,  monitoring  can  allow  us  to  test  and  refine 
our  models  in  addition  to  minimizing  the  risk  of 
resource  damage  from  poor  predictions.  We  can  thus 
have  a  cycle  of  prediction,  decision,  action,  and 
monitoring. 


This  approach  to  natural  resource  problems  has 
received  a  lot  of  attention  in  the  last  10  years  and 
has  been  variously  described  as  adaptive  environ- 
mental assessment,  cyclic-incrementalism,  "muddling 
through,"  or  even  "common-sense  management,"  as 
discussed  in  several  chapters  in  this  book.  It  seems 
to  be  a  logical  and  common-sense  approach  to  solv- 
ing a  wide  variety  of  natural  resource  problems. 
With  such  an  approach,  the  separate  but  interrelated 
tasks  of  problem  identification,  data  collection,  pre- 
diction, decision,  and  monitoring  can  also  be  linked 
in  a  cyclical  process.  This  process  can  result  in  bet- 
ter management  of  wildlife  resources  and  also  devel- 
opment of  better  tools  (models)  for  understanding 
wildlife  habitat  relationships.  Although  many  of  the 
following  chapters  describe  isolated  parts  of  this 
process,  they  should  be  viewed  in  context,  not  as 
ends  in  themselves,  but  as  necessary  steps  in  a  con- 
tinuing process  toward  better  management  of  our 
natural  resources. 


Allen  Y.  Cooperrider 


xviii 


I  PLANNING 


1  The  Inventory  and  Monitoring  Process 

2  Data  Types 

3  Literature  Review 

4  Habitat  Mapping 


TFIF 

INVENTORY 
AND 

MONITORING 
PROCESS 


Editor's  Note:  This  chapter  provides  an  overview  of 
the  entire  inventory  and  monitoring  process,  with 
emphasis  on  the  planning  stage.  Biologists  in  their 
hurry  to  get  to  the  field  to  gather  the  all-important 
data  often  neglect  to  spend  the  necessary  time  in 
planning  their  work.  They  may  also  neglect  to  al- 
low time  for  the  necessary  follow-up  procedures, 
such  as  data  analysis.  Such  actions  result  in  much 
wasted  time  and  money.  This  chapter  not  only  pro- 
vides an  overview  of  the  entire  process,  but  also 
directs  the  reader  to  other  sections  and  chapters  in 
the  book  that  cover  specific  topics  in  more  depth. 


K.  Bruce  Jones 

U.S.  Bureau  of  Land  Management 
Phoenix  Training  Center 
Phoenix,  AZ  85015 


"The  amount  of  time  spent  in  solving  a  problem  is 
inversely  proportional  to  the  amount  of  time  spent 
in  characterizing  the  problem." 

—  Alan  Speigel,  Speigel's  Laws  of  Management 


"Planning  without  action  is  futile.  Action  without 
planning  is  fatal." 

—  K.  Hamilton  and  E.P.  Bergersen,  Methods  to 
Estimate  Aquatic  Habitat  Variables 


INTRODUCTION 

The  most  critical  stage  of  implementing  and 
completing  an  inventory  or  monitoring  study  is  not 
data  collection,  presentation,  or  interpretation,  but 
rather  design.  Years  of  data  can  be  useless  if  a  study 
is  poorly  designed.  This  chapter  outlines  a  procedure 
for  any  inventory  or  monitoring  project.  By  follow- 
ing these  steps,  biologists  will  establish  a  flexible, 
systematic,  and  logical  approach  toward  solving  wild- 
life habitat  management  problems.  Use  of  this  proce- 
dure should  also  increase  the  effectiveness  of 
inventory  and  monitoring  studies,  reduce  costs,  and 
decrease  failures  caused  by  poor  design  or  imple- 
mentation. The  process  will  also  lead  to  improved 
interpretation  and  presentation  of  results.  By  focus- 
ing inventories  and  monitoring  studies  on  specific 
issues,  wildlife  resources  should  receive  greater  con- 
sideration in  the  decisionmaking  process  and  realize 
greater  benefits.  Therefore,  issue-driven  inventories 
and  studies,  such  as  resolving  conflicting  land-use 
opportunities,  are  emphasized. 

The  inventory  and  monitoring  process  consists 
of  a  series  of  events: 


(1)    Problem  definition  (scoping) — 

•  identifying  issues,  concerns,  or  opportuni- 
ties; 

•  reducing  general  problems  to  specific  ones; 

•  predicting  and  analyzing  extent  of  potential 
problems; 

•  identifying  specific  inventory  and  monitor- 
ing study  objectives; 

•  prioritizing  objectives;  and 

•  deciding  types  and  levels  of  data  needed. 


Current  address:  Office  of  Endangered  Species,  U.S.  Fish  and 
Wildlife  Service,  Washington,  DC.  20240. 


Inventory  and  Monitoring 


(2)  Data  collection  (inventory,  studies); 

(3)  Data  analysis,  interpretation,  evaluation,  pres- 
entation, and  storage; 

(4)  Management  decision; 

(5)  Monitoring  studies;  and 

(6)  Management  review. 

Although  the  stages  may  be  called  different 
names  in  different  organizations,  this  approach  to 
problem  solving  is  quite  basic.  This  chapter  covers 
this  process  with  emphasis  on  steps  1,  2,  3,  and  5 
which  are  major  concerns  of  staff  biologists  and 
specialists. 


PROBLEM  DEFINITION  (SCOPING) 

Problem  definition,  often  termed  scoping,  deter- 
mines the  detail  of  inventory  or  monitoring  studies 
and  is  critical  in  conducting  these  efforts  success- 
fully. It  is  the  process  of  determining  specific  prob- 
lems and  issues  that  need  to  be  addressed,  data 
needed  to  address  these  issues,  and  priority  issues 
and  data  needs.  It  consists  of  six  sequential  steps  as 
listed  in  the  Introduction  and  described  here. 

Identifying  Issues,  Concerns,  or 
Opportunities 

Wildlife-  and  habitat-related  problems  originate 
from  the  following  six  general  sources: 


( 1 )  Need  for  baseline  data  for  land-use  planning; 

(2)  Proposals  from  wildlife  biologists; 

(3)  Management  proposals  from  persons  outside 
the  wildlife  resource; 

(4)  Major  land-use  planning  guidance; 

(5)  Ongoing  impacts  from  resource  development, 
originating  within  the  agency;  and 

(6)  Ongoing  impacts  from  resource  development, 
originating  outside  the  agency. 

Baseline  data  for  land-use  planning  are  the  basic 
driving  force  of  wildlife  inventories  because  they 
show  conditions  as  they  currently  exist.  Baseline 
data  are  useful  in  determining  potential  impacts  and 
in  comparing  actual  impacts  after  development. 

Proposals  from  wildlife  biologists  to  improve 
habitat  for  one  species  may  also  produce  potential 


problems  for  other  species.  Consider,  for  example,  a 
habitat  management  plan  that  proposes  to  convert 
pinyon-juniper  (Pinus-Juniperus  sp.)  and  chaparral 
habitats  into  grassland  for  enhancing  pronghorn  (An- 
tilocapra  americana)  populations.  Although  this 
proposal  would  benefit  pronghorn,  it  would  also  de- 
crease habitat  available  to  raptors,  upland  game,  and 
nongame  species. 

Proposals  by  managers  outside  the  wildlife  re- 
source could  cause  many  potential  impacts.  These 
impacts  are  identified  by  biologists,  other  resource 
specialists,  and  managers.  For  example,  a  biologist 
identifies  a  general  problem  between  a  mining  plan 
and  wildlife  habitat;  a  proposed  open-pit  mineral 
operation  would  greatly  reduce  the  amount  and 
quality  of  wildlife  habitat  in  the  area. 

Some  of  the  more  general,  broad-scale  problems 
that  require  inventories  and  monitoring  studies  are 
identified  through  land-use  planning.  For  example,  an 
areawide  grazing  proposal,  if  implemented,  could 
negatively  affect  several  million  acres  of  wildlife 
habitat.  This  type  of  problem  could  be  identified  by 
a  wide  variety  of  resource  specialists  within  an 
agency  during  the  scoping  process.  Ongoing  impacts 
or  problems  warranting  inventories  and  monitoring 
can  also  be  identified  by  resource  specialists  inside 
an  agency  (e.g.,  range  conservationists)  or  outside. 
For  example,  biologists  within  an  agency  could  iden- 
tify that  current  livestock  grazing  practices  are  re- 
ducing the  quality  and  quantity  of  forage  available  to 
big  game.  Similarly,  these  kinds  of  problems  can  be 
identified  by  sources  outside  an  agency,  including 
other  federal  agencies,  state  game  and  fish  depart- 
ments, private  wildlife  organizations  (e.g.,  The  Wild- 
life Society),  and  local  citizens. 

Reducing  General  Problems  to  Specific 
Ones 

After  general  issues  have  been  identified,  the 
next  step  is  to  break  down  general  problems  (as  pre- 
viously discussed)  into  specific  ones.  For  example,  a 
biologist  finds  a  correlation  between  reduced  wild- 
life habitat  and  habitat  quality  and  increased  live- 
stock grazing.  The  next  step  is  to  identify,  more 
specifically,  factors  that  might  account  for  relation- 
ships between  grazing  pressure  and  loss  of  wildlife 
habitat  and  quality.  In  this  example,  the  agency  biol- 
ogist might  identify  three  specific  problems: 


( 1 )  Increased  stocking  of  livestock  may  reduce 
the  amount  and  number  of  perennial  plant 
species. 

(  2  )    Long-term  overstocking  of  livestock  may  re- 
duce the  quality  of  the  "A  horizon"  in  the 
soil,  resulting  in  large-scale  erosion  and  a 
drop  in  vegetation  productivity. 


Inventory  and  Monitoring 


Perennial  plant  species  may  be  reduced  by  increased  live- 
stock grazing. 


(3)  Overstocking  and  poor  livestock  distribution 
may  eliminate  cover  around  waterfowl  nest- 
ing ponds. 

Sections  II  and  III  of  this  book  can  provide  as- 
sistance in  identifying  habitat  and  species  problems. 
For  example,  a  new  biologist  initiating  an  inventory 
in  a  grassland  area  with  known  populations  of  prairie 
chickens  (Tympanuchus  cupido)  may  wish  to  con- 
sult Chapter  6,  Rangelands,  and  Chapter  20,  Upland 
Game  Birds,  for  guidance  on  habitat  and  featured 
species. 

In  identifying  specific  problems  from  general 
ones,  biologists  should  seek  expertise  from  other  re- 
source specialists.  In  the  example  given  above,  the 
soil  scientist  or  soil  conservationist  plays  a  major 
role  in  making  general  problems  specific.  In  this 
example,  range  conservationists,  hydrologists,  and 
surface  protection  specialists  would  also  play  major 
roles  in  identifying  specific  problems. 

Predicting  and  Analyzing  Extent  of 
Potential  Problems 

After  specific  problems  have  been  identified,  the 
biologist  should  attempt  to  make  some  preliminary 
predictions  as  to  the  extent  of  these  specific  prob- 
lems. General  predictions  can  be  formulated  from  a 
large  variety  of  information  sources: 


( 1 )  Existing  agency  data  (e.g.,  range  survey  files); 

(2)  Agency  expertise  in  resources  other  than 
wildlife  (e.g.,  soil  sciences); 

(3)  Review  of  literature  containing  data  and  re- 
sults on  similar  problems  (e.g.,  manuscripts 


on  impacts  to  wetland  vegetation  from  live- 
stock grazing); 

(4)  Public  input  (e.g.,  observations  of  citizens 
familiar  with  the  problem  and  area); 

(5)  Interest  group  input  (e.g.,  local  wildlife  advo- 
cacy groups  familiar  with  the  area  and/or 
problem); 

(6)  Experts  on  the  habitat  or  species  in  question. 


The  biologist  should  make  sure  he  or  she  has 
gathered  available  information  before  collecting  addi- 
tional data.  Chapter  3,  Literature  Review,  provides 
guidance  on  searching  existing  data  sources.  In  addi- 
tion to  sources  listed  above,  some  recently  described 
procedures  may  allow  biologists  to  predict  specific 
problems  that  could  exist  between  resources  and  the 
extent.  Thomas  (1979)  developed  life-form  tables 
that  group  species  according  to  components  of  habi- 
tat used  for  feeding  and  breeding.  Similarly,  Short 
(1982)  developed  a  procedure  that  groups  species 
into  guilds  based  on  habitat  layers  used  for  feeding 
and  breeding.  Both  procedures  allow  biologists  to 
use  varying  degrees  of  existing  information  to  assess 
the  complexity  of  wildlife  species  arranged  in  habitat 
space.  Biologists  can  then  predict  the  extent  of  im- 
pacts to  potential  species  that  would  occur  if  certain 
habitat  components  were  lost.  Although  these  proce- 
dures neither  predict  species  occurrence  in  an  area 
nor  generate  species-specific  data,  they  do  provide 
an  excellent  preliminary  analysis  of  potential  impacts 
to  most  wildlife  species  within  a  given  area. 

The  U.S.  Fish  and  Wildlife  Service  and  U.S.  For- 
est Service  have  developed  models  that  can  be  used 
for  analyzing  potential  impacts  to  wildlife.  Generally, 
these  models  provide  an  index  of  habitat  quality  for 
various  vertebrates.  By  knowing  the  amount  and 
quality  of  habitat  that  may  be  affected  by  a  proposed 
action,  the  biologist  can  use  these  models  to  deter- 
mine corresponding  impacts  on  specific  vertebrates. 
The  use  of  these  models  are  discussed  in  greater 
detail  in  Chapter  2,  Data  Types,  and  Chapter  38, 
Habitat  Evaluation  Systems. 

After  compiling  information  from  existing  data 
sources  and  making  some  preliminary  predictions, 
the  biologist  will  know  whether  an  on-the-ground 
inventory  or  study  is  needed  to  adequately  assess 
or  rectify  specific  problems.  Existing  data  and  pre- 
liminary analyses  may  be  enough  that  a  biologist 
needs  only  to  collect  data  on  one  or  a  few  specific 
problems. 

Based  on  existing  data  and  preliminary  predic- 
tions, the  agency  biologist  can  provide  decision- 
makers with  information  that  may  affect  a  proposed 


Inventory  and  Monitoring 


action.  The  impact  of  the  biologist's  recommenda- 
tions on  the  proposed  action  will  depend  on  the 
adequacy  of  existing  biological  data  and  economical 
and  political  factors.  Preliminary  predictions  influ- 
ence managers  to  make  the  following  decisions  rela- 
tive to  the  proposal: 


( 1 )  Drop  the  proposal  entirely; 

(2)  Leave  the  proposal  as  is;  or 

(  3  )    Modify  the  proposal  so  adverse  impacts  are 
reduced  or  eliminated. 

Providing  good  information  to  managers  at  an 
early  stage  can  be  quite  effective  in  mitigating  im- 
pacts. Proposals  can  be  dropped  or  modified  before  a 
project  has  progressed  too  far  in  planning. 

Modifying  a  proposed  action  often  involves  pro- 
posing new  alternatives  to  mitigate  predicted  ad- 
verse impacts.  The  degree  of  mitigation  will  depend 
on  the  accuracy  of  existing  data  and  techniques  used 
to  make  preliminary  predictions.  Broad,  incomplete 
wildlife  data  used  to  predict  impacts  of  a  site-specific 
proposal  would  not  produce  accurate  assessments. 


Identifying  Specific  Inventory  and 
Monitoring  Study  Objectives 

At  this  stage  in  the  process,  the  agency  biologist 
has  identified  specific  problems,  determined  the  ade- 
quacy of  existing  data,  and  made  some  preliminary 
predictions.  The  biologist  will,  therefore,  have  a  list 
of  specific  problems  requiring  on-the-ground  data. 

To  develop  a  series  of  procedures  and  analyses 
needed  to  address  the  identified  problems,  the  biolo- 
gist should  transform  each  specific  problem  into 
specific  objectives.  For  example,  a  specific  problem 
was  identified  when  the  biologist  suspected  in- 
creased livestock  grazing  was  reducing  habitat  struc- 
ture. He  or  she  must  then  ask  the  following  types 
of  questions: 


( 1 )  Is  livestock  grazing  reducing  habitat  struc- 
ture? 

(2)  If  so— 

•  Are  stocking  rates  related  to  reduced  habi- 
tat structure? 

•  How  much  livestock  grazing  causes  losses 
in  habitat  structure? 

•  Which  plants  are  structurally  affected  by 
livestock? 


These  are  examples  of  a  few  specific  questions 
that  can  be  asked  about  a  specific  problem.  The 
questions  then  should  be  transformed  into  specific 
objectives.  For  example,  determine  the  plant  species 
that  are  structurally  reduced  by  livestock.  This  infor- 
mation can  then  be  used  for  prioritizing  objectives. 

Agency  biologists  must  follow  this  logical  proce- 
dure in  formulating  specific  objectives,  because  the 
effectiveness  of  the  inventory  or  study  will  only  be 
as  good  as  the  biologist's  understanding  of  the  prob- 
lem(s)  and  his  or  her  ability  to  formulate  specific 
objectives  that  will  answer  specific  questions. 

Prioritizing  Objectives 

Typically,  funding  and  personnel  are  inadequate 
to  collect  on-the-ground  data  for  all  identified  objec- 
tives. Therefore,  the  agency  biologist  needs  to  priori- 
tize. In  many  instances,  agencies  have  identified 
priority  species  and  habitat;  however,  the  biologist 
may  need  to  consider  other  factors,  such  as — 

•  policies,  laws,  and  regulations; 

•  agency  priorities; 

•  potential  degree  of  impact; 

•  feasibility  (i.e.,  whether  the  problem  can  be 
logically,  biologically,  or  economically  re- 
solved in  the  time  given); 

•  public  interests. 

Deciding  Types  and  Levels  of  Data  Needed 

Once  a  decision  is  made  to  collect  data  on 
priority  objectives,  the  biologist  must  decide  the 
types  and  levels  of  data  needed.  Chapter  2,  Data 
Types,  provides  guidance  on  types  and  levels  of  data 
needed  for  inventory  and  monitoring  studies. 


DATA  COLLECTION  (INVENTORY,  STUDIES) 

After  all  objectives  and  methods  for  an  inven- 
tory or  study  are  determined,  the  biologist  should 
develop  a  schedule.  This  schedule  should  allot 
time  for — 

•  hiring  or  contracting; 

•  obtaining  necessary  equipment; 

•  holding  personnel  meetings; 

•  doing  preliminary  field  work; 

•  selecting  specific  methods; 

•  habitat  mapping  and  field  sampling;  and 

•  analyzing  data  and  interpreting,  evaluating, 
and  presenting  results. 


Inventory  and  Monitoring 


Although  data  analysis,  interpretation,  evalua- 
tion, and  presentation  are  covered  separately,  any 
schedule  should  include  enough  time  for  these  criti- 
cal tasks.  Some  efforts  require  more  than  one  person, 
but  even  a  single  biologist  should  follow  the  same 
sequence  of  events  when  conducting  an  inventory  or 
monitoring  study.  Section  IV  in  this  book,  Habitat 
Measurements,  provides  guidance  on  field  data  col- 
lection methods. 

Hiring  or  Contracting 

Competent  field  personnel  are  essential  in  ob- 
taining inventory  and  monitoring  study  objectives. 
The  degree  of  expertise  needed  will  vary  greatly, 
depending  on  the  type  of  animals  and  habitats  to  be 
sampled.  Vegetation  and  topography  can  be  sampled 
by  a  general  biologist  having  a  few  days  of  training. 
Animal  population  sampling,  however,  is  specialized 
and  usually  requires  someone  with  formal  education 
and  field  experience  with  the  taxa  to  be  sampled. 
Contracting  may  be  a  way  of  obtaining  the  expertise 
needed  to  conduct  a  highly  specialized  inventory 
or  monitoring  study.  However,  contractors  generally 
cost  more  than  in-house  personnel. 

Whether  an  inventory  or  monitoring  study  is 
conducted  in-house  or  by  contract,  a  biologist  must 
plan  specific  hiring  needs  at  least  3  months  before 
starting.  The  best  possible  study  design  and  funding 
level  could  fail  if  a  hiring  or  contracting  plan  was 
late  or  poorly  organized. 

Obtaining  Necessary  Equipment 

The  biologist  must  determine  the  types  of 
equipment  needed  to  meet  inventory  or  monitoring 
study  objectives.  The  biologist  must  order  any  ex- 
pensive or  hard-to-get  equipment  at  least  3  months 
before  the  inventory  or  monitoring  study. 

Holding  Personnel  Meetings 

Preliminary  meetings  of  all  involved  personnel 
are  perhaps  the  most  important  and  essential  part  of 
a  successful  inventory  or  monitoring  study.  At  the 
first  meeting,  the  lead  biologist  should  introduce 
everyone,  assure  objectives  are  clear,  and  outline 
agency  procedures.  The  lead  biologist  should  then 
request  the  specialists  to  review  existing  data  and  lit- 
erature pertinent  to  their  respective  inventory  or 
monitoring  study  objectives.  For  example,  a  raptor 
specialist  could  gather  locality  records  from  all  mu- 
seums that  have  made  collections  in  the  study  area. 
He  or  she  should  then  contact  regional  and  national 
experts  who  have  worked  with  the  species  in  ques- 
tion. Literature  specific  to  the  species  and  habitats 
should  also  be  reviewed.  Team  members  can  learn  a 
great  deal  about  the  general  ecology,  habitat  require- 
ments, and  population  dynamics  of  species  through 


literature  reviews.  This  knowledge  will  also  aid  them 
in  selecting  the  best  sampling  methods  and  analyses 
for  the  study. 


Doing  Preliminary  Field  Work 

While  team  members  are  accomplishing  all  the 
tasks  outlined  in  the  first  meeting,  a  field  trip  into 
the  study  area  should  be  conducted  to  familiarize  all 
personnel  with  habitat  diversity  and  size,  overall 
topography,  soil  types,  elevations  and,  most  impor- 
tantly, access.  Maps  should  be  supplied  to  all  person- 
nel before  this  initial  field  trip.  Biologists  should 
also  meet  with  appropriate  supervisors  and  land 
users  during  this  period  to  avoid  complications  dur- 
ing active  sampling. 

Selecting  Specific  Methods 

After  all  literature  and  museum  reviews  have 
been  completed,  a  series  of  meetings  should  be  con- 
ducted to  determine  the  methods  that  should  be 
used,  the  time  the  monitoring  study  or  inventory 
will  begin,  and  the  length  of  the  study.  At  this  point, 
the  team  leader  should  make  sure  that  all  team  mem- 
bers know  funding  and  schedule  limitations  and  reit- 
erate the  objectives.  The  following  are  the  types  of 
questions  that  should  be  addressed  during  this  type 
of  meeting: 

( 1 )  What  methods  will  be  used  for  each  habitat 
and  animal  type? 

(2)  What  sample  size  will  be  needed  to  meet 
each  objective? 

(3)  Which  groups  of  animals  and  habitats  will 
be  sampled  during  each  season? 

(4)  What  additional  equipment  is  needed  (e.g., 
aircraft  or  four-wheel  drive  vehicles)? 

(  5 )    Can  all  habitat  data  be  collected  by  using 
one  or  two  methods;  will  these  data  be 
sufficient  for  all  wildlife  groups? 

(6)  Can  all  wildlife  groups  be  sampled  at  the 
same  location  for  all  habitat  types? 

(7)  What  are  the  major  sampling  deficiencies? 

(8)  Is  access  to  sample  sites  a  problem? 

(9)  Has  everyone  contacted  the  appropriate 
specialists  in  other  federal  and  state  agen- 
cies? 

(10)  Can  other  ecological  questions  be  answered 
without  increasing  funding  or  time? 

(11)  How  will  the  data  be  analyzed  using  the 
chosen  sampling  methods? 

(12)  How  will  the  data  be  interpreted  and  pre- 
sented? 


Inventory  and  Monitoring 


(13)  How  and  where  will  raw  data,  summary 
analyses,  and  interpretations  be  stored? 

(14)  How  much  preliminary  sampling  is  needed? 

All  of  these  are  extremely  important  questions 
that  must  be  answered  before  an  inventory  or  moni- 
toring study  is  conducted.  Once  these  questions 
are  answered,  the  biologist  or  team  leader  should 
write  an  inventory  or  monitoring  plan  covering  ob- 
jectives, methods,  data  analysis,  and  storage. 


Habitat  Mapping  and  Field  Sampling 


Habitat  Mapping.  Following  this  meeting,  a  field 
trip  should  be  conducted  to  map  habitat  types  and 
locate  sample  sites.  Sample  sites  should  be  randomly 
located  within  the  area  in  question.  If  an  inventory 
or  study  is  conducted  by  habitat  type,  the  entire 
sample  area  should  be  stratified  into  subareas.  Each 
subarea  would  represent  a  habitat  type,  and  the 
number  of  samples  in  each  would  be  proportional  to 
the  amount  occupied  by  that  habitat  type  in  the 
sample  area.  Subareas  may  also  represent  controls 
and  different  intensities  of  land  use.  For  example,  a 
biologist  may  want  to  develop  a  study  to  determine 
the  effects  of  fire  on  vegetation  and  vertebrate 
species.  The  biologist  would  randomly  sample 
vegetation  and  wildlife  within  stratified  subgroups 
such  as  unburned,  burned,  and  partially  burned 
areas.  Chapter  4,  Habitat  Mapping,  provides  guidance 
on  this  important  step. 


Field  Sampling.  Once  sample  sites  are  located,  the 
biologist  should  develop  a  sampling  schedule  with 
deadlines  for  completing  each  sample.  At  this  point, 
preliminary  samples  should  be  taken.  This  allows 
each  biologist  to  test  the  methods  he  or  she  has 
selected  and  to  fine-tune,  method-specific  skills  (e.g., 
bird  transects). 

During  field  work,  the  lead  biologist  should 
conduct  weekly  meetings  to  identify  problems  and 
ensure  quality  control  over  the  specific,  planned 
objectives. 


DATA  ANALYSIS,  INTERPRETATION, 
EVALUATION,  PRESENTATION,  AND 
STORAGE 

Once  data  have  been  collected,  they  need  to  be 
analyzed  and  prepared  in  a  way  that  provides  useful 
information  to  a  decisionmaker.  These  tasks  are  just 
as  important  as  collecting  data;  unfortunately,  many 
biologists  do  not  allow  or  spend  enough  time  on  this 
follow-up.  Section  VI  in  this  book  provides  guidance 
on  these  efforts. 


Data  Analysis,  Interpretation,  and 
Evaluation 

When  field  work  is  completed,  the  team  should 
meet  to  analyze  the  data.  The  type  of  analyses  used 
should  have  been  decided  when  specific  methods 
were  selected,  before  the  inventory  or  study. 

Once  analyzed,  the  data  should  be  interpreted 
and  evaluated.  The  opinions  of  those  biologists  who 
conducted  the  inventory  or  study  should  be  pre- 
served because  the  biologists  are  most  familiar  with 
the  data  and  circumstances  of  collection.  Chapter  37, 
Statistical  Analysis,  provides  guidance  on  using  statis- 
tics; Chapter  38,  Habitat  Evaluation  Systems,  pro- 
vides guidance  on  systems  developed  especially  for 
evaluating  habitat;  Chapter  39,  Evaluation  and  Inter- 
pretation, provides  guidance  on  the  more  subjective 
analysis  procedures;  and  Chapter  40,  Economic  Anal- 
ysis, describes  how  changes  in  wildlife  productivity 
can  be  used  to  predict  changes  in  economic  value. 

Presentation 

A  biologist  must  effectively  communicate  results 
to  managers  and  others.  For  example,  detailed  statis- 
tical analyses  and  complicated  tables  and  figures 
should  not  be  used  to  convey  information  to  man- 
agers. These  types  of  data  analyses  and  displays  are 
best  suited  to  a  specialized  technical  audience.  Sim- 
ple, easily  accessible  tables,  figures,  and  interpreta- 
tions should  be  used  for  more  general  audiences. 

Generally,  documents  or  manuscripts  containing 
inventory  or  study  data  and  interpretations  should 
be  structured  in  the  following  format: 


( 1 )  Introduction,  describing  the  problem  and 
the  reasons  the  biologist  conducted  the  in- 
ventory or  study; 

(  2 )    Methods  Section,  describing  procedures 
used  to  collect  data;  and 

(3)    Results/Discussion/Summary,  describing 
results  of  the  inventory  or  study  and  an  inter- 
pretation of  those  results,  followed  by  a  sum- 
mary of  the  results  and  conclusions  (one  to 
three  sections). 

A  complete  manuscript  contains  all  these  compo- 
nents. This  is  the  best  way  to  ensure  all  elements  of 
the  inventory  or  study  are  preserved  for  future  con- 
siderations, especially  the  biologist's  interpretations 
of  the  data.  Chapter  41,  Written  Communications, 
provides  guidance  on  preparing  written  material. 

In  addition  to  writing  a  separate  manuscript,  the 
biologist  may  be  required  to  orally  present  results 
to  managers,  the  public,  and  others  in  briefings,  staff 


Inventory  and  Monitoring 


meetings,  public  forums,  and  other  gatherings.  Per- 
sonal communications  are  often  the  most  effective, 
and  the  biologist  should  not  overlook  their  impor- 
tance. Chapter  42,  Verbal  Presentations,  provides 
guidance  on  oral  presentations. 


Data  Storage 

Data  stored  in  files  and  on  computers  preserve 
results  of  inventories  and  studies.  Computers  also 
allow  biologists  to  continually  add  information  to  a 
file  in  a  standard  format  and  retrieve  it  easily.  This  is 
especially  important  in  monitoring  studies  where 
data  were  collected  infrequently  or  over  several 
years.  Chapter  36,  Data  Management,  provides  guid- 
ance on  managing  and  storing  data. 


MONITORING  STUDIES 

After  an  inventory  or  study  has  been  conducted 
and  the  results  presented  to  a  manager,  a  decision 
will  be  made  on  some  action.  Such  a  decision  should 
be  based  on  particular  objectives  for  wildlife  and 
other  resources,  and  on  the  predicted  effects  of  the 
chosen  wildlife  resource  alternative.  Predicted 
effects  are  normally  based  on  limited  data;  thus,  the 
biologist  needs  to  monitor  the  wildlife  resources 
to  determine  if  the  objectives  are  being  met.  This 
requires  additional  data  collection  and  may  result  in 
changes  in  management.  Hence,  monitoring  is  a 
cyclic  phenomena  (Figure  1 )  in  which  periodic  data 
collection  is  followed  by  re-evaluation  of  earlier 
management  decisions. 

The  design  and  implementation  of  monitoring 
studies  follow  the  procedure  listed  earlier  in  this 
chapter.  In  general,  monitoring  studies  are  more  site- 
and  problem-specific  than  inventories.  Inventories 
usually  involve  determining  the  species  and  habitats 
in  an  area  and,  less  frequently,  habitats  used  by  cer- 
tain species.  The  accuracy  of  wildlife  uses  of  habitats, 
taken  on  a  one-time  basis,  is  usually  low  because  of 
limited  sample  periods.  Longer  term  monitoring 
studies  better  estimate  fluctuations  in  wildlife  uses  of 
habitat.  To  determine  how  certain  land-use  practices 
affect  wildlife  and  habitat,  the  biologist  should  con- 
duct a  long-term  monitoring  study  rather  than  a  one- 
time inventory.  Several  inventories  should  be  con- 
ducted to  better  understand  the  wildlife  species  and 
habitats  occurring  in  an  area. 

Although  monitoring  may  occur  independently 
of  an  inventory  (e.g.,  bald  eagle  nesting  success  re- 
lated to  dewatering  of  an  aquatic  habitat ),  many 
monitoring  studies  resulting  from  land-use  plans  will 
be  based  on  inventory  results.  Therefore,  biologists 
may  need  to  sample  after  an  inventory  has  been 
conducted  in  an  area. 


Generally,  monitoring  studies  involve  collecting 
wildlife  and  habitat  information  over  time  to  deter- 


mine- 


(1) 


(2) 


(3) 


(4) 


Wildlife  use  of  habitat  components  (e.g.,  pre- 
ferred height  of  sagebrush  for  sage  grouse 
nesting); 

Effects  of  certain  land  uses  on  certain  wildlife 
and  habitats: 

•  individual  wildlife  species  (e.g.,  how  off- 
road  vehicle  [ORV]  use  affects  desert  tor- 
toise [Gopherus  agassizi]  density); 

•  individual  plant  species  (e.g.,  how  intensive 
livestock  grazing  affects  key  mule  deer 
browse); 

•  individual  habitat  components  (e.g.,  how  a 
mining  operation  increases  the  sediment 
load  in  a  trout  stream); 

•  portions  of  the  wildlife  community  (e.g., 
how  a  powerline  affects  migratory  raptors); 

•  entire  wildlife  communities  (e.g.,  how 
water  impoundment  affects  vertebrates  in  a 
riparian  community). 


Species  or  habitat  changes  caused  by  certain 
natural  environmental  conditions  (e.g.,  effects 
of  flooding  on  desert  fishes); 

Accuracy  of  predictive  models  (e.g.,  compari- 
son of  white-winged  dove  [Zenaida  asiatica] 
responses  to  habitat  changes  with  those  pre- 
dicted by  a  U.S.  Fish  and  Wildlife  Service 
Habitat  Suitability  Index  [HSI]  model); 


Off-road  vehicle  use  can  damage  wildlife  habitats. 


Inventory  and  Monitoring 


Broad-scale  Issue  &  Problem 
Identification 


I 


Reduction  of  General  Problems 
to  Specific  Ones 


Preliminary  Predictions  &  Analysis 


Decision    Made 

with  Existing  Data 


i 


Predictions  Made 
Based  on  Decis 


1 


Monitoring  Study 


Adjust 


i 


Predictions 


No   Decision,  No  New 
Data  Collection 


Decision   Adjusted 

Based  on  Monitoring 
Data 


/ 


No   New   Data 
Collection 


Specific    Problem 
Identification 


n 


Decision  not  Made  with 

Existing  Data-Request      -4- 
Additional  Data  (Monitoring 
Study  or  Inventory) 


t 
oritize  Field 

Objectives 


Collection 


1 


Decision  on   Levels  and 
Types  of  Data  Needed 


1 


Data   Collection   (Monitori 
Study  or  Inventory) 


1 


Data  Analysis  and 
Interpretation/ 
Recommendation 


Figure  1.     Inventory  and  monitoring  process. 
8  Inventory  and  Monitoring 


(  5  )    Improving  the  accuracy  of  predictive  models 
(e.g.,  possibly  modifying  the  HSI  model  for 
white-winged  dove,  by  using  on-the-ground 
data  collection); 

(6)  Additional  mitigation  to  protect  wildlife  and 
habitat  in  an  area  (e.g.,  monitoring  data  may 
have  indicated  a  mitigation  plan  was  not  pro- 
tecting nesting  habitat  for  black  hawks  [Bit- 
teogallus  anthracinus].  From  new  data,  new 
stipulations  may  be  recommended  to  protect 
the  raptor's  nesting  habitat.); 

(7)  Additional  habitat  improvement  to  benefit  a 
species  or  habitat  of  concern  (e.g.,  a  water 
development  program  may  not  be  increasing 
pronghorn  herds  in  an  area  as  predicted. 
From  the  new  monitoring  data,  the  plan  will 
probably  need  to  be  modified.). 

Some  of  the  above  objectives  related  to  improv- 
ing a  biologist's  predictive  ability  or  the  models 
used.  Thus,  monitoring  can  serve  dual  purposes:  (  1 ) 


determining  whether  management  objectives  are 
being  met  and  (  2  )  improving  a  biologist's  predictive 
capabilities.  This  approach  to  management  and  moni- 
toring is  variously  termed  "cyclic-incrementalism" 
(Bailey  1982),  "adaptive-environmental  assessment" 
(Holling  1978),  "adaptive-management"  (Barrett  and 
Salwasser  1982)  or,  more  simply,  "common-sense 
management"  (Barrett  and  Salwasser  1982)  or  "mud- 
dling through"  (Bailey  1982).  These  approaches  are 
fundamental  to  monitoring  and  are  discussed  in 
more  detail  in  many  chapters  throughout  this  book. 


SUMMARY 

In  summary,  biologists  are  strongly  advised  to 
follow  the  logical  thought  processes  developed  in 
this  chapter.  Successful  inventories  and  monitoring 
studies  do  not  come  easily.  Only  through  careful 
planning,  precise  problem  and  objective  identifica- 
tion, and  complete  compilation  of  results,  will  an 
inventory  or  monitoring  study  be  successful. 


Inventory  and  Monitoring 


LITERATURE  CITED  HOLLING,  C.S.,  ed.  1978.  Adaptive  environmental  assess- 

ment and  management.  Vol.  3,  International  Series  on 
Applied  Systems  Analysis.  John  Wiley  and  Sons,  New 
York,  NY.  377pp. 
SHORT,  H.L.  1982.  Techniques  for  structuring  wildlife 
guilds  to  evaluate  impacts  on  wildlife  communities. 
BAILEY,  J.A    1982.  Implications  of  "muddling  through"  for  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.,  Spec.  Rep.  244. 

wildlife  management.  Wildl.  Soc.  Bull.  10:363-369.  34pp. 

BARRETT,  R.H.  and  H.  SALWASSER.  1982.  Adaptive  man-  THOMAS,  J  W,  ed.  1979.  Wildlife  habitats  in  managed 

agement  of  timber  and  wildlife  habitat  using  dynast  forests — the  Blue  Mountains  of  Oregon  and  Washing- 

and  wildlife-habitat  relationship  models.  Western  ton.  U.S.  Dep.  Agric,  For.  Serv.  Agric.  Handbook  553- 

Assoc,  of  Fish  and  Wildl.  Agencies,  Proc.  62:350-365.  512pp. 


10  Inventory  and  Monitoring 


DATA  TYPES 


K.  Bruce  Jones 


Editor's  Note:  A  good,  well-planned  and  organized 
inventory  or  monitoring  program  cannot  be  based 
on  cookbook  methodology1.  The  biologist  must  use 
considerable  judgment  in  identifying  problems, 
specifying  objectives,  setting  priorities,  and  choos- 
ing methodology.  The  choice  of  data  to  be  collected 
is  critical.  Yet  the  biologist  can  collect  an  almost 
unlimited  number  of  types  of  data  This  chapter 
provides  an  overview  of  data  types  and  appropriate 
situations  for  collecting  a  particular  type. 


U.S.  Bureau  of  Land  Management 
Phoenix  Training  Center 
Phoenix,  AZ  8S015 


INTRODUCTION 

After  identifying  specific  objectives  and  ques- 
tions for  an  inventory  or  monitoring  study,  a  biolo- 
gist must  decide  on  the  types  and  levels  of  data 
needed.  This  decision  is  often  a  confusing  part  of  the 
inventory  and  monitoring  process.  Many  different 
types  of  data  can  be  collected  to  answer  specific 
questions.  In  addition,  the  intensity  of  data  collection 
must  be  determined.  For  example,  an  objective  is 
established  to  determine  the  influence  of  a  new 
water  development  on  an  area's  pronghorn  (Antilo- 
carpa  americana)  population.  Do  we  need  to  deter- 
mine the  density  of  pronghorns  in  the  area  before 
and  after  the  development?  Do  we  need  a  control? 
Would  relative  abundance  rather  than  density  an- 
swer the  question?  Decisions  made  on  sampling  de- 
sign and  data  types  are  extremely  important.  In  the 
previous  example,  why  spend  a  great  deal  of  time 
and  money  collecting  density  data  when  abundance 
information  will  suffice?  If  a  control  was  not  used, 
how  will  we  know  if  the  observed  changes  in 
pronghorn  abundance  result  from  our  habitat 
improvement? 

This  chapter  discusses  different  types  of  data 
and  sampling  considerations,  and  provides  some 
basic  guidelines  for  making  decisions  on  types  of 
data  and  intensity  of  data  collection  to  use.  It  also 
discusses  some  procedures  that  are  used  to  analyze 
existing  data. 


SAMPLING  CONSIDERATIONS 

Throughout  this  chapter,  I  highlight  some  basic 
data  types  and  sampling  intensities.  Some  basic  fac- 
tors related  to  sampling  design  and  intensity  apply  to 
all  types  of  data  collection. 

Before  a  biologist  starts  collecting  data,  a  deci- 
sion must  be  made  on  sampling  intensity.  The  num- 
ber of  samples  required  will  depend  on  the 
variability  of  animals  and  habitats  to  be  measured 
and  limitations  of  sampling  techniques  used.  For 


Current  Address:  Office  of  Endangered  Species,  U.S.  Fish  and 
Wildlife  Service,  Washington,  DC  20240 


Data  Types 


11 


example,  homogeneous  habitats  require  fewer  sam- 
ples because  they  are  less  variable.  Conversely,  het- 
erogeneous habitats  require  more  samples  because 
they  are  more  variable.  Sampling  techniques  and 
habits  of  animals  also  affect  sample  size.  For  exam- 
ple, bird  transects  detect  far  fewer  birds  in  closed 
habitats  than  in  open  ones.  Therefore,  to  make  the 
probability  of  detection  equal  in  the  two  habitats, 
sample  size  must  be  increased.  Also,  certain  bird 
species  are  secretive,  requiring  more  intensive  sam- 
pling to  adequately  cover  such  species. 

A  number  of  minimum  sample  size  equations 
have  been  developed  for  various  vertebrates.  Most  of 
these  are  based  on  population  variability  and  animal 
detectability.  Seber  (1973),  Smith  et  al.  (1975),  and 
Gysel  and  Lyon  (1980)  provide  formulas  and  discuss 
how  minimum  sample  size  is  determined. 

Because  most  inventories  and  monitoring  stud- 
ies are  restricted  by  funds  and  personnel,  biologists 
are  often  confronted  with  deciding  on  where  to 
place  samples.  Generally,  samples  should  be  placed 
randomly  within  habitats  or  areas.  Cause-and-effect 
studies  require  biologists  to  place  samples  within 
preselected  areas  (stratified  sampling).  However,  bi- 
ologists should  select  random  sites  within  these  pre- 
selected areas,  assuring  that  area-related  variability 
is  considered.  Random  sites  can  be  selected  by  as- 
signing numbers  to  grids  placed  over  a  map  of  the 
area  in  question.  Numbers  can  be  selected  from  ran- 
dom numbers  tables.  In  monitoring  studies  where 
the  biologist  wants  to  determine  animal  or  habitat 
changes  before,  during,  and  after  a  specific  land  use 
(e.g.,  prescribed  burn),  controls  should  be  estab- 
lished. Without  control,  there  is  no  way  to  deter- 
mine whether  specific  land  use  has  affected  habitat 
structure.  For  example,  improvements  in  vegetation 
structure  on  an  allotment  with  a  three-pasture,  rest- 
rotation  system  may  result  from  higher-than-normal 
precipitation  over  the  past  5  years  rather  than  from 
the  system.  In  this  example,  control  samples  taken 
outside  the  grazing  system  but  where  similar  physi- 
cal conditions  (vegetation,  soils,  to  name  two)  exist 
would  allow  the  biologist  to  account  for  other  possi- 
ble causes  of  observed  vegetation  structure  changes. 

In  cases  where  money  and  access  restrict  sam- 
pling, biologists  may  have  to  cluster  samples  rather 
than  spread  them  over  the  entire  habitat.  This  will 
reduce  the  probability  that  habitat  variability  was 
adequately  sampled.  For  example,  a  biologist  wants 
to  compare  vegetation  structure  between  pinyon- 
juniper  (Pinus-Juniperus  sp.)  and  chaparral,  but  is 
limited  to  two  samples  in  each  habitat  type  and  one 
habitat  is  five  times  larger  than  the  other.  Because 
habitat  size  is  not  considered  in  allocating  samples, 
the  biologist  may  not  be  able  to  determine  which 
habitat  has  greater  vegetation  structure,  even  if  the 
data  suggest  large  differences.  If  samples  of  habitat 
and  animal  variables  are  taken  from  the  same  site, 


then  the  biologist  can  determine  relationships  be- 
tween animals  and  their  habitats. 

The  types  and  numbers  of  animals  and  habitat 
characteristics  sampled  directly  depend  on  objec- 
tives established  for  an  inventory  or  study.  This  will 
determine  the  complexity  and  intensity  of  sampling. 
Intensity  can  range  from  simple  verification  of  one 
animal  in  one  habitat  to  large-scale  vertebrate  sam- 
ples taken  in  many  habitats  over  several  years.  (Ta- 
ble 1 )  provides  some  examples  of  different  sampling 
intensities,  given  some  specific  objectives.  In  addi- 
tion, Green  (1979)  provides  an  excellent  discussion 
of  sampling  design  and  data  intensity. 

Generally,  biologists  are  confronted  with  three 
types  of  data:  animal,  habitat,  and  animal/habitat. 

ANIMAL  DATA  COLLECTION 

Animal  data  are  characterized  by  two  basic 
types:  ( 1 )  individual  species  and  (  2  )  groups  of  spe- 
cies. Each  of  these  are  detailed  below. 

Individual  Species  Measurements 

There  are  many  types  of  individual  species  data, 
which  are  summarized  in  (Table  2).  Although  biolo- 
gists will  not  likely  collect  all  these  types  of  data, 
they  will  need  some  knowledge  of  them  to  interpret 
results  of  contracts  and  other  existing  studies. 

Presence.  The  simplest  measurement  of  individual 
species  populations  is  presence/absence.  In  many 
cases,  the  only  objective  is  to  verify  use  of  a  habitat 
or  site-specific  area  by  a  species  (e.g.,  bald  eagles 
[Haliaeetus  leucocephalns]  along  a  riparian  stretch). 
These  types  of  data  help  determine  potential  impacts 
to  a  species  and  its  habitat  from  development  or 
natural  factors  such  as  fire.  These  data  also  help 
focus  the  biologist's  efforts  on  site-specific  areas, 
especially  those  of  threatened  or  endangered  species. 

One-time  verification  of  a  species  in  an  area 
may  only  be  accurate  for  a  short  time,  especially 
when  a  species  uses  specific  habitats  only  during  cer- 
tain seasons.  For  example,  black  hawks  (Buteogallus 
anthracinus)  may  nest  in  a  riparian  area  during  a 
wet  year  when  aquatic  prey  is  abundant,  but  fail  to 
nest  in  the  same  area  in  the  following  year  because 
of  drought-reduced  prey  abundance.  Another  con- 
cern in  using  presence/absence  data  involves  one- 
time failure  to  identify  a  species  in  an  area.  Several 
factors  may  account  for  species  not  being  verified  in 
an  area:  unskilled  observers,  poor  sampling  tech- 
niques, weather  conditions  during  sampling,  and  ab- 
sence of  that  species  in  the  area  during  the  sampling 
period.  The  biologist  should  be  very  cautious  in 
using  one-time,  presence/absence  data  to  assess  an 
area's  fauna.  Multiple-year  verification  of  species 
in  habitats  is  preferred. 


12 


Data  Types 


Table  1.     Examples  of  specific  objectives  carried  through  the  design  process. 


Specific  Objective/Question 


Sampling  Questions 


Solution 


Determine  how  a  new  water 
development  will  affect  mule 
deer  use  of  key  browse  in  the 
area. 


What  are  the  key  browse  in  the 
area? 

How  does  water  development 
affect  use? 

How  did  mule  deer  use  in  the 
area  compare  before  installing 
the  new  development  versus 
after? 

Did  other  changes  occur  that 
might  account  for  differences 
before  and  after  development? 


Since  I  did  not  know  the  size  of  the 
area  (several  thousand  acres  of  contin- 
uous sagebrush-grass),  I  systematically 
located  my  sample  points  along  four 
lines,  out  from  the  water  development  at 
right  angles  (cross  shaped).  Samples 
were  located  at  100  m  (300  ft)  intervals 
along  these  lines  (10  samples/line). 
Angle  of  the  first  line  was  randomly  se- 
lected. At  each  point,  I  measured  key 
browse  by  a  modified  Cole  technique. 

Since  cows  graze  the  area  year-round,  I 
could  not  determine  whether  browse 
were  being  used  by  livestock  or  deer. 
To  assess  relative  food  consumption,  I 
took  deer  and  cow  fecal  samples  at 
each  point.  To  detemine  deer  use, 
I  established  20  points  along  each  line 
(50  m  [165  ft]  apart,  every  other  point  at 
a  browse  utilization  station),  where  I 
read  0.025  ha  (0.01  a.)  deer  pellet 
plots.  I  sampled  both  deer  and  browse 
utilization  twice  before  construction  of 
the  development,  30  days  after  devel- 
opment, and  then  once  each  session 
for  3  years. 


Determine  if  a  riparian  improve- 
ment plan  will  increase  the 
number  of  nesting  black  hawks 
and  nesting  success  on  Yankee 
Creek. 


What  was  the  black  hawk  nest- 
ing density  like  before  imple- 
menting the  plan? 

What  was  the  average  number 
of  young  produced  at  nests 
before  implementing  the  plan? 

What  was  the  black  hawk  nest- 
ing density  1,  3,  5,  10,  and  15 
years  after  implementation  of 
the  plan? 

How  many  new  nests  were  built 
1,  3,  5,  10,  and  15  years  after 
implementation  of  the  plan? 

What  was  the  impact  of  the  plan 
on  primary  black  hawk  prey? 

Did  factors  other  than  those 
provided  by  the  plan  affect  the 
results? 


In  order  to  determine  the  number  of 
black  hawk  nests  before  and  after  im- 
plementation of  the  plan,  I  counted 
all  nests  within  cottonwood-willow 
stands  along  the  creek.  To  estimate 
increases  in  the  amount  of  tree  canopy 
in  periods  after  implementation,  I  ob- 
tained infrared,  low-level  photos  of  the 
area  every  other  year  for  15  years 

To  determine  the  effects  of  the  plan  on 
the  fish  prey  base,  I  established  perma- 
nent sampling  stations  where  I  sampled 
fish  before  and  1,  3,  5,  10,  and  15 
years  after  the  plan  was  implemented. 
This  included  sampling  riffles,  runs  and 
pools,  and  sites  near  known  nests. 
Samples  were  taken  in  the  spring. 

Since  there  were  over  15  nests,  nest 
success  at  all  of  them  could  not  be 
monitored.  I,  therefore,  randomly  se- 
lected three  nests  to  monitor  at  1 ,  3,  5, 
10,  and  15  years  after  implementation 
to  determine  nest  success  and  habitat 
uses.  New  nests  were  plotted  against 
changes  in  vegetation.  Before  and  1,  3, 
5,  10,  and  15  years  after  implementa- 
tion, I  determined  the  total  number  of 
active  nests  in  the  creek.  I  also  moni- 
tored weather  and  flow  conditions  in  the 
creek  for  years  previously  listed  via  a 
hygrothermograph  and  a  U.S.  Geologi- 
cal Survey  gauging  station,  respectively. 


Data  Types 


13 


Table  2.     Summary  of  data  types,  some  procedures  and  limitations. 


Data  Type 

Method/Expression 

Application/Limitation 

1.    Individual  Species 

A.  Presence/absence 

1.  Searches  (visual) 

List  of  species  in  an  area  or 

Used  to  verify  key  species  in  certain 

habitat. 

areas  or  habitats.  Multi-year  samples 
needed  for  species  that  migrate  or  use 
habitats  seasonally.  To  compare  two 
or  more  areas,  sampling  intensity  must 
be  equal.  Random  or  systematic 
sampling. 

2.  Capture  and  harvest 

Same  as  above. 

As  above,  along  with  sampling  tech- 
niques; must  be  similar  to  compare  two 
or  more  areas  or  habitats  (e.g.,  similar 
depth  traps).  Indirect  sampling 
methods. 

3.  Sign 

Same  as  above. 

Feces,  nests,  and  other  physical  evi- 
dence that  a  species  occupies  an  area 
or  habitat.  Indirect  method.  Sampling 
limitations  as  above. 

4.  Auditory 

Same  as  above. 

Direct  and  indirect  methods  of  verifying 
species  in  a  habitat  or  area  via  sound. 
Sampling  limitations  as  above. 

5.  Museum  or  other  exist- 

Animal and  plant  records  by 

Indirect  method  of  verifying  species  in 

ing  records 

county,  township  and  range, 

habitats  or  areas.  Data  highly  variable. 

sections,  and  habitat  type  with 

Many  records  only  at  county  level. 

general  notes.  From  museums, 

Some  animals/plants  misidentified. 

literature,  and  individual  re- 

cords. Data  same  as  above. 

B.  Abundance  (indexes 

of  population  size) 

1.  Counts/searches 

Number  of  animals/unit  effort 

Direct  method  of  determining  abun- 

(usually time). 

dance  of  animals  in  certain  areas  or 
habitats.  Some  methods  such  as  road 
riding  are  inexpensive.  Data  good  for 
comparing  different  sites  and  impacts 
of  land  use  on  wildlife  and  habitats  if 
controls  are  used.  Good  techniques  for 
monitoring  if  permanent  sites  are  estab- 
lished. Subject  to  observer  biases, 
and  good  only  for  readily  observable 
species. 

2.  Capture/trapping 

Number  of  animals  caught/unit 

A  series  of  individual  methods  for  deter- 

time. 

mining  species  abundance.  Methods 
usually  specific  to  certain  species 
groups  (e.g.,  pit-fall  traps  and  lizards). 
Less  observer  bias  than  direct 
searches.  Some  procedures  excel  in 
verifying  secretive  species.  Relatively 
cost-effective.  Excellent  procedures  for 
monitoring,  provided  sampling  is 
systematic. 

14 


Data  Types 


Table  2.     Summary  of  data  types,  some  procedures  and  limitations  (continued). 


Data  Type 


Method/Expression 


Application/Limitation 


3.  Line  searches 


4.  Sign 


5.  Auditory 


Similar  to  searches  within  an 
area  or  habitat,  but  generally 
expressed  as  number  of  ani- 
mals seen/line  distance  (e.g., 
birds  per  mile). 


Includes  scent  stations,  nests, 
number  of  tracks  or  pellet 
groups.  Data  are  expressed  in 
number  of  sign/unit  area  or 
effort. 


Generally  the  recording  of  dif- 
ferent animal  calls/unit  time. 


Direct  methods  of  determining  abun- 
dance of  animals  in  certain  habitats  or 
an  area  as  a  function  of  distance  and 
time  spent  searching.  When  lines  are 
permanently  located,  these  methods 
are  excellent  for  long-term  monitoring. 
Similar  to  searches,  these  have  ob- 
server-associated biases. 

Commonly  used  as  indexes  of  certain 
big  game  abundance.  Also  used  for 
lizards,  snakes,  tortoises,  and  small 
mammals.  Most  of  these  indexes  pro- 
duce highly  variable  data.  Recommend 
these  procedures  only  when  technigues 
cannot  be  used. 

Commonly  used  for  small  birds,  raptors, 
frogs,  toads,  and  predators.  Often  diffi- 
cult to  count  or  estimate  the  number 
of  different  individuals.  Generally  used 
for  verifying  a  species  in  an  area. 


C.  Density/spatial 
1.  Counts/searches 


2.  Line  searches 


3.  Mark  and  recapture 


4.  Depletion 


5.  Capture/trapping 


Number  of  animals/unit  area. 


Animals/unit  area  as  a  function 
of  expected  versus  observed 
(Chi-sguared  test). 


Population  size/area  as  a  func- 
tion of  marked  versus  unmarked 
animals. 


As  above,  marked  animals  are 
considered  removed. 


Number  of  animals/unit  area. 


Common  procedure  used  for  key  spe- 
cies including  big  game  and  threatened 
or  endangered  species.  Labor-intensive 
and  effective  only  on  easily  verified 
species.  If  combined  with  point  proce- 
dure (e.g.,  variable  circle  used  for 
birds),  procedure  is  less  labor-intensive. 

Includes  methods  with  horizontal  belts 
along  a  line  (e.g.,  Emlen's  bird  tran- 
sects). Used  for  birds  and  other  spe- 
cies easily  observed.  Generally 
inexpensive  to  run,  but  procedures 
reguire  expert  observers.  Not  effective 
in  small  discontinuous  habitats  (e.g., 
riparian). 

Several  different  procedures.  Some 
one-time  sampling;  others  multiple  sam- 
ples. Commonly  used  to  monitor  terres- 
trial and  aguatic  populations.  Procedure 
allows  monitoring  of  individual  animals. 
Generally  labor-intensive. 

Similar  to  above.  Some  procedures 
involve  physical  removal  of  animals.  Re- 
moval usually  on  small,  easily  captured 
animals  (e.g.,  fish).  Generally  labor- 
intensive.  Best  results  with  closed 
populations. 

Live  and  kill  traps.  Used  in  intensive 
studies  where  population  size  is 
needed.  Most  effective  on  readily 
trapped  animals.  Also  used  in  conjunc- 
tion with  mark-and-recapture  proce- 
dures. Time-  and  labor-intensive. 


Data  Types 


15 


Table  2.     Summary  of  data  types,  some  procedures  and  limitations  (continued). 


Data  Type 

Method/Expression 

Application/Limitation 

6.  Telemetry 

Animal  movement/unit  time  or 
percentage  ot  time  spent  in  an 
area  or  habitat,  performing 
certain  behaviors. 

Effective  method  for  determining  behav- 
ior and  movement  patterns.  Initially 
labor-intensive.  Generally  limited  to 
priority  species. 

D.  Population  structure 

1 .  Sex  ratios 

2.  Age  ratios 

Ratio  males  to  females. 

Ratio  of  animals  in  certain  age 
groups. 

Some  animals  have  a  sex  ratio  range 
that  represents  a  fit  population.  Num- 
bers outside  these  ranges  can  indicate 
a  stressed  population.  Generally  used 
for  big  game.  Less  accurate  for  lower 
vertebrates  because  their  sex  ratios 
vary  with  different,  natural  environmen- 
tal conditions. 

Similar  to  sex  ratios.  Other-than-normal 
ratios  of  animals  in  certain  age  groups 
can  indicate  a  stressed  population. 
Generally  used  for  animals  whose  age 
structure  does  not  vary  with  environ- 
mental fluctuations.  Be  cautious  in  us- 
ing these  in  lower  vertebrates  for 
reasons  listed  for  sex  ratios. 

E.  Productivity 

(recruitment  rates) 

Generally,  reproductive  effort 
(number  of  recruits). 

Commonly  used  for  big  game  and  other 
species  whose  birth  or  recruitment 
rates  can  be  measured.  Generally,  la- 
bor-intensive. Many  populations  cyclic. 

F.  Population  condition 

Generally,  percentage  of  ani- 
mals in  certain  condition 
classes  (e.g.,  rankings  of  fat 
storage). 

Animals  conditions  used  as  indicators 
of  population  fitness  and  stress.  Limited 
to  priority  species.  Labor-intensive. 

II.  Wildlife  Communities 

A.  Species  richness 

1.  Presence/absence 
(habitat  type) 

2.  Presence/absence 
(habitat  components) 

List  of  species  in  a  habitat  type 
or  total  number  species/habitat 
type. 

Similar  to  above  but  expressed 
by  habitat  component  (e.g., 
number  of  species  using  rock 
for  feeding  substrate). 

Lists  complied  from  on-site  data  collec- 
tion and  existing  information.  Can  be 
used  to  compare  habitat  species  rich- 
ness. Commonly  used  for  land-use 
planning,  with  limited  utility  on  site-spe- 
cific assessments. 

Used  to  identify  critical  components  of 
habitat  based  on  species'  uses.  Also 
used  to  assess  impact  of  habitat  com- 
ponent loss  on  groups  of  wildlife.  Guild- 
ing  is  a  process  for  grouping  animals 
(see  text).  Use  for  land-use  planning, 
but  cautiously  for  site-specific  analysis. 

B.  Community  abundance 

Overall  abundance  of  a  taxo- 
nomic  group  (e.g.,  small  mam- 
mals) in  habitats.  Data 
expression  varies  with  sampling 
methods  (see  Individual  Spe- 
cies section). 

Used  to  assess  value  of  habitats  (e.g., 
pinyon-juniper  versus  chaparral)  for 
certain  groups  of  wildlife.  More  time- 
and  labor-intensive  than  presence/ 
absence.  Problems  similar  to  those  for 
sampling  individual  species  abundance 
(see  Individual  Species  section). 

16 


Data  Types 


Table  2.     Summary  of  data  types,  some  procedures  and  limitations  (continued). 


Data  Type 


C.  Community  density 


Method/Expression 


As  above  but  expressed  as 
animals/unit  area. 


Application/Limitation 


As  above.  Subject  to  problems  associ- 
ated with  individual  species  densities. 


D.  Biomass 


Weight  or  mass  (gms)  a  spe- 
cies contributes  to  the  overall 
community.  Often  expressed  in 
ratios. 


Used  to  estimate  species  contributions 
to  a  community  based  on  mass.  Also 
estimates  trophic  levels  and  food  avail- 
able to  certain  predators.  Limited  utility 
in  most  management-related  studies. 


E.  Species  similarities 
(similarity  coefficients) 


Index  generally  between  0  (no 
similarity)  and  1  (identical). 
Single  value. 


Used  to  compare  species  assemblages 
on  two  different  sites.  Calculations  are 
simple  and  do  not  require  abundance 
(presence/absence  only).  Some  formu- 
las emphasize  similarity;  others,  differ- 
ences. Excellent  for  comparing  wildlife 
communities. 


F.  Species  diversity  and 
habitat  use  diversity 


Index  generally  ranging  from  0 
(no  diversity)  to  10  (high  diver- 
sity). Certain  formulas  empha- 
size evenness  of  abundance 
shared  by  species  on  a  site 
(log10);  others,  richness  of  spe- 
cies (loge). 


Commonly  used  to  compare  species 
diversity  between  sites  or  the  diversity 
of  habitats  used  by  species  (habitat  use 
diversity  or  niche  breadth).  Abundance 
or  density  data  needed  for  calculations. 
Subject  to  biases  associated  with  abun- 
dance and  density.  When  sampling  is 
conducted  over  long  periods,  data 
provide  insight  into  changes  in  species 
dominance  relative  to  a  management 
prescription. 


G.  Species  overlap 


Like  similarity  coefficients,  but 
value  range  of  0  to  1  is  based 
on  shared  species  abundance 
versus  total  abundance.  Horn- 
Overlap  is  the  most  commonly 
used  procedure. 


Used  to  assess  differences  in  animal 
community  structure  between  habitats. 
Abundance  or  density  data  required. 
Also  assesses  species  contributions  on 
one  site  versus  those  on  other  sites. 
Biases  similar  to  those  of  abundance 
and  density  sampling. 


Habitat  (characteristics,  com- 
position, and  structural  condi- 
tions) 

A.  Point 


Percentage  of  plant  cover,  sub- 
strate type,  plant  frequency, 
plant  and  other  horizontal  and 
vertical  attributes  of  habitat. 


Used  for  gross  characterizations  of 
habitats  and  areas.  When  points  are 
permanently  located,  changes  in  habitat 
can  be  determined.  Often  combined 
with  a  line  to  determine  percentages. 
Generally  poor  for  grasses  and  forbs. 


B.  Line 


As  above,  but  as  a  function  of 
distance  along  a  line. 


Extremely  useful  for  characterizing  plant 
composition  and  frequency  and  habitat 
components  for  a  habitat  or  area.  Gen- 
erally quick  and  easy  to  read.  Excellent 
for  monitoring  changes  if  permanently 
marked.  Most  effective  on  habitats 
dominated  by  shrubs. 


C.  Plot 


Number  of  a  habitat  compo- 
nent/unit area.  Also,  weight 
contributed  by  a  plant  species/ 
unit  area. 


Commonly  used  for  grass  and  forb 
density,  composition,  and  biomass  on 
an  area  or  in  a  habitat.  Labor-intensive 
Not  generally  effective  for  woodland 
and  forest  habitats.  Good  for  monitoring 
rangeland  if  permanently  located. 


Data  Types 


17 


Table  2.     Summary  of  data  types,  some  procedures  and  limitations  (concluded). 


Data  Type 

Method/Expression 

Application/Limitation 

D. 

Plotless 

Number  of  plants  or  a  certain 

Fast  and  effective  way  of  determining 

habitat  component/unit  area  as 

absolute  densities  of  trees,  shrubs,  and 

a  function  of  distance  from  a 

other  habitat  components  such  as 

center  point. 

snags.  Structural  characteristics  of  hab- 
itat components  are  also  usually  mea- 
sured. Point-center-quarter  method  is  a 
commonly  used  plotless  method.  Excel- 
lent for  monitoring  if  points  are  perma- 
nently located. 

IV.  Animal '/Habitat 

A. 

Animal/habitat  indexes 

Usually  expressed  as  a  quality 

Indexes  are  suited  for  broad-scale  land- 

index  value  ranging  from  0 

use  planning,  especially  when  using  a 

(poor  quality  habitat)  to  1  (best 

generic  model.  Can  be  used  for  more 

possible  habitat).  Data  also 

site-specific  application  if  area-specific 

expressed  as  quality  (suitability) 

data  are  used  to  develop  the  model. 

index  x  the  area  =  habitat 

Habitat  Suitability  Index  models  use  a 

value. 

series  of  ecological  relationships  to 
develop  models  for  individual  species. 
Habitat  Relation  models  rank  serai 
stages  of  habitats  for  individual  species 
relative  to  performing  important  ecologi- 
cal functions. 

B. 

Animal/habitat  correlations 

Data  expression  varies  with 

Procedures  that  use  empirical  data  to 

procedures  (see  below). 

determine  relationships  between  ani- 
mals and  habitats. 

1.  Correlation  and  regres- 

Data expressed  as  a  linear  or 

If  animal  and  habitat  data  are  collected 

sion 

nonlinear  (e.g.,  polynomial) 

on  similar  sites,  these  equations  are 

relationship,  usually  between 

ideal  for  area-specific  monitoring  and 

animal  population  and  one 

development  of  predictive  equations  for 

or  several  habitat  variables. 

management  assessment.  SPSS  and 
BMDP  software  provide  these  types  of 
analysis.  With  practice,  these  are  easily 
developed. 

2.  Principle  components 

Data  generally  expressed  as 

Valuable  tool  for  illustrating  differences 

location  along  axes  (e.g., 

in  animal  uses  of  habitat.  Allows  the 

X,Y,Z).  Locations  on  axes  repre- 

biologists to  compress  several  habitat 

sent  combined  habitat 

variables  into  two  or  three  axes  which 

variables. 

then  are  easily  illustrated. 

18 


Data  Types 


Abundance  and  Density.  Abundance  and  density 
measurements  have  been  widely  used  in  wildlife 
inventories  and  monitoring  studies.  Generally,  there 
are  three  types  of  abundance  measurements:  total 
number  of  animals  in  a  population;  number  of 
animals  per  unit  area  (density);  and  abundance  and 
density  of  one  population  relative  to  another 
(Caughley  1977).  Rarely  will  biologists  measure  the 
absolute  number  of  animals  in  a  population  or  area 
and  need  actual  numbers  or  density.  Therefore,  most 
wildlife  inventories  and  monitoring  studies  use 
estimates  of  population  size  and  density.  Estimates 
are  usually  accomplished  by  indexes  that  are 
correlated  with  population  size  and  density 
(Caughley  1977;  Eberhardt  1978). 

Estimates  of  population  size  or  absolute  abun- 
dance are  commonly  used  in  wildlife  inventories. 
Because  these  estimates  are  indexes  of  populations 
and  not  ascribed  to  units  of  areas,  they  are  usually 
termed  abundance.  Abundances  include  animal  num- 
bers obtained  from  direct  observation,  sign  (tracks, 
etc.),  and  captures.  Certain  biases  are  associated  with 
each  of  these  methods.  Caughley  (1977)  and  Eber- 
hardt (  1978)  provide  excellent  summaries  of  these 
biases;  biologists  should  consult  these  and  other 
sources  before  deciding  a  method.  Estimates  of  abun- 
dance include,  but  are  not  limited  to — 

•  number  of  animals  seen  per  hour  of  observation 
(also  termed  time-restraint); 

•  number  of  animals  seen  per  linear  distance  (e.g., 
raptors  seen  per  mile  of  powerline ); 

•  number  of  animals  trapped  per  24  hours; 

•  number  of  tracks  counted  per  hour  of  observa- 
tion; 

•  number  of  calls  heard  per  hour  (e.g.,  frogs  and 
owls). 

All  these  can  be  measured  without  regard  to  area. 
Therefore,  they  are  estimates  that  presumably  are 
correlated  with  population  size. 

Relative  abundance  generally  refers  to  the  con- 
tribution a  species  makes  to  the  total  abundance 
of  that  wildlife  community.  For  example,  if  three  ob- 
servations of  species  A  and  one  observation  of  spe- 
cies B  were  made,  then  species  A  makes  up  75% 
of  the  abundance  and  species  B,  25%.  These  relative 
abundances  are  expressed  as  0.75  and  0.25,  respec- 
tively (see  Hendrickson  et  al.  1980  for  excellent  fish 
examples). 

Density  refers  to  animal  numbers  per  unit  area 
(Caughley  1977).  Estimates  of  density  are  obtained 
in  ways  similar  to  those  acquired  for  abundance 
estimates,  but  they  are  applied  to  a  naturally  or  artifi- 
cially ascribed  area  (e.g.,  habitat  boundary  or  marked 


hectare  or  acre  plot,  respectively).  For  example,  a 
biologist  estimates  lizard  density  by  searching  a  pre- 
marked  area  and  counting  all  lizards  seen  (Bury 
1982).  Similarly,  bird  density  can  be  estimated  by 
using  a  variable  plot  method  where  the  number  of 
birds  are  counted  within  1/100  ha  (0.025  a.)  circular 
plots  (see  Ralph  and  Scott  1981,  for  examples). 

Estimates  of  density  are  also  obtained  by  count- 
ing the  number  of  animals  seen  within  prescribed 
belts  established  along  a  fixed  line.  By  using  proba- 
bility statistics,  animal  densities  can  be  predicted 
in  prescribed  belts  outside  areas  adjacent  to  the 
transect  line.  Perhaps  the  most  commonly  used  line/ 
area  transect  is  that  developed  by  Emlen  ( 1971 ) 
for  birds.  Emlen's  bird  transect  has  been  modified  to 
reduce  sampling  biases,  including  statistical  applica- 
tions available  on  computer  software.  Ralph  and 
Scott  ( 1981 )  provide  a  series  of  papers  about  bird 
sampling. 

When  comparing  estimates  of  population  abun- 
dance and  density,  the  biologist  should  consider 
the  number  of  samples  to  be  allocated  to  each  area. 
This  is  especially  important  when  comparing  densi- 
ties. To  compare  two  sites,  biologists  should  com- 
pensate for  area-related  variability  by  allocating 
sample  size  relative  to  proportions  of  area.  For  exam- 
ple, if  area  A  is  1,000  ha  ( 2,500  a. );  area  B  is  5,000 
ha  (2,500  a.);  and  area  C  is  10,000  ha  (250,000  a.), 
then  area  A  would  have  1/10  the  number  of  samples 
of  area  C,  and  area  B,  1/2  the  number  of  area  C. 

Depletion  sampling  and  mark-and-recapture  are 
other  methods  commonly  used  to  estimate  popula- 
tion size  and  density.  Generally,  removal  methods 
involve  removing  animals  during  two  or  more  sam- 
ple periods,  where  only  two  factors  are  used:  popula- 
tion size  N  and  capture  probability.  Population  data 
are  obtained  by  comparing  expected  captures  to 
actual  captures,  generally  by  a  goodness-of-fit  test 
(Chi-squared  test).  One  problem  with  this  method  is 
that  it  assumes  all  animals  have  an  equal  probability 
of  capture.  In  many  cases,  this  is  not  true.  For  exam- 
ple, the  probability  of  capturing  large  fish  is  greater 
than  small  fish  when  electroshocking  is  used.  Otis  et 
al.  ( 1978)  and  White  et  al.  (1982)  have  developed 
more  generalized  removal  models  that  can  account 
for  unequal  capture  probability.  Removal  procedures 
are  common  in  estimating  the  number  of  fish  in  a 
stream,  primarily  because  the  biologist  is  sampling  a 
relatively  closed  system.  White  et  al.  (1982)  provide 
a  good  review  of  a  number  of  removal  methods. 

Mark-and-recapture  techniques  involve  marking 
animals  and  then  sampling  at  a  later  date  to  deter- 
mine the  ratio  of  marked  to  unmarked  animals.  Pop- 
ulation size  is  determined  from  the  ratio  of  marked 
to  unmarked  animals  in  one  or  more  successive  sam- 
ples. The  Petersen-Lincoln  estimator  is  perhaps  the 
most  commonly  used  mark-and-recapture  method. 


Data  Types 


19 


Mark-and-recapture  methods  require  several  assump- 
tions: (  1 )  no  loss  or  gain  of  marked  animals,  ( 2 )  no 
recruitment,  (3)  equal  catchability  in  marked  and 
unmarked  animals,  and  (4)  equal  mortality  rates  in 
marked  versus  unmarked  animals  (Tipton  1980). 
These  criteria  are  more  easily  met  in  samples  taken 
within  short  periods  (e.g.,  consecutive  days)  than 
over  longer  periods  (e.g.,  several  months).  Data  ob- 
tained by  these  methods  often  vary  (Tipton  1980), 
primarily  because  multiple  samples  have  been  taken 
under  different  climatic  conditions,  affecting  animal 
activity  and  behavior.  Otis  et  al.  (1978)  and  White  et 
al.  (  1982)  provide  detailed  discussions  of  these 
biases. 

Spatial  Arrangement  and  Movement.  In  addition 
to  estimating  abundance  and  density,  determining 
spatial  arrangement  and  movement  of  individual 
species  in  habitat  types  may  be  important.  For 
species  that  require  different  habitats  for  different 
functions,  such  as  breeding  or  feeding,  biologists 
should  collect  data  on  species  movements  and 
different  uses  of  habitats.  Individual  movements  can 
be  obtained  from  direct  observations  of  marked  or 
unmarked  animals  or  from  radiotelemetry.  Chapter 
33,  Radiotelemetry,  provides  guidance  on  such 
efforts.  These  types  of  data  help  biologists  determine 
habitat  needs  of  individual  species,  which  aid  in 
various  types  of  environmental  assessment. 
Limitations  of  mark-and-recapture  (resighting) 
animals  (see  previous  section)  and  of  radiotelemetry 
are  generally  related  to  cost,  the  former  being  labor- 
intensive  and  the  latter  requiring  expensive 
equipment. 

Population  Structure.  Sex  and  age  ratios  are  used 
to  determine  conditions  of  natural  populations, 
especially  big  game.  Sex  ratios  of  healthy  populations 
are  compared  with  those  of  populations  being 
studied,  thus  giving  the  biologist  an  indicator  of 
population  health.  Although  sex  ratios  may  indicate 
population  problems,  biologists  should  take  other 
population  data  to  support  conclusions  derived  from 
sex  ratios. 

Generally,  age  ratios  also  reflect  population  con- 
dition. Lack  of  certain  age  classes  such  as  juveniles 
and  breeding  adults  can  indicate  a  stressed  popula- 
tion. For  example,  biologists  are  concerned  about 
large  numbers  of  old  desert  tortoises  (Gopherus 
agasstzii)  and  lack  of  young  in  Arizona  populations. 
As  with  sex  ratios,  corroborative  data  such  as  abun- 
dance, density,  and  productivity  should  be  collected 
along  with  age  ratios. 

Biologists  should  be  cautious  when  using  and 
interpreting  population  structure  data.  Many  species' 
age  and  sex  ratios  fluctuate  naturally  with  changes 
in  environmental  conditions.  Therefore,  it  is  impor- 
tant to  obtain  multiple-year  samples  with  controls  to 
determine  human  changes  in  population  structure. 


Productivity.   Like  population  structure,  productiv- 
ity can  indicate  population  fitness.  Productivity  is 
usually  considered  the  number  of  young  (as  defined) 
recruited  into  the  population  per  adults,  adult  fe- 
males or  adult  pairs.  Examples  are  the  number  of 
young  fledged  per  nest  and  fawns  per  doe.  Although 
often  a  better  indicator  of  species'  fitness  than  popu- 
lation structure,  this  measurement  may  be  hard  to 
obtain  with  some  species. 

Condition.  Condition  of  individual  animals  can 
indicate  fitness  of  a  population,  especially  population 
stress.  For  example,  if  90%  of  the  animals  in  a 
bighorn  sheep  (Ovis  canadensis)  population  are  in 
poor  condition  (e.g.,  emaciated,  having  thin  coats  or 
high  incidences  of  sinusitus ),  then  a  biologist  might 
predict  that  this  population  is  in  trouble.  Although 
often  good  indicators  of  population  condition,  these 
measurements  are  generally  hard  to  obtain, 
expensive,  and  labor-intensive. 

General  Considerations.  Biologists  should  be 
careful  in  selecting,  making,  and  interpreting 
population  measurements,  particularly  estimates  of 
abundance,  density,  population  structure,  or 
productivity.  Population  numbers,  sex  and  age  ratios, 
and  productivity  all  vary  naturally.  If  a  biologist 
needs  to  collect  such  data  on  a  species  whose 
abundance  fluctuates  naturally  (e.g.,  rabbit 
populations),  then  samples  should  be  taken  over 
several  years. 

Knowledge  of  natural  population  fluctuations  is 
extremely  important  in  interpreting  data  from  cause  - 
and-effect  studies.  For  example,  observed  declines 
in  a  population  on  a  mined  area  may  result  from  low 
precipitation  during  sampling  rather  than  from  mine- 
caused  habitat  alteration. 

Community  Measurements 

Community  measurements  are  data  collected  on 
species  groups.  The  species  groups  may  be  taxo- 
nomic  (e.g.,  songbirds)  groups  that  use  habitat  in 
similar  ways  (guilds)  or  other  logical  groupings.  For 
example,  a  biologist  may  want  to  compare  species 
richness  (total  number  of  species)  between  two 
habitat  types  to  help  decide  where  to  place  a 
powerline. 

Community  data  may  also  involve  only  a  specific 
taxonomic  group.  For  example,  because  of  limited 
funding,  a  biologist  may  only  assess  species  diversity 
of  bird  communities  on  habitat  types.  Data  collection 
limited  to  a  taxonomic  group  or  community  may 
also  be  caused  by  objectives  established  during  pre- 
planning. For  example,  data  may  be  collected  only 
on  raptors  because  of  priorities  established  during 
preliminary  planning.  Community  measurements  are 
summarized  in  (Table  2). 


20 


Data  Types 


The  simplest  and  most  widely  used  community 
measurement  is  the  presence/absence  of  the  wildlife 
groups.  Presence/absence  data  allow  biologists  to 
determine  wildlife  community  composition  in  differ- 
ent habitat  types.  These  data  also  allow  biologists 
to  compare  faunal  richness  of  habitat  types.  These 
comparisons  may  be  useful  in  land-use  planning.  For 
example,  an  inventory  might  show  that  pinyon-juni- 
per  habitats  are  more  faunally  rich  than  adjacent 
desert  grassland  habitats  and  possess  a  greater  num- 
ber of  threatened  or  endangered  species.  These  data 
are  then  used  to  support  a  management  alternative 
of  placing  a  mining  road  through  desert  grassland 
habitat  rather  than  through  pinyon-juniper. 

While  primarily  used  at  a  habitat  level  (e.g., 
pinyon-juniper),  presence/absence  data  can  also  be 
used  to  determine  species  groups  that  depend  on 
certain  habitat  components  or  spatial  arrangement 
(e.g.,  vertical  vegetation  arrangement).  For  example, 
bird  species  can  be  verified  in  predetermined  incre- 
ments of  vertical  vegetation  layers,  thus  indicating 
bird  species  arrangements  within  a  habitat  type.  This 
type  of  data  can  also  be  useful  in  land-use  planning. 
A  biologist  may  find  that  shrub  and  old  growth  com- 
ponents of  pinyon-juniper  contain  the  greatest  vari- 
ety of  birds.  These  data  are  then  used  to  support  a 
management  alternative  that  would  limit  wood  cut- 
ting to  middle-aged  trees. 

Presence/absence  data  collection  is  far  less  ex- 
pensive than  abundance  or  density  data  collection, 
especially  when  censusing  an  entire  wildlife  commu- 
nity. Collecting  presence  data  for  wildlife  groups  is 
similar  to  that  for  single  species;  however,  the  analy- 
sis, interpretation,  and  use  of  such  data  are  often 
quite  different. 

Several  types  of  statistical  procedures  have  been 
used  to  analyze  presence/absence  data  of  wildlife 
communities.  Similarity  coefficients  are  the  simplest 
of  these  analyses.  A  summary  of  these  coefficients 
is  listed  in  (Table  2).  Generally,  similarity  coeffi- 
cients compare  the  number  of  species  common  to 
two  habitats  with  the  total  number  of  species.  Gen- 
erally, these  values  range  between  0  (no  similarity) 
and  1  (identical  communities).  Some  similarity  coef- 
ficients such  as  Simpson's  emphasize  similarities  be- 
tween two  sites,  and  others  such  as Jaccards 
emphasize  differences  between  sites.  See  Pimentel 
( 1979)  for  a  full  discussion  of  these  methods. 


Cluster  analysis  is  another  method  to  group 
species  that  use  similar  habitat  or  habitat  compo- 
nents. Cluster  analysis  also  provides  a  visual  perspec- 
tive of  degrees  of  similarity  between  groups  and 
variability  within  a  cluster  of  species.  Clusters  are 
depicted  in  two-dimensional  graphs  along  X  and  Y 
axes  or  in  histograms  based  on  separation  between 
groups.  Separation  or  distance  between  clustered 
species  groups  is  determined  several  ways  depending 
on  the  cluster  statistic  used.  For  example,  one  clus- 
tering statistic  determines  group  distances  by  calcu- 
lating a  central  location  in  the  X  and  Y  axes  of  each 
group.  Pimentel  ( 1979)  provides  a  detailed  explana- 
tion of  this  application  of  cluster  analysis,  and  Short 
and  Burnham  (  1982)  provide  an  example  of  this 
analysis,  demonstrating  how  bird  species  cluster  in 
terms  of  habitat  uses. 


Another  approach  to  determining  groupings  of 
species  is  to  determine  guilds,  which  are  groups 
of  species  that  use  similar  habitat  components.  Short 
(1983,  1984)  also  describes  a  simple  procedure  for 
determining  species  guilds  based  on  where  they 
breed  and  feed.  For  example,  one  guild  breeds  in 
vegetation  canopy  (above  4.5  m  [15  ft])  and  feeds 
on  the  surface,  whereas  another  guild  could  feed  in 
the  same  location  but  also  in  the  canopy.  These 
groups  are  different  in  their  use  of  habitat  for  feed- 
ing and  breeding.  Other  important  ecological  func- 
tions can  be  used  to  group  species  into  guilds, 
although  these  are  not  described  by  Short.  For  exam- 
ple, thermoregulation  is  important  in  reptile  commu- 
nities; therefore,  for  this  community,  it  is  important 
to  describe  where  and  how  they  thermoregulate 
in  addition  to  where  they  breed  and  feed. 


By  grouping  species  that  use  habitats  in  similar 
ways  (guilds),  the  biologist  can  predict  the  species 
groups  that  will  be  affected  when  a  habitat  compo- 
nent is  altered  or  lost.  The  detail  or  breakdown  of 
habitat  components  with  which  species  are  associ- 
ated can  vary.  For  example,  whereas  one  biologist 
identifies  3  vertical  layers  of  vegetation,  another  may 
identify  10.  The  biologist  who  identifies  three  layers 
may  be  more  concerned  with  horizontal  arrange- 
ment of  vegetation  for  determining  habitat  require- 
ments of  small  mammals  and,  therefore,  will  only  use 
three  vertical  layers,  whereas  another  biologist, 
studying  bird  communities,  may  need  greater  deline- 
ation of  vertical  space. 


In  interpreting  presence/absence  data,  biologists 
should  consider  the  intensity  of  the  sample.  In  short- 
term  inventories,  a  biologist  would  not  likely  ob- 
serve all  species  in  a  given  area,  especially  rare  spe- 
cies and  those  with  restricted  activity.  The  methods 
used  to  verify  species  in  an  area  may  also  be  incom- 
plete, and  this  will  affect  the  total  number  of  species 
observed. 


Simply  stated,  guilding  is  a  method  of  sorting 
wildlife  into  groupings  based  on  the  use  of  habitat 
components  for  important  ecological  functions.  The 
primary  function  of  placing  species  into  guilds  is 
to  provide  a  way  of  looking  at  how  species  groups 
might  be  affected  by  changes  in  habitat  structure, 
rather  than  determining  requirements  of  each  spe- 
cies and  then  projecting  how  each  might  be  affected 


Data  Types 


21 


by  a  proposed  action.  For  example,  a  group  of  spe- 
cies dependent  on  downed  logs  for  breeding,  feed- 
ing, and  thermal  cover,  would  be  negatively  affected 
by  removal  of  downed  and  dead  material  from  the 
area. 

Guild  data  can  be  effectively  used  during  early 
land  management  planning  stages,  especially  to  de- 
termine potential  conflicts  with  land-use  proposals. 
Guilding  also  helps  focus  attention  on  heavily  used 
components  of  habitat.  Biologists,  however,  should 
be  cautious  when  using  guild  data.  For  example, 
guild  data  do  not  account  for  the  ability  of  a  species 
to  shift  to  secondary  habitat  structures  when  pre- 
ferred habitat  is  lost.  In  addition,  guilds  are  artifi- 
cially defined  groups  described  by  biologists.  Mannin 
et  al.  (1984)  found  that  67%  of  birds  within  guilds 
in  Oregon  forest  habitats  responded  differently  to 
habitat  changes  than  predicted.  Some  biologists  ar- 
gued that  each  species  occupies  a  separate  guild  and, 
therefore,  reacts  differently  to  habitat  changes  (see 
discussions  of  Mannin  et  al.  1984).  Biologists  should 
not  attempt  to  use  guilding  data  to  determine  pre- 
cise, site-specific  changes  in  species  abundance  and 
composition.  Again,  these  data  are  best  used  to  focus 
attention  on  important  habitat  components,  rather 
than  as  precise  predictive  models. 

Another  related  procedure  is  use  of  indicator 
species.  One  use  of  this  procedure  involves  selecting 
a  species  termed  an  indicator  species,  which  repre- 
sents a  group  of  species  or  guild.  This  species  is  then 
sampled  to  determine  how  it  is  affected  by  land-use 
practices.  Data  obtained  for  this  species  are  then 
applied  to  the  entire  group  of  species  in  the  guild. 
For  example,  alligator  lizards  (Gerrhonotus  sp.)  rep- 
resent a  guild  of  species  that  require  downed  leaf 
litter  within  habitats  they  occupy.  Therefore,  in  the- 
ory, if  we  know  how  alligator  lizard  populations 
will  change  with  changes  in  leaf  litter  composition, 
we  know  how  all  other  members  of  that  guild  will 
respond.  Like  guilding,  this  procedure  assumes  that 
all  species  in  a  guild  use  habitat  in  an  identical  man- 
ner and  all  have  the  same  ability  to  use  secondary 
habitat.  As  mentioned  earlier,  Mannin  et  al.  (1984) 
found  that  species  within  guilds  did  not  react  simi- 
larly to  habitat  changes.  From  a  management  per- 
spective, the  indicator  species  process  saves  time 
and  money  since  it  allows  the  biologist  to  sample 
only  a  few  species  and  yet  determine  how  the  entire 
wildlife  community  responds  to  habitat  changes. 
Biologically,  however,  this  concept  does  not  consis- 
tently hold  up.  While  I  recommend  using  guilding 
procedures  to  focus  on  important  habitat  compo- 
nents, biologists  should  be  cautious  in  using  indica- 
tor species  to  represent  species  guilds,  especially 
when  accurate,  site-specific  predictions  are  needed. 

The  U.S.  Forest  Service,  Region  6  (Albuquerque, 
NM),  uses  indicator  species  in  a  different  way.  In- 
stead of  representing  a  group  or  guild  of  species,  an 


indicator  species  is  used  to  indicate  quality  of  a  cer- 
tain habitat  component  on  a  site.  A  group  of  indica- 
tor species  is  selected  so  that  all  important  habitat 
components  on  a  site  are  covered.  Then,  when  a 
certain  site  is  altered  by  a  land-use  practice,  the  biol- 
ogist can  look  at  changes  in  indicator  species  to 
determine  effects  on  important  habitat  components. 
Perhaps  an  easier  way  to  determine  the  same  effects 
would  be  to  measure  changes  in  important  habitat 
components.  The  biologist  identifies  these  important 
components  from  existing  studies  and  literature. 

When  it  is  possible  to  collect  data  on  abundance 
or  density  of  a  wildlife  or  taxonomic  community 
(e.g.,  raptors),  species  diversity  and  overlap  indexes 
can  be  calculated.  Of  the  several  types  of  diversity 
indexes,  the  most  commonly  used  is  the  Shannon- 
Weaver  index  (Hair  1980).  This  index  provides  a 
single  value  for  a  wildlife  community  based  both  on 
the  total  number  of  species  and  the  evenness  of 
abundance  among  the  species.  All  the  indexes  allow 
biologists  to  assess  relative  values  of  habitats  based 
on  species  diversity.  Since  the  indexes  are  based 
partially  on  relative  abundance  or  density,  they  can 
only  be  used  within  groups  having  abundance  meas- 
ured in  similar  ways.  For  example,  a  single  diversity 
index  cannot  be  computed  for  big  game  and  lizards, 
because  relative  abundance  or  density  is  determined 
in  different  ways. 

Whereas  similarity  coefficients  compare  species 
similarities  between  habitats,  overlap  indexes  com- 
pare proportions  of  species  and  their  abundance 
between  habitats.  Perhaps  the  most  commonly  used 
overlap  index  is  that  described  by  Horn  ( 1966). 
Routledge  (1980)  and  Ricklefs  and  Lau  (1980)  dis- 
cuss limitations  and  bias  associated  with  these  types 
of  indexes.  Biologists  should  consult  these  and  other 
sources  before  using  diversity  or  overlap  indexes. 

Another  type  of  data  used  to  assess  and  compare 
species  composition  and  influence  is  biomass.  In 
many  wildlife  studies,  where  the  biologist  is  trying  to 
determine  how  species  contribute  to  trophic  levels, 
biomass  is  measured.  For  example,  it  may  be  impor- 
tant to  know  how  the  loss  or  reduction  of  a  certain 
species  will  affect  food  available  to  other  species 
in  the  community.  In  trophic  level  and  food  niche 
studies,  biomass  information  is  generally  preferred 
over  abundance  data;  two  species  may  provide  simi- 
lar energy  to  a  system  even  though  one  is  small  and 
abundant  and  the  other  is  common  and  rare. 


HABITAT  DATA 

In  most  cases,  biologists  collect  habitat  data  to 
make  some  type  of  prediction  about  animal  popula- 
tions. Many  of  the  same  kinds  of  measurements  dis- 
cussed for  animals  and  animal  communities  can  also 
be  applied  to  plant  communities  (habitat  types)  and 


22 


Data  Types 


individual  plant  species.  In  addition,  habitat  measure- 
ments include  abiotic  physical  features  such  as  rock 
and  soils.  I  discuss  measurements  and  data  collection 
in  two  general  areas:  ( 1 )  habitat  area  or  type  and 
(2)  habitat  component. 

Habitat  Area  or  Type  Data 

These  types  of  data  apply  to  a  habitat  type  or 
area  as  a  whole.  Generally,  a  habitat  type  is  an  area, 
delineated  by  a  biologist,  that  has  certain  consistent 
abiotic  and  biotic  attributes  such  as  dominant  and 
subdominant  vegetation.  The  U.S.  Bureau  of  Land 
Management  (BLM)  terms  these  "habitat  sites"  (Kerr 
and  Brown  1977).  Groups  of  habitat  areas  with  simi- 
lar attributes  that  support  similar  groupings  of  spe- 
cies are  termed  "habitat  types"  or  in  BLM,  "standard 
habitat  sites." 

Similar  to  animal  community  measurements,  the 
simplest  measurement  of  habitat  type  is  presence/ 
absence.  In  most  cases,  this  involves  identifying  and 
mapping  habitat  sites  within  an  area. 

The  following  spatial  measurements  of  habitat 
types  are  commonly  used  to  assess  habitat  availabil- 
ity: 

•  size  of  habitat  sites  (area); 

•  location  and  position  of  habitat  sites,  including 
distance  between  habitat  sites  and  heterogeneity 
within  an  area; 

•  edge  influences  created  by  habitat  site  interfaces 
and  ecotones; 

•  temporal  availability  of  habitat  sites  (e.g.,  tempo- 
rary lentic  habitat  sites). 

All  these  data  help  biologists  assess  availability  of 
certain  habitats  and  the  potential  for  an  area  to  sup- 
port various  types  of  wildlife. 

Indexes  have  been  used  to  quantify  potential 
influences  of  edges  on  wildlife  habitat.  Patton  (  1975) 
provides  a  simple  index  of  edge. 


Habitat  Component  Data 

Habitat  component  data  refer  to  information 
taken  on  abiotic  and  biotic  attributes  of  habitat.  Gen- 
erally, data  collection  is  limited  to  those  components 
known  to  affect  wildlife  distribution  and  population 
fitness.  Data  can  include  spatial  arrangement  of  these 
components  at  a  single  point  in  time  or  over  several 
years,  or  physical  characteristics  of  components  such 
as  average  height  and  width.  Habitat  component 
data  can  represent  an  entire  habitat  type  or  a  specific 
area  occupied  by  species.    Table  2    provides  a  sum- 
mary of  these  data.  Examples  of  abiotic  data 
include — 

•  average  surface  rock  cover, 

•  average  size  of  surface  rocks, 

•  tallus  and  rock  outcrop  abundance  and  size, 

•  soil  depth, 

•  vertical  and  horizontal  heterogeneity  of  soils 
and  rock. 

Measurement  and  classification  of  terrestrial  and 
aquatic  physical  features  are  covered  in  Chapters 
27  and  28  in  this  book.  Examples  of  biotic  data 
include — 

•  average  litter  depth; 

•  average  litter  cover  per  unit  area; 

•  snag  density  and  size; 

•  average  height  of  different  vegetation  life-forms 
(grasses,  shrubs,  and  trees); 

•  vertical  and  horizontal  heterogeneity  of  litter, 
snags,  and  live  vegetation. 

Other  data  often  collected  include  attributes  such  as 
elevation,  latitude  and  longitude,  slope,  and  aspect. 

Many  sampling  methods  can  be  used  to  obtain 
presence/absence,  abundance,  density,  size,  spatial 
arrangement,  and  habitat  and  habitat  component 
availability  data.  These  methods  fall  into  four  general 
categories: 


Plant  species  richness,  diversity,  similarity,  and 
overlap  can  be  computed  for  habitat  sites  or  types  in 
ways  similar  to  animal  species  composition  in  animal 
communities.  Plant  species  richness  is  the  total  num- 
ber of  species  present  within  a  habitat  type.  Similar- 
ity coefficients  can  be  computed  by  comparing 
shared  plant  species  richness  between  habitats.  Plant 
species  diversity  and  overlap  between  habitat  types 
are  computed  from  diversity  indexes  identical  to 
those  previously  discussed  for  animals.  Often,  these 
data  help  assess  forage  quantity  and  quality  between 
habitat  sites  or  types. 


(1)  Point  (e.g.,  point  intercept); 

(2)  Line  (e.g.,  line  intercept); 
(  3 )  Area  or  plot; 

(4)  Plotless  (e.g.,  point-center-quarter). 


Generally,  point  data  can  be  obtained  quickly, 
depending  on  the  number  of  points  needed.  The 
number  of  samples  taken  will  depend  on  variability 


Data  Types 


23 


of  the  habitat  component  to  be  measured.  For  exam- 
ple, a  biologist  can  use  several  point  samples  to  de- 
termine leaf  litter  depth.  These  points  can  be 
systematically  (e.g.,  on  a  line)  or  randomly  located. 
One  of  the  most  widely  used  of  these  methods  is  the 
toe-pace  transect  (Hays  et  al.  1981).  Point  methods 
can  also  be  used  to  determine  frequency  and  relative 
abundance  of  certain  components.  For  example,  a 
toe-pace  transect  can  be  used  to  determine  percent- 
ages of  certain  substrate  on  habitat  surfaces.  This 
type  of  procedure  does  not  identify  the  density  or 
arrangement  of  habitat  components  on  the  surface 
but  rather  frequency  and  relative  abundance  of  these 
components  (expressed  as  a  percentage).  These 
types  of  data  are  best  used  as  a  rough  indicator  of 
component  composition  and  frequency  on  a  site. 

Line  transect  methods  also  generate  data  on 
habitat  component  composition,  frequency,  and 
abundance.  In  addition,  they  provide  data  on  hori- 
zontal arrangement  of  habitat  components  (a  func- 
tion of  distance  between  components).  Similar  to 
point  transects,  these  data  are  expressed  as  percent- 
ages and  can  be  used  to  characterize  habitat  compo- 
nents on  a  site.  The  most  common  line  transect 
method  is  the  line-intercept  (see  Hays  et  al.  1981 ). 

Plot  sampling  techniques  are  frequently  used  to 
describe  habitat  component  arrangement,  density, 
and  composition  within  an  area.  The  area  in  which 
habitat  components  are  measured  is  usually  prede- 
fined, and  area  size  can  vary  greatly.  Methods  used  to 
generate  these  data  are  usually  time-  and  labor-inten- 
sive, providing  relatively  accurate  data  but  only  for 
a  small  area.  They  are  most  commonly  used  when 
grass  and  forb  composition  and  density  data  are 
needed.  Several  plot  or  area  methods  have  been  de- 
veloped. See  Hays  et  al.  (1981 )  for  greater  detail. 
Data  generated  from  plot  methods  are  usually  ex- 
pressed as  the  number  of  plants  per  unit  area.  In 
many  forage  studies,  data  are  expressed  in  weight  by 
plant  species  per  unit  area. 

Plotless  sampling  techniques  produce  density 
estimates  for  habitat  components,  but  as  a  function 
of  distance  rather  than  density  in  a  predetermined 
area.  One  of  the  more  commonly  used  plotless  tech- 
niques is  the  point-center-quarter  method  (Phillips 
1959),  which  allows  many  samples  to  be  collected 
in  less  time  than  a  plot  method. 

When  set  up  as  permanent  samples  and  read 
over  several  years,  line,  plot,  and  plotless  methods 
yield  valuable  information  on  changes  in  habitat 
components  on  a  site.  Nonetheless,  these  habitat 
measurement  techniques  have  many  biases  and  elim- 
inations. Chapters  26  through  32  discuss  habitat 
component  sampling  methods  in  greater  detail. 

In  all  these  methods,  characteristics  of  individ- 
ual habitat  components  can  be  measured.  For  exam- 


ple, tree  density  determined  from  a  point-center- 
quarter  method  (plotless)  is  based  on  distances  of 
trees  from  the  point.  Each  tree  sampled  for  distance 
can  also  be  measured  for  physical  characteristics 
such  as  height  and  diameter.  In  some  cases,  habitat 
components  cannot  be  measured  randomly  or  sys- 
tematically. For  example,  a  biologist  might  need 
to  determine  physical  characteristics  of  snags  used 
by  nesting  spotted  owls  (Strix  occidentalis)  so  that 
adequate  habitat  is  provided  in  the  area's  snag  man- 
agement plan.  In  this  case,  the  biologist  would  meas- 
ure only  those  snags  with  known  spotted  owl  nests. 

Heterogeneity  of  both  abiotic  and  biotic  habitat 
components  can  be  computed  from  diversity  in- 
dexes. Foliage  height  diversity  and  patchiness  in- 
dexes are  two  of  the  more  commonly  used  (see 
Ralph  and  Scott  1981;  Hays  et  al.  1981;  Chapter  31 
in  this  book ).  Foliage  height  diversity'  data  give  the 
biologist  an  indication  of  how  evenly  vegetation 
is  distributed  in  vertical  space.  Patchiness  diversity 
indexes  provide  data  on  how  evenly  vegetation  is 
spread  on  horizontal  space.  Vertical  spatial  data  are 
extremely  helpful  in  determining  a  site's  ability  to 
support  certain  birds,  while  horizontal  spatial  data 
help  determine  which  small  mammals  and  lizards 
occupy  a  site. 

In  addition  to  other  types  of  habitat  data  dis- 
cussed, population  structure  (e.g.,  age  ratios),  pro- 
ductivity, and  condition  (e.g.,  degree  of  hedging) 
measurements  can  be  taken  on  certain  plant  species. 
These  types  of  measurements  are  similar  to  those 
taken  on  individual  animal  species,  but  are  generally 
easier  to  obtain  because  plants  are  stationary.  These 
measurements  help  biologists  determine  the  condi- 
tion and  fitness  of  certain  types  of  plants  such  as  key 
riparian  tree  species  or  important  big  game  browse. 


ANIMAL/HABITAT  CORRELATIONS 

Perhaps  the  most  important  consideration  in 
designing  and  conducting  a  habitat  inventory  or 
monitoring  project  is  ensuring  that  animal  and  habi- 
tat data  can  be  compared.  By  doing  so,  biologists  can 
then  predict  how  habitat  alterations  will  affect  ani- 
mal populations. 

Two  factors  are  extremely  important  if  animal 
and  habitat  data  are  to  be  compared.  First,  animal 
and  habitat  data  should  be  collected  on  the  same 
site.  As  samples  accumulate,  the  biologist  can  deter- 
mine how  animal  populations  fluctuate  with  differ- 
ences in  physical  and  spatial  arrangement  of  habitats. 
If  a  biologist  samples  habitats  and  animals  at  different 
sites,  habitat  and  animal  data  cannot  be  compared. 

Second,  the  level  of  sampling  needed  to  com- 
pare animal  and  habitat  data  should  be  carefully  de- 
termined. In  most  cases,  animal  and  habitat  sampling 


24 


Data  Types 


levels  do  not  need  to  be  similar.  For  example,  most 
inventories  of  wildlife  eommunities  in  habitat  types 
require  only  presence  and  absence  data  for  animals. 
However,  habitat  measurements  are  generally  more 
specific  than  simple  presence  or  absence  of  habitat 
components;  these  tell  little  about  habitat  quality 
and  condition.  More  often,  habitat  measurements  in- 
clude spatial  arrangement  and  abundance  of  habitat 
components. 

Conversely,  sometimes  biologists  may  need 
more  detailed  data  on  animals  than  habitat.  For  ex- 
ample, a  biologist  who  is  monitoring  the  effect  of 
new  water  developments  on  a  pronghorn  population 
in  an  area  would  require  detailed  information  on 
movement,  abundance  or  density,  spatial  arrange- 
ment, and  reproduction  of  pronghorns.  Habitat  data 
collection  would  probably  be  limited  to  habitat  map- 
ping and  presence  or  absence  of  water.  In  both 
cases,  the  level  or  complexity  of  animal  versus  habi- 
tat sampling  depend  on  specific  inventory  and  moni- 
toring objectives.  However,  habitat  data  are  usually 
less  costly  and  time-consuming  to  collect  and  more 
accurate  than  animal  data;  plants  are  easier  to  locate 
and  are  somewhat  stationary.  Therefore,  most  inven- 
tories and  studies  generally  have  more  detailed  habi- 
tat data.  (Table  2)  provides  a  summary  of  these  data. 

Many  statistical  procedures  compare  animal  and 
habitat  data.  Perhaps  the  simplest  of  these  proce- 
dures is  to  compile  a  list  of  species  observed  or  col- 
lected in  a  habitat  component  or  habitat  type.  This 
allows  biologists  to  weigh  relative  wildlife  values 
of  certain  habitat  components  and  habitat  types, 
which  can  then  be  used  in  evaluating  proposed  land 
management. 

Multivariate  analyses,  such  as  correlation,  regres- 
sion, and  principal  component  analyses,  can  be  used 
to  compare  animal  and  habitat  data.  For  example,  a 
biologist  may  want  to  determine  which  physical 
attributes  of  habitat  affect  species  richness.  Jones  et 
al.  (1985)  used  a  step-wise,  multiple  regression  to 
determine  how  reptile  species  richness  on  habitat  is- 
lands was  affected  by  habitat  size  and  habitat  compo- 
nent structure.  In  this  study,  animal  data  consisted 
of  only  presence  and  absence  of  species  on  sites, 
while  habitat  data  were  more  detailed  and  included 
size,  elevation,  and  detailed  horizontal  and  vertical 
habitat  structures.  Pimentel  ( 1979)  and  Green 
(1979)  discuss  these  statistical  procedures  and  pro- 
vide examples.  In  addition,  Jones  and  Glinski  ( 1983), 
discussing  lizards;  Rotenberry  and  Wiens  ( 1980), 
discussing  birds;  and  Matthews  and  Hill  (1981 ),  dis- 
cussing fish,  provide  examples  of  multivariate  anal- 
yses of  animal/habitat  data. 

The  U.S.  Fish  and  Wildlife  Service  and  U.S.  For- 
est Service  have  developed  models  that  assess  habitat 
capability  or  suitability  for  certain  species.  The  for- 
mer has  developed  the  Habitat  Evaluation  Proce- 


dures (HEP)  system,  consisting  of  a  series  of  habitat 
suitability  index  ( HSI )  models  developed  for  individ- 
ual animals  (Schamberger  et  al.  1982). 

The  HEP  process  and  HSI  models  have  some 
drawbacks.  First,  habitat  ratings  are  indexes,  and  are 
normally  derived  from  expert  opinion  and  literature 
review  rather  than  from  empirical  data.  Testing  these 
models  is  difficult,  since  the  index  cannot  be  verified 
in  the  field. 

Another  type  of  habitat  evaluation  model  was 
developed  by  Williams  et  al.  (1977).  These  are 
known  as  Pattern  Recognition  ( PATREC )  models, 
which,  unlike  HEP  models,  are  based  on  statistical 
inference.  Data  used  in  these  models  are  often  em- 
pirical (e.g.,  population  data).  Models  result  in  prob- 
ability statements  on  an  area's  suitability  to  support 
a  certain  species  (e.g.,  an  area  in  question  has  a 
75%  probability  of  supporting  dinosaurs ).  Because 
PATREC  models  are  developed  from  empirical  data 
(generally  a  population  factor  such  as  presence  or 
density),  they  are  easier  than  HSI  models  to  test. 

The  U.S.  Forest  Service  has  developed  a  series  of 
models  (habitat  capability  relation  models)  that  cor- 
relate species  with  habitat  components.  The  U.S. 
Forest  Service  has  also  developed  simulation  models 
that  simulate  changes  in  vegetation  over  time  as  a 
result  of  plant  succession  and  land-use  practices  on 
an  area.  By  combining  these  two  types  of  models, 
biologists  can  predict  how  wildlife  habitat  quality 
and,  thus,  wildlife  populations  will  change  over  time 
given  a  certain  land  use. 

This  approach  is  limited  in  that  habitat  quality  is 
evaluated  only  by  serai  stages  of  cover  types  (habitat 
types).  The  procedure  does  not  account  for  relation- 
ships between  site-specific  habitat  components  and 
wildlife,  and  it  is  based  on  models  developed  for 
individual  species.  Currently,  only  a  few  species  have 
been  modeled.  Therefore,  I  recommend  caution 
when  using  this  system,  especially  when  assessing 
management  alternatives  on  a  site-specific  basis.  As 
with  guilding,  I  believe  such  models  are  most  valua- 
ble during  major  land-use  planning,  especially  to 
focus  attention  on  major  issues. 

The  area  of  animal/habitat  correlations  is  cov- 
ered in  more  detail  in  Chapter  38,  Habitat  Evaluation 
Systems. 


DISCUSSION 

Generally,  there  are  two  approaches  to  data 
collection  and  their  use  in  recommending  wildlife 
and  habitat  alternatives.  These  two  approaches  have 
been  described  by  Bailey  (1982)  and  are  termed 
"Linear-Comprehensive  Management"  and  "Cyclic- 
Incrementalism." 


Data  Types 


25 


Linear-comprehensive  management  is  the  accu- 
mulation of  data  on  all  habitat  components  and  wild- 
life populations  in  all  ecosystems  (Figure  1).  This 
includes  detailed,  intensive  data  on  population  sizes 
and  productivities,  seasonal  distributions  of  animals, 
forage  productivity,  cover  availability,  precipitation, 
and  all  other  ecological  factors  affecting  biological 
systems.  This  management  philosophy  also  assumes 
clear  definitions  of  publicly  agreed-upon  manage- 
ment goals;  the  biologist  will  have  specific  data  that 
support  public  opinions  on  issues.  It  also  assumes 
that  all  population  and  habitat  treatment  methods 
are  clearly  understood  and  that  the  biologist  has 
a  comprehensive  knowledge  of  how  each  treatment 
will  affect  the  ecosystem  and  animals  in  question. 

In  general,  linear-comprehensive  management 
generates  large  amounts  of  site-specific  data  without 
regard  to  how  the  resulting  data  will  be  used  in 
management  decisions.  In  some  cases,  models  are 
developed  and  used  in  planning  without  local  test- 
ing. Linear-comprehensive  management  does  not 
work  in  most  wildlife  management  situations  for  sev- 
eral reasons.  First,  wildlife  is  diverse  in  most  ecosys- 


INVENTORY 

1 

CENSUS 

1 

YIELD 
DETERMINATION 

1 

DIAGNOSIS 

* 

CONTROL 

HABITAT 
MGMT 

POPULATION 
MGMT 

terns.  Simply  stated,  the  biologist  cannot  collect  data 
on  all  population  and  habitat  factors,  nor  can  a  biolo- 
gist determine  how  each  species  will  be  affected  by 
every  treatment.  Some  of  the  most  comprehensive 
studies  have  only  censused  a  tiny  fraction  of  the 
animal  population  in  an  ecosystem.  In  addition,  the 
relationships  between  animals  and  ecosystems  are 
very  complex.  Most  animals  have  several  habitat 
requirements,  and  these  requirements  are  often  in- 
terrelated in  complex  ways.  Therefore,  predicting 
how  any  particular  animal  will  react  to  habitat 
changes  is  difficult.  As  mentioned  earlier,  Mannan  et 
al.  ( 1984)  found  that  bird  species  within  well-de- 
fined guilds  responded  differently  to  certain  manage- 
ment prescriptions.  Finally,  wildlife  budgets  prohibit 
collection  of  large  amounts  of  wildlife  data.  Because 
knowledge  and  budgets  are  limited,  most  wildlife 
management  programs  are  extensive,  rather  than 
intensive. 

Cyclic-incrementalism  is  defined  as  a  slow,  sub- 
jective process  of  collecting  data  and  knowledge 
on  how  certain  management  prescriptions  affect 
wildlife  (Figure  2).  This  process  also  allows  for  con- 
tinuous adjustment  of  treatments  and  management 
based  on  monitoring  data.  In  this  process,  the  biolo- 
gist identifies  a  series  of  achievable  goals  (e.g.,  30% 
more  browse  in  an  area ).  These  goals  are  measured 
in  units  that  relate  directly  to  management  objec- 
tives. The  biologist  then  selects  a  series  of  methods 
that  are  believed  to  achieve  the  management  objec- 
tive (e.g.,  prescribed  burn).  Since  the  biologist  is  not 
100%  confident  that  the  prescription  will  have  the 
desired  result,  it  is  applied  on  a  limited  ( incremen- 
tal) basis,  and  then  the  result  is  monitored.  From  the 


PLANNED 

TEST  OF 

PUBLIC  RESPONSE 


POPULATION, 
HABITAT 
STATUS 


TREATMENT 


GOALS 
P 


PI   AMMFH 


PLANNED 

TEST  OF 

BIOTIC  RESPONSE 


Figure  1.     The  linear-comprehensive  approach  to 

wildlife  management  (after  Bailey  1982). 


Figure  2.     The  cyclic-incremental  approach  to 

wildlife  management  (after  Bailey  1982). 


26 


Data  Types 


monitoring  data,  the  biologist  can  determine  if  the 
habitat  management  prescription  or  land  use  needs 
to  be  modified. 

In  this  process,  the  biologist  determines  the 
concerns  and  needs  of  the  public  in  formulating 
management  objectives.  It  is  also  extremely  impor- 
tant to  inform  the  public  about  any  findings  made 
during  the  monitoring  of  a  management  prescription. 
This  will  help  the  biologist  reinforce  or  modify  man- 
agement objectives.  Because  prescriptions  are  con- 
ducted and  monitored  on  a  site-specific  basis,  the 
biologist  has  a  better  understanding  of  the  local  situ- 
ation. This  type  of  procedure  is  more  likely  to  gain 
public  support,  as  opposed  to  use  of  rigid  proce- 
dures developed  for  large  regions. 

In  his  description  of  these  concepts,  Bailey 
(1982)  points  out  that  wildlife  management  is  an  art, 
not  a  science.  Wildlife  managers  apply  the  science 
of  biology,  but  the  application  of  this  science  is  an 
art.  The  process  of  cyclic-incrementalism  allows  the 
biologist  to  accumulate  site-specific  data  relative  to 


publicly-accepted  wildlife  management  objectives, 
and  slowly  build  a  basis  for  management  decisions. 
Because  this  process  does  not  require  large-scale 
data  collection,  it  can  be  scaled  to  available  budget. 

The  key  to  successful  inventories  and  monitor- 
ing studies  is  in  establishing  achievable  and  manage- 
ment-related objectives,  and  in  orienting  the  type 
and  level  of  data  collected  to  these  objectives.  It 
serves  no  purpose  to  determine  the  distribution  of 
all  wildlife  in  several  habitat  types  if  the  objective  is 
to  determine  how  grazing  practices  affect  certain  key 
wildlife  species.  Data  should  be  collected  to  solve 
specific  management  problems. 

Throughout  the  chapter,  I  have  highlighted  dif- 
ferent types  and  levels  of  data.  This  information 
should  be  used  in  making  decisions  on  the  type  of 
data  needed  to  answer  specific  management  ques- 
tions. I  do  not  recommend  using  a  standard  set  of 
methods  and  data,  but  rather  an  assortment  of  tech- 
niques and  information,  depending  on  individual 
objectives,  goals,  priorities,  and  budget. 


Data  Types 


27 


LITERATURE  CITED 


BAILEY,  J.A.  1982.  Implications  of  "muddling  through"  for 
wildlife  management.  Wildl.  Soc.  Bull.  10(4):363-369. 

BROWN,  D.E.,  C.H.  LOWE,  and  C.  PASE.  1979.  A  digitized 
classification  system  for  the  biotic  communities  of 
North  America,  with  community  (series)  and  associa- 
tion examples  for  the  Southwest.  J.  Arizona — Nevada, 
Acad.  Sci.  14,  Supplement  1. 

BURY,  R.B.  1982.  Structure  and  composition  of  Mojave 
Desert  reptile  communities  determined  with  a  re- 
moval method.  Herpetological  Communities,  U.S.  Dep. 
Inter.,  Fish  and  Wildl.  Serv.,  Wildl.  Res.  Rep.  1 3. 

CAUGHLEY,  G.  1977.  Analysis  of  vertebrate  populations. 
John  Wiley  and  Sons,  London.  234pp. 

EBERHARDT,  L.L.  1978.  Appraising  variability  in  popula- 
tion studies.  J.  Wildl.  Manage.  42(2):207-238. 

EMLEN,  J.T.  1971.  Population  densities  of  birds  derived 
from  transect  counts.  Auk  88.323-342. 

GREEN,  R.H.  1979.  Sampling  design  and  statistical  meth- 
ods for  environmental  biologists.  John  Wiley  and 
Sons,  New  York,  NY. 

GYSEL,  L.W.  and  L.J.  LYON.  1980.  Habitat  analysis  and 
evaluation.  Pages  305-327  in  Schemnitz,  S.D.,  ed. 
Wildlife  Management  Techniques  Manual.  The  Wildl. 
Soc,  Washington,  DC. 

HAIR,  J. D.  1980.  Measurements  of  ecological  diversity. 
Pages  269-275  in  Schemnitz,  S.D.,  ed.  Wildlife  Man- 
agement Techniques  Manual.  The  Wildl.  Soc,  Wash- 
ington, DC. 

HAYS,  R.L.,  C.  SUMMERS,  and  W.  SEITZ.  1981.  Estimating 
wildlife  habitat  variables.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  FWS/OBS-81/47.  111pp. 

HENDRICKSON,  DA.,  W.L  MINCKLEY,  R.R.  MILLER,  D.J. 
SIEBERT,  and  PH.  MINCKLEY.  1980.  Fishes  of  the  Rio 
Yaqui  Basin,  Mexico  and  United  States.  J.  Arizona, 
Acad.  Sci.  15(3>65-106. 

HORN,  H.S.  1966.  Measurements  of  overlap  in  compara- 
tive ecological  studies.  Am.  Nat.  100:419-424. 

JONES,  KB.  and  C.  GLINSKI.  1983.  Microhabitats  of  lizards 
and  southwestern  riparian  community.  Pages  342- 
346  in  Johnson,  R.R.,  CD.  Ziebell,  DR.  Patton,  P.F. 
Ffolliott,  and  R.H.  Hamre,  tech.  coords.  Riparian  Eco- 
systems and  their  Management:  Reconciling  Conflict- 
ing Uses.  U.S.  Dep.  Agric,  Forest  Serv.,  Gen.  Tech. 
Rep.  RM-120. 

,  LP.  KEPNER,  and  T.E.  MARTIN.  1985.  Species  of 

reptiles  occupying  habitat  islands  in  western  Arizona: 
A  deterministic  assemblage.  Oecologia.  In  press. 

KERR,  R.M.  and  K  BROWN.  1977.  Data  requirements  for 
terrestrial  wildlife  habitat  inventory  in  Classification, 
Inventory,  and  Analysis  of  Fish  and  Wildlife  Habitat. 
Proc  Natl.  Symp.  Phoenix,  AZ,  January  1977.  U.S. 
Govt.  Printing  Office,  Washington,  DC. 

KUCHLER,  AW.  1964.  Map  of  potential  natural  vegetation 
of  the  conterminous  United  States.  Am.  Geophysical 
Soc,  Spec.  Publ.  36. 

MANNAN,R.W„  M.L  MORRISON,  and  EC.  MESLOW.  1984. 
Comment:  The  use  of  guilds  in  forest  bird  manage- 
ment. Wildl.  Soc.  Bull.  12(4):426-430. 

MATTHEWS,  W.J.  and  L.G.  HILL.  1980.  Habitat  partitioning 
in  the  fish  community  of  a  southwestern  USA  river. 
Southwestern  Nat.  25(1):  51-66. 


OTIS,  D.L.,  KP.  BURNHAM,  GC.  WHITE,  and  DR.  ANDER- 
SON. 1978.  Statistical  inference  from  capture  data 
on  closed  animal  populations.  Wildl.  Monogr. 
62:1-135. 

PATTON,  DR.  1975.  A  diversity  index  for  quantifying 
habitat  "edge."  Wild.  Soc.  Bull.  3(4):171-173. 

PHILLIPS,  E.A.  1959.  Methods  of  vegetation  study.  Henry 
Holt  and  Company,  Inc.  107pp. 

PIMENTEL,  R.A.  1979.  Morphometries:  The  multivariate 
analysis  of  biological  data.  Kendall/Hunt  Publ.  Co., 
Dubuque,  IA. 

RALPH,  C.J.  and  J.M.  SCOTT,  eds.  1981.  Estimating  num- 
bers of  terrestrial  birds.  Allen  Press,  Inc.,  Lawrence, 
KS. 

RICKLEFS,  RE.  and  M.  LAU.  1980.  Bias  and  dispersion  of 
overlap  indices:  Results  of  some  Monte  Carlo  simula- 
tions. Ecology  6 1(5):  10 19- 1024. 

ROTENBERRY,  J.T.  and  J.A.  WIENS.  1980.  Habitat  struc- 
ture, patchiness,  and  avian  communities  in  North 
America  steppe  vegetation:  A  multivariate  analysis. 
Ecology  61:1228-1250. 

ROUTLEDGE,  R.D.  1980.  Bias  in  estimating  the  diversity 
of  large,  uncensused  communities.  Ecology 
6l(2):276-281. 

SCHAMBERGER,  MA,  AH.  FARMER,  and J.W.  TERRELL. 
1982.  Habitat  suitability  index  models:  Introduction. 
U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  FWS/OBS-82/10. 

SEBER,  G.A.F.  1973  Estimation  of  animal  abundance  and 
related  parameters.  Griffin,  London.  506pp. 

SHORT,  H.L  1983.  Wildlife  guilds  in  Arizona  desert  habi- 
tats. U.S.  Dep.  Inter.,  Bur.  Land  Manage.,  Tech.  Note 
362.  258pp. 

.  1984.  Habitat  suitability  index  models:  The  Ari- 
zona guild  and  layers  of  habitat  models.  U.S.  Dep. 
Inter.,  Fish  and  Wildlife  Serv.  FWS/OBS-82/ 10.70. 
37pp. 

and  KP.  BURNHAM.  1982.  Technique  for  structur- 


ing wildlife  guilds  to  evaluate  impacts  on  wildlife 
communities.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv., 
Special  Sci.  Rep.  244.  Washington,  DC. 

SMITH,  M.H.,  R.H.  GARDNER,  J.B.  GENTRY,  D.W.  KAUF- 
MAN, and  M.J.  OTARRELL  1975.  Density  estimation 
of  small  animal  populations.  Pages  25-53  in  Golley, 
F.B.,  K  Petrusewicz,  and  L.  Ruszkowski,  eds.  Small 
Mammals:  Their  Production  and  Population  Dynamics. 
Int.  Biol.  Prog.  5,  Cambridge  Univ.  Press,  London. 

THOMAS,  J.W.  1979.  Wildlife  habitats  in  managed  forests: 
The  Blue  Mountains  of  Oregon  and  Washington.  U.S. 
Dep.  Agric,  For.  Serv.,  Agric.  Handbook  553 

TIPTON,  A.R.  1980.  Mathematical  modeling  in  wildlife 
management.  Pages  211-220  in  Schemnitz,  S.D.,  ed. 
Wildlife  Management  Techniques  Manual.  The  Wildl. 
Soc.  Washington,  DC. 

WHITE,  G.C.,  DR.  ANDERSON,  KP.  BURNHAM,  and  D.L 
OTIS.  1982.  Capture-recapture  and  removal  methods 
for  sampling  closed  populations.  Los  Alamos  Natl. 
Laboratory  LA-8787-NERP. 

WILLIAMS,  G.L.,  KR.  RUSSELL,  and  W.K  SEITZ.  1977. 

Pattern  recognition  as  a  tool  in  the  ecological  analysis 
of  habitat  in  Marmelstein,  A.,  ed.  Classification,  Inven- 
tory, and  Analysis  of  Fish  and  Wildlife  Habitat.  U.S. 
Dep.  Inter.,  Fish  and  Wildl.  Serv.  FWS/OBS-78/76. 
604pp. 


28 


Data  Types 


LITERATURE 
REVIEW 


Ora  Wagoner1 

Bureau  of  Land  Management 

Service  Center 

Denver,  CO  80225-0047 


Editor's  Note.  Too  many  biologists  duplicate  work 
done  by  others.  Libraries  can  help  you  determine 
what's  already  been  accomplished  in  your  field 

Consulting  existing  information  is  a  mandatory 
step  in  preparing  for  an  inventory  or  monitoring 
study.  This  involves  more  than  a  literature  review. 
Today's  libraries  search  literature  and  other  infor- 
mation sources  with  electronic  data  bases,  and  they 
pool  their  resources  by  operating  in  networks.  They 
can  also  provide  you  with  unpublished  data  and 
referrals  to  current  researchers  or  other  information 
sources,  such  as  museum  records. 


"Immediately  after  characterizing  a  problem,  and  im- 
mediately before  initiating  any  action,  find  out  who 
might  already  know  the  answer.  Remember  to  ask 
the  librarian." 

Alan  Speigel 


INTRODUCTION 

Wildlife  inventories  and  monitoring  studies 
should  begin  with  searches  of  existing  literature  to 
determine  past  research  and  current  studies.  The  tre- 
mendous growth  in  stores  of  knowledge  has  made 
searching  the  literature  a  complicated  procedure  re- 
quiring differing  techniques  and  special  tools. 

A  librarian's  approach  to  literature  searching  is 
systematic,  to  investigate  all  possible  sources  of  in- 
formation. This  chapter  provides  biologists  with 
guidance  on  how  scientific  literature  and  other  infor- 
mation are  accessed,  and  introduces  BLM  employees 
to  the  services  provided  by  the  BLM  Library  (see 
Appendix  I). 

Wildlife  biologists  need  to  use  libraries  and  their 
resources  to  ensure  that  they  perform  thorough  liter- 
ature searches  and  that  they  access  other  sources  of 
information  essential  to  wildlife  inventories  and 
monitoring  studies.  Library  users  must  allot  time  for 
literature  searches  to  be  performed,  as  materials 
must  often  be  obtained  from  distant  sources. 

The  processes  used  to  get  information  depend 
on  the  nature  and  amount  of  data  requested.  In  a 
typical  search,  an  information  specialist  first  checks 
printed  sources  available  at  the  library;  if  not  suc- 
cessful, he  or  she  then  uses  computer  networks  to 
inquire  about  information  from  other  locations.  The 
delivery  of  documents  from  other  locations  can  take 
4  to  6  weeks,  so  biologists  should  plan  on  this 
amount  of  time  to  conduct  their  searches. 

This  chapter  covers  published  materials  includ- 
ing serials,  reports,  books,  and  other  documents,  and 
specialized  systems  of  organized  data  such  as  com- 
puter data  bases  and  museums.  It  also  provides  sug- 
gestions on  accessing  unpublished  data. 


1  Present  Address:  Dudley  Knox  Library,  Naval  Postgraduate  School, 
Monterey,  CA  93943-5002. 


Literature  Review 


29 


PRINTED  MATERIALS 

Printed  materials  include  published  and  unpub- 
lished documents.  Information  about  published  mate- 
rials is  found  in  many  different  forms.  Announce- 
ments of  new  books  or  reviews  provide  the  best 
method  for  keeping  abreast  of  newly  published  mate- 
rial in  a  specific  subject  area.  These  can  be  found  in 
many  journals.  Information  from  book  publishers,  so- 
cieties, and  universities  that  publish  in  particular 
areas  of  interest  can  also  be  invaluable  for  learning 
of  new  publications.  A  review  of  Subject  Guide  to 
Books  in  Print  (published  by  R.  R.  Bowker  Com- 
pany, 205  E.  42nd  St.,  New  York,  NY  10017)  will  re- 
veal most  new  books  in  all  subject  areas  as  well  as 
older  books  that  are  still  available  from  the  publish- 
ers. Appropriate  subject  headings  to  use  in  searching 
Books  in  Print  are  wildlife  management  and  animal 
populations. 

Because  book  material  becomes  quickly  out- 
dated, the  most  current  printed  information  is  found 
in  serials  or  periodicals.  The  process  of  finding  ap- 
propriate information  begins  with  locating  the  appro- 
priate periodicals. 

There  are  various  processes  available  for  identi- 
fying desired  periodicals.  Most  libraries  have  access 
to  current  subscriptions  for  the  major  serial  publica- 
tions in  a  given  subject  area.  A  look  at  the  Table  of 
Contents  in  each  of  the  appropriate  periodicals  will 
identify  new  articles  covering  selected  areas  of  inter- 
est. For  assistance  in  selecting  periodical  titles  for 
any  subject  area,  as  well  as  obtaining  subscription  in- 
formation, Ulrich's  International  Periodicals  Di- 
rectory is  a  valuable  source  to  use.  Suggested  sub- 
ject headings  to  check  in  Ulrich's  Directory  are 
Biology-Zoology,  Environmental  Studies,  and 
Conservation. 

If  you  are  not  fortunate  enough  to  have  easy  ac- 
cess to  a  library  that  subscribes  to  the  desired  jour- 
nals, personal  subscriptions  must  be  considered. 
They  can  be  expensive,  however,  and  can  require 
time  and  space  for  maintenance. 

You  will  probably  want  to  consider  some  other 
alternatives.  One  is  to  access  Current  Contents. 
There  are  several  subject-specific  versions  of  Current 
Contents,  each  of  which  is  issued  monthly.  Each  is- 
sue reprints  the  Table  of  Contents  from  selected 
journals  within  its  specialty  area. 

Current  Contents/Agriculture,  Biology,  and 
Environmental  Sciences,  in  particular,  covers  over 
1 ,000  periodicals  and  300  books  of  interest  in  wild- 
life inventory  and  monitoring. 

Another  alternative  is  to  access  periodicals 
through  abstracting  and  indexing  services.  These  are 
subscription  services,  usually  appearing  monthly; 


they  provide  an  index  to  articles,  reports,  books,  and 
other  documents  related  to  specific  subject  areas. 
Separate  services  exist  for  most  subject  areas,  such 
as  Engineering  Index  for  engineers,  or  Chemical  Ab- 
stracts for  chemists. 

Indexes  reference  citations  under  subject  head- 
ings and  authors,  giving  full  citations.  For  periodical 
articles,  a  typical  reference  will  list  author,  article  ti- 
tle, title  of  the  periodical  in  which  it  was  printed, 
volume  number,  issue  number,  page  numbers,  and 
date  published. 

Abstracting  services  provide  brief  summaries 
about  items  in  addition  to  information  about  how 
they  can  be  located.  (Indexing  services  include  only 
title  and  bibliographic  information. )  Some  of  the 
more  useful  abstracting  and  indexing  services  to 
consult  for  a  literature  search  covering  wildlife  are 

1.  Agricultural  Index 

2.  Bibliography  of  Agriculture 

3.  Bio  Research  Index  (supplements  Biological 
Abstracts) 

4.  Biological  Abstracts 

5.  Biological  and  Agricultural  Index 

6.  Chemical  Abstracts 

7.  Ecology  Abstracts 

8.  Science  Citation  Index 

9.  Wildlife  Abstracts  (Cumulations  of  Wildlife 
Review) 

10.  Wildlife  Review 

1 1 .  Zoological  Record. 

Other  indexing  and  abstracting  services  that  can 
be  invaluable  in  a  literature  search  are 

1.  Cumulative  Subject  Index  to  the  Monthly  Cata- 
log of  U.S.  Government  Publications  1900-1971 

2.  Supplementary  Index  to  U.S.  Publications 

3.  Masters  Abstracts 

4.  Dissertation  Abstracts. 


These  services  cover  government  documents  and 
doctorate  and  masters  theses.  The  indexed  publica- 
tions may  be  published  or  unpublished  documents.  A 
biologist  should  not  overlook  the  usefulness  of  these 
types  of  publications  and  other  unpublished  mate- 
rials that  may  be  identified  through  other  sources, 
such  as  peer  contacts. 

Each  of  these  services  covers  a  different  selec- 
tion of  publications  that  deal  with  the  subject  of 
wildlife,  including  inventory  and  monitoring.  The 
BLM  Library,  any  large  public  library,  or  a  technical 
research  library  can  help  you  access  these  sources. 
(See  Appendix  I  for  more  detailed  information  on 
these  services.) 


30 


Literature  Review 


Bibliographies 

Bibliographies  covering  specific  subject  areas 
cannot  be  overlooked  as  sources  of  information.  De- 
pending upon  the  date  of  compilation,  a  bibliography 
may  be  excellent  for  background  information  if  the 
subject  in  question  must  include  older,  but  still  ap- 
plicable, materials.  Some  useful  bibliographies  for 
wildlife  subject  areas  are 

1.  U.S. — Canadian  Range  Management,  1935—1977: 
A  Selected  Bibliography  on  Ranges,  Pastures, 
Wildlife,  Livestock,  and  Ranching; 

2.  U.S. — Canadian  Range  Management,  1978-1980: 
A  Selected  Bibliography  on  Ranges,  Pastures, 
Wildlife,  Livestock,  and  Ranching;  and 

3.  Wildlife  Abstracts — the  entire  series,  from  1935 
through  1980. 

Other,  more  subject-specific  bibliographies  (exam- 
ple, a  bibliography  on  the  pronghorn  antelope)  are 
valuable  resources.  They  can  be  found  in  a  variety  of 
places  within  library  collections. 

Another  often-overlooked  source  of  bibliogra- 
phies is  the  references  listed  at  the  end  of  the  article 
or  report  the  biologist  already  has  on  an  individual 
topic. 

Once  an  article  is  identified,  a  biologist  can  re- 
quest a  copy  of  the  article  from  the  author,  library, 
or  other  appropriate  organization.  Preprinted  reprint 
request  cards  are  useful  for  this  purpose. 

Some  of  the  major  journals  published  in  wildlife 
research  are  outlined  in  Appendix  II.  Not  listed  be- 
cause of  their  number  are  journals  that  are  specific 
to  a  State;  examples  are  California  Fish  and  Game 
and  Wyoming  Wildlife.  These  journals  are  highly  use- 
ful even  for  areas  outside  of  the  specified  State.  They 
can  be  identified  through  the  use  of  indexes  at  the 
BLM  Library  or  other  large  library  facility. 


The  use  of  bibliographic  computer  data  bases  is 
the  most  efficient  method  to  develop  current  bibli- 
ographies on  any  topic.  Online  data  bases  have  ex- 
isted since  the  mid  1960s,  but  they  were  primarily 
private  files  accessible  only  to  special  groups.  In  the 
early  1970s,  online  literature  searching  data  bases 
were  made  available  to  the  general  public,  thus  limit- 
ing the  time-consuming  step  of  manually  searching 
tables  of  contents,  indexes,  and  abstracts.  With  a  few 
exceptions,  coverage  of  data  bases  is  generally  lim- 
ited to  literature  from  the  1970s  to  the  present. 
Therefore,  any  researching  into  literature  published 
before  1970  must  be  done  manually. 

To  access  the  hundreds  of  data  bases  available 
requires  the  use  of  a  vendor  with  a  software  system 
which  provides  the  commands  for  searching.  The 
three  most  frequently  used  vendors  for  literature 
searching  are 

DIALOG — operated  by  the  Lockheed  Corporation, 

ORBIT — operated  by  SDC  Search  Service,  and 

BRS — operated  by  the  Bibliographic  Retrieval  Ser- 
vices Company. 

Since  knowledge  of  software  is  necessary  to  search 
efficiently,  most  researchers  access  these  systems 
through  trained  librarians.  For  example,  the  com- 
puter version  of  Biological  Abstracts  on  DIALOG 
goes  back  to  1969,  whereas  the  printed  version  be- 
gan publication  in  1927.  Experience  has  shown  that 
researchers  sometimes  become  so  impressed  with 
the  computer  that  they  tend  to  forget  that  not  all  the 
published  literature  is  contained  within  its  files.  If 
you  desire  a  comprehensive  search,  be  sure  to  use 
both  manual  and  computer  searches.  Computer 
searches  may  be  adequate  if  only  recent  data  are 
desired. 


COMPUTER  SYSTEMS 

After  a  researcher  has  reviewed  the  appropriate 
manually-searchable  sources,  the  next  step  is  to  con- 
sult computer  data  bases.  These  data  bases  generally 
fall  into  two  major  categories — bibliographic  and 
nonbibliographic. 

Bibliographic  data  bases  produce  results  similar 
to  searching  the  indexing  and  abstracting  services 
previously  mentioned.  These  data  bases  contain  cita- 
tions to  specific  articles,  books,  and  reports,  and  can 
be  searched  by  keywords,  title  words,  authors,  or 
other  appropriate  identifiers.  Nonbibliographic  data 
bases  differ  in  that  they  contain  actual  data  and  usu- 
ally do  not  refer  the  user  to  a  printed  source. 


BLM  librarian  performing  online  computer  search. 


Literature  Review 


31 


For  interactive  searching,  librarians  use  termi- 
nals and  communicate  directly  with  computers  to 
get  the  desired  citations.  Searches  for  wildlife  infor- 
mation will  most  frequently  be  performed  in  the 
data  bases  AGRICOLA,  BIOSIS,  CAB,  and  CRIS,  ex- 
panding to  other  sources  depending  upon  the  topic. 

Commercial  sources  represent  only  a  portion  of 
the  many  computer  sources  which  provide  informa- 
tion on  wildlife  inventory  and  monitoring.  Other 
sources  of  interest  may  or  may  not  be  bibliographic, 
and  there  may  be  special  requirements  for  access. 
Noncommercial  bibliographic  files  are  basically 
structured  the  same  as  the  commercial  sources  dis- 
cussed above.  Data  bases  are  often  maintained  and 
updated  by  the  originating  organizations,  and  access 
to  the  data  bases  is  made  available  only  through 
their  approval. 

Several  noncommercial  data  bases  have  been  de- 
veloped by  Federal  agencies  to  assist  in  meeting 
their  missions.  Information  on  specific  noncommer- 
cial information  systems  can  be  found  in  DuBrock  et 
al.  (1981).  Three  data  bases  of  primary  interest  for 
inventory  and  monitoring  of  wildlife  habitat  are  the 
Fish  and  Wildlife  Reference  Service,  RUNWILD,  and 
the  Procedures  System.  Other  data  bases  pertaining 
to  wildlife-related  subjects  are  summarized  in 
Appendix  III. 


Fish  and  Wildlife  Reference  Service  (FWRS) 


FWRS  is  a  computer  index  of  documents  from 
the  Federal  Aid  in  Fish  and  Wildlife  Restoration  Pro- 
gram (Pittman-Robertson  and  Dingell-Johnson  Acts), 
the  Anadromous  Fish  Conservation  Program,  the  En- 
dangered Species  Grant  Program,  the  Cooperative 
Fishery  and  Wildlife  Research  Units,  and  State  game 
and  fish  agencies.  The  data  base  does  not  include 
most  routine  publications  of  the  U.S.  Fish  and  Wild- 
life Service. 

The  system  allows  access  to  research  informa- 
tion from  the  State  level  by  providing  an  index  to  se- 
lected State  agency  technical  reports.  Knowledge  of 
indexed  publications  can  be  gained  from  the  FWRS 
quarterly  newsletter,  which  lists  new  publications,  or 
by  obtaining  a  computer  search  that  will  also  include 
older  publications.  A  computer  search  can  be  or- 
dered directly  from  FWRS;  there  is  a  charge  unless 
the  requester  is  employed  by  the  U.S.  Fish  and  Wild- 
life Service  or  a  cooperating  agency.  This  data  base  is 
available  as  a  private  file  on  the  DIALOG  System.  The 
BLM  Library  has  received  permission  to  access  it 
through  DIALOG  and  can  accomplish  searches  for 
BLM  employees. 


RUNWILD 

RUNWILD  is  a  storage  and  retrieval  system  con- 
taining wildlife  information  that  can  be  accessed 
from  three  levels:  inventory,  species-habitat  associa- 
tions, and  management  data.  Developed  in  1973  by 
the  USDA  Forest  Service,  it  is  stored  on  a  computer 
in  Fort  Collins,  Colorado.  RUNWILD  can  be  used  for 
875  vertebrate  species  and  subspecies  in  the  South- 
west, but  its  format  is  applicable  to  any  geographical 
area.  Casner  et  al.  1978,  Lehmkuhl  and  Patton  1982, 
and  Patton  1978,  1979a,b  provide  details  of  the 
system. 

Access  to  RUNWILD  is  available  to  any  Federal 
employee  and  can  be  obtained  by  contacting  the 
USDA,  Fort  Collins  Computer  Center  (FCCC)  to  es- 
tablish an  account.  They  will  supply  a  form  that  re- 
quires the  requester  to  provide  information  neces- 
sary for  computer  access  codes  and  passwords. 
Accounting  and  use  information  is  periodically  is- 
sued to  the  user  from  the  FCCC.  Each  account  num- 
ber is  billed  for  use  charges  as  necessary.  The  data 
base  manager,  located  in  the  Region  3  Forest  Service 
in  Albuquerque,  New  Mexico,  is  the  contact  person 
for  questions  about  the  system  and  for  obtaining  a 
data  base  password.  Because  this  is  a  nonbiblio- 
graphic  data  base  and  the  user  is  usually  searching 
for  specific  raw  data,  the  BLM  Library  does  not  per- 
form these  searches  for  users.  I  recommend  that  one 
account  be  opened  for  use  by  interested  personnel 
at  a  specific  office. 

New  additions  and  updates  are  being  made  to 
the  file  on  a  continuing  basis.  RUNWILD  II  was  made 
available  in  November  1978  and  RUNWILD  III  in 
1982.  The  data  base  contents  have  been  made  avail- 
able on  microfiche  through  the  National  Technical 
Information  Service  (NTIS).  Information  on  the  use 
of  the  microfiche  is  given  in  Patton  ( 1979a). 


Procedures 

"A  Procedure  for  Describing  Fish  and  Wildlife" 
(nicknamed  Procedures)  was  developed  by  the  U.S. 
Fish  and  Wildlife  Service  for  uniformly  describing 
characteristics  of  fish  and  wildlife  populations 
(Cushwa  et  al.  1980;  Mason  et  al.  1979).  The  system 
outlines  procedures  for  collecting  ecological  facts 
concerning  species  of  vertebrates  and  selected  inver- 
tebrates and  can  be  applied  to  information  gathering 
for  species  in  any  area.  The  Procedures  system  has 
been  used  in  data  collection  in  several  States,  includ- 
ing Pennsylvania,  Colorado,  Wyoming,  Missouri,  Illi- 
nois, Virginia,  Kentucky,  and  Tennessee.  The  data 
bases  are  administered  by  the  game  and  fish  agency 
for  the  State  involved,  and  access  is  obtained  by  con- 
tacting the  data  base  manager  at  that  organization  to 
set  up  an  account  for  interactive  searching  or  for  re- 
questing a  batch  search. 


52 


Literature  Review 


The  Procedures  system  provides  a  useful  and 
valuable  methodology  for  maintaining  fish  and  wild- 
life species  data  regardless  of  the  geographic  area 
being  considered.  The  system  summarizes  informa- 
tion on  wildlife  habitat  use  and  species  distribution, 
status,  and  life  history.  Workbooks  that  provide  spe- 
cific coding  descriptions  are  used  when  preparing 
data  for  input  into  the  data  bases.  The  standard  clas- 
sifications required  for  coding  the  information  are 
widely  accepted,  recognized  standards  that  are  easily 
understood  and  applied.  For  categories  in  which 
there  were  no  accepted  classification  standards  al- 
ready in  use,  standards  were  developed  and  applied. 


ACCESSING  BIBLIOGRAPHIC  COMPUTER 
DATA  BASES 

Data  bases  may  be  accessed  by  anyone  who  has 
a  terminal  and  has  made  the  necessary  financial  ar- 
rangements. I  recommend,  however,  that  biologists 
obtain  bibliographic  information  through  the  ser- 
vices of  a  librarian  or  information  specialist.  These 
professionals  are  trained  in  techniques  necessary  for 
efficiently  and  effectively  using  the  variety  of  subject 
data  bases  available.  They  also  keep  current  on  data 
base  changes  by  reading  literature  and  attending  up- 
dated training  sessions  for  experienced  searchers.  Bi- 
ologists can  best  use  their  time  reviewing  the  sub- 
ject bibliographies  that  librarians  provide  for 
publications  of  interest. 

Many  organizations  perform  literature  searches 
upon  request.  Most  large  public  libraries  and  aca- 
demic libraries  offer  the  service.  In  most  large  cities, 
there  are  commercial  organizations  that  advertise 
this  kind  of  service.  These  sources  usually  charge  ac- 
tual costs  for  communications  and  offline  prints  and 
add  a  percentage  as  a  service  charge.  Some  private 
individuals  also  provide  computer  literature  search- 
ing for  a  fee.  Consequently,  anyone  can  have  access 
to  bibliographic  data  bases  and  thus  to  current  re- 
search data.  The  key  is  knowing  where  to  find  the 
organizations  or  individuals  that  will  perform  the 
searching. 

Within  BLM,  access  to  computer  literature 
searching  services  is  available  through  the  BLM  Li- 
brary. The  Library  maintains  contracts  with  DIALOG, 
SDC,  and  BRS,  enabling  access  to  more  than  250  data 
bases.  The  requester  simply  telephones  the  BLM  Li- 
brary with  details  of  the  request  or  sends  a  com- 
pleted Library  Search  Request  Form  to  the  Library 
(Figure  1). 

To  conduct  an  effective  search,  a  librarian  must 
obtain  specific  information  from  the  requester  con- 
cerning the  data  needed.  In  addition  to  a  general 
statement  of  the  overall  objective  of  the  search  re- 
quest, the  requester  must  provide  key  words.  Key 
words  are  any  meaningful  words  that  are  important 


to  the  topic  being  researched.  They  are  terms  or 
phrases  that  must  be  present  in  the  citations  for  the 
resultant  bibliography  to  be  pertinent  to  the  search 
request.  Names  of  authors  and  organizations  may 
also  serve  as  specific  keys  upon  which  a  search  can 
be  based. 

Librarians  develop  search  strategies  from  the  re- 
lationships that  exist  among  key  words,  grouping  like 
concepts.  Search  strategies  are  executed  in  various 
data  bases  that  cover  subjects  of  interest,  and  bibli- 
ographies are  returned  to  requesters.  Employees 
then  have  the  opportunity  to  review  the  listings, 
mark  those  items  of  interest,  and  return  the  lists  to 
the  Library.  The  BLM  Library  then  obtains  copies  or 
loans  of  the  desired  items.  (Appendix  I  describes 
some  of  the  data  bases  available  through  the  BLM  Li- 
brary that  contain  wildlife  information. ) 


OTHER  INFORMATION  SOURCES 

Information  from  other  than  published  and  elec- 
tronic information  sources  can  be  used  to  great  ad- 
vantage in  research  for  wildlife  inventory  and  moni- 
toring studies.  Two  major  areas  to  be  considered 
will  be  discussed  in  the  remainder  of  this  chapter. 

Museums 

Museum  data  can  be  a  useful  source  of  informa- 
tion for  the  wildlife  biologist  (See  Figure  2).  Several 
museums  have  important  systematic  biological  re- 
search resources  that  can  be  used  for  natural  history 
information  on  specimens.  Of  special  interest  in  the 
consideration  of  this  data  source  is  information  on 
species  location.  Information  such  as  place  of  collec- 
tion, date  of  collection,  specific  locality,  and  county 
are  required  in  the  records  for  a  species. 

Museum  collections  that  would  be  useful  for 
wildlife  inventory  research  are  found  in  many  differ- 
ent institutions  (see  Appendix  IV).  These  organiza- 
tions produce  many  publications  that  contain  infor- 
mation about  their  collections.  Information  regarding 
what  they  publish  can  be  obtained  from  a  variety  of 
sources,  mentioned  by  Moore  (1980).  Assistance  in 
obtaining  any  of  the  references  mentioned  can  be 
obtained  from  a  large  public  or  a  technical  library. 
BLM  personnel  should  seek  help  from  the  BLM  Li- 
brary at  the  Service  Center. 

Once  a  museum  source  is  identified,  it  is  possi- 
ble to  locate  catalogs  of  that  museum's  holdings  or 
specimen  listings.  One  can  then  access  the  museum's 
records  covering  the  species  or  geographical  areas  of 
interest.  The  searcher  should  be  aware  of  a  major 
problem  that  sometimes  exists  with  museum  data.  If 
only  general  localities  are  recorded  for  an  observa- 
tion, the  usefulness  of  the  record  for  studies  of  distri- 
bution and  habitat  may  be  lost.  Accurate  association 


Literature  Review 


33 


J 


4 


M/ 


Form  1279-2 
(April  1979) 


UNITED  STATES 
TMENT  OF  THE  W 
BUREAU  OF  LAND  MANAGEM 


DEPARTMENT  OF  THE  INTERIOR  ^    J--     V\  i..     \ 


lEQUESr^     ky   nA 


ftp  ^ 

w 


1.    Requested  by  (name)   £^J^    g ' /^ZJ^, 


Date 


9^^/ 7,  /far 


Office*  and  Division 


relepnone  (include  nieti  < 


£  t\ 


Telepffone  (include  aika  code) 

Oo'l)2V?-  757ft 


a.    Needed  by  ^,^^,/^r 


'        A    ,  z.    Subject  of  search  (describe) 

I  3.    Give5  specific  terms  and/or  phrases  that  are  considered  relevant.    Include  synonyms,   closely  related  phrases,   and 

generic  names  when  appropriate.    Use  scientific  and  technical  as  well  as  common  vocabulary.    (Attach  additional 
sheet,  if  necessary.)  ■ 

rgC\llDE    4£   HhHy^fc^t^,     dsiUU-rLfa ,    rruw^c^,   tYUun^TOfy    ^fattas, 

pftf  C  l  PZ.  I    fa  >State  limitations  desired  (e.g.  geographical,  language,  dates,  etc.) 


5.    Give  two  or  three  references  already  located  which  you  consider  relevant  to  the  subject  requested. 

iUO'tdt    A.L.     Icl7(j,.     £lMs    bt-hciis-Lrt-    i^   Su.£att.e-*s  1~o 


^ uUtcruo  T^oul^  ^curiae*    £r^    &bi*rTh,Cj,<*£Ci.eU.'   UJiftCrrv-nty- 


n&A/t    7JU.IAA,    r<Uvr<UiU<u    <%*i,     UQV^U  jrf^ 


P- 

-me    ^ta^cAor,    trvrf  ftu^  tuxt  not  rttusurJii . 

6.    Desired  scope  (check  appropriate  item) 

[T^fBroad  search  designed  to  retrieve  as  many  as  possible  of  the  relevant  citations,  but  which  might  also  retrieve 
many  irrelevant  citations. 
J  Narrow  search  designed  to  retrieve  some  of  the  relevant  citations,  but  with  few  accompanying  irrelevant 
citations. 


Remarks        JL    ,/W^       L 


:-y\  L*^- 


GPO   650-423 


Figure  1.  A  completed  BLM  Library  Search  Request  Form. 


34 


Literature  Review 


MUSEUM    OF     VERTEBRAU     ZGJLGGY    DATA    FUR    fclRDS    CGLLEUEU     IN    CULURAUO 


Ob/v-7/bb 


PAGt     1 


NC.    OF     ITLMS    IN    CUtRY    RESPONSE:     7 
NO.    OF     ITEMS    IN    DATA    BANK:    2743b 
PERCENTAGE    OF    kESPGhSE/TGTAL    DATA    BANK: 


0.03* 


x*v 


PARUL1NAE 

DENDR01CA         CORONATA 

MVZ  CAT. NO.:   37399 
SEX:  h       age  :  juv. 

COUNTRY:    COLORACO 


MEMGRABlLlS 
SKIN:     STUDY  SKEL.:    NONE 

COLLECTOR:    DILLE    FRED    M. 

STA*t"/CO./PRLV. 


CULL  .n  j.  ; 
LARlMtR    CU. 


COLL. NO. 
LARIMER    CO. 


ELEV:         ;    LOCALITY:    ESTES    PARK 

DATE:      4    AUGUST  19J?      M.: RlFRC.: 

FAT: MOLT:YES      S  TOM.COM  .  :NG        CLLORS:nO 

REMARKS:    ED:MGLT-StE    SPECIMEN. 
DENDR01CA  CGRUNATA  MEMLRAblLlS 

MVZ    CAT. NO.:       44533       SKIN:     STUDY  SKEL.:    NONE 

SEX:    K  AGE     :    U    AD.  COLLECTOR:    DILLE    FRED    Ml 

COUNTRY:    COLORADL  STAJE/CC  ./PROV , 

llev:        ---   --:    LOCALITY:    ESTES   park 

date:     6  july  1904     mt.:  reprc. 

fat:  molt:no       stum.ccnt . :n0      colcrs:no 

remarks:  ed:mst. 
oehorgica  coronata  mem0rab1l1s 

mvz  cat. no.:     78903     skin:   stuoy          skel.:  n[ 
sex:  f       age   :  u  ad.                                collector  :f£ktr  f.h.  coll. no.: 

country:  colorado  ,, state/co. /prov . :   larimtk  co. 

ELEV:  —    — ;    LOCALITY^fBERTHCUD   t 

DATE:    18    AUGUST  1897  ^W I ■ I  m"  REPRC. 

FAT: MOLT:YES      ST0M.COM.  :N0        COL0RS:NO 

REMARKS:    EO:MOLT-SEE    SPECIMEN. 
N1LS0MA  PUSILLA  P1LE0LATA 

MVZ    CAT. NO.:       76915)       SKIN:     STUOY  SKEL.:    NONE 

SEX:    L>         AGE     :    U  COLLECTOR:    FOWLER    F.H. 


COUNTRY: 

COLORADO 

ELEV: 

—  --; 

OATE: 

9    JULY 

FAT: 



COLL. NO.  ! 
LARIMER    CO. 


STATE/CC./PROV.: 

LOCALITY:    LONG'S    PEAK;    ROCKY    MOUNTAIN    NAT.    PARK 

1899      NT.:    kEPRO.:    

MOLT:NO        STQM.COM.:NO        COLORS:no 
remarks:  basin  above  timberline 
n1ls0ma  pusilla  pileolata 

mvz  cat. no.:     7e916     skin:   study  skel.:  nene 

SEX:    M        AGE    :    U  COLLECTOR:    FO*LER    F.H.  COLL.nu.: 

COUNTRY:    COLORADO  ST ATE/CO ./PROV . :    LARIMER    CO. 

ELEV:  --;    LOCALITY:    LONG'S    PEAK;    ROCKY    MOUNTAIN    NAT.    PARK 

DATE:     10    JULY  1899       HT . :     REFRU.:    

FAT: MOLT  i  Hi    — fcimj.CCM.:NO        COlCRS:nO 

REMARK  S/*6AS  IN    AEOVE    TIMBERLINE^ 
TURD IN AE 
MYADESTES  TCWNSENDI  TCmRSENCI 

MVZ    CAT.NC.:    107381       SKIN:    STUDY  sKi.:    NCNE 

SEX:    M  AGE     :    U  COLL  EC  TOR  r^^-  LULL.Nl| 

COUNTRY:    COLORACO  ^,-"-»^i^    ST  ATL/CO./PKOV  .  :    LAklMtk    CU  .     ..fl^ 

ELEV:   ;  locality:  lovelanc 

DATE:    --    MAY  1904      NT.:    kEPRL.: 

FAT:    ---  MQL1»NU       _ST0M.COM  .  :N0        C0LCRS:nO 

REM/ 


MQLl»NO  S 


U.: 

LAKlMcR    CU.    _.A\ 


MUiLl 


rt^bM     ZGjLGGY    DATA    FOR     clRDS    CCLLiCTtU    IK    COLORADO 


Ot>/C7/6b 


PAGt     _ 


TURD1NAE 

S1ALIAX  LUkRUCulb-S 

MVZ    CAT.NC.:       396t>i 
JEX:    F         AGE     :    U 

jUNTRY:    COLORADO 

■  i   LUCnia 

fit  :        1    APRIL  1900 

-f  AT  i  -■■  ' 

FEMARKS:     


SKIN:     STUOY  SKtL.:     NCNL 

CCLLEC10F:    DILLE     F.M, 

FT. COLL  INS 

M.: KEFKL.: 

NO       stom.ccm.:no       culcrs:nl 


STATE/Cu./PkLV.: 


CULL  .N^i. 
LARIMER  CU. 


144 


20t) 


it>2 


318 


319 


Figure  2.  Major  scientific  museums  have  computer  bases  on  their  collections  that  contain  valuable 

information  on  species  distribution;  an  example  from  the  Museum  of  Vertebrate  Zoology  at  the  University 
of  California,  Berkeley  is  shown  here. 


Literature  Review 


55 


of  the  specimen  with  a  particular  habitat  type  may 
be  impossible.  The  site  named  may  be  some  distance 
from  the  actual  collection  or  observation  point,  or, 
in  some  instances,  more  than  one  point  in  the  gen- 
eral area  might  have  the  same  name.  Care  should 
also  be  exercised  in  checking  the  records  to  be  sure 
they  correctly  represent  the  specimen  being 
reported. 

Personal  Contacts 

The  value  of  knowing  who  is  prominent  in  a 
specific  area  of  research  cannot  be  overlooked  as  a 
potential  source  for  valuable  information.  Generally 
these  specialists  are  willing  and  anxious  to  tell  any- 
one about  their  research  or  send  copies  of  their  pub- 
lications upon  request.  A  way  to  identify  specialists 
is  to  make  note  of  recurrent  names  in  subject  bibli- 
ographies. Libraries  can  conduct  author  searches  us- 
ing the  data  bases  discussed  earlier;  thus,  a  bibliog- 
raphy of  writings  done  by  a  specific  person  can  be 
generated.  In  addition,  the  CRIS  Data  Base  and  the 
Federal  Research  in  Progress  Data  Base,  which  are 
available  through  DIALOG,  are  data  bases  that  cover 
current  research  information.  To  learn  more  about 
the  research  project  listed  there,  the  specialist  can 
contact  the  researcher  whose  name  is  given. 

The  key  step  in  using  this  method  is  locating 
the  specialist.  Some  useful  non-computer  tools  in 
your  library  that  can  correctly  identify  and  locate  an 
individual  are  -- 

•  directories  of  organizations 

•  information  in  proceedings  of  meetings,  sympo- 
sia, and  conferences 

•  subject  matter  directories  such  as  the  Conserva- 
tion Directory 

•  who's  who-type  publications 

•  introductory  remarks,  short  biographical  infor- 
mation, and  title  page  information  in  some  pub- 
lications 

•  inquiries  to  people  who  have  already  been  iden- 
tified. 


In  many  instances,  conversation  with  a  colleague  can 
lead  to  information  sources  that  would  be  impossible 
to  otherwise  identify. 


The  telephone  becomes  an  important  research  tool. 


SUMMARY 

The  importance  of  thorough  literature  searches 
cannot  be  overemphasized.  When  biologists  have 
conducted  studies  that  are  similar  to  yours,  you  can 
realize  considerable  savings  in  time  and  resources.  In 
some  instances  there  may  be  little  or  no  information 
available  in  response  to  an  inquiry,  but  that  in  itself 
is  useful  to  know.  If  no  previous  studies  exist,  the  re- 
sults of  your  study  may  be  invaluable. 

One  of  the  major  focuses  of  libraries  today  is  in- 
terlibrary  cooperation  and  shared  resources.  Even  if 
one  has  no  idea  where  to  begin  to  look  for  some 
person,  publication,  or  resource,  a  check  with  your 
library  or  information  center  is  a  good  way  to  begin. 

Most  large  libraries  and  many  technical  libraries 
have  computer  services  that  enable  them  to  identify 
book  and  periodical  holdings  of  other  libraries 
throughout  the  Nation.  These  services  will  save  time 
and  increase  the  thoroughness  of  literature  reviews. 
At  the  BLM  Library,  the  needs  of  field  employees  are 
given  highest  priority.  For  library  and  information 
needs,  it  is  the  place  to  begin. 


36 


Literature  Review 


LITERATURE  CITED 


CASNER,  W.B.,  B.  KULONGOWSKI,  DR.  PATTON,  and  S.J. 
PINKERTON.  1978.  RUNWILD— For  the  UNIVAC 
1100  Series:  implementation  and  maintenance.  U.S. 
Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep.  RM-51A.  31pp. 

CUSHWA,  C.T.,  C.W.  DUBROCK,  N.D.  GLADWIN,  G.R. 
GRAVATT,  R.C.  PLANTICO,  R.N.  ROWSE,  and  LJ. 
SLASKI.  1980.  A  procedure  for  describing  fish  and 
wildlife  for  Pennsylvania:  summary  evaluation  report. 
U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  FWS/OBS-79/19A. 
15pp. 

DUBROCK,  C.W.,  D.N.  GLADWIN,  W.T.  MASON,  Jr.,  and 
C.T.  CUSHWA.  1981.  State-of-the-art  of  fish  and  wild- 
life species  information  systems  in  the  United  States. 
Trans.  North  Am.  Wildl.  Nat.  Resour.  Conf.  46:156- 
170. 

LEHMKUHL,  J.F.  and  DR.  PATTON.  1982.  RUNWILD 
Wildlife/habitat  relationships:  user's  manual  for  the 
RUNWILD  III  data  storage  and  retrieval  system.  U.S. 


Dep.  Agric,  For.  Serv.  Southwestern  Region.  Wildl. 

Unit  Tech.  Rep.  68pp. 
MASON,  W.T.  Jr.,  C.T  CUSHWA,  LJ.  SLASHI,  and  D.N. 

GLADWIN.  1979.  A  procedure  for  describing  fish  and 

wildlife:  coding  instructions  for  Pennsylvania.  U.S. 

Dep.  Inter.,  Fish  and  Wildl.  Serv.  FWS/OBS-79/19. 

21pp. 
MOORE,  J. L.  1980.  Wildlife  management  literature.  Pages 

7.38  in  S.D.  SCHEMNITZ,  ed.  Wildlife  management 

techniques  manual.  4th  ed.  Wildl.  Soc.  Washington, 

DC.  686pp. 
.  1978.  RUNWILD:  a  storage  and  retrieval  system  for 

wildlife  habitat  information.  U.S.  Dep.  Agric,  For.  Serv. 

Gen.  Tech.  Rep.  RM-51.  8pp 
PATTON,  DR.  1978.  RUNWILD:  a  storage  and  retrieval 

system  for  wildlife  habitat  information.  U.S.  Dep. 

Agric,  For.  Serv.  Gen.  Tech.  Rep.  RM-51.  8pp. 
.  1979a.  How  to  use  RUNWILD  data  files  stored  on 

microfiche.  U.S.  Dep.  Agric,  For.  Serv.  Res.  Note  RM- 

377.  2pp. 
.  1979b.  RUNWILD  II:  a  storage  and  retrieval  system 

for  wildlife  data.  Trans.  North  Am.  Wildl.  Nat.  Resour. 

Conf.  44:425-430. 


Literature  Review 


37 


APPENDIX  I.  BLM  Library 


History 

The  BLM  Library  was  established  as  part  of  the 
BLM  Service  Center  in  1970.  The  Library's  services 
continue  to  grow  with  advances  in  library  technol- 
ogy and  information  management.  Today  the  BLM  Li- 
brary offers  a  full  range  of  library  services,  including 
reference,  circulation,  aid  in  acquisition/purchasing, 
cataloging,  interlibrary  loan,  plus  automated  and 
manual  literature  searching. 

Library  Collection 

The  BLM  Library  contains  more  than  30,000  vol- 
umes and  subscribes  to  over  200  periodicals.  This 
collection  covers  a  wide  range  of  subjects  covering 
all  aspects  of  land  management,  research  and  devel- 
opment, computer  science,  and  administration.  The 
library's  collection  is  arranged  according  to  the  Li- 
brary of  Congress  Classification  System  and  is  cata- 
loged through  the  online  shared  cataloging  system  of 
Online  Computer  Library  Center  (OCLC)  Inc.,  a  so- 
phisticated online  computer  network  that  links  over 
2000  participating  libraries. 

Loans 

The  BLM  Library  can  deliver  any  published  ma- 
terial to  field  offices.  Field  personnel  are  encouraged 
to  use  their  local  resources  first;  if  a  publication  is 
not  easily  identified  or  obtained,  they  should  contact 
the  BLM  Library  for  assistance. 

Most  materials  in  the  Library's  collection  can  be 
loaned  for  4  weeks.  If  the  item  is  small,  a  photocopy 
may  be  made  and  sent  to  the  requester  for  retention. 

If  the  publication  you  need  is  not  owned  by  the 
Library,  it  will  be  located  at  another  library  and  bor- 
rowed on  interlibrary  loan.  If  the  publication  is  avail- 
able free  of  charge,  you  may  be  provided  with  the 
information  necessary  to  obtain  your  own  copy. 

Literature  Searching 

The  BLM  Library  has  literature  searching  capa- 
bilities that  enable  the  staff  to  access  over  250  data 
bases  and  quickly  prepare  bibliographies  on  almost 
any  subject.  This  service  is  available  to  all  BLM  per- 
sonnel and  may  be  requested  by  filling  out  a  "Library 
Search  Request"  Form  (BLM  Form  No.  1279-2),  or 
by  calling  the  Library. 

The  requester  provides  the  librarian  with  a  gen- 
eral statement  of  the  information  needed  and  with 
key  words  that  can  be  used  in  building  a  search 
strategy  for  entering  data  bases.  The  key  words  pro- 
vided are  important  in  that  they  determine  how  suc- 


cessful the  search  will  be.  Terms  must  be  pertinent 
to  the  subject  area  to  locate  citations  on  the  desired 
request. 

Reference 

A  full  range  of  reference  services  is  provided  by 
the  Library  staff.  Requests  involve  short,  specific  in- 
formation questions  that  may  be  answered  by  phone 
or  general  reference  questions  that  require  more  ex- 
tensive searching.  Comprehensive  information  re- 
quests require  identification  and  location  of  mate- 
rials; response  time  will  vary  from  several  hours  to 
many  days. 

Referral 

Often  the  most  current  or  relevant  information 
you  seek  may  not  yet  be  published  and  may  be  avail- 
able only  from  a  specialist.  The  Library's  staff  can  lo- 
cate researchers  or  organizations  that  specialize  in 
your  subject  area.  The  staff  has  a  good  knowledge  of 
potential  referrals  from  the  personal  contacts  they 
make,  and  from  the  computerized  networks  they 
access. 


Current  Awareness 

The  Library's  current  awareness  services  are  an 
extension  of  literature  searching.  Users  can  be  kept 
current  in  selected  subject  areas  through  computer 
based  services  called  Selective  Dissemination  of  In- 
formation (SDI).  Search  strategies  are  stored  and 
saved  as  SDIs  in  selected  data  bases.  As  updated  ma- 
terials are  added  to  these  data  bases,  the  SDI  pro- 
grams are  executed  and  the  Library  is  provided  with 
offline  printouts  that  are  forwarded  to  the  requester. 


Special  Collections 

Presidential  Executive  Orders  (Numbered)  June 
1985-Present 

Federal  Register-All  volumes  available — 
1936-Present 

Code  of  Federal  Regulations  (All  titles) 

Memoranda — Instruction  and  Information — 
Washington  Office:  1964-Present 

BLM  Service  Center:  1965-Present 


38 


Literature  Review 


Wildlife  information  data  bases  available  through  the  BLM  Library. 


Data  Base 

Description 

Coverage 
Dates 

Producer 

Update 
Frequency 

Comments 

AGRICOLA 

The  cataloging  and 
Indexing  data  base  of 
the  National 
Agricultural  Library; 
contains  publications 
of  the  USDA  Forest 
Service.  Covers 
worldwide  journal  and 
monographic  literature 
in  agriculture  and 
related  subjects, 
including  animal 
science,  forest  and 
plant-related  areas, 
entomology,  and 
agricultural 
engineering 

1970-Present 

National 

Agriculture 

Library 

Monthly 

Records  are  citations 
only. 

BIOSIS 
Previews 

Covers  all  aspects  of 
the  life  sciences, 
drawing  on  all  forms  of 
original  published 
literature  for  citations. 
Corresponds  to  the 
printed  publications 
Biological  Abstracts 
and  Biological 
Abstracts/RRM 
(formerly  entitled 
Bioresearch  Index), 

1969-Present 

Bioscience 
Information 
Service 

Semimonthly 

Records  primarily 
citation  only,  but 
abstracts  given  for  the 
Biological  Abstracts 
records  from  July  1976 
to  the  present. 

CAB 
Abstracts 

A  comprehensive  file  of 
agricultural  and 
biological  information 
containing  all  records 
in  the  26  main  abstract 
journals  published  by 
the  Commonwealth 
Agricultural  Bureau. 
Other  publications, 
including  books  and 
journals,  are  scanned. 
In  some  instances,  less 
accessible  literature  is 
abstracted  by 
scientists  working  in 
other  countries. 
Included  in  this  data 
base  are  Forestry 
Abstracts. 

1972-Present 

Commonwealth 

Agricultural 

Bureau 

Monthly 

Significant  papers 
abstracted;  less 
important  works 
reported  as  citations 
only. 

Literature  Review 


39 


Wildlife  information  data  bases  available  through  the  BLM  Library  (continued). 


Data  Base 

Description 

Coverage 
Dates 

Producer 

Update 
Frequency 

Comments 

Comprehensive 
Dissertation 
Index  (CDI) 

Provides  access  to 
records  of  more  than 
1,000  scientific  and 
technical  papers 
presented  at  over 
1,000  major  regional, 
national,  and 
international  meetings 
each  year.  Primary 
subject  areas  covered 
include  life  sciences 
and  chemistry. 

1973-Present 

Cambridge 

Scientific 

Abstracts 

Monthly 

Records  are  citations 
only,  but  availability 
and  ordering 
information  also  given. 

Current 

Research 

Information 

Service 

(CRIS) 

A  current  awareness 
data  base  for 
agriculturally-related 
research  sponsored  or 
conducted  by  U.S. 
Department  of 
Agriculture  research 
agencies,  State  forestry 
schools,  and  other 
cooperating  State 
institutions.  Currently 
active  and  recently 
completed  projects 
(within  the  last  2  years) 
are  in  the  file.  Included 
among  the  subjects 
reported  are  biological 
sciences,  natural 
resource  conservation 
and  management,  and 
environmental 
protection. 

Last  two  years 

U.S.  Department 
of  Agriculture, 
Washington,  DC 

Quarterly 

Abstracts  included  with 
the  records. 

ENVIROLINE 

Covers  more  than 
5,000  international 
primary  and  secondary 
source  publications 
reporting  on  all 
aspects  of  the 
environment.  Has  a 
section  on  wildlife 
publications,  as  well  as 
many  other  subject 
overages. 

1971 -Present 

Environment 
Information 
Center,  Inc. 

Monthly 

Records  contain 
abstracts. 

40 


Literature  Review 


Wildlife  information  data  bases  available  through  the  BLM  Library  (continued). 


Data  Base 

Description 

Coverage 
Dates 

Producer 

Update 
Frequency 

Comments 

Federal 
Research 
in  Progress 

Provides  access  to 
information  about 
ongoing  Federally- 
funded  research 
projects  in  physical 
sciences,  engineering, 
and  life  sciences. 
Project  descriptions 
include:  project  title, 
keywords,  start  date, 
estimated  completion 
date,  principal 
investigator,  performing 
and  sponsoring 
organizations, 
summary,  and 
progress  report. 

Current 
reserch 

National 
Technical 
Information 
Service 

Semiannual 
(reload) 

Abstracts  included  with 
the  records. 

Fish  and 
Wildlife 
Reference 
Service 

Includes  indexed 
documents  from  the 
Federal  Aid  in  Fish  and 
Wildlife  Restoration 
Program  (Pittman- 
Robertson  and  Dingell- 
Johnson  Acts),  the 
Anadromous  Fish 
Conservation  Program, 
the  Endangered 
Species  Grants 
program,  Cooperative 
Fishery  and  Wildlife 
Research  Units,  and 
State  fish  and  wildlife 
agencies.  Documents 
include  reports, 
published  papers, 
technical  publications, 
theses,  and  special 
materials,  such  as 
endangered  species 
recovery  plans. 

1950s-Present 

U.S.  Fish  and 
Wildl.  Serv., 
Division  of 
Federal  Aid 

Quarterly 

GPO  Monthly 
Catalog 

The  machine-readable 
equivalent  of  the 
printed  Monthly 
Catalog  of  United 
States  Government 
Publications.  Contains 
records  of  reports, 
studies,  fact  sheets, 
maps,  handbooks, 
conference 
proceedings,  etc., 
issued  by  all  U.S. 
Government  agencies, 
including  the  U.S. 
Congress. 

1976-Present 

U.S. 

Government 
Printing  Office 

Monthly 

Records  contain 
abstracts. 

Literature  Review 


41 


Wildlife  information  data  bases  available  through  the  BLM  Library  (concluded). 


Data  Base 

Description 

Coverage 
Dates 

Producer 

Update 
Frequency 

Comments 

Life  Sciences 
Collection 

Gives  worldwide 
coverage  of  journal 
articles,  books, 
conference 
proceedings,  and 
report  literature. 
Contains  information  in 
the  fields  of  animal 
behavior,  biochemistry, 
ecology,  entomology, 
genetics,  immunology, 
microbiology, 
toxicology,  and 
virology,  among  others. 

1978-Present 

Cambridge 

Scientific 

Abstracts 

Monthly 

Records  contain 
abstracts. 

SCISEARCH 

A  multidisciplinary 
index  to  the  literature 
of  science  and 
technology  that 
contains  all  the  records 
published  in  Science 
Citation  Index  and 
additional  records  from 
the  Current  Contents 
series  of  publications 
that  are  not  included  in 
the  printed  version  of 
Science  Citation  Index. 

1974-Present 

Institute  for 

Scientific 

Information 

Biweekly 

Records  are  citations 
only. 

Water 

Resources 

Abstracts 

Prepared  from 
materials  collected  by 
over  50  water  research 
centers  and  institutes 
in  the  United  States. 
Covers  a  wide  range  of 
water  resource  topics 
including  water 
resource  economics, 
ground  and  surface 
water  hydrology, 
metropolitan  water 
resources  planning 
and  management,  and 
water-related  aspects 
of  nuclear  radiation 
and  safety. 

1968-Present 

U.S.  Department 
of  the  Interior 

Monthly 

Citations  contain 
abstracts. 

Zoological 
Records 

Provides  extensive 
coverage  of  the  world's 
zoological  literature 
with  particular 
emphasis  on 
systematic/taxonomic 
information. 

1978-Present 

Biosciences 
Information 

Bimonthly 

Records  are  citations 
and  contain  systematic 
classification  of  up  to 
six  levels  for  the 
organism  cited. 

42 


Literature  Review 


, 


APPENDIX  II.  Publications  useful  for  wildlife  studies. 


Title 

Frequency 

Source 

Publications 

in  which  Indexed 

Comments 

American  Birds 

Bimonthly 

National  Audubon  Society 
950  Third  Avenue 
New  York,  NY  10022 

Biological  Abstracts 

Major  areas  of  interest  are 
the  changing  distribution, 
population,  migration,  and 
rare  occurrences  of  birds 
throughout  the  Americas. 
Formerly  issued  as 
Audubon  Magazine, 
Section  2;  Audubon  Field 
Notes. 

American  Fisheries 
Society  Transactions 

Bimonthly 

American  Fisheries 

Society 
5410  Grosvenor  Lane 
Suite  110 
Bethesda,  MD  20814 

Biological  Abstracts 
Chemical  Abstracts 
Current  Contents 
Excerpta  Medica 
Oceanic  Abstracts 
Pollution  Abstracts 

Objectives  of  the  Society 
are  conservation, 
development,  and  wise 
utilization  of  recreational 
and  commercial  fisheries, 
promotion  of  all  branches 
of  fisheries  science  and 
practice,  and  exchange 
and  dissemination  of 
knowledge  about  fish, 
fisheries,  and  related 
subjects. 

Auk 

Quarterly 

American  Ornithologists' 

Union 
%University  of  Illinois  at 

the  Medical  Center 
Department  of  Anatomy 
Box  6998 
Chicago,  IL  60680 

Biological  Abstracts 
Chemical  Abstracts 
Current  Contents 

Scholarly  articles  on 
various  topics  concerning 
birds  basic  to  the  study  of 
animal  or  bird  behavior. 
Covers  studies  throughout 
the  world,  but  articles  are 
all  in  English. 

Condor 

Quarterly 

Cooper  Ornithological 

Society,  Inc. 
Meriden  Rd. 
Lebanon,  NH  03766 

Biological  Abstracts 
Chemical  Abstracts 
Current  Contents 
Zoological  Record 

A  scholarly  journal 
devoted  to  the  biology 
and  behavior  of  birds  in 
the  wild. 

Copeia 

Quarterly 

American  Society  of 
Ichthyologists  and 
Herpetologists 

Department  of  Ichthyology 

American  Museum  of 
Natural  History 

New  York,  NY  10024 

Biological  Abstracts 
Chemical  Abstracts 
Current  Contents 
Oceanic  Abstracts 

Concerned  with  fishes, 
amphibians,  and  reptiles. 

Fisheries 

Bimonthly 

American  Fisheries 

Society 
5410  Grosvenor  Lane 
Bethesda,  MD  20814 

Current  Contents 

Articles  are  aimed  toward 
scientific  research  and 
enlightened  management 
of  aquatic  resources. 

Herpetologics 

Quarterly 

The  Herpetologists' 

League,  Inc. 
Department  of  Biology 
University  of  Southwestern 

Louisiana 
Lafayette,  LA  70504 

Biological  Abstracts 
Chemical  Abstracts 
Current  Contents 
Zoological  Record 

The  pumal  is  dedicated 
to  furthering  knowledge  of 
the  biology  of  amphibians 
and  reptiles. 

Journal  of 
Herpetology 

Quarterly 

Society  for  the  Study  of 
Amphibians  and 
Reptiles 

Ohio  University 

Dept.  of  Zoology 

Athens,  OH  45701 

Biological  Abstracts 
Current  Contents 
Zoological  Record 

Contains  original  research 
articles  and 

comprehensive  reviews  on 
amphibians  and  reptiles. 

Literature  Review 


43 


APPENDIX  II.  Publications  useful  for  wildlife  studies  (continued). 


Title 

Frequency 

Source 

Publications 

in  which  Indexed 

Comments 

Journal  of 
Mammology 

Quarterly 

American  Society  of 

Mammologists 
Vertebrate  Museum 
Shippensburg  State 

College 
Shippensburg,  PA  17257 

Biological  Abstracts 
Biological  and 

Agricultural  Index 
Current  Contents 
Index  Medicus 
Oceanic  Abstracts 
Pollution  Abstracts 
Selected  Water 

Resources  Abstracts 

Covers  the  entire  field  of 
mammalian  biology,  in- 
cluding marine  mammals. 
Papers  on  each  issue 
may  involve  taxonomy, 
ecology,  genetics,  behav- 
ior, physiology,  or  distri- 
bution of  species. 

Journal  of  Wildlife 
Management 

Quarterly 

The  Wildlife  Society 
5410  Grosvenor  Lane 
Bethesda,  MD  20814 

Biological  Abstracts 
Biological  and 

Agricultural  Index 
Chemical  Abstracts 
Science  Citation  Index 

Topics  ranging  from  habi- 
tat management  to  the 
effect  of  toxic  agents  on 
animals  and  plants  and 
other  wildlife  management 
issues. 

North  American 
Journal  of  Fisheries 
Management 

Quarterly 

American  Fisheries 

Society 
5410  Grosvenor  Lane 
Suite  110 
Bethesda,  MD  20814 

The  society  promotes  the 
development  of  all 
branches  of  fishery  sci- 
ence and  practice,  and 
the  conservation,  devel- 
opment, and  wise  utiliza- 
tion of  fisheries,  both 
recreational  and  commer- 
cial. Journal  reports  and 
papers  reflect  these  ob- 
jectives. 

North  American 
Wildlife  and  Natural 
Resources 
Conference 
Transactions 

Annually 

Wildlife  Management 

Institute 
709  Wire  Bldg. 
Washington,  DC  20005 

The  institute  promotes  the 
better  management  and 
wise  use  of  all  renewable 
natural  resources  in  the 
public  interest. 

Raptor  Research 

Quarterly 

Raptor  Research 
Foundation,  Inc. 

Dept.  of  Zoology 

161  WIDB 

Brigham  Young 
University 

Provo,  UT  84602 

Focuses  on  predatory 
birds.  Covers  all  aspects 
of  general  ecology,  natu- 
ral history,  management, 
and  conservation  of  diur- 
nal and  nocturnal  preda- 
tory birds. 

44 


Literature  Review 


APPENDIX  II.  Publications  useful  for  wildlife  studies  (concluded). 


Title 

Frequency 

Source 

Publications 

in  which  Indexed 

Comments 

Wildlife  Monographs 

Irregular 

The  Wildlife  Society 
5410  Grosvenor  Lane 
Bethesda,  MD  20814 

Current  Contents 

Each  Monograph  covers 
an  individual  subject 
relating  to  resource 
conservation  and  wildlife 
management. 

Wildlife  Review 

Quarterly 

Editorial  Office 

U.S.  Fish  and  Wildlife 

Service 
Aylesworth  Hall 
Colorado  State  University 
Fort  Collins,  CO  80523 

Alerts  wildlife  biologists  of 
current  worldwide 
literature  of  wildlife 
management  and 
conservation.  The 
individual  issues  of  the 
Wildlife  Review  are 
accumulated  into  Wildlife 
Abstracts. 

Wildlife  Society 
Bulletin 

Quarterly 

The  Wildlife  Society 
5410  Grosvenor  Lane 
Bethesda,  MD  20814 

Biological  Abstracts 

Covers  all  aspects  of 
management,  law 
enforcement,  education, 
economics,  administration, 
philosophy,  contemporary 
problems,  and  other 
topics  related  to  wildlife. 

Wilson  Bulletin 

Quarterly 

Wilson  Ornithological 

Society 
Department  of  Ornithology 
Royal  Ontario  Museum 
1000  Queen's  Park 
Toronto,  Ont.  M552C6 

Biological  Abstracts 
Current  Contents 
Science  Citation  Index 
Zoological  Record 

Articles  based  on 
research  or  field 
observations  throughout 
the  world. 

Literature  Review 


45 


APPENDIX  III.  Other  data  bases  useful  for  wildlife  research. 


Title 

Source 

Emphases 

BIO-STORET 

Environmental  Protection  Agency 

Zooplankton,  macroinvertebrates, 
some  vertebrates. 

Data  base  of  animal  and  plant 
species  of  concern  to  Office  of 
Endangered  Species 

Brookhaven  National  Laboratory 

Review  and  candidate  species  for 
endangered  list  and  those  listed  as 
proposed  threatened  or  endangered 

Information  system  for  regional 
energy-related  assessment  and 
planning  (GEOECOLOGY) 

Environmental  Sciences  Division  Oak 
Ridge  National  Laboratory 

Wildlife  data — includes  breeding  bird 
information,  endangered  and  mammal 
range  maps  and  distribution  data. 

Integrated  Habitat  Inventory  and 
Classification  System  (IHICS) 

Bureau  of  Land  Management 

Site-specific  wildlife  species 
information — amphibians,  birds, 
reptiles,  mammals. 

National  Data  Base  (RPA) 

Forest  Service 

Birds,  mammals,  reptiles,  amphibians, 
fish,  invertebrates 

Raptor  Management  Information 
System  (RMIS) 

Bureau  of  Land  Management 

Management  and  impact  information 
on  all  nocturnal  and  diurnal  raptors 
worldwide.  Provides  for  retrieval  of 
annotated  bibliographies  of  published 
and  unpublished  information 
concerning  raptors. 

Sensitive  Wildlife  Information  System 
(SWIS) 

U.S.  Army  Corps  of  Engineers 

Biology  and  distribution  of 
approximately  100  selected  mammal, 
bird,  reptile,  amphibian,  fish,  and 
invertebrate  species.  Emphasizes 
Federal  "endangered"  or  "threatened" 
species. 

Species  Data  Bases 

Migratory  Bird  and  Habitat  Research 
Laboratory  (MBHRL) 

Avian  Information 

Species  Data  Bases 

National  Coastal  Ecosystems  Team 

Aquatic  species — fish,  molluscs, 
aquatic  crustaceans 

Species  Data  Bases 

Office  of  Migratory  Bird  Management 
(MBMO) 

Avian  Information 

Species  Systems 

Northern  California  and  Pacific 
Northwest  Coastal  Characterization 
species  systems 

Vertebrate  species  and  selected 
invertebrates 

Terrestrial  Species  Database 

Western  Energy  Land  Use  Team 
(WELUT) 

Amphibians,  reptiles,  birds,  mammals 

Western  Sierra  Wildlife/Habitat 
Relationships  Program  (WHR) 

Forest  Service  Region  5 

Birds,  mammals,  reptiles,  amphibians 

WILD  RAM 

Forest. Service  Region  4 

Birds,  mammals,  reptiles,  amphibians, 
fish 

Wildlife-Habitat  Relationships  Data 
Base  (WILDHAB) 

Forest  Service  Region  6 

Birds,  mammals,  reptiles,  amphibians, 
fish 

Wildlife  Management  Information 
System  (WMIS) 

Forest  Service  Region  9 

Birds,  mammals,  reptiles,  amphibians, 
fish 

WILD01 

Forest  Service  Region  1 

Birds,  mammals,  reptiles,  amphibians, 
fish 

46 


Literature  Review 


APPENDIX  IV.  Museums  useful  for  wildlife  inventory  research. 


American  Museum  of  Natural  History 
79th  Street  and  Central  Park  West 
New  York,  NY  10024 
212-873-1300 

California  Academy  of  Science 

Golden  Gate  Park 

San  Francisco,  CA  941 18 

415-221-5100 

Carnegie  Museum  of  Natural  History 
4400  Forbes  Avenue 
Pittsburgh,  PA  15213 
412-622-3131 

Field  Museum  of  Natural  History 
Roosevelt  Road  at  Lake  Shore  Drive 
Chicago,  IL  60605 
312-922-9410 

Los  Angeles  County  Museum  of  Natural  History 
900  Exposition  Boulevard 
Los  Angeles,  CA  90007 
213-744-3414 


Museum  of  Vertebrate  Zoology 
University  of  California 
2593  Life  Sciences  Building 
Berkeley,  CA  94720 
415-642-3567 

Museum  of  Zoology 
University  of  Michigan 
1109  Washtenau 
Ann  Arbor,  MI  48109 
313-764-0476 

National  Museum  of  Natural  Sciences 
Victoria  Memorial  Building 
McLeod  and  Metcalfe  Streets 
Ottawa,  Ontario,  K1A  OMB 
CANADA 

U.S.  National  Museum  of  Natural  History 

Smithsonian  Institute 

10th  Street  and  Constitution  Ave.  NW 

Washington,  DC  20560 

202-357-2664 


Literature  Review 


47 


HABITAT 
MAPPING 

Richard  M.  Kerr 


Editor's  Note:  Mapping  and  classifying  land  areas 
into  distinct  habitat  types  is  a  necessary  part  of 
most  wildlife  activities.  This  is  particularly  true  of 
wildlife  inventories,  but  it  is  also  generally  neces- 
sary for  management,  monitoring  and  research. 
Although  the  principles  of  habitat  mapping  and 
classification  have  not  changed  since  Leopold 
(1933)  wrote  about  them,  many  new  tools  are  now 
available.  This  chapter  covers  the  principles  and 
provides  a  guide  to  the  new  techniques  available. 


U.S.  Bureau  of  Land  Management 
Las  Cruces  District  Office 
Las  Cruces,  NM  88004 


"Facts  concerning  game  distribution,  behavior,  his- 
tory, and  management  can  often  be  accumulated  on 
maps  or  tables  to  better  advantage  than  in  notes." 

— Aldo  Leopold,  Game  Management 


INTRODUCTION 

Aldo  Leopold  noted  that  facts  necessary  for 
wildlife  management  can  better  be  recorded  as  maps 
and  tables  rather  than  notes  because  notes  tend  to 
be  unorganized  and  difficult  to  analyze.  He  went 
further  to  describe  type  mapping  and  the  sorts  of 
wildlife  information  that  could  be  placed  on  base 
maps  (GLO  plats,  U.S.  Geological  Survey  quads,  or 
aerial  photography;  Leopold  1933). 


Mapping  is  done  for  several  purposes  or  uses  in 
wildlife  habitat  management: 

•  to  show  geographic  locations  of  wildlife  habitat 
types; 

•  to  show  relationships  of  types  to  other  types; 

•  to  show  community  ( types  of  habitat )  intersper- 
sion; 

•  to  quantify  types  of  wildlife  habitat; 

•  to  overlay  wildlife  habitat  types  with  other  re- 
source inventories;  and 

•  to  provide  geographic  locators  in  which  to  re- 
cord site-specific  animal  occurrence  and  use. 


Mapping  various  types  of  habitat  by  drawing 
boundaries  around  these  types  not  only  provides 
data  for  inventory  and  analysis,  but  also  provides  the 
habitat  manager  with  information  for  monitoring 
management  direction  or  problems.  With  habitat 
maps,  the  biologist  can  identify  and  locate  the  prob- 
lem, quantify  the  size  of  the  problem  and  provide, 
in  some  instances,  certain  types  of  baseline  data. 


Most  wildlife  data  bases  or  inventory  systems 
used  before  1970  described  the  life  history  of  a  spe- 
cies and,  as  part  of  that  discussion,  related  the  kinds 
of  habitat  that  the  species  used.  This  is  an  impracti- 
cal organization  of  data  for  a  land  manager  because 
as  many  as  400  vertebrate  species  may  occur  in  a 
U.S.  Bureau  of  Land  Management  (BLM)  District. 
Mapping  the  habitat  individually  for  each  one  of 


Habitat  Mapping 


49 


these  species  would  be  an  impossible  job.  Ranges 
would  have  to  be  delineated  by  observation  or  cap- 
ture, and  this  would  be  difficult  for  1  species  and 
impossible  for  400. 

A  system  that  delineates,  measures,  and  locates 
on  a  map  different  types  of  habitat  and  describes  the 
occurrence  of  animal  species  in  each  type  is  more 
useful  to  the  land  manager.  Such  a  system  can  allow 
sampling  of  habitats  and  extrapolation  of  data  to 
similar  habitats. 


In  addition,  land  managers  must  have  a  system 
that  allows  them  to  record  site-specific  information 
(e.g.,  200  elk  [Cervus  elaphus]  winter  in  this  area). 
Also,  special  features  in  the  landscape  like  cliffs, 
caves,  and  springs  need  to  be  recorded  in  the 
system. 


CLASSIFICATION  SYSTEMS  FOR  MAPPING 

Mapping  from  the  Top  Down:  Continental 
or  Regional  Classifications 

In  the  late  1890s,  Clinton  Hart  Merriam  pub- 
lished "Life  Zones  of  the  San  Francisco  Mountains." 
This  elevational  zone  classification  was  one  of  the 
first  ecosystem  classifications  to  be  used  by  ecolo- 
gists  in  North  America  (see  Figure  1)  and  is  still 
useful  in  ecological  analyses  (Brown  1982).  The 
early  ideas  of  Clements  (1928)  and,  later,  Kuchler 
(1964),  tended  to  locate  regional  ecosystems  de- 
scribed by  dominant  natural  potential  vegetation. 

The  recent  efforts  of  Brown  et  al.  (  1977),  al- 
though at  a  larger  scale  (1:1,000,000),  are  still  re- 
gional, geographic  vegetation  types  based  on  natural 
potential  vegetation  (Figure  2). 


N 


METERS 
3965 


Alpine   Tundra 
Timberline 


ARTIC- ALPINE 


HUDSONIAN 


Interior 
Chaparral 


3355 


2745 


2135 


-1525 


-915 


Figure  1.     Merriam's  Life  Zones,  although  helpful  as  a  general  characterization,  cannot  be  used  as  an  inven- 
tory because  of  individual  site  variations  within  elevational  zones  (Merriam  1890). 


50 


Habitat  Mapping 


/-       \ 


BIOTIC 
COMMUNITIES 

OF  THE 
SOUTHWEST 


TUNDRA    FORMATION 
H    ai  i-ini     TUNOnAS 
FOREST    FORMATION 
i^B   PtTRAN  '.ntiAi  ciNI    CONIIIH  f  Olll  ST 
■H    SIEHHAN  SUDAll'INt   CONIFER  FOREST 
9BK   PETHAN  MONTANf    CONIFER   FORES1 
HI    SlLHRAN  MONTANf  CONIFER  FORES1 
■H    S1NALOAN  DECIDUOUS  FOHEST 
WOODLAND   FORMATION 
ESS    GREAT  BASIN  CONNER  WOODLAND 
rSBS\    MAOntAN   EVERGREEN   WOODLAND 
■U    CALIFORNIAN   hVLHGRLEN  WOODLAND 

SCRUB   FORMATION 

■■  GREAT  UASIN  MONTANF   SCRIM 

131  t  CALIFORNIAN  CHAPARRAL 

IM::  CALIFORNIAN  COAS1ALSCRUB 

1JS53  INTERIOR  CHAPARRAL 

JSS2  SINALOAN  TMOHNSCRUD 

GRASSLAND   FORMATION 

■■     SUBALPINF   GRASSLAND 

■    PL  A 
PL  A 


^■BR     SFMIOFSIHI    (.HASSLAND 

DESERTSCRUB  FORMATION 
ISM-    GREAT  BASIN  UESEHTSCRU8 

111.1-  MOHAVF    DlblMTSCHUB 

■iU.tJ  CHtHUAHUAN  DLStRISCHUU 
SONONAN    VI  SIR  I SCHUU 

I5«-H  Uv«<>t  Colorado  fii*«i  Subdivision 

[TSfcJj,  A../on.i  Upland  SubdlviMon 

KHSS  PUtn*.  of  Sonora  Subdivision 

nSiJl  Ccot.jl  Goll  Coo-.!  Subdivision 

KH3H  Vi/catno  Subdivision 

(fSTfifi  Maodaleno  S«bd*vi%Mv> 


Figure  2.  A  regional  classification  map  called  Biotic  Communities  of  the  Southwest  (Brown  et  al.  1977), 
although  at  a  larger  scale  than  most  regional  maps,  is  still  too  small  for  field  management  information. 
The  State  of  Arizona  is  shown  as  an  example. 


Robert  Bailey's  "Ecoregions  of  the  United  States" 
(1976)  does  not  predominantly  use  vegetation  data 
as  most  do  but  uses  a  variety  of  biotic  and  abiotic 
factors  to  develop  a  single  geographic  classification 
at  a  scale  of  1:7,500,000.  This  again  could  be  consid- 
ered a  regional  ecosystem  map  or  classification 
(Figure  3). 


The  BLM  has  taken  AW.  Kuchler's  (1975)  map 
of  Potential  Natural  Vegetation  (Figure  4)  and  re- 
gionalized it  by  using  physiographic  region  bounda- 
ries based  on  soils,  vegetation,  and  regional 
topography.  These  regionalized  potential  vegetation 
types  are  used  as  ecosystem  boundaries  within 
which  to  aggregate,  analyze,  and  extrapolate  informa- 
tion or  data. 


The  trouble  with  most  of  these  delineations  and 
classifications  that  start  mapping  from  the  top  down 
(from  regional  or  continental  ecosystems)  is  that  the 
scale  is  extremely  small  for  use  by  field  biologists, 
and  mapping  is  based  on  potential  vegetation  and 
not  on  the  existing  vegetation.  Consequently,  accu- 
racy in  the  field  is  very  limited.  The  major  benefit 
then  of  these  preclassified  geographic  systems  is  to 
use  their  boundaries  for  aggregating  data  for  limiting 
extrapolation  of  information. 


Mapping  from  the  Ground  Up:  Integrated 
and  Component  Classifications 

The  classifications  of  Merriam  (1890),  Kuchler 
(1964),  Brown  et  al.  (1977),  and  Bailey  (1976) 


Habitat  Mapping 


51 


'*\7~~-7s'       'tif     /v    \'\ 

^^^xvV 

V\  ^of" 

/d\\  " 

44       dp= 

*A 

^9k 

r\  WARM 

1/                /r^ 

"VA  CONTINENTAL 

*0»    ^ 

3 

rO       c 

T'c 

vM              Z^— 

^C: 

«>  ^ 

\s 

/v^£] 

>j 

/    \^sX — "~-~-^* 

>  '       \  - 

-y^&t>v 

IA 

|/7_jAc/iC 

^      -o 
33 

s>7 

o  1/ 

> 
1       ^ 

><ff\ 

^^  O  V-         ■  1  /I    r" 

1      m     | 

l^"\                 ./        A 

-A  ^<C 

//   $\      •    P-^ 

y 

yk  ^  rn  S 

s^y 

< 

j\     y^ 

Highlands 

A/i\ 

S 

^*"^, 

^r^ 

\7_&M-  mountains 

.'.  jP  -  plutcau 

O^/^A/,  RAINFOREST 

i — — |  A  -  altiplano 

0 

400 

800  MILES                                     "^T^vJ 

L 

1 

Figure  3. 

terns. 


Ecoregions  of  the  United  States  are  a  classification  that  represents  a  hierarchy  of  regional  ecosys- 


might  be  considered  to  be  integrated  because  they 
are  geographically  specific,  and  their  boundaries 
encompass  all  habitat  components  within  the 
mapped  ecosystem.  Field-level  classifications  that  use 
all  habitat  components  for  delineating  mapping  units 
are  also  integrated  systems. 


Another  type  of  classification  system  has  been 
constructed  where  each  component  (vegetation, 
landform,  soil,  climate,  etc.)  is  mapped  individually 
and  independently  from  local  to  continental  levels, 
but  is  not  "integrated"  into  one  mapping  unit  to 
form  an  ecosystem.  This  is  called  a  hierarchical  com- 


52 


Habitat  Mapping 


I 


Figure  4.     AW.  Kuchler's  Potential  Natural  Vegetation  is  further  regionalized  by  using  physiographic  region 
boundaries. 


Habitat  Mapping 


53 


ponent  classification  system.  Hence  there  are  inte- 
grated and  component  classifications. 

Since  integrated  classifications  are  maps  or  de- 
pictions of  geographically  located  ecosystems,  they 
are  best  for  wildlife  habitat  mapping.  This  is  because 
animals  respond  to  all  components  within  an  area, 
not  just  one.  In  addition,  they  react  to  the  size  and 
location  of  these  small  ecosystems. 

Within  the  last  decade,  integrated  concepts  have 
been  so  mixed  with  hierarchical  component  con- 
cepts that  it  is  now  difficult  to  determine  a  partially 
integrated  system  from  the  hierarchical  component 
system.  A  good  example  of  this  is  "A  Component 
Land  Classification  for  the  United  States"  (Driscoll  et 
al.  1983).  This  system,  authored  by  five  government 
agencies,  is  meant  to  be  a  hierarchical  component 
classification  system;  yet  two  components,  soil  and 
vegetation,  are  integrated  through  habitat  types  and 
range  sites.  Other  component  hierarchies  are  not  yet 
developed.  What  is  called  a  component  system  now 
appears  to  be  an  integrated  system. 

In  the  field,  less  time  should  be  spent  on  re- 
gional classification  systems  and  more  time  on  local 
inventories,  as  far  as  wildlife  habitat  data  are  con- 
cerned. A  classification  can  be  built  from  a  good 
inventory,  but  an  inventory  cannot  be  built  from  a 
classification. 


MAPPING  TOOLS 

To  inventory  wildlife  habitat,  a  biologist  must 
construct  a  map  to  quantify  and  geographically  lo- 
cate habitat.  Describing  animal  communities  found 
within  plant  communities  by  composition  is  merely 
a  characterization  of  different  kinds  of  habitat.  An 
inventory  of  wildlife  habitat  has  not  been  completed 
until  a  reasonably  accurate  map  of  the  habitat  has 
been  constructed.  Similarly,  monitoring  studies  initi- 
ated must  have  a  map  to  be  of  continuing  value; 
anything  less  is  incomplete  or  unusable. 

The  map  may  be  an  overlay  of  a  base  map,  a 
computer  graphic  printout,  delineators  of  habitat  on 
a  U.S.  Geological  Survey  quadrangle  map,  a  typed 
aerial  photo  (see  Glossary  for  definitions),  a  Landsat 
image,  a  typed  planimetric  map,  a  plane  table  map, 
or  typed  orthophoto  maps. 

A  map  scale  for  inventorying  and  monitoring 
should  be  at  the  range  somewhere  between  1/60,000 
to  1/2,000,  depending  on  the  intensity  and  purposes 
of  the  study.  Maps  at  a  smaller  scale  than  1 :60,000 
may  provide  some  assistance  in  planning,  but  are  not 
large  enough  to  solve  actual  habitat  management 
problems. 


Some  maps  have  been  constructed  by  using  a 
scale  of  1/2  inch  to  equal  a  mile  or  1/100,000.  These 
were  commonly  used  in  BLM's  Unit  Resource  Analy- 
sis (URA)  system.  At  this  scale,  these  maps  can  only 
give  spatial  relationships  of  gross  kinds  of  habitat. 
They  are  not  suitable  for  recording  and  analyzing 
wildlife  habitat  data. 


Aerial  Photos 

The  most  commonly  used  mapping  tools  are 
aerial  photos  from  1/2,000  to  1/63,831,  and  7.5-  and 
1 5-minute  U.S.  Geological  Survey  quadrangles.  Ob- 
viously, the  larger  the  survey  area  in  acres,  the 
smaller  the  scale  will  be  to  meet  economic  criteria. 
The  easiest  and  best  photo-map  combination  for 
broad  inventories  over  large  acreages  of  land  8,000 
to  400,000  ha  (20,000  to  1  million  or  more  a.)  is 
the  U.S.  Geological  Survey  orthophoto  quadrangle 
(1/24,000)  (Figure  5)  and  the  7. 5-minute  topo- 
graphic quadrangle  (Figure  6).  The  orthophoto  quad 
is  a  corrected  copy  of  an  aerial  photo  at  1 :24,000 
scale,  which  allows  the  habitat  typing  to  be  placed 
directly  on  a  corrected  aerial  photo  which  is  the 
same  area,  scale,  and  control  as  the  base  map.  Habi- 
tat type  lines  can  be  traced  or  plotted  from  the  or- 
thophoto quadrangle  to  the  topographic  quadrangle 
of  the  same  area. 

Orthophoto  maps  (Figure  7)  are  corrected  aer- 
ial photos  with  topographic  map  information  super- 
imposed. They  thus  combine  all  features  of  both  7.5- 
minute  topographic  quadrangles  and  orthophoto 
quadrangles.  These  are  ideal  for  habitat  mapping  but 
are  only  available  for  a  few  areas. 

If  orthophoto  quadrangles  are  not  available  at 
1/24,000,  a  7. 5-minute  U.S.  Geological  Survey  quad- 
rangle map  is  used;  type  lines  are  transferred  from 
aerial  photos  with  a  plotter  or  zoom  transfer  scope. 

A  complete  description  of  the  principles  behind 
aerial  photography  can  be  found  in  Burr  (1976). 

The  kind  of  aerial  photos  used  may  vary  as  to 
the  situation  and  region.  The  first  habitat  inventories 
were  completed  with  black  and  white  aerial  photos 
(Figure  8),  and  these  are  still  very  useful  and  accept- 
able. A  recent  innovation  is  the  use  of  color-infrared 
film  which  is  particularly  useful  if  one  wants  to  em- 
phasize riparian  vegetation  or  seasonally  dominant 
vegetation.  This  film  produces  a  brighter  shade  of 
red  on  the  print  where  more  growth  is  taking  place 
(Figure  9). 

In  the  Southwest,  where  vegetation  is  sparse  and 
soil  highly  visible,  the  use  of  true  color  film  is  desira- 
ble. Color  often  gives  better  definition  between  habi- 
tat types  than  black  and  white  photos  give  (Figure 
10). 


54 


Habitat  Mapping 


Figure  5.     Orthophoto  quadrangles  like  these  are  excellent  corrected  photos  for  habitat  mapping. 


Habitat  Mapping 


55 


Figure  6.     U.S.  Geological  Survey  7.5-minute  topographic  quadrangles  are  the  mainstay  for  base  maps  for 
wildlife  habitat  inventory. 


56 


Habitat  Mapping 


Figure  7.     The  U.S.  Geological  Survey  7.5-minute  orthophoto  map  provides  a  corrected  aerial  photo  with 
topographic  and  other  map  features. 


Habitat  Mapping 


57 


Figure  8.     Black  and  white  aerial  photo  from  Piceance  Basin,  Colorado. 


58  Habitat  Mapping 


Figure  9.     Color  infrared  aerial  photo  of  same  area  as  Figure  8. 


Habitat  Mapping 


59 


Figure  10.     True  color  aerial  photo  of  same  area  as  Figure  8. 


60 


Habitat  Mapping 


The  photo  scale  recommended  for  overall  in- 
ventories on  public  lands  has  historically  been  close 
to  1/24,000  because  this  ties  in  well  with  the  scale 
of  U.S.  Geological  Survey  quads,  which  make  excel- 
lent base  maps  for  typing  vegetation.  This  medium- 
scale  photography  is  normally  available  from  various 
government  agencies  (Appendix  I). 

Larger-scale  photos  about  1/2,000  to  1/4,000  are 
beneficially  used  where  typing  is  necessary  on  highly 


valuable  habitats  such  as  riparian  areas  in  semiarid 
regions.  Color-infrared  aerial  photography  for  these 
areas  may  be  valuable  (Figure  11).  This  large  scale  is 
also  extremely  useful  for  habitat  monitoring  because 
individual  trees  and  shrubs  can  easily  be  seen  as 
well  as  stream  channel  changes.  This  large-scale  pho- 
tography is  normally  not  available  from  government 
agencies  and  must  be  obtained  by  contract.  In  some 
instances,  larger  scale  photography  has  been  used  for 
monitoring  vegetation  damage. 


Figure  11.     Large-scale  (1/2,000)  color  infrared  aerial  photo  of  riparian  areas  along  Black  Canyon  Creek, 
Arizona. 


Habitat  Mapping 


61 


A  complete  description  of  low-level  aerial  pho- 
tography for  riparian  inventory  and  monitoring  was 
given  by  Cuplin  (1981). 

In  addition  to  mapping,  aerial,  oblique,  and 
ground-based  photographs  are  indispensable  as  rec- 
ords for  baseline  data  and  for  composition,  density, 
and  structure  of  habitat.  A  properly  identified  photo- 
graph may  be  the  best  record  for  a  baseline  study 
of  a  variety  of  monitoring  studies.  The  use  of  a  cover 
or  profile  board  can  show  the  decreasing  density  of 
perennial  grass  on  a  duck-nesting  habitat  study, 
whereas  a  series  of  yearly  photos  from  a  photo  hub 
may  show  the  invasion  of  grass  into  a  shrub  type 
used  for  deer  winter  feeding.  "A  picture  is  worth  a 
thousand  words." 


Landsat  Images 

Landsat  images  are  sometimes  used  to  delineate 
gross  vegetation  cover  types.  Reflected  light  is  re- 
corded in  four  bands  and  can  be  used  to  generate 
Landsat  images  that  can  be  manually  interpreted  or 
the  digital  reflectance  data  can  be  computer-classi- 
fied. Pixels  of  about  0.04  ha  (1.1  a.)  picture  reflected 
light  and  are  density-sliced  and  artificially  classified 
into  false  colors  for  various  classifications.  Because  of 
the  characteristics  of  the  Landsat  reflections,  these 
scenes  do  not  provide  all  the  benefits  of  a  photo- 
graph at  medium  scales.  They  show  no  texture, 
shadow,  or  highlights,  but  just  the  color  of  a  particu- 
lar classification  (Figure  12).  Also,  the  discerning 
abilities  of  the  Landsat  scanners  are  not  capable  of 
separating  vegetation  species  or  areas  that  give  the 
same  reflectance.  Detailed  classifications  are  not  pos- 
sible with  Landsat  images. 

A  common  standard  used  to  describe  levels  of 
classification  detail  has  been  a  "Land  Use  and  Land 
Cover  Classification  System  for  Use  with  Remote 
Sensor  Data"  (Anderson  et  al.  1976).  (Note:  The  au- 
thor gives  description  titles  to  only  Levels  I  and  II, 
leaving  the  development  of  Levels  III  and  IV  to  local 
workers  [Table  1].) 

Information  given  in  the  latest  Manual  of  Re- 
mote Sensing  (Colwell  1983)  states  "Efforts  to  clas- 
sify Level  III  categories  using  Landsat  generally  result 
in  overall  accuracy  figures  of  less  than  70% ,  a  figure 
often  considered  unacceptable."  In  actuality,  vegeta- 
tion cover  classifications  for  land  management  use 
require  more  definition  than  Level  III  (Anderson  et 
al.  1976).  Level  IV  approaches  the  level  necessary 
for  management  use,  and  it  is  even  less  accurate  than 
Level  III,  perhaps  50  percent  or  less.  If  Level  II  was 
described  as  shrubland,  then  the  Level  III  classifica- 
tion descriptor  might  be  rabbit  brush  (Chtysotham- 
nus  sp.)  while  the  Level  IV  descriptor  might  be 
rabbit  brush-snakeweed  {Gutierrezia  sp.). 


Unlike  more  easily  interpreted  photographic 
scenes,  Landsat  scenes  are  highly  dependent  on 
ground  truthing.  The  more  ground  truthing,  the 
more  accurate  is  the  final  classification  scene.  The 
problem  of  signature  extension  has  not  been  solved 
for  all  areas  and  vegetation  types,  hence  a  poorly 
ground-truthed  scene  of  vegetation  may  be  highly 
inaccurate. 


Vegetation  signatures  (reflection  shades)  many 
times  do  not  remain  accurate  even  over  reasonable 
distances  (3  to  5  mi.);  therefore,  for  accurate  classifi- 
cations, signatures  must  be  classified  for  small  areas 
by  using  a  great  deal  of  ground  truthing  and  super- 
vised classifications  (Short  1982).  Because  of  this,  it 
may  be  more  economical  to  use  aerial  photos  to 
lend  more  accuracy  to  final  products.  Also,  because 
definitions  on  photos  exceed  those  of  pixel  classifica- 
tions, ground  control  (section  corners,  etc.)  is  better 
accomplished  with  photos  than  Landsat  images. 


Although  Landsat  images  may  be  an  efficient 
advance  in  providing  a  mapping  base  where  there 
are  no  existing  base  maps  (i.e.,  Alaska  or  Africa), 
most  areas  in  the  western  U.S.  are  already  accurately 
mapped  with  7.5-minute  U.S.  Geological  Survey  topo- 
graphic quadrangles.  These  maps,  constructed 
through  photo  interpretation,  in  many  instances  have 
some  vegetation  mapped  at  a  level  comparable  to 
the  Level  III  mapping  used  by  contemporary  Landsat 
projects.  This  mapping  is  much  more  accurate  and 
at  a  larger  scale  than  the  Landsat  images  available 
either  by  standard  order  or  by  computer  construc- 
tion and  enhancement.  Computer-enhanced  Landsat 
images  therefore  must  be  used  with  discretion  be- 
cause they  do  not  provide  mapping  at  the  accuracy 
and  level  (or  scale)  for  most  BLM  management 
needs.  However,  they  may  supplement  or  provide  an 
additional  tool  for  enhancing  management  data  such 
as  information  on  snow  cover,  cloud  cover,  soils,  and 
repetitive  synoptic  views.  If  aerial  photos  were  to 
be  totally  replaced  by  Landsat  scenes  as  a  basic  habi- 
tat inventory  tool,  the  land  management  profession 
would  have  taken  a  giant  step  backward. 


Sources  of  Supply 

Aerial  photos  and  Landsat  images  may  be  ob- 
tained from  affiliates  of  the  National  Cartographic  In- 
formation Center  (Appendix  I).  Assistance  for  BLM 
employees  may  be  obtained  from  the  BLM  Service 
Center.  For  U.S.  Geological  Survey  maps  of  all  kinds, 
including  orthophoto  quads  for  areas  west  of  the 
Mississippi,  write  to  the  U.S.  Geological  Survey,  Box 
25286,  Denver  Federal  Center,  Denver,  CO  80225. 


62 


Habitat  Mapping 


HIGH  DENSITY  SAGEBRUSH 

LOW  DENSITY  SAGEBRUSH 

SALTBUSH  /  GREASEWOOD  /  SAGE 

SALTBUSH 

ASPEN  /  OPEN  ASPEN 

CONIFER 

RIPARIAN  /  AGRICULTURE 

BARREN 

WATER 


Figure  12.     An  unsupervised  Landsat  classification. 


Habitat  Mapping 


63 


Table  1.     Land  use  and  land  cover  classification  system  for  use  with  remote  sensor  data  (modified  from 
Anderson  et  al.  1976). 


Level  I1 

Level  II2 

Urban  or  built-up  land 

11.  Residential3 

12.  Commercial  and  services 

13.  Industrial 

14.  Transportation,  communications,  and  utilities 

15.  Industrial  and  commercial  complexes 

16.  Mixed  urban  or  built-up  land 

17.  Other  urban  or  built-up  land 

Agricultural  land 

21.  Cropland  and  pasture 

22.  Orchards,  groves,  vineyards,  nurseries,  and  orna- 
mental horticultural  areas 

23.  Confined  feeding  operations 

24.  Other  agricultural  land 

Rangeland 

31.  Herbaceous  rangeland 

32.  Shrub  and  brush  rangeland 

33.  Mixed  rangeland 

Forest  land 

41 .  Deciduous  forest  land 

42.  Evergreen  forest  land 

43.  Mixed  forest  land 

Water 

51.  Streams  and  canals 

52.  Lakes 

53.  Reservoirs 

54.  Bays  and  estuaries 

Wetland 

61 .  Forested  wetland 

62.  Nonforested  wetland 

Barren  land 

71.  Dry  salt  flats 

72.  Beaches 

73.  Sandy  areas  other  than  beaches 

74.  Bare  exposed  rock 

75.  Strip  mines,  quarries,  and  gravel  pits 

76.  Transitional  areas 

77.  Mixed  barren  land 

Tundra 

81.  Shrub  and  brush  tundra 

82.  Herbaceous  tundra 

83.  Bare  ground  tundra 

84.  Wet  tundra 

85.  Mixed  tundra 

Perennial  snow  or  ice 

91.  Perennial  snowfields 

92.  Glaciers 

'Level  1  Typical  data  characteristics  at  this  level  would  be  found  on  Landsat  types  of  data. 

2Level  2  Typical  data  characteristics  at  this  level  would  be  found  on  high-altitude  data  at  40,000  ft.  (12,400  m)  or  above  (less  than 
1  80,000  scale) 
3Numbers  and  titles  refer  to  standard  land  use  codes  and  definitions  from  U.S.  Urban  Renewal  Administration  et  al.  1965. 


THE  INTEGRATED  MAPPING  PROCESS 

The  traditional  integrated  mapping  approach  fits 
well  with  the  concept  of  habitat  inventory,  which 
has  five  basic  steps: 

(1)  mapping; 

(2)  stratification  and  classification; 

(3)  inventory  of  habitat-species  relationships; 


(4)  site-specific  inventory;  and 

(5)  data  storage,  retrieval,  and  analysis. 


Mapping 

To  provide  for  the  aggregation  of  wildlife  habi- 
tat data  at  local  levels  and  up  into  higher  levels  of 
field  classification,  a  mapping  unit  is  used  to  collect 


64 


Habitat  Mapping 


; 


and  quantify  data.  This  mapping  unit  is  delineated  on 
aerial  photographs  as  a  homogeneous  polygon — 
homogeneous  as  to  existing  vegetation,  landform, 
and  soil.  Although  size  of  the  mapping  unit  has  no 
required  range,  it  varies  between  0.8  ha  and  perhaps 
4,000  ha  (2a.  and  10,000a.).  There  may  be  some 
misunderstanding  of  what  the  mapping  unit  is  called. 
In  BLM's  Integrated  Habitat  Inventory  and  Classifica- 
tion System  (IHICS),  it  is  called  a  habitat  site  (Hamil- 
ton et  al.  1983).  In  other  systems,  it  is  called  a 
mapping  unit,  whereas  in  the  BLM's  vegetation  in- 
ventory, it  is  called  a  site  writeup  area. 

Mapping  units  are  first  pretyped  on  aerial  photos 
(1:24,000)  or  orthophoto  quads.  This  is  done  by 
examining  the  photos  and  drawing  boundary  lines 
around  areas  homogeneous  in  terms  of  existing  vege- 
tation, landform,  and  soils  (if  soil  inventory  overlays 
or  ecological  site  delineations  are  available).  To  re- 
duce displacement  caused  by  relief,  the  effective 
area  or  central  portion  of  a  photo  pair  should  be 
used  for  interpretation. 

Once  all  areas  of  the  photos  are  pretyped,  the 
boundaries  or  delineations  that  were  put  on  in  the 
office  are  checked  in  the  field  by  visual  observation. 
It  is  also  desirable  to  use  some  management  bounda- 
ries to  break  the  mapping  unit  by  forming  one  of 
its  boundaries  (e.g.,  a  grazing  allotment  boundary 
should  not  be  crossed  when  delineating  mapping 
units  because  many  times  data  need  to  be  accumu- 
lated and  analyzed  according  to  grazing  allotment). 
Similarly,  Resource  Area  boundaries  should  not  be 
crossed  when  delineating  mapping  units.  During 
a  field  check,  section  corners  are  located  and  pin- 
pricked  through  the  photograph.  Identification  is 
placed  on  the  back  of  the  photo  for  control.  If  ortho- 
photo  quads  are  used,  this  is  not  necessary;  each 
mapping  unit  is  assigned  a  unique  designation  within 
the  inventory  area. 

Certain  habitat  data  are  recorded  for  each  map- 
ping unit  while  checking  boundaries  in  the  field. 
Landform,  site  elevation,  dominant  and  subdominant 
vegetation  (by  visual  observation  or  vegetation  data), 
aspect  (by  cardinal  direction),  and  slope  data  are 
recorded  and  entered  on  the  appropriate  form. 
Acreage  should  be  computed  and  recorded  as  soon 
as  the  unit  boundary  is  mapped  and  becomes  final. 

Once  mapping  units  have  been  checked  and  are 
considered  final,  they  are  transferred  to  a  1:24,000 
U.S.  Geological  Survey  quadrangle  map  if  they  are 
not  already  on  an  orthophoto  map.  This  is  done  by  a 
zoom  transfer  scope  or  other  type  of  plotter  and  by 
using  the  located  section  corners  for  control. 

Where  ecological  sites  (potential  vegetation) 
have  been  delineated  on  the  ground  and  units  of 
existing  vegetation  have  been  delineated  on  maps, 
the  same  units  and  numbers  can  be  used  for  habitat 


inventory  data  base.  If  this  option  is  available,  the 
work  described  in  the  preceding  paragraphs  is  not 
necessary  except  when  collecting  the  described 
habitat  data. 


Stratification  and  Classification 

After  all  mapping  units  have  been  given  an  iden- 
tification number  and  initial  habitat  information  has 
been  gathered  and  recorded,  the  mapping  units  are 
placed  in  strata  which  will  ultimately  be  a  field  level 
classification.  To  determine  the  number  and  types 
of  strata,  the  biologist  in  charge  will  need  to  make  a 
field  review,  noting  which  types  of  habitat  should 
be  designated  as  habitat  strata.  The  partially  com- 
pleted habitat  information  forms  can  be  used  for 
stratifying  by  reviewing  the  habitat  site  information 
on  them,  particularly  the  dominant  and  subdominant 
vegetation,  landform,  or  other  factors  based  on  the 
needs  of  the  inventory  situation.  Once  criteria  for 
grouping  strata  are  determined,  classifications  are 
made  by  assigning  each  mapping  unit  to  the  stratum 
that  applies. 

The  BLM  IHICS  can  take  input  forms  showing 
habitat  data  for  mapping  units  and  automatically  sort 
strata  by  selected  criteria.  For  example,  if  all  map- 
ping units  located  on  a  mesa  where  ponderosa  pine 
{Pinus  ponderosa)  is  dominant  and  Columbia  need- 
legrass  (Stipa  columbiana)  is  subdominant  are  se- 
lected as  a  stratum,  the  computer  will  easily  sort  and 
form  the  stratum.  Each  classification  (stratum)  is 
assigned  a  number  and  name.  Each  mapping  unit 
then  has  its  classification  number  recorded  on  the 
appropriate  form.  New  strata  may  be  devised  at  any 
time  by  using  different  criteria  and  reassembling 
the  mapping  units. 

Although  existing  vegetation  and  landform  are 
important,  criteria  for  forming  strata  may  vary.  If 
a  particularly  important  animal  is  present,  a  stratum 
of  habitat  may  be  made  for  it  by  using  an  important 
characteristic  of  its  habitat.  For  instance,  if  Gila 
monsters  (Heloderma  suspectum )  are  found  in  a 
creosote-bursage  habitat,  but  only  in  areas  that  con- 
tain abundant  ground  litter,  and  if  abundant  litter 
is  used  as  a  criterion  within  this  cover  type,  a  stra- 
tum is  constructed  that  will  represent  potential  Gila 
monster  habitat. 

Sufficient  habitat  data  (described  below)  are 
collected  for  each  map  unit.  Strata  can  be  con- 
structed by  using  habitat  data  in  manual  or  auto- 
matic data  processing  (ADP)  operations  (i.e.,  sort  all 
map  units  that  are  ponderosa  pine  and  pine  needle- 
grass  [Stipa pinetorum]  at  7,000  to  9,000  ft  [2,121 
to  2,727  m]  elevation).  Consequently,  biologists  have 
the  ability  to  construct  strata  for  sampling  based  on 
a  variety  of  criteria  once  map  units  are  delineated 
and  initial  habitat  information  is  recorded. 


Habitat  Mapping 


65 


The  mapping  unit  is  the  lowest  level  of  ecosys- 
tem classification  systems.  The  stratum  level  is  the 
aggregation  of  mapping  units  having  similar  vegeta- 
tion, landform,  soil,  etc.  The  two  lower  levels  (map 
unit,  stratum)  provide  a  means  of  categorizing  field 
inventory  data  to  facilitate  data  analysis  for  land- 
use  planning  and  decisionmaking.  They  provide  a 
long-term  accumulation  and  aggregation  of  new  data 
from  the  mapping  unit  level  to  upper  levels  of  conti- 
nental or  regional  systems. 


Inventory  of  Habitat-Species  Relationships 

Once  delineation  and  classification  of  mapping 
units  have  been  made,  the  next  step  is  to  collect 
information  that  will  reveal  species-habitat  relation- 
ships. For  each  stratum,  representative  mapping  units 
are  selected  for  sampling  vegetation  and  wildlife. 

Vegetation  composition  and  structural  data 
should  be  sampled  in  the  same  habitat  sites  where 
the  animal  sampling  occurs.  If  these  data  are  avail- 
able from  other  vegetation  inventories,  then  sam- 
pling is  not  needed. 

A  hypothetical  list  prepared  through  literature 
review  by  using  upper-level  classifications  (Kuchler 
1964;  Brown  et  al.  1977)  can  be  used  as  a  basis  for 
sampling  animal  occurrence.  A  sampling  plan  is  de- 
veloped that  verifies  the  occurrence  of  animal  spe- 
cies listed  hypothetically.  If  a  complete  wildlife 
inventory  is  conducted,  each  animal  class  is  invento- 
ried for  one  or  more  seasons,  depending  on  its  life 
history.  Generally,  inventories  are  conducted  for 
mammals,  birds,  reptiles,  and  amphibians.  When  time 
and  money  are  limited,  only  those  classes  deter- 
mined important  in  the  management  situation  are 
inventoried.  Species  thought  to  occur  in  the  selected 
habitat  sites  are  listed  in  the  data  base  as  hypotheti- 
cal. Unique  or  important  animals  expected  to  occur 
in  the  inventory  area  should  be  verified  by  using 
appropriate  inventory  methods. 

Sampling  of  animals  in  a  stratum  will  relate  oc- 
currence of  animals  to  certain  canopies  or  other 
habitat  structure,  or  other  special  features  within  the 
habitat.  One  way  to  do  this  is  to  use  inferential  sta- 
tistics and  compare  occurrence  of  certain  animals 
with  occurrence  of  a  particular  layer  of  habitat  or 
special  feature.  This  is  accomplished  best  if  the  biol- 
ogist completes  the  same  vegetation  (habitat)  inven- 
tory at  each  animal  sampling  location.  After  sufficient 
numbers  of  samples  have  been  taken,  a  statistical 
comparison  can  be  made.  Vegetation  or  habitat  sam- 
pling methods  may  be  developed  by  region  or  area 
of  inventory,  considering  those  things  in  the  habitat 
that  appear  to  be  important  in  that  particular  region. 

A  simple  list  of  animals  occurring  in  an  area  is 
helpful,  but  the  land  manager  needs  additional  infor- 


mation. This  includes  season  of  use;  function  of  the 
habitat  for  a  particular  species;  determination  of 
whether  the  habitat  is  crucial  for  the  species;  and 
the  relative  abundance  of  the  species,  compared 
with  other  habitats. 

If  stratification  and  sampling  have  revealed  that 
two  or  more  strata  have  the  same  animal  community 
and  uses,  then  these  may  be  combined  into  a  single 
habitat  stratum. 

If  the  purpose  of  mapping  is  to  monitor  a  wild- 
life habitat  problem,  then  the  same  procedures  of 
mapping  apply  except  that  vegetation  sampling  is 
done  by  a  system  that  monitors  those  aspects  of  the 
habitat  that  are  being  managed  or  affected. 

Because  the  BLM  IHICS  uses  inventories  of  ex- 
isting vegetation,  the  potential  vegetation  situation  is 
normally  portrayed  by  using  habitat  types  (plant 
associations)  for  forests  and  range  or  ecological  sites 
for  rangelands.  Comparing  the  existing  vegetation 
with  this  potential  vegetation  will  give  possibilities 
for  vegetation  management,  which  would  direct 
animal  use  or  animal  community  progression  toward 
a  desired  level.  This  potential  determination  may 
be  available  for  broad  areas  or  can  be  determined 
individually  for  small  management  areas  (such  as 
deer  feeding  areas). 

Much  money  and  time  can  be  spent  acquiring 
potential  information  on  the  total  area  of  rangelands 
within  a  BLM  administrative  unit  jurisdictional 
boundary,  when  in  practice,  only  a  small  amount  of 
this  information  will  be  used  in  habitat  management 
decisions.  Biologists,  therefore,  cannot  afford  to 
spend  their  time  delineating  potential  for  all  areas. 
Biologists  should  restrict  their  efforts  toward  deter- 
mining vegetation  potential  only  in  areas  where  in- 
formation is  not  available  from  the  U.S.  Soil 
Conservation  Service  (SCS)  range  site  system  or  the 
forest  habitat  type  system,  and  the  information  is 
necessary  for  specific  immediate  wildlife  habitat 
management  analysis. 

It  is  important  to  understand  that  vegetation 
progression  under  commercial  forest  management 
changes  structure  rapidly,  whereas  composition  and 
structural  changes  under  range  management  situa- 
tions take  long  periods  and  the  effects  are  much 
more  subtle  and  generally  less  predictable. 


Site-Specific  Inventories 

Mapping  units  provide  the  mechanism  for  re- 
cording site-specific  data  because  each  has  a  unique 
number  which  locates  it  geographically.  For  those 
species  important  to  the  management  of  public  lands 
and  for  which  information  cannot  be  safely  extrapo- 
lated, site-specific  data  must  be  collected.  These 


66 


Habitat  Mapping 


species  are  generally  highly  mobile  species  (such  as 
raptors  and  wild  ungulates)  whose  presence  is  not 
well  correlated  with  just  a  habitat  or  vegetation  type 
(e.g.,  because  all  ponderosa  pine  stands  do  not  con- 
tain deer,  we  must  find  ones  that  do  and  record  by  a 
mapping  unit  number). 

The  form  for  recording  information  for  mapping 
units  and  site-specific  data  should  also  allow  for  con- 
tinuing entry  of  animal  occurrence  data  as  they  are 
gathered,  either  in  complete  inventories  or  oppor- 
tune sightings. 

Sometimes  features  of  the  habitat  that  are  too 
small  to  map  (springs,  walls,  power  poles,  cliffs,  etc.) 
have  a  major  effect  on  whether  certain  species  of 
wildlife  occur  in  an  area.  The  effect  may  be  either 
negative  or  positive.  These  are  called  special  habitat 
features,  and  a  complete  habitat  inventory  should 
record  them.  Time  and  money  will  not  allow  a  total 
inventory  of  all  special  habitat  features.  The  biolo- 
gists must  choose  those  features  that  are  important 
to  wildlife  in  their  inventory  areas  and  inventory 
them.  (For  instance,  all  special  habitat  features  that 
are  water-related  should  be  inventoried.)  Species 
that  are  benefitted  or  discouraged  by  the  special  hab- 
itat feature  are  listed.  This  gives  the  biologists  the 
second  form  of  site-specific  data  for  their  data  bases. 


Data  Analysis,  Storage,  and  Retrieval 

In  the  next  stage,  field  forms  for  animal  use, 
occurrence,  relative  abundance,  and  crucial  habitat 
are  assembled  and,  depending  on  the  system,  sum- 
marized on  additional  forms.  At  this  stage,  the  biolo- 
gist may  need  to  modify  the  original  strata 
designations  based  on  the  additional  information 
collected. 

At  this  stage,  data  is  normally  entered  into  some 
sort  of  computer  system.  In  the  BLM  IHICS,  the  sum- 
mary forms  are  entered  into  a  computer  data  base. 
Regardless  of  the  system  used,  forms  should  be 
checked  carefully  prior  to  key-entry. 

In  the  BLM's  IHICS,  summary  data  are  edited 
automatically  and  then  used  to  create  data  bases  that 
are  available  in  the  field  on  either  a  time-sharing 
computer  system  or  on  a  field  office  microcomputer. 
Separate  data  bases  are  maintained  for  habitat  sites 
(mapping  units),  standard  habitat  sites  (mapping 
strata),  and  special  habitat  features  for  each  inven- 
tory area. 


The  use  of  ADP  has  generally  proceeded  in  two 
reasonable  directions  as  far  as  wildlife  habitat  inven- 
tory and  monitoring  are  concerned.  Some  agencies 
have  emphasized  digitized  graphic  displays  (maps)  of 
types  of  habitat.  This  was  done  by  using  systems 
such  as  the  Map  Overlay  and  Statistical  System 
(MOSS)  to  construct  polygons  (habitat  mapping 
units  from  digitized  data  derived  from  aerial  photos). 
For  large  study  areas,  400,000  ha  ( 1  million  a. )  or 
more  at  1/24,000  accuracy,  digitizing  needed  for 
constructing  computer-generated  maps  is  expensive 
and  time-consuming,  but  its  use  is  increasing  at  the 
field  level. 

Other  agencies  emphasized  the  measurement  of 
biological  components  within  strata  and  mapping 
units.  These  data  are  normally  input  as  tables  and 
other  narratives  whereas  the  maps  that  have  been 
used  with  them  have  been  manually  constructed. 
The  BLM  IHICS  is  a  good  example  of  a  system  that 
sorts  several  dozen  habitat  characteristics. 

Both  types  of  systems  are  valuable  where  large 
amounts  of  data  are  handled,  and  the  automation 
entices  field  offices  to  use  the  data  repeatedly  for  a 
variety  of  uses,  thereby  obtaining  greater  benefit. 

The  present  move  toward  balanced  emphasis  on 
computer  mapping  and  definitive  biological  data 
will  make  the  data  even  more  usable  and  valuable. 
This  type  of  balanced  emphasis  could  provide  a  very 
usable  Geographic  Information  System  (GIS).  Pro- 
cesses such  as  the  digital  elevation  model  (DEM)  can 
be  included  or  used  in  conjunction  with  GIS  thereby 
enhancing  the  utility  of  both  data  bases. 


DISCUSSION 

The  overall  purpose  of  mapping  is  to  organize 
wildlife  habitat  data  to  manage  wildlife  habitat  and  to 
mitigate  or  eliminate  impacts  of  other  uses.  The  hab- 
itat map  is  the  basic  tool  for  the  land-management 
biologist.  Geographic  locations,  quantification  of 
habitat,  and  specific  habitat  characterization  cannot 
be  accomplished  without  a  map.  In  addition,  maps 
are  necessary  to  compare  wildlife  habitat  with  other 
resources  for  multiple-use  decisions.  Even  though 
contemporary  map-making  methods  may  change 
from  a  manually-constructed  paper  map  to  a  com- 
puter-generated graphic,  a  map  is  only  as  accurate  as 
its  human  constructor. 


Habitat  Mapping 


67 


LITERATURE  CITED 


ANDERSON,  J.A.,  E.E.  HARDY,  J.T.  ROACH,  and  RE.  WIT- 
MER.  1976.  A  land  use  and  land  cover  classification 
system  for  use  with  remote  sensor  data.  U.S.  Dep. 
Inter.,  Geological  Survey,  Professional  Paper  964. 
28pp. 

BAILEY,  RG.  1976.  Ecoregions  of  the  United  States.  U.S. 
Dep.  Agric,  Forest  Service.  Ogden,  UT  84401.  77pp. 

BROWN,  D.E.,  ed.  1982.  Desert  plants,  biotic  communities 
of  the  American  Southwest,  United  States  and  Mexico. 
Vol.  4,  Numbers  1-4,  Special  issue.  University  of  Ari- 
zona, Tucson.  342pp. 

,  C.H.  LOWE,  and  C.P.  PASE.  1977.  Biotic  communi- 
ties of  the  Southwest.  U.S.  Dep.  Agric,  Gen.  Tech. 
Rep.  RM-41.  342pp. 

BURR,  R.D.  1976.  The  use  of  aerial  photographs.  Technical 
Note  287.  U.S.  Dep.  Inter.,  Bur.  Land  Manage.  BLM 
Service  Center,  Denver,  CO  80225.  34pp. 

CLEMENTS,  F.E.  1928.  Plant  succession  and  indicators. 
The  H.W.  Wilson  Co.  New  York,  NY.  453pp. 

COLWELL,  R.N.,  ed.  1983-  Manual  of  remote  sensing.  2nd 
Ed.,  Vol.  II.  American  Society  of  Photogrammetry. 
Falls  Church,  VA.  2440pp. 

CUPLLN,  P.  1 98 1 .  The  use  of  large-scale  color  infrared 
photography  for  stream  habitat  and  riparian  vegeta- 
tion inventory.  Technical  Note  325,  U.S.  Dep.  Inter., 
Bur.  Land  Manage.  BLM  Service  Center,  Denver,  CO 
80225.  7pp. 

DRISCOLL,  R.S.,  D.L.  MERKLE,  J.S.  HAGIHARA,  and  D.L. 
RADLOFF.  1983  A  component  land  classification  for 
the  United  States:  Status  report.  Technical  Note  360. 


U.S.  Dep.  Inter.,  Bur.  Land  Manage.  BLM  Service  Cen- 
ter, Denver,  CO  80225.  30pp. 

HAMILTON,  C.K.,  R.M.  KERR,  and  LA.  PETERSON.  1983. 
IHICS,  the  Bureau  of  Land  Management's  habitat  in- 
ventory system.  Presented  at  National  Workshop 
on  Computer  Uses  in  Fish  and  Wildlife  Programs — A 
State  of  the  Art  Review.  Virginia  Polytechnic  Institute 
and  State  University,  Blacksburg,  VA.  1 3pp. 

KUCHLER,  AW.  1964.  Potential  natural  vegetation  of  the 
conterminous  United  States,  map  and  manual.  Special 
Publication  36,  Am.  Geographic  Soc.  New  York,  NY. 
122pp. 

.  1975.  Potential  natural  vegetation  of  the  contermi- 
nous United  States,  revised  map.  Am.  Geographic 
Soc.  New  York,  NY. 

LEOPOLD,  A.  1947.  Game  management.  Charles  Scribner's 
Sons.  New  York,  NY.  481pp. 

MERRIAM,  C.H.  1890.  Results  of  a  biological  survey  of  the 
San  Francisco  Mountain  region  and  desert  of  the 
Little  Colorado  in  Arizona.  North  Am.  Fauna  No.  3- 
U.S.  Dep.  Agric,  Washington,  DC  (Reprinted  in  se- 
lected works  of  Clinton  Hart  Merriam.  Natural  Sci- 
ences in  America,  K.B.  Sterling,  ed.)  Arno  Press.  New 
York,  NY  1974. 

SHORT,  N.M.  1982.  The  Landsat  tutorial  workbook,  basics 
of  satellite  remote  sensing.  NASA  Reference  Publica- 
tion 1078.  National  Aeronautics  and  Space  Adminis- 
tration. Washington,  DC. 

U.S.  URBAN  RENEWAL  ADMINISTRATION,  HOUSING 
AND  HOME  FINANCE  AGENCY,  and  BUREAU  OF 
PUBLIC  ROADS.  1965.  Standard  land  use  coding  man- 
ual, a  standard  system  for  identifying  and  coding 
land  use  activities.  Washington,  DC.  1 1 1pp. 


68 


Habitat  Mapping 


APPENDIX  I.     Sources  of  Supply 


National  Affiliate  Offices 


State  Affiliate  Offices  (cont.) 


National  Cartographic  Information  Center 
U.S.  Geological  Survey 
507  National  Center 
Reston,  VA  22092. 


3.  Montana 

Montana  Bureau  of  Mines  and  Geology 

Montana  Tech 

Main  Hall,  Room  200 

Butte,  MT  59701 


Eastern  Mapping  Center 

National  Cartographic  Information  Center 

U.S.  Geological  Survey 

536  National  Center 

Reston,  VA  22092. 


Nebraska 

Director  and  State  Geologist 
Conservation  and  Survey  Division 
University  of  Nebraska 
Lincoln,  NE  68508 


Mid-Continent  Mapping  Center 

National  Cartographic  Information  Center 

U.S.  Geological  Survey 

1400  Independence  Road 

RoUa,  MO  65401 


5.  New  Mexico 

University  of  New  Mexico 
Technology  Applications  Center 
2500  Central  Avenue,  S.E. 
Albuquerque,  NM  87131 


Rocky  Mountain  Mapping  Center 
National  Cartographic  Information  Center 
U.S.  Geological  Survey 
Stop  504,  Box  25046,  Federal  Center 
Denver,  CO  80225. 


6.  Nevada 

Nevada  Bureau  of  Mines  and  Geology 
University  of  Nevada,  Reno 
Reno,  NV  89557-0088 


Western  Mapping  Center 

National  Cartographic  Information  Center 

U.S.  Geological  Survey 

345  Middlefield  Road 

Menlo  Park,  CA  94025. 


Oregon 

Oregon  State  Library 
Public  Services 
Salem,  OR  97310 


Alaska  Office  National  Cartographic  Information  Center 
U.S.  Geological  Survey 
Skyline  Building 
218  E  Street 
Anchorage,  AK  99501. 


Utah 

Utah  Geological  and  Mineral  Survey 

606  Black  Hawk  Way 

Research  Park 

Salt  Lake  City,  UT  84108-1280 


7.  EROS  Data  Center 
U.S.  Geological  Survey 
Sioux  Falls,  SD  57198 

State  Affiliate  Offices 


Washington 

(User  Services) 

Washington  State  Library 
Information  Services  Division 
Olympia,  WA  98504 


1.  Arizona 


Arizona  State  Land  Department 
Information  Resources  Division 
1624  West  Adams,  Room  302 
Phoenix,  AZ  85007 


(Data  Acquisitions) 

Division  of  Management  Services 
Photos,  Maps,  and  Reports  Section 
QW-21,  Public  Lands  Building 
Olympia,  WA  98504 


2    Idaho 

Idaho  State  Historical  Library 
325  West  State 
Boise,  ID  83702 


1 0.   Wyoming 
State  Engineer 
Barrett  Building 
Cheyenne,  WY  82002 


Habitat  Mapping 


69 


II  MAJOR  HABITATS 


5  Forests 

6  Rangelands 

7  Deserts 

8  Tundra 

9  Riparian  Habitats 

10  Marshes 

1 1  Streams 

12  Lakes 


FORESTS 


Jack  Ward  Thomas 

Pacific  Northwest  Forest  and  Range  Experiment  Station 
U.S.  Forest  Service 
LaGrande,  OR  97850 


Jared  Verner 


Editor's  Note:  This  is  the  first  in  a  series  of  chapters 
dealing  with  habitats — other  upland  habitats  in- 
clude deserts,  rangelands,  and  tundra  These  are  not 
clearly  demarked  ecosystems;  they  are  important 
basic  habitat  types  that  overlap  geographically  and 
descriptively.  As  such,  they  require  different  em- 
phases when  planning  inventory  or  monitoring 
studies  of  wildlife  resources. 

While  more  emphasis  has  been  placed  on  inventory 
and  monitoring  of  forested  habitats  than  on  others, 
there  are  no  generally  accepted  or  standardized 
techniques.  Biologists  still  need  to  design  studies 
based  on  local  needs,  while  taking  advantage  of  the 
experiences  of  others. 


Pacific  Southwest  Forest  and  Range  Experiment  Station 
U.S.  Forest  Service 
Fresno,  CA  93710 


"Forest  ecosystems  are  far  too  complicated  for  us  to 
ever  fully  understand  and  precisely  predict  future 
conditions.  Inventories  of  current  conditions,  and 
technologies  for  analyzing  resource  potentials  and 
conducting  management  actions,  are  often  incom- 
plete or  less  than  desired. . . .  None  of  these  are  ex- 
cuses for  lack  of  action.  They  are  simply  the 
circumstances  under  which  resource  managers 
operate." 

—  Hal  Salwasser 


INTRODUCTION 

The  question  to  be  answered  in  this  chapter  is, 
what  are  the  important  attributes  of  forest  habitat  to 
be  measured  if  inventorying  or  monitoring  of  wild- 
life habitat  is  the  goal?  The  answer  will  vary  depend- 
ing on  the  questioner's  interests.  For  example,  if 
the  aim  is  to  obtain  a  general  idea  of  the  habitat  and 
which  species  groups  are  (or  could  be)  present,  it 
may  be  sufficient  to  determine  the  extent  of  each 
forest  type  and  its  attendant  successional  stages  or 
structural  conditions  and  the  presence  or  absence  of 
animal  groups  of  interest.  This  could  be  accom- 
plished either  through  observation  or  from  informa- 
tion relating  vertebrate  species  to  forest  plant 
communities  and  attendant  seres  or  structural  condi- 
tions (see  Thomas  1979b  or  Verner  and  Boss  1980 
as  examples). 

If,  however,  one  desires  sufficient  information  to 
predict  the  response  of  a  single  species  to  habitat 
alterations,  more  specific  habitat  attributes  must  be 
considered.  For  example,  if  Rocky  Mountain  elk 
(Cervus  elaphus  nelsoni)  in  the  heavily  forested 
areas  of  the  intermountain  West  were  of  particular 
interest,  information  should  be  collected  on  stand 
size,  stand  height,  canopy  closure,  density  of  roads 
open  to  vehicular  traffic,  cover/forage  ratios,  distance 
to  edges  between  cover  and  forage,  and,  in  some 
instances,  quantity  and  quality  of  forage.  In  short,  the 
more  specific  the  question,  the  more  detailed  the 
habitat  information  required. 

Jenkins  ( 1977:43)  quoted  Frank  Egler  as  saying, 
"Ecosystems  are  not  only  more  complex  than  we 
think,  they  are  more  complex  than  we  can  think." 
He  then  said, 

". . .  to  deal  with  this  literally  incomprehensible 
complexity,  for  whatever  purpose,  none  of  us 
have  any  choice  but  to  take  a  strenuously  re- 
ductionist approach which  means  making 

hard  choices  about  which  aspects  to  emphasize 


Forests 


73 


and  which  to  disregard.  No  single  set  of 
choices  can  ever  satisfy  everyone  or  every 
purpose,  so  there  will  never  be  a  single  classifi- 
cation system  ...  it  is  both  necessary  and  legiti- 
mate to  tailor  our  efforts  to  our  own 
purposes . . ." 

First,  we  need  a  clear  idea  of  what  to  inventory 
or  monitor  and  why.  There  are  no  pat  answers  to 
the  questions — what  should  be  monitored?  What 
approach  should  be  used?  How  should  the  inventory 
and  monitoring  be  accomplished?  The  answers  de- 
pend on  the  legal  requirements,  management  needs, 
objectives  of  the  wildlife  management,  resources, 
and  skills  of  persons  available  to  do  the  work.  There 
is  not  and  should  not  be  a  list  of  species,  guilds,  or 
life  forms  that  should  be  monitored  on  each  forested 
area.  There  is  not  and  should  not  be  specification  of 
a  vegetative  classification  system  to  be  used.  There  is 
not  and  should  not  be  a  prescribed  list  of  the  right 
techniques  to  be  used  to  inventory  and  monitor 
either  habitats  or  associated  wildlife  species. 


It  sounds  trite,  but  the  only  realistic  answer 
it  all  depends.  This  entire  area  of  activity  is  so  new 
and  so  dynamic,  in  terms  of  management  need  and 
technology  available,  that  the  best  any  reviewer  can 
do  is  to  array  the  existing  information  and  let  the 
user  choose,  considering  the  particular  objectives, 
needs,  and  resources  available. 

One  needs  to  settle  on  a  forest  habitat  classifica- 
tion system  to  structure  the  effort.  This  is  necessary 
to  provide — 

•  a  basis  for  stratifying  forest  habitats  for 
sampling; 

•  a  framework  for  communication  among  various 
disciplines  and  the  public; 

•  a  mechanism  for  aggregating  information  from 
local  areas  into  district,  regional,  and  even 
larger  assessments. 

These  first  steps  of  deciding  what  and  why  to 
inventory  and  of  selecting  an  appropriate  habitat 
classification  system  to  guide  that  effort  are  critical 
and  must  be  carefully  considered.  Decisions  should 
be  predicated  upon  these  questions: 


( 1 )    How  much  will  it  cost  and  can  it  be  afforded? 


(2)    What  steps  can  be  taken  to  ensure  the  useful- 
ness of  a  data  base  that  can  be  developed  at 
a  realistic  cost? 

Intensive  and  frequent  sampling  of  forest  habitat 
is  expensive.  Sampling  that  includes  statistically  valid 
estimates  of  animal  numbers  can  be  prohibitively 


expensive  (Verner  1983).  In  many  situations,  then, 
only  relatively  simple,  nonintensive  surveys  are 
affordable. 

Further,  it  is  likely  that  wildlife  biologists  will 
continue,  at  least  in  the  foreseeable  future,  to  oc- 
cupy a  distinctly  minority  position  in  the  major  land 
management  agencies.  If  they  want  to  be  effective, 
therefore,  they  will  recognize  the  advantages  of  mak- 
ing their  data  bases  compatible  with  those  developed 
by  the  disciplines  that  dominate  the  agency — e.g., 
foresters  and  planners.  Among  the  advantages  are — 

•  increased  probability  that  forest  wildlife  and 
wildlife  habitat  can  be  more  readily  considered 
in  land-use  planning  and  management  decisions; 

•  better  communication  among  disciplines  and 
with  the  public; 

•  possible  reduction  in  the  cost  of  gathering 
needed  information  about  wildlife  habitat. 


CLASSIFICATION  SYSTEMS 

The  habitat  classification  system  used  in  inven- 
tory and  monitoring  of  forest  wildlife  and  their  habi- 
tats is  usually  selected  on  the  basis  of  the  classi- 
fication system  being  used  by  forest  inventory  or 
timber  managers,  usually  both.  Forest  habitat  classifi- 
cation is  complex  and  little  agreement  exists  on  a 
standard  approach  (Bailey  1977).  The  confused  sta- 
tus of  the  situation  is  apparent  from  a  perusal  of  a 
1977  symposium  on  "Classification,  Inventory,  and 
Analysis  of  Fish  and  Wildlife  Habitat"  (Marmelstein 
1977). 

Several  criteria  have  been  used  to  develop  sys- 
tems of  ecosystem  classification,  including  biological 
(Brown  et  al.  1980;  Garrison  et  al.  1977)  and  physi- 
cal (Godfrey  1977).  The  most  commonly  used  stand- 
ard classification  system  for  western  forests  based 
primarily  on  vegetation  was  developed  by  Dauben- 
mire  (1968).  With  some  modification,  Daubenmire's 
system  has  been  extended  to  half  or  more  of  the 
forested  land  of  the  western  U.S.  The  primary  classifi- 
cation systems  in  use  in  the  1 1  western  States  and 
Alaska  are  shown  in  Table  1. 

Bailey  (1982:16)  noted  that— 

"Past  wildlife  studies  and  inventories  have  pro- 
ceeded without  the  benefit  of  an  integrated 
system.  Biologists  often  had  to  depend  on  any 
available,  sometimes  inadequate,  information  or 
devise  their  own  classification,  usually  a  map 
featuring  forest  cover.  Many  investigators  gath- 
ered disconnected  bits  of  descriptive  informa- 
tion on  habitat  without  a  classification  frame- 
work to  give  them  meaning.  Without  such  a 


74 


Forests 


framework,  it  was  very  difficult  and  sometimes 
impossible  to  integrate  wildlife  information 
with  other  information  for  evaluating  trade-offs 
or  interactions  within  the  fish  and  wildlife  re- 
source and  between  it  and  other  resources.  As 
of  1970,  there  was  no  national  approach  to 
integrating  wildlife  information ..." 


What  was  true  in  1970  is  still  true  today.  But 
improvements  have  been  made  and  efforts  are  un- 
derway to  develop  an  acceptable  National  approach 
(Hoekstra  et  al.  1983). 

What  can  be  done  when  it  makes  sense  to  use  a 
finer  or  coarser  breakdown  of  habitats  than  fits  the 


Table  1.     Major  classification  systems  for  forests  of  the  western  States. 


System 

Description 

Source 

Habitat  types 

Daubenmire  1968; 
Pfister  1977 

Land  systems 

Involves  delineation, 
description,  and  evaluation  of 
relatively  homogeneous  units 
of  land  at  local  or  regional 
scale.  Hierarchical 
components  exert  the  most 
control  at  the  top. 

Wertz  and  Arnold  1972;  Bailey 
1976,  1978;  Crowley  1967 

ECOCLASS 

A  method  of  integrating 
several  systems  to  identify 
homogeneous  land  units. 
Developed  by  the  U.S.  Forest 
Service. 

Cortiss  1974;  Buttery  1978 

ECOSYM 

Links  classification  to  land 
management  needs. 

Henderson  et  al.  1978 

National  Site  (Land) 
Classification  System 

Like  ECOCLASS,  a  component 
system  that  involves 
vegetation,  soil,  landform,  and 
water.  Developed  jointly  by  the 
U.S.  Forest  Service,  Bureau  of 
Land  Management,  Fish  and 
Wildlife  Service,  Geological 
Survey,  and  Soil  Conservation 
Service. 

Driscoll  et  al.  1978 

Integrated  Habitat  Inventory 
Classification  System 

A  six-level  system  for 
organizing  species  occurrence 
data  from  plant  communities 
to  physiographic  regions. 
Developed  by  Bureau  of  Land 
Management. 

U.S.  Department  of  the  Interior, 
Bureau  of  Land  Management 
1978 

Land  resource  regions  and 
major  land  resource  areas 

Groups  organizationally  related 
and  geographic  factors  related 
to  land  use,  topography, 
climate,  water,  and  soil. 
Developed  by  the  Soil 
Conservation  Service. 

Austin  1965;  Kuchler  1964 

Land  use  and  cover  mapping 
program 

Broad-based  information,  i.e., 
not  intended  for  wildlife. 
Developed  by  the  U.S. 
Geological  Survey. 

Anderson  et  al.  1976 

Forest  cover  types 

Classification  based  on 
dominant  tree  species. 

Society  of  American  Foresters 
1954 

Forests  75 


classification  system  in  use  by  dominant  forces  in  the 
agency?  A  mechanism  exists  to  allow  data  collection 
at  the  most  meaningful  level  for  specific  wildlife 
purposes.  One  such  mechanism,  developed  for  use  in 
the  Blue  Mountains  of  the  Pacific  Northwest,  has 
proved  useful  there  and  elsewhere  (Table  2).  In 
such  a  system,  the  classifications  are  "crosswalked" 


so  that  information  gathered  under  one  system  can 
be  appropriately  applied  to  others. 

Given  such  information  as  that  shown  in  Table 
2,  data  on  wildlife  habitat  can  be  developed  at  the 
most  appropriate  level  for  the  wildlife  resource  and 
then,  if  appropriate,  can  be  applied  to  other  land 


Table  2.     Relationships  among  four  plant  community  classifications  in  the  Blue  Mountains  of  Oregon  and 
Washington  (adapted  from  Thomas  et  al.  1979a). 


Forest  and  range 
ecosystems  as 
described  by  Garrison  et 
al.  (1977) 

Plant  community 
designations  from 
Thomas  et  al.  (1979a) 

Plant  community  types 
of  the  Blue  Mountains  as 
described  by  Hall  (1973) 

Potential  natural 
vegetation  as  described 
by  Kuchler  (1964) 

Pinyon-juniper' 

Western  juniper 
Juniper/stiff  sage  scab- 
land 
Juniper/low  sagebrush 
Juniper/big  sagebrush 

Juniper/bunchgrass 

K24  Juniper  steppe 
woodland 

Ponderosa  pine 

Ponderosa  pine 

Ponderosa  pine/fescue 

Ponderosa  pine/bitter- 
brush/Ross  sedge 

Ponderosa  pine/blue 
wildrye 

Ponderosa  pine/wheat- 
grass 

K10  Ponderosa  shrub 
forest 

Douglas-fir 

Mixed  conifer 

Ponderosa  pine/Doug- 
las-fir/elk sedge 

Ponderosa  pine/Doug- 
las-fir/snowberry- 
oceanspray 

Ponderosa  pine/Doug- 
las-fir/ninebark 

K12  Douglas-fir  forest 
(interior) 

Larch 

Mixed  conifer/pinegrass- 

ash  soil 
Mixed  conifer/pinegrass- 

residual  soil 

K14  Grand  fir/Douglas-fir 
forest 

Spruce-fir 

White  fir  (grand  fir) 

White  fir/twinflower 
White  fir/big  huckleberry 
White  fir/grouse  huckle- 
berry 

Subalpine  fir/big  huckle- 
berry 

Subalpine  fir/grouse 

huckleberry 
Subalpine  fir/whitebark 

pine/sedge 

K15  Western  spruce-fir 
forest 

Lodgepole 

Lodgepole  pine 

Lodgepole  pine/pine- 
grass-grouse 
huckleberry 

Lodgepole  pine/big 
huckleberry 

Lodgepole/grouse 
huckleberry 

No  provision  in  Kuchler 
(1964) 

1See  specific  reference  citations  for  a  listing  of  scientific  names  of  plants  noted  in  this  table 


76 


Forests 


classification  systems.  The  rule-of-thumb,  then,  for 
selecting  the  appropriate  habitat  classification  system 
to  be  used  as  the  basis  for  inventorying  and  monitor- 
ing has  two  components: 

( 1 )  Select  the  system  in  use  by  the  dominant  profes- 
sional group  in  the  agency  in  order  to  enhance 
communication. 

(  2  )  If  another  finer  or  coarser  grained  classification 
is  used  as  the  framework  for  the  wildlife-habitat 
data  base  (for  whatever  reason),  make  sure  it 
can  be  "crosswalked"  with  other  classification 
systems  (Verner  1984). 

The  ultimate  objective  is  to  present  wildlife 
habitat  in  such  a  manner  that  it  will  be  effectively 
considered  by  land-use  planners  and  managers.  The 
selection  of  a  habitat  classification  system  that  may 
be  technically  superior  for  wildlife-habitat  evaluation 
purposes  is  a  strategic  error  if  it  cannot  be  easily 
translated  into  the  system  used  by  the  dominant 
resource. 

This  requirement  for  crosswalking  is  addressed 
by  Bailey  (1980:17)  in  his  discussion  of  the  ecosys- 
tem concept  of  classification: 

"This  concept  regards  the  earth  as  a  series  of 
interrelated  systems  in  which  all  components 
are  linked,  so  that  a  change  in  any  one  compo- 
nent may  bring  about  a  change  in  other  com- 
ponents and  in  the  operation  of  the  whole.  An 
ecosystem  approach  to  land  evaluation  stresses 
the  interrelationship  among  components 
. . .  Since  ecosystems  are  spatial  systems,  they 
will  be  consistently  inserted,  or  nested,  into 
each  other.  Each  level  subsumes  the  environ- 
ment of  the  system  at  the  level  below  it.  At 
each  level,  new  processes  emerge  that  were 
not  present  or  not  evident  at  the  next  lower 
level . . ." 

In  this  example  (Table  2 ),  the  dominant  land 
management  agency  (U.S.  Forest  Service)  active  in 
the  area  used  Hall's  ( 1973)  description  of  plant  com- 
munities as  the  basic  classification  system  to  guide 
forestry  and  range  management  decisions.  On  the 
whole,  the  divisions  were  too  fine  to  be  germane  to 
the  classification  of  information  about  wildlife-habitat 
relationships.  Therefore,  for  relating  wildlife  to  habi- 
tats, resource  professionals  developing  information 
on  wildlife-habitat  relationships  (Thomas  et  al. 
1979a)  modified  the  classification  system  derived  by 
Garrison  et  al.  (1977).  They  thought  that  these 
broader  community  descriptors  were  more  suitable 
to  the  existing  state-of-knowledge  concerning  wild- 
life-habitat relationships.  The  biologists  involved 
recognized  that  the  vegetative  classification  schemes 
developed  by  Kuchler  ( 1964  and  1970)  were  receiv- 
ing favorable  attention  by  persons  dealing  with  na- 


tional assessments,  but  they  thought  that  such  a 
system  was  too  broad  to  be  meaningful  in  the  con- 
text of  wildlife-habitat  relationships. 

The  decision  about  what  classification  system  to 
use  may  have  already  been  decided  by  the  existence 
of  a  wildlife-habitat  data  base  for  the  area  in  ques- 
tion. The  development  of  such  data  bases  was  stimu- 
lated by  requirements  of  the  National  Environmental 
Policy  Act  of  1969  (which  required  environmental 
impact  statements  for  federally  funded  projects),  the 
Forest  and  Rangelands  Renewable  Resource  Planning 
Act  of  1974,  the  National  Forest  Management  Act  of 
1976,  and  the  Federal  Land  Policy  and  Management 
Act  of  1976.  The  first  such  data  bases  for  the  west- 
ern States  were  developed  by  Patton  ( 1978)  for  the 
Southwest  and  Thomas  ( 1979b)  and  associates  for 
the  Pacific  Northwest.  These  initial  publications  were 
quickly  followed  by  similar  efforts  for  the  western 
Sierra  Nevada  in  California  (Verner  and  Boss  1980), 
the  Great  Basin  of  southeastern  Oregon  (Maser  and 
Thomas  1983),  and  western  Oregon  and  Washington 
(Brown,  in  press).  Parallel  efforts  are  underway  in 
several  other  areas  in  the  western  U.S.  and  Canada. 
The  Fish  and  Wildlife  Habitat  Relationships  Program 
of  the  U.S.  Forest  Service  was  established  to  encour- 
age the  development  of  similar  programs  throughout 
the  U.S.  and  to  enhance  development  of  additional 
habitat  analysis  techniques.  Cushwa  and  DuBrock 
(1982)  detailed  various  wildlife  data  bases  in  exist- 
ence or  being  developed  by  various  State  and  Fed- 
eral agencies. 

Various  permutations  of  this  basic  approach  to 
wildlife-habitat  analysis  seem  to  be  most  widely  used 
in  dealing  with  analysis  and  management  of  forested 
habitats  of  the  western  U.S.  The  basis  of  the  original 
system  was  to  consider  wildlife  habitats  at  three 
levels: 

(  1 )  The  relationship  of  groups  and  individual  spe- 
cies to  major  plant  communities  and  their 
successional  stages  or  structural  conditions  for 
activities,  including  feeding  and  reproduction; 

(2)  A  description  of  special  habitats  that  were  not 
adequately  considered  through  plant  communi- 
ties and  successional  stages; 

(3)  The  habitat  requirements  of  featured  species, 
i.e.,  the  species  that  receive  special  management 
attention. 


MAJOR  SPECIES  GROUPS 

The  purpose  of  this  section  is  to  review  and 
evaluate  the  existing  approaches  to  forming  groups 
of  species  for  inventory  and  monitoring  in  forests  so 
as  to  reflect  habitat  conditions. 


Forests 


77 


When  dealing  with  the  vertebrate  component  of 
the  forest  ecosystem,  it  is  necessary  to  determine 
which  species  or  group  of  species  need  to  be  inven- 
toried. Obviously,  there  are  insufficient  knowledge, 
skill,  and  resources  to  deal  with  inventory  and  moni- 
toring of  all  the  species. 

When  wildlife-habitat  data  bases  are  developed, 
each  individual  species  must  be  considered.  But  if 
the  relationships  of  all  vertebrate  species  within  an 
area  to  habitat  features  are  considered,  the  detail  and 
sheer  volume  of  data  can  be  overwhelming.  When 
possible,  it  would  be  more  convenient  and  cheaper 
to  use  groups  of  species  rather  than  individuals. 

Featured  Species 

The  desirability  of  the  featured  species  approach 
(examining  the  relationship  of  groups  of  species  to 
habitat  variables)  is  in  keeping  with  the  necessity  for 
the  "strenuous  reductionist  approach"  mentioned 
earlier  in  relation  to  habitat  classification.  A  group 
called  the  Committee  of  Scientists,  who  recom- 
mended regulations  as  required  by  the  National  For- 
est Management  Act  of  1976,  called  for  use  of 
indicator  species  in  setting  up  and  monitoring  forest 
management  plans.  These  indicators  were  to  serve 
as  surrogates  for  the  welfare  of  a  group  of  species. 

Assume  that  the  spotted  owl  (Strix  occiden- 
talis)  was  chosen  as  the  indicator  species  for  species 
that  thrive  in  old-growth  Douglas-fir  (Pseudotsuga 


menziesii)  forests  in  western  Washington.  If  the 
spotted  owl's  habitat  requirements  were  met,  ac- 
cording to  the  indicator  species  concept,  other  spe- 
cies that  require  old-growth  Douglas-fir  would  be 
provided  for  simultaneously. 

Verner  (1984)  reviewed  the  use  of  the  guild 
concept  as  applied  to  management  and  essentially  re- 
jected the  indicator  species  concept  on  practical 
and  technical  grounds.  A  guild  is  made  of  species 
within  an  order  that  jointly  exploit  the  same  re- 
source in  an  ecosystem.  He  did  believe  that  the  use 
of  whole  guilds  (i.e.,  each  bird  of  any  species  within 
a  guild  is  tallied  in  sampling)  to  evaluate  habitat 
conditions  might  be  useful  but  recommended  that 
(p.  1 )  "much  testing  must  be  done  before  it  is  ap- 
plied." This  means  that  the  research  community  has 
not  yet  tested  the  hypotheses  inherent  in  the  indica- 
tor species  approach.  Field  biologists  who  are  using 
the  approach  should  admit  that  to  themselves  and  to 
the  users  of  that  work. 

Guilds.  The  guild  concept  assumes  that  animal 
species  within  an  order  can  be  grouped  by  how 
similarly  they  use  environmental  resources  (Root 
1967).  Severinghaus  (1981 )  applied  the  guild 
concept  to  species  of  vertebrates  according  to  their 
response  to  habitat  and  habitat  change,  delineating 
30  mammal  and  31  bird  guilds.  He  said  (p.  189), 
"The  usefulness  of  guilds  . . .  relies  on  the  principle 
that . . .  actions  impacting  one  member  of  the  guild 
should  impact .  . .  other  members  in  a  similar  way." 

Short  and  Burnham  ( 1982)  and  Short  (1983) 
presented  a  technique  for  structuring  guilds  for  use 
in  evaluating  impacts  of  habitat  alterations  on  wild- 
life communities.  Described  are  mathematical  and 
computer  aids  to  form  guilds,  as  opposed  to  the 
more  intuitive  classifications  proposed  by  Thomas  et 
al.  ( 1979a),  Severinghaus  ( 1981),  and  others. 

The  Short-Burnham  and  other  mathematically 
derived  techniques,  although  holding  promise,  are 
probably  more  complex  than  needed  for  application 
by  most  field  biologists.  Mathematical  sophistication 
may  exceed  biological  sensibility  in  this  approach 
(Verner  1984).  It  might  be  best  to  contract  for  these 
services  if  this  approach  is  broadly  applied  in 
management. 

Thomas  et  al.  ( 1979a),  in  an  effort  to  respond  to 
the  recommendations  of  the  Committee  of  Scientists, 
modified  a  suggestion  of  ornithologist  Antti  Haapa- 
nen  ( 1965,  1966)  that  species  of  birds  could  be 
grouped  according  to  their  habitat  use  into  "life 
forms."  They  extended  the  concept  to  include  all 
vertebrates  and  grouped  the  species  (379)  that  oc- 
curred in  the  Blue  Mountains  of  Oregon  and  Wash- 
ington into  16  life  forms.  Life  forms  include 
groupings  of  vertebrates — regardless  of  order — that 
require  similar  habitat  attributes. 


78 


Forests 


Oversophisticated  mathematics  may  shed  great  darkness 
on  the  subject. 

What  those  initial  life  forms  were  is  not  as  im- 
portant as  the  concept;  for  habitat  management,  ver- 
tebrate species  could  be  grouped  by  the  general 
type  of  habitat  they  used  for  feeding  and  reproduc- 
tion, irrespective  of  taxonomic  considerations.  The 
concept  assumes  only  that  each  species  in  a  life  form 
will  be  present  in  suitable  habitat.  It  does  not,  how- 
ever, assume  that  species  within  the  life  form  use  the 
habitat  similarly  or  that  numbers  of  the  various  spe- 
cies will  rise  and  fall  simultaneously. 

Life  forms  are  constructed  to  fill  a  management 
need,  and  the  number  of  life  forms  is  limited  only  by 
the  amount  of  knowledge  available  on  the  relation- 
ship of  the  individual  species  to  habitat.  Life  forms 
can  be  created  or  eliminated,  or  species  can  be 
moved  between  life  forms  to  fit  the  need  or  to  re- 
spond to  new  information.  It  is  also  important  to 
understand  that  in  using  the  life  form  approach,  it  is 
the  habitat  condition  that  is  inventoried  and  moni- 
tored, not  the  species  in  the  life  form. 

The  three  approaches  discussed  for  grouping 
species  in  order  to  relate  them  to  habitats  (indicator 
species,  guilds,  and  life  forms)  share  the  idea  that 
animal  species  can  be  grouped  according  to  re- 
sponse to  habitat.  If  so,  that  grouping  can  be  a  real 
aid  in  the  simplification  of  habitat  analysis  across  the 
spectrum  of  the  vertebrate  species  present  in  the 
area  to  be  considered. 

Other  approaches  have  included  those  of  Root 
(1967),  MacMahon  et  al.  (1981),  andjaksic  (1981). 
Techniques  used  to  formulate  guilds  include  cluster 
analysis  (Crome  1978),  principal  components  analy- 
sis (Holmes  et  al.  1979;  Landres  and  MacMahon 
1980;  Short  and  Burnham  1982),  canonical  correla- 
tion (Folse  1981),  discriminant  function  analysis 
(DeGraaf  and  Chadwick  1984;  Dueser  and  Shugart 


1979;  Lindeman  et  al.  1980;  Nie  et  al.  1974;  Rao 
1966;  Williams  1980),  and  polar-ordination  (Ander- 
son 1971;  Austin  and  Orloci  1966;  Bannister  1968; 
Bray  and  Curtis  1957;  DeGraaf  and  Chadwick  1984; 
Gauch  1982).  Note  that  all  these  techniques  produce 
investigator-defined  guilds,  "a  fact  that  introduces 
circularity  into  their  use  for  examining  ecological 
questions"  (Verner  1984:1 ).  This  simply  means  that 
there  is  a  danger,  if  the  investigator  defines  the  guild, 
of  altering  the  expected  to  fit  the  observed  species- 
habitat  conditions  and  vice-versa. 

It  seems  inappropriate  to  present  lists  of  life 
forms  or  guilds  here.  They  should  be  developed  by 
the  user  from  the  available  wildlife  data  base  on  the 
occurrence  of  animal  species  and  their  relationship 
to  various  habitat  conditions.  The  only  advantage 
of  dealing  with  life  forms  or  guilds  is  the  enormous 
savings  that  can  result  from  monitoring  whole  guilds 
instead  of  indicator  species.  Conversely,  if  the  deci- 
sion is  made  to  monitor  indicator  species,  the  spe- 
cies groupings  delineate  those  species  that  the 
indicator  species  represents.  We  reject  the  indicator 
species  approach  to  monitoring  as  neither  reasonable 
nor  practical  for  the  reasons  presented  by  Verner 
(1984).  In  practice,  most  management  analysts  have 
fallen  back  on  consideration  of  individual  species — 
either  en  masse  or  of  a  select  few  of  particular 
interest. 

We  recommend  the  following: 

( 1 )  Do  not  use  the  indicator  species  approach  un- 
less regulations  require  it.  If  the  indicator  spe- 
cies approach  is  used,  recognize  that  it  is  merely 
a  hypothesis  and  that  the  users  must  provide 
the  test. 

(2)  Use  the  guild  approach,  as  it  seems  to  be  the 
most  economical  and  practical  approach  to 
monitoring  species  as  reflective  of  habitat 
condition. 

(3)  Use  the  life  form  approach  when  monitoring  is, 
for  financial  or  other  reasons,  restricted  to  vege- 
tative conditions;  it  is  essential  to  understand 
that  the  approach  involves  a  hypothesis  that 
must  be  tested  by  the  user. 


CRITICAL  HABITAT  FEATURES 

Soil 

Soil  is  a  primary  variable  contributing  to  the 
development  of  plant  communities.  With  the  excep- 
tion of  fossorial  vertebrates,  the  soil  type  per  se  does 
not  appear  to  be  a  controlling  factor  in  animal  habi- 
tat. The  relationships  between  soils  and  wildlife  are 
thoroughly  discussed  by  Robinson  and  Bolen  (1984). 


Forests 


79 


Thomas  et  al.  (1979a:26)  suggested  that — 

"Plant  communities  and  their  successional 
stages  have  unique  environmental  conditions 
that  are  ecologically  important  as  niches  for 
wildlife  species  . . .  The  niches  are  a  product  of 
the  plant  community,  its  successional  stages, 
and  other  environmental  factors — including  soil 
type,  moisture  regime,  microclimate,  slope, 
aspect,  elevation,  and  temperature.  The  com- 
plex interactions  of  site  and  plant  community 
structure  could  be  dissected  and  the  more 
precise  of  each  on  the  animal  community  de- 
termined. If  such  information  existed,  it  would 
probably  be  too  complex  to  use  readily.  The 
plant  community  type,  however,  can  be  consid- 
ered an  integrator  of  the  many  factors  interact- 
ing on  the  site." 

The  factors  that  plant  communities  integrate 
include  soils.  We  recommend  that  soils  not  be  a  part 
of  monitoring  unless  there  are  particular  reasons  for 
including  them,  i.e.,  the  soil  texture  has  particular 
significance  to  a  wildlife  species  of  sufficient  interest 
that  monitoring  is  deemed  appropriate.  Even  then, 
it  is  important  to  identify  what  soil  attribute  is  signif- 
icant to  the  occurrence  of  the  species  in  question. 
This  could  include  such  things  as  texture,  size  of  soil 
particles,  cohesiveness,  moisture  retention  proper- 
ties, depth,  and  drainage. 

Physical  Features 

Maser  et  al.  (1979a,  b)  noted  that  geomorphic 
and  edaphic  habitat  features  occurred  within  the 
general  vegetative  mosaic  and  have  special  values  to 
wildlife.  These  "unique"  or  "geomorphic"  habitats 
included  cliffs,  caves,  talus,  lava  flows,  sand  dunes, 
and  playas.  Each  of  these  habitats  supports  one  or 
more  species  not  found  in  the  general  forest 
environment. 


Talus  is  the  accumulation  of  rocks  at  the  base  of  steep 
slopes. 


Verner  and  Boss  (1980:2-3)  called  such  compo- 
nents "special  habitat  requirements,"  adding  ground 
burrows,  friable  soils,  moist  soil,  earthen  bank,  rock 
outcrops,  limestone  outcrops,  and  crevices.  The  spe- 
cies to  which  such  features  are  important  were  iden- 
tified. For  example,  friable  soils  are  important  to 
burrowing  animals;  dead-and-down  woody  material  is 
critical  to  reptiles,  amphibians,  and  small  mammals; 
caves  harbor  bats;  and  cliffs  provide  nest  sites  and 
roosts  for  raptors. 

Plant  Communities  and  Successional  Stages 

The  forest  habitat  types  (from  whatever  classifi- 
cation scheme  is  chosen  for  use)  are  the  first  vegeta- 
tive component  considered.  Forest  condition  within 
each  habitat  type  is  described  by  successional  stage. 
Thomas  et  al.  (1979a)  identified  six  such  stages: 

( 1 )  grass  forb 

(2)  shrub  seedling 

(3)  pole-sapling 

(4)  young 

(5)  mature 

(6)  old-growth 

Verner  and  Boss  ( 1980)  suggested  four  such  stages 
for  forest  habitats: 

(1)  grass/forb 

(2)  shrub/seedling/sapling 

(3)  pole/medium  tree 

(4)  large  tree 

Special  Habitat  Features 

According  to  Thomas  et  al.  (1979a:21),  special 
habitats  are  "biological  in  nature,  can  be  manipulated 
by  the  forest  manager,  and  play  a  critical  role  in  the 
lives  of  at  least  some  species.  These  habitats  include 
riparian  zones,  edges,  snags,  logs,  and  other  dead 
woody  material  on  the  forest  floor."  Verner  and  Boss 
(1980)  added  forest  openings,  elevated  perches, 
nest  cavities,  hollows,  and  litter  to  the  list  of  what 
they  called  "vegetation  elements."  Which  of  these 
special  habitats  should  be  included  in  inventories  or 
monitoring  operations  depends  on  the  wildlife  spe- 
cies or  groups  of  species  considered  important  by 
the  responsible  land  manager.  For  example,  snags 
would  be  important  if  cavity-nesting  birds  received 
management  emphasis,  and  logs  and  other  dead 
woody  material  would  be  of  interest  if  reptiles  and 
amphibians  were  of  interest.  Of  these,  we  believe 
that  only  riparian  zones,  edges,  snags,  logs,  and  forest 
openings  will  generally  be  considered  critical  inven- 
tory elements. 

Riparian  Zones.  Riparian  zones  are  those  areas 
identified  by  the  presence  of  vegetation  that  requires 
free  or  unbound  water.  When  they  occur  within 


80 


Forests 


forested  areas,  they  generally  create  well-defined 
mesic  areas  within  much  drier  surrounding  areas. 
They  make  up  a  minor  portion  of  the  area,  are  more 
productive  of  biomass  than  surrounding  forests,  and 
are  a  dramatic  source  of  diversity.  Of  the  378 
vertebrate  species  in  the  Blue  Mountains  of  Oregon, 
285  were  identified  as  dependent  on  riparian  zones 
or  as  using  them  more  frequently  than  other  habitats 
(Thomas  et  al.  1979b). 

Edges.  Edges  are  places  where  different  plant 
communities  or  successional  stages  meet  and  are 
identified  as  particularly  rich  in  wildlife.  Leopold 
(1933:131)  stated  that  "game  wildlife  is  a 
phenomenon  of  edges  . . .  We  do  not  understand  the 
reason  for  all  of  these  edge  effects,  but  in  those  cases 
where  we  can  guess  the  reason,  it  usually  harks  back 
either  to  the  desirability  of  simultaneous  access  to 
more  than  one  environmental  type,  or  the  great 
richness  of  border  vegetation  or  both."  Patton 
(1975)  suggested  that  the  amount  of  edge  per  unit 
of  area  might  serve  as  an  index  to  diversity. 

Thomas  et  al.  (1979d)  divided  edges  into  inher- 
ent edges  (those  resulting  from  the  meeting  of  two 
plant  communities)  and  induced  edges  (those  be- 
tween two  successional  stages  within  a  plant  com- 
munity). They  suggested  further  that  the  degree 
of  contrast  in  height  of  vegetation  along  an  edge 
would  be  correlated  with  the  diversity  of  animals 
encountered  along  the  edge  and  in  adjacent  stands. 
Initial  testing  supports  this  hypothesis. 

Snags.  Snags  are  dead  or  partly  dead  trees  at  least 
10.2  cm  (4  in.)  in  diameter  at  breast  height  (dbh); 
snags  at  least  1.8  m  (6  ft)  tall  are  critical  to  many 
species  of  forest  wildlife.  For  example,  in  the  Blue 
Mountains,  39  bird  and  23  mammal  species  (16%  of 
the  379  species  occurring)  use  snags  for  nesting  or 
shelter.  Snags  are  now  recognized  by  managers  of 
public  forest  land  as  a  critical  habitat  component. 

Recommendations  on  size,  species,  and  numbers 
of  snags  required  to  support  various  population  lev- 
els of  wildlife  dependent  on  snags  are  beginning  to 
emerge  (Thomas  et  al.  1979b).  The  following  infor- 
mation on  snags  should  be  gathered: 

•  species 

•  height 

•  dbh  inside  bark 

•  degree  of  decay  ( hard  or  soft ) 

•  top  condition  (whole  or  broken) 

•  surrounding  vegetative  structure 

•  density 

•  use  by  wildlife 

Logs.  Logs  (dead-and-down  woody  material)  are 
dead  trees  or  portions  of  trees  lying  on  the  forest 


Logs  partly  submerged  in  water  form  an  important  link 
between  terrestrial  and  aquatic  habitats  for  some  species. 


Snags  are  critical  habitat  components  for  many  species. 


Forests 


81 


floor.  Logs  are  important  wildlife  habitat  components 
(Maser  et  al.  1979c).  Elton  (1966:279)  said— 

"When  one  walks  through  the  rather  dull  and 
tidy  woodlands  . . .  that  result  from  modern 
forestry  practices,  it  is  difficult  to  believe  that 
dead  wood  provides  one  of  the  two  or  three 
greatest  resources  for  animal  species  in  a  natu- 
ral forest,  and  that  if  fallen  timber  and  slightly 
decayed  trees  are  removed,  the  whole  system 
is  gravely  impoverished  of  perhaps  more  than  a 
fifth  of  its  fauna." 

In  the  Blue  Mountains,  for  example,  5  amphibi- 
ans, 9  reptiles,  116  birds,  and  49  mammals  (47% 
of  the  379  species  present)  make  some  use  of  dead- 
and-down  woody  material  (Maser  et  al.  1979c)  (Fig- 
ure 1 ).  This  is  evidently  common  because  Elton 
(1966:279),  speaking  of  England,  said  "...  indexes  of 
the  Geological  Survey  contain  456  species  of  animals 
(including  invertebrates)  at  Wytham  living  in  wood 
or  under  bark  where  decay  had  begun  or  already 
gone  far.  Another  518  species  are  known  to  occur  in 
this  habitat  elsewhere  in  Britain."  Maser  et  al. 
(1979c)  presented  a  state-of-knowledge  review 
about  dead-and-down  woody  material  as  wildlife 
habitat.  They  also  presented  information  about  the 
role  of  such  material  in  forest  ecology. 

Forest  Openings.  Forest  opening  areas  within  a 
forest  that  do  not  support  trees  (for  whatever 
reason)  are  an  important  source  of  diversity  within 
the  forest.  They  frequently  support  species  that  do 


Openings  provide  diversity  in  a  forest. 


not  occupy  forested  habitats  per  se,  and  they  are 
important  to  species  associated  with  their  edges.  For 
example,  more  elk  and  mule  deer  (Odocoileus 
hemionus)  occur  within  about  183  m  (600  ft)  of 
edges  between  forest  cover  and  forage  areas 
(Reynolds  1966;  Harper  1969;  Leckenby  1984).  This 
was  a  key  factor  in  the  development  of  elk  and  deer 
habitat  models  that  are  receiving  wide  application 
(see  Thomas  et  al.  1979c  for  an  example).  The 
inventory  of  forest  openings  can  be  accounted  for 
under  the  inventory  of  plant  communities, 
successional  stages,  and  edges  discussed  earlier. 

Structural  Conditions  of  the  Forest 

Even  greater  detail  may  be  required  for  struc- 
tural conditions  of  the  forest  beyond  what  can  be 
inferred  from  knowledge  of  the  successional  stage  of 
a  given  plant  community.  Included  are  the  following: 

•  dbh  of  trees,  treating  conifers  and  broadleafs 
separately 

•  tree  height  and  height  of  the  base  of  the 
crown 

•  canopy  volume — the  space  occupied  by  the 
crowns  of  trees 

•  canopy  closure 

•  understory  volume — the  space  occupied  by 
the  leaves  and  limbs  treating  shrubs  separately 
from  grasses  and  forbs 

•  sight  distance 

•  stand  size 

Dbh.  Dbh  can  be  important  in  dealing  with 
management  of  snags  and  dead-and-down  woody 
material.  Living  trees  are  the  snags  and  logs  of  the 
future.  Division  into  conifers  and  broadleaved  trees 
can  be  important,  as  many  birds  and  some  arboreal 
mammals  are  influenced  by  this  variable  (Thomas  et 
al.  1977). 

Tree  Height  and  Height  of  Base  of  Crown.  Tree 
height  and  the  height  of  the  base  of  the  crown  must 
be  known  if  the  relationship  between  vegetative 
volume  and  wildlife  species  is  to  be  considered. 

Canopy  Volume.  The  importance  of  considering 
canopy  volumes,  particularly  for  purposes  of 
explaining  bird  occurrence,  has  been  emphasized  by 
a  number  of  investigators  (reviewed  in  Larson 
1981).  MacArthur  (1964)  believed  that  variability  in 
bird  occurrence  due  to  foliage  height  diversity  could 
be  accounted  for  by  considering  three  layers  of 
vegetation:  0-60,  60-750,  and  750  cm  (0-2,  2-25,  and 
25  ft).  Thomas  et  al.  (1977),  however,  found  that 
considering  vegetative  volumes  in  1.5-m  (5-ft)  layers 


82 


Forests 


The  trunk  provides 
a  food  source  for 
woodpeckers,  particularly 
pileated  woodpeckers. 


The  root  wad  is  used  by 
flycatchers  for  perching, 
by  grouse  for  dusting,  and 
by  juncos  for  nesting 


Elevated  areas  are 
used  as  lookouts 
and  feeding  sites. 


Limbs  are  used  as 
perches,  and  if  hollow, 
as  nest  cavities 


The  spaces  between  loose  bark 
and  wood  are  used  as  hiding 
and  thermal  cover  by  invertebrates 
and  small  vertebrates,  such  as  the 
Pacific  treefrog. 


Protected  areas  under  the 
log  are  used  as  nesting 
cover  by  grouse  and  as 
hiding  and  thermal  cover 
by  snowshoe  hares. 


Figure  1.     Downed  logs  furnish  many  structural  features  important  to  wildlife. 


77k-  myriad  of  openings  and  vegetation  types  in  this  urea  form  edges  important  to  wildlife. 


Forests 


83 


markedly  increased  the  variability  in  bird  occurrence 
explained  by  vegetative  volumes.  Mawson  et  al. 
(1976)  described  a  technique  called  HTVOL  (height- 
volume)  whereby  trees  were  measured  within  a 
plot  in  a  way  that  vegetative  volumes  in  various 
"slices"  could  be  determined.  Each  tree  to  be 
sampled  was  assigned  to  a  geometric  form  and 
appropriate  measurements  taken.  The  degree  of 
detail  needed  to  determine  vegetative  volumes 
depends  on  the  investigator's  need. 

Canopy  Closure.  Canopy  closure  provides 
information  that  is  useful  in  describing  the  stand 
quality  in  terms  of  thermal  cover,  i.e.,  cover  that  is 
used  by  ungulates  to  thermoregulate  during 
extremes  of  ambient  temperature  (Thomas  et  al. 
1979e)  (Figure  2).  Such  information  can  also  be 
useful  in  predicting  the  productivity  of  understory 
vegetation  within  such  stands. 


/Thermal  cover  conditions 

/        $          /               /o 

/          ••£■           /                  /  o 

/          <P>           /       £        /  £ 

'          £           /       §       /  $ 

„  IUU 

c 

CD 

o 

g  80 

1/5 

f  / 

1/1 

CD 

cB    60 

- 

'  J* 

> 

<F 

o 
%    40 

A< 

CD 

> 

O 

o     „„ 

_    20 

— 

CO 

E 

CD 

£     n 

■  i  i 

i    i 

•   i 

0 

20          40          60          80       100 

Canopy  closure  (percent) 

Figure  2.     Effectiveness  of  thermal  cover  for  deer 
and  elk  is  closely  related  to  canopy  cover. 


Understory  Volume.  The  volume  of  understory 
vegetation  (particularly  shrubs)  is  useful  in 
predicting  bird  occurrence  (MacArthur  1964;  also 
see  review  by  Larson  1981)  and  as  a  measure  of 
forage  available  for  herbivores.  Forage  availability 
can  be  expressed  as  kg/ha  (lb/a.)  of  air-dried  forage. 

Sight  Distance.  Sight  distance  is  the  distance  at 
which  90%  or  more  of  a  standing  deer  or  elk  is 
hidden  from  view.  Sight  distance  can  be  useful  in 
evaluating  the  effectiveness  of  hiding  cover  for  deer 
or  elk.  Hiding  cover  has  been  denned  as  existing 


when  an  elk  or  deer  is  90%  or  more  hidden  at  a 
distance  of  61  m  (200  ft)  or  less  from  an  observer 
(Black  et  al.  1976). 

Stand  Size.  Stand  size  is  an  important  habitat 
variable  for  several  reasons.  For  example,  optimum 
stand  sizes  are  recommended  to  ensure  use  by  elk 
and  deer.  These  sizes  usually  range  from  10  to  16  ha 
(25  to  40  a.).  Some  consider  stand  width  to  be  a 
more  important  measure  of  stand  size  than  area,  as 
this  characteristic,  because  of  edge  effect,  has  the 
most  influence  (positive  and  negative)  on  animal 
occurrence  and  density  (Leckenby  1984).  For 
example,  recommendations  for  optimum  habitat  for 
deer  and  elk  call  for  stand  width  to  be  held  at  183- 
366  m  (600-1,200  ft;  Thomas  et  al.  1979c).  In 
addition,  stand  size  is  related  to  the  number  of 
species  that  will  be  found  in  the  stand  and  the 
diversity  of  species  within  the  stand.  The  number  of 
species  that  are  found  within  a  stand  is  positively 
correlated  with  the  size  of  the  stand  up  to  some  size. 
This  relationship  is  called  the  species/area  curve; 
see  Cain  and  Castro  (1959),  Greig-Smith  (1964), 
Galli  et  al.  (1976),  and  Whitcombe  et  al.  (1981)  for 
reviews  of  this  concept. 


Species  Composition 

It  is  sometimes  desirable  to  consider  the  species 
of  plants  involved  in  a  habitat.  This  information  is 
gathered  in  addition  to  categorization  of  the  plant 
community  (identified  by  characteristic  dominant 
vegetation  in  the  climax  stage)  and  the  successional 
stage.  The  following  examples  show  why  data  about 
species  composition  may  be  valuable  in  wildlife  habi- 
tat assessment: 

•  variability  in  palatability  and  food  value  in  for- 
age plants  between  consumer  species; 

•  differences  in  values  of  various  species  of  trees 
to  cavity  nesters; 

•  indication  of  range  condition. 

Remember,  however,  that  collection  and  analysis  of 
inventory  data  including  species  composition  in- 
creases costs  many  times.  In  some  instances,  such  as 
the  classification  of  plant  communities,  it  may  be 
sufficient  to  record  plant  species  as  present  or  absent 
or  to  deal  with  estimates  of  relative  abundance. 


INVENTORY 

Broadscale  Systems 

The  broadscale  inventorying  and  monitoring  of 
wildlife  populations  and  wildlife  habitats  are  man- 
dated by  law  on  National  Forest  lands  but  little  has 


84 


Forests 


been  done  to  date  (Salwasser  et  al.  1983)-  There  are 
no  well  developed,  well  tested  systems  to  accom- 
plish those  tasks,  although  some  efforts  are  underway 
to  produce  the  tools  to  accomplish  them  (Hoekstra 
et  al.  1979,  1983;  Verner  1983).  The  systems  that 
are  available  have  been  called  "working  hypotheses," 
and  their  developers  advise  caution  in  their 
application. 


Special  Habitat  Features 

This  first  level  of  analysis  may  not  be  sufficient 
to  deal  with  special  habitat  features  or  with  the  de- 
tailed requirements  of  individual  species  of  special 
concern  to  management.  Special  and  unique  habitat 
features  are  then  considered.  These  were  listed  and 
discussed  earlier. 


The  basic  system  that  has  received  the  broadest 
application  on  forested  land  in  western  North  Amer- 
ica was  initially  described  by  Thomas  ( 1979a)  and 
associates  and  has  become  known  as  the  Wildlife  and 
Fish  Habitat  Relationship  (WFHR)  approach.  The 
system  has  been  and  is  being  modified  in  the  permu- 
tations being  developed  for  other  ecosystems  in  the 
U.S.  and  Canada.  Examples  include  the  application 
for  the  western  Sierra  Nevada  (Verner  and  Boss 
1980),  western  Oregon  and  Washington  (Brown,  in 
press),  and  the  Oregon  portions  of  the  Great  Basin 
(Maser  and  Thomas  1983).  This  general  approach 
has  been  used  in  many  other  areas  in  the  West,  but 
the  adapted  models  and  data  bases  have  not  been 
formally  published.  Scientists  in  British  Columbia  are 
working  on  the  system  to  help  describe  the  ecosys- 
tems of  British  Columbia. 


All  applications  of  this  general  approach  attempt 
to  identify  habitat  needs  for  feeding  and  reproduc- 
tion (some  adaptations  include  other  activities  such 
as  resting)  by  each  vertebrate  species  for  each  plant 
community,  according  to  successional  stage  or  struc- 
tural condition.  Most  approaches  use  the  formation 
of  life  forms  or  guilds  to  facilitate  analysis  and  moni- 
toring. Expected  changes  in  habitat  can  be  superfi- 
cially evaluated  by  comparing  the  wildlife  in  the 
existing  habitat  to  those  in  the  anticipated  habitat. 
Predicted  changes  in  species  occurrence  are,  indeed, 
the  evaluation.  Some  species  will  be  helped,  some 
will  be  hurt,  and  some  will  not  be  affected  by  any 
major  change  in  a  given  habitat.  The  impact  on  each 
species,  or  group  of  species,  can  be  evaluated  at  a 
crude  level. 

The  required  habitat  inventory  for  this  level  of 
analysis  is  relatively  quick  and  easy  to  accomplish  by 
simple  delineation  of  plant  communities  and  their 
successional  stages  or  structural  conditions.  Aerial 
photo-interpretation  is  easily  applied,  and  imagery 
from  multispectral  scans  obtained  from  satellites 
(LANDSAT)  shows  promise  in  this  regard  and  is 
being  effectively  applied  in  the  Blue  Mountains  of 
Oregon  and  Washington  by  the  Oregon  Department 
of  Fish  and  Wildlife  and  the  U.S.  Forest  Service  in 
monitoring  big  game  habitat  conditions.  It  has  been 
particularly  useful  in  deriving  forage-cover  ratios, 
sizing  and  spacing  of  cover  and  forage  areas,  and 
quantification  of  cover  quality. 


Here  we  use  snags  as  an  example.  Standard  for- 
est management  regimes  are  unlikely  to  provide 
snags  in  adequate  numbers  to  support  viable  popula- 
tions of  all  members  of  the  life  forms  that  excavate 
cavities  in  dead  trees  for  nesting  or  for  those  that 
occupy  existing  cavities  for  nesting.  Therefore,  a 
snag  inventory  should  be  routinely  done  in  managed 
stands  if  the  welfare  of  the  wildlife  species  that  use 
snags  is  of  concern.  This  can  be  accomplished  by 
regular  inventorying  by  forest  crews  whose  primary 
job  is  to  provide  data  to  support  the  timber  manage- 
ment program  (and  who,  importantly,  are  well  fi- 
nanced to  do  that  job)  or,  with  less  accuracy,  by  use 
of  aerial  photography. 

Admittedly,  snags  are  an  obvious  and  relatively 
easy  special  habitat  feature  to  inventory  and  monitor. 
They  are  also  a  feature  that  has  received  considera- 
ble management  attention.  Other  features,  such  as 
dead-and-down  woody  material,  edges,  talus,  and 
others  have  not  been  routinely  inventoried.  Whether 
these  features  should  be  inventoried  depends  on 
management  objectives.  If  there  are  wildlife  species 
that  depend  on  such  features  that  are  of  significant 
interest  to  management,  the  associated  habitat  fea- 
ture should  be  monitored.  The  relationship  of  wild- 
life species  to  several  special  habitat  features  is 
detailed  in  such  publications  as  Thomas  (1979b). 
The  techniques  for  inventory,  if  they  exist,  have  not 
become  standardized.  Also,  there  is  no  standard  list 
of  special  habitat  features.  Whether  to  inventory  and 
monitor  special  habitat  features,  and  which  ones  to 
address,  are  determined  by  the  habitat  management 
objectives  of  the  area  in  question.  Prime  candidates 
for  inventory  and  monitoring  are  those  that  are  criti- 
cal to  featured  species,  indicator  species,  or  manage- 
ment guilds.  Significance  and  priority  for  attention 
are  determined  by  emphasis  offered  by  the  manage- 
ment plan. 

When  the  needs  of  individual  species  must  be 
considered,  additional  habitat  features  germane  to 
management  of  that  species  will  become  important. 
This  is  illustrated  by  the  discussion  of  featured  spe- 
cies that  follows. 

Featured  Species 

Featured  species,  or  species  of  special  interest 
to  managers,  usually  require  special  consideration. 


Forests 


85 


These  usually  include  game  species,  threatened  or 
endangered  species,  management  indicator  species, 
and  ecological  indicator  species.  The  National  Forest 
Management  Act  and  the  regulations  issued  pursuant 
to  that  Act  require  the  maintenance  of  viable  popula- 
tions of  all  native  and  desirable  non-native  species 
in  National  Forest  management.  The  regulations 
specify  that  progress  toward  achievement  of  that 
goal  be  evaluated  by  monitoring  the  occurrence  and 
density  of  management  indicator  species.  Astute 
selection  of  management  indicator  species  and  assur- 
ance of  their  subsequent  welfare  were  assumed  to 
be  an  appropriate  mechanism  to  meet  legal  require- 
ments for  maintenance  of  viable  populations  of  all 
wildlife.  Included  within  this  group  are  some  rather 
vaguely  denned  "ecological  indicators"  which,  by 
implication,  represent  a  larger  group  of  species.  We 
believe  this  approach  to  be  impractical  and  unrealis- 
tic for  the  reasons  specified  by  Verner  ( 1 984 ).  It  is, 
however,  the  law  and  the  U.S.  Forest  Service  must 
attempt  to  comply.  A  rational  approach  to  carrying 
out  such  an  effort  has  been  spelled  out  by  Verner 
( 1983).  Agencies  not  bound  by  the  National  Forest 
Management  Act  are  advised  to  use  other  approaches 
such  as  management  guilds  (discussed  later),  or  stay 
with  individual  species  monitoring  without  inference 
to  other  species. 

No  general  comprehensive  list  of  habitat  attri- 
butes for  featured  species  can  be  listed  here.  These 
attributes  are  unique  to  the  species  being  consid- 
ered. The  description  of  the  system  by  Thomas  et  al. 
( 1979c)  used  elk  and  mule  deer  as  examples  of  fea- 
tured species.  The  habitat  attributes  to  be  measured, 
and  mentioned  throughout  this  chapter  include — 

•  percentage  of  the  area  consisting  of  thermal 
cover,  hiding  cover,  and  forage  areas, 

•  stand  size, 

•  density  of  roads  open  to  vehicular  traffic, 

•  cover/forage  area  ratios. 

If,  however,  the  pileated  woodpecker  (Dryoco- 
pus  pileatus)  were  chosen  as  a  featured  species, 
other  habitat  attributes  would  become  important, 
such  as — 

•  snags  greater  than  51  -cm  (20-in.)  dbh, 

•  species  of  snags, 

•  height  of  canopy, 

•  dead-and-down  woody  material, 

•  the  presence  of  carpenter  ants  ( Camponotus 
sp.). 

Each  featured  species  will  introduce  a  different  set  of 
habitat  attributes  to  be  considered. 


The  U.S.  Department  of  Interior,  Fish  and  Wild- 
life Service  (1980)  has  developed  "Habitat  Evalua- 
tion Procedures  (HEP)"  focusing  on  habitat 
requirements  of  fish  and  wildlife  species  applicable 
to  project  planning  and  impact  evaluation  (Scham- 
berger  and  Farmer  1978).  This  was  the  impetus  for 
development  of  Habitat  Suitability  Index  (HSI) 
models  (Schamberger  et  al.  1982)  for  select  wildlife 
species.  These  models  provide  a  good  source  of  in- 
formation on  habitat  variables  that  one  should  meas- 
ure when  using  one  of  the  particular  species  as  an 
indicator  or  featured  species.  Examples  of  species  for 
which  such  habitat  models  exist  include  beaver 
{Castor  canadensis)  (Allen  1982a),  marten  (Martes 
americana)  (Allen  1982b),  and  black-capped  chicka- 
dee (Parus  atricapillus)  (Schroeder  1983). 


The  relationship  between  these  habitat  evalua- 
tion procedures  and  the  habitat  evaluation  proce- 
dures of  Thomas  (1979b),  Verner  and  Boss  ( 1980), 
and  others  is  explored  in  detail  by  Thomas  (1982). 


MONITORING 

A  wag  once  said  that  monitoring  was  an  inven- 
tory done  over  and  over.  So  it  is.  The  National  Forest 
Management  Act  of  1976,  and  regulations  issued 
pursuant  thereto,  contain  instructions  to  monitor 
indicator  species  to  determine  ( 1 )  if  management 
objectives  are  being  met,  and  (2)  if  the  management 
indicator  species  are  changing  in  status.  The  costs 
of  such  monitoring  of  individual  wildlife  species  are 
likely  to  be  very  high  in  personnel  and  dollars.  Ver- 
ner (1983)  estimated  that  for  the  western  Sierra 
Nevada  it  would  take  some  300,000  10-minute  point 
counts  to  detect  a  10%  change  in  numbers  of  such 
species  of  special  interest  such  as  the  pileated  wood- 
pecker or  willow  flycatcher  (Empidonax  traillii).  If 
20  counts  per  observer  per  day  were  made,  1 5,000 
observer  days  would  be  required.  Because  the  suita- 
ble counting  period  within  any  one  season  is  about 
60  days,  some  250  temporary  employees  capable 
of  locating  the  sample  points  and  capable  of  recog- 
nizing birds  by  sight  and  sound  would  be  needed. 
Salary  costs  alone  are  estimated  at  $825,000  per 
year.  Verner  (1983:357)  summed  up  by  saying: 

"Not  only  are  the  costs  and  personnel  needs 
for  such  an  effort  out  of  reason,  but  also  it 
is  unreasonable  to  find  such  a  large  number  of 
counting  points  because  each  must  be  indepen- 
dent of  the  others  to  satisfy  assumptions  of  the 
statistical  models.  This  fantasy  can  be  extended 
to  include  all  of  the  various  species  we  may 
be  concerned  about  and  in  all  of  the  various 
types  of  habitats  where  they  may  occur.  Ob- 
viously other  approaches  are  required." 


86 


Forests 


Note  also  that  even  after  a  change  in  the  popula- 
tion is  detected,  this  does  not  assure  information 
about  cause  and  effect.  Is  the  change  due  to  weather, 
food  supply,  habitat  change,  or  some  other  factor  or 
combination  of  factors?  Are  such  changes  indicative 
of  something  that  requires  management  action  or 
merely  normal  perturbations  in  population  numbers? 


Toward  More  Practical  Monitoring 

Verner  (1983)  suggested  various  ways  to  make 
monitoring  more  feasible  in  terms  of  costs  and  man- 
agement need.  These  are  summarized  below: 

( 1 )  Separate  species  into  categories  of  high  and  low 
risk  to  be  reduced  below  acceptable  levels  in 
terms  of  range  or  density  by  forest  management 
activities.  We  believe  that  the  vast  majority  of 
species  will  fall  in  the  low  risk  category  and  that 
habitat  monitoring  will  suffice  to  monitor  trends 
in  their  status.  High  risk  species  will  probably 
require  monitoring  of  the  species  itself. 

(2)  Design  monitoring  activities  to  detect  only  de- 
clining trends.  Only  in  rare  circumstances,  such 
as  with  ungulates  (which  are  hunted )  or  go- 
phers (which  cause  damage  to  young  trees), 
would  increases  be  a  cause  for  concern.  Besides, 
data  on  animal  damage  are  very  apt  to  be  avail- 
able from  sources  such  as  forest  surveys  for 

the  forestry  program.  The  monitoring  should  be 
designed  to  protect  against  statistical  type  II 
errors  (concluding  that  a  population  is  stable 
when  it  has  really  declined). 

Costs  vary  with  the  power  of  the  test  and  the 
magnitude  of  the  decline  to  be  detected.  We 
recommend  a  power  of  0.8  (failure  to  detect  a 
specified  decline  20%  of  the  time)  as  a  good 
compromise  between  tolerable  costs  and  man- 
agement needs.  In  all  situations,  the  sampling 
design  should  be  standardized  to  produce  as 
nearly  identical  conditions  as  possible  every 
time  the  sampling  is  done  (probably  annually). 
The  following  should  be  considered: 

•  sampling  intensity 

•  time  of  day 

•  season 

•  sample  sites 

•  weather 

•  order  of  sampling  sites 

•  variability  due  to  observers 

(3)  Select  "high  probability"  sites  for  sampling  when 
indicator  species  are  monitored.  Management 
indicator  species,  with  the  exception  of  game 


species,  are  apt  to  be  uncommon;  in  fact,  they 
may  have  been  selected  because  they  were  rare, 
threatened,  or  endangered. 

Therefore,  average  counts  of  such  species  are 
generally  so  low  in  a  randomly  selected  set  of 
sampling  points  that  costs  preclude  sampling 
enough  points  to  detect  declining  trends  with 
statistical  confidence  (Verner  1984).  A  pre-mon- 
itoring  inventory  should  be  used  to  identify 
sites  where  target  species  occur  and  monitoring 
stations  should  be  randomly  selected  from 
among  those  sites. 

(4)  Monitor  "management  guilds"  rather  than  indi- 
vidual species.  Management  guilds  are  species 
that  respond  in  a  similar  way  to  changes  in  their 
environment.  Verner  (1983:360)  said,  "The 
primary  use  of  such  management  guilds  should 
be  to  monitor  trends  in  the  suitability  of  various 
zones  of  a  habitat  to  support  wildlife  species. 

As  such,  guilds  will  probably  make  their  greatest 
contribution  to  a  monitoring  system  as  indica- 
tors of  the  quality  and  quantity  of  certain  habi- 
tats that  are  most  likely  to  be  changed  by 
management  activities." 

One  obvious  advantage  of  monitoring  manage- 
ment guilds  is  that  counts  of  a  group  of  species 
are  higher  than  counts  of  any  of  the  component 
species.  Therefore,  fewer  samples  are  required. 
Further,  the  biologist  obtains  a  more  complete 
listing  of  the  species  assemblage  than  would 
result  if  only  indicator  species  were  counted.  At 
least  for  birds,  the  cost  of  obtaining  the  more 
complete  list  is  no  greater  than  tallying  just  one 
or  a  few  species  (Verner  1983).  This  assumes,  of 
course,  that  personnel  are  available  who  can 
readily  identify  the  species  by  both  sight  and 
sound.  This  gives  managers  at  least  some  hint  of 
the  changing  status  of  individual  species. 

(5)  Monitor  only  during  a  single  season.  Ideally, 
managers  should  have  knowledge  of  the  year- 
round  status  of  habitats  and  occupying  species 
that  would  require  at  least  four  monitoring  ef- 
forts per  year.  Likely,  only  one  effort  per  year 
can  be  afforded.  For  birds,  that  effort  should 
correspond  with  the  breeding  season  (Verner 
1983).  For  game  species,  particularly  big  game, 
monitoring  is  best  done  as  close  to  hunting 
season  as  feasible  to  allow  time  for  adjustment 
in  hunting  regulations. 

For  birds,  Verner  (1983:361)  hypothesized  that 
"If  transients  and  winter  residents  use  the  same 
zones  of  a  habitat  for  feeding  and  cover  in  basi- 
cally the  same  ways  as  the  breeders  do,  mainte- 
nance of  the  populations  of  all  breeding  species 
should  ensure  maintenance  of  suitable  habitat 
for  transients  and  winter  residents." 


Forests 


87 


(6)  Monitor  trends  in  habitat.  The  assumption  in 
such  monitoring  is  that  if  enough  is  known 
about  species-habitat  relationships,  trends  in 
populations  (or  potential  populations)  could  be 
inferred  from  monitoring  habitat  trends.  These 
systems  are  probably  adequate  for  low  risk  spe- 
cies, at  least  for  the  time  being.  Nonetheless 
they  are  still  in  a  rudimentary  stage.  They  re- 
quire testing,  verification,  and  continuous 
correction. 

Ultimately,  it  may  be  possible  to  effectively 
monitor  trends  in  most  wildlife  populations  almost 
exclusively  through  habitats,  even  for  management 
indicator  species.  This  will  be  the  most  cost-effective 
method  because  trends  in  habitats  must  be  moni- 
tored for  information  on  timber  resources.  Thus,  the 
cost  of  monitoring  wildlife  trends  can  be  mainly 
absorbed  by  monitoring  timber  resources. 


An  Example  of  a  Monitoring  System 

The  Sierra  National  Forest  has  developed  a  wild- 
life monitoring  plan  that  seems  well  conceived,  ade- 
quate, realistic,  and  promising  (Verner  1983).  The 
plan  calls  for  5  years  of  inventory  to  precede  institu- 
tion of  monitoring  efforts.  This  inventory  effort  will 
provide  the  sampling  framework  and  baseline  infor- 
mation on  which  monitoring  plans  are  based.  The 
plan  is  in  three  phases: 

( 1 )  Monitoring  of  management  indicator  species. 
This  is  required  by  law,  and  the  minimum  legal 
requirement  will  be  met  by  monitoring  seven 
species,  five  of  which  are  rare,  threatened,  or 
endangered,  and  two  of  which  are  game  species. 


Monitoring  nesting  habitats  of  rare,  threatened,  or  endan- 
gered species  should  be  done  yearly. 


DISCUSSION 


(2)  Monitoring,  at  2 -year  intervals,  management 
guilds  of  birds  in  three  habitats  considered  espe- 
cially vulnerable  to  alteration  by  management. 
These  habitats  include  late-successional  mixed- 
conifer  forests,  riparian  habitats,  and  meadow 
edges.  Each  would  include  200  sampling  points. 
It  is  hypothesized  that  stability  in  bird  assem- 
blages indicates  similar  stability  in  assemblages 
of  other  vertebrate  groups. 

(3)  Monitoring  of  most  wildlife  species  in  most 
habitats  by  monitoring  trends  in  habitat  defined 
by  plant  community,  successional  stage,  and 
special  and  unique  habitat  features  such  as  snags 
and  dead-and-down  woody  material.  Such  an 
approach  requires  the  development  of  data 
bases  of  site-specific  wildlife-habitat  relationships 
such  as  those  described  by  Patton  (  1978), 
Thomas  (1979b),  or  Verner  and  Boss  ( 1980). 


The  inventorying  and  monitoring  of  forest  wild- 
life habitats  and  forest  wildlife  (except  for  game 
species)  are  in  their  infancies.  The  literature  reveals 
more  speculation  than  action  at  this  point.  In  fact, 
most  of  what  has  been  discussed  in  this  chapter  is 
based  on  hypothesized  wildlife-habitat  relationships 
formulated  from  the  synthesis  of  ecological  theory 
and  some  basic  data  on  how  different  species  of 
wildlife  react  to  their  habitats.  The  result  has  been 
called  "a  working  hypothesis,  a  place  to  start,  and  a 
way  to  derive  tentative  answers  to  questions  for 
which  there  are  no  certain  answers"  (Thomas 
1979b:7).  But  we  do  have  a  place  to  start.  We  are 
confident  that,  as  speculation  gives  way  to  action,  ex- 
perience will  allow  improvements  in  the  currently 
primitive  systems,  and  the  art  of  monitoring  will 
evolve  to  a  science.  An  overriding  consideration  in 
all  of  this  must  be  its  cost,  because  if  economical 
ways  cannot  be  found  to  do  the  job,  it  will  not  get 
done. 


88 


Forests 


LITERATURE  CITED 


ALLEN,  AW.  1982a.  Habitat  suitability  index  models: 

beaver.  FWS/OBS-82/ 10.30.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  Washington,  DC.  20pp. 

.  1982b.  Habitat  suitability  index  models:  marten. 

FWS/OBS-82/ 10. 11.  U.S.  Dep.  Inter.,  Fish  and  Wildl. 
Serv.  Washington,  DC.  9pp. 

ANDERSON,  A.J.B.  1971.  Ordination  methods  in  ecology. 
J.  Ecol.  59:713-726. 

ANDERSON,  JR.,  E.E.  HARDY,  J.T.  ROACH,  and  RE.  WIT- 
MER.  1976.  A  land  use  and  land  cover  classification 
system  for  use  with  remote  sensor  data.  U.S.  Dep. 
Inter.,  Geol.  Surv.  Prof.  Pap.  964.  Reston,  VA.  28pp. 

AUSTIN,  M.E.  1965.  Land  resource  regions  and  major  land 
resource  areas  of  the  United  States  (exclusive  of 
Alaska  and  Hawaii).  USDA  Soil  Conserv.  Serv.  Agric. 
Handb.  296.  Washington,  DC.  82pp. 

AUSTIN,  MP.  and  L.  ORLOCI.  1966.  Geometric  models  in 
ecology.  II.  An  evaluation  of  some  ordination  tech- 
niques. J.  Ecol.  54:217-227. 

BAILEY,  R.G.  1976.  Ecoregions  of  the  United  States  (map). 
U.S.  Dep.  Agric.  For.  Serv.,  Intermountain  Region. 
Ogden,  UT. 

.  1 977.  A  new  map  of  the  ecosystem  regions  of  the 

United  States.  Pages  1 21- 1 28  in  Classification,  Inven- 
tory, and  Analysis  of  Fish  and  Wildlife  Habitat — the 
Proceedings  of  a  National  Symposium.  FWS/OBS- 
78/76.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  Washing- 
ton, DC. 

.  1978.  Description  of  the  ecoregions  of  the  United 

States.  U.S.  Dep.  Agric,  Forest  Service,  Intermountain 
Region.  Ogden,  UT.  77pp. 

.  1980.  Integrated  approaches  to  classifying  land  as 

ecosystems,  in  Proc.  IUFRO/ISS  workshop  on  land 
evaluation  for  forestry.  Wageningen,  The  Netherlands. 
1982.  Classification  systems  for  habitat  and  ecosys- 


tems. Pages  16-26  in  Mason,  W.T.  ed.  Research  on 
Fish  and  Wildlife  Habitat.  Office  of  Research  and 
Development,  U.S.  Environmental  Protection  Agency. 
Washington,  DC. 

BANNISTER,  P.  1968.  An  evaluation  of  some  procedures 
used  in  simple  ordinations.  J.  Ecol.  56:27-34. 

BLACK,  H.  Jr.,  R  SCHERZINGER,  and  J.W.  THOMAS.  1976. 
Relationships  of  Rocky  Mountain  elk  and  Rocky 
Mountain  mule  deer  habitat  to  timber  harvest  in  the 
Blue  Mountains  of  Oregon  and  Washington.  Pages 
105-118  in  Proceedings  of  the  Elk-Logging-Roads 
Symposium.  Univ.  Idaho,  Moscow. 

BRAY,  JR.  and  J.T.  CURTIS.  1957.  An  ordination  of  the 
upland  forest  communities  of  southern  Wisconsin. 
Ecol.  Monogr.  27:325-349. 

BROWN,  D.E.,  C.H.  LOWE,  and  C.P.  PASE.  1980.  A  digi- 
tized systematic  classification  for  ecosystems  with  an 
illustrated  summary  of  the  natural  vegetation  of  North 
America.  U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep. 
RM-73.  93pp. 

BROWN,  H.R.,  ed.  In  Press.  Wildlife  habitats  in  managed 
forests — western  Oregon  and  Washington.  U.S.  Dep. 
Agric,  For.  Serv.,  Pacific  Northwest  Region.  Portland, 
OR. 

BUTTERY,  R.F.  1978.  Modified  ECOCLASS— A  Forest  Ser- 
vice method  for  classifying  ecosystems.  Pages  157- 
168  in  Integrated  Inventories  of  Renewable  Natural 
Resources.  Tucson,  AZ.  U.S.  Dep.  Agric,  For.  Serv. 
Gen.  Tech.  Rep.  RM-55. 


CAIN,  S.A.  and  G.M.  DE  OLIVERIA  CASTRO.  1959.  Manual 
of  vegetation  analysis.  Harper  and  Brothers.  New 
York,  NY.  325pp. 

CORTISS,  J.C.  1974.  ECOCLASS — a  method  for  classifying 
ecosystems.  Pages  264-271  in  Foresters  in  Land-Use 
Planning.  Proc.  of  the  1973  Natl  Conv.,  Soc  Am.  For. 
Washington,  DC. 

CROME,  F.H.J.  1978.  Foraging  ecology  of  an  assemblage  of 
birds  in  lowland  rainforest  in  northern  Queensland. 
Aust.  J.  Ecol.  3:195-212. 

CROWLEY,  J.M.  1967.  Biogeography.  Can.  Geogr.  11:312- 
326. 

CUSHWA,  C.T.  and  C.W.  DUBROCK.  1982.  Design  of  com- 
puterized fish  and  wildlife  species  data  bases  by  state 
and  federal  agencies.  Pages  37-46  in  Mason,  W.T. 
ed.  Research  on  Wildlife  Habitat.  EPA-600/8-82-022. 
U.S.  Environmental  Protection  Agency.  Washington, 
DC. 

DAUBENMIRE,  R.  1968.  Plant  communities:  a  textbook  of 
plant  synecology.  Harper  and  Row.  New  York,  NY. 
300pp. 

DEGRAAF,  R.M.  and  N.L.  CHADWICK.  1984.  Habitat  clas- 
sification: a  comparison  using  avian  species  and 
guilds.  Environ.  Manage.  8(6):51 1-518. 

DRISCOLL,  R.S.,  J.W.  RUSSELL,  and  M.C.  MEIER.  1978. 
Recommended  national  land  classification  system  for 
renewable  resource  assessments.  U.S.  Dep.  Agric,  For. 
Serv.  Rocky  Mtn.  For.  and  Range  Exp.  Sta.  Fort  Col- 
lins, CO  (unpublished  manuscript).  44pp. 

DUESER,  R.D.  and  H.H.  SHUGART,  Jr.  1979.  Niche  pattern 
in  a  forest -floor  small-mammal  fauna.  Ecology  60:108- 
118. 

ELTON,  C.S.  1966.  Dying  and  dead  wood.  Pages  279-305 
in  The  Pattern  of  Animal  Communities.  John  Wiley 
and  Sons,  Inc.  New  York,  NY. 

FOLSE,  L.J.  Jr.  1981.  Ecological  relationships  of  grassland 
birds  to  habitat  and  food  supply  in  east  Africa.  Pages 
160-166  in  Capen,  D.E.  ed.  The  Use  of  Multivariate 
Statistics  in  Studies  of  Wildlife  Habitat.  U.S.  Dep. 
Agric,  For.  Serv.  Gen.  Tech.  Rep.  RM-87. 

GALLI,  A.E.,  C.F.  LECK,  and  R.T.T.  FORMAN.  1976.  Avian 
distribution  patterns  in  forest  islands  of  different  sizes 
in  central  New  Jersey.  Auk  93(2  ):356- 364. 

GARRISON,  G.A.,  J.J.  BJUGSTAD,  DA.  DUNCAN,  M.E. 
LEWIS,  and  DR.  SMITH.  1977.  Vegetation  and  envi- 
ronmental features  of  forest  and  range  ecosystems. 
U.S.  Dep.  Agric,  For.  Serv.  Agric  Handb.  475.  U.S. 
Govt.  Print.  Off.  Washington,  DC.  68pp. 

GAUCH,  H.G.  Jr.  1982.  Multivariate  analysis  in  community 
ecology.  Cambridge  Univ.  Press,  Cambridge.  298pp. 

GODFREY,  A.E.  1977.  A  physiographic  approach  to  land 
use  planning.  Environ.  Geol.  2:43-50. 

GREIG-SMITH,  P.  1964.  Quantitative  plant  ecology.  2nd 
ed.  Butterworths,  London.  256pp. 

HAAPANEN,  A.  1965.  Bird  fauna  of  the  Finnish  forests  in 
relation  to  forest  succession.  I.  Ann.  Zool.  Fenn. 
2(3):153-196. 

.  1 966.  Bird  fauna  of  the  Finnish  forests  in  relation 

to  forest  succession.  II.  Ann.  Zool.  Fenn.  3(3):  176- 
200. 

HALL,  F.C.  1973-  Plant  communities  of  the  Blue  Mountains 
in  eastern  Oregon  and  southeastern  Washington.  U.S. 
Dep.  Agric,  For.  Serv.  Reg.  6  Area  Guide  3-1.  62pp. 

HARPER,  JAMES  A.  1969.  Relationship  of  elk  to  reforesta- 
tion in  the  Pacific  Northwest.  Pages  67-71  in  Black, 
H.C.  ed.  Wildlife  and  Reforestation  in  the  Pacific 
Northwest.  School  of  Forestry,  Oreg.  State  Univ. 
Corvallis. 


Forests 


89 


HENDERSON,  J.A.,  L.S.  DAVIS,  and  EM.  RYBERG.  1978. 
ECOSYM:  a  classification  and  information  system  for 
wildland  resource  management.  Utah  State  Univ. 
Logan.  30pp. 

HOEKSTRA,  T.W.,  D.E.  CHALK,  C.L.  HAWKES,  and  S.A. 
MILLER.  1983-  Monitoring  regional  wildlife  and  fish 
habitats  and  populations  for  national  assessments  and 
appraisals.  Pages  308-314  in  Trans.  48th  North  Am. 
Wildl.  and  Nat.  Resour.  Conf. 

,  D.L  SCHWEITZER,  C.T.  CUSHWA,  S.H.  ANDER- 
SON, and  R.B.  BARNES.  1979.  Preliminary  evaluation 
of  a  national  wildlife  and  fish  data  base.  Pages  308- 
314  in  Trans.  44th  North  Am.  Wildl.  and  Nat.  Resour. 
Conf. 

HOLMES,  R.T.,  RE.  BONNEY,  Jr.,  and  S.W.  PACALA.  1979. 
Guild  structure  of  the  Hubbard  Brook  bird  commu- 
nity: a  multivariate  approach.  Ecology  60:512-520. 

JAKSIC,  F.M.  1981.  Abuse  and  misuse  of  the  term  "guild" 
in  ecological  studies.  Oikos  37:397-400. 

JENKINS,  R.E.  1977.  Classification  and  inventory  for  the 
perpetuation  of  ecological  diversity.  Pages  41-51 
in  Classification,  Inventory,  and  Analysis  of  Fish  and 
Wildlife — the  Proceedings  of  a  National  Symposium, 
U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  Washington,  DC. 

KUCHLER,  AW.  1964.  Potential  natural  vegetation  of  the 
conterminous  United  States  (map  and  manual).  Am. 
Geogr.  Spec.  Publ.  36.  11 6pp. 

.  1970.  Potential  natural  vegetation.  (Map,  1966.) 

Pages  90-91  in  The  National  Atlas  of  the  United  States 
of  America.  U.S.  Dep.  Inter.,  Geological  Survey.  Wash- 
ington, DC. 

LANDRES,  P.B.  and  J.A.  MACMAHON.  1980.  Guilds  and 
community  organization:  analysis  of  an  oak  woodland 
avifauna  in  Sonora,  Mexico.  Auk  97:351-365. 

LARSON,  T.A.  1981.  Ecological  correlates  of  avian  commu- 
nity structure  in  mixed-conifer  habitat:  an  experimen- 
tal approach.  Ph.D.  Dissertation.  Illinois  State  Univ. 
Normal. 

LECKENBY,  D.A.  1984.  Elk  use  and  availability  of  cover 
and  forage  habitat  components  in  the  Blue  Mountains, 
Northeast  Oregon,  1976-1982.  Wildl.  Res.  Rep.  14. 
Oregon  Dep.  of  Fish  and  Wildl.  Portland.  40pp. 

LEOPOLD,  A.  1933.  Game  management.  Charles  Scribner 
Sons.  New  York,  NY.  481pp. 

LINDEMAN,  R.H,  P.F.  MERENDA,  and  R.F.  GOLD.  1980. 
Introduction  to  bivariate  and  multivariate  analysis. 
Scott,  Foresman  and  Co.  Glenview,  IL  444pp. 

MACARTHUR,  R.H.  1964.  Environmental  factors  affecting 
bird  species  diversity.  Am.  Nat.  98:387-397. 

MACMAHON,  J.A,  D.J.  SCHIMPF,  DC.  ANDERSON,  KG. 
SMITH,  and  R.L.  BAYNJr.  1981.  An  organism-centered 
approach  to  some  community  and  ecosystem  con- 
cepts. J.  Theor.  Biol.  88:287-307. 

MARMELSTEIN,  A.  1977.  Classification,  inventory,  and 
analysis  of  fish  and  wildlife  habitat — proceedings  of  a 
national  symposium.  U.S.  Dep.  Inter.,  Fish  and  Wildl. 
Serv.  Washington,  DC.  604pp. 

MASER,  C,  J.M.  GEIST,  DM.  CONCANNON,  R.  ANDER 
SON,  and  B.  LOVELL  1979a.  Wildlife  habitats  in  man- 
aged rangelands — the  Great  Basin  of  southeastern 
Oregon:  geomorphic  and  edaphic  habitats.  U.S.  Dep. 
Agric,  For.  Serv.  Gen.  Tech.  Rep.  PNW-99.  84pp. 

,  J.E.  RODIEK  and  J.W.  THOMAS.  1979b.  Cliffs,  talus, 

and  caves.  Pages  96-103  in  Wildlife  Habitats  in  Man- 
aged Forests — the  Blue  Mountains  of  Oregon  and 
Washington.  U.S.  Dep.  Agric,  Agric.  Handb.  553-  U.S. 
Govt.  Print.  Off.  Washington,  DC. 


— ,  R.G.  ANDERSON,  K  CROMACK  Jr.,  J.T.  WILLIAMS, 
and  RE.  MARTIN.  1979c.  Dead  and  down  woody 
material.  Pages  78-95  in  Wildlife  Habitats  in  Managed 
Forests — the  Blue  Mountains  of  Oregon  and  Washing- 
ton. U.S.  Dep.  Agric,  Agric  Handb.  553-  U.S.  Govt. 
Print.  Off.  Washington,  DC. 

—  and  J.W.  THOMAS.  1983.  Wildlife  habitats  in  man- 


aged rangelands — the  Great  Basin  of  southeastern 
Oregon:  Introduction.  U.S.  Dep.  Agric,  For.  Serv.  Gen. 
Tech.  Rep.  PNW-160.  15pp. 

MAWSON,  J.C.,  J.W.  THOMAS,  and  R.M.  DEGRAAF.  1976. 
Program  HTVOL — the  determination  of  crown  vol- 
ume by  layers.  U.S.  Dep.  Agric,  For.  Serv.  Res.  Pap. 
NE-354.  9pp. 

NIE,  N.H.,  C.H.  HULL,  J.G.  JENKINS,  K  STEINBRENNER, 
and  D.H.  BENT.  1974.  SPSS  statistical  package  for  the 
social  sciences.  2nd  ed.  McGraw-Hill.  New  York, 
NY.  675pp. 

PATTON,  DR.  1975.  A  diversity  index  for  quantifying 
habitat  "edge."  Wildl.  Soc  Bull.  3(4):  171- 173- 

.  1978.  RUNWILD:  a  storage  and  retrieval  system  for 

wildlife  habitat  information.  U.S.  Dep.  Agric,  For. 
Serv.  Gen.  Tech.  Rep.  RM-51.  8pp. 

PFISTER,  R.D.  1977.  Ecological  classification  of  forest  land 
in  Idaho  and  Montana.  Pages  329-358  in  Ecological 
Classification  of  Forest  Land  in  Canada  and  North- 
western USA.  Univ.  British  Columbia.  Vancouver, 
Canada. 

RAO,  C.R.  1966.  Influence  on  discriminant  function  coeffi- 
cients. Pages  587-602  in  R.C.  Rose,  I.M.  Chakrovarti, 
P.C.  Mahalanobis,  C.R.  Rao,  and  KJ.C  Smith  eds.  Es- 
says in  Probability  and  Statistics.  Univ.  North  Carolina 
Press,  Chapel  Hill. 

REYNOLDS,  H.G.  1966.  Use  of  openings  in  spruce-fir  for- 
ests of  Arizona  by  deer,  elk  and  cattle.  U.S.  Dep. 
Agric,  For.  Serv.  Res.  Note  78,  4pp. 

ROBINSON,  W.L  and  E.B.  BOLEN.  1984.  Wildlife  ecology 
and  management.  MacMillan  Publ.  Co.  New  York, 
NY.  478  pp. 

ROOT,  R.B.  1967.  The  niche  exploitation  pattern  of  the 
blue-gray  gnatcatcher.  Ecol.  Monogr.  37:317-350. 

SALWASSER,  HAL,  C.K  HAMILTON,  W.B.  KROHN,  J.  LIP- 
SCOMB, and  C.H.  THOMAS.  1983.  Monitoring  wildlife 
and  fish:  mandates  and  their  implications.  Pages  297- 
306  in  Trans.  48th  North  Am.  Wildl.  and  Nat.  Resour. 
Conf. 

SCHAMBERGER,  M.  and  A.  FARMER.  1978.  The  habitat 
evaluation  procedures:  their  application  in  project 
planning  and  impact  evaluation.  Pages  274-283  in 
Trans.  43rd  North  Am.  Wildl.  and  Nat.  Resour.  Conf. 

, ,  and  J.W.  TERRELL.  1982.  Habitat  suitability 

index  models:  introduction.  FWS/OBS-82/10.  U.S. 
Dep.  Inter.,  Fish  and  Wildl.  Serv.  Washington,  DC. 
2pp. 

SCHROEDER,  R.L  1983.  Habitat  suitability  index  models: 
black-capped  chickadee.  FWS/OBS-82/10. 37.  U.S.  Dep. 
Inter.,  Fish  and  Wildl.  Serv.  Washington,  DC.  1 2pp. 

SEVERINGHAUS,  WD.  1981.  Guild  theory  development 
as  a  mechanism  for  assessing  environmental  impact. 
Environ.  Manage.  5:187-190. 

SHORT,  H.L  1983.  Wildlife  guilds  in  Arizona  desert  habi- 
tats. U.S.  Dep.  Inter.,  Bur.  Land  Manage.  Serv.  Cen. 
Tech.  Note  362.  258pp. 

and  KP.  BURNHAM.  1982.  Technique  for  structur- 
ing wildlife  guilds  to  evaluate  impacts  on  wildlife 
communities.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv. 
Spec.  Sci.  Rep.  244.  Washington,  DC.  34pp. 


90 


Forests 


SOCIETY  OF  AMERICAN  FORESTERS.  1954.  Forest 
cover  types  of  North  America.  Society  of  American 
Foresters.  Washington,  DC.  67pp. 

THOMAS,  J. W.  1979a.  Introduction.  Pages  10-21  in  Wild- 
life Habitats  in  Managed  Forests — the  Blue  Mountains 
of  Oregon  and  Washington.  U.S.  Dep.  Agric,  Agric. 
Handb.  553-  U.S.  Govt.  Print.  Off.  Washington,  DC. 

ed.  1979b.  Wildlife  habitats  in  managed  forests — 

the  Blue  Mountains  of  Oregon  and  Washington.  U.S. 
Dep.  Agric,  Agric.  Handb.  553-  U.S.  Govt.  Print.  Off. 
Washington,  DC.  510pp. 

.  1982.  Species/habitat  relationships — a  key  to  con- 
sidering wildlife  in  planning  and  land  management 
decisions.  Pages  27-36  in  Mason,  W.T.  ed.  Research 
on  Wildlife  Habitat.  EPA-600/8-82-022.  U.S.  Environ- 
mental Protection  Agency.  Washington,  DC. 

,  R.M  DEGRAAF,  and  J.C  MAWSON.  1977.  Determi- 
nation of  habitat  requirements  for  birds  in  suburban 
areas.  U.S.  Dep.  Agric,  For.  Serv.  Res.  Pap.  NE-357. 
15pp. 

,  R.J.  MILLER,  C.  MASER,  R.G.  ANDERSON,  and  B.E. 

CARTER.  1979a.  Plant  communities  and  successional 
stages.  Pages  22-39  in  Wildlife  Habitats  in  Managed 
Forests — the  Blue  Mountains  of  Oregon  and  Washing- 
ton. U.S.  Dep.  Agric,  Agric.  Handb.  553.  U.S.  Govt. 
Print.  Off.  Washington,  DC. 

,  R.G.  ANDERSON,  C  MASER,  and  EL.  BULL.  1979b. 

Snags.  Pages  60-77  in  Wildlife  Habitats  in  Managed 
Forests — the  Blue  Mountains  of  Oregon  and  Washing- 
ton. U.S.  Dep.  Agric,  Agric.  Handb.  553-  U.S.  Govt. 
Print.  Off.  Washington,  DC. 

,  H.  BLACK  Jr.,  R.J.  SCHERZINGER,  and  R.J.  PEDER 

SEN.  1979c  Deer  and  elk.  Pages  104-127  in  Wildlife 
Habitats  in  Managed  Forests — the  Blue  Mountains 
of  Oregon  and  Washington.  U.S.  Dep.  Agric,  Agric. 
Handb.  553.  U.S.  Govt.  Print.  Off.  Washington,  DC. 

,  C.  MASER,  and  J.  RODIEK.  1979d.  Edges.  Pages  48- 

59  in  Wildlife  Habitats  in  Managed  Forests — the  Blue 
Mountains  of  Oregon  and  Washington.  U.S.  Dep. 


Agric,  Agric  Handb.  553.  U.S.  Govt.  Print.  Off.  Wash- 
ington, DC. 

, ,  and .  1979c  Riparian  zones.  Pages 

40-47  in  Wildlife  Habitats  in  Managed  Forests — the 
Blue  Mountains  of  Oregon  and  Washington.  U.S.  Dep. 
Agric,  Agric  Handb.  553.  U.S.  Govt.  Print.  Off.  Wash- 
ington, DC. 

U.S.  DEPARTMENT  OF  INTERIOR,  BUREAU  OF  LAND 
MANAGEMENT.  1978.  Integrated  habitat  inventory 
and  classification  system,  BLM  Manual  Section  6602. 
Washington,  DC.  37pp. 

U.S.  DEPARTMENT  OF  INTERIOR,  FISH  AND  WILD- 
LIFE SERVICE.  1980.  Habitat  evaluation  procedures 
(HEP).  ESM  121.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv. 
Washington,  DC. 

VERNER,  J.  1983-  An  integrated  system  for  monitoring 
wildlife  on  the  Sierra  National  Forest.  Pages  355-366. 
in  Trans.  48th  North  Am.  Wildlife,  and  Nat.  Resour. 
Conf. 

.  1984.  The  guild  concept  applied  to  management 

of  bird  populations.  Environ.  Manage.  8(  1):1-14. 

and  A.S.  BOSS.  1980.  California  wildlife  and  their 

habitats:  western  Sierra  Nevada.  U.S.  Dep.  Agric,  For. 
Serv.  Gen.  Tech.  Rep.  PSW-37.  439pp. 

WERTZ,  W.A.  and  J.F.  ARNOLD.  1972.  Land  systems  inven- 
tory. U.S.  Dep.  Agric,  For.  Serv.,  Intermountain  Re- 
gion. Ogden,  UT.  12pp. 

WHITCOMB,  R.F.,  C.S.  ROBBINS,  J.F  LYNCH,  B.L.  WHIT- 
COMB,  M.K.  KLIMBIEWICZ,  and  D.  BYSTRAK  1981. 
Effects  of  forest  fragmentation  on  avifauna  of  the 
eastern  deciduous  forest.  Pages  125-205  in  Burgess, 
R.L.  and  DM.  Sharpe  eds.  Forest  Island  Dynamics 
in  Man-Dominated  Landscapes.  Springer-Verlag.  New 
York,  NY. 

WILLIAMS,  B.K.  1980.  Discriminant  analysis  in  wildlife 
research:  theory  and  applications.  Pages  59-71  in 
Capen,  D.E.  ed.  The  Use  of  Multivariate  Statistics  in 
Studies  of  Wildlife  Habitat.  U.S.  Dep.  Agric,  For.  Serv. 
Tech.  Rep.  RM-87. 


Forests 


91 


RANGELANDS 


Henry  L.  Short 

U.S.  Fish  and  Wildlife  Service 
Western  Energy  and  land  Use  Team 
Ft.  Collins,  CO  80526 


Editor's  Note:  Most  range  studies  have  focused  on 
problems  of  livestock  management.  Most  wildlife- 
oriented  efforts  have  concentrated  on  alleviating 
impacts  from  grazing  and  have  focused  on  prob- 
lems such  as  forage  allocation  between  livestock 
and  wild  ungulates. 

This  chapter  provides  techniques  to  inventory  and 
monitor  habitat  structure.  Because  structure  (trees, 
shrubs,  etc.)  on  range  lands  is  very  limited,  it  is 
extremely  important  to  wildlife.  Areas  with  struc- 
ture are  where  biologists  should  concentrate  their 
efforts.  Systems  based  on  structural  diversity  will  be 
the  most  useful  on  rangelands. 


"As  demands  have  grown  for  the  products  of  range- 
lands,  it  has  become  obvious  that  such  cliches  as 
'good  range  management  is  good  wildlife  manage- 
ment' will  no  longer  suffice." 

— Chris  Maser  and  Jack  W.  Thomas  from  Wild- 
life Habitats  in  Managed  Rangelands — The 
Great  Basin  of  Southeastern  Oregon 


INTRODUCTION 

Rangelands,  as  discussed  in  this  chapter,  are 
considered  to  be  lands — 

•  where  the  potential  natural  vegetation  is  mostly 
grass,  grasslike  plants,  forbs,  or  shrubs; 

•  where  grazing  and  browsing  were  important 
influences  during  prehistoric  times;  and 

•  that  are  more  suitable  for  management  by  eco- 
logical than  by  agronomic  principles  (after 
Schwarz  et  al.  1976). 

Western  range  and  Great  Plains  ecosystems  total 
over  258  million  ha  (646  million  a.)  (Table  1)  in  17 
western  and  Great  Plains  States.  Eighty-six  percent 
is  grazed  by  domestic  livestock.  More  than  1 36  mil- 
lion kg  (  300  million  t )  of  herbage  and  browse  are 
produced  annually  on  these  rangelands,  providing 
nearly  149  million  Animal  Unit  Months  (AUMs)  of 
domestic  livestock  use  (Table  1 ).  The  approximate 
location  of  the  potential  natural  vegetation  types  that 
comprise  rangelands  is  illustrated  in  Figure  1. 

About  65%  of  the  rangeland  area  listed  in  Table 
1  occurs  in  1 1  western  states  where  shrublands, 
mountain  grassland,  and  desert  grasslands  predomi- 
nate. These  167  million  ha  (418  million  a.)  provide 
only  about  48%  of  the  herbage  and  browse  and  38% 
of  the  AUMs  produced  in  the  western  range  and 
Great  Plains  ecosystems  (Table  1).  The  Bureau  of 
Land  Management  administers  about  68.4  million  ha 
(171  million  a.)  within  11  western  states  and  author- 
izes the  use  of  about  1 2  million  AUMs  of  forage  for 
3.5  million  cattle  and  4.9  million  sheep  (U.S.  Depart- 
ment of  Interior,  Bureau  of  Land  Management  1975). 

Rangelands  are  vast  areas  that  provide  great 
diversity  and  present  significant  problems  to  man- 
agers. Many  of  the  western  rangelands  are  fragile,  oc- 
curring on  poorly  developed,  shallow,  and  highly 
alkaline  soils  in  arid  or  semiarid  climates  with  a 


Rangelands 


93 


Table  1.     Production  of  forage  and  Animal  Unit  Months  (AUMs)  in  1970  by  ecosystem  type  (after  U.S.  Forest 
Service  1972). 


Herbage  and 

browse  produc- 

Animal Unit 

Total  land  area 

tion  by  ecosys- 

Month production 

by  ecosystem 

tem  (thousands 

Total  land  area 

by  ecosystem 

(thousands 

of  tons  per 

grazed  (thou- 

(thousands of 

Ecosystems 

of  acres) 

year) 

sands  of  acres) 

AUMs) 

Sagebrush 

94,219 

35,840 

90,453 

10,850 

Desert  shrub 

86,043 

9,491 

58,762 

1,742 

Southwestern  shrub-steppe 

38,601 

3,526 

35,382 

1,958 

Chaparral-mountain  shrub 

32,801 

11,559 

30,118 

1,957 

Pinyon-juniper 

42,677 

4,934 

34,163 

1,715 

Mountain  grasslands 

79,839 

60,858 

75,206 

21,441 

Mountain  meadows 

4,045 

5,482 

3,785 

4,309 

Desert  grasslands 

26,098 

6,481 

26,098 

5,073 

Annual  grasslands 

6,700 

6,886 

6,696 

7,003 

Alpine 

8,322 

97 

162 

33 

Shinnery 

2,004 

927 

1,982 

456 

Texas  savanna 

15,221 

8,352 

14,870 

5,042 

Plains  grasslands 

173,260 

91,408 

163,709 

50,454 

Prairie 

37,533 

54,252 

36,539 

36,814 

Total 

646,643 

300,092 

557,925 

148,848 

short  growing  season.  Only  1 8%  of  the  western 
range  acreages  are  considered  in  good  condition, 
whereas  50%  are  rated  fair  and  32%  are  rated  poor 
(U.S.  Forest  Service  1972).  The  U.S.  Bureau  of  Land 
Management  considers  1 7%  of  its  land  in  satisfactory 
or  better  condition;  83%  produces  vegetation  at  a 
rate  less  than  its  potential  (U.S.  Bureau  of  Land  Man- 
agement 1975).  Forty-eight  percent  of  the  ranges 
in  the  more  mesic  Great  Plains  are  considered  in 
good  condition,  45%  in  fair  condition,  and  only  7% 
in  poor  condition  (U.S.  Forest  Service  1972).  Range- 
lands  in  the  1 1  western  states  provide  a  variety  of 
products  in  addition  to  forage.  Rangeland  surveys 
evaluate — 


•  the  current  year's  production  of  herbage  and 
browse; 

•  the  yield  of  water  in  acre -feet  per  acre  per  year; 

•  the  quantity  of  sediments  deposited  in  stream 
channels  per  year; 

•  changes  in  the  fertility,  structure,  or  drainage  of 
soils; 

•  carbon  and  particulate  content  in  the  atmos- 
phere, as  measures  of  air  quality; 


•  aesthetic  quality; 

•  changes  in  populations  of  rare  and  endangered 
wildlife  species; 

•  diversity  of  habitat  for  nongame  birds; 

•  resident  big  game  populations; 

•  outdoor  recreation  use;  and 

•  cultural  values. 

Results  of  these  surveys  indicate  that  western 
rangelands  have  poor  soil  stability;  poor  to  fair  non- 
hunting  recreation  opportunities;  fair  soil  quality;  fair 
to  good  air  quality;  fair  to  good  habitat  for  carni- 
vores, raptors,  and  rare  and  endangered  species  of 
wildlife  and  nongame  birds;  and  fair  to  good  hunting 
and  scenic  and  cultural  values  (U.S.  Forest  Service 
1972). 

Rangeland  habitats  are  immense  and  complex. 
Some  sort  of  common  denominator  for  rangelands 
seems  necessary  so  that  range  surveys  can  provide 
results  that  are  comparable  and  interpretable  be- 
tween areas.  The  concept  of  layers  of  habitat  appears 
to  be  a  useful  denominator  for  comparing 
rangelands. 


94 


Rangelands 


Figure  1.     Occurrence  of  potential  natural  vegetation  types  that  comprise  rangeland  habitats  in  the  western 
U.S.  (after  Kuchler  1964). 


Rangelands 


95 


The  processes  described  in  this  chapter  are 
based  on  the  assumption  that  the  niches  of  individ- 
ual wildlife  species  can  be  positioned  within  the 
vertical  structure  of  habitats.  This  vertical  structure 
can  be  represented  in  terms  of  habitat  layers.  The 
abstraction  of  representing  cover  types  as  habitat  lay- 
ers provides  a  means  for  performing  a  variety  of 
quantitative  assessments  of  habitat.  For  example,  the 
structure  of  several  land  units  can  be  compared  by 
measuring  the  areas  of  the  different  layers  of  habitat 
present  on  those  land  units.  Monitoring  the  area  of 
habitat  layers  will  describe  the  general  direction 
of  change  over  time  in  the  quantity  of  habitat  avail- 
able for  the  wildlife  community. 

The  layers  of  habitat  present  on  a  land  unit  can 
be  mapped  and  measured.  Wildlife  species  can  be 
associated  with  particular  layers  of  habitat  through 
application  of  the  species-habitat  matrix,  where 
groups  or  guilds  of  wildlife  species  that  share  de- 
pendencies on  the  same  general  structure  of  habitat 
are  formed.  Maps  of  layers  of  habitat  and  the  guilds 
associated  with  those  layers  of  habitat  are  distribu- 
tion maps  for  wildlife  guilds  and  species.  The  areas 
in  the  distribution  maps  can  be  monitored  over  time 
to  describe  trends  in  the  quantity  of  habitat  available 
for  particular  wildlife  guilds  and  their  component 
wildlife  species.  Land-use  changes  often  can  be  con- 
sidered in  terms  of  impact  to  layers  of  habitat.  The 
impacts  of  those  land  use  changes  on  wildlife  species 
can  be  predicted  because  wildlife  species  have  been 
associated  with  habitat  layers  through  the  wildlife 
guilding  process. 

A  layer  of  habitat  can  be  thought  of  as  providing 
a  large  array  of  habitat  conditions.  Guilds  of  wildlife 
species  that  have  a  dependency  on  a  habitat  layer 
can  be  disaggregated  to  describe  the  dependency  of 
individual  species  on  specific  conditions  along  this 
array  or  habitat  gradient.  This  disaggregation  is  done 
with  species-habitat  models  (or  HSI  models)  devel- 
oped for  individual  wildlife  species.  These  models 
emphasize  features  that  seem  important  to  individual 
wildlife  species.  These  features  can  be  measured 
during  inventories  to  predict  whether  or  not  a  habi- 
tat apparently  is  suitable  for  an  individual  species. 

The  absence  of  a  wildlife  species  in  a  planning 
unit  during  a  wildlife  inventory  frequently  can  be 
attributed  to  either  the  absence  of  a  habitat  layer  or 
inadequate  habitat  conditions  within  that  layer  of 
habitat.  It  is  possible,  in  this  way,  to  predict  why  a 
species  is  present  or  absent  within  a  habitat  and, 
presumably,  to  develop  remedial  management  prac- 
tices to  make  that  habitat  more  suitable. 


CLASSIFICATION  SYSTEMS 

The  vegetation  in  western  rangelands  has  been 
extensively  studied.  Sagebrush  and  pinyon-juniper 


ecosystems  make  up  about  one-third  of  the  area  of 
western  rangelands  (Table  1 );  each  has  been  the 
subject  of  recent  technical  symposia  (Utah  State  Uni- 
versity 1975,  1979).  These  symposia  emphasized 
the  distribution  of  these  major  vegetational  cover 
types  in  the  western  U.S.,  results  of  autoecological 
and  synecological  studies  of  the  ecosystems,  faunal 
communities,  processes  (e.g.,  fire,  mechanical,  and 
herbicidal  techniques)  for  modifying  the  structure  of 
the  ecosystems,  impacts  of  ecosystem  modification 
on  other  land  uses,  and  procedures  for  increasing 
herbage  production  in  modified  habitats.  The  distri- 
bution of  shrubs  within  the  U.S.,  uses  of  shrub  tis- 
sues, physiology  and  nutritive  quality  of  shrubs, 
techniques  for  regenerating  shrubs,  and  the  future  of 
shrubs  in  arid  lands  have  been  emphasized  in  other 
studies  (McKell  et  al.  1972).  Additional  publications 
about  rangelands  have  emphasized  floral  and  faunal 
relationships  in  Great  Basin  habitats  (Harper  and 
Reveal  1978)  and  topics  such  as  seedbed  preparation 
and  planting  and  the  management  of  seeded  ranges 
(Plummer  et  al.  1955,  1968),  impacts  of  fire  on 
shrubs  (Wright  1972;  Wright  et  al.  1979),  and  im- 
pacts of  drought  on  grasslands  (Albertson  and 
Weaver  1946). 

Studies  of  rangelands  that  describe  the  structure, 
composition,  distribution,  and  ecological  relation- 
ships among  vegetative  components  have  been  done 
on  a  State  (Nichol  1937;  Costello  1944)  and  a  re 
gional  basis  (Weaver  1954;  Weaver  and  Albertson 
1956).  A  map  of  the  natural  vegetation  in  the  U.S. 
and  a  description  of  associated  cover  types  was  pub- 
lished in  the  Atlas  of  American  Agriculture  (U.S.  De- 
partment of  Agriculture  1936).  The  presumed 
distribution  of  potential  range  vegetation  types  in 
the  U.S.  was  published  in  map  form  by  Kuchler 
(1964;  see  Figure  1).  Kuchler  (1964)  also  provided 
a  photographic  description  of  different  potential 
rangeland  types.  A  more  intensive  mapping  of  range 
cover  types  in  the  southwestern  U.S.  was  done  by 
Brown  and  Lowe  (1980),  and  a  text  description  of 
range  types  was  provided  by  Brown  (  1982). 

Many  other  authors  have  provided  specialized 
descriptions  of  rangeland  vegetation.  These  descrip- 
tions vary  from  statewide  assessments  to  intensive 
treatments  of  small  range  areas.  Several  range  vegeta- 
tion classification  systems  are  summarized  in  Table 
2  by  area  of  application,  general  type  of  rangeland 
vegetation  considered,  a  brief  description  of  the 
vegetative  classification  system,  and  the  apparent 
utility  of  the  system  in  inventories  and  assessments 
of  wildlife  habitat. 

The  references  in  Table  2  will  be  especially 
helpful  to  persons  working  in  unfamiliar  range  habi- 
tats. The  publications  listed  frequently  provide  maps 
of  the  distribution  of  range  vegetation,  descriptions 
of  vegetation  communities,  and  lists  of  wildlife  spe- 
cies that  use  different  range  cover  types. 


96 


Rangelands 


Table  2.     Range  vegetation  classification  systems  developed  for  grassland,  shrubland,  and  woodland  cover 
types  in  the  western  U.S.  and  their  utility  in  inventories  and  assessments  of  wildlife  habitat  quality. 


Area  of 
application 

Type  of 
vegetation 

Classification  system 

Utility  in  inventorying 
and  monitoring 

Reference 

Grasslands 

1.  Southeastern 
Washington, 
northeastern  Or- 
egon, west- 
central  Idaho 

Fescue-wheatgrass 

Descriptions  of  vege- 
tative associations 

Describes  a  variety 
of  climax  communi- 
ties in  Pacific  North- 
west bunch-grass 
types 

Tisdale  (1983) 

2.   Idaho 

Fescue-wheatgrass 

Description  of  vege- 
tative associations 
similar  to  those  of 
Tisdale  (1983) 

Provides  map  of 
potential  vegetation 
types  in  Idaho 

Sharp  and  Sanders 
(1978) 

3.  Snake  River 
Canyon,  Idaho 

Fescue-wheatgrass 

Dichotomous  key  to 
grasslands  in  the 
Snake  River  Canyon 

Lists  specific  habitat 
criteria  for  vegetative 
association 

Tisdale  (1979) 

4    Southwestern 
U.S. 

Alpine  and  subalpine 
grasslands 

Description  of  vege- 
tative associations 

Text,  in  conjunction 
with  map  (Brown  and 
Lowe  1980),  pro- 
vides a  description 
of  the  location  and 
structure  of  this 
vegetative  commu- 
nity in  the  Southwest 

Brown  (1982) 

5.  Southwestern 
U.S. 

Plains  and  Great 
Basin  grasslands 

Same  as  above 

Same  as  above 

Brown  (1982) 

6.  Southwestern 
U.S. 

Semidesert  grass- 
land 

Same  as  above 

Same  as  above 

Brown  (1982) 

7.  Southwestern 
U.S. 

California  valley 
grassland 

Same  as  above 

Same  as  above 

Brown  (1982) 

8.  Southwestern 
U.S. 

Sonoran  savanna 
grassland 

Same  as  above 

Same  as  above 

Brown  (1982) 

9.  California 

Grasslands 

Dichotomous  key  to 
grasses;  illustrations 
of  common  species 

Describes  appear- 
ance, distribution 
and  use  of  individual 
species 

Sampson  et  al. 
(1951) 

Shrublands 

1 .  Southern  Califor- 
nia 

Chamise  chaparral 
and  mixed  chaparral 

Lists  of  plant  compo- 
nents of  different 
vegetative  types 

Describes  some 
vegetative  associa- 
tions along  moisture 
and  temperature 
gradients 

Mooney  and  Parsons 
(1973) 

2    Southwestern 
U.S. 

Oak  brush  (Quercus 
sp.) 

Description  of  vege- 
tative associations 

Text,  in  conjunction 
with  map  (Brown  and 
Lowe  1980),  pro- 
vides a  description 
of  the  location  and 
structure  of  this 
community  in  the 
Southwest 

Brown  (1982) 

3.  Southwestern 
U.S. 

California  coastal 
scrub 

Same  as  above 

Same  as  above 

Brown  (1982) 

4.  Southwestern 
U.S. 

California  coastal 
chaparral 

Same  as  above 

Same  as  above 

Brown  (1982) 

5.  Southwestern 
U.S. 

Interior  chaparral 

Same  as  above 

Same  as  above 

Brown  (1982) 

6    New  Mexico 

Saltbush  (Atnplex 
sp.) 

Dichotomous  key, 
photographs,  species 
descriptions,  range 
maps 

Describes  species  of 
Atnplex  and  identi- 
fies their  range 

Wagner  and  Aldon 
(1978) 

Rangelands 


97 


Table  2.     Range  vegetation  classification  systems  developed  for  grassland,  shrubland,  and  woodland  cover 
types  in  the  western  U.S.  and  their  utility  in  inventories  and  assessments  of  wildlife  habitat  quality 
(continued). 


Area  of 
application 

Type  of 
vegetation 

Classification  system 

Utility  in  inventorying 
and  monitoring 

Reference 

Shrublands  (continued, 

7.   Intermountain 
area 

Sagebrush  (Artemi- 
sia sp),  rabbitbrush 
(Chrysothamnus 
sp),  horsebrush  (Te- 
tradymia  sp.),  and 
matchbrush  or 
snakewood  (Xantho- 
cephalum  sp.) 

Description  and 
dichotomous  key  for 
species  of  the  four 
shrubs 

Describes  appear- 
ance, distribution, 
and  use  of  individual 
species;  provides 
illustrations  that  may 
be  helpful  in  species 
identification 

McArthur  et  al. 
(1979) 

8.  Western  Mon- 
tana 

Sagebrush- 
grasslands 

Key  based  on  un- 
derstory  and  mid- 
story  components 

Describes  distribu- 
tion and  vegetative 
composition  of  cover 
types,  impacts  of 
grazing  on  cover 
types,  and  response 
of  cover  types  to 
range  management 

Mueggler  and  Stew- 
art (1980) 

9.   Intermountain 
region 

Sagebrush- 
grasslands 

Variety  of  taxonomic 
keys  based  on  vege- 
tative characteristics 

Describes  species  of 
sagebrush;  manage- 
ment of  sagebrush 
for  livestock;  and  the 
presumed  impacts 
of  sagebrush  man- 
agement on  livestock 

Blaisdell  et  al.  (1982) 

10.  Columbia  River 
Basin-Idaho, 
Oregon,  Wash- 
ington 

Sagebrush-grass- 
lands 

Lists  of  midstory- 
understory  compo- 
nents 

Association  of  sage- 
brush-grassland 
complexes  with 
edaphic  and  climatic 
factors 

Hironaka  (1979) 

1 1 .   Northern  Great 
Plains 

Sagebrush 
species 

Lists  ecological 
relationships  of  some 
common  sagebrush 
species  and  associ- 
ates 

Describes  factors 
contributing  to  the 
distribution  of  sage- 
brush species  in 
Montana,  Wyoming, 
and  the  Dakotas 

Johnson  (1979) 

12.  Blue  Mountains 
in  eastern  Ore- 
gon and  south- 
eastern 
Washington 

Sagebrush-domi- 
nated rangelands 

Describes  environ- 
mental and  edaphic 
conditions  character- 
istic of  range  type 

Lists  species  compo- 
nents, range  appear- 
ance, and  criteria 
for  describing  range 
condition 

Hall  (1973) 

13.  Western  U.S. 

Sagebrush 

Dichotomous  key 

Photographs  of 
herbarium  types  and 
distribution  maps  of 
12  sagebrush  taxa 

Beetle  (1960) 

14    Southern  Idaho 

Sagebrush- 
grasslands 

Dichotomous  key 

Describes  wildlife 
use  of  habitat  types 
and  provides  sug- 
gestion for  manage- 
ment of  habitats 

Hironaka  et  al. 
(1983) 

15.  Oregon 

Sagebrush 

Diagrammatic  key 

Describes  habitats 
where  different  sage- 
brush species  occur 

Winward  (1980) 

98 


Rangelands 


Table  2.     Range  vegetation  classification  systems  developed  for  grassland,  shrubland,  and  woodland  cover 
types  in  the  western  U.S.  and  their  utility  in  inventories  and  assessments  of  wildlife  habitat  quality 
(concluded). 


Area  of 
application 

Type  of 
vegetation 

Classification  system 

Utility  in  inventorying 
and  monitoring 

Reference 

Woodlands 

1.  Southwestern 
US 

Pinyon-juniper 

Descriptions  of  vege- 
tative associations 

Text,  in  conjunction 
with  map  (Brown  and 
Lowe  1980),  pro- 
vides a  description 
of  the  location  and 
structure  of  the 
pinyon-juniper  com- 
munity in  the  South- 
west 

Brown  (1982) 

2    Southwestern 
U.S. 

Oak-woodlands 

Same  as  above 

Same  as  above 

Brown  (1982) 

MAJOR  SPECIES  GROUPS 

Western  range  ecosystems  provide  habitat  for  at 
least  10  rare  and  endangered  terrestrial  vertebrate 
wildlife  species  (U.S.  Forest  Service  1972): 

•  Utah  prairie  dog  {Cynomys parvidens) 

•  Grizzly  bear  (Ursus  arctos  horribilis) 

•  Yuma  clapper  rail  (Rallus  longirostris  yuma- 
nensis) 

•  Light-footed  clapper  rail  (R  1.  levipes) 

•  Masked  bobwhite  (Colinus  virginianus  ridg- 
wayi) 

•  California  condor  {Gymnogyps  calif ornianus) 

•  Sonoran  pronghorn  (Antilocapra  americana 
sonoriensis) 

•  San  Joaquin  kit  fox  (Vulpes  macrotis  mutica) 

•  Black-footed  ferret  (M  us  tela  nigripes) 

•  Whooping  crane  (Grus  americana) 

The  use  of  rangeland  habitats  by  terrestrial  verte- 
brate wildlife  species  is  probably  mostly  dependent 
on  the  structure  of  the  rangeland  vegetative  cover. 
For  example,  treeland  habitats  generally  support  a 
more  diverse  wildlife  community  than  do  shrubland 
or  grassland  habitats,  and  shrubland  habitats  fre- 
quently support  a  more  diverse  wildlife  community 
than  do  grassland  habitats.  Other  factors  influencing 
the  diversity  of  the  wildlife  community  on  a  range- 
land  include — 

•  latitude; 

•  climate; 


•  topography; 

•  distance  to  water; 

•  distance  to  cliff  faces,  caves,  or  other  features; 

•  density,  height,  and  vigor  of  range  vegetation; 
and 

•  presence  or  absence  of  particular  plant  species. 

Many  of  these  factors  are  described  in  greater  detail 
in  the  following  section. 

CRITICAL  HABITAT  FEATURES 

Range  surveys  are  often  conducted  to  describe 
soil  profiles  and  to  estimate  ( 1 )  the  potential  natural 
vegetation  that  can  occur  on  a  range  habitat,  (2) 
the  condition  of  range  vegetation,  and  (3)  the  wild- 
life community  that  is  present. 

Range  surveys,  until  World  War  II,  usually  were 
conducted  entirely  by  ground  methods,  such  as  tra- 
versing compass  lines  across  landscapes  and  sketch- 
ing in  planimetric  and  vegetation  details  (Poulton 
1970).  Cost  effectiveness  and  usefulness  of  range 
surveys  were  enhanced  by  the  widespread  accept- 
ance of  interpretive  aerial  photography.  Interpreta- 
tion of  aerial  photographs,  combined  with  field 
surveys,  offers  an  increased  capability  to  evaluate 
large  land  areas  to — 

( 1 )  identify  many  plant  species  and  soil  surface 
features; 

(2)  classify  surface  features  into  specific  habitat 
types; 


Rangelands 


99 


(3)  measure  vegetation  characteristics  such  as 
cover,  density,  height,  and  spatial  distribution; 
and 

(4)  monitor  changes  in  vegetation  and  soil  sur- 
face conditions  over  time. 

Samples  of  current  annual  growth  collected 
during  range  surveys  frequently  are  subjected  to 
chemical,  calorimetric,  and  digestive  analyses  to  de- 
velop estimates  of  range  carrying  capacity  for  domes- 
tic livestock  and  selected  wildlife  species.  Range 
conditions  often  are  evaluated  by  determining  if  the 
plants  that  characteristically  should  grow  in  the  area 
are  present  and  in  good  vigor  (Stoddard  et  al.  1975). 
Surveys  to  determine  the  presence  of  wildlife  in 
range  cover  types  usually  involve  on-site  assessments 
with  standardized  trapping  techniques,  especially 
for  small  mammals  and  reptiles,  and  census  lines  for 
avian  surveys  or  for  locating  pellet  group  plots  for 
large  mammals.  Animal  survey  techniques  sometimes 
are  supplemented  with  interpreted  aerial  photo- 
graphs that  show,  for  example,  waterfowl  on  lakes, 
deer  and  elk  against  snowfield,  and  domestic  live- 
stock on  range. 

Rangelands  are  complex  ecosystems  that  can  be 
studied  at  many  levels  of  detail.  Inventories  can  be 
very  intensive  and  applicable  to  small  areas  or  be 
extremely  general,  providing  limited  information  to 
managers.  It  seems  desirable,  in  this  chapter,  to  pro- 
vide a  general  framework  for  conducting  surveys 
to  inventory  rangelands  and  monitor  their  wildlife 
values.  This  framework  is  a  conceptual  representa- 
tion of  the  structure  of  rangeland  within  which  gen- 
eral survey  information  applicable  to  large  land  areas 
can  be  gathered  and  analyzed  in  a  timely  manner 
and  within  which  very  specific  information  also  can 
be  gathered.  I  have  not  included  extensive  lists  of 
range  variables  and  techniques  for  their  measure- 
ment; this  information  is  available  elsewhere.  Instead, 
I  have  provided  a  general  inventory  process  that — 


is  universally  applicable  to  a  wide  variety  of 
terrestrial  rangeland  types; 

considers  all  the  vertebrate  wildlife  species  that 
might  occur  on  a  rangeland; 

allows  a  manager  to  predict  the  impacts  that 
land  use  change  might  have  on  wildlife  species; 
and 

uses  state-of-the-art  technologies  to  provide 
quantitative  appraisals  of  rangelands  so  that 
rangeland  resources  can  be  monitored  through 
time. 


This  inventory  process  includes  a  way  to  abstractly 
represent  the  structure  of  range  vegetation,  associate 
wildlife  species  with  the  structure  of  range  vegeta- 
tion, and  relate  some  types  of  inventory  information 
to  range  management. 


It  is  unrealistic  to  consider  the  total  wildlife 
community  in  inventories  and  assessments  of  range- 
lands,  unless  some  general  criteria  can  be  identified 
that  determine  how  wildlife  species  use  habitats  and 
are  easily  measured.  Layers  of  habitat,  discussed  be- 
low, is  an  easily-measured  criterion  that  is  a  critical 
habitat  feature  for  terrestrial  vertebrate  species. 


Any  terrestrial  habitat,  including  a  rangeland, 
can  be  considered  a  volume  of  space  that  contains 
structure,  such  as  vegetation  which  may  be  useful  to 
wildlife  species.  The  volume  of  space  in  a  mixed 
rangeland  habitat  containing  shrubland,  grassland, 
and  treeland  components  is  represented  in  Part  A  of 
Figure  2.  The  structure  of  this  rangeland  habitat  also 
can  be  considered  in  terms  of  habitat  layers,  as  in 
Part  B  of  Figure  2,  which  can  be  used  to  inventory 
and  evaluate  rangeland  habitats.  The  different  habitat 
layers  are  defined  in  Short  and  Burnham  (1982) 
and  in  Table  3. 


B 


Midstory 


Midstory 


Tree  bole 


Understory 


rflftiii&KkA 


Terrestrial  subsurface 


Figure  2.     Cover  types  within  a  rangeland  habitat  can  be  considered  in  terms  of  layers  of  habitat. 
100  Rangelands 


Table  3.     Suggested  criteria  for  determining  the  presence  of  different  layers  of  habitat. 


Layer 

Criteria 

Tree  canopy  or  overstory 

Vegetation  structure  is  8  m  (25  ft)  or  more  above  the  terrestrial  or  aquatic  sur- 
face and  provides  at  least  5%  cover  when  projected  to  the  surface  (500  m2/ 
ha  or  2,200  ft2/a.). 

Tree  bole 

Tree  trunks  have  a  dbh  20  cm  (8  in.)  and  occur  at  a  density  12/ha  (5/a). 

Shrub  midstory 

Vegetation  height  from  50  cm  (20  in.)  up  to  8  m  (25  ft),  which  provides  at  least 
5%  cover  when  projected  to  the  surface  (500  m2/ha  or  2,200  ft2/a.). 

Understory 

Layer  extends  from  10  cm  (4  in.)  below  the  apparent  surface  up  to,  but  not 
including,  50  cm  (20  in.)  above  the  apparent  surface  and  provides  at  least  5% 
cover  when  projected  to  the  surface  (500  m2/ha  or  2,200  ft2/a). 

Terrestrial  subsurface 

Extends  from  more  than  10  cm  (4  in.)  below  the  apparent  surface  down. 

Surface  water  layer 

Land  surface-water  interface  and  shallow  water  up  to  25  cm  (10  in.)  deep. 

The  surface  water  layer  includes  the  shoreline 
and  shallow  water  (<  25cm  [10  in.]  deep)  areas  and 
the  terrestrial  substrate  that  is  under  the  shallow 
water.  Vegetation  that  emerges  through  shallow 
water  and  is  less  than  0.5  m  (1.6  ft)  high  is  consid- 
ered a  product  of  the  surface  water  layer,  as  is  float- 
ing vegetation  in  shallow  water. 

The  terrestrial  subsurface  layer  extends  upward 
to  within  10  cm  (4  in.)  of  the  apparent  surface  and 
includes  caves  and  deep  crevices  that  are  critical 
habitat  features  for  some  wildlife  species  and  a  soil 
substrate  suitable  for  the  construction  of  burrows 
and  dens. 

The  terrestrial  surface  layer  extends  from  10  cm 
(4  in.)  below  to  0.5  m  (1.6  ft)  above  the  apparent 
surface.  This  layer  includes  bare  ground,  talus,  cliff 
faces,  litter,  herbaceous  vegetation,  and  dwarf  or 
supine  woody  vegetation  as  habitat  features. 

The  shrub  or  midstory  layer  extends  from  0.5  to 
8.0  m  (1.6  to  26.4  ft)  above  the  apparent  surface 
and  includes  the  canopies  of  shrubs  and  dwarf  trees, 
as  well  as  other  vegetative  plant  parts  that  extend 
into  this  strata. 

The  tree  bole  layer  consists  of  live  and  dead 
tree  trunks  that  provide  both  a  foraging  substrate 
and  a  variety  of  nest  substrates,  including  bark  and 
cavities.  Tree  boles  comprise  a  vertical  feature  that 
extends  through  horizontal  layers.  Snags  are  special- 
ized tree  boles. 

The  tree  canopy  extends  above  8.0  m  (26.4  ft) 
and  provides  large  and  small  branches  of  living  and 
dead  trees  as  nest  substrates  and  includes  leaves, 
buds,  and  fruit  as  food  for  wildlife  species. 

Considering  habitat  in  terms  of  layers  has  sev- 
eral advantages.  The  concept  seems  intuitively  cor- 


rect, insofar  as  the  presence  of  layers  is  readily 
perceivable,  even  though  attempts  to  define  the  lay- 
ers may  seem  arbitrary  (Table  3).  Layers  of  habitat 
can  be  mapped,  wildlife  species  can  be  identified 
with  particular  habitat  layers,  and  many  management 
actions  impact  habitat  layers,  which  can  be  used  to 
predict  the  impacts  of  land  use  change  on  wildlife. 

Wildlife  species  can  be  associated  with  layers  of 
habitat  because  the  niche  of  an  individual  species 
occurs  within  one  or  more  habitat  strata.  Grasslands 
contain  only  the  terrestrial  subsurface  and  terrestrial 
surface  layers,  defined  above.  Shrublands  provide 
terrestrial  subsurface,  terrestrial  surface,  and  mid- 
story layers;  woodland  habitats  may  also  provide  tree 
boles  for  cavity  makers  and  users  and  a  tree  canopy 
layer.  Rangeland  habitats  that  provide  only  a  few 
layers  of  habitat  have  a  limited  volume  of  space 
within  which  wildlife  species  can  find  niches.  More 
niches  are  potentially  available  as  more  layers  of 
habitat  occur  in  cover  types,  so  more  wildlife  spe- 
cies potentially  are  supported  by  more  structurally 
complex  habitats. 

Rangelands  can  be  described  in  terms  of  habitat 
layers.  For  example,  Kuchler  (1964)  listed  27  poten- 
tial natural  vegetation  types  in  the  western  U.S.  that 
can  be  considered  rangelands.  These  types  are  ar- 
rayed in  Figure  3  according  to  the  number  of  habitat 
layers  that  each  type  provides  for  the  wildlife  com- 
munity. Considerable  similarity  exists  in  the  struc- 
ture of  the  wildlife  communities  that  occur  in  the  14 
grassland-dwarf  shrub  cover  types,  which  contain 
terrestrial  subsurface  and  terrestrial  surface  layers  of 
habitat.  Great  similarity  also  is  evident  in  the  struc- 
ture of  the  wildlife  communities  that  occur  in  the 
seven  shrubland  cover  types,  which  contain  terres- 
trial subsurface,  terrestrial  surface,  and  midstory 
layers  of  habitat.  The  similarities  in  the  wildlife  com- 
munities within  grasslands  and  the  communities 
within  shrublands  occur  even  when  the  vegetation 


Rangelands 


101 


Sandsage-bluestem  prairie 

Bluestem-grama  prairie 

Wheatgrass-needlegrass 

Grama-buffalograss 

Grama-needlegrass-wheatgrass 

Foothills  prairie 

Galleta-three  awn  shrubsteppe 

Wheatgrass-needlegrass- 
shrubsteppe 

Grama-galleta  steppe 

Wheatgrass-bluegrass 

Fescue-wheatgrass 

California  steppe 

Blackbrush 

Fescue-oatgrass 


Juniper  steppe  woodland 
Sagebrush  steppe 
Saltbush-greasewood 
Great  Basin  sagebrush 
Mountain  mahogany-oak  scrub 
Coastal  sagebrush 
Chaparral 


Oak-juniper  woodland 
Juniper-pinyon  woodland 
Mesquite-savanna 
Mesquite-acacia  savanna 


Mesquite-live  oak  savanna 
Montane  chaparral 


Figure  3.     Rangeland  cover  types  provide  different  habitat  layers  for  the  wildlife  community. 


1 02 


Rangelands 


species  differ  among  grasslands  or  among  shrublands. 
Certain  broad  habitat  changes  within  grasslands  or 
within  shrublands  are  expected  to  affect  similar  func- 
tional segments  of  the  wildlife  community  in  a  pre- 
dictable manner.  The  capability  to  predict  impacts 
on  the  wildlife  community  resulting  from  changes  in 
a  cover  type  is  important  to  resource  managers. 

Layers  of  habitat  can  be  mapped  and  used  to 
describe  potential  wildlife  habitat  values  in  a  plan- 
ning unit.  As  the  quantity  of  a  habitat  layer  changes, 
the  quantity  of  habitat  available  for  species  depend- 
ent on  that  habitat  layer  also  changes.  Thus,  a  meas- 
ure of  the  structure  of  a  rangeland  provides  a  first 
estimate  of  the  utility  of  that  rangeland  as  habitat  for 
wildlife  species. 


Habitat  Layer  Index  (HLI) 


The  interpretation  of  aerial  photography  has 
become  an  important  tool  in  range  surveys  because 
it  is  the  best  way  to  represent  the  physical  relation- 
ships between  surface  cover  features.  For  example, 
aerial  photographs  have  often  been  used  to  describe 
where  shrubland,  grassland,  and  treeland  cover 
types,  depicted  in  Figure  2A,  occur  within  a  study 
area.  An  aerial  photograph  that  has  been  interpreted 
to  describe  cover  types  is  provided  in  Figure  4, 
which  indicates  the  location  and  identification  of 
surface  cover  types  on  a  selected  land  area  within 
the  Piceance  Basin  of  northwestern  Colorado.  The 
photograph  has  been  interpreted  according  to  the 


423PJBS 


212 


522SG 


523SGHR 


522SG 


523SGHR 


Figure  4.     The  cover  type  polygons  in  this  figure  were  traced  from  an  interpreted  aerial  photograph  of  a 

small  portion  of  the  Piceance  Basin  in  northwestern  Colorado.  Identification  of  the  vegetation  within  indi- 
vidual polygons  can  be  made  from  vegetation  classes  listed  in  Table  5.  Segments  A  and  B  are  described 
in  detail  in  the  text. 


Rangelands 


103 


hierarchical  classification  system  listed  in  Table  4, 
coupled  with  ground  truthing,  which  is  the  on-site 
examination  of  photo  points  to  correlate  images  on 
the  photograph  with  cover  features  on  the  ground. 

Maps  of  cover  types,  like  that  in  Figure  4,  can 
provide  useful  information  to  persons  trying  to  in- 
ventory and  assess  wildlife  habitat  on  rangelands 
because  some  wildlife  species,  within  their  distribu- 
tion range,  are  closely  associated  with  the  habitat 
structure  provided  by  specific  cover  types.  For  ex- 
ample, species  like  sage  grouse  (Centrocercus  uro- 
phasianus),  Brewer's  sparrows  (Spizella  breweri), 
sage  sparrows  (Amphispiza  belli),  and  sagebrush 
voles  (Lagurus  curtatus)  may  be  dependent  on  the 
presence  of  sagebrush  (Artemesia  sp. )  habitat, 
whereas  bushtits  (Psaltriparus  minimus),  ash- 
throated  flycatchers  (Myiarchus  cinerascens),  gray 
vireos  (Vireo  vicinior),  black-throated  gray  warblers 
(Dendroica  nigrescens),  and  pinyon  mice  (Peromys- 
cus  truei)  may  be  dependent  on  the  presence  of 
pinyon-juniper  (Pinus-Juniperus  sp.)  and  oak  brush 
(Quercus  sp. ).  Lark  buntings  (Calamospiza  melano- 
corys),  horned  larks  (Eremophila  alpestris),  chest- 
nut-collared longspurs  (Calcarius  ornatus), 
McCown's  longspurs  (Calcarius  mccownii),  western 


meadowlarks  (Sturnella  neglecta),  Baird's  sparrows 
(Ammodramus  bairdii),  swift  foxes  (Vulpes  velox), 
northern  grasshopper  mice  (Onchomys  leucogaster), 
plains  harvest  mice  (Reithrodontomys  montanus), 
Ord's  kangaroo  rats  (Dipodomys  ordii),  thirteen- 
lined  ground  squirrels  (Spermophilus  tridecemlinea- 
tus),  and  pronghorns  (Antilocapra  americana), 
may  be  dependent  on  the  presence  of  prairie  grasses. 

Interpreted  aerial  photographs  can  be  used  to 
determine  the  presence  of  other  important  habitat 
features  like  high  cliffs  and  cliff  faces,  which  may  be 
habitat  for  peregrine  falcons  (Falco  peregrinus), 
prairie  falcons  {Falco  mexicanus),  black  swifts  (Cyp- 
seloides  niger),  white-throated  swifts  (Aeronautes 
saxatalis),  cliff  swallows  (Hirundo  pyrrhonota),  and 
canyon  wrens  (Catherpes  mexicanus);  boulder 
fields,  talus  slopes,  and  rock  slides  in  mountainous 
areas,  which  may  be  habitat  for  pikas  (Ochotona 
princeps);  caves,  which  may  be  habitat  for  a  variety 
of  bats;  and  prairie  dog  towns,  which  may  be  habitat 
for  black-tailed  prairie  dogs  (Cynomys  ludovici- 
anus),  black-footed  ferrets,  and  burrowing  owls 
(Athene  cunicularia);  and  water  sources  essential  to 
the  reproduction  of  amphibians  and  a  life  requisite 
for  numerous  other  species.  The  absence  of  these 


Table  4.     Hierarchical  classification  system  for  interpreting  aerial  photographs,  developed  for  the  Piceance 
Basin  of  Northeastern  Colorado. 


Level  1 

Level  2 

Level  3 

Level  4 

Level  5 

1.  Urban  or 

built-up  land 

1.  Commercial 

2.  Residential 

1.  Transportation 
(update  of 
USGS 
maps  only) 

2.  Agricultural/ 
reclaimed 
land 

1.  Agricultural 

2.  Reclaimed  land 

1.  Irrigated  crops 

2.  Nonirrigated 

crops 

3.  Orchards, 

groves, 
nurseries 

4.  Other 

1.  Reclaimed  mine 

land 

2.  Other  reclaimed 

land 

3.  Chained  land 

4.  Other 

3,  Wetlands 

(interpreted 

according  to 

National 

Wetland 

Inventory 

Standards) 

Percent  cover  of 
dominant  cover 
type 

Dominant  cover 

Subdominant 
cover 

104 


Rangelands 


Table  4.     Hierarchical  classification  system  for  interpreting  aerial  photographs,  developed  for  the  Piceance 
Basin  of  Northeastern  Colorado  (continued). 


Level  1 

Level  2 

Level  3 

Level  4 

Level  5 

4.  Forest  (>20% 

1.  Deciduous 

1. 

Closed  80  +  % 

AS  Aspen 

BS  Bare  soil 

trees) 

2. 

Open  50  to  80% 

UD  Upland 

BR  Bare  rock 

3. 

Sparse  20  to 

deciduous 

HR  Herbaceous 

50% 

OT  Other 

SG  Sagebrush 
OB  Oakbrush 
US  Upland  shrub 
HA  Halophytic 

shrub 
OT  Other 

2.   Evergreen 

1. 

Closed  80  +  % 

PD  Ponderosa 

BS  Bare  soil 

2. 

Open  50  to  80% 

pine 

BR  Bare  rock 

3. 

Sparse  20  to 

JN  Juniper 

HR  Herbaceous 

50% 

PY  Pinyon  pine 
DF  Douglas  fir 
SP  Spruce  sp. 
SA  Subalpine  fir 
LP  Lodgepole 

pine 
PJ  Pinyon  pine/ 

juniper  assn. 
SF  Spruce/fir  assn. 
CN  Coniferous 
OT  Other 

SG  Sagebrush 
OB  Oakbrush 
US  Upland  shrub 
HA  Halophytic 

shrub 
OT  Other 

3.  Mixed 

1. 

Closed  80  +  % 

AS  Aspen 

BS  Bare  soil 

deciduous/ 

2. 

Open  50  to  80% 

PD  Ponderosa 

BR  Bare  rock 

evergreen 

3. 

Sparse  20  to 

pine 

HR  Herbaceous 

(>33%  de- 

50% 

JN  Juniper 

SG  Sagebrush 

ciduous  and 

PY  Pinyon  pine 

OB  Oakbrush 

>33%  ever- 

DF Douglas  fir 

US  Upland  shrub 

green) 

SP  Spruce  sp. 
SA  Subalpine  fir 
LP  Lodgepole 

pine 
PJ  Pinyon  pine/ 

juniper  assn. 
SF  Spruce/fir  assn. 
UD  Upland 

deciduous 
CN  Coniferous 
OT  Other 

HA  Halophytic 

shrub 
OT  Other 

5.  Shrubland 

1.  Deciduous 

1. 

Closed  80  +  % 

OB  Oakbrush 

BS  Bare  soil 

(>20% 

2. 

Open  50  to  80% 

US  Upland  shrub 

BR  Bare  rock 

shrubs 

3. 

Sparse  20  to 

OT  Other 

HR  Herbaceous 

and  <20% 

50% 

OT  Other 

trees) 

2.   Evergreen 

1. 

Closed  80  +  % 

SG  Sagebrush 

BS  Bare  soil 

2 

Open  50  to  80% 

SL  Saltbrush 

BR  Bare  rock 

3. 

Sparse  20  to 

GR  Greasewood 

HR  Herbaceous 

50% 

HA  Halophytic 

shrub 
OT  Other 

OT  Other 

3.   Mixed 

1. 

Closed  80  +  % 

SG  Sagebrush 

BS  Bare  soil 

deciduous/ 

2. 

Open  50  to  80% 

OB  Oakbrush 

BR  Bare  rock 

evergreen 

3. 

Sparse  20  to 

SL  Saltbrush 

HR  Herbaceous 

(>33%  de- 

50% 

GR  Greasewood 

OT  Other 

ciduous  and 

US  Upland  shrub 

>33%  ever- 

HA Halophytic 

green) 

shrub 

^__ 

OT  Other 

Rangelands 


105 


Table  4.     Hierarchical  classification  system  for  interpreting  aerial  photographs,  developed  for  the  Piceance 
Basin  of  Northeastern  Colorado  (concluded). 


Level  1 

Level  2 

Level  3 

Level  4 

Level  5 

6.  Grassland 

(>20%  grass 
and  <20% 
trees  and 
<20%  shrubs) 

1.  Plains 

grassland 

2.  Mountain 

grassland 

1 .  Closed  80% 

2.  Open  50  to  80% 

3.  Sparse  20  to 

50% 

1.  Closed  80  +  % 

2.  Open  50  to  80% 

3.  Sparse  20  to 

50% 

TL  Tallgrass 
SH  Shortgrass 
BN  Bunchgrass 
DS  Desert  grass 
AN  Annual  grass 
OT  Other 

TL  Tallgrass 
SH  Shortgrass 
BN  Bunchgrass 
DS  Desert  grass 
AN  Annual  grass 
OT  Other 

BS  Bare  soil 
BR  Bare  rock 
OT  Other 

BS  Bare  Soil 
BR  Bare  rock 
OT  Other 

7.  Barren  land 
(>40%  bare 
ground  and 
<20%  trees 
and  <20% 
shrubs  and 
<20% 
grasses) 

1.  Natural 

2.  Man-made 

1.  Bare  exposed 
rock 

2  Bare  exposed 

soil 

3.  Spires/cliffs 

1.  Quarries 

2.  Strip  mine 

3  Gravel  pit 

4.  Mine  spoils 

5.  Nonvegetated 

HR  Herbaceous 
SG  Sagebrush 
US  Upland  shrub 
HA  Halophytic 

shrub 
PJ  Pinyon  pine/ 

juniper  assn. 
SF  Spruce/fir  assn. 
AS  Aspen 
CN  Coniferous 
OT  Other 

8.  Tundra 

1.  Shrub  and 

brush  tundra 

2.  Herbaceous 

tundra 

3.  Bare  ground 

tundra 

4.  Wet  tundra 

5.  Mixed  tundra 

9.  Perennial  snow 
or  ice 

1 .  Perennial 

snowfield 

2.  Glacier 

important  habitat  features  in  a  planning  unit  or  study 
site  suggests  that  wildlife  species  dependent  on  the 
structure  of  these  physical  features  will  be  uncom- 
mon or  absent. 

Many  wildlife  species  seem  to  utilize  a  habitat 
structure  more  general  than  that  provided  by  a  sin- 
gle plant  species.  Consequently,  it  may  be  easier 
to  produce  maps  like  that  in  Figure  4  than  to  deter- 
mine what  the  maps  really  mean  in  terms  of  species 
presence.  For  example,  it  is  difficult  to  list  the  wild- 
life species  expected  to  use  each  of  the  polygons 
in  Figure  4  or  to  compare  the  structure  of  the  habi- 
tats in  segments  A  and  B.  However,  the  concept  of 
habitat  layers  provides  a  framework  for  interpreting 


the  usefulness  of  the  structure  of  such  habitats  to  the 
wildlife  community. 

HLI  provides  a  way  to  quantitatively  describe 
the  relative  structural  complexity  of  the  vegetative 
cover  occurring  in  a  study  area.  It  provides  a  num- 
ber between  0.0  (no  structural  diversity)  and  1.0 
(high  structural  diversity)  that  can  be  used  to  char- 
acterize vegetative  diversity  on  a  parcel  of  land, 
compare  vegetative  diversity  on  several  parcels  of 
land,  or  provide  a  basis  for  statistical  assessments 
that  show  direction  and  rate  of  change  in  habitat 
structure  over  time.  The  index  does  not  predict  the 
suitability  of  habitats  for  individual  wildlife  species, 
although  wildlife  species  that  tend  to  occur  in  struc- 


106 


Range-lands 


;. 


turally  complex  habitats  may  occur  where  HLIs  are 
high,  whereas  other  wildlife  species  requiring  simple 
habitats  may  occur  where  HLIs  are  low.  High  HLIs 
should  not  be  equated  to  "good"  habitat  and  low 
HLIs  to  "poor."  A  grassland,  for  example,  might  pro- 
vide excellent  habitat  for  grassland  wildlife  species 
and  still  receive  a  low  HLI  because  of  its  limited 
structural  diversity.  Disturbance  to  a  grassland  could 
result  in  the  invasion  of  shrub  species,  which  might 
produce  increased  structural  diversity  and  a  high 
HLI,  but  result  in  a  poor  quality  habitat  for  grassland 


wildlife  species.  The  HLI  would  be  considered  sim- 
ply as  a  way  to  numerically  represent  change  in 
structure  in  order  to  enhance  habitat  assessments. 


The  cover  types  depicted  in  Figure  2A  were 
represented  in  terms  of  layers  of  habitat  in  Figure 
2B.  Likewise,  the  aerial  photograph  of  the  land  area 
within  the  Piceance  Basin  can  be  represented  in 
terms  of  cover  types  ( Figure  4  )  or  in  terms  of  layers 
of  habitat.  This  interpretation  (Figure  5)  is  based 


Two  layers  of  habitat 


Three  layers  of  habitat 


Four  layers  of  habitat 


Figure  5.     Aerial  photograph  of  the  land  area  represented  in  Figure  4  has  been  reinterpreted  to  indicate  the 
number  of  layers  of  habitat  present  in  different  polygons. 


Rangelands 


107 


on  the  assumption  that  the  niche  of  a  wildlife  spe- 
cies can  be  associated  with  the  structure  of  habitat, 
represented  in  terms  of  layers  of  habitat.  Figure  5 
identifies  the  number  of  habitat  layers  present  in 
each  polygon,  but  not  the  identity  of  the  individual 
layers.  This  information  is  sufficient  to  calculate  the 
HLI. 


The  determination  of  the  areas  of  the  tree  can- 
opy, tree  bole,  shrub  midstory,  understory,  subsur- 
face layer  suitable  for  the  establishment  of  burrows, 
and  the  water  surface  (if  it  exists)  on  a  study  area 
constitutes  an  inventory  of  habitat  resources  on  that 
study  area.  Results  of  such  an  inventory  can  be  ex- 
pressed in  terms  of  hectares  (acres)  of  each  habitat 
layer  within  the  study  area. 


Results  of  this  type  of  inventory  for  segments  A 
and  B  of  the  habitat  illustrated  in  Figure  5  are  pre- 
sented in  Table  5.  The  inventory  summarized  in 
Table  5  was  accomplished  by  determining  the  layers 
of  habitat  present  in  the  different  cover  types. 


The  formula  for  calculating  the  HLI  for  a  study 
area  is — 


HLI  = 


HLI    = 


i=l 


(6)  (5) 


n 
i  =  i 


where: 

X  =   the  number  of  layers  of  habitat  present 
within  some  bounded  area, 

Ai   =   the  area  of  layer  of  habitat  i  within  the 
bounded  area, 

Aj   =   the  surface  area  of  cover  type  j  within  the 
bounded  area, 

n   =   the  number  of  different  cover  types  present 
within  the  bounded  area, 

6   =   the  maximum  number  of  habitat  layers  that 
could  occur  in  a  unit  of  structurally  com- 
plex terrestrial  habitat,  and 

5    =    the  maximum  number  of  units  of  area  of 

habitat  layers  that  could  occur  within  a  unit 
of  structurally  complex  terrestrial  habitat. 

The  calculation  of  HLIs  for  segments  A  and  B  of 
Figure  5  is  presented  below.  There  are  four  layers 
of  habitat  present  in  segment  A  (terrestrial  subsur- 
face, understory,  midstory,  and  tree  bole);  the  sum 
of  the  area  of  the  layers  of  habitat  equals  836.7  ha 
(2091.8  a.)  (Table  5).  The  denominator  equals  six 
layers  of  habitat  and  5  x  281.1  ha  (702.8  a.)  of  habi- 
tat layers. 


number  of  actual  layers  actually  present  x 

actual  area  of  those  habitat  layers 

number  of  habitat  layers  potentially  present 

x 

potential  area  of  those  habitat  layers 


4  x  2091.8 
6  x  5  X  702.8 


=  0.397  or  39.7% 


Segment  A  contains  a  structural  diversity  measure  of 
39.7%.  This  HLI  can  be  compared  with  that  of  other 
land  units  to  compare  the  structural  diversity  of 
different  study  areas. 

The  HLI  for  segment  B  in  Figure  5  is  calculated 
in  a  similar  manner.  The  same  four  layers  of  habitat 
are  present  in  B  as  in  A;  the  area  of  the  habitat  layers 
in  B  equals  732.7  ha  (18317  a.)  (Table  5).  The  de- 
nominator equals  six  layers  of  habitat  and  5  x  269.4  ha 
(673.6  a.). 


HLI  = 


4  x   1831-7 
6  x  5  x  673.6 


0.362  or  36.2% 


The  HLI  for  segment  B  is  reduced  because  a 
substantial  block  of  sagebrush  habitat  was  cleared  to 
produce  a  dry  land  pasture  (polygon  212  in  section 
B  of  Figure  4 ).  The  loss  of  sagebrush,  which  pro- 
vided a  midstory  layer  on  this  land  unit,  made  this 
polygon  unsuitable  for  species  with  obligate  ties 
to  sagebrush  and  for  species  requiring  a  midstory 
structure. 


Habitat  Layers — Wildlife  Guilds 

The  inventory  and  assessment  of  the  structure 
of  habitat  is  one  level  of  range  analysis;  the  determi- 
nation of  the  quantity  of  particular  habitats  suitable 
for  particular  wildlife  guilds  is  a  second  range  evalua- 
tion process.  The  map  in  Figure  5  shows  where  two, 
three,  or  four  habitat  layers  occur  on  the  study  area 
in  the  Piceance  Basin,  although  the  particular  habitat 
layers  present  in  any  one  polygon  are  not  identified. 
Five  combinations  of  habitat  layers  actually  occur 
in  the  mapped  area.  Polygons  identified  as  212  in 
Figure  4  provide  subsurface  and  understory  layers  of 
habitat  (Table  5).  Polygons  identified  as  522  SG  pro- 
vide subsurface  and  midstory  layers;  polygons  identi- 
fied as  523  SGHR  and  522  SGHR  provide  subsurface, 
understory,  and  midstory  layers.  Polygons  identified 
as  423  PJBS  provide  subsurface,  midstory,  and  tree 
bole  layers;  and  polygons  identified  as  423  PJSG  pro- 
vide subsurface,  understory,  and  midstory,  and  tree 
bole  layers  of  habitat.  These  different  habitat  struc- 
tures may  vary  in  their  suitability  for  wildlife  species. 

The  location  of  each  of  the  different  habitat 
structures  can  be  represented  on  a  map.  Data  for 
drawing  such  maps  can  be  obtained  from  terrestrial 


108 


Rangelands 


Table  5.     Areas  (acres)  of  each  cover  type  and  individual  layers  of  habitat  present  in  habitat  segments  A  and 
B  of  Figure  5. 


Segment 

Cover  type 
(Figure  4) 

Individual  layers  of  habitat  present 

Total 

Sub- 
surface 

Under- 
story 

Mid- 
story 

Tree 
bole 

Total  area 
of  layers 

A 

212 

523  SGHR 
423  PJSG 
522  SGHR 
522  SG 
423  PJBS 

5.3 
352.9 
144.9 

22.0 
156.2 

21.5 

5.3 
352.9 
144.9 

22.0 
156.2 

21.5 

5.3 

352.9 

144.9 

22.0 

352.9 
144.9 

22.0 
156.2 

21.5 

144.9 
21.5 

10.6 

1058.7 

579.7 

66.0 
312.4 

64.5 

Total 

702.8 

702.8 

525.1 

697.5 

166.4 

2091.8 

B 

212 

523  SGHR 
423  PJSG 
522  SGHR 
522  SG 
423  PJBS 

139.6 

328.9 

48.5 

0.0 

98.0 

58.6 

139.6 

328.9 
48.5 

98.0 
58.6 

139.6 

328.9 

48.5 

328.9 
48.5 

98.0 
58.6 

48.5 
58.6 

279.2 
986.7 
194.0 
0.0 
196.0 
175.8 

Total 

673.6 

673.6 

517.0 

534.0 

107.1 

1831.7 

The  numerator  of  the  Habitat  Layer  Index  is  calculated  by  determining  the  product  of  the  number  of  habitat  layers  actually 
present  in  a  study  area,  like  that  in  Figure  5,  and  the  area  of  the  individual  habitat  layers  present  on  that  area.  The  denominator  is  the 
hypothetical  value  that  would  be  realized  if  all  six  layers  of  habitat  (water  surface,  terrestrial  subsurface,  terrestrial  surface,  midstory, 
tree  bole,  and  tree  canopy)  occurred  on  the  study  area  and  five  of  these  layers  extended  throughout  the  study  area 


surveys  of  the  structure  of  cover  types  or  from  in- 
terpretive aerial  photography  with  ground  truthing. 
Each  polygon  in  the  map  of  habitat  structure  within 
the  Piceance  Basin  study  area  ( Figure  6 )  provides 
subsurface,  surface,  midstory,  and  tree  bole  layers  of 
habitat.  Wildlife  species  can  be  associated  with  the 
structure  of  habitats  like  that  represented  in  Figure  6 
through  the  formation  of  wildlife  guilds  (described 
below).  Maps  of  habitat  structure  and  lists  of  wildlife 
guilds  dependent  on  that  habitat  structure  will  show, 
for  example,  that  the  notated  polygons  in  Figure  6 
constitute  a  distribution  map  of  suitable  habitat  for 
cavity-using  species  of  wildlife  in  this  study  area.  The 
habitat  layer-wildlife  guild  maps  help  explain  the 
distribution  of  wildlife  species  throughout  the  range- 
land  and  provide  a  rationale  for  inventorying  habitat 
conditions  suitable  for  groups  of  wildlife  species. 

The  association  of  wildlife  species  with  polygons 
of  habitat  can  be  accomplished  with  the  guilding 
procedure  designed  by  Short  and  Burnham  ( 1982), 
as  modified  by  Short  (1983).  The  vertical  layers  of 
habitat  are  represented  in  that  procedure  as  axes  of  a 
two-dimensional  matrix.  The  x-axis  of  the  matrix 
represents  layers  of  habitat  where  nesting,  birthing, 
or  hatching  occur;  the  y-axis  of  the  matrix  repre- 
sents habitat  layers  where  wildlife  species  forage  for 
food.  This  type  of  "species-habitat"  matrix  is  illus- 
trated in  Figure  7.  The  y-axis  of  the  matrix  is  divided 
so  that  layers  where  primary  consumption  (plant 
materials)  occurs  can  be  differentiated  from  habitat 


layers  where  secondary  consumption  ( animal  mate- 
rials) occurs.  The  process  described  below  produces 
both  guilds  of  primary  consumers  and  guilds  of  sec- 
ondary consumers.  Omnivorous  species  occupy  two 
guilds,  one  as  a  primary  consumer  and  one  as  a  sec- 
ondary consumer. 

The  blocks  formed  by  the  intersection  of  the 
lines  that  differentiate  the  layers  of  habitat  on  the  x- 
and  y-axes  of  the  matrix  are  called  "guild  blocks." 
The  species-habitat  matrices  for  the  variety  of  range- 
land  habitats  are  different  because  the  number  of 
guild  blocks  actually  available  for  use  by  members  of 
the  wildlife  community  varies  among  range  cover 
types.  The  guild  blocks  available  to  wildlife  in  the 
three  basic  range  cover  types  are  indicated  in  Figure 
7  for  grassland,  shrubland,  and  woodland  habitats. 
Woodland  habitats  obviously  provide  more  guild 
blocks  than  a  shrubland  habitat  which,  in  turn,  pro- 
vides more  guild  blocks  than  a  grassland  habitat. 

Guild  blocks  represent  the  different  ways  in 
which  wildlife  species  utilize  the  resources  of  habi- 
tats. For  example,  wildlife  species  in  grassland  habi- 
tats (Figure  7)  can  breed  in  the  terrestrial  subsurface 
layer  and  forage  in  the  subsurface  layer,  on  the  ter- 
restrial surface,  or  in  the  air.  Or  they  can  breed  on 
the  terrestrial  surface  and  feed  underground,  on  the 
terrestrial  surface,  or  in  the  air.  Wildlife  species  can 
breed  in  other  habitats  and  forage  in  the  subsurface 
layer,  on  the  terrestrial  surface,  or  in  the  air  above 


Rangelands 


109 


Figure  6.     Map  of  segments  A  and  B  of  the  Piceance  Basin  study  area  showing  polygons  containing  subsur- 
face, surface,  midstory,  and  tree  bole  layers  of  habitat. 


the  grassland.  Shrubland  and  woodland  communities 
are  more  complex  because  more  layers  of  habitat  are 
available  for  the  location  of  niche  spaces  for  wildlife 
species. 

Groups  or  guilds  of  wildlife  species  are  devel- 
oped from  a  species-habitat  data  base.  The  data  base 
describes  the  layers  of  habitat  where  breeding  (nest- 
ing, birthing,  or  hatching)  occurs,  as  well  as  the  lay- 
ers of  habitat  where  foraging  occurs,  for  each 
wildlife  species  potentially  occurring  in  a  study  area. 
The  information  within  the  wildlife  species  data  base 
is  processed  (Short  1983)  as  follows:  the  numerical 
codes  that  indicate  the  layers  of  habitat  used  for 
breeding  and  for  feeding  (listed  along  the  x-  and  y- 
axes  of  Figure  7)  are  entered  into  a  computer  for 
each  vertebrate  wildlife  species  inhabiting  a  cover 
type.  These  coded  data  are  subjected  to  standard 
"merge-sort"  routines  that  produce  lists  of  species 
with  dependencies  on  the  same  guild  block  or  the 


same  group  of  guild  blocks.  An  example  of  the  wild- 
life guilds  produced  with  this  technique  is  provided 
in  Table  6,  which  lists  the  guilds  of  primary  con- 
sumers occupying  chaparral  rangelands  habitats  in 
the  Hualapai-Aquarius  Planning  Units  in  west-central 
Arizona.  This  list  indicates  the  dependency  of  the 
wildlife  community  on  the  structure  of  the  chaparral 
habitat  in  this  study  area. 


Wildlife  guilds  occurring  in  pinyon-juniper  habi- 
tats in  the  Piceance  Basin  could  be  developed  in 
the  same  way  that  guilds  were  developed  in  Table  6 
and  associated  with  a  map  of  specific  layers  of  habi- 
tat like  that  in  Figure  6.  Wildlife  species  requiring 
understory  and  midstory  vegetation  could  occur  in 
the  polygons  in  Figure  6,  as  well  as  in  the  polygons 
with  understory  and  midstory  vegetation.  The  poly- 
gons listed  in  Figure  6,  however,  would  be  the  only 
habitat  suitable  for  cavity-making  and  cavity-using 


110 


Rangelands 


10.  Feeds  elsewhere 

G  W 

G  W 

W 

w 

G  W 

9.  Air 

S 

S 

s 

S 

8.  Tree  canopy 

W 

w 

W 

w 

W 

7.  Tree  bole 

E 

CO 

w 

w 

w 

w 

w 

c 
o 

6.  Shrub  layer 

s 

s 

s 

s 

G  W 

G  W 

w 

W 

G  W 

c 
o 

5.  Terr,  surface 

S 

S 

s 

S 

o 

CD 
C/3 

4.  Terr,  subsurface 

G  W 
S 

G  W 
S 

G 

s 

3.  Water  surface 

2.  Water  column 

o 
o 

1 .  Bottom  water  column 

c 

T3 

LL 

8.  Tree  canopy 

w 

W 

w 

w 

w 

7.  Tree  bole 

w 

W 

w 

w 

w 

CO 

6.  Shrub  layer 

s 

S 

s 

s 

CD 

E 

G  W 

G  W 

w 

w 

G  W 

CO 

5.  Terr,  surface 

c 

S 

S 

s 

s 

o 
o 

CO 

4.  Terr,  subsurface 

G  W 
S 

G  W 
S 

G 
S 

E 

a. 

3.  Water  surface 

2.  Water  column 

1 .  Bottom  water  column 

CO 

CD 

o 

C 

E 

CD 

o 

3 

o 

CO 

O 

o 

m 

CD 

CD 

Si 
co 

S 

CO 

c 
E 

CD 
O 
CO 

3 

CO 
CO 

O 

ca 

CD 

Q. 

CD 

5 

CD 

CO 

CO 

o 

o 

=J 

CO 

CO 

_co 
n 

3 

CD 

O 

CD 

o 

Q. 

E 
o 

o 

CD 

CO 
CD 

to 
CD 

to 
CD 

o 

CD 

CO 

o 

CD 

CO 

"D 
CD 

CD 

I— 

o 

CD 

CO 

5 

CO 

CD 

1— 

CD 

1— 

co 

CD 

CD 

h1 

CD 
CO 

1. 

2. 

3. 

4. 

5. 

6. 

7. 

8. 

9. 

10. 

Breedi 

ng  loc 

i 

Figure  7.     Form  of  the  species-habitat  matrix  used  in  the  development  of  wildlife  guilds,  and  use  of  guild 
blocks  by  wildlife  species  in  grassland  (G),  shrubland  (S),  and  woodland  (W)  habitats. 


Rangelands 


111 


Table  6.     Wildlife  guilds  of  primary  consumers  in  chaparral  habitats  of  west-central  Arizona  (from  Short 
1983). 


Guild 
No. 

Guild 
members 

Feeding  loci 

Breeding  loci 

CD 
O 

re 

=3 
CO 

n 

z> 

CO 

H 

to 
CD 

53 
1— 

CD 

o 
m 

D 

go 
to 

to 
CD 

CD 

CD 
>. 

_C0 

>. 
O 
CO 

"O 

E 
o 
n 

SZ 

C/) 

CD 
O 

n 

CD 
CD 

h1 

CL 

o 

c 

CO 

o 

CD 
CD 

< 

CD 
CD 

_C 

S 

CD 
CO 
CD 
00 

"O 
CD 
CD 

m 

CD 
O 

ra 

=i 

CO 

n 

CO 

to 

CD 

CD 

H 

CD 
O 
cfl 

3 
CO 

to 
to 

CD 

CD 

1— 

CD 

a 
>. 

o 

to 
"D 

E 
o 
n 

sz 
CO 

CD 
O 

n 

CD 
CD 

h1 

>s 

CL 

o 

c 

CO 

o 

CD 
CD 

1 

Black  bear 
Collared  peccary 
Deer  mouse 
Cactus  mouse 
Coyote 

• 
• 
• 
• 
• 

2 

Striped  skunk 
Hog-nosed  skunk 

• 
• 

3 

Botta's  pocket 
gopher 

• 

• 

4 

White-throated 
wood  rat 

• 

• 

• 

• 

5 

Bighorn  sheep 
Gray  fox 

Stephen's  woodrat 
Ringtail 
Brush  mouse 

• 
• 
• 
• 
• 

6 

Harris'  antelope 
squirrel 

• 

7 

Brown-headed 

cowbird 
House  finch 
Mockingbird 
Cactus  wren 
Costa's  hummingbird 
Western  harvest 

mouse 
Crissal  thrasher 

• 

8 

Mule  deer 
Rock  squirrel 
Cattle 

9 

Scrub  jay 
Bushtit 

• 
• 

10 

Scott's  oriole 
American  robin 
Lesser  goldfinch 
Townsend's  solitaire 
Anna's  hummingbird 
Lewis'  woodpecker 
Blue  grosbeak 
Black-chinned 
hummingbird 
Acorn  woodpecker 

* 

• 

112 


Rangelands 


Table  6.     Wildlife  guilds  of  primary  consumers  in  chaparral  habitats  of  west-central  Arizona  (from  Short 
1983)  (continued). 


Guild 
No. 

Guild 
members 

Feeding  loci 

Breeding  loci 

CD 
O 

as 

CO 
-O 

13 
CO 

to 
CD 

q3 

o 
to 

CO 

to 
CD 

CD 
h- 

CD 
03 
>. 
O 

00 

"O 

E 

o 
n 

co 

CD 
O 
JD 

CD 
CD 

h1 

o. 
o 

c 

<T3 
O 

CD 
CD 

h1 

< 

CD 

q3 
.c 
5 

CD 

oo 

CD 
c/) 
"O 
CD 
CD 

m 

<D 
O 

m 

=j 

CO 

n 

00 

to 

CD 

CD 

(— 

CD 
O 
CD 

=3 
00 

CO 
CD 

o3 

1— 

CD 
>. 

a 

>, 

o 

to 
"D 

E 
o 

-Q 

=3 

.C 

CO 

CD 

o 
n 

CD 
CD 

h1 

Q. 

o 

c 

o 

CD 
CD 

H 

10 
(cont.) 

Broad-tailed 

hummingbird 
Cassin's  kingbird 
Mountain  bluebird 
Rufous  hummingbird 
Starling 
Pinyon  jay 

Calliope  hummingbird 
Green-tailed  towhee 
Black-headed 

grosbeak 
American  goldfinch 
Common  flicker 
Steller's  jay 
Ladder-backed 

woodpecker 
Lazuli  bunting 
Yellow-rumped  warbler 
Western  kingbird 
Swainson's  thrush 
Elk 

Hermit  thrush 
White-winged  dove 
Western  bluebird 

11 

Desert  spiny  lizard 
Canyon  mouse 
Western  spotted 

skunk 
Cliff  chipmunk 
Arizona  woodrat 

12 

Southern  grass- 
hopper mouse 

Ord's  kangaroo  rat 

Northern  grass- 
hopper mouse 

Desert  tortoise 

13 

Mourning  dove 

• 

14 

Rufous-sided 
towhee 

Eastern  cottontail 

Black-tailed  jack- 
rabbit 

Common  raven 

Rock  dove 

Gambel's  quail 

House  mouse 

Rangelands 


113 


Table  6.     Wildlife  guilds  of  primary  consumers  in  chaparral  habitats  of  west-central  Arizona  (from  Short 
1983)  (concluded). 


Guild 
No. 

Guild 
members 

Feeding  loci 

Breeding  loci 

CD 
O 

ro 

CO 

n 

CD 

q3 

(— 

CD 
O 

CO 

to 

CD 

CD 

I— 

CD 
>. 

ro 
>~ 
o 

CO 

"O 
E 

o 
n 

sz 

CD 
O 

n 

CD 
CD 

H 

D. 
O 
C 
CO 
o 

CD 
CD 

H 

< 

CD 
CD 

sz 

CD 
CO 
CD 
CO 

-a 
o 

CD 

CD 
O 
03 

(/> 

n 

zj 

CO 

03 

CO 

CD 

CD 

1— 

CD 
U 

■2 

CO 

to 

CD 

q3 

l— 

CD 

m 
>. 
o 

CO 

E 

o 
n 

x: 

CD 
O 
_Q 

CD 
CD 

H1 

>. 

Q. 
O 

c 

CO 

o 

CD 
CD 

h1 

14 
(cont.) 

Desert  cottontail 
Rofous-crowned 
sparrow 

• 
• 

• 
• 

15 

Brown  towhee 
Black-chinned 

sparrow 
Black-throated 

sparrow 

• 
• 

• 

• 
• 

• 

16 

Western  meadowlark 
Lark  sparrow 
Chipping  sparrow 
Fox  sparrow 
Purple  finch 
Brewer's  sparrow 
Dark-eyed  junco 
White-crowned  sparrow 
Water  pipit 

• 
• 
• 
• 
• 
• 
• 
• 
• 

17 

Say's  phoebe 

• 

• 

18 

Verdin 

• 

19 

Hooded  oriole 
Warbling  vireo 
Phainopepla 
Ruby-crowned  kinglet 
Plain  titmouse 
Ash-throated  flycatcher 
Wied's  crested 

flycatcher 
Northern  oriole 
Yellow-bellied 

sapsucker 
Hermit  warbler 

114 


Rangelands 


species  that  require  the  tree  bole  layer  habitat.  The 
map  in  Figure  6  would  identify  the  distribution  of 
habitat  favorable  to  this  group  of  species.  The  land 
units  containing  particular  layers  of  habitat  can  be 
surveyed  to  determine  the  quantity  of  habitat  that 
exists  for  specialized  wildlife  species  or  groups  of 
wildlife  species.  For  example,  66.6  ha  ( 166.4  a.)  of 
habitat  containing  tree  boles  suitable  for  cavity  nest- 
ing species  occurred  in  Segment  A,  and  42.8  ha 
(107.1  a.)  of  this  habitat  occurred  in  Segment  B  of 
Figure  6  (Table  5).  The  chaining  of  pinyon- juniper 
on  Segments  A  or  B  would  reduce  the  quantity  of 
the  tree  bole  layer  and  adversely  affect  wildlife 
guilds  having  a  dependency  on  the  tree  bole  layer. 


Use  of  Habitat  Layers  by  Individual  Species 

Maps  of  habitat  layers,  coupled  with  lists  of  the 
wildlife  guilds  that  use  those  habitat  layers,  provide 
one  way  to  evaluate  a  habitat  cover  type.  Approxi- 
mating where  a  niche  space  of  a  species  occurs  in  a 
layer  of  habitat  is  a  second  appraisal  technique,  al- 
though a  more  arduous  one.  This  technique,  how- 
ever, yields  an  estimate  of  the  utility  of  a  range 
habitat  for  a  particular  wildlife  species. 

A  layer  of  habitat  can  vary  greatly  in  its  compo- 
sition and  utility  to  wildlife.  For  example,  the  terres- 
trial surface  layer  can  provide  a  variety  of  structural 
features  for  wildlife  (listed  under  code  6  in  Table  7). 
The  understory  can  ( 1 )  be  devoid  of  cover  on  salt 
playas  and  some  beaches,  (2)  have  the  appearance  of 
desert  pavement,  (3)  contain  rocks  or  be  a  rock 
surface  in  talus  fields  and  on  cliffs,  (4)  be  covered 
with  tree  litter  under  a  forest  overstory,  and  (  5  )  be 
covered  with  grasses,  forbs,  and  dwarf  shrubs  of 
different  densities  and  heights.  The  large  array  of 
structured  conditions  along  the  terrestrial  surface 
layer  can  be  considered  as  constituting  a  gradient  of 
conditions  for  wildlife  species  having  a  dependency 
on  this  layer  of  habitat.  Range  surveys  that  describe 


surface  cover,  plant  cover  type,  and  plant  vigor  pro- 
vide information  that  helps  position  a  study  area 
along  such  a  terrestrial  surface  gradient. 

The  terrestrial  surface  layer  in  the  native  grass- 
lands of  North  America  provides  a  variety  of  struc- 
tures corresponding  to  those  listed  in  subunits  6.14 
to  6.19  of  Table  7.  The  grasslands  might  vary  from 
bunchgrasses,  providing  low,  scattered  herbaceous 
cover,  to  sod  grasses,  providing  a  tall,  dense  herba- 
ceous cover.  Wildlife  species  may  generally  select  a 
particular  habitat  layer  (e.g.,  terrestrial  surface  layer) 
and  a  particular  structure  of  that  habitat  layer  (e.g., 
shortgrass  prairie  with  sparse  cover)  and  then  fine 
tune  their  habitat  selection  on  criteria  like  percent- 
age bare  ground  or  density  and  height  of  particular 
plant  species. 

Thus,  the  greatest  abundance  of  wildlife  species 
A  may  occur  at  point  X  along  the  terrestrial  surface 
habitat  gradient,  whereas  the  greatest  abundance 
of  wildlife  species  B  may  occur  at  point  Y  along  the 
gradient. 

A  hypothetical  habitat  layer  gradient  for  grass- 
lands is  presented  in  Figure  8.  Horned  larks  (Eremo- 
phila  alpestris)  occupy  a  position  along  the  habitat 
gradient  characterized  as  open  shortgrass  prairie; 
western  meadowlarks  occupy  a  position  character- 
ized by  medium-dense  mid  to  tall  grasses;  and  bobo- 
links (Dolichonyx  oryzivorus )  occupy  a  position 
characterized  by  dense,  tall  stands  of  grasses,  weeds, 
or  alfalfa.  These  three  species  might  belong  to  the 
same  grassland  wildlife  guild,  even  though  their  indi- 
vidual presence  in  a  survey  would  be  dependent  on 
the  nature  of  the  herbaceous  cover  in  the  grassland 
being  inventoried. 

Individual  wildlife  species  can  be  approximately 
positioned  within  a  gradient  developed  for  a  layer 
of  habitat,  if  simple  word  models  or  HSI  models  (U.S. 
Fish  and  Wildlife  Service  1980)  are  constructed. 


Low,  scattered,  herbaceous  cover    •*■ 


-►   High,  dense,  herbaceous  cover 


Figure  8.     Hypothetical  habitat  gradient  that  illustrates  how  a  variety  of  birds  select  grasslands  with  slightly 
different  structures. 


Rangelands 


115 


Table  7.     Habitat  conditions  that  help  describe  gradients  for  the  terrestrial  subsurface,  terrestrial  surface, 

midstory,  tree  bole,  and  tree  canopy  layers  of  habitat.  The  codes  5—9  cross-reference  to  the  identification 
of  habitat  layers  listed  along  the  x-axes  of  the  species-habitat  matrices  in  Figure  7. 


5.  Terrestrial  subsurface  (to  10  cm  [4  in.]  below  apparent  surface) 


5  01  Rocky  substrate  unsuitable  for  burrow  or  tunnel  construction 

5.02  Hydric  or  heavy  soils  unsuitable  for  burrow  or  tunnel  construction 

5.03  Powdery  soils  unsuitable  for  burrow  or  tunnel  construction 

5.04  Soils  suitable  for  burrow  or  tunnel  construction 

5.05  Cave  and  deep  crevice 

5.06  Artificial  (man-made)  structures,  such  as  mine  shafts  and  out  buildings,  where  interior  use  is  similar  to  that  of  burrows  or 
caves 


6.  Terrestrial  surface — Understory  layer  (from  10  cm  [4  in]  below  to  0.15  m  [6  in]  above  apparent  surface) 


6.01  Salt  playas  or  flats  with  hydric  soils 

6.02  Beaches  (mud,  sand,  or  rock)  without  hydrophytes 

6.03  Marshy  areas  with  hydrophytes  but  not  hydric  soils 

6.04  Bare  ground  (sand  to  rubble,  up  to  305  mm  [12  in]  particles) 

6.05  Boulder  (>  305  mm  [12  in.]) — covered  surface 

6.06  Talus — unvegetated 

6.07  Talus — vegetated 

6.08  Cliff — on  ledge  near  valley  floor 

6.09  Cliff — in  cavity  near  valley  floor 

6.10  Cliff — on  ledge  near  mesa  or  mountaintop 
6  1 1  Cliff — in  cavity  near  mesa  or  mountaintop 
6  12  Herbaceous  litter 

6  13  Woody  litter  (includes  shrub  branches,  tree  branches,  and  stumps) 

6.14  Shortgrass  prairie  vegetation  with  sparse  (<  50%)  cover 

6.15  Shortgrass  prairie  vegetation  with  open  or  closed  (>  50%)  cover 
6  16  Midgrass  prairie  vegetation  with  sparse  (<  50%)  cover 

6  17  Midgrass  prairie  vegatation  with  open  or  closed  (>  50%)  cover 
6  18  Tallgrass  prairie  vegetation  with  sparse  (<  50%)  cover 

6.19  Tallgrass  prairie  vegetation  with  open  or  closed  (>  50%)  cover 

6.20  Supine  or  dwarf  woody  vegetation  or  woody  vegetation  within  0.5  m  (19  in.)  of  apparent  surface 

6.21  Cactus  stems  and  pads 

6.22  Artificial  (man-made)  structures — ground  debris  and  artifacts,  bridges,  trestles,  and  rooftops  where  external  use  is  analo- 
gous to  that  of  the  horizontal  and  vertical  surface  of  natural  objects 


7.  Midstory  layer  (from  0.5  to  8.0  m  [  19  to  26  ft]  above  apparent  surface) 


7  01  Canopy  of  deciduous  shrubs  or  small  trees  (<  50%  cover) 

7.02  Canopy  of  deciduous  shrubs  or  small  trees  (>  50%  cover) 

7.03  Canopy  of  evergreen  shrubs  or  small  trees  (<  50%  cover) 

7.04  Canopy  of  evergreen  shrubs  or  small  trees  (>  50%  cover) 

7.05  Grass  and  grasslike  vegetation  (includes  bamboo)  extending  into  midstory  layer 

7.06  Forb  vegetation  extending  into  midstory  layer 

7.07  Cactus  stems  and  pads  extending  into  midstory  layer 

7.08  Artificial  (man-made)  structures  extending  into  midstory  layer  and  used  in  a  manner  analogous  to  that  of  natural  vegeta- 
tion 


8.  Tree  bole  layer 


8.01  Snag — trunk  of  dead  deciduous  tree 

8.02  Snag — trunk  of  dead  evergreen  tree 

8.03  Trunk — living  deciduous  tree 

8.04  Trunk — living  evergreen  tree 
8  05  Cactus  bole  or  stem 

8.06  Artificial  (man-made)  structure — telephone  and  power  poles,  chimneys,  nest  boxes 


9.  Tree  canopy 


9.01  Small  branches— live  deciduous  trees  (<  50%  cover) 

9.02  Small  branches — live  deciduous  trees  (>  50%  cover) 

9.03  Small  branches — live  evergreen  trees  (<  50%  cover) 

9.04  Small  branches — live  evergreen  trees  (>  50%  cover) 

9.05  Large  branches — live  deciduous  trees  (<  50%  cover) 

9.06  Large  branches — live  deciduous  trees  (>  50%  cover) 

9.07  Large  branches — live  evergreen  trees  (<  50%  cover) 

9.08  Large  branches — live  evergreen  trees  (>  50%  cover) 

9.09  Large  branches — dead  deciduous  trees 

9.10  Large  branches — dead  evergreen  trees 


116  Rangelands 


Screech  owl  (Otus  kennicottii) 


1.  Cover  types  inhabited 


Northern  floodplain  forest 
Pinyon-juniper  woodland 
Eastern  ponderosa  forest 
Black  Hills  pine  forest 
Pine-Douglas  fir  forest 
Grama- buffalograss 
Wheatgrass-needlegrass 
Bluestem-grama  prairie 
Sandsage-bluestem  prairie 
Grama-needlegrass-wheatgrass 


2.  Habitat  layers  where  nesting,  birthing,  or  hatching  occur 


Tree  bole  layer 

Substrate  8.01,  8.02,  8.03,  8.04,  8.06  (from  Table  8).  Nest  is  nearly  always  in  a  woodpecker/flicker  hole  in  a  Cottonwood  or 
other  large  deciduous  tree.  Occasionally  found  in  cavity  in  pinyons  and  junipers. 


3.  Habitat  layers  where  foraging  occurs 


Secondary  consumers 

Terrestrial  surface  layer:  feeds  on  rodents,  small  birds,  amphibians,  and  reptiles 


4.  Species-habitat  model  (references) 


Requires  dense  cottonwoods  or  other  deciduous 
woodlots  along  watercourses  from  lower  moun- 
tains throughout  plains.  May  select  other  trees, 
including  isolated  cottonwoods,  in  a  variety  of 
range  habitats. 


Cottonwoods  or  other  trees  are  at  least  20  cm 
(8  in  )  dbh. 


■No- 


Yes 


Trees  are  of  negligible  value  as  nesting  habitat 
for  the  screech  owl. 


Trees  possess  woodpecker/flicker  holes  suitable 
for  use  as  nest  sites. 


■No- 


Yes 


Trees  may  represent  useful  nesting  habitat  for 
the  screech  owl. 


Figure  9.     Information  to  approximate  conditions  along  a  layer  of  habitat  useful  to  the  screech  owl  (Otus 
kennicottii)  in  Wyoming  and  Colorado  (after  U.S.  Forest  Service  1981 ). 


Rangelands 


117 


Least  chipmunk  (Tamias  minimus) 


1.  Cover  types  inhabited 


Mountain  mahogany-oak  scrub 
Sagebrush  steppe 
Eastern  ponderosa  pine 
Black  Hill  pine  forest 
Pine-Douglas  fir  forest 
Pinyon-juniper  woodland 
Great  Basin  sagebrush 
Douglas  fir  forest 
Saltbush-greasewood 


2.  Habitat  layers  where  nesting,  birthing,  or  hatching  occur 


Terrestrial  subsurface  layer — Substrate  5.04 

Terrestrial  surface  layer — Substrate-6  06,  6.07,  6.13,  6.22  (from  Table  8)  Nests  are  found  under  stumps,  logs,  or  rocks. 


3.  Habitat  layers  where  foraging  occurs 


Primary  consumer 

Terrestrial  surface  layer:  consumes  fruit,  nuts,  seeds,  berries,  mushrooms,  leaves,  stems. 

Secondary  consumer 

Terrestrial  surface  layer:  consumes  invertebrates,  especially  insects 


4.  Species-habitat  model 


Uses  surface  layer  within  a  variety  of  cover 
types,  including  semiarid  sagebrush  and  grease- 
wood  cover  types,  browse  slopes,  brushy 
ravines,  and  coniferous  mountain  forests. 


Yes 

_L 


Terrestrial  surface  has  sparse  (20  to  50%)  cover. 


■No 


Yes 


Terrestrial  surface  provides  stumps,  logs,  rocks, 
and  roots  under  which  chipmunks  can  burrow. 


Yes 


Understory,  midstory,  and/or  overstory  plant 
species  provide  a  source  of  fruit,  nuts,  berries, 
seeds,  and  leaves  and  the  terrestrial  surface 
layer  seems  capable  of  providing  a  source  of 
invertebrates  as  food  sources. 


Habitat  is  of  negligible  value  to  the  least  chip- 
munk. 


■No 


Yes 


Cover  type  may  represent  useful  habitat  to  the 
least  chipmunk. 


Figure  10.     Information  to  approximate  conditions  along  a  layer  of  habitat  useful  to  the  least  chipmunk 
(Tamias  minimus)  in  Wyoming  and  Colorado  (after  U.S.  Forest  Service  1981). 


118 


Rangelands 


Brewer's  sparrow  (SpizeUa  brewer!) 


1.  Cover  types  inhabited 


Grama-buffalo  grass 
Sandsage-bluestem  prairie 
Mountain  mahogany-oak  scrub 
Great  Basin  sagebrush 
Wheatgrass-needlegrass  shrub  steppe 
Wheatgrass-needlegrass 
Grama-needlegrass-wheatgrass 
Saltbush-greasewood 
Sagebrush  steppe 


2.  Habitat  layers  where  nesting,  birthing,  or  hatching  occur 


Terrestrial  surface  layers 

Substrate  6.20,  7.01 ,  7.02,  7.03,  7.04  (from  Table  8).  Brewer's  sparrows  frequently  construct  their  nest  about  15  to  25  cm  (6 
to  10  in.)  above  ground  in  dense  woody  shrubs,  especially  sagebrush. 


3.  Habitat  layers  where  foraging  occurs 


Primary  consumer:  terrestrial  surface  layer.  The  Brewer's  sparrow  feeds  on  grass  and  seeds  on  the  ground. 

Secondary  consumer:  terrestrial  surface  layer.  The  Brewer's  sparrow  gleans  insects  from  the  ground  and  from  low  shrubs. 
Midstory  layer:  the  Brewer's  sparrow  gleans  insects  from  large  shrubs. 


4.  Species-habitat  model  (references) 


Living  woody  shrub  vegetation  within  the  breed- 
ing range  of  the  Brewer's  sparrow  is  present  in  a 
habitat  block  greater  than  or  equal  to  0.2  ha 
(0.5  a.)  in  area 


No 


Yes 
t 


Living  woody  shrub  vegetation  occurs  on  terrain 
whose  slope  does  not  exceed  30°. 


Habitat  block  is  of  negligible  value  as  breeding 
habitat  for  the  Brewer's  sparrow. 


■No 


Yes 


Living  woody  shrub  vegetation  occurs  on  gener- 
ally well-drained  soils  rather  than  on  rock  out- 
crops. 


■No 


Yes 


Habitat  may  be  of  high  quality  if  woody  shrub 
species  are  50-70  cm  (20-28  in.)  in  height,  pro- 
vide a  dense  limb  structure  25  cm  (10  in.)  above 
ground,  and  provide  a  shrub  canopy  30%.  Habi- 
tat may  be  of  low  to  moderate  quality  if  these 
conditions  do  not  occur. 


Yes 


Habitat  may  be  of  high  quality  as  nesting  habitat 
to  the  Brewer's  sparrow  if  woody  shrubs  are 
sagebrush;  habitat  may  be  of  medium  quality  if 
woody  shrubs  are  hawthorn,  plum,  serviceberry, 
bitterbrush,  or  rabbitbrush,  habitat  may  be  of  low 
quality  if  shrubs  are  saltbush,  greasewood,  hop- 
sage,  ceanothus,  manzanita,  currant,  or  other. 


Figure  11.     Information  to  approximate  conditions  along  a  layer  of  habitat  useful  to  the  Brewer's  sparrow  in 
(Spizella  breweri)  Wyoming  and  Colorado  (after  U.S.  Forest  Service  1981). 


Rangelands 


119 


The  intent  of  either  model  is  to  describe  habitat  con- 
ditions favorable  to  the  species.  Word  models  are 
listed  in  Figures  9  through  1 1  for  the  screech  owl 
{Otis  kennicottii),  least  chipmunk  (Tamias  mini- 
mus), and  Brewer's  sparrow,  respectively.  The 
models  describe  the  cover  types  inhabited  by  the 
species.  This  information  can  be  compiled  from 
many  sources,  including  field  guides,  range  maps, 
museum  records,  and  other  published  literature.  The 
best  situation  is  where  such  a  list  already  exists.  For 
example,  U.S.  Forest  Service  personnel  already  have 
associated  wildlife  species  with  range  cover  types  in 
Colorado,  Wyoming,  Kansas,  Nebraska,  and  South 
Dakota  (U.S.  Forest  Service  1981);  for  Arizona  and 
New  Mexico  on  the  computerized  RUNWILD  format 
(Patton  1978);  and  for  the  western  Sierra  Nevadas 
(Verner  and  Boss  1980). 

The  models  also  identify  the  layers  of  habitat 
where  nesting,  hatching,  or  birthing  of  the  species 


occur  and  where  foraging  takes  place  and  special 
structural  features  or  plant  species  apparently  se- 
lected by  the  species  within  a  habitat  layer.  The 
numerical  designation  of  the  major  structural  fea- 
tures within  layers  of  habitat  corresponds  to  the 
coded  subunits  in  Table  8.  Sound,  professional  judg- 
ment is  needed  when  developing  word  models  for 
placing  wildlife  species  along  habitat  gradients  be- 
cause species-habitat  information  usually  is  not  well- 
synthesized  in  the  literature. 

Models  like  those  in  Figures  9  to  1 1  try  to  de- 
fine the  niche  space  of  a  species  in  sufficient  detail 
so  that  the  usefulness  of  habitat  for  particular  species 
can  be  qualitatively  estimated  in  inventories.  Field 
surveys  can  be  used  to  estimate  the  relative  quantity 
of  habitat  for  particular  species.  It  may  also  be  possi- 
ble to  determine,  from  the  word  model,  those  life- 
requisite  needs  of  the  species  that  need  to  be  pro- 
vided or  modified  by  remedial  management  actions. 


A  typical  sagebrush  rangeland  in  the  western  United  States. 


120 


Rangelands 


LITERATURE  CITED 


ALBERTSON,  F.W.  and  J.E.  WEAVER.  1946.  Reduction  of 
ungrazed  mixed  prairie  to  short  grass  as  a  result  of 
drought  and  dust.  Ecol.  Monogr.  l6(4):449-463- 
BEETLE,  A.A.  I960.  A  study  of  sagebrush.  The  section 
tridentatae  of  Artemisia.  Wyoming  Agric.  Exp.  Sta. 
Bull.  368.  Laramie.  83pp. 
BLAISDELL, J.P,  R.B.  MURRAY,  and  ED.  MCARTHUR. 
1982.  Managing  intermountain  rangelands — sage- 
brush-grass ranges.  U.S.  Dep.  Agric,  For.  Serv.  Gen. 
Tech.  Rep.  INT- 134.  4  lpp. 
BROWN,  D.E.  1982.  Biotic  communities  of  the  American 
Southwest — United  States  and  Mexico.  Boyce  Thomp- 
son Southwestern  Arboretum.  Superior,  AZ.  342pp. 

and  C.H.  LOWE.  1980.  Biotic  communities  of  the 

Southwest.  U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep. 
RM-78.  lpp.  (map). 
COSTELLO,  D.F.  1944.  Important  species  of  the  major 

forage  types  in  Colorado  and  Wyoming.  Ecol.  Monogr. 
14(1):107-134. 
HALL,  F.C.  1973-  Plant  communities  of  the  Blue  Mountains 
in  eastern  Oregon  and  southwestern  Washington. 
U.S.  Dep.  Agric,  For.  Serv.  PNW-R6  Area  Guide  3-1. 
62pp. 
HARPER,  K.T.  and  J.L.  REVEAL  (symp.  organizers).  1978. 
Intermountain  biography:  a  symposium.  Great  Basin 
Naturalist  Memoirs.  2.  Brigham  Young  Univ.  Provo, 
UT.  268pp. 
HIRONAKA,  M.  1979.  Basic  synecological  relationships  of 
the  Columbia  River  sagebrush  type.  Pages  27-32  in 
The  Sagebrush  Ecosystem — A  Symposium.  Utah  State 
Univ.,  College  of  Natural  Resources.  Logan.  251pp. 
,  MA.  FOSBERG,  and  A.H.  WINWARD.  1983.  Sage- 
brush-grass habitat  types  of  southern  Idaho.  Forest, 
Wildlife,  and  Range  Exp.  Sta.  Bull.  35.  Univ.  of  Idaho, 
Moscow.  44pp. 
JOHNSON,  K.L.  1979-  Basic  synecological  relationships  of 
the  sagebrush  types  on  the  high  plains  of  Montana, 
Wyoming,  and  the  Dakotas.  Pages  42-49  in  the  Sage- 
brush Ecosystem — A  Symposium.  Utah  State  Univ., 
College  of  Natural  Resources.  Logan.  251pp. 
KUCHLER,  AW.  1964.  Manual  to  accompany  the  map. 
Potential  natural  vegetation  of  the  conterminous 
United  States.  Am.  Geogr.  Soc  Spec.  Publ.  36.  39pp. 
+   116  plates,  (map). 
MCARTHUR,  ED.,  AC.  BLAUER,  A.P.  PLUMMER,  and  R. 
STEVENS.  1979.  Characteristics  and  hybridization  of 
important  intermountain  shrubs.  III.  Sunflower  family. 
U.S.  Dep.  Agric,  For.  Serv.  Res.  Pap.  INT-220.  82pp. 
MCKELL,  CM.,  J.P.  BLAISDELL,  and  JR.  GOODIN,  tech. 
eds.  1972.  Wildland  shrubs — their  biology  and  utiliza- 
tion. U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep.  INT- 
1.  494pp. 
MOONEY,  H.A.  and  D.J.  PARSONS.  1973.  Structure  and 

function  of  the  California  chaparral — an  example  from 
San  Dimas.  Pages  83-1 12  in  di  Castra,  F.  and  H.A. 
Mooney,  eds.  Mediterranean  Type  Ecosystems:  Origin 
and  Structure.  Springer-Verlag.  New  York,  NY. 
MUEGGLER,  W.F.  and  W.L.  STEWART.  1980.  Grassland 
and  shrubland  habitat  types  of  western  Montana.  U.S. 
Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep.  INT-66.  154pp. 
NICHOL,  A.A.  1937.  The  natural  vegetation  of  Arizona. 
Ariz.  Agric.  Exp.  Sta.  College  of  Agric,  Univ.  Ariz. 
Tucson.  Tech.  Bull.  68:181-222. 


PATTON,  DR.  1978.  RUNWILD:  A  storage  and  retrieval 
system  for  wildlife  habitat  information.  U.S.  Dep. 
Agric,  For.  Serv.  Gen.  Tech.  Rep.  RM  51.  8pp. 

PLUMMER,  A.P,  DR.  CHRISTENSEN,  and  SB.  MONSEN. 
1968.  Restoring  big-game  range  in  Utah.  Publ.  No.  68- 
3.  Utah  Div.  Fish  and  Game.  Salt  Lake  City.  183pp. 

,  AC.  HULL,  Jr.,  G.  STEWART,  and  J.H.  ROBERTSON. 

1955.  Seeding  rangelands  in  Utah,  Nevada,  southern 
Idaho,  and  western  Wyoming.  U.S.  Dep.  Agric,  For. 
Serv.  Agric.  Handb.  71.  73pp 

POULTON,  C.E.  1970.  Practical  applications  of  remote 
sensing  in  range  resources  development  and  manage- 
ment. Pages  179-189  in  Range  and  Wildlife  Habitat 
Evaluation — A  Research  Symposium.  U.S.  Dep.  Agric, 
For.  Serv.  Misc.  Publ.  1147.  Washington,  DC. 

SAMPSON,  AW,  A.  CHASE,  and  D.W.  HEDRICK.  1951. 
California  grasslands  and  range  forage  grasses.  Calif. 
Agric  Exp.  Sta,  College  of  Agric.  Univ.  Calif.  Bull. 
724.  130pp. 

SCHWARZ,  C.F.,  C.E.  THOR,  and  G.H.  ELSNER.  1976. 

Wildlife  planning  glossary.  U.S.  Dep.  Agric,  For.  Serv. 
Gen.  Tech.  Rep.  PSW-13.  252pp. 

SHARP,  LA.  and  K.D.  SANDERS.  1978.  Rangeland  re- 
sources of  Idaho:  A  basis  for  development  and  im- 
provement. Idaho  Rangeland  Comm.  and  College  of 
FWR.  FWR  Exp.  Sta.  Misc.  Publ.  6.  Moscow,  ID.  74pp. 

SHORT,  H.L.  1983.  Wildlife  guilds  in  Arizona  desert  habi- 
tats. U.S.  Dep.  Inter,  Bur.  Land  Manage.  Tech.  Note 
362.  258pp. 

and  K.P.  BURNHAM.  1982.  Technique  for  structur- 
ing wildlife  guilds  to  evaluate  impacts  on  wildlife 
communities.  U.S.  Dep.  Inter,  Fish  and  Wildl.  Serv. 
Spec.  Sci.  Rep.  -  Wildl.  244.  34pp. 

STODDARD,  LA,  AD.  SMITH,  and  T.W.  BOX.  1975. 

Range  management.  Third  ed.  McGraw-Hill  Book  Co. 
New  York,  NY.  532pp. 

TISDALE,  E.W.  1979.  A  preliminary  classification  of  Snake 
River  Canyon  grasslands  in  Idaho.  Univ.  Idaho  FWR 
Exp.  Sta.  Note.  32.  Moscow. 

.  1983.  Grasslands  of  western  North  America:  The 

Pacific  Northwest  bunchgrass  type.  Pages  223-245  in 
Nicholson,  A.C,  A.  McLean,  and  T.E.  Baker,  eds.  Grass- 
land Ecology  and  Classification  Symposium  Proceed- 
ings. B.C.  Min.  For.  Victoria,  B.C. 
DEPARTMENT  OF  AGRICULTURE.  1936.  Atlas  of 
American  agriculture.  U.S.  Govt.  Printing  Office. 
Washington,  DC. 

DEPARTMENT  OF  INTERIOR,  BUREAU  OF  LAND 
MANAGEMENT.  1975.  Range  condition  report  pre- 
pared for  the  Senate  Committee  on  appropriations. 
U.S.  Dep.  Inter..  Bur.  Land  Manage.  Washington,  DC. 
(unpubl). 

FISH  AND  WILDLIFE  SERVICE.  1980.  Habitat  Evalu- 
ation Procedures  (HEP),  ESM  102.  U.S.  Dep.  Inter, 
Fish  and  Wildl.  Serv.  Div.  of  Ecological  Services. 
Washington,  DC. 

FOREST  SERVICE.  1972.  The  Nation's  range  re- 
sources— a  forest-range  environmental  study.  For. 
Resour.  Rep.  19.  U.S.  Dep.  Agric,  For.  Serv.  Washing- 
ton, DC.  147pp. 

.  1981.  Wildlife  and  fish  habitat  relationships.  Vol.  I 

Narratives.  Vol.  II  Matrices.  U.S.  Dep.  Agric,  For.  Serv. 
Rocky  Mtn.  Region.  Denver,  CO. 
UTAH  STATE  UNrVERSITY.  1975.  The  pinyon- juniper 

ecosystem:  a  symposium.  Utah  State  Univ.,  College  of 
Natural  Resources,  Utah  Agric.  Exp.  Sta,  Logan. 
194pp. 


U.S. 


U.S. 


U.S 


U.S. 


Rangelands 


121 


.1979.  The  sagebrush  ecosystem:  a  symposium.  Utah 

State  Univ.,  College  of  Natural  Resources,  Logan. 

251pp. 
VERNER,  J.  and  A.S.  BOSS.  1980.  California  wildlife  and 

their  habitats:  Western  Sierra  Nevada.  U.S.  Dep.  Agric, 

For.  Serv.  Gen.  Tech.  Rep.  PSW-37.  439pp. 
WAGNER,  W.L.  and  E.F.  ALDON.  1978.  Manual  of  the 

saltbushes  (Atriplex  sp.)  in  New  Mexico.  U.S.  Dep. 

Agric,  For.  Serv.  Gen.  Tech.  Rep.  RM-57.  50pp. 
WEAVER,  J.E.  1954.  North  American  prairie.  Johnson  Publ. 

Co.  Lincoln,  NE.  348pp. 
and  F.W.  ALBERTSON.  1956.  Grasslands  of  the 

Great  Plains.  Johnson  Publ.  Co.  Lincoln,  NE.  395pp. 


WINWARD,  A.H.  1980.  Taxonomy  and  ecology  of  sage- 
brush in  Oregon.  Agric.  Exp.  Sta.  Bull.  642.  Oregon 
State  Univ.,  Corvallis.  15pp. 

WRIGHT,  HA.  1972.  Shrub  response  to  fire.  Pages  204- 
217  in  McKell,  C.M.J.P  Blaisdell,  and  JR.  Goodin, 
tech  eds.  Wildland  Shrubs — Their  Biology  and  Utiliza- 
tion. U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep.  INT- 
1.  494pp. 

,  L.F.  NEUENSCHWANDER,  and  CM.  BRITTON. 

1979.  The  role  and  use  of  fire  in  sagebrush-grass  and 
pinyon-juniper  plant  communities.  A  state-of-the-art 
review.  U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep. 
INT- 58.  48pp. 


122 


Rangelands 


DESERTS 


K.  Bruce  Jones1 

U.S.  Bureau  of  Land  Management 
Phoenix  Training  Center 
Phoenix,  AZ  85015 


"But  the  desert  is  much  more  than  merely  warm. 
It  is  a  consistent  world  with  a  special  landscape,  a 
special  geography,  and  to  go  with  them,  a  special 
flora  and  fauna  adapted  to  that  geography  and  that 
climate." 

—Joseph  Wood  Krutch,  from  The  Voice  of  the 
Desert 


Editors  Note:  Deserts  compose  a  large  portion  of 
the  western  U.S.  Wildlife  are  one  of  their  most  im- 
portant resources.  Animals  of  the  desert  have 
unique  adaptations  to  the  extremes  of  heat  and 
precipitation  that  characterize  such  regions.  As  the 
author  emphasizes,  effective  inventory  and  moni- 
toring studies  must  take  into  account  these  ex- 
tremes and  desert  animals'  adaptations  to  them. 


INTRODUCTION 

Deserts  are  characterized  by  low,  erratic  precip- 
itation and  highly  variable  temperatures,  and  extend 
from  southeastern  Oregon  and  southern  Idaho 
through  Nevada  and  Utah,  except  at  higher  eleva- 
tions, continuing  south  through  southern  California 
and  Arizona,  and  eastward  through  central  and 
southern  New  Mexico  (Oosting  1956).  Desert  habi- 
tats range  from  homogeneous  stands  of  sagebrush 
{Artemisia  sp. )  in  the  Great  Basin  Desert  to  highly 
diverse,  structurally  rich  vegetation  of  the  Sonoran 
Desert.  To  a  large  degree,  vegetation  structure 
within  deserts  reflects  the  areas'  precipitation  pattern 
and  temperature  regimes.  For  example,  structural 
simplicity  within  the  Great  Basin  Desert  reflects  the 
area's  short  growing  season,  low  precipitation,  and 
precipitation  pattern  (60%  of  the  precipitation  is  in 
the  form  of  snow ).  Conversely,  parts  of  the  Sonoran 
Desert  are  structurally  rich  because  of  a  year-round 
growing  season  (few  freezing  temperatures),  bian- 
nual precipitation  patterns,  and  precipitation  in  the 
form  of  rain. 


Northern,  cooler  desert  regions,  such  as  the 
Great  Basin  Desert,  support  far  fewer  wildlife  species 
than  more  southern,  warmer  regions  such  as  the  So- 
noran Desert  (Bender  1982;  Brown  1982b).  Certain 
groups  of  wildlife,  such  as  large  herbivores,  are  gen- 
erally more  numerous  in  northern  deserts.  Similar  to 
their  effect  on  vegetation,  shorter  growing  seasons 
affect  the  diversity  of  wildlife,  especially  small  birds, 
mammals,  reptiles,  and  amphibians  by  reducing  the 
diversity  and  availability  of  insect  prey.  Thermal  re- 
gimes in  northern  deserts  limit  ectothermic  (cold- 
blooded) and  small  homeothermic  (warm-blooded) 
wildlife  activity  to  short  periods  (generally  May  to 
early  September). 

Desert  habitats  possess  some  of  the  most  unu- 
sual wildlife  in  North  America.  Many  desert  animals 
are  physiologically  and  morphologically  adapted  to 
survive  under  extreme  environmental  conditions 
(low,  erratic  rainfall  and  highly  variable  tempera- 
tures). Many  small  mammals  require  no  free  water; 
these  animals  survive  on  their  own  metabolic  water 
and  through  water  conservation  strategies  such  as 


•Current  Address:  U.S.  Fish  and  Wildlife  Service.  Office  of  Endan- 
gered Species,  Washington,  DC  20240. 


Deserts 


123 


nocturnal  activity  patterns  (Golley  et  al.  1975). 
Many  desert  animals  such  as  lizards  conserve  water 
by  excreting  uric  acid  rather  than  urea  (Porter 
1972). 

Perhaps  the  single  most  important  adaptation 
employed  by  desert  wildlife  is  behavior.  Many  small 
mammals  and  snakes  are  entirely  nocturnal  and  fos- 
sorial.  They  avoid  high  temperatures  by  remaining 
inactive  in  burrows,  under  rock,  or  in  caves  (e.g., 
bats)  during  the  day,  and  perform  feeding  and  repro- 
ductive activities  at  night.  Even  diurnal  animals  such 
as  some  birds,  big  game,  and  lizards  are  inactive  dur- 
ing hot  parts  of  the  day. 

In  desert  regions,  most  animals  hibernate  or  re- 
main inactive  during  cold  winter  months.  This  is  es- 
pecially true  in  northern  sections  of  the  Great  Basin 
Desert  where  winter  temperatures  drop  below  0°  F. 
In  the  Sonoran  Desert,  where  temperatures  rarely 
reach  the  freezing  point,  many  small  mammals  and 
small  lizards  remain  active  year-round. 


as  vegetation  patches,  rock,  soil,  and  surface  debris 
(e.g.,  logs).  The  arrangement  and  abundance  of  these 
microhabitats  determine  the  diversity  and  abundance 
of  desert  wildlife;  small  changes  in  these  microhabi- 
tats can  alter  species  abundance  and  diversity  (see 
Heatwole  1982  and  Short  1983  for  examples).  Be- 
cause animals  are  closely  associated  with  microhabi- 
tats, man  can  easily  affect  the  composition  and  abun- 
dance of  wildlife  in  deserts. 

With  rapid  expansion  of  man  into  desert  habi- 
tats, many  habitat  components  crucial  to  species  ex- 
istence are  being  altered,  especially  those  important 
to  both  man  and  wildlife.  For  example,  dewatering  of 
perennial  streams  and  springs  for  domestic  and  live- 
stock water  has  drastic  effects  on  wildlife,  especially 
aquatic  organisms.  Recreation  activities  have  become 
very  popular  in  deserts  near  metropolitan  areas,  and 
these  activities  have  both  direct  and  indirect  effects 
on  wildlife  and  their  habitats  (physical  disturbance 
and  habitat  alteration,  respectively).  In  addition, 
many  regions,  especially  within  the  Great  Basin  Des- 
ert, have  large  coal,  oil,  and  gas  deposits.  The  devel- 
opment of  these  resources  can  have  significant  ef- 
fects on  desert  habitats  and  their  faunas. 


<-.*■:,;.    '.-•$•'   ->■<  «*-?m.  *j?r 


Great  Basin  Desert  in  the  winter. 


Response  (return  to  original  state)  of  desert 
habitats  after  man-caused  physical  changes  is  slow, 
and  in  severe  instances,  where  soils  are  lost,  habitats 
may  never  return  to  their  original  state. 

Inventories  and  monitoring  studies  of  wildlife 
communities  in  desert  habitats  help  biologists  and 
managers  understand  the  sensitive  and  complex  in- 
terrelationships between  wildlife  species  and  com- 
munities, and  resources  of  desert  habitats.  With 
these  data,  one  can  protect,  and  in  some  situations, 
enhance  the  integrity  of  our  deserts.  This  chapter  de- 
scribes desert  habitats,  including  habitat  compo- 
nents, and  provides  general  guidance  for  conducting 
inventories  and  monitoring  studies  in  deserts. 


Although  many  desert  wildlife  require  no  free- 
standing water  (see  Golley  et  al.  1975  for  examples), 
permanent  and  temporary  water  sources  are  critical 
to  certain  wildlife.  For  example,  big  game  species 
such  as  bighorn  sheep  (Ovis  canadensis)  occupy 
areas  with  scattered,  but  reliable  free-standing  water 
(Wilson  et  al.  1980).  Many  habitat  management  strat- 
egies are  developed  around  this  concept;  bighorn 
sheep  populations  can  be  increased  by  increasing  the 
distribution  and  reliability  of  free-standing  water. 
Small  streams  transecting  through  desert  habitats  are 
inhabited  by  several  small  fish  and  amphibians.  These 
animals  are  entirely  dependent  on  permanent,  run- 
ning (lotic)  water  within  these  regions. 

Because  of  extreme  environmental  conditions, 
desert  wildlife  are  highly  dependent  on  microhabi- 
tats, especially  those  that  provide  thermal  cover  such 


Desert  bighorn  sheep  at  man-made  water  supply. 


124 


Deserts 


CLASSIFICATION  AND  DESCRIPTIONS  OF 
MAJOR  DESERTS 

Although  deserts  cover  large  sections  of  western 
North  America,  they  possess  certain  similar  environ- 
mental characteristics  throughout  (Oosting  1956).  All 
deserts  share  the  following  general  characteristics: 

1.  low,  erratic  precipitation,  generally  less  than 
300  mm  (12  in. )  per  year 

2.  high,  daily  air  and  soil  temperatures  that  drop 
rapidly  at  night 

3.  highly  variable,  seasonal  temperatures 

4.  low  atmospheric  moisture  throughout  most  of 
the  year 

5.  few  overcast  or  cloudy  days 

Shreve  (  1942)  and  Oosting  ( 1956)  characterized  des- 
erts into  two  general  groups  based  on  climate:  cold 
formation  and  warm  formation.  Lowe  and  Brown 
( 1982)  further  subdivided  warm  formation  deserts 
into  warm-temperate  and  tropical-subtropical  based 
on  climate  and  plant  species  affinities  and  origin: 

1.  Cold-temperate  Deserts 
a.  Great  Basin  Desert 

2.  Warm- temperate  Deserts 

a.  Mojave  Desert 

b.  Chihuahuan  Desert 

3.  Tropical-subtropical  Deserts 
a.  Sonoran  Desert 

Because  most  authors  agree  on  the  four  major 
deserts  listed  above,  I  will  base  my  discussions 
around  this  classification.  Finer  delineation  of  these 
four  major  deserts  is  available.  Perhaps  the  most  ex- 


tensive delineation  of  North  America's  four  major 
deserts  is  that  of  Brown  (  1982b).  In  his  publication, 
deserts  are  broken  down  to  association  and  commu- 
nity (series)  levels,  following  the  digitized  format  of 
Brown  et  al.  (1979).  Other  delineations  of  deserts 
include  Province  and  Region  descriptions  by  Bailey 
(1978),  and  potential  natural  vegetation  descriptions 
to  the  association  level  by  Kuchler  ( 1 964 ). 

The  following  are  brief  descriptions  of  each  des- 
ert (see  Figure  1  for  locations). 


Great  Basin  Desert 

The  Great  Basin  Desert  is  the  most  extensive 
desert  in  the  U.S.,  stretching  from  southeastern  Ore- 
gon and  Wyoming,  south  to  northern  Arizona  and 
New  Mexico,  east  to  northern  New  Mexico,  and 
west  into  extreme  eastern  California.  Topography  of 
the  Great  Basin  Desert  varies,  but  generally  consists 
of  wide  valley  floors  between  1,200  and  1,800  m 
(4,000  and  6,000  ft),  interrupted  by  mountains. 
Temperatures  drop  much  lower  than  any  other  U.S. 
desert,  with  a  short  frost-free  season  and  very  cold 
winters  (Oosting  1956),  and  precipitation  ranges 
from  76  to  300  mm  (4  to  12  in.).  Two  major  vegeta- 
tion communities  occur  within  this  desert,  both  of 
which  are  structurally  and  floristically  simple:  sage- 
brush associations  dominated  by  sagebrush,  Artemi- 
sia sp.,  and  shadscale  or  saltbush  associations  domi- 
nated by  saltbushes,  Atriplex  sp.  (Oosting  1956; 
Turner  1982).  Species  with  evolutionary  affinities  to 
warmer  climates  such  as  rabbitbush  {Chrysothamnas 
sp.)  and  blackbrush  (Coleogyne  sp. )  are  also  present 
in  the  Great  Basin  Desert  (Turner  1982a). 


Great  Basin  Desert  in  Utah,  dominated  by  sand  sagebrush. 


Deserts 


125 


Great  Basin 


Mojave 

Sonoran 

Chihuahuan 


Figure  1.  Locations  of  major  North  American  deserts. 


126  Deserts 


Mojave  (Mohave)  Desert 

The  Mojave  Desert  is  the  smallest  of  the  desert 
formations,  occurring  in  southeastern  California, 
southern  Nevada,  northwestern  Arizona,  and  extreme 
southwestern  Utah  at  elevations  of  300-1,200  m 
(1,000-4,000  ft;  Oosting  1956).  Summers  are  hot  and 
dry;  precipitation  throughout  this  desert  is  generally 
meager  (less  than  120  mm  [5  in.]  per  year)  although 
some  areas  receive  up  to  22.5  cm  (9  in.)  per  year 
(Thome  et  al.  1981).  Floristically,  faunally,  and  geo- 
graphically, the  Mojave  Desert  is  an  intermediate  be- 
tween Great  Basin  and  Sonoran  Deserts,  although  its 
fauna  more  closely  resembles  that  of  the  Sonoran 
Desert  (Turner  1982b).  The  Mojave  Desert  is  per- 
haps the  most  structurally  and  floristically  variable  of 
all  the  deserts  although  most  regions  consist  of  low- 
height  (<1.5  m  [5  ft])  shrubs  consisting  of  homoge- 
neous stands  of  creosotebush  (Larrea  sp.)  and  black- 
brush;  other  regions  include  structurally-diverse  flora 
such  as  the  Joshua  tree  (  Yucca  brevifolia;  Turner 
1982a).  There  are  few  Mojave  Desert  endemics 
(plant  species)  that  distinguish  this  desert  from  oth- 
ers (with  the  exception  of  the  Joshua  tree).  Classifi- 
cation of  the  Mojave  Desert  as  a  distinct  desert  is 
primarily  based  on  geographic  position  and  climate 
rather  than  plant  or  animal  associates  (Turner 
1982b). 


'  -  ■  :  ' '  j* 


Yucca  in  bloom. 


A  Mojave  Desert  Joshua  tree  "woodland"  community  in  Arizona. 


Deserts 


127 


Chihuahuan  Desert,  with  mesquite  (New  Mexico). 


Chihuahuan  Desert 

The  Chihuahuan  Desert  is  floristically  and  struc- 
turally variable  and  has  great  regional  differences  in 
plant  species  composition  (Brown  1982a).  Most  of 
this  desert  occurs  in  Mexico  (80%  )  with  northern 
boundaries  extending  into  southeastern  Arizona, 
southern  and  central  New  Mexico,  and  western 
Texas  at  elevations  of  900-1,800  m  (3,000-6,000  ft; 
Brown  1982a).  Similar  to  the  Great  Basin  Desert,  this 
desert  is  interrupted  by  high  mountains  (Oosting 
1956).  Precipitation  is  variable  and  ranges  from 
76-300  mm  (4-12  in.)  per  year.  The  Chihuahuan  Des- 
ert has  expanded  in  the  past  200  years  at  the  ex- 
pense of  semidesert  grassland,  primarily  due  to  man's 
activities  (York  and  Dick-Peddie  1969;  Brown 
1982a).  Although  present  in  the  Mojave  and  Sonoran 
Deserts,  mesquite  (Prosopis  sp.)  and  creosotebush 
dominate  most  sites  within  the  U.S.  Chihuahuan  Des- 
ert. Whereas  most  areas  are  structurally  homogene- 
ous and  simple,  whitethorn  acacia  (Acacia  con- 
stricta),  ocotillo  (Fouquieria  splendens),  yuccas 
(Yucca  sp. ),  agaves  (Agave  sp.),  and  cactus  (Opuntia 
sp.)  add  to  structural  diversity  in  mountainous  areas 
and  canyons  (Brown  1982a).  Solid  mesquite  dune 
stands  are  common  in  southern  New  Mexico,  pri- 
marily on  sites  previously  occupied  by  semidesert 
grassland  (Brown  1982a). 

Sonoran  Desert 

The  Sonoran  Desert,  occupying  areas  in  south- 
ern Arizona,  southeastern  California,  and  northwest- 
ern Mexico,  is  perhaps  the  most  structurally  and  flo- 
ristically diverse  desert  in  the  world.  It  is  also  the 
only  U.S.  desert  with  most  of  its  flora  derived  from 


tropical  climatic  zones,  and  appears  to  have  formed 
less  than  10,000  years  ago  (Turner  and  Brown 
1982).  There  are  two  major  subdivisions  of  the  So- 
noran Desert  within  the  U.S.:  Lower  Colorado  River 
Valley  and  Arizona  Upland  (Turner  and  Brown 
1982).  The  Lower  Colorado  subdivision  occurs  pri- 
marily in  southwestern  Arizona  and  southeastern  Cal- 
ifornia at  elevations  of  0-600  m  (0-2,000  ft).  Precipi- 
tation is  low  (less  than  175  mm,  7  in.  per  year)  and 
temperatures  moderate  (mostly  frost-free).  Creosote- 
bush  and  bursage  (Franseria  sp. )  dominate  most  of 
this  subdivision.  Structural  and  floristic  diversity  is 
achieved  in  these  areas  along  intermittent  washes 
consisting  of  ironwood  (Olneya  tesota),  blue  palo 
verde  (Cercidium  floridum),  and  mesquite  (Turner 
and  Brown  1982). 

The  Arizona  Upland  subdivision  of  the  Sonoran 
Desert  is  the  most  structurally  diverse  of  all  U.S.  des- 
erts. It  occurs  in  southern  Arizona  and  northern  So- 
nora  at  elevations  of  300-1,200  m  (1,000-4,000  ft). 
Similar  to  the  Lower  Colorado  subdivision,  this  sub- 
division receives  rain  biannually  (summer  and  win- 
ter), but  unlike  its  other  Sonoran  counterpart,  has 
greater  moisture  (due  to  higher  elevations  and  rug- 
ged, broken  terrain;  Turner  and  Brown  1982). 
Turner  and  Brown  ( 1982)  attributed  dominance  of 
arboreal  and  succulent  vegetation  within  this  region 
to  available  moisture.  Overstory  structure  is  pro- 
vided by  ironwoods,  foothill  palo  verde  (Cercidium 
microphyllum),  saguaro  cactus  (Carnegiea  gigan- 
tea ),  and  mesquite.  A  wide  variety  of  understory 
components  may  also  be  present:  cactus  (Opuntia 
sp. ),  bursage,  ratany  (Krameria  sp. ),  buckwheat  (Er- 
iogonum  sp. ),  brittlebush  (Encelia  sp. ),  and  jojoba 
(Simmondsia  chinensis;  Turner  and  Brown  1982). 


128 


Deserts 


Sonoran  Desert  in  Arizona. 


MAJOR  WILDLIFE  SPECIES  GROUPS 

Wildlife  species  assemblages  of  North  American 
deserts  range  from  relatively  undiverse  in  homogene- 
ous, northern  sections  of  the  Great  Basin  Desert  to 
highly  diverse  (Table  1 )  in  the  Arizona  Upland  sub- 
division of  the  Sonoran  Desert  (see  Bender  1982  and 
Brown  1982b).  Great  differences  in  faunas  of  north- 
ern portions  of  the  Great  Basin  Desert  and  southern 
deserts  such  as  the  Sonoran  Desert  primarily  result 
from  differences  in  temperature,  moisture,  and  struc- 
tural characteristics  associated  with  each.  As  dis- 
cussed in  the  previous  section,  the  Great  Basin  Des- 
ert is  extremely  homogeneous,  structurally  simple, 
cold,  and  precipitation  falls  primarily  in  the  winter 


(60%  )  when  plants  are  dormant.  Conversely,  the  So- 
noran Desert  is  structurally  diverse,  warm,  and  re- 
ceives precipitation  during  both  winter  and  summer. 
Moderate  temperatures  allow  plants  to  use  winter 
and  summer  precipitation  for  growth  (Oosting  1956; 
Daubenmire  1974).  Lack  of  available  moisture  and 
short  growing  seasons  may  preclude  existence  of 
many  arboreal  plants  from  the  Great  Basin  Desert, 
common  in  southern  deserts. 

The  Great  Basin  Desert  has  fewer  wildlife  species 
than  any  of  the  other  deserts.  However,  this  desert 
supports  larger  populations  of  pronghorn  (Antiloca- 
pra  americana),  mule  deer  {Odocoilens  hemionus), 
and  elk  {Cetvus  elapbus )  than  other  deserts  (see 


Table  1.  Qualitative  assessment  of  species  diversity  for  the  four  major  U.S.  deserts. 
£  =  highly  diverse,    Q  =  moderately  diverse,     Q  =  low  diversity. 


Desert 

Total 
Diversity 

Raptor 

Fish 

Amphibians 

Reptiles 

Small 
Birds 

Mammals 
(Non-game) 

Upland 
Game 

Water- 
fowl 

Big 
Game 

Great  Basin 

o 

Q 

O 

o 

O 

O 

Q 

Q 

9 

• 

Mojave 

Q 

Q 

O 

Q 

Q 

9 

Q 

Q 

O 

o 

Chihuahuan 

Q 

Q 

O 

• 

• 

9 

Q 

9 

O 

* 

9 

Sonoran 

• 

• 

Q 

• 

• 

• 

• 

9 

o 

9 

'Moderate  diversity  results  primarily  from  introduced  species. 


Deserts 


129 


Schmidt  and  Gilbert  1978).  It  also  supports  large  up- 
land game  populations  (Johnsgard  1973)  such  as  chu- 
kars  (Alectoris  chukar)  and  sage  grouse  (Centrocer- 
cus  urophasianus ).  Several  bird  species  move  into 
the  Great  Basin  Desert  during  spring  and  summer  to 
breed,  although  breeding  bird  diversity  and  abun- 
dance is  generally  less  than  any  of  the  other  three  des- 
erts  (see  Brown  and  Gibson  1983)-  Breeding  and  res- 
ident raptor  and  bird  species  are  greatest  within  the 
Great  Basin  Desert  in  areas  adjacent  to  more  structur- 
ally diverse  habitats,  such  as  riparian  areas  and  cliffs. 
For  example,  riparian  sites  provide  structural  diversity 
necessary  for  nesting,  whereas  the  adjacent  Great 
Basin  Desert  provides  feeding  grounds.  Areas  with 
cliffs  increase  diversity  in  this  desert  by  increasing 
structural  habitat  diversity. 

Reptile,  amphibian,  and  small  mammal  diversity 
is  relatively  low  in  the  Great  Basin  Desert  when 
compared  with  the  three  other  deserts  (see  Brown 
and  Gibson  1983).  As  previously  discussed,  low  di- 
versity of  small  nongame  wildlife  results  primarily 
from  a  short  growing  season  and  low  environmental 
temperatures. 

The  Mojavc  Desert  supports  an  array  of  wildlife. 
Whereas  this  desert's  big  and  upland  game  diversity 
and  abundance  is  not  as  great  as  the  Great  Basin  De- 
sert, it  does  support  a  diverse  small  mammal  and 
reptile  fauna.  Similar  to  the  Great  Basin  Desert,  small 
bird  and  raptor  diversity  and  abundance  are  greatly 
enhanced  in  areas  with  added  structural  diversity 
(cliffs,  riparian  habitats,  canyons,  and  Joshua  trees). 

The  Chihuahuan  and  Sonoran  Deserts  have 
highly  diverse  faunas,  including  large  numbers  of  am- 
phibians, reptiles,  small  birds,  mammals,  and  raptors. 
High  species  diversity  results  from  topographical  and 
vegetational  structural  diversity,  a  long  growing  sea- 
son, and  influence  of  Mexican  and  southcentral  U.S. 


faunas.  In  southern  New  Mexico,  Ibex  (Capra  aega- 
grus )  and  Oryx  ( Oryx  gazella )  have  been  intro- 
duced into  parts  of  the  Chihuahuan  Desert.  Javelina 
( Tayassu  tajacn )  have  also  been  introduced  into 
several  parts  of  both  the  Chihuahuan  and  Sonoran 
Deserts.  These  introductions,  combined  with  native 
big  game  and  upland  game  species,  give  these  des- 
erts relatively  diverse  game  faunas. 


Siberian  ibex  in  southern  New  Mexico  habitat.  Released  in 
1975  in  the  Canadian  River  Canyon. 


Oryx,  commonly  called  Gemsbok,  in  southern  New  Mex- 
ico habitat.  Released  in  1969  on  White  Sands  Missile 
Range. 


The  Sonoran  Desert  has  the  highest  wildlife  di- 
versity of  any  North  American  desert  ( see  Brown 
and  Gibson  1983),  primarily  resulting  from  a  year- 
round  growing  season  and  great  vegetation  structural 
diversity. 

Bender  ( 1982 )  and  Brown  ( 1982b)  provided 
lists  of  wildlife  and  plants  associated  with  North 
American  deserts.  Table  1  provides  a  general  assess- 
ment of  species  richness  in  the  four  major  U.S. 
deserts. 


CRITICAL  HABITAT  FEATURES 

Deserts  consist  of  several  physical  habitat  com- 
ponents that  contribute  to  wildlife  species  diversity. 
These  components  are  used  by  wildlife  to  perform  a 
number  of  ecological  functions  such  as  mating,  feed- 
ing, nesting,  and  thermoregulation,  and  they  also 
provide  resting  substrate  and  escape  cover.  Habitat 
components  or  resources  are  partitioned  in  time  and 
space  by  desert  wildlife  (Creusere  and  Whitford 
1982).  Because  deserts  arc  environmentally  extreme, 
many  wildlife  species  use  specific  habitat  compo- 
nents, especially  those  components  that  provide 
greater  moisture  and  lower  temperatures.  For  exam- 
ple, the  desert  night  lizard  (Xantusia  vigilis)  is  re- 
stricted to  desert  regions  with  downed  litter  of 
yucca  and  agave  plants  (Zweifel  and  Lowe  1966; 
Jones  1981a). 


130 


Deserts 


Several  other  factors  determine  wildlife  commu- 
nity composition  and  affect  availability  of  habitat 
components.  These  include  slope,  aspect,  precipita- 
tion, ecotones,  geographic,  and  man-caused  factors. 

Microhabitat  components,  such  as  soil,  rock, 
vegetation,  and  water,  vary  greatly  between  and 


within  our  four  U.S.  deserts.  The  following  discusses 
each  habitat  component  with  examples  of  their  rela- 
tions to  various  wildlife.  This  will  help  determine 
which  microhabitats  should  be  sampled  when  sur- 
veying and  monitoring  certain  desert  wildlife.  Tables 
2  through  4  summarize  important  habitat  compo- 
nents in  deserts. 


Major  Vertebrate  Associations  and  Relationship 
of  Species'  Ecology  to  Components 

Table  2.  Important  habitat  components  of  desert  habitats. 

CD 
03 

to 
Q 

CD 

> 

o 
O 

CD 
Q. 
CO 

o 
en 
111 

>< 

CO 

a. 
o 

o 
o 

LL 

03 

to 

3 

(D 
O) 

c 

T3 
CD 

CD 
LL 

cn 

c 
!c 

t: 

in  .cd 
i-  73 

O  i: 

4_      </3 

C/3  J3 

CO    3 

z  c/> 

CO 

o 
cn 
o 
o 

03 
>- 

-C 

Q- 

c 
o 

o 

3 
"O 

o 

a. 

CO 

<r 

03 

to 

3 

CO 

cn 

c 

u5 

CO 
LT 

o 
to 

3 

cn 

CD    CD 
O    CO 

CO  -Q 

SI     3 

h  in 

Microhabitat 
Components 

Description 

Variables/Factors 

Water 
(Lotic) 

Rivers  and 
streams, 
flowing 
springs 

Riffle/run/pool  water  temperature, 
turbidity,  DO,  organic  content, 
siltation,  pollutants,  interface  with 
other  habitats,  substrate 

Fl 

• 

• 

• 

• 

• 

• 

RE 

• 

• 

• 

• 

• 

• 

BG 

• 

RA 

Water 

(Permanent 

Lentic) 

Ponds, 

marshes, 

lakes, 

potholes, 

bogs, 

seeps, 

springs, 

agricultural 

runoff, 

natural  and 

man-made 

catchments 

Water  temperature,  DO,  organic 
content,  siltation,  pollutants, 
interface  with  other  habitats  (e.g., 
nesting  islands),  emergent 
vegetation,  shoreline  vegetation, 
substrate,  water  distributors 

Fl 

• 

• 

• 

• 

• 

• 

AM 

• 

• 

• 

• 

• 

• 

RE 

• 

• 

• 

• 

• 

RA 

WA 

• 

• 

BA 

• 

BG 

• 

Water 

(Temporary 

Lentic) 

Temporary 
rain  pools, 
irrigation 
ditches 

Water  temperatures,  duration, 
frequency,  siltation,  pollutants, 
DO,  emergent  vegetation, 
substrate  interface  with  other 
habitats,  organic  content,  water 
distribution 

AM 

• 

RE 

• 

• 

• 

• 

• 

WA 

• 

• 

BG 

• 

Rock 

Talus 

slopes, 

caves,  cliffs, 

boulders, 

substrate, 

outcorps, 

instream 

structure 

Rock  size,  heterogeneity, 
interfaces  with  other  habitats, 
origin,  vertical  and  horizontal 
structure 

Fl 

• 

• 

AM 

• 

• 

RE 

• 

• 

SM 

• 

BA 

• 

BG 

• 

• 

RA 

• 

SB 

• 

• 

Soil 

Surface  and 
subsurface 
soil  types 

Soil  types,  depth,  heterogeneity, 
horizontal  and  vertical  structure, 
interfaces  with  other  habitats 

AM 

• 

RE 

SB 

RA 

SM 

Vegetation- 
Litter/ 
Debris 

Leaves, 
logs,  limbs 
and  other 
persistent 
and  non- 
persistent 
litter/debris 

Litter  size,  depth,  heterogeneity, 
horizontal  and  vertical  structure, 
type,  moisture  retention, 
temperature. 

Fl 

AM 

• 

RE 

SB 

SM 

• 

UG 

• 

Deserts 


131 


vlajor  Vertebrate  Associations  and  Relationship 
Df  Species'  Ecology  to  Components 

Table  2.  Important  habitat  components  of  desert  habitats 
(concluded). 

CD 
f> 

C 

a> 

"cd 

Q 

CD 

> 

o 
O 

CD 
Q. 
03 
O 

</> 
LU 

CD 
CL 

O 
TD 

O 
O 
U_ 

to 
.a 

3 

CO 

a> 

c: 

T3 
CD 

CD 
U. 

en 

c 
jz 
tr 
in  a> 

>-    03 

—  in 
a>  3 

CO 

o 
a) 
o 
o 

</) 

>% 

-C 
0- 

g 
o 

o 

Q. 
CD 

<r 

£ 
o5 

3 
CO 
en 

c 

o5 

CD 
CL 

>. 
O 

re 

3 
05 

CD    CD 
O    CO 

E^ 

CD  -Q 

SI     3 

H  CO 

Microhabitat 
Components 

Description 

Variables/Factors 

Vegetation- 
Dead 

Standing 

vegetation, 

roots 

Size,  interface  with  other 
habitats,  heterogeneity,  vertical 
and  horizontal  structure,  soils 
(roots  only). 

Fl 

AM 

BA 

UG 

RE 

• 

SB 

• 

RA 

• 

SM 

• 

A 

Vegetation- 
Live 

All 

vegetation 
including 
roots 

Horizontal  and  vertical  structure, 
interfaces  with  abiotic  habitats, 
heterogeneity 

AL 

• 

• 

• 

• 

• 

Vegetation- 
Plant 
Species 

Individual 
plant 
species 
associations 

Individual  plant  species 
occurrence  and  abundance 

RE 

• 

• 

• 

SB 

• 

RA 

UG 

Animal- 
Created 

Burrows 
and  other 
cover  sites 

Cover  site  size,  shape,  animal 
species  activity  patterns,  animal 
species  size 

AM 

• 

• 

RE 

• 

• 

RA 

• 

• 

SM 

• 

• 

CA 

Man-made 
Structures 
(Other  than 
water  and 
agricultural) 

Bridges, 

towers, 

electrical 

lines, 

buildings, 

mine  shafts, 

posts, 

fences 

Size,  type,  frequency. 

AL 

• 

• 

• 

• 

• 

• 

Codes  for  major  vertebrate  associations: 

AL  —  all  wildlife 
AM —  amphibians 
BA  —  bats 
BG  —  big  game 
CA  —  carnivores 
Fl   —fish 
RA  —  raptors 
RE  —  reptiles 
SB  —  small  birds 
SM —  small  mammals 
UG—  upland  game 
WA —  waterfowl 


132 


Deserts 


Table  3.  Important  macrohabitat  components/factors  in  deserts. 


Macrohabitat 
Components/Factors 

Description 

Variables/Factors 

Microhabitats  Affected 

Slope 

%  angle  of  area  from 
horizontal 

%  slope,  moisture  avail., 
thermal  regimes,  vegetation 
structure 

Water,  soils,  vegetation 
debris,  vegetation  (live  and 
dead) 

Aspect 

South,  north,  east,  and  west 
facing 

Direction,  moisture  avail., 
temperature  regimes, 
vegetation  structure 

Water,  soils,  vegetation 
debris,  vegetation  (live  and 
dead) 

Elevation 

Vertical  above  or  below  sea 
level 

Vertical  distance,  moisture 
avail.,  thermal  regimes, 
vegetation  structure 

Water,  vegetation  debris, 
vegetation  (live  and  dead) 

Precipitation 

All  forms 

Quantity,  type,  duration, 
frequency,  moisture  avail., 
thermal  regimes,  vegetation 
structure 

Water,  soils,  vegetation 
debris,  vegetation  (live  and 
dead) 

Ecotones/Habitat 
Juxtaposition 

Habitat  interfaces  and 
locations 

Heterogeneity,  interface  size, 
quantity,  position 

Vegetation  debris  and 
vegetation  (live  and  dead) 

Geographic  Location 

Major  geographic  boundaries 
and  barriers,  habitat  location 

Size,  location,  and  frequency, 
habitat  size,  disjunction, 
influence  of  major  faunal 
groups  in  adjacent  areas 

Direct  effect  on  species 
reproductive  capabilities 

Human  Influences 

Disturbance  and  alterations  of 
habitat  due  to  man 

Habitat  loss  and  structural 
change 

Loss  of  habitat  necessary  to 
meet  ecological  needs.  Less 
diverse  structure  associated 
with  habitat  alterations 

Table  4.  Abundance  and  diversity  of  habitat  components  in  four  deserts. 

Code  for  abundance  and  diversity:    #  =  high      Q  =  moderate    Q  =  low 


Habitat  Component 

Great  Basin 

Mojave 

Chihuahuan 

Sonoran 

WATER 
Lotic 

Permanent  Lentic 
Temporary  Lentic 

abundance      diversity 

o       o 
o       o 

©      © 

abundance       diversity 

o       o 
o       o 
o       o 

abundance       diversity 

o       o 
o        © 
o        © 

abundance       diversity 

•  © 

•  • 
©                © 

ROCK 

©      © 

•      • 

©      • 

•                • 

SOIL 

•      © 

•      © 

•      • 

•                 • 

VEGETATION 
Litter/Debris 
Dead 
Live 
Plant  Species 

oooo 
oooo 

o        © 

©      © 
©      © 
o       o 

o       © 

©      © 
©      © 
o       o 

©                 • 

•  • 

•  • 

o       o 

ANIMAL-CREATED 

©      © 

©      © 

©      © 

©      © 

Deserts 


133 


Water  (Lotic) 

Running,  permanent  water  (lotic  water)  is  an 
extremely  important  component  in  desert  ecosys- 
tems (Naiman  and  Soltz  1981).  Several  species,  espe- 
cially aquatic  wildlife  such  as  fish,  amphibians,  and 
some  reptiles,  are  physically  dependent  on  lotic 
water  sources  (Jones  1981a;  Lowe  1964;  Minckley 
1973;  Naiman  and  Soltz  1981).  Without  this  compo- 
nent, these  types  of  wildlife  cannot  exist  within  des- 
erts. Aquatic  amphibians,  turtles,  and  snakes  require 
running  water  for  physiological  functions  such  as 
thermoregulation,  water  balance,  escape  cover,  and 
food  (Stebbins  1966). 


Lotic  (running-water)  habitat. 


Substrate  of  running  streams  often  determine 
the  presence  and  abundance  of  certain  fish  species. 
For  example,  the  desert  sucker  (Catostomus  clarki) 
feeds  on  rock-bound  algae  that  is  present  only  in 
streams  and  rivers  with  cobble  bottoms  (Minckley 
1973). 

Aquatic  sites  and  riparian  areas  also  provide  an 
abundant  food  source.  For  example,  certain  raptor 


species  such  as  black  hawks  (Buteogallus  anthraci- 
nus)  and  bald  eagles  (Haliaeetus  leucocephalus)  fre- 
quent deserts  only  where  aquatic  food  is  available 
(see  Bent  1961  and  Millsap  1981).  Generally,  these 
raptors  require  large  fish  such  as  desert  suckers,  and 
these  types  of  prey  are  restricted  to  permanent 
streams,  rivers,  and  lakes.  For  these  raptors  to  suc- 
cessfully capture  prey  and  nest  along  streams,  rivers, 
and  lakes,  vertical  structures  such  as  snags,  cliffs,  and 
trees  must  be  present.  This  is  an  example  of  more 
than  one  habitat  component  that  must  be  available 
for  a  species  to  use  an  area.  It  also  demonstrates  im- 
portance of  habitat  edges  and  land/water  ecotones 
(see  Thomas  et  al.  1979). 

Running  streams  also  provide  drinking  water  for 
many  species.  Many  big  game  species  inhabiting  des- 
erts are  dependent  on  flowing  streams  for  physiologi- 
cal water,  although  lentic  (still)  water  distribution  is 
usually  more  important.  Yoakum  ( 1980)  and  Kind- 
schy  et  al.  ( 1982)  stated  that  American  pronghorns, 
within  the  Great  Basin  Desert,  require  permanent 
water  at  less  than  8-km  (5-mi.)  intervals. 

The  quality  of  water  in  desert  streams  also  limits 
species'  occurrence  and  abundance.  Naiman  and 
Soltz  (  1981 )  discussed  the  importance  of  physio- 
chemical,  heavy  metal,  stream  structure,  and  macro- 
invertebrate  ( benthic )  characteristics  in  determining 
fish  diversity  and  abundance.  Ratios  of  riffles,  runs, 
pools,  and  backwater  subhabitats  also  determine  fish 
occurrence  in  streams  much  like  vertical  and  hori- 
zontal structures  in  terrestrial  habitats  (see  discus- 
sions in  Naiman  and  Soltz  1981 ). 


Water  (Permanent  Lentic) 

Permanent,  lentic  habitats  are  important  to  large 
numbers  of  desert-dwelling  wildlife.  Lentic  dwelling 
fish,  including  introduced  fishes  such  as  bass  (Per- 


»'_. 

ssfif 

Kr 

flfinP 

' 

;           .          ' 

Lentic  (still-water)  habitat. 


134 


Deserts 


cichthyidae  sp. ),  catfish  (Ictaluridae  sp. ),  and  sunfish 
(Centrarchidae  sp.),  are  totally  restricted  to  and  are 
dependent  on  lakes,  stock  ponds,  and  natural  ponds 
such  as  cienegas  within  deserts  (Minckley  1973;  Nai- 
man  and  Soltz  1981 ).  Cienegas,  springs,  bogs,  and 
potholes  provide  the  only  habitat  within  deserts  for 
fish  such  as  desert  pupfish  (Cyprinodon  macularius) 
and  Gila  top  minnows  (Poeciliopsis  occidentalis). 

Aquatic  amphibians  (e.g.,  Rio  Grande  leopard 
frogs,  Rana  berlatideri),  turtles  (e.g.,  mud  turtles, 
Kinosternon  sp. ),  and  snakes  (e.g.,  checkered  garter 
snake,  Thamnophis  marcianus)  are  totally  depen- 
dent on  lentic  waters  within  deserts  (when  lotic 
habitats  are  not  present )  for  reproduction,  food,  es- 
cape cover,  physiological  processes  such  as  thermo- 
regulation and  water  regulation,  and  egg  and  tadpole 
development  (frogs  and  toads). 

Lentic  waters,  like  running  streams,  also  provide 
an  abundant  and  diverse  prey  base.  For  example,  cer- 
tain bats  (Chiroptera)  require  large  amounts  of  in- 
sect prey,  which  in  many  desert  regions  are  pro- 
vided directly  above  lentic  water  sources  such  as 
reservoirs  and  lakes  (see  Barbour  and  Davis  1969  for 
Chiroptera  requirements ).  Raptors  such  as  ospreys 
(Pandion  haliaetus)  and  bald  eagles  are  highly  de- 
pendent on  fish  (Bent  1961;  Millsap  1981),  and  lakes 
are  often  the  only  source  of  these  prey  in  desert 
regions. 

Most  waterfowl  are  limited  to  permanent,  lentic 
habitats  within  deserts  (see  Fleming  1959  for  exam- 
ples). Lentic  habitats  provide  waterfowl  with  food, 
escape  cover,  and  resting  sites. 

Interfaces  of  lentic  habitat  with  adjacent  habitat, 
especially  shoreline  and  emergent  vegetation,  en- 
hance and  increase  use  of  lentic  habitats  by  fish, 
frogs,  toads,  snakes,  waterfowl,  and  small  birds. 
Emergent  and  shoreline  vegetation  increase  nesting 
opportunities  for  frogs,  toads,  and  waterfowl,  and 
provide  excellent  escape  cover  and  feeding  substrate 
for  all  wildlife.  Floating  logs  and  islands  also  provide 
excellent  in-water  structures  for  basking  turtles 
(thermoregulation  and  escape  cover)  and  waterfowl 
(nesting  substrate  and  isolation  from  predators), 
respectively. 

Permanent  natural  or  man-made  lentic  habitats 
(e.g.,  stock  tanks  and  rainwater  catchments)  are  ex- 
tremely important  drinking  waters  for  big  game  spe- 
cies inhabiting  deserts.  Although  relatively  drought 
tolerant,  most  desert  bighorn  sheep  and  pronghorn 
require  free-standing  water,  especially  during  the 
breeding  season  (Wilson  et  al.  1980;  Yoakum  1980). 

Water  quality,  as  discussed  for  lotic  habitat,  is 
equally  important  in  lentic  habitats.  Dissolved  oxy- 
gen, temperature,  pollutant  concentrations,  habitat 
structure  (water  depth  and  horizontal  arrangement), 


and  water  availability  and  distribution  determine  suc- 
cess of  many  wildlife  species  in  desert  habitats  (see 
Naiman  and  Soltz  1981  for  examples). 


Turtle  basking  on  log. 


Man-made  water  catchment. 


Water  (Temporary  Lentic) 

Temporary  surface  water  accumulation  is  com- 
mon throughout  U.S.  deserts,  especially  on  clay  soils 
( slow  percolation )  during  summer  convectional 
storms.  In  particular,  the  Chihuahuan  Desert  is  domi- 
nated by  clay  soils,  providing  large  areas  of  tempo- 
rary surface  water.  Large  amounts  of  surface  water 
are  reflected  in  this  desert's  abundant  and  diverse 
amphibian  fauna  (see  Conant  1978).  Toad  and  sala- 
mander (e.g.,  tiger  salamander,  Ambystoma  ti- 
grimim )  breeding  is  tied  to  summer  rain  and  agri- 
cultural pools.  Water  is  necessary  for  embryonic 
development  of  these  animals  (see  Stebbins  1966  for 
species'  examples ).  These  waters  also  provide  escape 


Deserts 


135 


cover  and  food  for  many  forms  of  wildlife.  For  exam- 
ple, yellow  mud  turtles  (Kinosternon  flavescens)  and 
waterfowl  use  temporary  pools  created  by  summer 
rains  or  agricultural  runoff  for  escape  and  resting 
cover,  and  for  food  (see  Fleming  1959  and  Stebbins 
1966). 

Rocky  and  boulder-strewn  canyons  also  retain 
intermittent  surface  water.  Red-spotted  toads  (Bufo 
punctatus)  use  these  canyon  pools  for  breeding  (see 
Jones  1981a).  These  natural  temporary  catchments 
are  also  important  physiological  water  sources  for 
big  game  such  as  bighorn  sheep,  especially  during 
warm  summer  months  (Wilson  et  al.  1980). 


Temporary  lentic  habitat. 


Rock 

Rock  arrangement,  heterogeneity,  and  proximity 
to  other  habitat  components  affect  nearly  every  spe- 
cies of  wildlife  in  deserts  (see  discussions  of  Thomas 
et  al.  1979  and  Maser  et  al.  1979a). 

Talus  slopes,  cliffs,  and  rock  outcrops  provide 
nesting  and  feeding  substrate,  thermal  and  escape 
cover,  and  resting  sites  for  many  forms  of  wildlife  in- 
cluding small  birds,  lizards,  snakes,  small  mammals, 
big  game,  bats,  predatory  mammals,  and  raptors. 
Rock  structure  is  an  extremely  important  component 
for  desert  bighorn  sheep.  Deep,  rugged  cliffs  are  nec- 
essary for  lambing,  escape,  and  thermal  cover  (Wil- 
son et  al.  1980).  Raptors  also  use  these  sites  as 
perches  for  locating  and  capturing  prey  (Call  1978; 
Bent  1961).  Caves  provide  excellent  resting  sites  for 
bats.  Several  million  bats  have  been  documented  in  a 
single  large  cave  (Connor,  personal  comm.)  because 
caves  provide  stable,  moderate  temperatures 
throughout  the  year  (see  Barbour  and  Davis  1969). 

Type  and  origin  of  rock  also  determine  species' 
use  of  rock  structures  (Maser  et  al.  1979a).  For  ex- 
ample, chuckwallas  (Sauromalus  obesus)  use  rock 


structures  that  flake  and  crack  (mostly  granitic  and 
volcanic  rock )  as  escape  cover,  bloating  themselves 
between  cracked  walls  making  it  extremely  hard  for 
predators  to  remove  them  (Stebbins  1966). 

Arrangement  of  different  size  rock  and  depth  be- 
low the  surface  affect  wildlife  use  in  deserts.  Burrow- 
ing wildlife  are  especially  affected  by  these  factors. 
Generally,  small  burrowing  mammals  such  as  kanga- 
roo rats  (Dipodomys  sp. )  inhabit  only  areas  where 
surface  and  subsurface  rock  is  broken  or  soils  are 
deep.  Other  small  mammals  such  as  pocket  mice 
{Perognathas  sp. )  and  certain  ground  squirrels  (Ci- 
tellus  sp. )  exploit  rocky  hills  and  mountain  slopes 
because  rocks  are  broken  (e.g.,  volcanic  soils),  pro- 
viding excellent  escape,  resting,  and  nesting  cover 
(see  Burt  and  Grossenheider  1976  and  Golley  et  al. 
1975). 

Boulder-strewn  surfaces  (large  boulders)  pro- 
vide excellent  feeding,  nesting,  and  resting  sites  for 


Color  variations  in  collared  lizards  can  be  linked  to  differ- 
ent habitats. 


Another  color  variation  of  collared  lizard. 


136 


Deserts 


many  wildlife.  Desert  tortoises  (Gopherus  agassizii) 
are  common  in  boulder-strewn  areas  of  west-central 
Arizona  (Burge  1979).  Boulders  seem  to  provide  tor- 
toises with  excellent  cover  sites  (located  below 
rocks)  and  a  cryptic  substrate.  Large  boulders  also 
provide  small  birds  and  predatory  lizards  (e.g.,  col- 
lared lizards,  Crotaphytus  collaris )  with  perching 
sites  for  obtaining  food. 

Rocks  are  also  extremely  important  for  thermo- 
regulation in  reptiles,  especially  lizards.  Simon  and 
Middendorf  ( 1976)  showed  that  Yarrow's  spiny  liz- 
ards (Sceloporus  jarrovi)  were  able  to  attain  core 
temperatures  necessary  for  activity  during  winter 
months  because  rocky,  south-facing  slopes  acted  as 
heat  sinks.  Rocky  hillsides  also  provide  caverns  in 
which  rattlesnakes  (Crotalns  sp. )  and  other  snakes 
hibernate  (see  Hirth  et  al.  1969). 

Rocky  slopes  and  hillsides  also  contain  many 
small,  moist  microhabitats  created  by  water  buildup 
and  shading.  Cumulatively,  this  phenomenon  creates 
a  high  degree  of  heterogeneity  in  vegetation  struc- 
ture and  composition  which  in  turn  increases  inver- 
tebrate production  and  diversity  (see  Maser  et  al. 
1979a). 


Soil  depth,  type,  heterogeneity,  and  interface 
with  other  habitat  components  such  as  rock  and  veg- 
etation, determine  subsurface  space  for  wildlife. 
Some  soils  such  as  clay,  and  in  areas  with  bedrock, 
are  impenetrable  to  certain  burrowing  wildlife  (e.g., 
green  toad,  Bufo  debilis).  On  these  sites,  roots  of 
vegetation  are  important  in  breaking  up  rock  and 
soil.  In  parts  of  the  Chihuahuan  and  Sonoran  Deserts, 
creosotebush  breaks  up  clay  soils  and  bedrock,  pro- 
viding wildlife  access  to  otherwise  unavailable  space 
(Mares  and  Hulse  1977;  Barbour  et  al.  1977). 

Similar  to  surface  and  arboreal  edges  created  by 
landform  and  vegetation,  greater  soil  heterogeneity 
and  interfaces  with  other  habitat  components  (e.g., 
rock)  provide  greater  numbers  of  spatial  niches,  result- 
ing in  higher  wildlife  biomass  and  diversity,  especially 
amphibians,  reptiles,  and  small  mammals  (see  Thomas 
et  al.  1979  for  surface  and  arboreal  examples). 

Wildlife  use  subsurface  niches  for  a  wide  variety 
of  ecological  functions,  and  several  species  depend 
on  soil  characteristics  for  exploitation  of  desert  re- 
gions. For  example,  spadefoot  toads  (Scaphiopus  sp. ) 
are  dependent  on  deep,  loose  soils  to  burrow  and  es- 
tivate  (McClanahan  1967).  Spadefoot  toads  remain 


Soil 

The  number  of  subsurface  (fossorial)  niches  in 
deserts  is  largely  dependent  on  soil  characteristics. 
Lack  of  vegetation  structure  in  deserts  is  often  offset 
by  subsurface  space  created  by  deep  and  diverse 
soils,  especially  in  warm  deserts  (see  discussions  of 
Bender  1982,  Brown  1982b,  and  Attenborough 
1976).  Creosotebush  stands  of  the  Chihuahuan  and 
Sonoran  Deserts  have  little  vegetation  structure,  but 
have  a  rich  fauna,  especially  amphibians,  lizards, 
snakes,  and  small  birds  and  mammals  (Bender  1982; 
Brown  1982b),  primarily  because  of  soil  diversity 
and  access  to  subsurface  space. 


Spadefoot  toad. 


Badger  at  burrow. 


burrowed  in  these  soils  for  the  entire  year,  except 
during  warm-season  convectional  storms  when  they 
surface  to  feed  and  breed.  Similarly,  small  mammals 
are  affected  by  soil  type,  depth,  and  heterogeneity 
(Golley  et  al.  1975).  For  example,  many  kangaroo 
rats  are  generally  incapable  of  burrowing  into  hard, 
coarse,  shallow  soil,  except  along  deeply  rooted 
plants  (see  Burt  and  Grossenheider  1976  for  exam- 
ples), and  are  limited  to  deserts  with  deep,  loose, 
and  gravel  substrate. 


Deserts 


137 


In  addition  to  providing  resting  sites,  certain 
soils  provide  nesting,  thermoregulatory  substrate, 
and  escape  cover.  For  example,  fringe-toed  lizards 
{Uma  sp.)  exploit  deserts  with  fine  sandy  soils.  Be- 
sides supplying  nesting  substrate,  fine  textured  sand 
provides  fringe-toed  lizards  cryptic  background  and 
warm  thermal  regimes  necessary  for  activity  (see 
Norris  1958). 

Water  movement  through  different  soils  (perco- 
lation) also  affects  species'  existence  in  deserts. 
Jones  et  al.  (  1983)  and  Conant  (  1978)  showed  that 
deserts  with  clay  soils  accumulated  more  surface 
water  than  areas  with  loose,  gravelly,  and  rocky  soils, 
resulting  in  larger  desert  amphibian  faunas. 

Soil  depth,  origin  (e.g.,  volcanic),  and  heteroge- 
neity also  affect  vegetation  composition  and  struc- 
ture within  deserts  (Davis  1976).  Vegetation  compo- 
sition and  structure  are  discussed  later  in  this 
chapter. 

Vegetation — Litter/Debris 

Ground  debris  or  litter  provides  excellent  feed- 
ing, thermoregulatory,  nesting  substrate,  and  escape 
cover  for  wildlife  inhabiting  desert  regions.  Vegeta- 
tion debris  consists  of  dead  plant  material  (e.g.,  tree 
limbs  and  leaves)  that  accumulate  on  the  surface, 
and  it  is  generally  correlated  with  vegetation  compo- 
sition and  abundance  (Daubenmire  1974).  Litter 
abundance  and  diversity  are  probably  highest  in  the 
Sonoran  Desert  and  least  in  sections  of  the  Great 
Basin  Desert. 

In  deserts,  large  numbers  of  invertebrate  species 
inhabit  ground  litter,  providing  abundant  prey  for 
ground-foraging  amphibians,  lizards,  and  small  birds 
(see  Pianka  1970  and  Tomoff  1974  for  examples). 

Downed  limbs  and  logs  are  important  nest  sub- 
strate for  many  lizards,  snakes,  and  small  mammals, 
providing  cool,  moist,  well-concealed  sites  within 
hot  deserts.  Certain  species  such  as  Gilbert's  skink 
(Eumeces  gilberti)  are  highly  dependent  on  these 
microhabitats  for  maintaining  relatively  low  pre- 
ferred body  temperatures  within  deserts  (Jones  and 
Glinski  1985).  Pack  rats  (Neotoma  sp. )  require  var- 
ious forms  of  litter  to  build  nests,  and  their  occur- 
rence within  desert  regions  seems  to  be  related  to 
occurrence  of  vegetation  debris  (Burt  and  Grossen- 
heider  1976). 

Vegetation  debris  is  extremely  important  in 
aquatic  desert  habitats.  Downed  logs,  limbs,  and 
other  litter  provide  excellent  thermal  and  resting 
cover  for  amphibians,  turtles,  snakes  and  fishes,  and 
contribute  organic  nutrients  to  aquatic  systems  (Nai- 
man  and  Soltz  1981). 

Because  wildlife  respond  to  several  physical 


Pack  rat  nest. 


attributes  of  litter  in  deserts,  it  is  important  to  re- 
cord the  type  of  litter,  depth,  width,  and  frequency. 

Vegetation — Standing — Dead 

Standing,  dead  vegetation  or  snags  provide  ex- 
cellent vertical  structure  for  nesting,  feeding,  resting, 
thermoregulatory  substrate,  and  escape  cover  within 
deserts.  Many  cavity  nesting  birds  (e.g.,  Gila  wood- 
peckers, Melanerpes  uropygialis)  and  raptors  (e.g., 
screech  owl,  Otus  asio)  use  standing,  dead  vegeta- 
tion for  nest  substrate,  roosting,  and  resting  sites 
(see  Knopf  1977  for  examples).  Small  lizards  such  as 
tree  lizards  (Urosaurus  ornatus)  use  standing,  dead 
vegetation  for  foraging,  defense,  and  thermoregula- 
tion substrate  (Vitt  et  al.  1981 ). 

Similar  to  downed  litter,  standing,  dead  vegeta- 
tion abundance  and  diversity  are  positively  associ- 
ated with  vegetation  composition  and  structure  (see 
Daubenmire  1974).  Therefore,  Sonoran  Desert  re- 
gions, especially  the  Arizona  Upland  subdivision, 
probably  possess  the  highest  abundance  and  diver- 
sity of  snags  among  major  U.S.  deserts. 

Vegetation — Live 

Live  vegetation  structure  is  perhaps  the  most 
important  habitat  component  in  desert  regions,  al- 
though soils  and  rock  may  be  more  significant  where 
vegetation  is  low  and  homogeneous.  A  number  of 
vegetation  variables  should  be  sampled  when  con- 
ducting inventories  and  monitoring  in  deserts. 

Arrangement  of  vertical  and  horizontal  vegeta- 
tion structure  in  deserts,  and  its  relations  to  wildlife, 
has  been  thoroughly  studied  (see  Rottenberry  and 
Wiens  1980  for  an  example).  Most  authors  agree 
that  high  wildlife  abundance  and  diversity  accom- 
pany deserts  with  high  horizontal  and  vertical  vege- 


138 


Deserts 


tation  structure,  especially  birds.  Tomoff  (1974)  cor- 
related avian  species  diversity  increases  to  increases 
in  vegetation  structure  within  the  Sonoran  Desert. 
Greater  vertical  structure  increases  the  number  of 
avian  nesting,  feeding,  and  breeding  niches  within 
these  habitats.  Conversely,  deserts  with  little  vertical 
vegetation  structure  provide  few  spatial  niches.  Rap- 
tors are  also  positively  associated  with  vegetation 
structure  (Millsap  1981). 

Certain  lizards  and  bats  also  benefit  from  verti- 
cal vegetation  structure  (Pianka  1966;  Barbour  and 
Davis  1969).  Vitt  et  al.  (1981)  demonstrated  that 
woody  vegetation  was  important  reproductive  and 
feeding  substrate  for  arboreal  lizards  (tree  lizard, 
long-tailed  brush  lizard  [Urosaurus  graciosus],  and 
the  desert  spiny  lizard  [Sceloporus  magister]).  Verti- 
cal vegetation  structure  provides  bats  with  greater 
amounts  of  roosting  and  resting  space  (Barbour  and 
Davis  1969),  and  it  also  increases  abundance  of  prey 
(flying  insects)  for  insectivorous  lizards,  small  birds, 
and  bats.  Vertical  vegetation  structure  is  also  used  by 
lizards  for  thermoregulation.  Norris  (1953)  de- 
scribed climbing  behavior  of  desert  iguanas  (Dipso- 
saurus  dorsal  is)  into  creosote  to  escape  rapidly  in- 
creasing surface  temperatures;  temperatures  1-2  m 
(3-6  ft)  above  the  surface  are  considerably  lower. 

Horizontal  vegetation  structure,  in  conjunction 
with  vertical  structure,  also  affects  species  abun- 
dance and  diversity  (Rottenberry  and  Wiens  1980). 
Generally,  greater  varieties  of  horizontal  patchiness 
are  associated  with  greater  wildlife  diversity.  Rotten- 
berry  and  Wiens  ( 1980)  suggested  that  certain  birds 
prefer  either  open  or  closed  habitats.  By  providing 
highly  variable  horizontal  structure,  desert  habitats 
support  species  that  prefer  both  open  and  closed 
habitats.  Deserts  possessing  homogeneous,  horizontal 
arrangement,  and  horizontal  structure  that  is  ex- 
tremely opened  or  closed  probably  have  lower  avian 
diversity  than  those  that  are  structurally  variable. 

Similarly,  lizards  and  snakes  respond  to  horizon- 
tal patchiness,  although  their  relationship  to  patchi- 
ness is  related  to  foraging  and  thermoregulatory  re- 
quirements. Certain  lizard  species  require  alternating 
open  and  closed  habitat  configurations  (  Pianka 
1966).  These  lizards  forage  in  open  spaces  and  re- 
treat to  large,  dense  bushes  between  foraging  bouts 
to  lower  their  body  temperatures.  Deserts  transected 
by  intermittent  washes  provide  these  alternative 
structures. 

Horizontal  patchiness  also  conceals  nests  of  up- 
land game  such  as  Gambel's  quail  {Lophortyx  gam- 
belli)  and  sage  grouse,  and  provides  escape  cover 
for  individual  birds  (see  Goodwin  and  Hungerford 
1977  for  Gambel's  quail  and  Eng  and  Schladweiler 
1972  for  sage  grouse). 

Composition  and  availability  of  vegetation  in  des- 


erts affects  species'  distribution  and  diversity.  Pres- 
ence and  abundance  of  certain  plant  species  affect 
horizontal  and  vertical  structure  (as  previously  dis- 
cussed), and  also  provide  food  for  commercially  im- 
portant species  such  as  bighorn  sheep,  Gambel's 
quail,  herbivorous  lizards,  birds,  and  mammals. 

Availability  of  perennial  and  annual  vegetation 
affects  bird  species  abundance  and  diversity  in  desert 
regions  (Serventy  1971;  Tomoff  1974).  Generally, 
bird  species'  diversity  increases  with  greater  mix- 
tures of  annual  and  perennial  vegetation  because  of 
great  variety  in  avian  food  preferences  (Martin  et  al. 
1951).  Similarly,  abundance,  type,  and  diversity  of 
vegetation  affects  food  available  to  herbivorous,  des- 
ert rodents;  greatest  herbivorous  rodent  diversity  ex- 
ists on  sites  with  high  annual  and  perennial  diversity 
and  mixture  (Black  1968;  Brown  1973). 

Big  game  use  of  desert  regions  are  affected  by 
plant  species  composition  and  abundance.  For  exam- 
ple, desert  bighorn  sheep  require  a  high  diversity  of 
forbs,  perennial  grass,  and  shrubs  in  desert  habitats 
to  meet  basic  metabolic  needs  (Wilson  et  al.  1980). 


Mi*53*'        ■    ■'  ***&* 


i$ 


Desert  sheep  select  areas  near  rocky  escape  cover. 


Although  wildlife  diversity  generally  increases 
with  increases  in  plant  species  diversity,  increased 
plant  diversity  may  reduce  numbers  of  important 
game  species.  For  example,  Eng  and  Schladweiler 
(  1972)  found  that  sage  grouse  preferred  closed 
stands  of  sagebrush. 

Extensive  root  systems  of  certain  plants,  such  as 
creosotebush,  provide  access  to  subsurface  space  for 
toads,  salamanders,  lizards,  snakes,  and  small  mam- 
mals. Species  diversity  is  often  increased  in  areas 
with  extensive  plant  root  systems. 


Deserts 


139 


Vegetation — Plant  Species  Associations 

Certain  wildlife  species  have  evolved  morpho- 
logical, physiological,  and  behavioral  characteristics 
that  are  adaptive  to  certain  plant  species  attributes. 
These  should  be  carefully  considered  when  sampling 
desert  wildlife.  There  are  several  good  examples  of 
these  relationships  in  deserts.  Desert  iguanas  forage 
heavily  on  creosotebush  buds,  especially  in  the 
spring,  and  their  distribution  is  closely  tied  to  occur- 
rence of  creosotebush  (Norris  1953).  Sage  grouse 
rely  heavily  on  sagebrush  buds,  especially  in  the 
spring,  and  this  species  is  highly  associated  with 
dense  stands  of  sagebrush  throughout  its  range  ( Eng 
and  Schladweiler  1972).  Although  both  the  desert 
iguana  and  sage  grouse  are  dependent  on  these  plant 
species  for  food,  they  also  benefit  from  escape  and 
thermoregulatory  cover  provided  by  these  plants. 


Wintering  sage  grouse  feed  extensively  on  sagebrush  buds 
and  seedheads. 


Cactus  wren  nest  in  cholia  cactus. 


Certain  plant  species  also  provide  nesting  struc- 
ture for  desert  wildlife.  Occurrence  of  elf  owls  (Mi- 
crathene  whitneyi)  in  the  Sonoran  Desert  is  associ- 
ated with  nesting  cavities  presence  in  saguaro  cactus 
(Knopf  1977;  Bent  1961 ).  Similarly,  cactus  wrens 
(Campylorhynchus  brunneicapillus)  are  associated 
with  cactus  and  yuccas,  primarily  because  these 
plants  provide  nesting  structure  (Knopf  1977). 

Animal-Created  Structures 

Structures  created  by  certain  wildlife  provide 
other  desert  animals  with  habitat  necessary  for 
breeding,  feeding,  escape,  resting  and  thermoregula- 
tory substrate,  and  escape  cover. 

Fossorial  mammals  construct  vast  networks  of 
subsurface  burrows  that  provide  resting  and  nesting 
habitat  structure  for  a  number  of  desert  dwelling 
wildlife  such  as  toads,  lizards,  snakes,  and  raptors 
(burrowing  owl  {Athene  cunicularia]).  Subsurface 
space  contributes  greatly  to  species  diversity  in  des- 
erts with  little  vegetation  structure. 

Primary  cavity  nesting  species  such  as  wood- 
peckers provide  nesting  habitat  in  cactus  and  woody 
vegetation  (e.g.,  mesquite)  for  a  number  of  second- 
ary cavity  nesting  birds  such  as  elf  owls  (see  Knopf 
1977  for  examples). 

Beavers  modify  flow  of  desert  streams,  providing 
pool  and  backwater  habitat,  and  enhance  populations 
of  fish,  toads,  frogs,  aquatic  turtles,  shorebirds,  and 
waterfowl  (Burt  and  Grossenheider  1976).  It  is  im- 
portant to  consider  animal-created  structures  when 
attempting  to  sample  and  predict  wildlife  occurrence 
within  deserts. 

Man-Made  Structures 

Man-made  structures  have  both  positive  and 
negative  effects  on  wildlife  in  desert  environments. 
Although  power  lines  provide  nesting,  resting,  and 
roosting  structure  for  birds  and  raptors  (e.g.,  red- 
tailed  hawk,  Buteo  jamaicensis),  they  also  cause 
mortality  due  to  electrocution  and  obstruction  (see 
Olendorff  et  al.  1981). 

Man-made  structures  such  as  buildings,  barns, 
mine  shafts,  and  fences  provide  a  wide  range  of  nest- 
ing, feeding  and  thermoregulatory  substrate,  and  es- 
cape cover  for  wildlife  such  as  barn  owls  (Tyto 
alba),  fence  lizards  (Sceloporus  sp. ),  bats,  mourning 
doves  (Zenaida  macroura),  and  house  mice  (Mus 
mmculus).  Vertical  mine  shafts  and  fences  have 
some  negative  effects  on  desert  wildlife.  Many  small 
wildlife  species  are  trapped  in  shafts,  and  fences  pre- 
sent physical  hazards  and  movement  restriction  to 
big  game  species  (see  Yoakum  1980  for  pronghorn 
examples). 


140 


Deserts 


A  number  of  man-made  structures  such  as  water 
catchments  and  nesting  platforms  have  been  used  in 
desert  regions  to  enhance  wildlife.  Maser  et  al. 
( 1979b)  provided  an  excellent  synopsis  of  man- 
made  habitats  and  their  effect  on  wildlife.  All  man- 
made  habitats  should  be  considered  when  sampling 
deserts.  They  often  affect  the  presence  or  absence  of 
certain  species  within  deserts. 

Slope 

Because  of  diverse  topography,  slopes  vary 
greatly  within  desert  habitats.  Slope  variability  and 
mixture  in  deserts  contribute  greatly  to  overall  habi- 
tat variability  and  diversity  of  wildlife  faunas  (see 
Maser  et  al.  1979a  for  examples).  Wind  currents,  di- 
rect and  indirect  heating  of  substrate,  and  moisture 
retention  associated  with  different  slope  angles  di- 
rectly (animal  heat  and  water  exchange,  and  ease  of 
movement)  and  indirectly  (vegetation  differences) 
affect  wildlife  (see  Maser  et  al.  1979a).  The  interrela- 
tionship between  slope  and  soils,  vegetation,  aspect, 
and  elevation  determines  the  physical  structure  of 
desert  ecosystems.  Therefore,  effects  of  slope  on 
physical  structures  in  deserts  will  vary,  depending  on 
interfaces  with  these  other  habitat  components.  Dif- 
ferences in  slopes  often  contribute  to  presence,  ab- 
sence, and  abundance  of  certain  wildlife  species. 
Whereas  steep,  rugged  slopes  are  generally  a  require- 
ment of  bighorn  sheep  (Wilson  et  al.  1980),  they 
preclude  occurrence  of  many  other  species  such  as 
horned  lizards  (Phtynosoma  sp. ).  Slopes  should  be 
measured  and  described  when  conducting  invento- 
ries and  monitoring  in  deserts. 


The  Sonoran  Desert  exhibits  many  diverse  habitat  features. 


Aspect 

Aspect  involves  the  compass  direction  of  land- 
scape. Similar  to  slope,  aspect  affects  moisture  availa- 
bility, air  and  soil  temperatures,  which  in  turn  affect 
vegetation  composition  and  structure,  and  wildlife 
thermal  and  moisture  regimes  (see  previous  discus- 
sions on  these  microhabitat  components  for  effects 
on  desert  wildlife).  Effects  of  aspect  on  desert  envi- 
ronments also  depend  on  interfaces  with  microhabi- 
tat components,  especially  soil  and  vegetation,  and 
macrohabitat  factors  such  as  slope,  precipitation,  and 
elevation.  Generally,  vegetation  structure  is  greatest 
along  north-  and  east-facing  slopes  because  of  lower 
temperatures  and  higher  moisture  regimes.  However, 
soil  type,  rock,  and  elevation  may  produce  contrary 
results.  Aspect  should  be  considered  when  selecting 
sample  locations  within  deserts. 

Elevation 

Elevation  affects  moisture  availability  and  temper- 
atures that  in  turn  affect  vegetation  structure  and  com- 
position (see  Lowe  1964  for  examples).  It  also  affects 
thermal  and  moisture  regimes  of  air,  soil,  and  litter, 
which  can  affect  wildlife  (see  earlier  discussions). 


Generally,  higher  elevations  have  cool  weather, 
greater  soil  and  litter  moisture,  and  more  precipita- 
tion. Increased  vegetation  structure  resulting  from 
greater  moisture  and  cooler  weather  can  be  offset  by 
low  temperatures  associated  with  higher  elevations. 
For  example,  high  desert  regions  of  Wyoming  have 
very  few  species  of  amphibians,  reptiles,  and  small 
mammals  because  environmental  temperatures  are 
too  low  for  species'  thermal  maintenance.  Food  nec- 
essary for  birds  and  other  herbivorous  and  insectivo- 
rous wildlife  are  too  infrequent  (short  growing  sea- 
son) for  physiological  maintenance  in  these  cold 
desert  regions. 

Similar  to  slope  and  aspect,  elevational  attributes 
are  affected  by  association  with  soil,  vegetation, 
slope,  aspect,  and  degree  of  mountain  rain  shadow. 

Precipitation 

Precipitation  varies  greatly  within  deserts  (see 
Bender  1982),  and  this  directly  affects  vegetation 
structure  and  moisture  available  to  wildlife  in  air, 
soils,  and  litter  (see  previous  discussions  on  these 
microhabitat  components;  Oosting  1956).  Precipita- 


Deserts 


141 


tion  within  deserts  is  primarily  affected  by  mountain 
rain  shadow  and  elevation  (see  Bender  1982). 

Frequency  and  type  of  precipitation  also  affect 
vegetation  structure  and  moisture  available  to  wild- 
life. Northern  deserts  (Great  Basin)  receive  large 
amounts  of  snow  (Bender  1982).  Moisture  is  gener- 
ally not  available  during  warm  spring  and  summer 
months  for  plant  growth  and  wildlife  consumption. 
Conversely,  southern  deserts  that  receive  large  pro- 
portions of  their  precipitation  as  rain  (e.g.  Chihua- 
huan  and  Sonoran  Deserts)  in  summer  months  (see 
Bender  1982),  provide  water  for  plant  growth,  insect 
production,  amphibian  breeding,  and  wildlife  con- 
sumption. Large  snow  packs  at  higher  elevations  in 
the  Great  Basin  Desert  (e.g.,  Wyoming)  also  reduce 
plant  food  available  to  wildlife,  except  on  south-fac- 
ing winter  ranges  (see  Skovlin  1982  for  an  example). 

Because  precipitation  quantity  type  and  fre- 
quency can  radically  affect  wildlife  activity,  year-to- 
year  fluctuations  in  precipitation  are  important  when 
interpreting  inventory  and  monitoring  results. 

Ecotones/Habitat  Juxtaposition 

Throughout,  I  have  mentioned  how  edges  and 
position  of  habitats  affect  habitat  structure  and  wild- 
life diversity.  Generally,  wildlife  diversity  is  highest 
in  deserts  with  high  mixtures  of  different  microhabi- 
tat  components,  and  favorable  thermal  and  moisture 
regimes  (warm  and  moist).  The  Sonoran  Desert  fully 
demonstrates  this  relationship.  Great  mixtures  of  to- 
pography, soils,  rock  and  vegetation  structure,  warm 
thermal  regimes,  and  moisture  provide  many  differ- 
ent spatial  and  temporal  niches  to  wildlife  (reflected 
in  one  of  the  world's  most  diverse  wildlife  faunas ). 

Ecotones  consisting  of  habitat  elements  from  des- 
erts and  adjacent  habitats  (e.g.,  pinyon-juniper)  cre- 
ate distinct  physical  elements  (e.g.,  structure)  that 
enhance  species  diversity  (Thomas  et  al.  1979).  For 
example,  breeding  bird  diversity  is  increased  in  eco- 
tones between  sagebrush  and  pinyon-juniper  habitats 
because  of  greater  habitat  structural  diversity  (both 
tree  and  shrub  life  forms). 

Geographic  Location 

Geographic  location  and  past  geological  history 
account  for  present  species  distribution  and  diversity 
of  some  wildlife  within  desert  regions.  For  example, 
species  diversity  in  the  Sonoran  Desert  is  greatly  en- 
hanced by  contributions  of  adjacent  subtropical  and 
tropical  avifauna  ( Brown  1982b;  Bender  1982).  The 
evolution  of  the  Sonoran  Desert  from  mostly  subtropi- 
cal vegetation  also  accounts  for  occurrence  of  other 
wildlife  with  more  southerly  ties.  Similarly,  Great  Basin 
wildlife  diversity  is  increased  by  species  from  adjacent 
pinyon-juniper  and  grassland  communities. 


Geographic  boundaries  associated  with  deserts 
are  also  responsible  for  present  distributions.  Abrupt 
mountain  ranges  and  canyons  (e.g.,  Grand  Canyon) 
have  isolated  certain  desert  wildlife  (Golley  et  al. 
1975).  Current  scattered  fish  distributions  within 
many  desert  regions  reflect  a  drying  trend  since  the 
last  ice  age  (10,000  years  ago;  Naiman  and  Soltz 
1981 ).  Many  streams  have  dried  up,  leaving  disjunct 
fish  populations. 

Past  and  present  weather  and  associated  vegeta- 
tion patterns  have  also  affected  species'  occurrences 
throughout  deserts.  Van  Devender  and  Spaulding 
(1979)  demonstrated  recent  drying  trends  in  the 
West  (past  10,000  years),  resulting  from  recession  of 
the  last  ice  age.  This  drying  trend  left  relict  and  dis- 
junct woodland  on  mountains  throughout  southwest- 
ern deserts.  Several  woodland  vertebrates  remain  on 
these  relict  sites,  with  diversity  dependent  on  habitat 
size  and  distance  from  major  stands  (see  Martin 
1980,  1981;  Jones  et  al.  1985). 

Human  Influences 

Man  has  drastically  affected  plant  and  animal 
composition  and  diversity  throughout  U.S.  deserts.  Ir- 
rigation, urbanization,  and  other  developments  have 
severely  reduced  natural  desert  habitats.  Artificial,  ur- 
ban, and  agricultural  habitats  support  fewer  species 
than  do  native  surrounding  deserts  (Davis  1973; 
Landcaster  and  Rees  1979).  Other  land  uses  have 
also  reduced  quality  of  desert  habitats  such  as  off- 
road  vehicles,  livestock  grazing,  and  mineral  develop- 
ment (Webb  and  Wilshire  1983;  Jones  1981b;  and 
Dunaway  1971).  Fire  prevention  and  channelization 
of  desert  streams  and  rivers  have  also  reduced  natu- 
ral characteristics  in  desert  ecosystems.  See  the  fol- 
lowing section  for  greater  detail  on  impacts  to  desert 
habitats. 


MAJOR  IMPACTS  ON  DESERT  HABITATS 

There  are  two  major  types  of  impacts  on  desert 
habitats,  natural  and  man-caused.  I  have  already  dis- 
cussed, with  the  exception  of  fire  and  flooding,  how 
natural  factors  affect  wildlife  distributions  and  com- 
position in  deserts. 

Fire  affects  wildlife  within  deserts  by  altering 
habitat  structure  (Wright  and  Bailey  1982).  Gener- 
ally, vegetation  on  a  site  is  limited  to  annual  grasses 
and  sprouts  from  perennial  shrubs  and  grasses  up  to 
4  years  after  a  fire.  This  favors  species  that  prefer 
open  sites  and  areas  with  annual  grasses  and  forbs, 
and  is  less  favorable  to  species  requiring  shrubs  and 
trees.  However,  fires  in  deserts  do  not  generally 
carry  over  large  areas;  burns  are  usually  patchy 
(Wright  and  Bailey  1982).  This  result  tends  to  in- 
crease species  diversity  in  an  area  due  to  increased 
horizontal  and  vertical  vegetation  structural  diver- 


142 


Deserts 


sity.  For  example,  in  deserts  such  as  the  sagebrush 
communities  of  the  Great  Basin  Desert,  fire  opens  up 
homogenous  vegetation  stands. 

Fire  does  not  seem  to  result  in  high  direct  mor- 
tality to  wildlife  (Vogl  1977).  Most  animals  are  able 
to  escape  rapidly  burning,  low  intensity  desert  fires. 
In  summer  months  following  a  wet  winter  or  spring, 
annual  grasses  are  more  abundant,  resulting  in  larger 
and  more  frequent  fires.  However,  since  fuels  in  des- 
erts are  mostly  annual  grass,  fires  burn  very  quickly 
over  a  site,  with  little  direct  wildlife  mortality. 

Flooding  is  another  natural  phenomena  that  af- 
fects wildlife  habitat,  especially  in  desert  riparian  sys- 
tems. For  the  most  part,  flooding  is  beneficial  to 
wildlife  habitat.  In  cottonwood-willow  (Populns- 
Salix  sp. )  communities  of  the  Southwest,  flooding  is 
required  for  establishment  of  deciduous  tree  seed- 
lings. When  flooding  is  reduced  or  eliminated  by  in- 
stream  structures  such  as  dams,  reproduction  of 
these  important  tree  species  are  reduced  or  elimi- 
nated. Reduction  or  loss  of  trees  will  reduce  wildlife 
diversity  on  the  site. 


Flooding  occurs  as  a  result  of  short,  intense  summer  thun- 
derstorms. Most  of  this  precipitation  runs  off  without  pen- 
etrating the  soil. 


Similar  to  fire,  species  that  occupy  riparian  habi- 
tats evolved  with  these  natural  perturbations.  There 
is  probably  little  direct  mortality  of  wildlife  during 
most  floods,  although  100-  and  500-year  floods  may 
affect  sedentary  wildlife  such  as  small  mammals  and 
reptiles. 

Like  natural  perturbations  of  desert  habitats, 
man-caused  alterations  affect  wildlife  both  directly 
and  indirectly.  Noise  and  activity  from  off-road  vehi- 
cle use  and  mining  activity  such  as  drilling  have 
been  shown  to  affect  reproductive  success  of  wildlife 
within  an  area.  For  example,  drilling  activities  con- 
ducted in  a  bighorn  sheep  lambing  ground  may 


cause  ewes  to  abort  their  young  (Wilson  et  al. 
1980).  Similarly,  Bury  and  Luckenbach  ( 1983)  found 
that  continual  off-road  vehicle  use  in  an  area  reduces 
the  number  of  breeding  birds. 

Fences  and  roads  also  have  a  direct  effect  on 
wildlife  within  desert  habitats.  Both  types  of  devel- 
opment reduce  or  eliminate  wildlife  movement 
within  an  area.  For  example,  the  fitness  of  a  prong- 
horn  population  in  a  desert  area  can  be  reduced  by 
cutting  off  natural  migrational  routes  by  fencing 
(Yoakum  1980). 

The  most  extensive  and  severe  impacts  to  wild- 
life are  those  that  occur  from  loss  of  habitat  and  hab- 
itat quality.  Of  all  habitats,  deserts  are  probably  the 
most  severely  affected  by  domestic  livestock  grazing. 
Low,  erratic  precipitation  and  extreme  environmen- 
tal temperatures  reduce  the  ability  of  most  desert 
plants  to  handle  persistent  livestock  grazing,  espe- 
cially around  water  developments  where  livestock 
tend  to  congregate.  In  addition,  the  Mojave,  Sonoran, 
and  Chihuahuan  Deserts  historically  had  sparse  wild 
ungulate  populations  (Schmidt  and  Gilbert  1978). 
Therefore,  most  desert  plants  have  not  evolved  with 
grazing  pressures.  An  indicator  of  excessive  livestock 
use  in  most  deserts  is  reduced  amounts  of  forbs  and 
perennial  grasses.  In  some  parts  of  the  Sonoran  Des- 
ert, perennial  grasses  are  absent,  primarily  due  to 
long-term,  year-round  livestock  use.  Other  condi- 
tions such  as  reduced  shrub  and  tree  cover  (Sonoran 
Desert)  and  plant  species  diversity  can  indicate  ex- 
cessive livestock  use. 

The  direct  impact  of  livestock  on  desert  vegeta- 
tion results  in  reduced  wildlife  diversity  primarily 
due  to  habitat  structural  losses.  Excessive  livestock 
grazing  also  reduces  food  available  to  herbivorous 
wildlife,  often  resulting  in  reduced  wildlife  abun- 
dance and  diversity. 

Other  impacts  that  indirectly  affect  wildlife 
through  habitat  disturbance  or  loss  include  mining, 
road  construction,  off-road  vehicle  use,  and  other  ac- 
tivities that  cause  surface  disturbance. 

In  addition  to  the  type  of  land  use,  the  severity 
of  the  impact  depends  on  the  size  of  the  develop- 
ment relative  to  the  total  area  of  habitat  and  the  du- 
ration of  the  project.  Generally,  a  short-duration  pro- 
ject that  disturbs  only  a  small  portion  of  the  total 
habitat  does  not  significantly  reduce  the  area's  over- 
all species  diversity.  However,  even  the  smallest  de- 
velopment can  significantly  reduce  or  eliminate  a 
species  that  historically  depends  on  a  specific  area 
(e.g.,  bighorn  sheep  lambing  ground  within  the  So- 
noran Desert  or  elk  winter  range  on  a  south  slope 
within  the  Great  Basin  Desert). 

When  a  land  use  modifies  or  eliminates  a  large 
portion  of  a  habitat,  even  for  a  short  period  of  time, 


Deserts 


143 


species  diversity  can  be  drastically  reduced.  Recent 
studies  in  animal  biogeography  indicate  that  species 
require  some  minimum  habitat  size  (see  Martin 
1980,  1 981;  Jones  et  al.  1985). 

When  a  large  portion  of  a  habitat  is  lost,  certain 
species  are  extirpated  because  habitat  size  is  re- 
duced below  that  required  to  maintain  minimal  via- 
ble populations.  This  type  of  impact  is  most  acute  in 
small,  isolated  habitats  where  loss  of  only  a  few 
hundred  acres  represents  a  significant  portion  of  the 


Off-road  vehicle  damage  to  wildlife  habitats. 


total  habitat.  For  example,  disturbance  and  loss  of 
sand  dune  habitat  in  the  Algodones  dunes  of  south- 
eastern California  from  intensive  off-road  vehicle  use 
has  significantly  reduced  fringe-toed  lizard  popula- 
tions (Bury  and  Luckenbach  1983).  Other  small,  iso- 
lated habitats  within  North  American  deserts  whose 
faunas  can  be  drastically  reduced  by  small  man- 
caused  perturbations  include  woodland  on  top  of 
mountains  and  riparian  sites.  Jones  et  al.  (  1985) 
found  that  reptile  diversity  on  island  woodlands  of 
the  Sonoran  and  Mojave  Deserts  were  highly  depen- 
dent on  the  size  of  the  island. 

In  summary,  impacts  of  land  use  on  desert  habi- 
tats are  generally  longer-lived  than  in  other  ecosys- 
tems, primarily  because  low,  erratic  precipitation 
precludes  rapid  recovery. 


SUMMARY  OF  FACTORS  AFFECTING 
SAMPLING  IN  DESERTS 

Several  factors  affect  sampling  of  desert  wildlife 
faunas  and  their  habitats.  First  and  most  important, 
biologists  must  be  aware  that  most  desert  wildlife  re- 
spond quickly  to  environmental  changes,  especially 
amphibians,  reptiles,  and  small  mammals.  The  life-his- 
tory strategies  of  most  of  these  animals  are  "boom 


and  bust,"  taking  advantage  of  favorable  environmen- 
tal conditions  (Attenborough  1976).  Because  of 
these  strategies,  results  of  surveys  and  studies  may 
vary  directly  with  seasonal  and  yearly  climatic  condi- 
tions under  which  samples  were  taken.  Therefore,  bi- 
ologists should  be  cautious  in  interpreting  inventory 
and  monitoring  studies  in  deserts.  A  one-time  inven- 
tory of  an  area,  preceded  by  a  dry  year,  may  fail  to 
verify  several  species.  Similarly,  results  of  a  monitor- 
ing study  may  be  affected  by  fluctuations  in  precipi- 
tation. For  example,  a  biologist  attributes  reduced 
abundance  of  small  mammals  on  a  site  to  increases 
in  off-road  vehicle  use,  when  in  fact,  reduction  in 
small  mammal  numbers  is  caused  by  a  previously  dry 
year.  It  is  very  important  for  one  to  consider  re- 
sponses of  animals  to  environmental  fluctuations 
when  designing  wildlife  studies  in  deserts. 

In  collecting  data  on  distribution  and  habitat  as- 
sociations of  desert  wildlife,  biologists  must  be  aware 
of  the  diverse  ways  in  which  these  animals  use  des- 
ert resources,  especially  when  deciding  which  habi- 
tat components  to  measure.  For  example,  soil  types, 
vegetation  debris,  and  water  availability  play  large 
roles  in  desert  species  diversities. 


Microhabitat  components  should  be  considered 
when  selecting  sample  site  location.  For  example, 
the  Arizona  Upland  subdivision  of  the  Sonoran  Des- 
ert has  animals  arranged  in  several  microhabitats. 
Therefore,  to  adequately  sample  the  faunas  of  this 
area,  the  biologist  should  select  sample  sites  so  that 
all  microhabitats  are  represented  (see  earlier  discus- 
sion for  important  components). 

Because  animal  abundance  and  diversity  also 
vary  with  aspect  and  slope  within  deserts,  it  is  im- 
portant to  select  sample  sites  that  will  represent  vari- 
ation in  these  factors. 

Sample  size  for  inventories  and  monitoring  stud- 
ies within  deserts  vary  with  habitat  size  and  hetero- 
geneity. For  example,  fewer  samples  are  needed  to 
represent  large,  homogeneous  stands  of  sagebrush  in 
the  Great  Basin  Desert  or  creosote  in  the  Chihua- 
huan  Desert,  than  in  relatively  diverse  stands  of  the 
Sonoran  Desert.  Generally,  larger  areas  require  more 
samples  than  smaller  areas  to  compensate  for  area-re- 
lated sample  variability. 


Temporal  ( time )  partitioning  of  desert  resources 
also  plays  a  large  role  in  the  composition  of  desert 
wildlife  (see  Creusere  and  Whitford  1982).  There 
are  many  different  activity  phases  of  desert  wildlife, 
especially  nocturnal  ones.  One  must  become  familiar 
with  different  activity  periods  to  obtain  complete 
lists  of  desert  wildlife,  or  to  study  responses  of  cer- 
tain species  to  land  management. 


144 


Deserts 


Certain  sampling  techniques  are  more  effective 
in  deserts  than  others.  Because  many  reptiles,  am- 
phibians, and  small  mammals  are  nocturnal,  these  an- 
imals are  most  effectively  sampled  by  live  and  kill 
trapping,  although  roadriding  is  also  a  popular  way 
of  sampling  these  animals.  Trapping  also  allows  the 
biologist  to  capture  secretive,  burrowing  forms. 

Although  effective  in  woodland  and  grassland 
habitats  with  large  big  game  herds,  pellet  group  tran- 
sects are  not  effective  for  sampling  the  smaller  herds 
of  desert  big  game  species.  For  example,  in  an  area 
with  10-15  deer  per  section,  several  hundred  1/100- 
acre  circular  plots  would  be  needed  to  adequately 
sample  pellet  groups.  Aerial  count  methods  may  pro- 
vide a  better  estimate  of  big  game  density. 

Because  most  deserts  are  dominated  by  shrubs, 
and  to  a  lesser  extent  trees  and  grasses,  I  recom- 
mend using  a  line-intercept  transect  method,  or 
point-intercept  method  along  a  line  to  sample  wild- 
life habitat  variables  within  deserts.  This  will  give  a 
good  estimate  of  habitat  frequency,  abundance,  and 
horizontal  and  vertical  cover.  When  monitoring,  sam- 
ple points  should  be  permanently  marked.  If  grass 
and  forb  estimates  are  needed,  a  combination  of  a 
point-intercept  along  a  line  and  plots  is  useful.  If 
shrub  density  is  desired  (e.g.,  for  sage  grouse),  a 
point-center  quarter  method  is  recommended.  See 
Chapter  5  for  greater  detail  of  habitat  measurement 
techniques. 


Great  Basin  Desert  inventories  and  monitoring 
should  be  conducted  between  May  and  September, 
except  when  working  with  wintering  big  game  pop- 
ulations. In  the  Mojave,  Chihuahuan,  and  Sonoran 
Deserts,  sampling  is  most  effective  between  April 
and  June,  and  again  in  September  and  October,  al- 
though certain  raptors  nest  in  the  late  winter  and 
early  spring  within  these  deserts.  Summer  months 
are  generally  too  warm  to  sample  within  southern 
U.S.  deserts. 


Summer  rains  within  deserts  also  provide  an  ex- 
cellent opportunity  to  sample  many  small  nongame 
wildlife.  For  example,  many  burrowing  snakes  and 
toads  are  active  immediately  after  rain.  Toads  com- 
monly breed  in  temporary  rainwater  within  deserts. 
Temporary  rainwater  pools  should  be  checked  at 
night  to  verify  certain  toad  species. 

Deserts  are  extremely  warm  during  summer 
months  and  can  be  bitter  cold  during  winter,  espe- 
cially in  the  Great  Basin  Desert.  Generally,  most 
wildlife  greatly  reduce  their  activity  during  these  pe- 
riods. Surveys  and  studies  should  be  scheduled 
around  extreme  environmental  conditions. 

Most  desert  regions  are  remote  with  limited  ac- 
cess and  few  towns.  A  wide  variety  of  transportation 
and  sampling  schemes  may  be  necessary  to  counter- 
act limited  access.  For  example,  helicopters  provide 
rapid  and  efficient  means  of  mapping  desert  habitats. 
One  might  also  consider  placing  several  samples 
closely  together  rather  than  spreading  them  out,  pro- 
vided random  sampling  criteria  are  considered. 

Rattlesnakes  are  very  common  within  deserts, 
especially  in  southern  deserts  during  the  spring.  Care 
should  be  taken  when  moving  through  densely  vege- 
tated areas  while  walking  bird  transects  in  the  early 
morning  or  while  setting  small  mammal  trap  lines  in 
the  early  evening. 

As  in  other  habitats,  inventories  and  monitoring 
studies  of  deserts  are  affected  by  objectives.  Chapter 
1  provides  a  discussion  of  how  survey  and  study  ob- 
jectives affect  sampling.  Objectives  developed  for 
studies  and  surveys  will  also  affect  species  and  habi- 
tats to  be  sampled. 

In  summary,  deserts  provide  tremendous  oppor- 
tunities to  study  relationships  between  animals  and 
their  environment.  But  careful  consideration  must  be 
given  if  inventories  and  monitoring  are  to  answer 
specific  management  questions. 


Deserts 


145 


LITERATURE  CITED 

ATTENBOROUGH,  D.,  ed.  1976.  Deserts  and  Grasslands. 
Doubleday  and  Company,  Inc.  New  York,  NY. 

BAILEY,  R.G.  1978.  Description  of  ecoregions  of  the 
United  States.  U.S.  Dep.  Agric,  For.  Serv.  Ogden,  UT. 
77pp. 

BARBOUR,  M.G.,  J.A.  MAC  MAHON,  S.A.  BAMBERG,  and 
J.A.  LUDWIG.  1977.  The  structure  and  distribution  of 
Larrea  communities.  Pages  227-251  in  Mabry,  T.J.,  J.H. 
Hunziker,  and  DR.  DiFreo, Jr.,  eds.  Creosote  Bush. 
U.S./IBP  Synthesis  Series  6.  Dowden,  Hutchison  and 
Ross,  Inc.  Stroudsburg. 

BARBOUR,  R.W.  and  W.H.  DAVIS.  1969.  Bats  of  America. 
Univ.  Kentucky  Press.  Lexington.  286pp. 

BENDER,  G.L.,  ed.  1982.  Reference  handbook  on  the  de- 
serts of  North  America.  Greenwood  Press.  Westport, 
CT.  594pp. 

BENT,  AC.  1961.  Life  histories  of  North  American  birds  of 
prey.  Vol.  1  398pp.  Vol.  2  466pp.  Dover  Publ.,  Inc. 
New  York,  NY. 

BLACK,  H.L.  1968.  Populations  of  small  rodents  in  relation 
to  grazing  by  cattle  on  foothill  rangeland.  Masters 
Thesis,  Univ.  of  Utah.  Provo. 

BROWN,  D.E.  1982.  Introduction.  Pages  8-16  in  Brown, 
D.E.,  ed.  Desert  Plants:  Biotic  Communities  of  the 
American  Southwest — United  States  and  Mexico.  Univ. 
of  Arizona  Press.  Tucson. 

.  1982a.  Chihuahuan  desert  scrub.  Pages  169-179  in 

Brown,  D.E.,  ed.  Desert  Plants:  Biotic  Communities  of 
the  American  Southwest — United  States  and  Mexico. 
Univ.  of  Arizona  Press.  Tucson. 

.  1982b.  Desert  plants:  biotic  communities  of  the 

American  Southwest — United  States  and  Mexico.  Univ. 
of  Arizona  Press.  Tucson.  342pp. 

,  C.H.  LOWE,  and  C.P.  PASE.  1979.  A  digitized  clas- 
sification system  for  the  biotic  communities  of  North 
America,  with  community  (series)  and  association  ex- 
amples for  the  Southwest.  J.  Arizona  Acad.  Sci.  14 
(Suppl):l-16. 

BROWN,  J.A.  1973-  Species  diversity  of  seed-eating  desert 
rodents  in  sand  dune  habitats.  Ecology  54:775-787. 

BROWN,  J.H.  and  AC.  GIBSON.  1983-  Biogeography.  C.V. 
Mosby  Co.  St.  Louis,  MO  492523pp. 

BURGE,  B.L  1979.  Survey  of  the  present  distribution  of 
the  desert  tortoise,  Gophenis  agassizii,  in  Arizona. 
U.S.  Dep.  Inter.,  Bur.  Land  Manage.  Contract  YA-5 12- 
CT8-108.  Denver,  CO. 

BURT,  W.H.  and  R.P.  GROSSENHEIDER.  1976.  A  field 

guide  to  the  mammals.  Houghton-Mifflin.  Boston,  MA. 
289pp. 

BURY,  R.B.  and  R.A.  LUCKENBACH.  1983.  Vehicular  recre- 
ation in  arid  land  drives,  biotic  responses  and  manage- 
ment alternatives.  Pages  217-221  in  Webb,  R.H.  and 
H.G.  Wilshire,  eds.  Environmental  Effects  of  Off-Road 
Vehicles.  Impacts  and  Management  in  Arid  Regions. 
Springer-Verlag.  New  York,  NY. 

CALL,  M.W.  1978.  Nesting  habitats  and  surveying  tech- 
niques for  common  western  raptors.  Tech.  Note  316. 
U.S.  Dep.  Inter.,  Bur.  Land  Manage.  BLM  Service  Cen- 
ter. Denver,  CO. 

CONANT,  R.  1978.  Semiaquatic  reptiles  and  amphibians  of 
the  Chihuahuan  Desert  and  their  relationships  to 
drainage  patterns  of  the  region.  Pages  445-491  in 
Waver,  R.H.  and  D.W.  Riskind,  eds.  Trans.  Symp.  Biol. 
Resour.  Chihuahuan  Desert  Region,  U.S.  and  Mexico. 
U.S.  Dep.  Inter.,  Nat.  Park  Serv.  Proc.  Trans.  Ser.  3- 


CREUSERE,  F.M.  and  W.G.  WHITFORD.  1982.  Use  of  time 
and  space  by  lizards,  in  Scott,  N.,  ed.  Herpetological 
communities,  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv. 
Wildl.  Res.  Rep.  1 3. 

DAUBENMIRE,  R.F.  1974.  Plants  and  environment:  a  text- 
book of  autecology.  Wiley  and  Sons,  Inc.  New  York,  NY. 

DAVIS,  G.P.  Jr.  1973-  Man  and  wildlife  in  Arizona:  the  pre- 
settlement  era,  1823-1864.  MS  Thesis.  Univ.  of  Ari- 
zona. Tucson. 

DAVIS,  R.M.,  ed.  1976.  National  range  handbook.  U.S.  Dep. 
Agric,  Soil  Cons.  Serv.  Washington,  DC. 

DUNAWAY,  D.  1971.  Human  disturbances  as  a  limiting 
factor  of  Sierra  Nevada  bighorn  sheep.  N.  A.  Wild. 
Sheep  Conf.  Trans.  1:165-173. 

ENG,  R.L.  and  P.  SCHLADWEILER.  1972.  Sage  grouse  win- 
ter movements  and  habitat  use  in  central  Montana.  J. 
Wildl.  Manage.  36(  1  ):14 1-146. 

FLEMING,  W.B.  1959.  Migratory  waterfowl  in  Arizona. 
Ariz.  Game  and  Fish  Dep.  Wildl.  Bull.  5.  Phoenix. 

GOLLEY,  F.B.,  K  PETRUSEWICZ,  and  L.  RYSZKOWSKI. 
1975.  Small  mammals:  their  productivity  and  popula- 
tion dynamics.  Cambridge  Univ.  Press.  Cambridge. 

GOODWIN,  J.G.  Jr.  and  C.R.  HUNGERFORD.  1977.  Habitat 
use  by  native  Gambels  and  scaled  quail  and  released 
masked  bobwhite  quail  in  southern  Arizona.  U.S.  Dep. 
Agric.  For.  Serv.  Res.  Paper  RM-197. 

HEATWOLE,  H.  1982.  A  review  of  structuring  in  herpeto- 
logical assemblages.  Pages  1-19  in  Scott,  N.,  ed.  Her- 
petological Communities.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  Wildl.  Res.  Rep.  1 3. 

HIRTH,  HE,  R.C.  PENDLETON,  AC.  KING,  and  T.R. 

DOWNARD.  1969.  Dispersal  of  snakes  from  a  hiberna- 
culum  in  northwestern  Utah.  Ecology  50(2):332-339. 

JOHNSGARD,  PA.  1973.  Grouse  and  quails  of  North 
America.  Univ.  of  Nebraska  Press.  Lincoln.  553pp. 

JONES,  KB.  1981a.  Distribution,  ecology,  and  habitat 
management  of  the  reptiles  and  amphibians  of  the 
Hualapai  Aquarius  planning  area,  Mojave  and  Yavapai 
Counties,  Arizona.  U.S.  Dep.  Inter.,  Bur.  Land  Manage. 
Tech.  Note  353.  Denver,  CO. 

.  1 98 1  b.  Effects  of  grazing  on  lizard  abundance  and 

diversity  in  western  Arizona.  Southwest  Nat.  26(  2 ): 
107-115. 

and  PC.  GLINSKI.  1985.  Microhabitats  of  lizards  in 

a  Southwestern  riparian  community  in  North  Ameri- 
can Riparian  Communities:  Resolving  Conflicting  Uses. 
U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Note.  In  press. 

,  L.M.  KEPNER,  and  W.G.  KEPNER.  1983.  Anurans  of 

Vekol  Valley,  Arizona.  Southwest  Nat.  28(4):469-470. 

,  LP.  KEPNER,  and  T.E.  MARTIN.  1985.  Species  of 

reptiles  occupying  habitat  islands  in  western  Arizona: 
a  deterministic  assemblage.  Oecologia  In  press. 

KINDSCHY,  R.R.,  C.  SUNDSTROM,  and  J.D.  YOAKUM. 
1982.  Wildlife  habitats  in  managed  rangelands — the 
Great  Basin  of  southeastern  Oregon.  Pronghorns.  U.S. 
Dep.  Agric,  For.  Serv.  Gen.  Tech.  Report  PNW-145. 

KNOPF,  A.A.  1977.  The  Audubon  Society  field  guide  to 
North  American  birds — western  region.  Alfred  A. 
Knopf,  Inc.  New  York,  NY. 

KUCHLER,  AW.  1964.  The  potential  natural  vegetation  of 
the  conterminous  United  States.  American  Geographic 
Society.  Special  Publ.  361.  116  pp.  map. 

LANDCASTER,  R.K  and  WE.  REES.  1979.  Bird  communi- 
ties and  the  structure  of  urban  habitat.  Can.  J.  Zool. 
57:2358-2368. 

LOWE,  C.H.,  ed.  1964.  The  vertebrates  of  Arizona.  Univ.  of 
Arizona  Press.  Tucson.  342pp. 


146 


Deserts 


MARES,  MA.  and  AC.  HULSE.  1977.  Patterns  of  some  ver- 
tebrate communities  in  creosote  bush  deserts.  Pages 
209-226  in  Mabry,  T.J.,  J.H.  Hunziker,  and  DR.  DiFreo 
Jr.,  eds.  Creosote  Bush.  U.S./IBP  Synthesis  Series  6, 
Dowden,  Hutchison  and  Ross,  Inc.  Stroudsburg. 
MARTIN,  A.C.,  H.S.  ZIM,  and  A.L.  NELSON.  1951.  American 
wildlife  and  plants:  a  guide  to  wildlife  food  habits. 
Dover  Publ.,  Inc.  New  York,  NY.  500pp. 
MARTIN,  T.E.  1980.  Diversity  and  abundance  of  spring  mi- 
gratory birds  using  habitat  islands  on  the  Great  Plains. 
Condor  82:430-439. 

.  1981.  Limitation  in  small  habitat  islands:  chance  or 

competition?  Auk  98:715-734. 
MASER,  C,  J.M.  GEIST,  DM.  CONCANNON,  R.  ANDER- 
SON, and  B.  LOVELL.  1979a.  Wildlife  habitats  in  man- 
aged rangelands — the  Great  Basin  of  southeastern  Ore- 
gon— geomorphic  and  edaphic  habitats.  U.S.  Dep. 
Agric,  For.  Serv.  Gen.  Tech.  Report  PNW-99. 

, J.W.  THOMAS,  ID.  LUMAN,  and  R.  ANDERSON. 

1979b.  Wildlife  habitats  in  managed  rangelands — the 
Great  Basin  of  southeastern  Oregon — man-made  habi- 
tats. U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Report 
PNW-86. 
MC  CLANAHAN,  L.  196^.  Adaptations  of  the  spadefoot 
toad,  Scaphiopus  conchi,  to  desert  environments. 
Comp.  Bio.  Chem.  Physiol.  20:73-99. 
MILLSAP,  B.A.  1981.  Distributional  status  of  Falconiformes 
in  west-  central  Arizona:  with  notes  on  ecology,  repro- 
ductive success,  and  management.  U.S.  Dep.  Inter., 
Bur.  Land  Manage.  Tech.  Note  355.  Denver,  CO. 
MINCKLEY,  W.L.  1973.  Fishes  of  Arizona.  Ariz.  Game  and 

Fish  Dept.  Phoenix.  293pp 
NAIMAN,  RJ.  and  D.L.  SOLTZ,  eds.  1981.  Fishes  in  North 
American  deserts.  Wiley  and  Sons,  Inc.  New  York,  NY. 
552pp. 
NORRIS,  K.S.  1953-  The  ecology  of  the  desert  iguana,  Dip- 
sosaurus  dorsalis.  Ecology  34:265-287. 

.  1958.  The  evolution  and  systematica  of  the  iguanid 

genus  Uma  and  its  relation  to  the  evolution  of  other 
North  American  desert  reptiles.  Bull.  Am.  Mus.  Nat. 
Hist.  H4(3):247-326. 
OLENDORFF,  R.R.,  AD.  MILLER,  and  R.N.  LEHIV      ..  1981. 
Suggested  practices  for  raptor  protection  on  power 
lines:  the  state  of  the  art  in  1981.  Raptor  Res.  Rep.  4, 
Raptor  Research  Foundation.  Dep.  Vet.  Biol.,  Univ.  of 
Minnesota,  St.  Paul. 
OOSTING,  H.J.  1956.  The  study  of  plant  communities. 

W.  H.  Freeman  and  Company.  San  Francisco,  CA.4-i0pp. 
PIANKA,  E.R.  1966.  Convexity,  desert  lizards,  and  spatial 
heterogeneity.  Ecology  -t7(  6):  1055- 1059. 

.  1970.  Comparative  autecology  of  the  lizard  Cnem- 

idophorus  tigris  in  different  parts  of  its  geographic 
range.  Ecology  51:703-720. 
PORTER,  KR.  1972.  Herpetology.  W.  B.  Saunders  Co.  Phil- 
adelphia. PA.  281pp. 
ROTTENBERRY,  J.T.  and  J.A.  WIENS.  1980.  Habitat  struc- 
ture, patchiness,  and  avian  communities  in  North 
American  steppe  vegetation:  a  multivariate  analysis. 
Ecology  61:1228-1250. 
SCHMIDT,  J.L.  and  D.L.  GILBERT,  eds.  1978.  Big  game  of 
North  America:  ecology  and  management.  Stackpole 
Books.  Harrisburg,  PA.  512pp. 
SERVENTY,  D.L.  1971.  Biology  of  desert  birds.  Pages  287- 
339  in  Farmer,  D.S.,  and  JR.  King,  eds.  Avian  Biology: 
Vol.  I  Academic  Press.  New  York,  NY. 
SHORT,  ILL  1983.  Wildlife  guilds  in  Arizona  desert  habi- 
tats. BLM  Tech  Note  302.  Denver,  CO.  258pp. 


SHREVE,  F.  1942.  The  desert  vegetation  of  North  America. 
Bot.  Rev.  8:195-246. 

SIMON,  C.A.  and  G.A.  MIDDENDORF.  1976.  Resource  par- 
titioning by  an  iguanid  lizard:  temporal  and  microhabi- 
tat  aspects.  Ecology  57<6):1 317-1 320. 

SKOVLIN,  J.M.  1982.  Habitat  requirements  and  evalua- 
tions. Pages  369-413  in  Thomas,  J.W.  and  D.E.  Toweill, 
eds.  Elk  of  North  America:  Ecology  and  Management. 
Stackpole  Books,  Harrisburg,  PA. 

STEBBINS,  R.C.  1966.  A  field  guide  to  western  reptiles 
and  amphibians.  Houghton-Mifflin  Co.  Boston,  MA. 
279pp. 

THOMAS,  J.W.,  C.  MASER,  andJ.E.  RODIEK.  1979.  Wildlife 
habitats  in  managed  rangelands — the  Great  Basin  of 
southeastern  Oregon — edges.  U.S.  Dep.  Agric,  For. 
Serv.  Gen.  Tech.  Rep.  PNW-85.  17pp. 

THORNE,  R.F.,  B.A.  PRIGGE,  and  J.  HENDR1CKSON.  1981.  A 
flora  of  the  higher  ranges  and  the  Kelso  Dunes  of  the 
eastern  Mohave  Desert  in  California.  Aliso  10:F- 1-186. 

TOMOFF,  C.S.  1974.  Avian  species  diversity  in  desert 
scrub.  Ecology  55(2):396-403. 

TURNER,  R.M.  1982a.  Great  Basin  desert  scrub.  Pages  145- 
155  in  Brown,  D.E.,  ed.  Desert  Plants:  Biotic  Commu- 
nities of  the  American  Southwest — United  States  and 
Mexico.  Univ.  of  Arizona  Press.  Tucson. 

.  1982b.  Mohave  desert  scrub.  Pages  157-168  in 

Brown,  D.E.,  ed.  Desert  Plants:  Biotic  Communities  of 
the  American  Southwest — United  States  and  Mexico. 
Univ.  of  Arizona  Press.  Tucson. 

and  D.E.  BROWN.  1982.  Sonoran  desert  scrub,  in 

Brown,  D.E.,  ed.  Desert  Plants:  Biotic  Communities  of 
the  American  Southwest — United  States  and  Mexico. 
Univ.  of  Arizona  Press.  Tucson.  342pp. 

VAN  DEVENDER,  T.R  and  W.G.  SPAULDING.  1979.  De- 
velopment of  vegetation  and  climate  in  the  southwest- 
ern United  States.  Science  204:701-710. 

VITT,  L.J.,  R.C.  VAN  LOBEN  SELS,  and  R.D.  OHMART. 

1981.  Ecological  relationships  among  arboreal  desert 
lizards.  Ecology  62(2):  398-4 10. 

VOGL,  RJ.  1977.  Fire:  A  destructive  menace  or  a  natural 
process?  Pages  261-289  in  Cairns,  J.  Jr.,  KL.  Dickson, 
and  E.E.  Herricks,  eds.  Recovery  and  Restoration  of 
Damaged  Ecosystems.  Univ.  Press  of  Virginia.  Char- 
lottesville. 

WEBB,  R.H  and  H.G.  WILSHIRE,  eds.  1983.  Environmental 
effects  of  off-road  vehicles:  impacts  and  management 
in  arid  regions.  Springer-Verlag.  New  York,  NY.  534pp. 

WILSON,  L.O.,  J.  BLAISDELL,  G.  WELSH,  R.  WEAVER,  R. 
BR1GHAM,  W.  KELLY,  J.  YOAKUM,  M.  HINKES,  J. 
TURNER,  and  J.  DE  FORGE.  1980.  Desert  bighorn  hab- 
itat requirements  and  management  recommendations. 
Pages  1-7  in  Des.  Bighorn  Coun.  Trans.  24. 

WRIGHT,  HA.  and  AW.  BAILEY.   1982.  Fire  ecology.  Wiley  - 
Interscience,  New  York,  NY.  501pp. 

YOAKUM,  J.  1980.  Habitat  management  guides  for  the 
American  pronghorn  antelope.  U.S.  Dep.  Inter.,  Bur. 
Land  Manage.  Tech.  Note  347.  Denver,  CO. 

YORKJ.C.  and  W.A.  DICK-PEDDIE.  1969.  Vegetation 
changes  in  southern  New  Mexico  during  the  past 
hundred  years.  Pages  157-166  in  McGinnies,  W.G.  and 
B.J.  Goldman,  eds.  Arid  Lands  in  Perspective.  Univ.  of 
Arizona  Press.  Tucson. 
ZWEIFEL,  R.G.  and  C.H.  LOWE.  1966.  The  ecology  of  a 
population  of  Xantusia  vigilis,  the  desert  night  lizard. 
Am.  Mus.  Novit  2247;1  S~ 


Deserts 


147 


8 


TUNDRA 


Peter  C.  Lent 

Bureau  of  land  Management 
920  Valley  Road 
Reno,  NV  89512 


"When  we  reach  the  Arctic  regions,  or  snow-capped 
summits,  the  struggle  for  life  is  almost  exclusively 
with  the  elements." 

— Charles  Darwin,  from  On  the  Origin  of 
Species 


Editor's  Note:  Animals  of  tundra  habitat  must 
struggle  for  life.  Because  of  the  harsh  working  con- 
ditions and  inaccessibility  of  many  tundra  areas, 
few  biologists  have  conducted  systematic  habitat 
inventories  or  surveys  of  general  vertebrate  habitat 
relationships  in  these  severe  environments.  Most 
of  the  work  in  such  areas  has  concentrated  on  a 
few  large,  economically  important  wildlife  species. 
Nevertheless,  the  conditions  are  so  different  from 
those  in  more  temperate  areas  that  the  subject  de- 
serves a  separate  chapter.  This  chapter  describes  the 
state-of-the-art  for  both  arctic  and  alpine  tundra; 
the  biologist  working  in  such  areas  will  find  much 
room  for  innovation  and  development  of  new 
systems. 


INTRODUCTION 

The  origin  of  the  English  word  "tundra"  is  said 
to  come  from  the  Finnish  "tunturi"  then  from  the 
Russian  "tundra,"  meaning  a  marshy,  treeless  plain. 
In  a  general  sense  the  term  refers  to  any  cold  cli- 
mate landscape  having  vegetation  without  trees. 
Such  landscapes  occur  predominantly  in  the  north- 
ern hemisphere  north  of  the  tree  line  (arctic  tundra) 
and  in  mountainous  areas  of  both  hemispheres  above 
the  tree  line  (alpine  tundra).  The  vegetation  commu- 
nities are  characterized  by  short-stemmed  perennial 
herbaceous  plants,  stunted  or  prostrate  shrubs,  li- 
chens, and  mosses.  Similar  vegetation,  where  trees 
are  absent  due  to  low-growing  season  temperatures, 
also  occurs  in  mid-  to  high-latitude  maritime  envi- 
ronments such  as  the  Aleutian  Islands  and  on  certain 
islands  in  the  southern  hemisphere,  such  as  South 
Georgia  and  the  Macquaries  (Jenkins  and  Ashton 
1970). 


Tundra  areas  constitute  slightly  over  8%  of  the 
total  U.S.  land  area  and  about  5%  of  the  earth's  land 
surface.  Of  the  U.S.  total,  about  74  million  ha  (183 
million  a.)  are  in  Alaska  and  3  million  ha  (7.5  million 
a. )  of  alpine  tundra  are  in  the  contiguous  48  states 
(Figure  la).  Of  the  Alaskan  tundra  area,  Brown  et  al. 
( 1978)  estimated  that  52%  is  arctic  and  48%  is  al- 
pine tundra  (Figure  lb). 

The  manager  or  biologist  new  to  arctic  ecosys- 
tems will  find  a  detailed  and  comprehensive  body  of 
literature  available.  The  purpose  of  this  chapter  is 
to  help  the  manager  get  started  but  is  by  no  means 
exhaustive.  In  contrast,  the  managers  concerned 
with  alpine  areas  will  find  a  more  scattered  body  of 
literature  and  will  need  to  make  a  greater  effort  to 
seek  out  material  specific  to  their  area.  Two  general 
treatments  of  North  American  alpine  areas  are  worth 
noting:  Zwinger  and  Willard  (1972)  and  Price  (1981). 
A  glossary  of  terms  broadly  relevant  to  arctic  and 
alpine  environments  has  been  prepared  by  Gabriel 
and  Talbot  ( 1984 ).  The  biologist  or  manager  working 
with  alpine  areas  will  find  some  useful  material  in 


Tundra 


149 


1 .  Sierra  Nevada 

2.  Southern  Cascades,  Siskiyou  and  Klamath  Mountains 

3.  Olympic  Mountains 

4.  Northern  Cascade  Mountains 

5.  Northern  Rocky  Mountains 

6.  Great  Basin  Ranges 

a)  Blue  and  Wallowa  Mountains:  Oregon 

b)  Ruby,  Wheeler,  Charleston  Mountains:  Nevada 

c)  Uinta  and  Wasatch  Mountains:  Utah 

d)  San  Francisco  Peak  and  White  Mountains:  Arizona 

e)  Mounts  San  Jacinto  and  San  Gorgonio:  California 

f)  White,  Inyo  and  Panamint  Mountains:  California 

7.  Southern  Rocky  Mountains 


Figure  la.     Alpine  tundra  areas  in  the  contiguous  48  states  (adapted  from  Zwinger  and  Willard  1972). 


150 


Tundra 


Moist  tundra 
Wet  tundra 

Alpine  tundra, 
ice  and  snow 


Figure  lb.     Arctic  and  alpine  tundra  areas  in  Alaska. 


the  arctic  literature,  but  must  use  caution  in  its  ap- 
plication; there  are  some  fundamental  differences 
between  arctic  and  alpine  ecologies. 

All  tundra  areas  are  characterized  by  short 
growing  seasons,  usually  fewer  than  100  days,  and 
low  summer  temperatures.  In  extreme  situations  the 
daily  temperature  may  average  only  2°  to  5°  C  (36° 
to  41°  F).  Although  all  share  this  characteristic  of 
"cold-dominated,"  the  vegetation  types  encompassed 
by  tundra  are  exceptionally  diverse.  At  one  extreme 
is  the  high  arctic  polar  desert,  where  vegetative 
cover  may  be  less  than  10%,  composed  primarily  of 
cushion  plant  forms  and  mosses  with  a  few  other 
annuals  scattered  sporadically.  Polar  desert  and  semi- 
desert  are  rare  within  the  U.S.  They  are  best  repre- 
sented in  the  arctic-alpine  environment  of  the 
Brooks  Range  in  northern  Alaska.  At  the  other  end  of 
the  spectrum,  over  most  of  North  America,  the  tun- 
dra grades  into  forest  in  a  forest-tundra  ecotone. 
Here  vegetative  biomass  can  be  relatively  high  with 
several  well-developed  canopy  layers. 


Discussion  in  this  chapter  will  encompass  these 
transition  zones  because  of  the  unique  qualities  they 
have  in  common  with  tundra,  such  as  prevalence 
and  importance  of  lichens  and  potential  presence  of 
permafrost,  and  because  important  migratory  species, 
such  as  elk  (Cervus  elaphus)  and  caribou  (Rangifer 
tarandus),  depend  on  a  single  system  composed 
of  both  tundra  and  forest-tundra  zones. 

Some  writers  have  contended  that  it  is  better  to 
reserve  the  term  "tundra"  to  refer  to  areas  with 
permafrost,  either  continuous  or  discontinuous.  Bill- 
ings ( 1973),  for  example,  preferred  the  use  of  the 
terms  "meadow"  and  "fell-field"  for  the  mesic  and 
xeric  alpine  equivalents  of  tundra,  respectively.  As 
he  noted,  important  differences  exist,  such  as  the 
more  intense  solar  radiation  in  alpine  environments 
coupled  with  extreme  diurnal  temperature  fluctua- 
tions, as  opposed  to  the  continuous  summer  daylight 
in  the  high  latitude  tundra  environment.  But  Billings' 
own  work  revealed  that  50%  of  the  vascular  plant 
species  in  the  alpine  zone  of  the  Beartooth  Plateau 


Tundra 


151 


of  Montana  were  species  also  present  in  arctic  tun- 
dra. A  similar  large  overlap  of  species  does  not  occur 
between  arctic  and  alpine  faunas,  however 
(Hoffmann  1974,  1984). 

Aside  from  the  similarities  in  low  growth  forms, 
which  suggest  that  similar  vegetation  sampling  tech- 
niques are  applicable  in  alpine  and  arctic  situations, 
there  are  some  basic  similarities  arising  from  the 
cold-dominated  environment  of  both,  such  as  low 
primary  productivity,  low  decomposition  rates,  low 
stress  tolerance,  and  high  susceptibility  to  physical 
disturbance — even  where  permafrost  is  not  present. 
The  complex  spatial  heterogeneity  of  arctic  and  al- 
pine vegetation,  a  "patchiness"  that  is  based  on  re- 
sponses to  microrelief  features,  also  dictates  that 
similar  problems  are  to  be  encountered  in  inventory 
and  monitoring  of  arctic  and  alpine  habitats. 

The  presence  of  permafrost  is  certainly  charac- 
teristic of  most  arctic  tundra  areas.  Discontinuous 
permafrost  is  also  found  in  about  100,000  knT 
(40,000  mi.  )  of  alpine  areas  in  the  western  U.S. 
exclusive  of  Alaska  (Pewe  1983).  However,  perma- 
frost is  not  universally  present  in  tundra  nor  does 
the  presence  of  permafrost,  if  it  is  at  sufficient  depth, 
necessarily  preclude  tree  growth.  Furthermore,  even 
in  the  absence  of  permafrost,  per  se,  cryopedogenic 


forces  (frost  action)  play  a  significant  ecological  role 
in  arctic  and  alpine  habitats.  Thus  the  position  taken 
by  Bliss  ( 1981 )  and  the  one  taken  in  this  chapter  is 
that  the  presence,  absence,  and  the  characteristics  of 
the  permafrost  layer  are  among  the  major  influences 
on  the  nature  of  a  given  tundra  area  and  are  major 
considerations  in  inventory  and  management.  They 
are  not,  however,  decisive  in  determining  whether 
an  area  is  considered  to  be  tundra. 


A  current  tendency  exists  to  shift  tundra  to 
lower  levels  in  hierarchical  vegetation  classification 
systems.  For  example,  Viereck  and  Dyrness  (1980) 
distinguished  between  tundra  and  grass/herbaceous 
formations  in  the  first  dichotomy  (Level  I)  of  their 
hierarchical  system.  However,  in  a  later  version  of 
their  classification  system  (Viereck  et  al.  1982),  all 
tundra  types  were  reduced  to  subcategories  under 
two  of  their  Level  I  classes  (herbaceous  and  shrub). 
They  classified  communities  of  the  Aleutian  Islands 
and  other  maritime  environments  that  are  dominated 
by  Calamagrostis  canadensis  and  Elymus  arenarius 
as  grasslands  and  not  tundra.  They  also  referred  to 
shrubland's  several  types  that  have  been  included 
with  tundra  or  shrub-tundra  by  other  workers.  Dris- 
coll  et  al.  (1983)  also  included  tundra  as  a  subdivi- 
sion in  their  grassland/herbaceous  class. 


MMfaT  'l^ftg 


V#  ';r;v^;-- 


*..*v>* 


»43ur*  l    ,  Eft  * s  $fa  i*<m  *T m  .    .  - .  ^  ■ 


/*»■ 


^^  -'<•*< 


•  •* 


<  ■ 


rW 


Caribou  on  tundra  riparian  area  in  July. 


152 


Tundra 


GENERAL  HABITAT  CLASSIFICATION 
SYSTEMS 

Various  wildlife  habitat  classification  systems 
have  been  employed  in  both  arctic  and  alpine  areas. 
There  are  no  generally  or  widely  used  systems  and 
none  of  them  has  been  systematically  tested  as  a 
general  system.  There  is,  however,  some  general  cor- 
respondence among  some  of  these,  particularly  for 
terrestrial  birds  and  small  mammals  (Table  1).  Most 
of  these  habitat  classes  represent  various  subdivi- 
sions of  the  soil  moisture  gradient  from  flooded  and 
saturated  through  desiccated  and  rocky  (Pruitt  1966; 
Hoffmann  1974).  This  reflects  the  dominant  role  of 
soil  moisture  as  a  niche  variable  both  directly  and 
indirectly  through  its  interrelations  with  the  perma- 
frost regime  and  vegetation.  Winter  habitats  of  small 
mammals  also  reflect  similar  snow-cover  gradients 
(Pruitt  1966).  These  general  habitat  classes  and 
other  habitat  features  and  their  relationships  to  one 
another  are  illustrated  in  Figures  2  and  3- 


Moitoret  et  al.  (1985)  performed  the  most  de- 
tailed analysis  of  arctic  tundra  habitats  to  date.  Using 
the  clustering  and  ordination  analysis  program 
TWINSPAN,  they  found  flooded  and  riparian  to  be 
the  most  distinct  of  their  seven  coastal  tundra  terres- 
trial bird  habitats.  Drier  types,  other  than  riparian, 
had  lower  and  less  diverse  bird  use.  Their  analysis 
suggested  that  habitat  factors  relating  to  microrelief 
and  interspersion  characteristics  may  be  important  in 
defining  habitats.  Thus,  habitat  classes  so  defined  do 
not  necessarily  show  one-to-one  correspondence 
with  the  broad  vegetation-based  classes  ascertained 
from  satellite  imagery  such  as  that  of  Landsat. 

Several  more  specialized  tundra  habitat  classifi- 
cations have  been  employed.  One  relating  specifi- 
cally to  wetlands  is  described  in  the  discussion 
regarding  migratory  waterfowl.  A  system  used  for 
impact  analysis  in  arctic  and  subarctic  Alaska  is  de- 
scribed in  the  section  treating  inventory  and  moni- 
toring for  impact  assessment. 


Table  1.     Habitat  classes  used  in  some  studies  of  arctic  and  alpine  terrestrial  birds  and  mammals. 


Locale 

Habitat  Classes 

Reference 

Birds  and  small 

Moist 

Sedge  tussock 

Willow  sedge 

Dry 

Krumm- 

Fellfield 

Cliff,  talus 

Hoffmann 

mammals, 

meadow 

meadow1 

holz2 

(1974) 

Beartooth 

Plateau,  MT 

Breeding  habi- 

Moist, wet 

Heath  tussock 

Riparian 

Dry 

Combined 

Cliffs,  talus 

Williamson 

tats  of  birds, 

meadow 

meadow 

with  dry 

(1961) 

Cape  Thomp- 

Fellfield 

meadow 

son,  Alaska 

Bird  habitats, 

Sedgegrass 

Tussock  heath 

Tall  shrub 

Dwarf 

Dry  tundra 

Kessel  and 

Upper  Sheen- 

marsh 

shrub2 

Schaller 

jek  R.  (Subarc- 

(1960) 

tic-alpine) 

Alaska 

Bird  habitats, 

5  hydric 

Tussock 

Riparian 

Moitoret  et 

coastal  Arctic 

types3 

al.  1985 

National  Wild- 

life Refuge, 

Alaska 

Selected 

Long-tailed, 

Lapland,  long- 

Arctic  warblers, 

Golden 

Willow 

Horned 

Marmots, 

Characteristic 

parasitic 

spur,  Calcarius 

Dendroica  sp. 

plover,  Plu- 

ptarmigan 

lark,  Ere- 

Marmota 

Species 

jaegers, 

lapponicus 

vialis  dom- 

mophila 

sp. 

Sterconarius 

Red-backed 

Water  shrew, 

inica 

Chip- 

alpestris 

sp. 

vole,  Clethrion- 

Sorex  palustris 

munks, 

Mountain 

omys  rutilus 

Pocket 

Eutamias 

Dall's 

goat 

Tundra  vole, 

Moose,  Alces 

gophers, 

sp. 

sheep 

Microtus 

alces 

Thomomys 

Golden 

Oeconomus 

sp. 

Dall's 
sheep 

eagle, 
Aquila 
chrysaetos 

Hoffmann  (1974)  also  distinguishes  rock  polygons,  these  being  dry  meadow  with  rock  stripes  or  patterns.  Characteristic  species  are 

pikas,  Ochotona  sp. 

2Krummholz  is  a  coniferous  ecotone  below  the  alpine  tundra.  These  two  are  only  crudely  structural  counterparts. 
3Five  hydric  types  identified:  Flooded.  Wet  sedge,  Moist  sedge,  Mosaic,  and  Moist-sedge  shrub. 


Tundra 


153 


MAJOR  WILDLIFE  GROUPS 

It  is  useful,  particularly  for  wildlife  species  in 
arctic  tundra  habitats,  to  consider  three  categories  of 
species:  ( 1 )  resident  species  that  remain  active  year- 
round,  (2)  resident  species  hibernating  in  winter, 
and  (3)  migratory  species  present  for  only  a  portion 
of  the  year.  Tundra  areas  tend  to  have  a  larger  pro- 
portion of  migratory  and  otherwise  highly  mobile 
species  than  most  other  broad  habitat  classes. 

Resident  arctic  mammal  species  include  some 
medium-sized  hibernators,  such  as  the  arctic  ground 
squirrel  {Spermophilns parry  ii)  and  hoary  marmot 
(Marmota  caligata),  or  large  ones,  such  as  grizzly 
bears  (Ursus  arctos),  but  include  a  larger  variety  of 
species  that  remain  active  year-round. 

The  smaller  species,  rodents  and  shrews  primar- 
ily, tend  to  be  "chionophiles"  (Pruitt  1978),  that  is, 
they  require  snow  cover  to  ameliorate  an  otherwise 
unsurvivable  winter  environment.  Others  are  "chion- 
ophobes,"  such  as  the  musk-ox  (Ovibos  moschatus), 


requiring  habitats  of  restricted  snow  cover  for  win- 
ter survival.  The  importance  of  snow  in  the  tundra 
environment  is  treated  in  a  separate  section  of  this 
chapter. 

The  number  of  resident  arctic  bird  species  is 
extremely  small,  but  may  include  ptarmigan  (Lago- 
pus  lagopus  and  L  mutus),  ravens  (Corvus  corax), 
snowy  owls  (Nyctea  scandiaca),  and  gyrfalcons 
(Falco  rusticolus).  The  great  majority  of  the  97  or 
so  bird  species  using  the  northern  Alaska  tundra  are 
migratory  (Pitelka  1979).  Their  dependency  on  tun- 
dra habitats,  however  critical,  is  a  brief  one  and 
timed  precisely  according  to  the  adaptations  of  each 
species.  Thus  there  is  a  complex  and  compressed 
pattern  of  arrivals  and  departures  that  makes  the 
correct  timing  of  inventories  critical  to  their  success. 
In  northern  Alaska  the  peak  of  arrival  dates  is  in  late 
May;  some  early  breeding  species  may  leave  as  early 
as  late  July. 

Waterfowl  and  shorebirds  are  almost  entirely 
migratory  except  for  a  few  species  that  may  associ- 
ate with  open  water  marine  environments  in  winter 


GLACIER 


BEACH 


DE-ROCK, 
TALUS 


FORB   MEADOW 


SANDY  BLUFF 


POLYGON 


THAW  LAKE 
r  MEADOW 


WOODED 
TUNDRA      4 


>   A>*    4 


Figure  2.     Schematic  representation  of  arctic  and  arctic-alpine  tundra  habitat  types  and  features  (after  Hoff- 
mann 1984). 


154 


Tundra 


Musk-ox  winter  tundra  habitat  with  low  snow-cover. 


in  some  arctic  areas.  Most  waterbird  species  are 
clearly  tied  to  wet  tundra  habitats.  However,  there 
are  some  shorebird  species  clearly  associated  with 
dry,  upland,  or  even  arctic-alpine  habitats  (Kessel 


and  Schaller  I960;  King  1979).  The  arctic-alpine  tun- 
dra supports  a  surprisingly  diverse  array  of  summer 
breeding  birds,  estimated  by  Pitelka  (1979)  to  be  50 
species,  of  which  nearly  half  are  passerines. 

Because  all  migratory  species  cannot  be  sampled 
at  an  optimal  time  in  a  single  inventory,  it  is  neces- 
sary to  have  a  clear  statement  of  objectives  and  clear 
familiarity  with  the  regional  literature  to  strive  for 
the  optimal  timing  for  specific  needs  or  specific  spe- 
cies. Moulting  concentrations  are  common  after 
breeding  in  several  waterfowl  species,  and  moulting 
areas  may  be  used  only  briefly,  yet  are  of  critical 
importance  to  entire  populations  (Pitelka  1979;  NPR- 
A  Task  Force  1978;  Derksen  et  al.  1982). 

Systematic  surveys  of  arctic  tundra  waterfowl 
with  small  aircraft  began  as  early  as  1948  (Smith  and 
Allen  1948)  and  have  been  greatly  improved  since 
then  (King  1970).  Bergman  et  al.  (1977)  devised 
a  system  for  classifying  those  tundra  wetlands  that 
support  large  breeding  bird  populations.  The  system 
uses  data  on  size  of  wetland,  vegetation,  water  depth, 
and  water  chemistry  to  identify  eight  key  habitat 


Alpine  Stand  Types 


IGNEOUS  ROCK  AREA 


SEDIMENTARY  ROCK   AREA 


SOLIF 


Figure  3.     Schematic  representation  of  alpine  habitat  types  and  features  (after  Hoffmann  1984). 

Tundra 


155 


classes  of  importance  to  waterbirds  (Table  2).  Pre- 
liminary studies  suggest  that  identification  of  the 
class  of  wetland  associated  with  a  given  tundra  area 
will  be  an  acceptable  predictor  of  the  waterbird 
species,  using  such  an  area  without  a  direct  survey 
of  the  bird  populations  themselves  (U.S.  Fish  and 
Wildlife  Service  1982). 

Some  species  may  have  both  migratory  and  resi- 
dent populations,  even  within  the  same  general  area, 
i.e.,  the  caribou.  In  other  situations,  a  population 
may  shift  from  resident  to  migratory  status  depend- 
ing on  certain  environmental  conditions.  In  the  arc- 
tic, the  gyrfalcon  is  one  such  species  in  which  prey 
availability  can  shift  behavior. 


As  with  the  arctic,  alpine  tundra  areas  show  low 
species  diversity.  A  major  difference  is  that,  espe- 
cially with  resident  species,  alpine  populations  often 
exist  in  small  islands  of  habitat.  Habitats  formerly 
occupied  by  extirpated  populations  will  not  be 
quickly  recolonized  and  genetic  conservation  is 
more  likely  to  be  a  management  concern.  For  exam- 
ple, an  endemic  species  of  butterfly  (Boloria  acrone- 
mid)  has  been  found  only  in  three  alpine  meadows 
in  the  San  Juan  Mountains  of  Colorado  (Gall  1984). 


Thirteen  species  of  herbivorous  mammals  in- 
habit the  alpine  area  of  the  Beartooth  Mountains, 
Montana,  and  only  six  species  of  herbivorous  or 


Table  2.     Wetland  classification  system  of  Bergman  et  al.  (1977)  with  two  additional  types  from  Derksen 
(1979). 


Wetland  Class 

Description 

Class  1:          Flooded  tundra 

Shallow  waters  formed  during  spring  thaw  when  melt  water 
overflows  stream  basins  or  is  trapped  in  vegetated  tundra 
depressions.  These  temporary  wetlands  form  in  shallow 
basins,  polygons,  or  broad  meadows. 

Class  II:         Shallow-Carex 

Shallow  ponds  with  a  gently  sloping  shore  zone  surrounded 
by  and  usually  containing  emergent  Carex  aquatilis  in  a 
central  open  water  zone.  These  ponds,  formed  in  low 
centers  of  polygonal  ground,  often  produce  a  mosaic  pattern 
of  ridges  and  ponds. 

Class  III:        Shallow-AYcfopfr/'/a 

Ponds  or  pools  in  beaded  streams  containing  Arctophila 
fulva  in  the  central  zone  and  shoreward  stands  of  A.  fulva  or 
Carex  aquatilis.  Shores  are  more  abrupt  than  those  of  Class 
II  ponds.  Maximum  water  depths  range  from  20  to  50  cm. 

Class  IV:        Deep- Arctophil 'a 

Wetlands  of  either  large  pond  or  lake  size  that  lack 
emergents  in  the  central  zone  and  contain  stands  of 
Arctophila  fulva  near  the  shore. 

Class  V:         Deep-open 

Large,  deep  lakes  that  have  abrupt  shores,  sublittoral 
shelves,  and  a  deep  central  zone. 

Class  VI:        Basin-complex 

Large,  partially  drained  basins  that  may  contain  nearly 
continuous  water  in  spring  due  to  flooding  of  the  bottom  by 
melt  water.  A  composite  of  several  other  classes. 

Class  VII:       Beaded  stream 

Small,  often  intermittent,  streams  consisting  of  a  series  of 
channels  formed  in  ice-wedges  and  linked  to  pools  that 
develop  at  ice-wedge  intersections. 

Class  VIII:      Coastal  wetlands 

Aquatic  habitats  that  occupy  low  areas  within  a  zone  directly 
influenced  by  sea  water. 

Class  IX:        Upland  tundra 

Ponds  characterized  by  depressions  in  upland  tussock 
vegetation  {Eriophorum  vaginatum)  that  fill  with  melt  water 
and  evolve  from  ephemeral  pools  to  permanent  ponds. 
When  mature,  these  ponds  are  typically  less  than  10  by  3  m 
with  a  maximum  depth  of  1  m. 

Class  X:         Ice- wedge  pools 

Small  (10  m  wide)  pools  formed  in  ice-wedge  troughs  in  both 
high  and  low  center  polygon  areas.  Older,  enlarged  pools 
resemble  Class  II  ponds  and  are  more  diverse  than  smaller, 
deep,  acidic  pools. 

156 


Tundra 


Uncompahgre  frittilary  butterfly. 


Pingo  and  associated  tundra  ponds. 


insectivorous  birds  regularly  breed  above  timberline. 
Eight  carnivorous  mammal  species  and  seven  to 
eight  raptor  species  spend  at  least  part  of  the  year  in 
this  alpine  environment  (Hoffmann  1974;  Price  1981). 
Carnivorous  mammals  and  predatory  hawks  and  owls 
seem  to  constitute  a  relatively  greater  proportion  of 
the  overall  species  array  compared  with  the  propor- 
tion in  temperate  or  tropical  biomes. 

In  alpine  environments  the  large  mammals,  such 
as  grizzly  bears,  bighorn  sheep  (Ovis  canadensis), 
elk,  Dall's  sheep  (Ov  is  dalli),  and  mountain  goat 
(Oreamnos  americanus),  generally  make  altitudinal 
movements,  descending  to  more  wooded  habitats 
in  winter.  Dall's  sheep  and  mountain  goats  may  on 
occasion  use  tundra  ranges  in  winter. 

Some  bird  species  also  make  seasonal  altitudinal 
movements.  In  the  western  contiguous  states,  white- 
tailed  ptarmigan  (Lagopus  leucurus)  use  a  broad 
array  of  alpine  habitats,  moving  in  winter  to  lower 
elevation  shrub  communities.  Farther  north,  where 
there  are  three  sympatric  ptarmigan  species,  the 
white-tailed  occupies  a  narrower  niche. 

Altitudinal  movements  may  even  be  on  a  daily 
basis  (Hoffmann  1974).  Many  raptors  nest  at  lower 
elevations,  but  take  advantage  of  upwelling  wind 
currents  to  move  up  to  open  alpine  terrain  for  forag- 
ing during  the  day. 


SPECIAL  HABITAT  FEATURES 

In  areas  of  continuous  or  near-continuous 
permafrost,  suitable  habitat  for  denning  or  burrowing 
species  may  be  very  limited.  The  depth  of  the  active 
layer  (seasonally  thawed  soil  zone)  can  be  an  impor- 
tant variable  influencing  wildlife  habitat  values.  Den- 
ning or  burrowing  species  depend  upon  well-drained 
or  exposed  substrates,  such  as  pingos,  where  the 


permafrost  zone  has  receded  or  is  ice-free.  Pingos 
are  conical-shaped  mounds  pushed  up  out  of  the  flat 
tundra  by  hydrostatic  pressure.  These  features  reach 
65  m  (213  ft)  in  height  and,  because  of  drainage 
and  aspect,  may  provide  relatively  high  substrate 
temperatures  as  well  as  lookouts  (Hoffmann  1974). 
Like  pingos,  coastal  bluffs,  cut  banks,  gravel  bars, 
talus  slopes,  and  sand  dunes  can  be  important  habitat 
features.  Their  number  and  extent  may  actually  be 
the  limiting  factor  for  certain  populations.  Sandy 
banks  of  only  1  m  (1  yd)  high  may  be  used  by  arctic 
foxes  (Alopex  lagopus)  for  denning  (Bee  and  Hall 
1956). 

Because  of  the  generally  flat,  treeless  character 
of  the  arctic  tundra,  river  and  coastal  bluffs  are  often 
the  principal  sites  for  nesting  raptors  outside  of  the 
mountainous  zone.  Specialized  systems  have  been 
developed  for  classifying  these  raptor  habitats  (Rit- 
chie 1979)  which  also  include  escarpments,  rock 
outcrops,  and  talus  slopes  as  one  moves  into  the  arc- 
tic-alpine environment.  Any  habitat  surveys  or  other 
intensive  activities  with  aircraft  should  be  conducted 
so  as  to  have  minimal  disturbance  on  nesting  raptors, 
peregrine  falcons  (Falco  peregrinus)  in  particular 
(Fyfe  and  Olendorff  1976).  Because  such  sites  are  fre- 
quently also  preferred  sites  for  extraction  of  con- 
struction materials  (gravel  or  sand)  or  sometimes 
even  as  camp  or  construction  sites,  it  is  important  to 
inventory  these  features  and  assess  their  value  as 
habitat. 

The  presence  of  deep  snowdrifts  is  an  important 
habitat  feature  for  denning  wolverines  {Gulo  gulo), 
polar  bears  (Ursus  maritimus),  and  brown  bears  (U. 
arctos).  The  drift  itself  may  form  the  den  or  it  may 
form  an  additional  insulative  cover  over  shallow  rock 
or  earth  dens.  On  the  other  hand,  Reynolds  et  al. 
1976  suggested  that  brown  bears  in  the  arctic  do 
not  den  until  about  the  top  10  cm  (4  in.)  of  soil  is 
frozen  because  otherwise  the  coarse-textured  soils 
would  collapse  on  the  excavation. 


Tundra 


157 


Some  habitat  features  may  be  transient.  For  ex- 
ample, certain  arctic  river  deltas  may  provide  the 
only  open  water  for  early  arriving  waterfowl  and 
shorebirds  at  a  time  when  tundra  lakes  are  still  fro- 
zen. These  critical  areas  of  open  water  are  easily 
discernible  from  aerial  reconnaissance  or  from  Land- 
sat  imagery  (NPR-A  Task  Force  1978;  U.S.  Fish  and 
Wildlife  Service  1982;  Lent  1985).  Aufeis  or  naled, 
the  sheets  of  ice  remaining  in  river  and  braided 
streambeds  late  into  summer,  are  another  special 
habitat  feature  that  provides  cool,  insect  relief  areas 
for  caribou,  moose  (Alces  alces),  and  other  large 
mammals. 

Thus,  the  special  importance  of  riparian  habitats 
and  riparian-associated  habitat  features  within  arctic 
tundra  landscapes  cannot  be  stressed  too  strongly. 
Not  only  do  they  provide  microrelief  and  thawed 
soils  for  burrowers  and  bluffs  and  banks  for  nesting 
sites,  they  also  provide  narrow  zones  of  brushy  habi- 
tats that  are  critical  to  the  requirements  of  other 
species  (NPR-A  Task  Force  1979;  Pamplin  1979;  Moi- 
toret  et  al.  1985). 


INVENTORY  AND  MONITORING  SYSTEMS 

Range  Surveys 

Most  of  the  early  efforts  in  the  1920s  and  1930s 
to  map,  describe,  and  inventory  arctic  and  subarctic 
tundra  ranges  were  conducted  because  of  their  eco- 
nomic importance  for  domestic  reindeer  in  the 
U.S.S.R.  and  in  Alaska  (Palmer  1926).  With  the  col- 
lapse of  the  reindeer  industry  in  Alaska  and  the  dis- 
ruptions resulting  from  World  War  II  in  the  Soviet 
Union,  there  was  a  hiatus  of  scientific  effort.  In  the 
late  1950s,  systematic  attention  was  first  turned  to 
the  study  of  caribou  and  their  habitats  in  Alaska  and 
Canada.  This  was  made  possible  by  improvements 
in  the  performance  and  availability  of  small  aircraft, 
providing  biologists  with  a  mobility  and  a  perspec- 
tive not  previously  possible.  This  work  tended  to 
focus  on  taiga  ranges  used  by  wintering  caribou. 

In  the  1950s,  scientific  studies  of  Scandinavian 
domestic  reindeer  got  underway  and  Soviet  work 
was  intensified.  Banfield  (1954)  and  Tener  (1965) 
also  began  their  classic  studies  in  the  Canadian  Arc- 
tic. In  Alaska,  two  major  reviews  of  literature  and 
field  work  relating  to  caribou  range  were  done  by 
Courtright  (1959)  and  Pegau  (1968).  Pegau  put 
emphasis  on  specific  comparisons  of  vegetation  in- 
ventory techniques  suitable  for  tundra  ranges. 

Among  the  pioneer  workers  in  Scandinavia, 
Skuncke  (1969)  devised  a  detailed  system  for  aerial 
surveys  of  reindeer  pastures  to  estimate  pasture  car- 
rying capacities  and  trends.  Much  of  the  system  is 
specialized  for  the  Scandinavian  situation,  however, 
and  also  requires  very  experienced  observers. 


Beaded  streams  furnish  good  nesting  habitat  for  white- 
fronted  geese  and  ducks. 

Special  inventory  techniques  are  required  for 
tundra  areas  where  lichens  are  a  significant  compo- 
nent or  likely  to  be  important  forage  plants.  Pegau 
( 1970)  used  a  919-cnT  (1-ft")  frame  and  visually  esti- 
mated lichen  cover  based  on  random  throws.  Using 
this  simple  technique,  he  was  able  to  assess  the  dam- 
age to  lichens  caused  by  holding  or  herding  reindeer 
on  lichen  ranges  during  dry  periods.  Pegau  (1968) 
also  measured  growth  rates  for  lichens  in  Alaska. 
Although  the  randomness  of  random  throws  has 
been  legitimately  criticized,  the  technique  has 
proven  useful  (Greig-Smith  1983). 

Changes  over  time  in  arctic  tundra  vegetation 
have  been  successfully  measured  in  other  studies  as 
well.  In  a  study  of  the  effects  of  grazing  by  confined 
musk-oxen  on  tussock  and  mat  cushion  tundra,  Mc- 
Kendrick  ( 1981 )  used  the  walking  point  method 
of  Owensby  (1973)  to  measure  basal  cover  and  spe- 
cies composition.  With  these  techniques,  he  was  able 
to  demonstrate  significant  reduction  in  lichen  and 
shrub-standing  crops  caused  by  musk-oxen.  Thus  the 
measured  difference  with  regard  to  musk-oxen  influ- 
ence on  shrubs  was  significantly  different  from  that 
reported  from  numerous  reindeer  studies.  Intense 
grazing  pressure  by  reindeer  tends  to  promote  inva- 
sion by  shrub  species.  Racine  (1979)  compared  soil 
condition,  permafrost  depth,  and  vegetation  on  plots 
before  and  after  tundra  fires  on  the  Seward  Penin- 
sula. He  used  simple  square  meter  (square  yard) 
plots,  recording  stem  density  and  percentage  cover 
to  measure  significant  differences  before  and  after 
the  fires. 

Thilenius  (1979)  presented  the  best  recent  dis- 
cussion of  the  problems,  concepts,  and  inexact  state- 
of-the-art  for  assessing  and  monitoring  alpine  range 
communities.  He  noted  that  classifying  range  condi- 
tion is  a  great  problem  in  alpine  range  management. 
Range  condition  classes  often  seem  to  be  confused 
with  site  climax  classes.  For  example,  cushion  plant 
communities  dominated  by  Geum  rossii  have  been 


158 


Tundra 


treated  by  range  managers  as  poor  condition  com- 
munities when  in  fact  such  forb-dominated  commu- 
nities, common  on  xeric  sites,  may  represent 
excellent  range  condition.  He  recommended  the 
system  of  Lewis  (1970)  who,  he  stated,  properly  rec- 
ognized the  unique  microclimate  and  soil  features 
of  alpine  ranges.  Thilenius  (1979)  also  emphasized 
that  alpine  range  monitoring  must  be  based  on  an- 
nual sampling  because  of  the  high  degree  of  year-to- 
year  variability. 

As  with  any  class  of  vegetation  or  habitat,  de- 
scription and  mapping  of  soils  types  is  an  important 
part  of  habitat  analysis.  The  characterization,  inven- 
tory, and  mapping  of  tundra  soils  has  become  partic- 
ularly important  in  the  last  two  decades  because 
arctic  development  has  required  an  increased  com- 
mitment to  land  planning  and  engineering  skills 
(Everett  1982). 

In  arctic  soils,  the  presence  of  permafrost,  its 
depth  and  water  content,  as  well  as  the  nature  of  the 
overlying  vegetation  are  important  considerations. 


When  permafrost  is  present,  it  is  frequently  associ- 
ated with  areas  of  relatively  little  relief.  This,  plus 
the  generally  fine  texture  of  soils  and  spongy  moss 
cover,  limits  water  movement  and  results  in  satu- 
rated soils  (Figure  4).  Gley  (wet)  soils  are  indicated 
by  cottonsedge  (Eriophorum  sp.)  tussock  and  dwarf 
heath  communities.  Bog  soils  are  characterized  as 
cottongrass  sedges  (Carex  aquatilis)  and  Dupontia 
Plant  communities  dominated  by  lichens,  mosses, 
and  dwarf  heaths  with  scattered  herbs  prevail  over 
shallow  dry  or  rocky  soils  where  the  permafrost 
is  also  dry  and  generally  1  m  (1  yd)  or  more  deep. 
Similar  vegetation  also  occurs  on  dry  raised  patches 
of  peaty  material  (Linell  and  Tedrow  1981). 


Remote  Sensing  Applications 

Because  of  the  vast  areas  involved  in  arctic  tun- 
dra management,  the  lack  of  access,  except  by  ex- 
pensive aircraft,  and  the  short  field  season,  there  is  a 
strong  rationale  for  the  use  of  remote  sensing  sys- 
tems for  mapping  and  inventory  of  vegetation  and 


ACTIVE 
LAYER 


WELL-DRAINED  SITE 


POORLY  DRAINED  SITE 


BOG 


Figure  4.     Idealized  cross-section  of  tundra  landscape  from  higher  well-drained  ground  on  left  to  peaty 
ground  on  right  (after  Linell  and  Tedrow  1981). 


Tundra 


159 


wildlife  habitats.  Several  efforts  have  been  directed 
in  recent  years  toward  the  use  of  satellite  sensors  for 
that  purpose,  particularly  Landsat.  These  trial  efforts 
are  described  by  Anderson  and  Belon  (1973),  Ander- 
son et  al.  (  1974),  LaPerriere  (  1978),  George  et  al. 
(1977),  and  Nodler  et  al.  (1978).  Anderson  and  Be- 
lon (1973),  Lent  and  LaPerriere  (1974),  and  Hall 
et  al.  ( 1980)  described  the  use  of  Landsat  imagery  to 
delineate  and  monitor  change  in  areas  that  have 
been  subject  to  recent  wildfires.  This  then  provides  a 
useful  tool  to  monitor  changes  in  wildlife  habitats 
due  to  the  wildfires  and  subsequent  recovery  of  tun- 
dra vegetation.  Lent  (1980)  also  used  Landsat  im- 
agery to  monitor  gross  changes  in  extent  of  snow 
cover.  More  recently,  Acevedo  et  al.  (1982)  used 
Landsat  for  environmental  mapping  and  related  base- 
line studies  on  a  portion  of  the  coastal  tundra  of 
the  Arctic  National  Wildlife  Refuge.  To  a  certain  ex- 
tent, wildlife  biologists  have  used  wildlife  habitat 
inventory  and  mapping  efforts,  and  use  classification 
schemes  that  take  advantage  of  the  broad  land  cover 
types  discriminated  from  Landsat  imagery  (U.S.  Fish 
and  Wildlife  Service  1982,  1984).  As  noted  earlier, 
Moitoret  et  al.  ( 1985)  reported  on  some  of  the  po- 
tential problems  with  such  approaches. 


Satellite  imagery  from  Landsat  and  weather  satel- 
lites has  also  been  used  to  assess  the  snow  and  ice 
conditions  encountered  by  nesting  geese  in  the  Ca- 
nadian Arctic.  By  assessing  the  year-to-year  timing  of 
snow  and  ice  melt  on  a  broad  scale,  this  nesting 
habitat  information  has  been  used  to  successfully 
predict  productivity  of  species  such  as  brant  {Branta 
bernicla),  snow  geese  {Chen  caerulescens),  arctic- 
breeding  Canada  geese  (B.  canadensis),  and  greater 
white-fronted  geese  (Anser  albifrons)  (Reeves  et 
al.  1976). 


A  major  limitation  to  the  use  of  satellite  imagery 
over  much  of  the  arctic  is  the  high  frequency  of 
overcast  days.  The  large  number  of  overcast  days  is 
partially  compensated  for  by  the  polar  convergence 
of  Landsat  orbits,  resulting  in  bursts  of  imagery  on 
successive  days  at  high  latitudes.  Because  the  phen- 
ology changes  so  rapidly  and  the  windows  of  good 
weather  are  frequently  so  short,  it  is  difficult  to  ob- 
tain comparable  imagery  in  successive  years  to  docu- 
ment long-term  trends. 


The  use  of  satellite  imagery  for  mapping  and 
monitoring  alpine  habitats  is  at  present  much  more 
difficult  because  of  the  problems  resulting  from  top- 
ographic complexity,  including  the  effects  of  shad- 
ows on  surface  albedo.  However,  Craighead  et  al. 
(1982)  classified  and  mapped  alpine  and  subalpine 
vegetation  in  western  Montana,  using  Landsat  im- 
agery; a  computer-assisted,  multi-spectral  analysis 
system;  and  multiple-level,  ground-truthing  to  inven- 
tory grizzly  bear  habitat. 


Simple  remote  sensing  techniques  involving 
small  aircraft  are  also  useful.  For  example,  good  suc- 
cess in  mapping  tundra-vegetation  types  has  been 
reported  for  false-color,  infrared  photography  and 
simple  hand-held,  35  mm  cameras  in  fixed- wing  air- 
craft (Holt  1980).  Hesjedal  and  Larsson  (1975)  re- 
ported color-infrared  film  to  be  useful  for  the 
detection  of  eroded  and  overgrazed  areas  in  Fenno- 
scandian  alpine  regions,  but  recommended  use  of 
stereo-panchromatic  photography  as  most  cost-effec- 
tive for  generalized  alpine  surveys. 

Use  of  stereoscopic  pairs  of  medium-  to  large- 
scale  imagery  is  helpful  for  raptor  nesting  habitat 
surveys.  Some  habitat  features  such  as  pingos  are  eas- 
ily mapped  from  small-scale  aerial  photographs  (Sims 
1983). 

Sims  (1983)  used  large-scale,  color-infrared,  70- 
mm  aerial  photography  to  map  lichen  types  which 
revealed  reindeer  winter  range  types  and  carrying 
capacities  on  the  Tuktoyaktuk  Peninsula,  Canada. 
Microdensitometric  readings  on  the  original  film 
were  diagnostic  for  areas  with  over  5%  lichen  cover. 
About  81%  of  sample  sites  were  correctly  classified. 

Inventory  and  Monitoring  for  Impact 
Assessment 

The  manager  of  wildlife  and  wildlife  habitat  in 
tundra  systems  is  increasingly  involved  in  issues  and 
decisions  involving  human  uses  of  these  lands.  In  the 
arctic,  exploration,  development,  transportation, 
and  production  of  oil,  gas,  and  mineral  ores  are 
prime  examples  of  these  land-use  problems.  In  alpine 
environments,  issues  are  frequently  related  to  inten- 
sive recreational  use,  livestock  use,  or  mineral  devel- 
opment. Thus  this  chapter  places  considerable 
emphasis  on  inventory  and  monitoring  techniques 
that  have  had  demonstrable  applications  in  such 
management  situations. 

Surface  disturbance  and  alteration  of  permafrost 
regime  have  clearly  had  undesirable  aesthetic  effects 
in  arctic  areas,  as  well  as  caused  localized  erosion 
problems.  The  long-term  ecological  consequences  of 
such  changes  have  been  less  clear-cut. 

Nevertheless,  for  purposes  of  planning  and  in- 
ventory before  proposed  developments  in  the  arctic, 
mapping  of  permafrost  temperature,  depth,  moisture 
content,  and  protective  vegetative  cover  have  re- 
ceived paramount  attention  (Linell  and  Tedrow 
1981).  Similarly,  monitoring  efforts  have  often  fo- 
cused on  changes  in  permafrost  status  and  the  over- 
lying insulating  vegetation  (Brown  et  al.  1978).  The 
fragility  or  sensitivity  to  damaged  tundra  areas  is 
directly  proportional  to  the  ice  content  of  the 
permafrost  layer  and  to  the  mean  annual  ground 
temperature  (Ives  1970),  as  well  as  to  the  insulative 
qualities  of  the  surface  cover.  The  ability  to  predict 


160 


Tundra 


such  sensitivity  is  now  good.  A  monitoring  program 
may  track  the  success  of  stipulations  relating  to  sur- 
face disturbance  or  recovery  programs  necessitated 
by  unavoidable  disturbance  (Hok  1969,  1971;  van 
Cleve  1977;  Brown  and  Berg  1980). 

During  the  major  monitoring  effort  associated 
with  construction  of  the  Trans-Alaska  Pipeline  Sys- 
tem, terrain  along  the  corridor  was  classified  into  1 2 
wildlife  habitat  types.  Of  these,  four  (subalpine,  al- 
pine tundra,  tussock  tundra,  and  wet-meadow  tun- 
dra) are  clearly  associated  with  or  are  subcategories 
of  tundra.  Three  others  (wetlands,  riparian  willow, 
and  unvegetated  floodplains)  frequently  occur  as 
islands  or  narrow  strips  within  tundra  habitats.  A 
panel  of  biologists  gave  independent  ratings  for  the 
relative  quality  of  these  wildlife  habitats,  weighing 
the  importance  of  24  wildlife  groups  (mostly  taxa  at 
the  family  level)  equally.  Based  on  this  relatively 
crude  evaluation  process,  wet-meadow  tundra  and 
wetlands  received  the  highest  quality  ratings.  Ripar- 
ian willow  and  tussock  tundra  were  rated  high  to 
moderate.  The  alpine  and  subalpine  types  were  rated 
low-moderate.  A  comparison  of  preconstruction  and 
post-construction  imagery  (aerial  photography)  re- 
vealed that  12,709  ha  (31,403  a.)  of  wildlife  habitat 
were  altered  or  destroyed  during  this  major  project. 
Nearly  half  of  these  lands  were  in  high  quality  wet- 
lands and  wet-meadow  tundra  habitats.  However,  the 
greatest  deviation  from  impacts  predicted  in  the 
environmental  impact  statement  (EIS)  was  in  the 
land  altered  by  construction  material  sites.  The  EIS 
predicted  that  the  disturbance  for  material  (primarily 
gravel)  sites  would  be  about  2,400  ha  (6,000  a.) 
compared  with  the  actual  figure  of  nearly  4,800  ha 
( 1 2,000  a. ),  as  determined  in  this  monitoring  study 
(Pamplin  1979). 

Because  of  the  rapid  succession  of  seasonal 
changes  in  tundra  areas  and  the  importance  of  these 
events  to  phenology,  animal  abundance,  and  sensitiv- 
ity, it  is  important  that  the  biologist  prepare  prelimi- 
nary charts  of  these  events.  Freezing  over  of  lakes, 
rivers,  and  coastal  lagoons;  the  timing  of  critical  ice 
depths;  establishment  of  permanent  winter  snow 
cover;  and  rate  of  spring  ablation  are  all  events  criti- 
cal to  planning  inventories  and  environmental  assess- 
ments. These  events  also  constitute  essential 
information  for  any  environmental  impact  analysis 
(U.S.  Fish  and  Wildlife  Service  1982;  Lent  1985). 

In  what  is  certainly  the  most  elaborate  mapping 
project  undertaken  for  a  tundra  area  and  perhaps 
one  of  the  most  elaborate  done  anywhere,  Walker 
et  al.  (1980)  produced  a  geobotanical  atlas  of  the 
Prudhoe  Bay  region  in  Alaska.  Among  the  many  maps 
are  ones  showing  relative  sensitivity  of  areas  to  oil 
spills  and  sensitivity  to  damage  by  vehicles.  Also 
shown  are  lichen  cover,  breeding  bird  density,  and 
the  thickness  of  the  active  layer  (layer  subject  to 
seasonal  freezing  and  thawing).  One  of  the  more  use- 


ful maps,  an  oil  spill  sensitivity  map  and  the  proce- 
dures for  its  production,  is  described  in  detail  in 
Walker  et  al.  1978.  They  were  able  to  demonstrate 
great  differences  in  the  sensitivity  of  various  plant 
communities  to  either  diesel  or  crude  oils  by  meas- 
uring live  cover  on  1-m"  ( 10.9-ft")  quadrats  before 
and  1  year  after  experimental  "spills."  These  map 
types  are  useful  to  those  who  must  assess  and  inven- 
tory wildlife  habitats  and  who  are  involved  in  plan- 
ning for  tundra  areas  in  relation  to  potential  impacts 
on  wildlife. 

The  production  of  an  atlas  like  that  of  Walker  et 
al.  (1980)  is  beyond  the  budget  of  most  land  man- 
agers. Furthermore,  many  of  the  methodologies  used 
in  producing  the  maps  are  not  described  in  detail 
and  the  reader  must  refer  to  the  references  cited  in 
the  atlas. 

Sensitivities  of  alpine  tundra  areas  to  human 
impacts  are  also  well-documented  by  Willard  and 
Marr  (1971 ).  Using  sample  species  cover  measures, 
they  monitored  the  disturbance  and  recovery  of 
alpine  vegetation.  In  general  they  reported  ex- 
tremely slow  recovery  rates;  in  some  instances  no 
recovery  had  occurred  after  4  years  of  protection, 
based  on  total  cover  and  species  composition  meas- 
ures. For  Kobresia  sp.  meadows  it  was  estimated  that 
damage  to  the  turf  caused  such  severe  erosion  and 
loss  of  the  thin  humus-enriched  horizons  that 
hundreds  of  years  might  be  necessary  to  restore  the 
original  climax  system. 

Bell  and  Bliss  (1973)  used  10x50  cm  (4x20  in.) 
clipping  plots  and  line  intercept  sampling  of  plant 
cover  to  measure  recovery  rates  of  alpine  plant  com- 
munities after  human  trampling.  Dry  tundra  and 
scree  slopes  were  most  fragile  and  slowest  to  re- 
cover. Recovery  rates  were  significantly  more  rapid 
in  meadows  and  snowbed  communities.  They  also 
measured  rates  of  downslope  movement  as  indica- 
tors of  disturbance. 

Snow  Cover  and  Tundra  Habitats 

Any  survey  or  inventory  of  winter  habitats  im- 
portant to  wildlife  must  include  the  collection  and 
analysis  of  information  on  snow  cover  if  it  is  to  pro- 
vide the  manager  with  a  proper  assessment  of  carry- 
ing capacity,  relative  importance  of  various  subareas, 
and  temporal  patterns  of  use.  In  their  simplest  form, 
inventories  of  snow  cover  may  be  limited  only  to 
recording  the  presence  or  absence  and  the  timing  of 
accumulation  and  ablation.  Obviously,  data  on  depth 
or  thickness  greatly  increase  the  value  of  a  survey. 
Data  on  density,  hardness,  ice,  crust  layers,  and  pat- 
terns of  drifting  or  distribution  in  relation  to  topog- 
raphy are  also  desirable. 

Pruitt  ( 1959)  was  the  first  in  North  America  to 
employ  extensive  aerial  surveys  and  snow  measure- 


Tundra 


161 


merits  to  relate  snow-cover  characteristics  to  caribou 
movements  and  distribution.  The  study  area  for  his 
classic  investigation  of  migratory  caribou  ecology  lay 
in  the  taiga,  with  only  a  small  portion  in  the  tundra 
zone.  Nevertheless,  the  concepts  and  techniques 
he  pioneered  are  applicable  to  both  zones. 

Pruitt  made  snow  measurements  using  a  set  of 
standard  snow  instruments  developed  and  adopted 
by  the  National  Research  Council  of  Canada  (Klein 
et  al.  1950).  With  these  instruments,  several  meas- 
ures of  snow-cover  morphology  may  be  made,  of 
which  thickness,  density,  and  hardness  are  usually 
considered  most  pertinent  for  ecological  investiga- 
tions. Richens  and  Madden  (1973)  reported  on  modi- 
fications to  these  instruments  to  make  them  more 
suitable  for  wildlife  biologists. 

Other  workers  (LaPerriere  and  Lent  1977;  Lent 
and  Knutsen  1971;  Skogland  1978;  Brooks  and  Collins 
1985)  carried  out  snow  surveys  using  the  Ramsonde 
penetrometer  (Figure  5).  The  penetrometer  was 
developed  in  the  European  Alps  for  avalanche  stud- 
ies and  is  most  suitable  for  use  with  dense,  wind- 
blown snow  with  complex  thin  crusts  and  ice  layers, 
such  as  is  typical  of  alpine  and  some  tundra  environ- 
ments. It  measures  the  force  required  to  penetrate 
a  given  layer  from  above. 


HAMMER 


GUIDE  ROD 


SHAFT 


60°      CONE 


V 


Figure  5.     Schematic  representation  of  Ramsonde 
penetrometer. 


Pruitt's  studies  and  those  of  later  investigators 
working  with  barren-ground  caribou  (LaPerriere  and 
Lent  1977)  and  mountain  reindeer  (Skogland  1978) 
have  confirmed  the  preference  of  winter  ranges  with 
relatively  low  depths,  densities,  and  hardnesses. 
Musk-oxen  have  shown  even  stronger  selection  for 
such  winter  ranges  (Lent  and  Knutsen  1971).  For 
musk-oxen,  it  has  been  shown  repeatedly  that  an  un- 
derstanding and  knowledge  of  the  prevalence  of 
crusts,  ice  layers,  and  groundfast  ice  on  winter 
ranges  is  critical  to  proper  habitat  evaluation  (Vibe 
1967;  Lent  and  Knutsen  1971;  Thomas  et  al.  1981). 

Miller  et  al.  (1982)  combined  surveys  of  cari- 
bou winter  trails  and  feeding  sites  with  surveys  of 
snow  cover  and  ice  conditions  with  subsequent  veg- 
etation sampling.  Vegetation  was  recorded  in  sum- 
mer as  percentage  frequency  of  occurrence  in 
quadrats  at  the  feeding  sites  and  at  points  away  from 
the  feeding  sites.  This  investigation  showed  clearly 
that  several  range  types  with  desirable  forage  species 
were  being  underutilized  because  of  difficulty  of 
access.  Winter  feeding  was  directed  primarily  at 
those  poor  forage  species  that  were  exposed  and 
directly  available.  Similarly,  Gaare  and  Skogland 
(1975)  concluded  that  only  25%  of  the  potential 
winter  forage  on  Norwegian  alpine  reindeer  ranges 
was  actually  available  for  consumption.  Thus,  sum- 
mer-only inventories  of  winter  ranges  can  greatly 
exaggerate  the  winter  range  carrying  capacities. 
Brooks  and  Collins  (1985)  reported  on  snow  survey 
methods  to  assess  forage  availability. 

Although  the  presence  of  a  deep,  dense  snow 
cover  may  restrict  or  prevent  winter  use  of  certain 
tundra  ranges  by  ungulates,  it  may  also  restrict  use 
by  other  species,  such  as  wolves  (Canis  lupus),  wol- 
verine, and  ptarmigan.  The  same  thick  snow  cover 
may  provide  protection  from  the  abrasive,  desiccat- 
ing actions  of  winter  winds  and  windblown  snow. 
This  "subnivean"  environment  is  relatively  warm, 
moist,  and  stable  and  supports  small  mammals  and 
invertebrates.  Special  techniques  for  studying  this 
environment  were  described  by  Schmidt  (1984). 

The  best  lichen  growths  on  reindeer  ranges  are 
associated  with  moderately  deep  snow  cover  be- 
cause of  the  protection  and  the  spring  moisture  pro- 
vided. Snow  accumulations  in  streambeds  and  other 
depressions  provide  specialized,  critical  habitats  for 
numerous  species  by  protecting  tundra  shrubs  from 
winter  damage.  The  vegetative  associations  of  these 
snowbeds  in  turn  support  some  specialized  wildlife 
species  (for  example,  the  singing  vole  [Microtus 
miurus])  and  provide  high  quality  late  summer  for- 
age for  certain  herbivores  after  other  forage  plants 
have  gone  to  senescence.  In  a  detailed  study  of  an 
alpine  area,  Thorn  ( 1982)  mapped  and  inventoried 
the  distribution  of  the  pocket  gopher  {Thamomys 
talpoides)  and  the  distribution  of  related  plant  com- 


162 


Tundra 


munities.  He  found  that  the  plant  communities  were 
most  highly  correlated  with  the  mean  duration  and 
mean  winter  depth  of  the  snowpack. 

OTHER  SPECIAL  CONSIDERATIONS  FOR 
WILDLIFE  MANAGEMENT  IN  THE  ARCTIC 

The  low  primary  production  rates  for  vascular 
plants  in  tundra  communities  cannot  be  emphasized 
too  strongly.  Annual  net  production  estimates  be- 
tween 50  and  500  g/m    (2.1  and  21.0  oz/yd  )  are 
typical.  Lichens  and  mosses  may  add  200  g/m    (8.4 
oz/yd  )  to  these  figures.  Furthermore,  in  both  arctic 
and  alpine  environments,  the  ratio  of  belowground 
to  aboveground  biomass  for  vascular  plants  is  ex- 
tremely high  in  comparison  with  temperate  zone 
communities;  values  as  high  as  30:1  have  been  re- 
ported (Wielgolaski  1975). 

Not  only  are  animal  populations  constrained, 
directly  or  indirectly,  by  this  low  primary  produc- 
tion but  many  wildlife  populations  in  the  arctic  are 
particularly  difficult  to  inventory  or  monitor  because 
their  numbers  display  extreme  fluctuations  under 
natural  conditions.  Population  peaks  can  markedly 
alter  vegetation  and  other  habitat  features  in  some 
instances,  leading  to  declines  and  completion  of 
a  population  cycle. 


Willow  ptarmingan  in  dry-  tundra  area. 


The  brown  lemming  {Lemmns  sibiricus)  is  the 
classic  example  of  a  cyclic  species  with  extreme 
fluctuations  in  numbers.  In  identical  trapping  efforts 
for  over  a  decade  from  less  than  1/ha  (2.5/a.)  to  highs 
in  2  years  when  it  reached  about  200/ha  ( 500/a. )  in 
the  wet  tundra  of  coastal  Alaska,  density  of  this  spe- 
cies has  varied.  Even  in  tussock  tundra  the  cyclic 
variations  can  be  pronounced.  Pruitt  (1966)  de- 
scribed variation  of  about  700%  in  biomass  of  small 
mammals  trapped  on  the  same  plots  over  3  years 
of  study. 

Certain  predatory  birds  that  are  closely  tied  to 
small  mammal  numbers,  such  as  pomarine  jaegers 
(Stercorarius pomarinus)  and  snowy  owls,  may  be 
absent  over  large  areas  during  years  of  low  numbers 
of  small  mammals.  The  implications  of  such  extreme 
fluctuations  for  the  design  of  monitoring  systems 
are  obvious:  baseline  conditions  are  difficult  if  not 
impossible  to  establish,  and  relating  population 
change  to  casual  management  factors  is  also  an  often 
impossible  task. 

Clumping  phenomena  also  complicate  inventory 
and  monitoring  efforts.  As  noted  earlier,  even  so- 
called  resident  species  may  be  highly  mobile  in  arc- 
tic tundra  regions.  Arctic  foxes  may  travel  hundreds 
of  miles,  deserting  previous  denning  areas  (Bannikov 
1970).  Ptarmigan  may  coalesce  by  the  thousands  in 
certain  favorable  river  valleys  in  winter  but  disperse 
widely  over  hundreds  of  square  miles  of  upland  tun- 
dra in  summer  (Irving  et  al.  1967;  NPR-A  Task  Force 
1979).  Caribou  may  go  several  years  without  making 
significant  use  of  certain  winter  ranges.  Then,  for 
reasons  yet  unknown,  they  may  converge  in  the 
thousands  on  those  same  wintering  areas  for  several 
successive  years. 

Not  only  does  such  clumping,  erratic  distribu- 
tion patterns  and  population  fluctuations  make  it 
difficult  to  measure  natural  baselines  by  which  man's 
influences  can  be  determined,  they  also  make  it  diffi- 
cult to  apply  concepts  of  sustained  yield  for  either 
population  or  habitat  management  (Beddington  and 
May  1977).  Sustained  stability  is  rare  and  generally 
unachievable  as  a  management  goal  in  arctic 
environments. 

Dunbar  (1973)  stated  the  problem  in  somewhat 
different  terms  by  pointing  out  that  arctic  systems 
show  great  instability  on  small  geographic  scales,  but 
overall  stability  on  a  large  geographic  scale.  Thus, 
large  expanses  are  necessary  for  recovery  from 
either  natural  or  man-made  perturbations.  Dunbar 
(1973:180)  sums  this  up  succinctly:  "Arctic  systems 
must  have  space.  They  must  also  have  time."  The 
latter  statement  is  a  reference  to  the  slow  recovery 
rates  of  these  systems  following  stresses.  Dunbar 
considered  arctic  lakes,  because  of  their  slow  turn- 
over rates  and  because  they  are  often  isolated  (that 
is,  lacking  the  possibilities  for  rapid  restoration  from 


Tundra 


163 


a  broader  system),  as  the  most  vulnerable  parts  of 
arctic  systems.  Migratory  shorebirds  and  waterfowl 
are  a  principal  class  of  consumers  dependent  in  large 
part  on  these  lakes  and  ponds  and  their  relatively 
productive  invertebrate  populations  and  emergent 
vegetation.  They  are  thus  vulnerable  because  of  their 
ties  to  the  aquatic  system.  This  localized  vulnerabil- 
ity is  usually  compensable  because  of  the  broad  spa- 
tial distribution  of  the  populations.  Critical 
vulnerabilities  occur  whenever  major  portions  of 
populations  exist  in  concentrations  (as  with  the 
molting  brant  in  the  Teshekpuk  Lake  area  of  Alaska) 
or  as  isolated  clumps  with  little  interchange  (as  with 
snow  goose  colonies).  Snow  goose  nesting  colonies 
represent  situations  analogous  to  islands  or  to  iso- 
lated lakes.  Because  of  the  strong  traditions  tying 
each  breeding  group  with  a  specific  location,  they 
function  as  isolated  units  not  readily  restored  by 
inputs  from  the  broader  system. 


productivity,  high  degree  of  annual  variability,  short 
growing  season,  and  habitat  patchiness.  Migratory 
species  are  relatively  large  components  of  the  faunas. 
The  short  season  for  conventional  field  work  and 
the  rapid  phenological  changes  during  that  season 
present  special  challenges. 

A  major  limitation  to  the  design  of  ecologically 
based  monitoring  systems  in  arctic  and  alpine  habi- 
tats is  the  lack  of  understanding  of  successional  pro- 
cesses and  stages. 

The  distribution,  duration,  depth,  and  physical 
attributes  of  snow  cover  are  of  particular  importance 
in  tundra  habitats  and  contiguous  boreal  and  alpine 
areas.  Special  techniques  are  available  to  describe 
and  measure  these  elements  of  wildlife  habitats.  Dis- 
tribution and  physical  attributes  of  permafrost  and 
frost  phenomena  are  also  of  special  concern  in  tun- 
dra areas. 


CONCLUSION 

Tundra  habitats  are  dominated  by  low  herba- 
ceous growth  forms;  shrubs  are  important  compo- 
nents in  restricted  areas,  and  lichens  and  mosses  are 
relatively  important  components  in  comparison  with 
other  major  habitat  types.  These  arctic  and  alpine 
habitats  present  many  unique  ecological  attributes 
for  the  wildlife  manager  to  consider  in  inventory, 
monitoring,  and  assessment  work.  These  include  low 


The  immense  geographic  scale  of  arctic  tundra 
areas  and  the  unusually  great  costs  in  undertaking 
field  work  there  leads  to  special  emphasis  on  the  use 
of  remote  sensing  techniques,  especially  those  em- 
ploying satellite-borne  sensors.  Such  techniques  are 
generally  less  useful  in  alpine  areas,  but  small-scale 
aerial  photography  has  had  several  applications.  Al- 
pine areas  present  special  problems  because  of  the 
frequent  existence  of  small  habitat  units  and  isolated 
populations. 


Arctic  fox. 


164 


Tundra 


LITERATURE  CITED 


ACEVEDO,  D.,  W  WALKER,  L.  GAYDOS,  and  J.  WRAY. 
1982.  Vegetation  and  land  cover,  Arctic  National 
Wildlife  Refuge,  Coastal  Plain,  Alaska.  U.S.  Geological 
Survey  Map  1-1443-  Misc.  Investigations  Ser.  1  sheet. 

ANDERSON,  J.G.,  C.H.  RACINE,  and  H.R.  MELCHIOR. 
1974.  Preliminary  vegetation  map  of  the  Espenberg 
Peninsula,  Alaska,  based  on  an  Earth  Resources  Tech- 
nology Satellite  image,  Pages  290-310  in  Melchior, 
H.R.  ed.  Final  Report.  Chukchi-Imuruk  Survey.  Coop. 
Park  Studies  Unit,  Univ.  of  Alaska,  Fairbanks.  5 1 7pp. 

ANDERSON,  J.H.  and  A.E.  BELON.  1973-  A  new  vegetation 
map  of  the  western  Seward  Peninsula,  Alaska,  based 
on  ERTS-1  imagery.  No.  E73- 10305.  NTIS.  Springfield, 
VA.  20pp. 

BANFIELD,  A.W.F.  1954.  Preliminary  investigation  of  the 
barren-ground  caribou.  In  two  parts.  Wildl.  Manage. 
Bull.  Ser.  1.  10A  and  10B.  Can.  Dep.  of  Northern 
Affairs  and  Nat.  Resour.  79pp.  and  1 1 2pp. 

BANNIKOV,  A.G.  1970.  Arctic  fox  in  the  U.S.S.R.  Pages 

121-129  in  Fuller,W.A.  and  P.G.  Kevan  eds.  Productiv- 
ity and  Conservation  of  Northern  Circumpolar  Lands. 
IUCN  Publ.  New  Ser.  16.  Morges,  Switzerland. 

BEDDINGTON,  JR.  and  RM.  MAY.  1977.  Harvesting  natu- 
ral populations  in  a  randomly  fluctuating  environ- 
ment. Science  197:463-465. 

BEE,  J.W.  and  E.R.  HALL.  1956.  Mammals  of  northern 
Alaska  on  the  Arctic  Slope.  Misc.  Publ.  8.  Mus.  Nat. 
Hist.,  Univ.  Kansas,  Lawrence.  309pp. 

BELL,  KL.  and  L.C.  BLISS.  1973.  Alpine  disturbance  studies: 
Olympic  National  Park.  Biol.  Cons.  5:25-32. 

BERGMAN,  R.D.,  R.L.  HOWARD,  KF.  ABRAHAM,  and  M.W. 
WELLER.  1977.  Water  birds  and  their  wetland  re- 
sources in  relation  to  oil  development  at  Storkerson 
Point,  Alaska.  Resource  Publ.  129.  U.S.  Dep.  Inter., 
Fish  and  Wildl.  Serv.  Washington,  DC.  38pp. 

BILLINGS,  W.D.  1973-  Arctic  and  alpine  vegetations:  simi- 
larities, differences  and  susceptibility  to  disturbance. 
Bio  Science  23:697-704. 

BLISS,  L.C.  1981.  Tundra  biome:  past  and  present.  Pages  3- 
24  in  Bliss,  L.C.,  O.W.  Heal,  and  J.J.  Moore  eds.  Tun- 
dra Ecosystems:  A  Comparative  Analysis.  Cambridge 
Univ.  Press,  Cambridge. 

BROOKS,  J.  and  W.  COLLINS.  1985.  Snow  cover  and  inter- 
pretation of  vegetation  habitat  inventions.  Pages  203- 
210  in  LaBau,  V.J.  and  C.L.  Kerr  eds.  Inventorying 
Forest  and  Other  Vegetation  of  the  High  Latitude  and 
High  Altitude  Regions.  Proc.  Int.  Symp.  Am.  For.  Reg. 
Conf.  Soc.  Am.  For.  Washington,  DC. 

BROWN,  J.  and  R.L.  BERG  eds.  1980.  Environmental  engi- 
neering and  ecological  baseline  investigations  along 
the  Yukon  River — Prudhoe  Bay  Haul  Road.  CRREL 
Report  80-19.  Hanover,  NH.  187pp. 

BROWN,  R.W.,  R.S.  JOHNSTON,  and  K  VAN  CLEVE.  1978. 
Rehabilitation  problems  in  alpine  and  arctic  regions. 
Pages  23-44  in  Reclamation  of  Drastically  Disturbed 
Land.  Soil  Sci.  Soc.  Am.  Madison,  WI. 

COURTRIGHT,  A.M.  1959.  Range  management  and  the 
genus  Rangifer:  A  review  of  selected  literature.  Un- 
publ.  M.S.  Thesis,  Univ.  Alaska.  Fairbanks.  172pp. 

CRAIGHEAD,  J.J.,  J.S.  SUMNER,  and  G.B.  SCAGGS.  1982.  A 
definitive  system  for  analysis  of  grizzly  bear  habitat 
and  other  wilderness  resources.  Wildlife- Wilderness 
Inst.  Monogr.  1,  University  of  Montana,  Missoula. 
279pp. 


DERKSEN,  D.V.,  W.D.  ELDRIDGE,  and  M.W.  WELLER. 

1982.  Habitat  ecology  of  Pacific  black  brant  and  other 
geese  moulting  near  Teshekpuk  Lake,  Alaska.  Wild- 
fowl 33:39-57. 

DRISCOLL,  R.R.,  D.L.  MERKEL,  D.L.  RADLOFF,  D.E.  SNY- 
DER, and  J.S.  HAGIHARA.  1983.  An  ecological  land 
classification  framework  for  the  United  States.  U.S. 
Dep.  Agric,  For.  Serv.  Misc.  Publ.  1439. 

DUNBAR,  M.  1973.  Stability  and  fragility  in  arctic  ecosys- 
tems. Arctic.  26:179-185. 

EVERETT,  KR.  1982.  Some  recent  trends  in  the  physical 
and  chemical  characterization  and  mapping  of  tundra 
soils,  Arctic  Slope  of  Alaska.  Soil  Sci.  133:264-280. 

FYFE,  R.  and  R  OLENDORFF.  1976.  Minimizing  the  danger 
of  nesting  studies  to  raptors  and  other  sensitive  spe- 
cies. Can.  Wildl.  Serv.  Occ.  Publ.  23-  Ottawa,  Canada. 
17pp. 

GAARE,  E.  and  T.  SKOGLAND.  1975.  Wild  reindeer  food 
habits  and  range  use  at  Hardangervidda.  Pages  196-215 
in  Wielgolaski,  F.G.  ed.  Fennoscandian  Tundra  Ecosys- 
tems. Part  2.  Springer  Verlag,  Berlin. 

GABRIEL,  H.W.  and  S.S.  TALBOT.  1984.  Glossary  of  land- 
scape and  vegetation  ecology  for  Alaska.  BLM-Alaska 
Tech.  Rep.  10.  U.S.  Dep.  Inter.,  Bur.  of  Land  Manage. 
Anchorage,  AK  137pp. 

GALL,  L.J.  1984.  Population  structure  and  recommenda- 
tions for  conservation  of  the  narrowly  endemic  alpine 
butterfly  (Boloria  acronemia  [Lepidoptira:  Nymphali- 
dae]).  Biol.  Conserv.  28:111-138. 

GEORGE,  T.H.,  W.J.  STRINGER,  J.E.  PRESTON,  W.R.  FI- 
BIEL,  and  PC.  SCORUP.  1977.  Reindeer  range  inven- 
tory in  western  Alaska  from  computer-aided  digital 
classification  of  LANDSAT  data.  28th  Alaska  Science 
Conf.  Anchorage,  AK 

GREIG-SMITH,  P.  1983.  Quantitative  plant  ecology.  3rd. 
ed.  Univ.  Calif.,  Berkeley.  359pp. 

HALL,  D.K,  J.P.  ORMSBY,  L.  JOHNSON,  and  J.  BROWN. 

1980.  Landsat  digital  analysis  of  the  initial  recovery  of 
burned  tundra  at  Kokolik  River,  Alaska.  Remote  Sen- 
sing of  Environment  10:263-272. 

HESJEDAL  O.  1974.  Terrestrial  vertebrates.  Pages  475-570 
in  Ives,  J.D.  and  R.G.  Barry  eds.  Arctic  and  Alpine 
Environments.  Methuen,  London. 

and  J.  LARSSON.  1975.  Remote  sensing  of  vegeta- 
tion in  conservation  of  tundra  landscapes.  Pages  237- 
242  in  Wielgolaski,  F.E.  ed.  Fennoscandian  Tundra 
Ecosystems;  Part  2,  Animals  and  Systems  Analysis. 

HOFFMANN,  R.S.  1974.  Terrestrial  vertebrates.  Pages  475- 
570  in  Ives,  J.D.  and  R.G  Barry,  eds.  Arctic  and  Al- 
pine Environments.  Methven,  London. 

1984.  Small  mammals  in  winter.  The  effects  of 

altitude,  latitude  and  geographic  history.  Pages  9-23 
in  Merritt,  J.E.  ed.  Winter  Ecology  of  Small  Mammals. 
Carnegie  Mus.  Spec.  Publ.  10. 

HOK,  J.  R.  1969-  A  reconnaissance  of  tractor  trails  and 
related  phenomena  on  the  North  Slope  of  Alaska.  Un- 
publ.  U.S.  Dep.  Inter.,  Bur.  of  Land  Manage.  Anchor- 
age, AK. 

HOLT  S.  1971.  Some  effects  of  vehicle  operation  on  Alaska 
arctic  tundra.  M.S.  Thesis,  Univ.  of  Alaska,  Fairbanks. 
85pp. 

1980.  Vegetation  patterns  and  effects  of  grazing  on 

caribou  ranges  in  the  Sondre  Stromfford  area,  West 
Greenland.  Pages  57-63  in  Reimers,  E.,  E.  Gaare,  S. 
Skjenneberg  eds.  Proc.  2nd  Int.  Reindeer/Caribou 
Symp.  Roros,  Norway. 


Tundra 


165 


IRVING,  L.,  G.C.  WEST,  L.J.  PEYTON,  and  S.  PANEAK 

1967.  Migration  of  willow  ptarmigan  in  arctic  Alaska. 
Arctic  20:77-85. 

IVES,  J. D.  1970.  Arctic  tundra:  how  fragile?  A  geomorphol- 
ogist's  point  of  view.  Pages  39-42  in  Hare,  E.K  ed. 
The  Tundra  Environment.  Royal  Soc.  Canada.  Univ. 
Toronto  Press,  Toronto. 

JENKINS,  J.F.  and  D.H.  ASHTON.  1970.  Productivity  stud- 
ies on  Macquarie  Island  vegetation.  Pages  851-863 
in  Holdgate,  M.W.  ed.  Antarctic  Ecol.,  Vol.  2.  Aca- 
demic Press,  London. 

KESSEL,  B.  and  G.B.  SCHALLER.  I960.  Birds  of  the  upper 
Sheenjek  Valley,  Northeastern  Alaska.  Biol.  Pap.,  Univ. 
Alaska,  Fairbanks.  4:1-59- 

KING,  J.G.  1970.  The  swans  and  geese  of  Alaska's  arctic 
slope.  Wildfowl.  21:11-17. 

KING,  R.  1979.  Results  of  serial  surveys  of  migratory  birds 
on  NPR-A  in  1977  and  1978.  Pages  187-226  in  Lent, 
P.C.  ed.  Studies  of  Selected  Wildlife  and  Fish  and  their 
Use  of  Habitats  on  and  Adjacent  to  the  National  Pe- 
troleum Reserve — Alaska — 1977-1978.  Field  Study 
Rep.  3-  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  Anchor- 
age, AK 

KLEIN,  G.J.,  DC.  Pearce,  and  L.W.  Gold.  1950.  Method  of 
measuring  the  significant  characteristics  of  a  snow 
cover.  Nat.  Res.  Coun.  Can.  Tech.  Mem.  18.  56pp. 

LaPERRIERE,  A.J.  1978.  Use  of  Landsat  imagery  for  wild- 
life habitat  mapping  in  northeastern  and  east-central 
Alaska.  Final  Report.  NASA  Contract  NAS520915. 
NTIS.  Springfield,  VA.  36pp. 

and  P.C.  LENT.  1977.  Caribou  feeding  sites  in  rela- 
tion to  snow  characteristics  in  northeastern  Alaska. 
Arctic.  30:101-108. 

LENT,  P.C.  1980.  Synoptic  snowmelt  patterns  in  arctic 
Alaska  in  relation  to  caribou  habitat  use.  Pages  71-76 
in  Reimers,  E.,  E.  Gaare,  S.  Skjenneberg  eds.  Proc. 
Second  Int.  Reindeer/Caribou  Symp.  Roros,  Norway. 
Direktoratet  for  Vilt  of  Verskvannsfisk,  Trondheim. 

1985.  Cold  region  vegetation  information  needs 

from  the  perspective  of  wildlife  and  fisheries.  Pages 
20-27  in  LaBau,  V.J.  and  CL.  Kerr  eds.  Inventorying 
Forest  and  Other  Vegetation  of  the  High  Latitude  and 
High  Altitude  Regions.  Proc.  Int.  Symp.  Am.  For.  Reg. 
Conf,  Soc.  Am.  For.  Washington,  DC. 

and  D.  KNUTSEN.  1 971.  Musk-ox  and  snow  cover 

on  Nunivak  Island,  Alaska.  Pages  50-62  in  Haugen,  A. 
ed.  Proc.  Snow  and  Ice  in  Relation  to  Wildlife  and 
Recreation  Symp.  Iowa  State  Univ.,  Ames. 

and  A.J.  LaPERRIERE.  1974.  Applications  of  ERTS 


imagery  to  the  study  of  caribou  movements  and  win- 
ter habitat.  Final  Rep.  NASA  Contract  NAS521833. 
NTIS.  Springfield,  VA.  50pp. 

LEWIS,  ME.  1970.  Alpine  rangelands  of  the  Uinta  Moun- 
tains, Ashley  and  Wasatch  National  Forests.  U.S.  Dep. 
Agric,  For.  Serv.,  Reg.  4.  Ogden,  UT  75pp. 

LINELL,  K.A.  and  J.C.F.  TEDROW.  1981.  Soil  and  permafrost 
surveys  in  the  Arctic.  Clarendon  Press,  Oxford. 
279pp. 

MCKENDRICK,  J.D.  1981.  Responses  of  arctic  tundra  to 
intensive  musk-ox  grazing.  Agroborealis  13:49-55. 

MILLER,  F.L.,  E.J.  EDMONDS,  and  A.  GUNN.  1982.  Forag- 
ing behaviour  of  Peary  caribou  in  response  to  spring- 
time snow  and  ice  conditions.  Canadian  Wildl.  Serv. 
Occ.  Pap.  48.  41pp. 

MOITORET,  C.S.,  PA.  MILLER,  R.M.  OATES,  and  MA. 
MASTELLER.  1985.  Terrestrial  bird  populations  and 
habitat  use  on  coastal  plain  tundra  of  the  Arctic  Na- 


tional Wildlife  Refuge.  Pages  362-446  in  1984  Update 
Report,  Arctic  National  Wildlife  Refuge  Coastal  Plain 
Assessment.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv. 
Anchorage,  AK 

NODLER,  FA.,  A.J.  LaPERRIERE,  and  DR.  KLEIN.  1978. 
Vegetation  type  mapping  in  northwestern  Alaska  in 
relation  to  caribou  and  reindeer  range  potentials. 
Alaska  Coop.  Wildlife  Research  Unit  Spec.  Rep.  2. 
Univ.  of  Alaska,  Fairbanks.  33pp. 

NPR-A  (National  Petroleum  Reserve — Alaska)  Task  Force. 
1978.  Fish  and  Wildlife  resources.  Values  and  Re- 
source Analysis.  Section  6,  Vol.  3,  Study  Rep.  2.  U.S. 
Dep.  Inter.  224pp. 

1979.  Ecological  Profile.  National  Petroleum  Re- 
serve in  Alaska.  Study  Rep.  4.  U.S.  Dep.  Inter. 

OWENSBY,  C.E.  1973.  Modified  step-point  system  for 
botanical  composition  and  basal  cover  estimates.  J. 
Range  Manage.  26:302-303. 

PALMER,  L.J.  1926.  Progress  of  reindeer  grazing  investiga- 
tions in  Alaska.  U.S.  Dep.  Agric.  Bull.  1423-  37pp. 

PAMPLIN,  W.L  1979.  Construction-related  impacts  of  the 
Trans-Alaska  Pipeline  System  on  terrestrial  wildlife 
habitats.  Special  Report  24,  Joint  State  Federal  Fish 
and  Wildlife  Advisory  Team.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  Anchorage,  AK  1 32pp. 

PEGAU,  RE.  1968.  Reindeer  range  appraisal  in  Alaska. 
Arctic  21:255-259. 

1970.  Effect  of  reindeer  trampling  and  grazing  on 

lichens.  J.  Range  Manage.  23:95-97. 

PEWE,  T.  1983-  Alpine  permafrost  in  the  contiguous 

United  States:  A  review.  Arctic  and  Alpine  Research 

15:145-156. 
PITELKA,  F.A.  1979.  An  avi-faunal  review  for  the  Barrow 

region  and  North  Slope  of  Arctic  Alaska.  Arctic  and 

Alpine  Research  6:161-184. 
PRICE,  L.W.  1981.  Mountains  and  Man.  Univ.  of  Calif.  Press, 

Berkeley.  506pp. 
PRUITT,  W.O.,  Jr.  1959.  Snow  as  a  factor  in  the  winter 

ecology  of  barren-ground  caribou.  Arctic.  12:159-179- 

1 966.  Ecology  of  terrestrial  mammals.  Pages  519- 

564  in  Environment  of  the  Cape  Thompson  Region, 
Northwest  Alaska.  U.S.  Atomic  Energy  Commission. 
U.S.  Govt.  Print.  Off. 

1978.  Boreal  Ecology.  Edward  Arnold  Publ.  Ltd. 


London,  iv  +  73pp. 

RACINE,  C.H.  1979.  The  1977  tundra  fires  in  the  Seward 
Peninsula,  Alaska:  effects  and  initial  revegetation. 
BLM-Alaska  Technical  Report  4.  U.S.  Dep.  Inter.,  Bur. 
Land  Manage.  Anchorage,  AK  51pp. 

REEVES,  H.M.,  F.G.  COACH,  AND  RE.  MUNRO.  1976. 
Monitoring  arctic  habitat  and  goose  production  by 
satellite  imagery.  J.  Wildl.  Manage.  40:532-541. 

REYNOLDS,  H.V.,  J.  CURATOLO,  and  R.  QUIMBY.  1976. 
Denning  ecology  of  grizzly  bears  in  northeastern 
Alaska.  Pages  403-409  in  Pelton,  M.J.,  J.W.,  Lentfer, 
and  E.  Folk  eds.  Bears  -  their  biology  and  Manage- 
ment. IUCN  New  Series  40.  Morges,  Switzerland. 

RICHENS,  V.B.  and  C.G.  MADDEN.  1973.  An  improved 
snow  study  kit.  J.  Wildl.  Manage.  37:109-113. 

RITCHIE,  R.  1979.  A  survey  of  cliff-nesting  raptors  and 
their  habitats.  Pages  313-336  in  Lent,  P.C.  ed.  Studies 
of  Selected  Wildlife  and  Fish  Populations  and  their 
Use  of  Habitats  on  and  Adjacent  to  NPR-A.  1977- 
1978.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  Anchor- 
age, AK 

SCHMIDT,  WD.  1984.  Materials  and  methods  of  subnivean 
sampling.  Pages  25-32  in  Merritt,  J.E.  ed.  Winter  Ecol- 


166 


Tundra 


ogy  of  Small  Mammals.  Carnegie  Mus.  Nat.  Hist.  Spec. 
Publ.  10. 

SIMS,  R.A.  1983-  Ground-truth  and  large  scale  70  mm  aerial 
photographs  in  the  study  of  reindeer  winter  range- 
land,  Tuktoyaktuk  Peninsula  area,  NWT.  PhD.  disserta- 
tion. Univ.  of  British  Columbia,  Vancouver.  1 78pp. 

SKOGLAND,  T.  1978.  Characteristics  of  the  snow  cover 
and  its  relationship  to  wild  mountain  reindeer  (Ran- 
gifer  tarandus  tarandiis  L. )  feeding  strategies.  Arctic 
and  Alpine  Research  10:569-580. 

SKUNCKE,  F.  1969.  Reindeer  ecology  and  management  in 
Sweden.  Biol.  Paper.  Univ.  Alaska,  Fairbanks.  82pp. 

SMITH,  R.H.  AND  R.P.  ALLEN.  1948.  An  aerial  waterfowl 
reconnaissance  of  the  far  north.  U.S.  Dep.  Inter.,  Fish 
and  Wildl.  Serv.  Spec.  Rep.  60:5-12. 

TENER,  J.S.  1965.  Musk-oxen  in  Canada.  Can.  Wildl.  Serv. 
Monogr.  Ser.  2.  Dept.  of  Northern  Affairs  and  Nat. 
Resources.  Ottawa.  166pp. 

THILENIUS,  J.F.  1979.  Range  management  in  the  alpine 
zone.  Pages  43-64  in  Johnson,  D.A.  ed.  Special  Man- 
agement Needs  of  Alpine  Ecosystems.  Range  Science 
Serv.  Society  for  Range  Management,  Denver,  CO. 
5:43-64. 

THOMAS,  DC,  F.L.  MILLER,  R.H.  RUSSELL,  and  G.R.  PAR- 
KER. 1981.  The  Bailey  Point  region  and  other  musk- 
ox  refugia  in  the  Canadian  Arctic:  A  short  review. 
Arctic.  34:34-36. 

THORN,  C.E.  1982.  Gopher  disturbance:  Its  variability  by 
Braun-Blanquet  vegetation  units  in  the  Niwot  Ridge 
alpine  tundra  zone,  Colorado  Front  Range,  U.S.A. 
Arctic  and  Alpine  Research  14:45-51. 

U.S.  FISH  AND  WILDLIFE  SERVICE.  1982  Initial  report 
Baseline  study  of  the  fish,  wildlife  and  their  habitats. 
Arctic  National  Wildlife  Refuge  coastal  plain  resource 
assessment.  U.S.  Fish  and  Wildl.  Serv.,  Anchorage, 
AK.  507pp. 

.  1983-1984.  Update  report.  Baseline  study  of  the 

fish,  wildlife  and  their  habitats.  Arctic  National  Wild- 
life Refuge  coastal  plain  resource  assessment.  U.S. 
Dep.  Inter.,  Fish  and  Wildl.  Serv.  Anchorage,  Alaska. 


613pp. 

VAN  CLEVE,  K  1977.  Recovery  of  disturbed  tundra  and 
taiga  surfaces  in  Alaska.  Pages  422-455  in  Cairns, 
J.,  KL.  Dickson,  and  E.  Herricks  eds.  Proc.  of  an  Int. 
Symp.  on  the  Recovery  of  Damaged  Ecosystems.  Vir- 
ginia Polytechnic  Inst,  and  Univ.  Blacksburg,  VA. 

VIBE,  C.  1967.  Arctic  animals  in  relation  to  climatic  fluc- 
tuations. Meddelelser  om  Gronland  170(5):l-227. 

VIERECK,  LA.  and  C.T.  DYRNESS.  1980.  A  preliminary 

classification  system  for  vegetation  of  Alaska.  U.S.  Dep. 
of  Agric,  For.  Serv.  General  Tech.  Rep.,  PNW-106. 

,  C.T.  DYRNESS,  and  A.R.  BATTEN.  1982.  Revision 

of  preliminary  classification  for  vegetation  of  Alaska. 
Inst.  North.  Forestry,  Univ.  Alaska,  Fairbanks.  72pp. 

WALKER,  DA.,  K.R.  EVERETT,  P.J.  WEBBER,  and  J. 

BROWN.  1980.  Geo-botanical  atlas  of  the  Prudhoe 
Bay  Region,  Alaska.  CRREL  Report  80-14.  Cold  Re- 
gions Research  and  Engineering  Laboratory.  Hanover, 
NH.  69pp. 

,  P.J.  WEBBER,  KR.  EVERETT,  and  J.  BROWN.  1978. 

Effects  of  crude  and  diesel  oil  spills  on  plant  commu- 
nities at  Prudhoe  Bay,  Alaska,  and  the  derivation  of  oil 
spill  sensitivity  maps.  Arctic  31:242-258. 

WIELGOLASKI,  F.E.  1975.  Productivity  of  tundra  ecosys- 
tems. Pages  1-12  in  Productivity  of  World  Ecosys- 
tems. Proc.  of  a  Symposium.  National  Academy  of 
Sciences.  Washington,  DC. 

WILLARD,  BE.  and  W.  MARR.  1971.  Recovery  of  alpine 
tundra  under  protection  after  damage  by  human 
activities  in  the  Rocky  Mountains  of  Colorado.  Biol. 
Cons.  3:181-190. 

WILLIAMSON,  F.S.L.,  M.C.  THOMPSON,  and  J.Q.  HINES. 
1966.  Avi-faunal  investigations.  Pages  437-480  in  Wili- 
movsky,  N.J.  and  J.N.  Wolfe  eds.  Environment  of  the 
Cape  Thompson  Region,  Alaska.  U.S.  Atomic  Energy 
Comm.  Washington,  DC. 

ZWINGER,  A.H.  and  BE.  WILLARD.  1972.  Land  above  the 
trees:  a  guide  to  American  alpine  tundra.  Harper 
and  Row,  New  York,  NY.  489pp. 


Tundra 


167 


RIPARIAN 
HABITAT 

Robert  D.  Ohmart  and  Bertin  W.  Anderson 


Center  for  Environmental  Studies 
Arizona  State  University 
Phoenix,  AZ  85281 


Editor's  Note:  This  chapter  is  the  first  of  two  on 
wetland  habitats.  These  habitats  are  extremely  im- 
portant, not  only  because  of  their  high  inherent 
wildlife  value,  but  also  because  of  their  effects  on 
the  adjacent  upland  and  aquatic  areas  and  their 
associated  biota  This  chapter  covers  wetland  areas 
associated  with  running  water,  while  the  following 
chapter  on  marshes  covers  areas  of  standing  water. 
These  are  not  universally  accepted  definitions  but 
provide  a  convenient  breakdown  for  this  book.  The 
length  of  this  chapter  reflects  the  high  importance 
of  and  current  interest  in  riparian  areas.  However, 
much  still  needs  to  be  learned  about  these  areas, 
and  a  critical  need  exists  for  better  management  of 
them.  Inventorying  and  monitoring  riparian  habi- 
tats will  be  a  central  part  of  such  efforts. 


INTRODUCTION 

One  of  the  most  important  assignments  in  the 
career  of  a  wildlife  biologist  is  to  monitor  or  inven- 
tory riparian  ecosystems.  Only  a  decade  ago  few 
people,  including  wildlife  biologists,  had  any  appreci- 
ation or  knowledge  of  these  very  limited  and  highly 
valuable  wildlife  habitats.  Even  today,  books  on  wild- 
life habitats  or  plant  communities  seldom  separate 
riparian  ecosystems  from  adjacent  upland  plant  com- 
munities. Before  undertaking  an  assignment,  we 
recommend  obtaining  copies  of  four  riparian  sym- 
posiums (Johnson  and  Jones  [1977];  Johnson  and 
McCormick  [1978];  Warner  and  Hendrix  [1984];  and 
Johnson  et  al.  [1985])  to  use  as  references. 

Riparian  as  an  adjective  is  defined  as  "relating  to 
or  living  or  located  on  the  bank  of  a  natural  water- 
course (as  a  river)  or  sometimes  of  a  lake  or  a  tide- 
water" (Webster's  New  Collegiate  Dictionary 
1979:991 ).  To  many,  riparian  is  synonymous  with 
wetland,  but  wetland  is  often  defined  as  consisting 
primarily  of  emergent  or  marsh  communities,  which 
are  not  discussed  in  this  chapter.  (See  Chapters  10, 
1 1 ,  and  1 2  for  treatment  of  marshes,  streams,  and 
lakes,  respectively.)  For  this  chapter,  which  ad- 
dresses only  terrestrial  riparian  ecosystems,  we  will 
adhere  to  the  following  definition:  "A  riparian  associ- 
ation of  any  kind  [excluding  marshes]  is  one  which 
is  in  or  adjacent  to  drainageways  and/or  their  flood- 
plains  and  which  is  further  characterized  by  species 
and/or  life-forms  different  than  that  of  the  immedi- 
ately surrounding  non-riparian  climax"  (Lowe 
1964:62).  This  definition  includes  plant  communities 
along  permanently  flowing  or  intermittent  drainages. 
Some  of  these  drainages  may  not  flow  for  many  years 
or  even  in  our  lifetime,  but  they  are  riparian  commu- 
nities if  the  plant  species  along  these  drainages  are 
different  from  those  of  the  adjacent  upland. 

Obligate  or  riparian-dependent  species  such  as 
cottonwoods  (Populus  sp. )  and  willows  (Salix  sp. ) 


Riparian  Habitats 


169 


are  frequently  referred  to  in  the  literature  as  phrea- 
tophytes,  referring  to  vegetation  species  having  their 
roots  in  perennial  ground  water  or  in  the  capillary 
fringe  above  a  water  table.  Most  of  these  species 
transpire  large  quantities  of  water,  and  water  man- 
agers believe  that  if  the  streamside  vegetation  is 
removed,  this  water  will  be  saved  or  remain  in  the 
aquifer. 

In  past  water  management  practices,  thousands 
of  acres  of  riparian  vegetation  have  been  removed  to 
prevent  this  wicking  of  water  into  the  atmosphere. 
For  example,  between  1967  and  1971,  about  21,600 
ha  (54,000  a.)  of  riparian  vegetation  was  removed 
along  the  Pecos  River  in  New  Mexico  to  save  water. 
Although  about  18,800  ha  (47,000  a.)  of  the  origi- 
nally cleared  area  has  remained  cleared  since  1971 
(U.S.  Department  of  Interior,  Bureau  of  Reclamation 
1979),  preliminary  results  indicate  that  the  amount 
of  water  saved  is  probably  insignificant.  Riparian 
ecosystems  do  not  stand  alone;  they  are  fed  by  wa- 
tersheds, which  when  destroyed,  also  destroy  the 
riparian  ecosystem. 

Nutrients,  water,  and  detrital  materials  are  trans- 
ported into  the  riparian  system  from  its  watershed. 


A  healthy  watershed  generally  indicates  a  healthy 
riparian  system.  Degraded  watersheds  produce  high 
surface  runoff  carrying  valuable  soil  into  the  stream, 
which  reduces  productivity  of  both  the  aquatic  and 
terrestrial  portions  of  the  system.  As  any  competent 
hydrologist  knows,  to  reduce  the  volume  and  sever- 
ity of  floods,  one  must  start  with  the  point  source — 
the  watersheds;  other  efforts  are  "treating  symptoms 
and  not  the  disease." 

The  importance  of  western  riparian  habitats  to 
humans  has  long  been  recognized,  as  indicated  by 
the  early  settlement  patterns  of  native  North  Ameri- 
cans and  Europeans.  During  drought  periods,  settle- 
ments that  were  not  near  permanent  water  sources 
were  forced  to  relocate;  many  died  in  the  process. 
Riparian  habitats  provided  and  still  provide  water, 
rich  fertile  soils  for  agriculture,  lush  forage  for  do- 
mestic livestock,  recreation,  and  home  sites.  Their 
importance  is  amplified  in  the  arid  western  states, 
but  is  also  obvious  in  the  East  where  riparian  areas 
are  termed  bottomland  hardwood  habitats. 

Riparian  habitats  in  the  West  are  very  limited 
when  compared  to  the  amount  of  acreage  they  con- 
stitute versus  upland  habitats.  Riparian  ecosystems 


Degraded  riparian  area. 


170 


Riparian  Habitats 


when  compared  with  upland  habitats  may  total  up  to 
0.5%  of  the  landscape  or  <  0.1%.  Because  of  the 
small  and  finite  nature  of  riparian  habitats,  their  vital- 
ness  to  human  survival  in  arid  environments,  their 
recreational  values,  and  their  high  fishery  and  wild- 
life values,  they  should  receive  critical  concern  in  all 
land-planning  and  management  efforts.  For  example, 
Johnson  (1978)  reported  that  64  wildlife  species 
presently  listed  as  endangered,  and  47  more  species 
being  considered  for  listing,  are  dependent  on  ripar- 
ian habitats.  In  the  past,  riparian  habitats  have  been 
treated  as  sewage  transport  systems  and  refuse  land- 
fill sites  and  have  been  subjected  to  numerous  other 
types  of  habitat  degradation. 

The  treatment  of  riparian  habitats  in  the  past 
and  their  current  condition  is  alarming,  which  ampli- 
fies the  need  to  pay  special  management  concern  to 
these  ecosystems.  For  example,  estimates  are  that 
70-90%  of  the  natural  riparian  ecosystems  in  the  U.S. 
have  been  lost  to  human  activities  (U.S.  Council  on 
Environmental  Quality  1978;  Warner  1979a).  Re- 
gional losses  in  these  ecosystems  have  been  esti- 
mated to  exceed  98%  in  the  Sacramento  Valley  in 
California  (Smith  1977)  and  95%  in  Arizona  (Warner 
1979b).  Johnson  and  Carothers  (1981)  estimated 


that  in  the  Rocky  Mountains/Great  Plains  region,  90- 
95%  of  the  cottonwood-willow  riparian  ecosystems 
of  the  plains  and  lower  foothills  have  been  lost.  Pos- 
sibly as  much  as  80%  of  the  remaining  riparian  eco- 
systems in  the  U.S.  (both  privately  and  publicly 
owned)  are  in  unsatisfactory  condition  and  are  domi- 
nated by  human  activities  (Almand  and  Krohn  1978; 
Warner  1979b).  In  the  West,  these  ecosystems  sup- 
port a  disproportionately  greater  number  of  wildlife 
species  than  their  upland  counterparts. 

Public  support  for  wildlife  and  its  habitat  needs 
must  be  interpreted  with  caution.  Kellert  (1980), 
reporting  on  a  nationwide  opinion  poll  that  focused 
on  land-use  allocations  for  wildlife,  showed  that 
many  Americans  were  willing  to  make  economic 
sacrifices  in  commodity  resource  production  for  en- 
dangered nongame  species  and  certain  featured  big 
game  species.  Sixty  percent  agreed  that  livestock 
grazing  on  public  lands  should  be  limited  if  it  de- 
stroyed vegetation  used  by  wildlife,  even  if  it  re- 
sulted in  higher  meat  prices;  34%  disagreed.  Over 
75%  believed  that  logging  should  be  done  in  a  man- 
ner to  enhance  wildlife  even  if  lumber  and  paper 
prices  rose.  Conversely,  almost  50%  believed  that 
natural  resources  must  be  developed,  even  if  it 


Same  riparian  area,  1 0-years  later. 


Riparian  Habitats 


171 


meant  less  wilderness  and  lowered  wildlife  popula- 
tions. These  results  indicate  public  awareness  for  the 
needs  of  wildlife  and  management  changes  that  need 
to  be  enacted  on  public  lands  to  maximize  natural 
resource  values.  The  public  at  large  generally  sup- 
ports the  concept  of  multiple-use  management  on 
public  lands,  but  interest  groups  strongly  disagree  on 
specific  issues. 


Sports  enthusiasts  in  eastern  Oregon  favorably 
responded  to  improved  grazing  management  prac- 
tices (Megank  and  Gibbs  1979).  Almost  70%  of 
anglers  surveyed  stated  that  their  recreation  experi- 
ences would  be  reduced  by  management  practices 
that  further  degraded  riparian  systems.  Their  reac- 
tion to  fences  was  positive  in  that  it  represented 
better  livestock  control.  Hunters  thought  that  man- 
agement that  improved  forage  production  for  live- 
stock would  also  help  deer  and  elk. 


Many  values  of  riparian  habitats  to  our  society, 
seldom  considered  by  terrestrial  wildlife  biologists, 
should  be  factored  in  as  sound  arguments  toward 
managing  these  systems  in  their  natural  state.  The  ri- 
parian vegetation  adjacent  to  streams  or  even  large 
rivers  is  extremely  important  as  an  energy  source  to 
the  aquatic  organisms  (see  Streams,  Chapter  1 1 ).  In 
small  headwater  streams,  99%  of  the  energy  for  het- 
erotrophic organisms  comes  from  the  vegetation 
along  the  stream,  whereas  only  1%  comes  from  pho- 
tosynthesizing  autotrophs  (Cummins  1974).  In  large 
river  systems,  such  as  the  Missouri  River,  as  much 
as  54%  of  the  organic  matter  consumed  by  fish  is  of 
terrestrial  origin  (Benner  in  Kennedy  1977).  A  fac- 
tor that  has  not,  to  our  knowledge,  been  seriously 
considered  by  fishery  biologists  is  the  quality  of  the 
organic  input  or  species  composition  of  the  stream- 
side  vegetation.  Trees  may  be  desirable  as  shade 
to  prevent  large  fluctuations  in  water  temperature, 
but  some  tree  species  will  ultimately  prove  more 
important  as  energy  input  sources  than  others,  e.g., 
cottonwood  and  willow  leaves  and  other  tree  parts 
are  probably  of  more  value  to  aquatic  detritivores 
than  leaves  and  other  parts  of  the  exotic  salt  cedar 
(Tamarix  chinensis).  The  former  would  certainly 
impose  less  change  on  total  dissolved  solids  and 
water  chemistry  than  the  latter.  Biologists  must  man- 
age for  healthy  streamside  vegetation  and,  as  knowl- 
edge progresses,  some  efforts  should  be  directed 
toward  encouraging  establishment  of  tree  species 
that  have  higher  nutrient  input  value  to  the  aquatic 
fauna.  This  approach  toward  managing  terrestrial 
vertebrates  is  already  underway  along  large  river 
systems. 


Streamside  vegetation  is  very  important  in  deter- 
mining the  structure  and  function  of  stream  ecosys- 
tems (Knight  and  Bottorff  1984).  Mahoney  and 
Erman  (1984)  found  that  riparian  vegetation  is  an 


important  source  of  food  to  stream  organisms,  pro- 
vides shade  over  small-order  streams,  and  serves 
to  stabilize  banks  in  preventing  excessive  sedimenta- 
tion and  intercepting  pollutants.  Asmussen  et  al. 
(1977)  reported  that  vegetation  buffer  strips  were 
very  effective  in  reducing  pollution  from  agricultural 
chemicals.  Karr  and  Schlosser  (1977,  1978)  reported 
that  proper  management  of  streamside  vegetation 
and  the  channel  may  substantially  improve  water 
quality  in  agricultural  watersheds.  Corbett  and  Lynch 
(  1985)  stressed  the  importance  of  streamside  zones 
in  water-quality  management  for  municipal  water 
supplies.  Haupt  (1959)  presented  guidelines  for 
buffer  strip  widths  in  road-building  projects,  and 
Benoit  (1978)  similarly  presented  guidelines  for  tim- 
ber harvest  operations  in  Oregon.  Treating  water  to 
bring  it  to  potable  standards  is  expensive;  stopping 
pollution  at  its  source  and  managing  for  productive 
riparian  vegetation  will  significantly  reduce  these 
costs.  Finally,  riparian  vegetation  can  be  important  in 
flood-control  efforts  (Chaimson  1984)  by  reducing 
water  velocity  and  its  erosive  energy  during  flood 
stage  (Li  and  Shen  1973).  The  vegetation  may  also 
reduce  streambank  damage  from  ice,  log  debris,  and 
animal  trampling  (Platts  1979;  Swanson  et  al.  1982); 
armor  levees;  and  prevent  channel  changes  during 
high  flows. 


Importance  of  riparian  systems  to  wildlife  has 
not  been  quantified  or  demonstrated  to  any  convinc- 
ing extent  until  the  past  1 5  years.  The  efforts  of 
Carothers  et  al.  (1974)  to  quantify  avian  densities  in 
cottonwood  habitats  along  the  Verde  River  in  central 
Arizona  and  those  of  Ohmart  and  Anderson  ( 1974) 
along  the  lower  Colorado  River  were  beginnings. 
A  riparian  symposium  (Johnson  and  Jones  1977; 
Hehnke  and  Stone  1978;  Thomas  et  al.  1979)  along 
with  others  were  fruitful  in  focusing  attention  on 
these  ecosystems.  These  studies  and  subsequent  ones 
indicated  that  some  of  the  highest  densities  of  breed- 
ing birds  in  North  America  were  found  in  riparian 
habitats,  and  more  than  60%  of  the  vertebrates  in 
the  arid  Southwest  were  obligate  to  this  ecosystem 
(see  Ohmart  and  Anderson  1982  for  a  review).  An- 
other 10-20%  of  the  vertebrates  were  facultative 
users  (present  for  a  portion  of  the  annual  cycle  but 
not  fully  dependent  on  riparian  habitats)  of  stream- 
side  vegetation.  Mosconi  and  Hutto  (1982)  reported 
that  in  western  Montana,  59%  of  the  species  of  land 
birds  use  riparian  habitats  for  breeding,  and  36% 
breed  only  in  riparian  habitats.  Similarly,  of  the  363 
species  of  land  vertebrates  in  the  Great  Basin  of 
southeast  Oregon,  299  either  directly  depend  on 
riparian  habitats  or  utilize  them  more  than  any  other 
habitat  types  (Thomas  et  al.  1979).  Therefore,  if 
these  ecosystems  were  totally  lost  or  continued  to 
be  reduced  to  vestiges  of  their  original  state,  con- 
ceivably 60-80%  of  our  native  wildlife  species  could 
be  lost  in  the  western  U.S.  The  Colorado  River  is  a 
classic  example  of  this,  in  that  at  least  four  species  of 


172 


Riparian  Habitats 


birds  have  recently  been  extirpated  and  unless  some 
dramatic  management  changes  are  made,  another 
six  species  could  be  lost  in  the  next  20  years 
(Hunter  et  al.  unpubl.  data). 


RIPARIAN  HABITAT  CLASSIFICATION 

Small  riparian  ecosystems  occurring  in  wa- 
tersheds at  higher  elevations  eventually  connect  to 
form  major  drainage  systems.  These  major  drainages, 
such  as  the  Green  River  in  Wyoming,  connect  into 
larger  systems,  such  as  the  Colorado  River.  To  aid  in 
understanding  these  western  riparian  ecosystems  at 
different  elevations,  we  recommend  using  the  Life 
Zone  concept  developed  by  Merriam  (  1 890 ),  which 
is  widely  used  in  the  West  for  its  utilitarian  value. 

The  Brown  et  al.  (1979)  concept  for  naming 
riparian  ecosystems  in  Merriam's  Life  Zones  is  also 
widely  accepted  in  the  West  because  its  hierarchical 
system  for  North  American  biomes  has  an  evolution- 
ary and  genetically  based  approach.  It  is  also 
digitized,  which  makes  it  computer-compatible.  Im- 
portantly, it  is  concordant  with  the  Life  Zone 
concept. 

Succession 

In  highly  modified  or  managed  rivers,  such  as 
the  lower  Colorado,  Pecos,  and  Rio  Grande,  there  is 
little  evidence  of  classical  succession.  If  an  area  is 
burned  or  cleared,  it  tends  to  return  as  trees  or 
shrubs  and  remains  in  that  state.  Succession  is  poorly 
studied  and  understood,  and  it  may  be  that  riparian 
floodplains  are  so  rich  in  nutrients  and  water  that 
classical  plant  succession  does  not  occur.  Brady  et  al. 
(1985)  presented  the  developmental  continuum  of 
a  riparian  gallery  forest  ranging  from  a  nursery  bar  to 
a  mature  forest.  Bock  and  Bock  ( 1985)  presented 
data  on  patterns  of  reproduction  in  a  species  of  syca- 
more (Platanns  urightii)  in  southeastern  Arizona. 

Communities 

Naming  individual  plant  communities  varies 
depending  on  the  biologist  mapping  and  sampling 
the  communities,  but  is  generally  based  on  one  or 
more  dominant  species  (either  by  density  or  stat- 
ure). For  example,  a  cottonwood-willow  community 
may  only  have  1  cottonwood  for  every  10  willows  in 
a  stand,  but  because  of  the  size  and  presence  of  cot- 
tonwoods,  it  may  be  called  a  cottonwood-willow 
community.  In  another  example,  a  community  may 
be  called  honey  mesquite  (Prosopis  glandulosa) 
although  it  contains  mostly  shrubs.  What  a  commu- 
nity is  called  is  unimportant  as  long  as  the  commu- 
nity names  and  accompanying  written  descriptions 
can  be  interpreted  by  others.  Iaurenzi  et  al.  (1983) 


presented  a  habitat  classification  system  for  mixed 
broadleaf  riparian  forests  in  the  Upper  Sonoran  Life 
Zone. 


Structural  Types 

Some  riparian  ecosystems  may  be  extremely 
dynamic  through  time  whereas  others  are  relatively 
stable.  Recognizing  and  classifying  structural  stages 
(young  through  mature  stages)  allows  a  quick  assess- 
ment of  the  riparian  ecosystem's  health.  If  no  young 
plant  communities  develop  to  replace  the  old,  deca- 
dent communities  of  similar  species  composition, 
then  animal  species  dependent  on  those  communi- 
ties may  be  lost  to  the  fauna.  Also,  young  plant  com- 
munities may  support  a  fauna  much  different  from 
that  found  in  a  mature  plant  community.  These 
structural  differences  (in  plant  community  age 
classes  and  foliage  layers)  are  very  important  in  man- 
aging for  maximum  riparian  productivity  and  verte- 
brate species  richness. 

The  concept  of  structural  classification  of  ripar- 
ian communities  is  an  important  one  and  not  difficult 
to  understand  if  an  area  is  envisioned  as  going  from 
bare  soil  to  supporting  a  mature  cottonwood-willow 
forest;  all  the  structural  types  would  then  be  present 
over  this  continuum.  Figure  1  shows  the  structural 
types  in  lower  elevational  systems;  Type  VI  is  the 
young  or  beginning  community,  and  as  it  grows  it 
passes  to  a  V,  IV,  etc.,  until  it  becomes  a  Type  I 
which  is  a  mature  cottonwood-willow  community.  In 
Type  VI,  most  of  the  foliage  volume  is  in  the  grass 
and  shrub  layers;  as  the  community  matures,  a  good 
overstory  (Type  II)  shades  out  much  of  the  under- 
story.  Type  I  contains  overstory,  midstory,  and  un- 
derstory  as  some  trees  die  opening  the  overstory  for 
ingression  of  shorter  trees,  shrubs,  and  annuals.  By 
dividing  a  continuous  process  into  stages  or  types, 
both  plant  communities  and  structural  types  can  be 
assessed  together  in  demonstrating  user-oriented 
impacts.  However,  structural  complexity  and  mean 
canopy  height  are  generally  reduced  where  riparian 
systems  are  under  heavy  water  management,  live- 
stock grazing,  pollution,  and  recreational  activities. 

In  the  following  example  of  structural  types,  a  cot- 
tonwood-willow community  was  used  to  demon- 
strate Types  VI  (young)  through  I  (mature).  At 
higher  elevations  this  community  could  consist  of 
sycamores,  narrow-leaf  cottonwoods  (Popnlns  an- 
gustifolia),  quaking  aspen  (P.  tremuloides),  or  other 
tree  species  that  have  similar  vertical  and  foliage 
density  attributes  as  cottonwood-willow.  Not  all  tree 
species  form  communities  that  have  the  same 
growth  pattern  as  those  discussed  above,  but  their 
community  development  shows  some  of  the  same 
structural  stages.  For  example,  salt  cedar  communi- 
ties generally  reach  maturity  as  structural  Type  II, 


Riparian  Habitats 


173 


1   J         v 

/(^     [A 

I  $  I  *«JL 

J^j^ 

j^^yinkj 

W^\ 

\  / ^ 

is-A 

*\r\ 

j\~  > 

~fir 

JqL 

L             •» 

.vKj      ^~ 

JU,/,-^ 

VI 


$$*% 


Figure  1.     Vegetation  structural  types  found  in  lower  elevations. 


honey  mesquite  as  Type  III,  and  arrowweed  (Tes- 
saria  sericea)  and  other  shrub  communities  are  al- 
ways Type  VI.  Managers  may  find  this  classification 
system  less  confusing  than  trying  to  establish  a  more 
complex  and,  hence,  more  complicated  approach 
to  recognizing  structural  types.  Most  importantly,  the 
structural  types  are  real  to  wildlife  and  its  use  of  the 
vegetation. 

In  habitat  discussions  today  of  the  Colorado 
River,  managers  have  wanted  to  know  the  commu- 
nity and  structural  type  under  consideration.  They 
can  easily  visualize  the  community  and  structural 
types,  and  we  have  provided  a  document  that  gives 
species  richness,  densities,  and  other  wildlife  use 
values  seasonally  for  3  to  5  years  along  with  mean 
values.  These  data  are  available  for  only  a  few  ripar- 
ian ecosystems,  and  our  approach  in  this  chapter 
is  to  help  managers  initiate  development  in  their  re- 
source management  areas.  The  system  can  always 
be  expanded  and  improved  toward  better  managing 
these  invaluable  wildlife  habitats. 


IMPORTANT  HABITAT  ATTRIBUTES  TO 
WILDLIFE 

Discerning  habitat  attributes  that  are  most  im- 
portant to  wildlife  is  not  an  easy  task,  and  reliable 


identification  of  those  habitat  components  only 
comes  from  long-term,  in-depth  studies.  Seasonal  and 
annual  variances  of  wildlife  numbers  are  generally 
very  high,  and  until  a  number  of  seasons  and  years 
have  been  studied,  the  habitat  components  that  are 
truly  important  to  wildlife  may  not  be  discernible  or 
fully  understood. 

Four  habitat  components,  not  discussed  in  depth 
in  this  section  but  important  to  wildlife,  are  plant 
community  size  (number  of  acres),  continuity  of 
riparian  habitat  along  the  streambed,  edge  or  eco- 
tone,  and  water.  The  streambed  has  been  widely 
discussed  in  island  biogeography  theory  and  in  terms 
of  habitat  fracturing.  The  more  extensive  stands  of  a 
riparian  forest  are  reduced,  the  less  wildlife  value 
they  have  for  some  species.  As  the  tract  gets  smaller, 
its  wildlife  values  diminish  to  the  point  that  the  for- 
est contributes  little,  if  any,  to  overall  wildlife  values. 
It  probably  reaches  its  zenith  of  importance  in  ripar- 
ian habitats  at  lower  elevations  where  expansive 
alluvial  floodplains  support  extensive  stands  of  con- 
tinuous wildlife  habitat.  T.E.  Martin  (Arizona  State 
University,  unpubl.  ms)  identified  area  size  as  the  pri- 
mary factor  accounting  for  variation  in  number  of 
breeding  pairs  of  birds  in  five  of  seven  of  his  eco- 
logical groups  in  high-elevation,  riparian  habitats  in 
central  Arizona.  As  knowledge  expands  and  under- 
standing of  the  needs  of  wildlife  increases,  we  ob- 


174 


Riparian  Habitats 


serve  that  continuous  areas  of  some  particular 
riparian  community  are  more  than  twice  the  value  of 
that  same  community  in  0.8-  to  4-ha  (2-  to  10-a.) 
blocks.  For  example,  many  bird  species  have  been 
lost  along  the  Colorado  River  and  others  will  be  lost 
in  the  near  future  as  island  size  stands  of  cotton- 
wood-willow  are  reduced.  Stands  of  28  ha  (70  a.) 
only  begin  to  fill  the  needs  of  some  bird  species, 
allowing  the  presence  of  a  few  breeding  pairs  at 
best.  For  these  reasons,  habitat  island  size  is  impor- 
tant regardless  of  whether  it  is  along  small  streams  at 
high  elevations  or  along  rivers  in  the  desert. 


An  example  of  continuous  riparian  habitat. 


Continuity  of  riparian  vegetation  along  the 
floodplain  is  extremely  important  to  some  species 
such  as  small  mammals,  reptiles,  and  amphibians. 
Small,  discontinuous  blocks  may  not  fulfill  the  needs 
of  many  of  these  species,  causing  reductions  in  num- 
bers and,  possibly,  extirpation  of  some  populations 
attempting  to  use  highly  fractured  riparian  habitats 
as  movement  corridors.  Amphibians  and  reptiles  that 
need  riparian  habitats  are  extremely  vulnerable  to 
this  fracturing  and,  as  pointed  out  by  Brode  and 
Bury  (1984),  what  were  once  continuous  populations 
of  some  species  in  California  are  now  isolated  relict 
populations.  Dispersal  routes  for  pioneering  individu- 
als and  gene  exchange  have  essentially  been  halted 
because  once  continuous  riparian  habitats  have  been 
disrupted. 

The  edge  or  ecotone  component  may  be  of  less 
concern  because  the  very  nature  of  riparian  ecosys- 
tems or  ribbons  of  habitat  running  through  uplands 
is  edge  at  its  maximum.  Some  believe  that  riparian 
ecosystems  are  all  edge,  regardless  of  alluvial  flood- 
plain  width,  because  the  river  constantly  meanders 
across  the  first  terrace,  scouring  old  plant  communi- 
ties and  depositing  new  soils  that  are  revegetated 


with  early  stage  plant  communities.  Continuing  this 
productive,  valuable  wildlife  habitat  as  edge  for  wild- 
life should  be  the  most  important  factor  in  good 
management  decisions. 

Water  may  or  may  not  be  present  aboveground 
in  riparian  ecosystems.  When  present,  it  is  an  impor- 
tant component  for  large  mammals  and,  to  a  lesser 
degree,  small  animals.  Most  smaller  vertebrates  gain 
adequate  moisture  in  their  diet  but  will  drink  or 
bathe  when  the  opportunity  arises.  The  primary  im- 
portance of  water  to  terrestrial  wildlife,  whether 
above  or  below  ground,  is  that  it  supplies  terrestrial 
vegetation  with  the  quality  and  quantity  needed 
for  health  and  growth. 


Birds 

As  stated  earlier,  habitat  features  most  important 
to  wildlife  are  very  difficult  to  extract  from  data 
sets  and  virtually  impossible  to  extract  and  confirm 
from  1-  to  2-year  studies.  What  may  appear  to  be  an 
important  habitat  component  one  year  may  not  even 
be  a  significant  component  again  in  a  5-  or  10-year 
study.  Bird  habitat  components  were  derived  from 
one  large  river  system  studied  for  at  least  7  years 
with  intensive  monthly  sampling  over  that  time  span. 
These  data  have  been  tested  on  other  large  desert 
riparian  systems,  including  a  revegetation  site  in 
which  the  plant  community  contained  all  of  the 
most  important  habitat  components  except  snags  for 
nest  cavities.  These  were  added  by  erecting  nest 
boxes.  The  revegetation  site  is  now  about  6  years 
old,  the  cottonwood  trees  are  more  than  15  m  (50 
ft)  high,  and  the  area  is  replete  with  the  avian  spe- 
cies that  it  was  designed  to  attract.  Unfortunately, 
the  area  is  too  small  (28  ha  [70  a.])  to  attract  and 
house  many  pairs  of  species  that  require  large  areas 
for  breeding  territories — again  demonstrating  the 
importance  of  island  size  effect. 

Keep  in  mind  that  these  habitat  components 
were  derived  for  birds  in  desert  riparian  systems.  Al- 
though similar  data  in  a  long-term  study  are  being 
collected  in  quaking  aspen  stands  in  Colorado  (Win- 
ternitz  1973,  1976;  Winternitz  and  Cahn  1983),  to 
our  knowledge  these  kinds  of  data  do  not  exist  for 
birds  at  higher  elevations.  Not  all  of  these  compo- 
nents may  be  important  at  higher  elevations  and 
if  they  are,  their  order  of  importance  may  also  differ. 
These  habitat  components  should  be  used  with  cau- 
tion at  higher  elevations,  and  where  supporting  data 
are  available  we  have  included  the  references.  Again, 
other  and  better  references  may  be  available  and 
should  be  used. 

Frequently,  biologists  need  bird  species  lists  to 
illustrate  the  importance  of  riparian  habitats,  to  refer- 
ence particular  species,  or  for  other  reasons.  A  list 
of  desert  riparian  species  in  each  desert  and  their 


Riparian  Habitats 


175 


dependency  on  riparian  revegetation  species  in  each 
desert  is  presented  by  Ohmart  and  Anderson  (1982). 
Knopf  ( 1985)  provided  a  list  of  birds  observed  in  a 
2-year  study  along  the  Front  Range  in  Colorado. 

We  have  found  the  following  habitat  compo- 
nents, in  the  order  given,  most  important  for  entire 
avian  communities  in  desert  riparian  areas: 

1.  Tree  species 

a.  Fremont  cottonwoods  (Populus  fremontii) 
and  Goodding  willows  (Salix  gooddingii) 

b.  Honey  mesquite 

2.  Shrubs 

a.  Quail  bush  (Atriplex  lentiformes) 

b.  Iodine  bush  or  inkweed  (Suaeda  torreyana) 

3.  Mistletoe  (Phoradendron  californicum) 

4.  Foliage  density 

5.  Foliage  height  diversity  (FHD) 

6.  Snags 

7.  Patchiness  (PI) 

Tree  Species. 


Cottonwoods  and  Willows.  This  tree 
component  was  widespread  along  most  permanent 
streams  and  rivers  at  lower  elevations  in  the  western 
U.S.  before  European  settlers.  Its  widespread 
distribution  and  antiquity  in  western  North  America 
(Axelrod  1958)  undoubtedly  provided  many 
opportunities  for  an  evolving  insectivorous  avifauna. 


A  lush  canopy  provided  shade,  cover,  and  a  myriad 
of  insects.  The  rough,  ever-sloughing  bark  attracts 
wood-boring  larvae  plus  a  number  of  other 
arthropods,  which  provide  forage  for  bark-gleaning 
and  trunk-scaling  birds.  The  soft  wood  is  easily 
excavated  by  woodpeckers  and,  when  abandoned, 
secondary  cavity- nesting  species  such  as  Lucy's 
warblers  (Vermivora  luciae),  brown-crested 
flycatchers  (Myiarchus  tyrannulus),  and  elf  owls 
(Micrathene  whitneyi)  have  an  array  of  vacant 
cavities  for  their  nesting  activities.  Even  in  winter 
after  leaf  drop,  the  rough,  splitting  bark  provides 
foraging  opportunities  for  numerous  bird  species.  In 
early  spring  the  flowers,  laden  with  pollen,  are 
swarmed  by  arthropods,  and  wintering  and  migrant 
birds  consume  the  pollen  and  insects  attracted  to 
the  flowers.  As  gallery  forests  of  these  trees  age, 
some  die  providing  snags  and  light  penetration  to 
the  forest  floor.  Shrubs,  other  trees,  and  annuals 
invade  to  provide  patchiness  that  attracts  other  bird 
species. 

These  large  (18-24  m  [60-80  ft]  tall  and  2-3  m 
[6-10  ft]  DBH),  branching  trees  with  their  attendant 
insect  fauna,  bird  life,  and  proximity  to  a  stream  with 
its  aquatic  life,  also  attract  a  number  of  wintering 
and  breeding  raptors.  These  birds,  at  the  top  of  the 
food  chain,  find  a  rich  and  varied  prey  base  upon 
which  to  feed. 

Studies  of  this  community  type  in  California 
(Gaines  1977),  Arizona  (Carothers  et  al.  1974;  An- 
derson and  Ohmart  1984;  Rice  et  al.  1984),  New 
Mexico  (Hubbard  1971;  Hink  and  Ohmart  1984), 
and  in  Texas  (Wauer  1977;  Engel-Wilson  and 
Ohmart  1978)  attest  to  its  wildlife  value.  As  this 
community  type  becomes  extirpated  from  the  West 
Coast  eastward,  at  least  10  bird  species  will  be  lost 
as  well. 


sA   m 


■■■■  'f*^m'-- 

m4     8|§p&;    '  v "■'■'* 


Cottonwoodwillow  association. 


Woodpeckers  can  easily  excavate  cavities  in  soft-wood 
trees,  such  as  the  cottonwoods. 


176 


Riparian  Habitats 


Deciduous  trees  at  higher  elevations  may  be  as 
important  to  birds  as  cottonwoods  and  willows  are 
in  desert  riparian  systems.  Bull  and  Skovlin  (1982), 
working  in  Oregon,  reported  that  bird  diversity  and 
species  composition  changed  with  the  amount  of 
deciduous  vegetation  as  it  ranged  from  high 
(>  40%  ),  to  moderate  (15-30%  ),  to  low  (<  1%  ). 
Birds  using  deciduous  vegetation  in  the  area  sampled 
were  the  only  group  highly  dependent  on  this  habi- 
tat component.  A  long-term  study  in  quaking  aspen 
stands  in  Colorado  has  documented  the  value  of  this 
deciduous  tree  to  birds  (Winternitz  1976). 


A  2-year  study  by  Knopf  (1985),  at  elevations 
ranging  between  1,200  and  2,750  m  (3,973  and 
9,022  ft),  indicated  that  breeding  birds  in  riparian 
habitats  were  more  simplistically  structured  to  habi- 
tat components  at  higher  elevations  than  at  lower 
elevations.  He  reported  dramatic  changes  in  species 
richness  at  intermediate  elevations,  which  could 
indicate  less  structuring  of  avian  communities  at 
these  elevations  or  that  his  study  period  was  too 
short  to  fully  document  what  was  occurring  at  these 
elevations.  At  elevations  above  1,909  m  (6,263  ft), 
the  importance  of  riparian  habitat  declined;  the  up- 
lands contained  a  more  diverse  avifauna  than  the 
riparian  areas.  Further  testing  of  these  correlations 
are  needed,  but  before  wildlife  biologists  de-empha- 
size intermediate  elevation  habitats,  we  strongly 
recommend  that  the  manager  fully  determine 
whether  the  wildlife  values  are  indeed  low  and  the 
potential  for  maximum  riparian  productivity  on  a 
site  has  been  achieved. 


Possibly,  the  entire  western  riparian  avifauna  has 
evolved  more  closely  with  deciduous  tree  species  at 
various  elevations  than  other  habitat  components. 
These  trees  provide  the  essentials  for  life,  e.g.,  food, 
cover,  and  space  for  these  insectivorous  birds  that 
can  also  obtain  needed  water  from  the  riparian  sys- 
tem. Box  elder  (Acer  negundo),  walnut  (juglans 
major),  sycamore,  narrow-leaf  cottonwood,  and  oth- 
ers may  be  ecologically  equivalent  to  cottonwood- 
willow  at  lower  elevations.  The  importance  of  these 
trees  to  wildlife  cannot  be  overstressed  if  they  are 
true  surrogates;  riparian  habitats  should  be  managed 
to  ensure  healthy  communities  with  young  replace- 
ment structural  types. 


Honey  Mesquite.  This  tree  is  deciduous  in 
desert  riparian  habitats.  It  grows  slowly  and  is  found 
primarily  on  the  highest  terrace  (second  terrace) 
away  from  the  river  where  channel  cutting  by  the 
river  seldom  occurs.  Again,  the  flowers  attract  wild- 
life, both  by  the  pollen  and  the  insects  feeding  on 
the  flowers.  The  fruits  or  beans  produced  by  this 
tree  are  rich  in  carbohydrates  and  are  consumed  by 
a  wide  range  of  wildlife  species. 


IS 


Honey  mesquite. 


In  low  areas  where  floodwaters  are  trapped, 
colloidal  materials  are  deposited  and  form  heavier 
soils  that  support  a  mixture  of  shrub  species,  primar- 
ily quail  bush,  four-winged  saltbush  (Atriplex  canes- 
cens),  wolfberry  (Lycium  sp),  and  inkweed.  These 
shrubs  provide  important  values  to  a  number  of 
wildlife  species  and  enhance  the  avian  productivity 
of  the  honey  mesquite  community. 

Honey  mesquite  communities  are  also  enhanced 
by  numerous  annual  plants  that  have  high  seed  pro- 
duction after  a  wet  summer  or  winter.  Seeds  pro- 
duced by  these  annuals  are  utilized  by  a  large 
granivorous  guild  that  includes  white-crowned  spar- 
rows (Zonotrichia  leucophyrs),  Gambel's  quail  (Cal- 
lipepla  gambelii),  and  Brewer's  sparrows  (Spizella 
breweri). 


Riparian  Habitats 


177 


Shrubs. 


Quail  Bush.  In  optimum  growing  conditions, 
quail  bush  reaches  heights  of3to4m(10to  14ft) 
and  a  single  plant  could  cover  13m    ( 1 40  ft  ). 
Although  an  evergreen,  this  shrub  drops  leaves  and 
stems  that  over  the  years  form  a  thick  layer  of 
composting  material.  This  dense  shrub  shades  the 
soil  to  help  hold  moisture  that  in  turn  expedites 
decomposition  of  the  litter  accumulation.  Ground- 
dwelling  birds,  such  as  quail,  thrashers  (Toxostoma 
sp.),  and  towhees  (Pipilo  sp.),  find  protection  and 
productive,  insect  foraging  areas  in  the  decomposing 
litter  under  this  shrub.  Small  insectivorous  birds 
forage  extensively  in  and  among  the  dense  stems  and 
leaves. 

When  this  shrub  is  mixed  with  a  community  of 
trees,  such  as  honey  mesquite  or  salt  cedar,  it  greatly 
enhances  the  wildlife  values  of  the  trees.  Also,  it 
forms  a  monoculture  where  quail  and  other  ground- 
dwelling  bird  species  can  attain  high  densities.  Its 
seeds  are  important  to  these  birds  as  well  because 
the  plant  is  a  prolific  seed-producing  species  and  the 
seeds  fall  to  the  ground  throughout  the  year. 

Iodine  Bush  or  Inkweed.  This  0.5-  to  1-m  (2- 
to  3-ft)  tall  shrub  is  used  by  numerous  ground-dwell- 
ing birds,  but  is  most  important  to  sage  sparrows 
(Amphispiza  belli}.  Wintering  sage  sparrows  ac- 
tively select  habitats  containing  this  plant  or  revege- 
tated  areas  containing  inkweed  (Meents  et  al.  1982). 
Seeds  and  plant  parts  of  inkweed  were  found  in  sage 
sparrow  gizzards,  but  why  they  actually  select  habi- 
tats with  a  preponderance  of  inkweed  is  unknown. 

Mistletoe.  Mistletoe  is  a  parasite,  widespread  on 
trees  and  shrubs  in  the  pea  family  (Leguminosae  or 
Fabaceae).  Although  it  can  be  found  on  many 
different  host  species,  it  appears  to  do  best  on  honey 
mesquite  trees.  The  dense  clumps  provide  shelter 
for  perching  birds,  nesting  cover  for  breeding 
species,  and  berries  for  a  number  of  frugivorous 
(fruit-eating)  birds,  especially  Phainopepla 
{Phainopepla  nitens;  Anderson  and  Ohmart  1978). 
Although  mistletoe  may  eventually  kill  some 
mesquite  trees,  its  value  to  wildlife  appears  to  offset 
its  negative  effects  to  mesquite.  In  healthy  riparian 
systems,  life  and  death  are  integral  parts  of  a  produc- 
tive community.  As  mesquite  trees  die,  they  produce 
hardwood  snags.  The  snags  attract  wood  boring 
insects  and  can  become  potential  nest  sites  for  the 
ladder-backed  woodpecker  {Picoides  scalaris). 
These  snags  also  provide  perches  for  hunting  raptors. 

Foliage  Density.  Until  recently,  the  importance  of 
foliage  density  (surface  area  of  leaves  and  stems/ 
area  )  or  foliage  height  diversity  could  not  be  readily 
separated.  In  recent  strip-clearing  studies,  foliage 


density  was  reduced  by  20-40% ,  whereas  foliage 
height  diversity  did  not  change.  However,  bird  life  in 
the  area  was  drastically  reduced  (Anderson  and 
Ohmart  unpubl.  ms).  This  seems  to  indicate  that 
foliage  density  is  more  important  in  the  higher 
vertical  layers  than  foliage  height  diversity,  as  it 
provides  better  forage  substrate  for  insect-gleaning 
birds,  concealment  from  predators,  and  sites  for 
nesting.  Canopy  and  mid-canopy  layers  of  vegetation 
are  low  quality  to  wildlife  unless  foliage  density  is 
moderate  or  high.  It  is  generally  believed  that  by 
increasing  foliage  density  in  each  layer  of  vegetation, 
the  carrying  capacity  of  insectivorous  birds  is 
increased  and  new  niches  are  added. 


Foliage  Height  Diversity  (FHD).  Foliage  height 
diversity  measures  how  evenly  foliage  is  distributed 
among  the  vertical  layers.  A  plant  community  with 
little  or  no  foliage  density  in  the  tree  canopy  layer 
would  have  a  lower  FHD  value  than  a  similar  plant 
community  with  dense  foliage  at  the  understory, 
midstory,  and  canopy  levels.  Thus,  foliage  height 
diversity  increases  when  foliage  density  values  are 
nearly  equal  among  all  vertical  layers  of  vegetation. 


:.      ■ 


>C   '■', 


Salt  cedar  community. 


Figure  2  shows  two  communities  with  differing 
height  diversities;  the  one  on  the  left  shows  low-, 
mid-,  and  upper-canopy  volumes  while  the 
community  on  the  right  shows  some  midstory 
volume  and  a  little  overstory. 

This  habitat  component  is  unquestionably  valu- 
able to  wildlife,  but  less  so  than  foliage  density. 
Other  species  of  trees  can  have  height  density  values 
similar  to  cottonwood  and  willow  habitat,  yet  not 
have  similar  species  richness  values  and  bird  densi- 
ties. Rice  et  al.  (1984)  examined  the  importance 
of  FHD,  patchiness  (PI),  and  individual  tree  species 
to  avian  communities.  Response  of  individual  bird 


178 


Riparian  Habitats 


ft? 


42? 


Figure  2.     Two  plant  communities  showing  different  foliage  height  diversities. 


species  was  significantly  higher  with  greater  fre- 
quency of  individual  tree  species  than  with  other 
vegetation  variables  depicting  structure.  Although 
height  densities  correlated  well  with  other  tested 
habitat  components,  they  were  strengthened  by  add- 
ing individual  tree  species.  Generally,  it  is  assumed 
that  as  one  adds  layers  of  vegetation,  these  additional 
layers  provide  niches  for  additional  species  of  birds. 

Snags.  Snags  are  extremely  important  to  a  number 
of  species  in  riparian  ecosystems.  Many  species  use 
snags  for  nesting,  feeding,  roosting,  as  hunting  and 
loafing  perches,  for  hibernacula  while  overwintering, 
and  as  a  moist  refuge  in  dry  summer  months.  Snags 
can  be,  and  have  been  reported  as,  limiting  to  the 
presence  and  abundance  of  cavity-nesting  birds 
(Haapanen  1965;  Balda  1975;  Conner  et  al.  1975; 
Evans  and  Conner  1979;  Scott  1979;  Mannan  et  al. 
1980;  Dickson  et  al.  1983).  A  wildlife  snag 
symposium  (Davis  et  al.  1983)  provided  a 
compilation  of  papers  on  this  subject  and  should  be 
referred  to  for  more  specific  information. 

Many  of  the  above  studies  and  others  examining 
the  importance  of  snags  to  birds  have  been  con- 
ducted in  upland  habitats,  but  snags  in  riparian  habi- 
tats are  as  limiting,  if  not  more  so,  than  those  in 
upland  areas  and  have  the  same  effect  on  many  pri- 
mary and  secondary  cavity-nesting  species.  T.E.  Mar- 
tin (Arizona  State  University,  unpubl.  ms)  examined 
the  importance  of  snags  in  1 3  riparian  habitats 
where  the  riparian  species  were  primarily  big  tooth 
maple  (Acer  grandidentatum),  quaking  aspen,  and 
New  Mexico  locust  (Robinia  neomextcana),  with 
an  understory  of  golden  pea  (Thermopsis pineto- 
rum)  and  raspberry  (Rubus  strigosus).  Snags  were 
primarily  aspen  and  were  significantly  (P  <  0.001) 
more  abundant  along  streams  on  north-facing  slopes 
than  on  south-facing  slopes.  Densities  of  snag-using 
species  were  greatest  in  snag-rich  habitats,  and  more 
snag-using  species  were  present. 

Brush  et  al.  (1983)  found  that  this  habitat  com- 
ponent, especially  softwood  snags,  is  generally  a 


limiting  factor  in  heavily  managed  desert  riparian 
systems.  Fires,  floods,  and  even  removal  of  snags  to 
protect  water  skiers  has  virtually  eliminated  this 
important  component  in  the  lower  Colorado  River 
riparian  ecosystem. 

Primary  and  secondary  cavity-nesting  species  are 
vulnerable  unless  management  is  aware  of  their 
needs  and  places  high  value  on  ensuring  large,  qual- 
ity snags  for  wildlife.  Unfortunately,  a  prerequisite 
to  snags  is  living  trees;  this  handicap  has  been  partly 
overcome  and  artificial  snags  are  being  excavated 
by  some  woodpeckers  (Grubb  et  al.  1983;  Peterson 
and  Grubb  1983).  Presence  or  absence  of  snags 
should  be  noted  as  riparian  habitats  and  assessed  for 
wildlife. 

Snags  provide  essential  habitat  for  about  85  bird 
species  (Scott  et  al.  1977)  that  use  natural  cavities, 
excavate  their  own,  or  use  holes  excavated  by  other 
species.  Many  of  these  species  are  obligate  riparian 
forms.  In  quaking  aspen  stands  in  Colorado,  almost 
40%  of  the  breeding  bird  species  using  this  type 
used  nest  holes,  and  trees  containing  these  cavities 
were  usually  well  over  100  years  old  (Winternitz 
and  Cahn  1983). 

Patchiness.  Horizontal  foliage  diversity  or 
intraplant  community  patchiness  on  the  horizontal 
scale  has  long  been  recognized  as  valuable  to 
wildlife.  Figure  3  attempts  to  show  patchiness 
in  a  plant  community;  the  open  areas  between 
interlocking  trees  and  some  midcanopy  vegetation 
produce  a  patchy  effect  in  this  forested  habitat. 
Patchiness,  like  foliage  height  diversity,  is  thought  to 
create  additional  niches  for  birds.  Some  species  are 
found  primarily  in  continuous  riparian  forests, 
whereas  others  are  attracted  to  openings  in  the 
canopy  where  lesser  trees  or  shrubs  provide  patches 
in  an  otherwise  continuous  canopy.  These  patches 
offer  habitat  space  for  new  species  that  would 
otherwise  be  absent,  because  as  patchiness  increases 
so  does  species  richness.  Anderson  et  al.  (1983) 
recently  developed  a  method  to  quantify  patchiness. 


Riparian  Habitats 


179 


1^  VP    \W 


Jk,  k-^Ok^  JWii'a* 


Figure  3.      Patchiness  in  a  plant  community 
(diagrammatic). 

Mammals 

Large  Mammals.  The  total  importance  of  riparian 
habitats  to  large  mammals  is  not  fully  understood, 
but  undoubtedly  these  ecosystems  provide  three 
essential  resources  to  this  group:  food,  water,  and 
cover.  Dealy  et  al.  (1981)  pointed  out  that  riparian 
habitats  in  general,  and  quaking  aspens  in  particular, 
provide  both  thermal  cover  and  forage  for  ungulates. 


In  the  Modoc  National  Forest  in  northern  Cali- 
fornia, Salwasser  and  Shimamoto  (1984)  examined 
animal  use  of  wetlands  that  have  developed  since  the 
1920s  behind  water  storage  dams.  They  examined 
wetland  use  for  three  classes  of  large  mammals:  feral 
horses,  domestic  cattle,  and  pronghorn  antelope 
(Antilocapra  amertcana).  Although  these  wetlands 
made  up  10%  of  the  available  habitats,  they  were 
used  heavily  by  all  classes  from  midsummer  through 
fall.  Pronghorn  antelope  use  peaked  at  80%  in  Au- 
gust and  remained  over  40%  into  October.  Feral 
horse  use  peaked  at  78%  in  July  and  remained  over 
40%  into  early  August.  Cattle  use  peaked  at  80% 
in  August  and  exceeded  50%  since  late  June.  Annual 
use  varied  by  the  three  classes,  but  all  used  the  de- 
veloped wetlands  far  more  extensively  than  the 
wetlands  that  were  available  among  habitats  in  mid- 
summer to  fall. 


Studies  examining  habitat  use  by  elk  in  mixed- 
forest  types  have  shown  that  of  1 5  associated  habitat 
types,  riparian  zones  were  most  heavily  used  (Peder- 
son  et  al.  1979).  Riparian  corridors  also  served  as 
travel  routes  between  areas. 

Although  elk  will  graze  areas  used  by  cattle, 
domestic  livestock  grazing  appears  to  diminish  use  of 
the  area  by  elk  (Blood  1966;  Nelson  and  Burnell 
1975;  Burbridge  and  Neff  1976;  Skovlin  et  al.  1976). 
Diminished  elk  use  of  areas  where  cattle  have  grazed 
is  probably  related  to  season  of  use  and  grazing  in- 
tensity by  cattle  before  or  during  the  time  of  elk  use. 
Fall  and  winter  use  by  elk  declined  significantly  fol- 
lowing cattle  use  in  late  spring  and  summer  on  pine- 
bunchgrass  uplands  near  riparian  zones  in  the  Blue 
Mountains  of  Oregon  (Skovlin  1984). 

How  much  vegetation  is  removed  in  riparian 
habitats  by  big  game  is  not  well  studied,  but  Skovlin 
(1984)  indicated  that  in  northeastern  Oregon,  deer 
and  elk  accounted  for  about  one-third  of  the  total 
browsing  on  riparian  trees  and  shrubs.  Cattle  appar- 
ently removed  the  remaining  two-thirds.  He  also 
reported  that  studies  conducted  near  the  above  sites, 
where  deferred  or  seasonal  livestock  grazing  has 
been  practiced,  showed  similar  vegetation  removal 
rates  for  big  game  and  livestock. 


The  extent  to  which  elk  (Cervus  elaphiis)  use 
and  depend  on  riparian  habitats,  other  than  as  travel 
routes,  in  the  nonsummer  months  is  not  well  known. 
However,  in  summer  in  western  Montana,  80%  of 
elk  use  in  July  was  within  0.4  km  (0.25  mi.)  of  per- 
manent water  (Marcum  1975).  Skovlin  (1984)  re- 
ported heavy  preference  for  summer  range  within 
0.8  km  (0.5  mi.)  from  water.  Whether  the  elk  are  at- 
tracted to  riparian  habitats  for  the  green  lush  forage, 
water,  or  both  is  unclear;  perhaps  only  lactating 
cows  depend  on  free  water  sources  at  this  time 
(Marcum  1975;  Black  et  al.  1976;  Thomas  et  al. 
1976). 


Soil  compaction  from  grazing  by  wild  ungulates 
is  normally  not  a  problem  in  riparian  zones  in  fall 
and  winter,  but  some  compaction  damage  of  satu- 
rated or  wet  soils  in  the  upland  range  in  early  spring 
may  occur  (Skovlin  1984).  He  also  reported  that 
shrub  use  by  elk  in  early  spring  and  summer,  before 
grazing  by  domestic  livestock,  can  significantly  affect 
shrub  survival  during  the  ensuing  growing  season. 
Cattle  browsing  on  the  same  shrubs  grazed  by  elk 
could  add  to  the  shrub  mortality  problem. 

Medium-Sized  Mammals.  Numerous  medium-sized 
species  of  mammals  are  either  obligate  or  facultative 


180 


Riparian  Habitats 


users  of  riparian  systems  at  all  elevations.  For 
example,  in  Big  Bend  National  Park  in  west  Texas, 
1 5  medium-sized  species  were  reported  using 
riparian  ecosystems  and  one,  the  beaver  {Castor 
canadensis),  was  a  true  obligate  (Boeer  and 
Schmidly  1977).  Williams  and  Kilburn  (1984) 
reported  that  of  the  502  recent  native  species  and 
subspecies  of  land  mammals  in  California,  about  25% 
(133  taxa)  were  limited  to  or  largely  depended  on 
riparian  and  other  types  of  wetland  communities. 
They  further  stated  that  "No  other  general  type  of 
mammalian  habitat  in  California  approaches  riparian 
and  other  wetland  communities  in  importance  to 

mammals "Of  the  1 5  species  they  listed,  5  were 

in  the  medium-sized  category.  Similarly,  in  Mexico, 
seven  medium-sized  mammals,  threatened  with 
extinction,  are  confined  to  riparian  ecosystems 
(Ceballos  1985). 

Lack  of  any  systematic  and  long-term  studies  of 
medium-sized  mammals  makes  it  difficult  for  man- 
agers to  establish  specific  habitat  management 
criteria  that  would  ensure  these  species  are  not  ex- 
tirpated by  water  management,  timber  harvest,  or 
grazing  practices.  As  in  most  past  management  deci- 
sions, the  best  approach  is  to  review  natural  history 
studies  of  these  species  and  avoid  eliminating  habitat 
components  that  have  been  identified  as  important. 

Many  of  these  species  are  aquatic  or  semi- 
aquatic  and  feed  on  plant  and  animal  matter  in  or 
along  the  stream.  Consequently,  instream  flows  and 
water  quality  represent  important  habitat  factors. 
Streamside  vegetation  also  provides  food  and  cover 
for  these  species,  so  severe  vegetation  removal  by 
domestic  livestock  can  be  important.  Therefore,  to 
provide  optimum  conditions  for  these  species,  one 
must  be  cognizant  of  the  specific  needs  for  each 
species  found  in  the  area. 

In  California,  five  medium-sized  mammal  species 
considered  obligate  to  riparian  habitats  have  been 
recommended  for  special  consideration  (Williams 
and  Kilburn  1984).  One,  the  snowshoe  hare  (Lepus 
americanus),  occurs  in  dense  thickets  of  alders  (Al- 
nus  sp. ),  willows,  and  other  shrubs  in  the  Sierra 
Nevada  range  (Orr  1940).  Dense  thickets  of  young 
conifers  also  serve  as  cover  for  this  hare  (Williams 
1985).  This  species  may  not  depend  on  riparian 
habitat  in  other  portions  of  its  range,  which  may  not 
be  unusual  for  some  high-elevation  forms. 

The  mountain  beaver  (Aplodontia  rufa),  also 
found  along  moist,  forested  habitats  along  the  Pacific 
Slope,  occurs  in  small  colonies.  It  favors  moist 
slopes,  supporting  lush  growths  of  forbs,  and  often 
excavates  its  burrows  next  to  a  stream.  Its  diet  in- 
cludes a  variety  of  forbs  and  the  buds,  twigs,  and 
bark  of  willow  and  dogwood  (Cornus  sp.).  It  also 
cuts  forbs  to  pile  as  hay  (Grinnell  and  Storer  1924). 


A  number  of  medium-sized  mammals  use  ripar- 
ian ecosystems  extensively  although  they  occur  in 
the  uplands.  Densities  are  lower  in  upland  habitats 
than  along  riparian  systems,  as  noted  by  studies  of 
the  gray  fox  (Urocyon  cinereoargenteus)  along 
Putah  Creek  near  Davis,  California  (Hallbery  and 
Trapp  1984 ).  Two  males  and  females  with  radio  col- 
lars spent  76%  of  their  nocturnal  and  92%  of  their 
diurnal  time  in  riparian  habitats  and  the  remainder  in 
agricultural  habitats.  An  example  of  the  latter  is  ex- 
emplified in  ringtails  {Bassariscus  astutus ),  where 
this  species'  densities  range  from  2.5  to  5  times 
greater  in  riparian  ecosystems  as  opposed  to  upland 
habitats  (Belluomini  and  Trapp  1984).  Numerous 
other  facultative  riparian  species  obviously  fit  into 
these  categories,  stressing  the  importance  of  these 
habitats  to  medium-sized  mammals. 

The  fisher  (Martes  pennanti)  is  poorly  studied 
especially  in  the  western  U.S.,  but  in  general,  is 
thought  to  depend  on  riparian  habitats  for  travel  and 
escape  routes  (de  Vos  1951,  1952;  Kelly  1977;  Buck 
1982;  Mullis  1985).  Of  the  studies  conducted  thus 
far,  only  Kelly  (1977)  has  reported  that  the  long  axis 
of  most  home  ranges  in  New  Hampshire  have  tended 
to  parallel  drainages.  Regardless,  these  riparian  habi- 
tats are  believed  to  be  very  important  to  this  species. 

In  the  West,  fisher  populations  have  been 
thought  to  be  declining  because  of  activities  such  as 
timber  harvesting,  road  building,  and  overtrapping. 
In  California  the  trapping  season  was  closed  in  1946, 
and  in  recent  years  these  populations  appear  to  have 
increased  (Mullis  1985).  Indications  are  that  the 
marten  (M.  americana)  may  similarly  need  riparian 
habitats. 

Beavers,  once  a  dominant  aquatic  mammal  in 
riparian  systems,  have  virtually  been  eliminated  in 
western  streams  through  trapping,  shooting,  instream 
flow  reductions,  and  other  factors.  The  beaver  needs 
streams  1  m  (3  ft)  or  more  deep,  and  timber  for 
food,  lodge,  and  dam  construction.  Its  propensity  to 
construct  dams  has  frequently  made  it  an  undesira- 
ble riverine  species.  Although  primarily  preferring  to 
live  in  lodges,  the  beaver  will  excavate  dens  in  the 
riverbank  (Nelson  and  Hooper  1976).  Its  damming 
activities  frequently  have  flooded  agricultural  lands, 
forest  lands,  and  damaged  irrigation  systems. 

Parker  et  al.  ( 1985)  recently  suggested  that 
beavers  might  have  played  an  important  role  in  re- 
sisting minor  perturbations  in  lower-order  streams. 
Their  model  and  arguments  are  not  too  difficult  to 
visualize  when  the  impact  this  species  can  have  on 
streams  and  riparian  trees  is  considered.  Lang  and 
Weider  (1984)  suggested  that  beavers  have  altered 
the  structure  of  forests  in  West  Virginia.  Platts  et 
al.  (1985)  supported  the  beaver  model  for  large 
storm  events.  However,  the  Parker  et  al.  (1985) 


Riparian  Habitats 


181 


model  needs  further  testing  before  it  can  be  used  as 
a  possible  management  tool  in  resisting  mild 
perturbations. 

Scott  (1984)  justified  the  importance  of  beaver, 
mink  (Mustela  vison),  and  muskrat  (Ondatra  zibeth- 
icus)  by  examining  the  dollar  return  of  these  species 
in  California.  In  some  instances,  the  fur-return  value 
of  these  species  may  represent  up  to  71%  of  an  indi- 
vidual's annual  income.  In  economically  depressed 
communities  this  could  be  an  important  source  of 
income. 

Small  Mammals.  Soil  texture,  structure,  and 
moisture  seem  to  be  important  in  habitat  selection 
of  many  small  mammals  that  burrow.  Other  rodents 
respond  to  riparian  habitats  in  varying  manners: 
sciurids  respond  more  to  tree  species  and  height  of 
trees;  heteromyids,  more  to  soil  and  open  habitats; 
castorids,  to  water  and  forage  availability;  cricetids, 
to  vegetation  density  and  structure;  and  zapodidae, 
possibly  to  grass  height  and  stem  densities  in  moist 
soils. 

Although  identifying  habitat  components  for  all 
small  mammals  in  riparian  communities  is  far  from 
complete,  there  is  good  evidence  to  demonstrate  the 
importance  of  these  habitats  to  small  mammals. 
Cross  (1985)  found  that  riparian  habitats  in  south- 
western Oregon,  composed  of  mixed  conifer  and 
deciduous  broadleaf  trees,  invariably  had  greater  spe- 
cies richness  and  total  small  mammal  biomass  than 
upland  sites.  Studies  (Stamp  and  Ohmart  1979)  in 
riparian  habitats  in  the  Sonoran  Desert  of  Arizona 
show  similar  results. 


Herptofauna 

To  our  knowledge,  there  are  no  in-depth,  long- 
term  data  from  a  riparian  ecosystem  at  any  elevation 
dealing  with  habitat  factors  important  to  an  entire 
herptofauna.  This  is  a  serious  omission  in  riparian 
studies  in  that  reptiles  and  amphibians  are  probably 
as  important,  and  possibly  more  so,  than  birds  and 
mammals  in  energy  flow  and  nutrient  cycling.  There 
are  a  number  of  autecological  studies  on  specific 
species,  however,  which  we  used  to  delineate  some 
of  the  most  important  habitat  features  for  amphibians 
and  reptiles.  Until  in-depth,  long-term  community 
ecology  studies  are  conducted,  we  will  be  making 
many  assumptions. 

In  the  Hubbard  Brook  Experimental  Forest  in 
New  Hampshire,  Burton  and  Likens  (1975)  esti- 
mated salamander  densities  of  2,950/ha  (  1,180/a.), 
which  exceeded  density  estimates  of  birds  and  mam- 
mals. Salamander  biomass  exceeded  that  of  birds  by 
216  times  and  approximated  that  of  small  mammals. 
Similar  densities  of  Siskiyou  Mountain  salamanders 
(Plethodon  stormi)  have  been  estimated  at  densities 


of  2,700/ha  (1,080/a.;  Nussbaum  1974)  in  optimum 
habitats  in  Oregon;  and  Murphy  and  Hall  (1981) 
reported  that  in  Oregon,  the  Pacific  giant  salamander 
(Dicamptodon  eusatus)  may  make  up  as  much  as 
99%  of  the  total  predator  biomass  in  some  streams. 

Reptile  densities  are  equally  high  in  the  West. 
The  western  pond  turtle  (Clemmys  marmorata)  has 
been  reported  at  densities  of  425/ha  ( 170/a.;  Bury 
1979)  and  Sonoran  mud  turtles  (Kinosternon  sono- 
riense)  in  Arizona  at  825/ha  (330/a.;  Hulse  1974). 
Southern  ringnecked  snakes  (Diadophis  punctatus) 
have  been  reported  at  densities  of  1,000-1,500/ha 
(400-600/a.;  Sullivan  1981). 

In  general,  amphibians  as  a  species  group  are 
more  dependent  on  riparian  ecosystems  than  are 
reptiles.  Much  of  this  revolves  around  their  evolu- 
tionary life-style;  they  are  aquatic  or  semi-aquatic 
and  lay  open  eggs  (nonshelled)  in  water  or  very 
moist  areas.  For  aquatic  species,  many  of  the  same 
needs  of  fishes  would  apply  to  these  forms,  e.g., 
shaded  stream  for  cooler  water  and  higher  oxygen 
levels,  productive  stream  bottoms,  and  escape  cover. 
For  more  terrestrial  forms,  such  as  some  salaman- 
ders, the  presence  of  rotting  logs  and  dense  ground 
cover  (litter  or  vegetation)  is  essential. 

Because  most  species  in  this  group  are  insecti- 
vorous and  carnivorous,  and  most  spend  their  active 
periods  in  shallow  water,  on  the  ground,  or  in  trees, 
any  action  that  reduces  densities  of  trees,  shrubs, 
or  other  vegetation  and  reduces  high  insect  biomass 
has  to  be  considered  negative.  On  the  Colorado 
River,  actions  that  help  birds  were  also  considered 
to  help  this  group,  especially  reptiles.  This  is  proba- 
bly not  totally  true,  but  revegetation  sites  that  are 
primarily  designed  for  birds  and  small  mammals  also 
support  high  populations  of  reptiles — higher  than 
those  found  in  most  natural  communities.  Because 
many  biologists  believe  that  birds  are  nothing  more 
than  glorified  reptiles,  these  results  should  not  be 
surprising. 


-  -  -   — *»-»v. 


182 


Riparian  Habitats 


For  some  amphibians,  the  maintenance  of  back- 
waters with  dense-to-moderate  emergent  stands  of 
vegetation  is  essential.  There  should  be  stable  levels 
of  instream  flow  with  good  bank  development  stabi- 
lized with  vegetation.  Large,  decaying  logs  in  the 
floodplain  and  in  the  adjacent  uplands  are  necessary 
habitat  components  for  numerous  amphibians,  such 
as  salamanders. 

Many  herpetile  species  either  overwinter  in  the 
soil  of  the  floodplain  or  in  decaying  wood  under 
logs.  Others  overwinter  in  hibernacula  in  downed, 
decaying  logs;  snags;  and  even  live  trees  with  natural 
cavities  or  heart  rot.  These  habitat  features  are  im- 
portant to  amphibians  for  their  overwinter  survival. 
Loss  of  one  or  more  of  these  critical  habitat  features 
can  break  the  annual  life  cycle  of  a  species,  resulting 
in  its  local  extirpation. 

Another  important  feature  of  riparian  habitats  to 
amphibians  and  reptiles,  frequently  overlooked  by 
managers,  is  the  use  of  these  systems  as  corridors  for 
dispersal  and  genetic  continuity  between  popula- 
tions. Brode  and  Bury  (1984)  stressed  the  impor- 
tance of  continuous  riparian  ecosystems  to  help 
maintain  genetic  heterogeneity,  and  noted  that  habi- 
tat disruption  has  resulted  in  isolated  populations 
of  many  species  in  California  riparian  habitats. 

Although  all  the  habitat  components  needed  for 
species  survival  may  be  present  along  a  stream,  a 
population  or  species  can  be  extirpated  for  a  num- 
ber of  reasons:  ( 1 )  their  density  is  so  low  that  an 
adequate  number  of  matings  do  not  occur  to  sustain 
population  levels;  (2)  genetic  drift;  (3)  genetic  het- 
erozygosity is  not  adequate  to  sustain  a  healthy  pop- 
ulation; and  (4)  some  density-independent  mortality 
factor  eliminates  the  population,  e.g.,  severe  and 
prolonged  freezing  or  drought.  For  this  group,  habi- 
tat continuity  and  patch  size  should  be  considered 
when  making  management  decisions. 


Habitat-Wildlife  Correlations 

A  few  years  ago,  MacArthur  and  MacArthur 
(  1961 )  reported  a  strong  correlation  between  east- 
ern forest  birds  and  FHD.  Over  the  years,  ecologists 
and  managers  have  looked  for  one  or  more  habitat 
features  that  highly  correlated  with  wildlife.  Bird 
species  diversity  (BSD)  is  identical  in  basic  concept 
to  FHD,  but  is  based  on  the  number  of  bird  species 
(species  richness)  in  the  plant  community  and  how 
evenly  distributed  the  density  of  each  species  is 
among  the  total  number  of  species.  An  avian  commu- 
nity containing  1 5  species,  where  2  or  3  make  up 
90%  of  the  total  birds,  would  have  low  BSD  values 
compared  with  an  avian  community  where  the  total 
bird  numbers  were  more  evenly  distributed  among 
the  15  species. 


The  BSD  and  FHD  values  and  their  relationships 
have  not  been  well  understood.  For  a  brief  period 
it  was  thought  that  the  key  to  good  management  was 
to  manage  for  high  BSD  values.  The  weakness  of 
this  approach  becomes  fairly  obvious  in  the  follow- 
ing example.  In  most  habitat  situations  the  density  of 
rare  or  federally  and  state-listed  species  is  low.  Con- 
sequently, in  the  earlier  example  of  1 5  bird  species 
in  a  community,  the  2  or  3  species  in  very  low  den- 
sity could  be  lost  and  hardly  change  the  BSD  value. 
Therefore,  one  can  see  the  fallacy  of  trying  to  man- 
age habitats  or  ecosystems  based  on  BSD  values. 
Keep  in  mind  that  BSD  and  species  richness  (num- 
ber of  species  present )  are  very  different  terms  and 
should  not  be  used  interchangeably. 

Since  high  BSD-FHD  relationships  were  reported 
by  MacArthur  and  MacArthur  ( 1961 )  in  eastern  for- 
ests, many  studies  have  supported  or  refuted  this 
relationship  throughout  North  America.  Many  scien- 
tists collecting  data  have  not  looked  at  other  habitat 
components  and  have  been  satisfied  with  relatively 
good  correlations  between  BSD  and  FHD.  A  few 
other  studies  looked  at  important  vegetation  fea- 
tures, but  these  research  results  suffer  from  one  or 
more  of  the  following:  lack  of  tight  experimental 
design;  limited  data  collection,  such  as  only  during 
the  breeding  season;  and  short-term  studies,  con- 
ducted for  only  1  or  2  years.  These  are  central  crite- 
ria in  judging  the  value  of  field  data  as  applied  to 
management  situations.  Much  money,  time,  and  ef- 
fort have  been  and  can  be  lost  when  management 
decisions  are  based  on  poorly  conducted  studies. 

Many  federal  and  state  agencies  have  tried  to  re- 
duce ecosystems  to  one  or  a  few  variables  for  man- 
agement purposes.  Our  long-term  research,  plus  that 
of  others,  argues  persuasively  that  ecosystems,  plant 
communities,  and  even  small  habitats  are  too  com- 
plex to  reduce  to  one  or  a  few  numbers  for  manage- 
ment purposes.  We  know  of  no  shortcut  approaches 
to  good  wildlife  and  habitat  management  and  seri- 
ously doubt  any  will  be  found.  Unless  managing 
agencies  discover  this  and  proceed  with  the  task  at 
hand,  which  is  the  collection  of  in-depth,  long-term 
changes  in  plant  and  animal  communities,  we  as 
biologists  will  always  be  playing  "catch-up  biology" 
or  patching  data  together  along  with  our  "gut  feel- 
ings" to  make  important  management  decisions  and 
recommendations. 


DATA  COLLECTION  PRIORITIES 

Biologists  are  so  frequently  enthusiastic  about 
wildlife  that  when  presented  with  the  opportunity  to 
study  an  area,  they  immediately  begin  thinking  of 
ways  to  census  various  classes  of  vertebrates,  study 
life  histories  of  animals,  or  collect  fauna.  Although 
important,  highly  desirable,  and  needed,  these  are 
probably  the  last  steps  that  should  be  undertaken. 


Riparian  Habitats 


183 


We  have  developed  a  list  of  priorities  that  we  be- 
lieve will  help  the  land  manager  develop  a  data  base 
on  the  riparian  ecosystem  and  also  document  pro- 
cesses occurring  or  about  to  occur  in  the  ecosystem. 
Since  the  limiting  factor  in  any  study  is  money, 
which  is  people,  vehicles,  equipment,  or  some  other 
factors  requiring  funding,  our  list  assumes  that  funds 
will  be  limited. 

Developing  Vegetation  Maps 

Develop  a  fine-grained  map  of  the  riparian  vege- 
tation ecosystem(s)  under  management  considera- 
tion. By  fine-grained,  we  mean  a  map  depicting  and 
naming  the  plant  communities  at  a  resolution  of 
possibly  0.2  ha  (0.5  a.)  or  less,  if  necessary,  to  show 
discrete  plant  communities.  Common  sense  must 
prevail  because  overzealous  type  mapping  could 
reduce  a  community  to  one  tree,  then  to  a  limb,  and 
to  a  leaf.  If  two  tree  species  occur  together  along  a 
stream  or  river,  the  community  might  be  delineated 
as  a  cottonwood-willow  association.  Or  if  a  tree  wil- 
low grows  in  one  area  and  a  shrub  willow  grows 
in  another,  you  would  have  Goodding  willow  (Salix 
gooddingii)  as  a  separate  community  from  coyote 
willow  (S.  exigua)  and  so  forth.  Because  riparian 
systems  are  linear,  and  frequently  vary  in  width  from 
a  few  yards  to  many  miles,  maps  must  be  scaled 
upward  to  include  small  communities  that  may  sup- 
port endangered  wildlife  species  or  very  highly  spe- 
cialized species. 


If  numerous  riparian  ecosystems  are  under  man- 
agement consideration  and  funding  only  allows  half 
of  them  to  be  adequately  mapped  or  all  to  be  super- 
ficially mapped,  then  we  recommend  prioritizing 
the  streams  and  intensively  mapping  only  half  of 
them.  Otherwise,  a  poor  job  on  all  of  them  provides 
little,  if  any,  good  information  from  which  to  make 
crucial  management  decisions  and  will  probably 
be  meaningless  to  future  managers  attempting  to  in- 
terpret and  use  the  maps. 

Good  vegetation  type  maps  should  be  accom- 
panied by  the  criteria  used  in  naming  each  major 
community.  For  example,  if  you  elect  to  name  cot- 
tonwood-willow associations  based  on  cottonwoods 
being  abundant  to  scattered  in  the  community,  this 
information  should  accompany  the  maps.  Or  you 
may  elect  to  name  the  community  as  such  if  about 
50%  of  the  trees  are  cottonwoods  and  the  other  half 
willows.  The  important  point  is  that  good  quantita- 
tive criteria  should  be  established,  adhered  to  in 
mapping,  and  always  accompany  the  maps  as  legends 
or  as  an  appendix.  These  maps  will  be  invaluable  to 
future  managers  as  they  attempt  to  assess  and  under- 
stand habitat  changes.  The  probability  is  high  that 
you  will  be  transferred  in  a  few  years  and  the  person 
replacing  you  must  be  able  to  interpret  the  informa- 
tion that  you  collected  while  working  on  riparian 
systems. 


^>Kr,-.%v- 


mmmm 


WWkMm™ 


A  large  scale  airphoto  (1:2,400)  provides  a  basis  for  map  - 
ping  (B)  brush,  (H)  herbaceous  vegetation,  and  (X)  bare 
ground,  and  (T)  trees  where  present. 


Classifying  Plant  Community  and  Structural 
Types 

Quantify  vegetation  structure  and  tree  species 
composition  in  a  subset  of  available  plant  com- 
munities along  the  stream  under  management  con- 
sideration. These  data  are  important  in  naming 
communities  and  assessing  structural  configurations 
of  vegetation.  As  riparian  ecosystems  are  perturbed 
they  tend  to  support  less  foliage  volume  at  higher 
layers.  For  example,  fires  or  floods  tend  to  remove 
trees,  reducing  canopy  foliage  volume.  Incoming 
shrubs  or  trees  increase  foliage  volume  at  the  herba- 
ceous and  shrub  layers.  Also,  as  man  perturbs  these 
habitats  through  reduction  of  instream  flow  or  in- 
creased silt  concentrations,  the  perturbations  be- 
come manifested  through  vegetation  structure  as 
reduced  foliage  volumes  in  higher  layers  of  vegeta- 
tion or  the  tree  canopy. 

In  our  work  with  community  structural  types, 
we  have  used  structural  Type  I  as  the  mature,  most 
structurally  diverse  community,  and  Type  VI  as  the 
early  stage  of  returning  vegetation.  A  healthy  riparian 
ecosystem  will  contain  mature  plant  communities 
and  structural  types,  intermediate  types,  and  Type  VI 
communities  that  will  eventually  replace  old,  deca- 
dent stands.  In  heavily  grazed  ecosystems,  such  as 


184 


Riparian  Habitats 


those  along  the  Salt  and  Verde  rivers  in  central  Ari- 
zona, replacement  cottonwood  communities  have 
been  virtually  eliminated  and  only  mature,  decadent 
stands  remain.  No  replacement  communities  are 
forthcoming  and  remedial  efforts,  such  as  planting  of 
cuttings,  have  been  attempted  to  provide  replace- 
ment cottonwood  stands.  Although  riparian  ecosys- 
tems can  be  revegetated  more  economically  through 
natural  flooding,  federal  agencies  are  supplementing 
regrowth  through  fencing  and  revegetation  efforts  to 
ensure  riparian  habitats  are  not  lost  forever. 


Censusing  Vertebrates 

Establish  census  lines  for  birds,  snap-trap  small 
mammals  (make  voucher  specimens),  and  install  can- 
trap  grids  for  reptiles  and  amphibians.  Bird  transects 
must  be  censused  no  less  than  nine  times  per  plant 
community  and  structural  type  per  season.  Number 
of  census  transects  for  birds  should  be  in  about 
equal  proportion  to  the  aerial  extent  of  the  commu- 
nity and  structural  types.  Small-mammal  trapping  and 
can  traps  should  also  follow  the  same  protocol. 


Original  type  maps  can  be  modified  to  show 
new  plant  communities  and  structural  types.  Your 
data  on  classification  and  criteria  depicting  structural 
types  should  be  archived  for  subsequent  use  in  the 
area. 


Data  should  be  collected  for  at  least  2  years;  3 
years  is  better.  Seasonal  and  annual  variation  in  ani- 
mal numbers  should  be  determined  before  realistic 
densities  and  habitat  affinities  can  be  made. 


Determining  Vertebrate  Species  Richness 
and  Relative  Abundance  (Breeding) 

Determine  vertebrate  species  richness  and  spe- 
cies composition  for  each  plant  community  and 
structural  type  during  the  breeding  season,  and  year- 
round  when  possible.  Relative  densities  determined 
by  a  gradation  of  rare-to-common  for  each  bird  spe- 
cies is  adequate.  Small  mammals  can  be  assessed 
through  snap  trapping  for  three  consecutive  nights 
and  expressing  each  species'  relative  values  as  X 
number  of  each  species  caught/number  of  trap  nights 
(cumulative  total).  If  small  mammal  trapping  is  un- 
dertaken, voucher  specimens  must  be  prepared  and 
housed  in  a  specimen  collection  where  they  are 
properly  cared  for.  If  certain  community  types  are 
important  to  large  mammals  for  feeding,  thermal 
cover,  or  breeding  grounds,  these  should  be  noted. 


Attempt  to  determine  relative  densities  of  am- 
phibians and  reptiles  through  trap-and-release  with 
can  traps  and  observations.  Values  can  be  based  on 
can-trap  days,  species  and  numbers  observed  per 
hour,  or  some  other  relative  index. 


We  believe  if  managers  concentrated  on  priori- 
ties 1  through  3,  they  would  be  in  a  strong  position 
to  defend  management  decisions  relating  to  domes- 
tic livestock  grazing,  presence  or  absence  of  endan- 
gered species,  wildlife  values  of  community  and 
structural  types,  or  wildlife  values  of  areas  behind 
proposed  dams.  As  it  is,  managers  often  do  not  have 
enough  data  to  demonstrate  the  high  wildlife  value 
of  riparian  ecosystems  that  are  the  most  productive 
wildlife  habitats  in  western  North  America. 


EFFECTS  OF  LAND  MANAGEMENT 
ACTIVITIES  ON  RIPARIAN  SYSTEMS 

Under  the  multiple-use  concept,  managers  must 
be  aware  of  impacts  or  potential  impacts  of  ongoing 
activities  on  the  riparian  resource.  Activities  of  recre- 
ationists,  once  considered  harmless,  now  could  be 
shown  to  have  some  profound  effects  on  some  verte- 
brate groups  and  plant  species  in  the  riparian  zone. 
As  with  other  activities,  such  as  domestic  livestock 
grazing  and  impoundments,  we  know  many  impacts 
can  be  mitigated  to  avoid  or  reduce  some  of  the 
damage  to  riparian  systems.  We  hope  some  of  the 
activities  covered  in  this  section  will  help  you  avoid 
some  of  the  major  management  problems  or  provide 
solutions  to  existing  problems. 


Determining  Vertebrate  Species  Richness 
and  Relative  Abundance  (Winter) 

Determine  vertebrate  species  richness  and  spe- 
cies composition  for  each  plant  community  and 
structural  type  in  winter.  These  values  will  vary  from 
winter  to  winter  depending  on  the  severity  of  the 
winter  and  the  previous  breeding  success  of  the 
overwintering  species.  Rodents  may  hibernate  and 
large  mammals  may  move  to  lower  elevations. 


Domestic  Livestock  Grazing 

Livestock  grazing  is  a  historic  use  on  public 
lands.  Overgrazing  has  been  a  problem  in  many  areas 
of  the  West,  however,  and  overgrazing  in  upland 
areas  has  caused  the  removal  of  virtually  all  nontoxic 
forage  in  adjacent  floodplains.  Reid  and  Pickford 
(1946)  reported  that  cattle  congregate  and  utilize 
riparian  forage  much  more  intensively  than  the  vege- 
tation of  adjacent  ranges.  Biologists  must  be  aware 


Riparian  Habitats 


185 


that  the  riparian  ecosystem  can  potentially  be  much 
more  productive  for  wildlife  under  better  grazing 
strategies.  Although  other  land-management  activities 
have  imposed  losses  or  serious  reductions  in  produc- 
tivity of  wildlife  habitat,  livestock  grazing  may  be 
the  major  factor  negatively  affecting  wildlife  in  the 
1 1  western  states  (Oregon- Washington  Interagency 
Wildlife  Council  1978).  We  agree  with  Skovlin  et  al. 
( 1977)  that  because  of  the  highly  limited  area  of 
riparian  habitats  in  planning  large  pasture  grazing  ap- 
proaches, riparian  systems  are  sacrificed  on  areas  of 
domestic  livestock  grazing  on  most  state  and  federal 
land  today. 


The  value  of  these  riparian  systems  to  domestic 
livestock  grazing  (essentially  the  permittee)  can  only 
be  appreciated  through  forage  production  values. 
Further,  the  difficulty  in  bringing  about  proper  stock- 
ing levels  is  deeply  embedded  in  historical  and  polit- 
ical realms.  There  are  1.62  million  ha  (4  million  a.) 
of  mountain  meadows  at  elevations  between  1 ,890 
and  150  m  (6,200  and  6,400  ft)  in  the  11  western 
states  (U.S.  Department  of  Agriculture,  Forest  Service 
1972).  These  moist  meadows  support  more  beef 
per  acre  than  any  other  range  type  (Skovlin  1984). 
In  northeast  Oregon,  these  meadows  are  so  produc- 
tive that  1  ha  (2.5  a.)  is  equal  in  forage  production 
to  10  to  15  ha  (25  to  38  a.)  of  forested  rangeland 
(Reid  and  Pickford  1946).  Although  only  1  to  2%  of 
the  summer  range  is  meadow  in  the  Pacific  North- 
west, it  potentially  produces  20%  of  the  forage  (Reid 
and  Pickford  1946;  Roath  and  Krueger  1982).  Be- 
cause of  livestock  concentrations,  topographical  con- 
straints on  livestock,  and  distribution  of  water,  the 
forage  in  the  riparian  zone  accounted  for  81%  of 
that  removed  by  livestock  in  the  Blue  Mountain  graz- 
ing allotment  (Roath  and  Krueger  1982). 


Cattle  are  probably  attracted  to  the  riparian 
ecosystem  for  the  same  reasons  as  other  large  mam- 
mals (Ames  1977;  Severson  and  Boldt  1978).  Many 
riparian  plant  species  remain  green  and  succulent 
longer  than  upland  vegetation,  and  sedges  (Carex 
sp.)  contain  higher  sustained  protein  and  energy 
content  than  important  upland  plant  species  (Mc- 
Lean et  al.  1963;  Skovlin  1967;  Paulsen  1969). 


Two  alternatives  that  may  hold  promise  are 
changing  season  of  forage  use  and  changing  kinds 
and  classes  of  livestock.  Siekert  et  al.  (1985)  re- 
ported that  spring  grazing  showed  no  significant 
changes  in  channel  morphology,  whereas  summer 
and  fall  grazing  did.  The  level  of  grazing  (30  cow-calf 
pairs  on  48  ha  [120  a.]  for  10  days)  and  duration 
are  probably  unrealistic  for  most  permittees.  Further, 
in  many  riparian  ecosystems,  most  if  not  all,  tree 
seedlings  would  be  eliminated  under  this  approach. 
Marlow  and  Pogacnik  (1985)  reported  that  grazing 
of  a  riparian  system  in  Montana  after  the  stream- 
banks  had  dried  (<  10%  soil  moisture)  protected 
the  stream  channel  from  damage.  They  recom- 
mended fencing  riparian  habitat,  rest-rotation,  light 
grazing  (20%  forage  removal),  and  grazing  after 
streambanks  have  dried  to  10%  moisture.  Fencing 
has  been  ruled  out  by  Skovlin  ( 1984)  in  that  it  is  too 
expensive  regardless  of  ownership. 

Stocking  moderately  with  steers  or  trying  differ- 
ent breeds  might  reduce  damage  to  riparian  ecosys- 
tems caused  by  cow-calf  operations.  Lactating  cows 
with  calves  appear  to  concentrate  in  areas  with 
green  forage  and  water,  whereas  steers  might  range 
more  into  the  uplands  and  not  concentrate  in  ripar- 
ian habitats.  Neither  of  the  above  approaches  will 
work  unless  stocking  rates  are  reduced  as  well.  Most 
examined  riparian  systems  and  their  watersheds 


Platts  (1984)  listed  eight  major  approaches 
managers  should  consider  in  managing  riparian- 
stream  habitats  under  multiple-use  systems.  These 
range  from  eliminating  grazing  permanently  or  until 
after  recovery  to  rehabilitating  through  revegetation 
and  artificial  stream  structures.  Most  of  these  ap- 
proaches are  unacceptable  to  permittees  today  but 
may  seem  more  reasonable  with  time  or  as  public 
concern  forces  political  action. 


Cowand-calf  herd  in  riparian  zone. 


186 


Riparian  Habitats 


L 


were  unquestionably  overgrazed,  and  use  of  any 
management  approach  without  AUM  reductions  ap- 
pears to  be  a  waste  of  time  and  money.  Streams  in 
overgrazed  pastures  are  characterized  by  being 
wider  and  shallower;  they  contain  more  fine  sedi- 
ment and  have  more  unstable  banks,  less  bank  un- 
dercut, and  higher  summer  water  temperatures 
(Marcuson  1977;  Van  Velson  1979;  Platts  1979, 
1984). 

Large  storm  events  and  the  response  of  riparian 
habitats  have  been  of  interest  to  biologists  because 
of  the  effects  of  those  events  on  riparian  ecosystems 
(Gregory  and  Madew  1982;  Lyons  and  Beschta 
1983).  An  interesting  data  set  has  been  reported  for 
three  streams  under  study  in  Nevada  and  Utah  where 
historically  the  watersheds  of  these  streams  have 
been  heavily  grazed  (Platts  and  Nelson  1983;  Platts  et 
al.  1983).  Peak  flows  (2-14  times  normal)  occurred 
in  Gance  Creek  in  1983  and  1984.  Big  Creek  and 
Chimney  Creek  were  not  grazed,  but  Big  Creek  emp- 
ties into  Bear  Creek.  In  1983,  flows  of  3,630  cfs  ex- 
ceeded all  recorded  flows  for  the  past  40-year 
records  (Platts  et  al.  1985). 

Chimney  Creek  had  received  heavy  grazing  until 
1981  and  showed  little  bank  development.  It  was 
rested  during  1982  and  1983,  and  the  banks  were 
developing  some  overhanging  vegetation  (Platts  et 
al.  1985).  Large  decomposing  aspen  logs  in  the 
stream  and  on  the  banks  were  evidence  of  past  for- 
ests that  once  lined  the  banks  of  Chimney  Creek. 
Aspen  forest  return  has  probably  been  prevented  by 
blowdown,  beaver,  and  heavy  grazing  of  sprouts  by 
livestock  (Platts  et  al.  1985).  The  severe  floods  in 
1983  and  1984  straightened  the  meandering  channel 
and  widened  the  streambank,  reducing  the  develop- 
ing bank  overhanging  vegetation.  The  large  aspen 
logs  that  helped  hold  the  stream  previously  were  de- 
composed or  flushed  from  the  stream  by  the  high 
discharges. 

The  riparian  vegetation  along  Gance  Creek  was 
dominated  by  large  trees  from  the  floodplain  to  the 
streambanks.  Flood  damage  was  previously  vertical 
cutting  and  some  lateral  movement  of  the  channel 
(Platts  et  al.  1985).  Damage  was  minimal  and  the 
authors  believed  that  had  beavers  still  controlled  the 
stream  as  they  had  done  in  the  1950s  and  1960s, 
the  flood  damage  would  have  been  lessened. 

Changes  that  occurred  on  Big  Creek  were  most 
interesting  because  a  portion  of  the  study  stream  had 
been  rested  for  about  10  years  and  was  comparable 
to  two  other  reaches  (one  above  and  one  below  the 
rested  site)  that  had  received  normal  grazing  pres- 
sures. The  rested  section  had  dramatically  recovered 
and  showed  good  floodplain  vegetation  and  stream- 
bank  development.  During  the  flood  years,  stream 
width  in  the  grazed  portions  increased  by  40%  with 
extensive  lateral  stream  movement  and  redeposition 


of  bedload  sediments.  The  rested  section  with  im- 
proved streambanks  was  able  to  contain  the  high 
flows  and  showed  only  a  slight  increase  in  channel 
width  (Platts  et  al.  1985).  Floodplain  vegetation  and 
soils  were  altered  dramatically  in  the  grazed  sections 
following  the  storm  events,  whereas  the  rested  sec- 
tion showed  little  evidence  of  vegetation  change 
or  newly  eroded  sediments.  Results  from  this  study 
amplify  the  admonitions  of  Heede  ( 1985)  that  man- 
agers should  understand  the  interrelationships  be- 
tween vegetation  and  hydrologic  processes  in 
riparian  ecosystems  before  attempting  any  type  of 
management  change  that  alters  these  natural  systems. 

The  economic  values  of  healthy  riparian  ecosys- 
tems and  their  attendant  wildlife  are  difficult  to  es- 
tablish (Everest  1977),  but  approaches  have  been 
made  based  solely  on  fisheries.  Olson  and  Armour 
(1979)  suggested  that  a  hypothetical  reach  of  14,484 
km  (9,000  mi.)  of  depleted  fishable  streams  on  U.S. 
Bureau  of  Land  Management  land  be  set  aside  exclu- 
sively for  recreation.  Based  totally  on  increased  fish- 
ery visitor  days  due  to  restored  habitat,  they 
estimated  a  first-year  benefit-cost  ratio  of  1:1.66.  Or, 
for  every  $1.00  spent  to  fence  the  riparian  corridor, 
there  would  be  $1.66  generated  by  fishermen.  Other 
values  such  as  backpacking,  camping,  bird  watching, 
erosion  control,  and  improved  water  quality  were 
not  included  in  this  economic  return  value. 

The  above  approach  needs  further  study  and 
confirmation  because,  if  true,  the  cost  of  fencing 
could  be  replaced  by  the  economic  return,  making  it 
a  valid  alternative.  If  fencing  is  not  economical 
(Skovlin  1984),  then  it  is  entirely  possible  that  only 
light  or  no  livestock  grazing  of  riparian  systems  is 
the  ultimate  answer  if  maximum  wildlife  productiv- 
ity is  the  management  goal.  However,  Armour 
(1977)  quoted  Harmay  as  stating: 


Vegetation  in  meadows  and  drainages  is 
closely  utilized  (by  domestic  livestock)  under 
any  stocking  rate  or  system  of  grazing  Reduc- 
ing the  livestock  or  adjusting  grazing  season 
usually  will  not  solve  the  problem 


Findings  by  Severson  and  Boldt  (1978)  support 
Harmay  in  that  in  the  northern  Great  Plains  the  ri- 
parian habitats  were  excessively  used,  regardless  of 
stocking  rates.  In  eastern  Oregon,  Gillen  (  1981 ) 
reported  that  at  a  moderate  rate,  the  meadows  only 
produced  3  to  16  times  more  forage,  again  support- 
ing the  contention  that  regardless  of  stocking  rate 
the  riparian  habitats  will  receive  the  greatest  grazing 
pressures.  Under  continuous  grazing  (60-100%  utili- 
zation) in  Nevada  and  Utah,  there  was  no  evidence 
of  riparian  improvement  and  under  present  stocking 
rates  with  continuous  grazing,  the  riparian-stream 
ecosystem  continually  deteriorated  (Platts  1984). 


Riparian  Habitats 


187 


If  not  limited  by  factors  such  as  high  salinity  and 
no  flooding,  riparian  habitat  can  recover  following 
heavy  grazing  (Davis  1977;  Glinski  1977;  Crouch 
1978).  Eliminating  grazing  for  10-12  years  may  be 
necessary  at  higher  elevations  (2,650  m  [4,800  ft]) 
in  willow  communities  where  grazing  pressure  was 
heavy  and  chronic  (Knopf  and  Cannon  1981).  Knopf 
and  Cannon  ( 1981 )  further  pointed  out  that  it  is 
more  difficult  to  improve  a  damaged  riparian  ecosys- 
tem by  eliminating  grazing  than  to  maintain  good 
conditions  in  one  that  is  being  grazed. 

A  number  of  studies  have  shown  a  dramatic 
increase  in  wildlife  values  where  riparian  systems 
were  abused  by  domestic  livestock  grazing;  the  areas 
were  fenced  and  monitored  a  number  of  years  after 
domestic  livestock  removal.  Numbers  of  small  mam- 
mals, songbirds,  and  raptors  increased  by  350%  (Wi- 
negar  1977;  Duff  1979;  Van  Velson  1979)  in  an  area 
fenced  for  8  years  after  grazing.  Game  animals  such 
as  ring-necked  pheasants  (Phasianus  colchicus), 
deer,  and  waterfowl  increased  as  well  (Van  Velson 
1979).  On  the  South  Platte  River  in  Colorado, 
Crouch  (1982)  found  more  ducks,  upland  game,  and 
twice  as  many  terrestrial  birds  in  areas  fenced  for  7 
years  compared  with  adjacent  grazed  habitats.  Signifi- 
cant differences  in  bird  species  richness  and  foraging 
guilds  have  been  reported  between  heavily  grazed 
2.5  cow-calf  units/ha  ( 1/a.)  and  lightly  grazed  (0.3 
cow-calf  units/ha  (0.75/a.)  riparian  habitats.  Total 
density  was  not  significantly  different  indicating  in- 
creases in  some  species  that  were  already  present 
and  the  extirpation  of  some  species  such  as  flycatch- 
ers, ground-foraging  thrushes,  and  foliage-gleaning 
insectivores. 

Small  mammals  are  also  adversely  affected  by 
domestic  livestock  grazing  in  riparian  communities. 
Small  mammal  densities  before  and  after  grazing  with 
a  stocking  rate  of  5  to  6.25  a./AUM  (2.0  to  2.5  ha/ 
AUM)  declined  from  320  to  33/a.  (800  to  83/ha)  in  a 
Douglas  hawthorn  {Crataegus  douglasii >dominated 
community,  from  180  to  24/a.  (450  to  60/ha)  in  a 
riparian  meadow,  and  from  52  to  17/a.  (129  to  42/ 
ha)  in  a  black  cottonwood  (Populus  trichocarpa)- 
mixed  conifer  community.  Ten  months  after  grazing 
ceased,  no  significant  difference  was  found  between 
the  small  mammal  densities  in  the  grazed  versus 
ungrazed  plots. 

Some  grazing  investigators  have  reported  in- 
creased rodent  species  richness  under  moderate  or 
heavy  grazing  pressures  (Moulton  1978).  We  do  not 
doubt  some  of  these  results  but  point  out  that  small 
mammal  species  that  are  added  or  increase  in  num- 
bers are  usually  habitat  generalists  whose  habitat 
requirements  are  broad.  Habitat  specialists,  such  as 
many  microtine  rodents,  are  usually  reduced  or  elim- 
inated when  grazing  pressures  are  high.  Under  these 
grazing  conditions,  species  in  the  genus  Peromyscus 


and  Perognathus  may  increase  or  be  added  because 
the  former  are  generalists  and  the  latter  require 
more  open  habitat. 

Moulton  (1978)  reported  that  moderate  grazing 
(2.3  a./AUM  [0.9  ha/AUM])  for  6  months  in  a  cotton- 
wood  riparian  system  reduced  prairie  vole  (Microtus 
ochrogaster)  numbers  and  increased  deer  mouse 
(Peromyscus  maniculatus)  numbers.  Eight  species 
of  small  mammals  were  trapped  in  the  grazed  plot 
and  four  in  the  ungrazed  control.  The  control  site 
had  been  ungrazed  for  11+  years,  and  the  vegeta- 
tion had  moved  toward  a  uniform,  dense  grass  struc- 
ture, unsuitable  for  a  number  of  rodent  species.  Light 
or  moderate  grazing  would  have  altered  plant  struc- 
ture and  species  composition,  making  the  habitat 
suitable  to  other  rodent  species. 

Where  grazing  can  be  controlled  in  riparian 
habitats  and  seasonally  light-to-moderate  forage  re- 
moval is  practiced,  the  impact  can  be  small  to  ripar- 
ian vegetation  and  wildlife.  But,  as  pointed  out,  by 
incorporating  riparian  areas  into  large  pastures,  these 
productive  wildlife  habitats  become  sacrifice  areas 
where  most,  if  not  all,  of  the  annual  plant  production 
is  removed.  As  suggested  by  May  and  Davis  (1982), 
riparian  habitats  should  be  separated  and  managed  as 
distinct  units. 

Preliminary  data  on  riparian  populations  of  the 
wandering  garter  snake  (Thamnophis  elegans  va- 
grans)  on  fenced  (1972  and  1975)  and  unfenced 
plots  on  the  Rio  de  los  Vacos  near  Santa  Fe,  New 
Mexico,  provided  some  interesting  results  (Szaro  et 
al.  1985).  The  fenced  plots  (only  cattle  excluded) 
supported  a  stand  ( 18%  )  of  trees  and  shrubs  com- 
posed of  thin-leaf  alder  (Alnus  tenuifolia),  irrorata 
willow  (5.  irrorata),  Scouler  willow  (S.  scouleriana), 
coyote  willow,  and  Mexican  cliff-rose  (Cowania 
mexicana).  In  contrast,  0.1%  of  the  grazed  plots 
supported  a  mixture  of  thin-leaf  alder  and  irrorata 
willow.  Herbaceous  ground  cover  (71%  versus 
88%  )  and  down  and  dead  debris  (0.4%  versus  5%  ) 
was  significantly  (P  <  0.05)  different  in  grazed  ver- 
sus ungrazed  plots,  respectively.  Streamside  shrubs  in 
the  fenced  plots  filtered  out  debris  during  floods  to 
form  debris  piles  up  to  4  m  ( 12  ft)  in  diameter  and 
2  m  (6  ft)  high.  These  decomposing  piles  supported 
numerous  worms  and  slugs  that  made  up  62%  and 
18%  of  snakes'  diet.  Snakes  were  five  times  more 
abundant  in  the  ungrazed  versus  the  grazed  plots, 
even  though  they  were  more  difficult  to  find  in  the 
vegetation  and  debris.  Other  species  of  reptiles  are 
undoubtedly  affected  by  domestic  livestock  grazing 
as  foliage  for  insects  and  cover  for  the  reptiles  are 
reduced  or  eliminated. 

Impacts  to  wildlife  by  heavy  domestic  livestock 
grazing  vary  from  moderate  to  extreme  depending 
on  whether  grazing  is  seasonal  or  yearlong.  Seasonal 


188 


Riparian  Habitats 


grazing  may  allow  limited  tree  and  shrub  regenera- 
tion that  provides  some  habitat  and  forage  for  wild- 
life, whereas  heavy,  yearlong  grazing  eventually  leads 
to  removal  of  most,  if  not  all,  of  the  palatable  ripar- 
ian vegetation.  In  the  latter  instance,  forage  and  ther- 
mal cover  for  large  mammals  are  slowly  eliminated 
along  with  food  and  habitat  for  medium-sized  and 
small  mammals,  birds,  and  herps.  In  seasonal  heavy 
grazing,  some  forage  and  thermal  cover  may  be  left 
for  large  mammals,  but  food  and  cover  for  medium- 
sized  and  small  mammals  are  generally  eliminated. 
Mid-canopy  and  understory  birds  may  be  affected  to 
the  point  of  exclusion.  Yearlong,  heavy  grazing  on 
the  Verde  River  in  central  Arizona  has  resulted  in  re- 
maining stands  of  cottonwood-willow  communities 
of  structure  Type  I  or  mature  communities  tending 
toward  decadence  (Higgins  and  Ohmart  1981).  Al- 
though floods  produce  good  seedling  development, 
these  seedlings  are  consumed  before  they  reach 
0.6  m  (  2  ft )  in  height.  Unless  corrective  measures 
are  taken,  in  a  few  years  the  old  decadent  communi- 
ties will  expire  and  there  will  be  no  young  replace- 
ment communities.  Crouch  (1978)  reported  a  50% 
reduction  of  cottonwoods  in  a  grazed  stream  in  Col- 
orado over  an  18-year  period. 

Birds,  for  example  (Ohmart  and  Anderson  1982), 
are  associated  with  four  layers  of  vegetation  on  the 
lower  Colorado  River: 

19  species  are  associated  with  the  7.6  m  (25  ft) 
or  taller  layer, 

10  species  with  the  4.6  to  7.6  m  (15  to  20  ft) 

layer, 

13  species  with  the  1.5  to  4.6  m  (5  to  15  ft) 
layer,  and 

11  species  with  the  0.15  to  1.5  m  (0.5  to  5  ft) 

layer. 


The  overstory  or  canopy  group  (19  species)  are 
specialists  and  were  absent  or  poorly  represented 
when  this  layer  was  absent  or  foliage  density  highly 
reduced.  The  23  species  in  the  two  middle  layers 
and  some  of  the  1 1  species  in  the  understory  group 
were  generalists.  Some  of  these  species  will  be  pres- 
ent even  when  their  foliage  layer  is  absent  or  poorly 
represented.  Heavy  grazing  not  only  affects  the  her- 
baceous and  shrub  layers,  but  over  time  affects  the 
upper  canopy  layers  as  riparian  tree  regeneration 
is  stopped  or  curtailed. 

Domestic  livestock  grazing  in  riparian  habitats 
may  be  used  as  a  management  tool  to  enhance  areas 
for  wildlife.  This  approach  has  potential,  but  one 
would  have  to  know  more  about  the  wildlife  and  its 
habitat  needs  than  presently  known  or  would  have 
to  experiment  with  different  levels  of  forage  removal 
to  bring  about  the  desired  wildlife  results.  Moulton 
( 1978)  suggested  that  grazing  in  a  riparian  area  in 


eastern  Colorado  may  increase  rodent  species  num- 
bers by  creating  microhabitat  diversity.  He  found 
eight  species  of  small  mammals  in  a  grazed  plot  (2.3 
a./AUM  [0.9  ha/AUM])  and  four  species  in  the  un- 
grazed  plot.  The  grazed  area  had  been  stocked  from 
July  through  December  with  spring  use  deferred 
for  17  years.  This  type  of  seasonal  grazing,  after  max- 
imum spring  plant  growth,  allowed  livestock  selec- 
tion of  preferred  species  creating  a  patchy 
microhabitat.  The  ungrazed  plot  ( 1 1  or  more  years 
of  rest)  promoted  a  uniform  vegetation  structure 
without  any  horizontal  patchiness.  Prairie  vole  num- 
bers were  higher  in  the  protected  site  but  not  ex- 
cluded in  the  grazed  site. 


Prairie  vole. 


Another  possible  use  of  cattle  grazing  as  a  man- 
agement tool  has  been  reported  by  Krueger  and 
Anderson  (1985)  in  dense  shrub-willow  communi- 
ties at  high  elevations  on  the  North  Platte  drainage 
in  Wyoming.  Apparently,  fish  populations  are  not 
harmed  by  this  activity,  at  least  on  the  Little  Des- 
chutes River  in  Oregon  (Lorz  1974),  but  this  was 
not  examined  in  the  study  by  Krueger  and  Anderson 
(1985).  Tunnels  created  by  cattle  grazing  through 
the  dense  shrub-willow  altered  plant  community 
structure,  creating  a  more  diverse  set  of  ecological 
conditions  for  birds.  Many  species  respond  positively 
to  this  treatment  except  for  the  green-tailed  towee 
(Pipilo  chlorums),  which  inhabits  dense  stands  of 
vegetation. 

These  management  tools  need  to  be  examined 
carefully  to  ensure  that  wildlife  species  targeted 
for  enhancement  receive  the  benefits.  Domestic  live- 
stock grazing  is  so  entrenched  on  public  land  that 
we  doubt  that  use  of  domestic  livestock  grazing  will 
become  an  important  management  tool  in  many 
ecological  situations.  First,  we  must  reverse  the  gen- 
eral downward  trend  in  riparian  vegetation  condi- 
tions before  cattle  are  needed  as  a  management  tool 


Riparian  Habitats 


189 


on  a  broad  basis.  Fish  habitat  appears  to  be  more 
sensitive  to  livestock  grazing  impacts  than  terrestrial 
habitats. 

Mining 

Mining  can  have  profound  effects  on  riparian 
ecosystems  ranging  from  total  sterility  of  the  riparian 
system  to  intermittent  effects  following  heavy  pre- 
cipitation. The  effects  vary  in  areas  in  the  West  and 
depend  on  the  mineral(s)  being  mined.  In  the  arid 
Southwest,  the  material  is  frequently  sand  and  gravel; 
removal  of  these  products  has  caused  extensive 
channel  cutting,  reduced  water  quality,  and  even 
flooding  and  loss  of  homes.  Ruptured  holding  ponds 
of  leached  materials  from  copper  mines  has  resulted 
in  virtually  permanent  losses  of  the  flora  and  fauna  in 
some  streams. 

Little  advice  other  than  extreme  caution  and  a 
full  knowledge  of  the  mining  operation,  proximity  of 
potential  toxins  to  the  stream,  safeguards  to  avoid 
pollutants,  and  good  common  sense  must  reign  in 
this  situation.  Also,  be  aware  of  potential  and  real 
secondary  impacts. 

An  excellent  approach  is  to  inventory  the  re- 
source intensively  before  any  mining  disturbance,  in- 
cluding seasonal  water  quality  samples  (have  a 
control  stream  if  at  all  possible),  and  then  monitor 
every  3-5  years  thereafter.  The  control  will  provide 
normal  variance  data  and  a  comparison  should  litiga- 
tion ever  occur. 


Recreational  Activities 

Recreational  activity  and  its  effects  on  wildlife 
can  range  from  relatively  minor  to  so  severe  that 
virtually  all  the  vegetation  is  destroyed  locally.  In 
many  instances,  the  agency  developing  the  recrea- 
tional opportunities  in  riparian  habitats  builds  roads 
through  the  habitats  allowing  vehicles  total  access  to 
the  recreational  area.  Consequently,  users  drive  off 
the  roads,  camp  at  random,  and  many  assume  an 
attitude  of  "destroy  anything  you  want,  we  won't 
ever  return."  Wood  gathering  for  firewood  consumes 
down  and  dead  trees,  limbs,  snags,  and  many  times 
standing  live  trees.  Many  forms  of  wildlife  leave  the 
area  and  others,  such  as  lizards,  snakes,  frogs,  and 
salamanders,  are  destroyed  by  children  and  pets. 

The  impacts  of  recreational  use  are  poorly  docu- 
mented, but  simply  by  viewing  some  recreational 
areas  one  is  left  with  the  impression  that  only  the 
heartiest  and  persistent  wildlife  are  left.  Aitchison 
(1977)  studied  bird  densities  and  species  composi- 
tion in  a  seasonal-use  campground  in  Oak  Creek 
Canyon  in  Arizona  for  3  years.  His  control  and  recre- 
ational site  primarily  supported  ponderosa  pine 
(Pinus  ponderosa),  cottonwoods,  and  Arizona  wal- 


nut at  a  1,646-m  (5,400-ft)  elevation.  The  camp- 
ground was  opened  from  about  Memorial  Day  to 
Labor  Day  each  year  which  spans  the  bird  breeding 
season.  In  the  first  year  of  study  (first  year  the  devel- 
oped campground  opened),  there  was  a  40%  de- 
crease in  bird  density  on  opening  day.  Agency 
personnel  destroyed  20%  of  the  nests  of  the  Steller's 
jay  (Cyanocitta  stelleri)  by  removing  and  slashing 
trees.  Aitchison  (1977:178)  reported:  "Campers  de- 
stroyed 30  percent  more  of  the  Steller's  jay  nests 
and  20  percent  of  the  robin  (Turdus  migratorius) 
nests  by  removing  branches  for  firewood,  making 
room  for  tents,  and  other  reasons."  Many  species 
abandoned  their  nests  but  foraged  in  the  camp- 
ground area.  Ultimately,  bird  species  remaining  in 
the  campground  were  larger,  different,  and  fewer 
than  on  the  control  site. 

In  highly  stressed  riparian  systems,  trailer  park 
development  can  be  positive  to  wildlife,  especially 
birds,  as  was  exemplified  in  one  of  our  studies.  Along 
the  Colorado  River,  the  cottonwood-willow  associa- 
tion is  rapidly  disappearing  and  a  wise  developer 
planted  native  trees  in  the  park  for  better  tree  sur- 
vival and  growth,  and  to  attract  birds  for  the  enjoy- 
ment of  the  residents.  This  small  oasis  supports  a  few 
pairs  of  birds  that  were  once  common  along  the 
lower  river  (Grinnell  1914),  but  are  now  rapidly  ap- 
proaching extirpation  (Anderson  and  Ohmart  1977). 
This  type  of  action  should  be  encouraged  in  devel- 
opments where  trees  will  be  protected,  but  man- 
agers must  plan  for  these  developments. 

Riparian  systems  are  very  attractive  to  recrea- 
tionists  in  that  the  systems  contain  water,  interesting 
plants  and  animals,  shade,  and  numerous  other  en- 
joyable features  in  the  otherwise  arid  and  semiarid 
environments.  Hoover  et  al.  (1985)  reported  in  a 
visitor  information  study  that  environmental  attri- 
butes receiving  highest  user  ratings  were  primarily 
ecological  features  present  in  healthy  riparian  eco- 
systems. Managers  should  educate  the  public  of  the 
fragile  nature  and  unique  values  of  these  systems. 
This  may  seem  impossible,  but  there  are  some  guide- 
lines available  such  as  the  Recreational  Carrying 
Capacity  in  the  California  Desert  (U.S.  Department  of 
the  Interior,  Bureau  of  Land  Management  1978)  and 
the  California  Desert  Area  Conservation  Plan  (U.S. 
Department  of  the  Interior,  Bureau  of  Land  Manage- 
ment 1980). 

Martin  (1984)  provided  an  excellent  approach 
in  using  recreation  planning  to  restore  and  protect 
riparian  systems.  He  recommended  ways  to  control 
visitor  use  subtly  and  directly  in  an  intensively  used 
riparian  system  replete  with  wildlife  values,  water 
recreation,  and  large  metropolitan  areas  nearby  on 
the  American  River  in  California.  Approaches  and 
successes  in  the  California  State  Park  System  (Barry 
1984)  and  potential  problems  and  questions  in  the 
proposed  wildlife  enhancement  and  recreational 


190 


Riparian  Habitats 


development  at  Oristimba  Creek  in  California  (Mor- 
ris 1984)  may  also  be  helpful  in  better  planning.  If 
new  campgrounds  are  absolutely  necessary,  Aitchi- 
son's  (1977)  suggestions  may  apply:  locate  new 
campgrounds  in  nonsensitive  areas,  periodically 
close  to  the  campground  to  allow  revegetation  and 
reduce  stress  on  wildlife;  open  the  campground  be- 
fore or  after  the  height  of  the  breeding  season;  con- 
trol visitors  and  agency  habitat  destruction;  and 
educate  the  public  through  signs  showing  good 
camping  procedures.  We  might  add  to  disallow  col- 
lection of  wood  for  any  reason. 


Impoundment  Construction 

Impoundments  are  constructed  for  a  number  of 
reasons  and  some  are  multiple-use  structures.  Gener- 
ally, each  has  regulations  that  outline  its  purpose 
and  function  such  as  flood  control,  hydroelectric 
power,  and  water  storage  for  agriculture  or  munici- 
palities. During  construction,  roads  are  built,  recrea- 
tional facilities  may  be  installed,  and  numerous  other 
secondary  impacts  to  wildlife  may  occur,  along  with 
the  eventual  inundation  of  the  vegetation  in  the  stor- 
age reservoirs.  In  some  instances,  the  secondary 
impacts  of  reservoirs  to  wildlife  can  equal  or  exceed 
primary  impacts. 


Reservoirs.  Depending  on  how  rapidly  water 
surface  levels  fluctuate  behind  reservoirs,  there  may 
be  a  potential  for  productive  wildlife  habitats  to 
develop.  Rapidly  and  wildly  fluctuating  water  levels 
are  not  conducive  to  the  development  of  emergent 
plant  communities  such  as  cattails  (Typha  sp.)  and 
bulrushes  (Scirpus  sp. ).  These  types  of  reservoirs 
develop  a  "bathtub  ring"  for  a  shoreline  where  only 
annual  plants  grow  and  perennials  are  drowned. 
Reservoirs  with  slow  fluctuating  water  levels 
generally  develop  good  emergent  plant  communities 
that  support  animals  such  as  muskrat,  beaver,  rails, 
and  gallinules.  Waterfowl  tend  to  more  heavily  use 
reservoirs  that  have  an  abundance  of  emergent 
vegetation  that  provides  cover  and  greater  foraging 
opportunities.  Reservoirs  with  rapidly  fluctuating 
shorelines  tend  to  attract  only  a  few  diving  ducks. 

Downstream.  The  wildlife  value  of  areas  below 
reservoirs  tends  to  degrade  slowly  over  the  years. 
Generally,  instream  flows  are  lower  than  natural 
flows.  Natural  floods  that  provide  new  soil  deposition 
and  enrichment  are  stopped,  and  riparian  plant 
health  and  vigor  slowly  decline.  If  controlled 
releases  from  the  dam  or  floods  do  occur,  they  are 
generally  greater  than  would  have  occurred  without 
the  dam,  and  the  health  of  the  riparian  system  may 
have  degraded  to  a  point  that  is  no  longer  resilient 
to  a  heavy  flood. 


Recreational  activities  can  affect  riparian  areas. 


Riparian  Habitats 


191 


Another  problem  tends  to  occur  below  dams 
where  instream  flows  are  highly  regulated  and  cessa- 
tion of  natural  floods  prevent  leaching  and  soil  reju- 
venation. Total  Dissolved  Solids  (TDS)  tend  to 
increase  in  areas  where  the  water  table  is  near  the 
soil  surface.  If  the  stream  carries  high  TDS  loads,  the 
process  is  relatively  rapid  and  the  TDS  are  wicked 
to  the  soil  surface  where  they  accumulate  as  the  soil 
moisture  evaporates.  Sodium  or  salt  levels  eventually 
accumulate  to  a  point  that  most  native  species,  ex- 
cept halophytes  (salt-tolerant  plants),  cannot  germi- 
nate or  survive.  Natural  floods  generally  leached 
and  removed  these  deposits,  and  new  soils  were  de- 
posited in  low  areas. 

In  some  situations  below  dams  the  riverbed  is 
essentially  dewatered  until  another  major  perennial 
tributary  enters.  The  Salt  and  Gila  rivers  in  central 
Arizona  and  the  Rio  Grande  from  about  El  Paso  to 
Presidio,  Texas,  are  classic  examples.  Water  return- 
ing to  these  dry  reaches  is  in  one  of  two  forms:  agri- 
cultural waste  waters  high  in  chemicals  and  salts  or 
impoundment  releases  because  reservoirs  are  near  or 
at  capacity.  Agriculture  waste  waters  generally  poi- 
son productive  riparian  vegetation  and  create  condi- 
tions that  favor  growth  of  less  desirable  trees  or 
shrubs.  These  waters  seldom  flow  in  the  original 
channel  and  eventually  the  channel  is  obliterated. 

During  high  rainfall  years,  releases  from  up- 
stream dams  must  occur,  and  the  dense,  low-growing 
trees  and  shrubs  that  cover  the  floodplain  and  block 
the  channel  form  a  living  dam  that  spreads  the  re- 
leased water  laterally  to  inundate  everything  in  the 
floodplain.  Floodwaters  drain  slowly  (generally  tak- 
ing months)  and  frequently  relict,  productive  native 
plant  communities  drown. 


Logging  and  Roads 

We  stated  earlier  that  riparian  ecosystems  can 
be  and  are  affected  by  any  major  perturbation, 
whether  it  be  natural  (fire,  storms)  or  man-made 
(logging,  roadbuilding,  or  grazing).  Therefore,  man- 
agement must  consider  all  disturbances  that  could 
potentially  or  actually  affect  riparian  ecosystems. 
Productive  fisheries  can  be  lost  to  high  stream  sedi- 
ment loads,  to  stream  channel  and  streamside  vegeta- 
tion destruction  by  floods  following  abuse  of 
watersheds,  and  to  perennial  streams  becoming  in- 
termittent because  of  continued  abuse  of  watersheds. 

Logging  and  roadbuilding  on  the  watersheds 
and  near  riparian  systems  destroy  the  natural  ground 
cover  and  churn  and  mix  the  soil  to  produce  trans- 
portable sediment.  Sediment  entering  a  stream 
comes  from  both  natural  and  man-made  activities 
and  its  rate  of  passage  varies  depending  on  slope, 
size  of  area  disturbed,  severity  of  disturbance,  kind 
and  type  of  streamside  vegetation  to  stabilize  the 


transported  materials,  instream  sediment  traps,  and 
the  periodicity  and  duration  of  large  streams.  Dunn 
and  Leopold  (1978)  estimated  that  5  or  more  years 
are  needed  for  a  transportable  sediment  load  to  to- 
tally pass  through  a  stream  system.  Mahoney  and 
Erman  ( 1984)  reported  that  the  sediment  load  cur- 
rently moving  through  any  stream  is  the  product 
of  past  years'  land-use  activity  and  major  storms. 

Leaving  buffer  strips  near  riparian  vegetation  is 
apparently  successful.  Aubertin  and  Patric  (1974) 
studied  a  34-ha  (14-a.)  clearcut  in  West  Virginia  and 
reported  only  slight  increases  in  stream  turbidity 
following  timber  harvest.  They  attributed  the  success 
to  leaving  a  10  to  20-m  (33  to  66-ft)  forested  strip 
adjacent  to  the  riparian  vegetation.  Moring  (1975) 
demonstrated  similar  results  in  his  15 -year  study 
in  Oregon.  He  showed  a  3-8-fold  reduction  in  sus- 
pended stream  sediments  in  the  clearcut  with  a 
buffer  strip  versus  a  clearcut  without  the  buffer  strip. 

A  large  study  on  streams  in  northern  California 
examined  macroinvertebrate  changes  relative  to 
logging  with  buffered  and  unbuffered  strips  (Erman 
et  al.  1977;  Ruby  et  al.  1977;  Newbold  et  al.  1980). 
Where  buffer  strips  were  >  30  m  (  >  98  ft )  on 
logged  sites,  there  were  no  differences  between  in- 
vertebrate populations  in  experimentals  and  con- 
trols. Where  buffer  strips  were  less,  differences  of 
invertebrate  populations  were  detectable  between 
experimentals  and  controls. 

Buffer  strips  also  reduce  pollutants  and  other 
chemical  substances  from  surface  runoff  (Young  et 
al.  1980).  Karr  and  Schlosser  (1977)  extensively 
reviewed  literature  on  the  value  of  near-stream  vege- 
tation on  water  quality  and  stream  biology  and 
should  be  consulted  for  more  in-depth  coverage. 

Not  only  are  buffer  strips  effective  in  reducing 
physical  and,  ultimately,  biological  damage  to  lower 
animals,  they  also  protect  small  mammals  from  inten- 
sive logging  operations  (Cross  1985).  Where  buffer 
or  leave  strips  varying  from  12  to  70  m  (39  to  230 
ft)  wide  were  retained,  those  remaining  habitats 
supported  small  mammal  communities  comparable 
to  undisturbed  sites.  These  studies  were  conducted 
in  southwestern  Oregon  in  mixed-coniferous  riparian 
vegetation.  Harris  (1984)  suggested  maintaining 
riparian  corridors  as  a  means  of  connecting  forest 
habitat  islands  in  similar  stands  of  old-growth  Doug- 
las fir  (Pseudotsuga  menziesii). 

Riparian  Ground-water  Withdrawals 

Ground-water  pumping  may  become  the  most 
serious  threat  to  North  American  riparian  systems. 
Water  diversions  and  reduced  instream  flows  can  be 
devastating  to  riparian  habitats,  but  pumping  of  ripar- 
ian ground  waters  for  industrial  and  municipal  devel- 
opment will  totally  annihilate  most,  if  not  all, 


192 


Riparian  Habitats 


riparian  plant  species.  Extensive  mesquite  bosques 
were  killed  around  Casa  Grande  Ruins  or  the  Casa 
Grande  National  Monument  by  ground-water  pump- 
ing in  central  Arizona  (Judd  et  al.  1971 ).  The  water 
table  has  receded  at  2+  m(7+  ft)  per  year  which, 
in  turn,  resulted  in  the  death  of  this  large  mesquite 
forest. 


Long-term  vegetation  changes  are  well  docu- 
mented by  aerial  photography  for  a  32-km  (2-mi. ) 
reach  of  the  Carmel  River  near  the  Monterey  Penin- 
sula in  California  (Groeneveld  and  Griepentrog 
1985).  Time-series  documentation  for  a  24-year  pe- 
riod (1956-80)  conclusively  demonstrated  a  marked 
decline  in  riparian  trees  such  as  red  willow  (S.  laevi- 
gata), black  cottonwood  {P.  trichocarpa),  California 
sycamore  {Platanus  racemosa),  and  white  alder 
(Alnus  rhombifolia).  Along  with  reduced  riparian 


plant  cover  was  an  invasion  of  weedy  perennials, 
annuals,  and  xerophylic  (arid-adapted)  shrubs.  The 
riverbanks  have  become  noticeably  eroded  due  to 
increased  channel  width. 

Groeneveld  and  Griepentrog  (1985)  cited  other 
unpublished  studies  where  lowering  of  the  water 
table  by  pumping  had  caused  ecological  change.  This 
is  a  real  and  final  threat  to  many  riparian  ecosystems 
in  the  West.  The  U.S.  Geological  Survey  is  currently 
studying  the  effects  of  ground-water  withdrawal  on 
native  riparian  species  along  the  Owens  River  south 
of  Bishop,  California  (Dileanis  et  al.  1985).  Unfortu- 
nately, no  tree  species  were  involved  in  the  study, 
but  data  for  Nevada  saltbush  (Atriplex  torreyi)  and 
rubber  rabbitbush  (Cbrysothamnus  nauseosus)  may 
provide  insight  into  how  sensitive  obligate-riparian 
species  are  to  gradual  and  drastic  declines  of  the 
water  table. 


Riparian  Habitats 


193 


LITERATURE  CITED 


AITCHISON,  S.W.  1977.  Some  effects  of  a  campground  on 
breeding  birds  in  Arizona.  Pages  175-182  in  Johnson, 
R.R.  and  D.A.  Jones,  tech.  coords.  Importance,  Preser- 
vation and  Management  of  Riparian  Habitat:  A  Sym- 
posium. U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep. 
RM-43.  217pp. 

ALMAND,  J.D.  and  W.B.  KROHN.  1978.  The  position  of 
the  Bureau  of  Land  Management  on  the  protection 
and  management  of  riparian  ecosystems.  Pages  359- 
361  in  Johnson,  R.R.  andJ.F.  McCormick,  tech. 
coords.  Strategies  for  Protection  and  Management  of 
Floodplain  Wetlands  and  Other  Riparian  Ecosystems: 
Proc.  of  the  Symposium.  U.S.  Dep.  Agric,  For.  Serv. 
Gen.  Tech.  Rep.  WO-12.,  Washington,  DC.  4 10pp. 

AMES,  C.R.  1977.  Wildlife  conflicts  in  riparian  manage- 
ment: grazing.  Pages  49-51  in  Johnson,  R.R.  and  D.A. 
Jones,  tech.  coords.  Importance,  Preservation  and 
Management  of  Riparian  Habitat:  A  Symposium.  U.S. 
Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep.  RM-43-  217pp. 

ANDERSON,  B.W.  and  R.D.  OHMART.  1977.  Breeding 
bird  use  of  a  residential  development.  Pages  196-201 
in  Annual  Report,  submitted  to  U.S.  Dep.  Inter., 
Bureau  of  Reclamation,  Boulder  City,  NV. 

and ,  1978.  Phainopepla  utilization  of  honey 

mesquite  forests  in  the  Colorado  River  valley.  Condor 
80:334-338. 

and ,  1984.  Vegetation  management  study 


for  the  enhancement  of  wildlife  along  the  lower  Colo- 
rado River.  U.S.  Dep.  Inter.,  Bur.  of  Rec,  Boulder 
City,  NV. 

-,  and  W.C.  HUNTER.  1983.  Quantifying 


variables  for  classifying  desert  riparian  vegetation. 
Pages  32-44  in  Moir,  W.H.  and  L.  Hendzel,  tech. 
coords.,  Proc.  of  the  Workshop  on  Southwestern  Hab- 
itat Types.  U.S.  Dep.  Agric,  For.  Serv.,  Southwest 
Reg.,  Albuquerque,  NM. 

ARMOUR,  C.L.  1977.  Effects  of  deteriorated  range  streams 
on  trout.  U.S.  Dep.  Inter.,  Bur.  Land  Manage.,  Idaho 
State  Off.,  Boise.  7pp. 

ASMUSSEN,  L.E.,  AW.  WHITE,  Jr.,  E.W.  HANSON,  and  J.M. 
SHERIDAN.  1977.  Reduction  of  2,4-d  load  in  surface 
runoff  down  a  grassed  waterway.  J.  Environ.  Qual. 
6:159-162. 

AUBERTIN,  G.M.  and  J.H.  PATR1C.  1974.  Water  quality 
after  clearcutting  in  a  small  watershed  in  West  Vir- 
ginia. J.  Environ.  Qual.  3:243-249. 

AXELROD,  D.I.  1958.  Evolution  of  the  Madro-Tertiary 
geoflora.  Bot.  Rev.  24:433-509. 

BALDA,  R.  1975.  The  relationship  of  secondary  cavity 
nesters  to  snag  densities  in  western  coniferous  for- 
ests. U.S.  Dep.  Agric,  For.  Serv.,  Southwest  Reg.,  Wildl. 
Habitat  Tech.  Bull.  1,  Albuquerque,  NM.  37pp. 

BARRY,  W.J.  1984.  Management  and  protection  of  riparian 
ecosystems  in  the  state  park  system.  Pages  758-766 
in  Warner,  R.E.  and  KM.  Hendrix,  eds.  California 
Riparian  Systems:  Ecology,  Conservation,  and  Produc- 
tive Management.  Univ.  California  Press,  Berkeley. 
1035pp. 

BELLUOMINI,  L.  and  G.  TRAPP.  1984.  Ringtail  distribution 
and  abundance  in  the  Central  Valley  of  California. 
Pages  906-914  in  Warner,  R.E.  and  KM.  Hendrix,  eds. 
California  Riparian  Systems:  Ecology,  Conservation, 
and  Productive  Management.  Univ.  California  Press, 
Berkeley.  1035pp. 


BENOIT,  C.  1978.  Fluvial  sediment  delivery  as  percent  of 
erosion:  The  relationship  between  landslope  and 
effective  streamside  bufferstrip  width.  U.S.  Dep.  Agric, 
For.  Serv.,  Portland,  OR.  (typescript) 
BLACK,  N.,  R.J.  SCHERZINGER,  and  J.W.  THOMAS.  1976. 
Relationship  of  Rocky  Mountain  elk  and  Rocky  Moun- 
tain mule  deer  to  timber  management  in  the  Blue 
Mountains  of  Oregon  and  Washington.  Pages  11-31  in 
Hieb,  S.R.,  ed.  Proc.  Elk-Logging-Roads  Symposium, 
Univ.  Idaho,  Moscow.  142pp. 
BLOOD,  DA.  1966.  Range  relationships  of  elk  and  cattle 
in  Riding  Mountain  National  Park,  Manitoba.  Canadian 
Wildl.  Serv.,  Wildl.  Manage.  Bull.  Ser.  1,19.  62pp. 
BOCK,  J.H.  and  C.E.  BOCK.  1985.  Patterns  of  reproduction 
in  Wright's  sycamore.  Pages  493-494  in  Johnson, 
R.R.,  CD.  Ziebell,  DR.  Patton,  P.F.  Ffolliott,  and  R.H. 
Hamre,  tech.  coords.  Riparian  Ecosystems  and  Their 
Management:  Reconciling  Conflicting  Uses.  Proc.  First 
North  Am.  Riparian  Conf.  U.S.  Dep.  Agric,  For.  Serv. 
Gen.  Tech.  Rep.  RM-120.  523pp. 
BOEER,  W.J.  and  D.J.  SCHMIDLY.  1977.  Terrestrial  mam- 
mals of  the  riparian  corridor  in  Big  Bend  National 
Park.  Pages  212-217  in  Johnson,  R.R  and  DA.  Jones, 
eds.  Importance,  Preservation  and  Management  of 
Riparian  Habitat:  A  Symposium.  U.S.  Dep.  Agric,  For. 
Serv.  Gen.  Tech.  Rep.  RM-43-  217pp. 
BRADY,  W.,  DR.  PATTON,  and  J.  PAXSON.  1985.  The 

development  of  southwestern  riparian  gallery  forests. 
Pages  39-43  in  Johnson,  R.R.,  CD.  Ziebell,  DR.  Pat- 
ton,  R.F.  Ffolliott,  and  R.H.  Hamre,  tech.  coords.  Ri- 
parian Ecosystems  and  Their  Management: 
Reconciling  Conflicting  uses.  Proc.  First  North  Am. 
Riparian  Conf.  U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech. 
Rep.  RM-120.  523pp. 
BRODE,  J.M.  and  R.B.  BURY.  1984.  The  importance  of 

riparian  systems  to  amphibians  and  reptiles.  Pages  30- 
36  in  Warner,  R.E.  and  KM.  Hendrix,  eds.  California 
Riparian  Systems:  Ecology,  Conservation,  and  Produc- 
tive Management.  Univ.  California  Press,  Berkeley. 
1035pp. 
BROWN,  D.E.,  C.H.  LOWE,  and  CP.  PASE.  1979.  A  digi- 
tized classification  system  for  the  biotic  communities 
of  North  America,  with  community  (series)  and  asso- 
ciation examples  for  the  Southwest.  J.  Arizona-Nevada 
Acad.  Sci.  14  (Suppl.  1):1-16. 
BRUSH,  T.,  B.W.  ANDERSON,  and  R.D.  OHMART.  1983- 
Habitat  selection  related  to  resource  availability 
among  cavity-nesting  birds.  Pages  88-89  in  Davis, 
J.W.,  GA.  Goodwin,  and  RA.  Ockenfels,  tech.  coords. 
Snag  Habitat  Management:  Proc.  Symposium.  U.S.  Dep. 
Agric,  For.  Serv.  Gen.  Tech.  Rep.  RM-99.  226pp. 
BUCK,  S.G  1982.  Habitat  utilization  by  fisher  near  Big 
Bear,  California.  M.S.  Thesis,  Humboldt  State  Univ., 
Areata,  CA.  178pp. 
BULL,  E.L.  and  J.N.  SKOVLEN.  1982.  Relationships  be- 
tween anifauna  and  streamside  vegetation.  Trans. 
North.  Am.  Wildl.  Nat.  Resour.  Conf.  47:496-506. 
BURBRIDGE,  W.R.  and  D.J.  NEFF.  1976.  Coconino  Na- 
tional Forest — Arizona  Game  and  Fish  Department  co- 
operative roads — wildlife  study.  Pages  44-57  in  Hieb, 
S.R.,  ed.  Proc  Elk-Logging-Roads  Symposium,  Univ. 
Idaho,  Moscow. 
BURTON,  T.M.  and  G.E.  LIKENS.  1975.  Salamander  popula- 
tion and  biomass  in  the  Hubbard  Brook  experimental 
forest,  New  Hampshire.  Copeia  1975:541-546. 
BURY,  R.B.  1979.  Population  ecology  of  freshwater  turtles 
in  Harless,  M.  and  H.  Morlock,  eds.  Turtles:  Prospec- 


194 


Riparian  Habitats 


tives  and  Research.  John  Wiley  &  Sons,  Inc.,  New 
York,  NY.  695pp. 

CAROTHERS,  S.W.,  R.R.  JOHNSON,  and  S.W.  AITCHISON. 
1974.  Population  structure  and  social  organization 
of  southwestern  riparian  birds  Am.  Zool.  14:97-108. 

CEBALLOS,  G.  1985.  The  importance  of  riparian  habitats 
for  the  conservation  of  endangered  mammals  in  Mex- 
ico. Pages  96-100  in  Johnson,  RR,  CD.  Ziebell,  DR. 
Patton,  P.F.  Ffolliott,  and  R.H.  Hamre,  tech.  coords. 
Riparian  Ecosystems  and  Their  Management:  Recon- 
ciling Conflicting  Uses.  Proc.  First  North  Am.  Riparian 
Conf.  U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep. 
RM  120.  523pp. 

CHAIMSON,  J.F.  1984.  Riparian  vegetation  planting  for 

flood  control.  Pages  120-123  in  Warner,  RE.  and  K.M. 
Hendrix,  eds.  California  Riparian  Systems:  Ecology, 
Conservation,  and  Productive  management.  Univ. 
California  Press,  Berkeley.  1035pp. 

CONNOR,  R.N.,  R.G.  HOOPER,  H.S.  CRAWFORD,  and  H.S. 
MOSBY.  1975.  Woodpecker  nesting  habitat  in  cut  and 
uncut  woodlands  in  Virginia.  J.  Wildl.  Manage.  39:144- 
150. 

CORBETT,  E.S.  andJ.A.  LYNCH.  1985.  Management  of 
streamside  zones  on  municipal  watersheds.  Pages 
187-190  in  Johnson,  R.R,  CD.  Ziebell,  DR.  Patton, 
R.F.  Ffolliott,  and  R.H.  Hamre,  tech.  coords.  Riparian 
Ecosystems  and  Their  Management:  Reconciling  Con- 
flicting Uses.  Proc.  First  North  Am.  Riparian  Conf. 
U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep.  RM-120. 
523pp. 

CROSS,  S.P.  1985.  Responses  of  small  mammals  to  forest 
riparian  perturbations.  Pages  269-275  in  Johnson, 
R.R.,  CD.  Ziebell,  DR.  Patton,  R.F.  Ffolliott,  and  R.H. 
Hamre,  tech.  coords.  Riparian  Ecosystems  and  Their 
Management:  Reconciling  Conflicting  Uses.  Proc.  First 
North  Am.  Riparian  Conf.  U.S.  Dep.  Agric,  For.  Serv. 
Gen.  Tech.  Rep.  RM-120.  523pp. 

CROUCH,  G.L.  1978.  Effects  of  protection  from  livestock 
grazing  on  a  bottomland  wildlife  habitat  in  northeast- 
ern Colorado.  Pages  118-125  in  Graul,  WD.  and  S.J. 
Bissell,  tech.  coords.  Lowland  River  and  Stream  Habi- 
tat in  Colorado:  A  Symposium.  Colorado  Chap.  Wildl. 
Soc  and  Colorado  Audubon  Council.  195pp. 

.  1982.  Wildlife  on  ungrazed  and  grazed  bottom- 
lands on  the  South  Platte  River  in  northeastern  Colo- 
rado. Pages  186-197  in  Wildlife-Livestock 
Relationships  Symposium.  Proc.  10,  Univ.  Idaho,  For. 
Wildl.  Range  Exper.  Sta.,  Moscow,  ID. 

CUMMINS,  K.W.  1974.  Structure  and  function  of  stream 
ecosystems.  BioScience  24:631-641. 

DAVIS,  GA.  1977.  Management  alternatives  for  the  ripar- 
ian habitat  in  the  Southwest.  Pages  59-67  in  Johnson, 
R.R.  and  DA.  Jones,  tech.  coords.  Importance,  Preser- 
vation and  Management  of  Riparian  Habitat:  A  Sym- 
posium. U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep. 
RM-43.  217pp. 

DAVIS,  J.W.,  GA.  GOODWIN,  and  R.A.  OCKENFELS,  tech. 
coords.  1983-  Snag  habitat  management:  Proc.  of 
the  Symposium.  U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech. 
Rep.  RM-99.  226pp. 

DEALY, J.E.,  DA.  LECKENBY,  and  DM.  CONCANNON. 
1981.  Plant  communities  and  their  importance  to 
wildlife.  Pages  1-66  in  Wildlife  Habitats  in  Managed 
Rangelands — The  Great  Basin  of  Southeast  Oregon. 
U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep.  PNW-120. 

DE  VOS,  A.  1951.  Recent  findings  in  fisher  and  marten 

ecology  and  management.  Trans.  16th  North  American 


Wildl.  Conf.  16:498-507. 

1952.  Ecology  and  management  of  fisher  and  mar- 


ten in  Ontario.  Ontario  Dep.  Lands  For.  Tech.  Bull. 
90pp. 

DICKSON,  J.G.,  R.N.  CONNER,  and  J.H.  WILLIAMSON. 
1983  Snag  retention  increases  birds  in  a  clearcut.  J. 
Wildl.  Manage.  47:799-809. 

DILEANIS,  P.D.,  FA.  BRANSON,  and  S.K.  SORENSON. 

1985.  Methods  for  determining  effects  of  controlled 
dewatering  of  shallow  aquifers  on  desert  phreato- 
phytes  in  Owens  Valley,  California.  Pages  197-200  in 
Johnson,  R.R,  CD.  Ziebell,  DR.  Patton,  R.F.  Ffolliott, 
and  R.H.  Hamre,  tech.  coords.  Riparian  Ecosystems 
and  Their  Management:  Reconciling  Conflicting  Uses. 
Proc  First  North  Am.  Riparian  Conf.  U.S.  Dep.  Agric, 
For.  Serv.  Gen.  Tech.  Rep.  RM-120.  523pp. 

DUFF,  DA.  1979.  Riparian  habitat  recovery  on  Big  Creek, 
Rich  County,  Utah.  Pages  91-92  in  Proceedings  of  the 
Forum — Grazing  and  Riparian/Stream  Ecosystems. 
Trout  Unlimited,  Inc.  Vienna,  VA.  94pp. 

DUNN,  T  and  LB.  LEOPOLD.  1978.  Water  in  environmen- 
tal planning.  W.H.  Freeman  Co.,  San  Francisco,  CA. 
818pp. 

ENGEL-WILSON,  R.W.  and  R.D.  OHMART.  1978.  Floral 
and  attendant  faunal  changes  on  the  lower  Rio 
Grande  between  Fort  Quitman  and  Presidio,  Texas. 
Pages  139-147  in  Johnson,  R.R.  and  J.F.  McCormick, 
tech.  coords.  Strategies  for  Protection  and  Manage- 
ment of  Floodplain  Wetlands  and  Other  Riparian 
Ecosystems.  Proc  of  the  Symposium.  U.S.  Dep.  Agric, 
For.  Serv.  Gen.  Tech.  Rep.  WO- 12.  4 10pp. 

ERMAN,  D.C.,  J.D.  NEWBOLD,  and  K.R.  RUBY.  1977.  Eval- 
uation of  streamside  bufferstrips  for  protecting 
aquatic  organisms.  Contr.  165,  California  Water  Re- 
sour.  Center,  Univ.  California,  Davis.  48pp. 

EVANS,  K.E.  and  R.N.  CONNOR.  1979.  Snag  management. 
Pages  214-225  in  DeGraff,  RM.  and  K.E.  Evans,  com- 
pilers. Management  of  North-Central  and  Northeast- 
ern Forests  for  Nongame  Birds.  U.S.  Dep.  Agric,  For. 
Serv.  Gen.  Tech.  Rep.  NC-51. 

EVEREST,  F.H.  1977.  How  to  demonstrate  the  importance 
of  fishery  resources  to  interdisciplinary  planning 
teams.  Fisheries  4:15-19. 

GAINES,  D.  1977.  The  valley  riparian  forests  of  California: 
Their  importance  to  bird  populations.  Pages  57-85 
in  Sands,  A.,  ed.  Riparian  Forests  in  California:  Their 
Ecology  and  Conservation.  Inst.  Ecology,  Univ.  Cali- 
fornia, Davis.  122pp. 

GILLEN,  R.  1981.  1980  progress  report  to  U.S.  Forest 
Service  Pacific  Northwest  Forest  and  Range  Experi- 
ment Station.  Under  Coop.  Aid  Program  218,  PNW 
Suppl.  218.  Oregon  State  Univ.,  Corvallis.  10pp. 

GLINSKI,  R.L.  1977.  Regeneration  and  distribution  of 

sycamore  and  cottonwood  trees  along  Sonoita  Creek, 
Santa  Cruz  County,  Arizona.  Pages  1 16-123  in  John- 
son, R.R.  and  DA.  Jones,  tech.  coords.  Importance, 
Preservation  and  Management  of  Riparian  Habitat:  A 
Symposium.  U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech. 
Rep.  RM-43.  217pp. 

GREGORY,  K.J.  and  JR.  MADEW.  1982.  Land  use  changes, 
flood  frequency  and  channel  adjustment.  Pages  757- 
781  in  Hey,  R.D.,  J.C  Bathurst,  and  C.R.  Thome,  eds. 
Gravel-bed.  John  Wiley  &  Sons,  New  York,  NY. 

GRINNELL,  J.  1914.  An  account  of  the  mammals  and  birds 
of  the  lower  Colorado  Valley  with  special  reference 
to  the  distributional  problems  presented.  Univ.  Cali- 
fornia Publ.  Zool.  12:51-294. 


Riparian  Habitats 


195 


and  T.I.  STORER.  1924.  Animal  life  in  the  Yosemite. 

Univ.  California  Press,  Berkeley.  952pp. 

GROENEVELD,  DP.  and  T.E.  GRIEPENTROG.  1985.  Inter- 
dependence of  groundwater,  riparian  vegetation, 
and  streambank  stability:  A  case  study.  Pages  44-48  in 
Johnson,  R.R.,  CD.  Ziebell,  D.R.  Patton,  P.F.  Ffolliott, 
and  R.H.  Hamre,  tech.  coords.  Riparian  Ecosystems 
and  Their  Management:  Reconciling  Conflicting  Uses. 
Proc.  First  North  Am.  Riparian  Conf.  U.S.  Dep.  Agric, 
For.  Serv.  Gen.  Tech.  Rep.  RM-120.  523pp. 

GRUBB,  T.C,  Jr.,  D.R.  PETIT,  and  D.L.  KRUSAC.  1983- 
Artificial  trees  for  primary  cavity  nesters.  Pages  151- 
154  in  Davis,  J.W.,  G.A.  Goodwin,  and  R.A.  Ockenfels, 
tech.  coords.  Snag  Habitat  Management:  Proceedings 
of  the  Symposium.  U.S.  Dep.  Agric,  For.  Serv.  Gen. 
Tech.  Rep.  RM-99.  226pp. 

HAAPANEN,  A.  1965.  Bird  fauna  of  the  Finnish  forests  in 
relation  to  forest  succession.  I.  Ann.  Zool.  Fenn. 
2:153-196. 

HALLBERY,  D.L.  and  G.R.  TRAPP.  1984.  Gray  fox  temporal 
and  spatial  activity  in  a  riparian/agricultural  zone  in 
California's  Central  Valley.  Pages  920-928  in  Warner, 
R.E.  and  KM.  Hendrix,  eds.  California  Riparian  Sys- 
tems: Ecology,  Conservation,  and  Productive  Manage- 
ment. Univ.  California  Press,  Berkeley.  1035pp. 

HARRIS,  L.D.  1984.  The  fragmented  forest— island  bio- 
geography  theory  and  the  preservation  of  biotic  di- 
versity. Univ.  Chicago  Press,  Chicago,  IL.  211pp. 

HAUPT,  H.F.  1959.  A  method  for  controlling  sediment 
from  logging  roads.  U.S.  Dep.  Agric,  For.  Serv.  Inter- 
mountain  For.  Range  Exp.  Sta.,  Misc.  Paper  22,  Ogden, 
UT.  22pp. 

HEEDE,  B.  1985.  Interactions  between  streamside  vegeta- 
tion and  stream  dynamics.  Pages  54-58  in  Johnson, 
R.R,  CD.  Ziebell,  D.R.  Patton,  R.F.  Ffolliott,  and  R.H. 
Hamre,  tech.  coords.  Riparian  Ecosystems  and  Their 
Management:  Reconciling  Conflicting  Uses.  Proc.  First 
North  Am.  Riparian  Conf.  U.S.  Dep.  Agric,  For.  Serv. 
Gen.  Tech.  Rep.  RM-120.  523pp. 

HEHNKE,  M.  and  CP.  STONE.  1978.  Value  of  riparian 

vegetation  to  avian  populations  along  the  Sacramento 
River  system.  Pages  228-235  in  Johnson,  R.R.  and 
J.F.  McCormick,  tech.  coords.  Strategies  for  Protection 
and  Management  of  Floodplain  Wetlands  and  Other 
Riparian  Ecosystems:  Proc.  of  the  Symposium.  U.S. 
Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep.  WO-12.,  Wash- 
ington, DC.  4 10pp. 

HIGGINS,  A.E.  and  R.D.  OHMART.  1981.  Riparian  habitat 
analysis:  Tonto  National  Forest.  U.S.  Dep.  Agric,  For. 
Serv.,  Tonto  Natl.  For.,  Phoenix,  AZ. 

HINK,  V.  and  R.D.  OHMART.  1984.  Middle  Rio  Grande 
biological  survey  final  report.  U.S.  Army  Corps  of 
Engineers,  Albuquerque,  NM. 

HOOVER,  ST.,  DA.  KING,  and  W.J.  MATTER.  1985.  A 

wilderness  riparian  environment:  Visitor  satisfaction, 
perceptions,  reality,  and  management.  Pages  223- 
226  in  Johnson,  R.R.,  CD.  Ziebell,  D.R.  Patton,  P.F. 
Ffolliott,  and  R.H.  Hamre,  tech.  coords.  Riparian  Eco- 
systems and  Their  Management:  Reconciling  Conflict- 
ing Uses.  Proc.  First  North  Am.  Riparian  Conf.  U.S. 
Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep.  RM-120. 
523pp. 

HUBBARD,  J.P.  1971.  The  summer  birds  of  the  Gila  Val- 
ley, New  Mexico.  Nemouria  2.  35pp. 

HULSE,  A.C  1974.  An  autoecological  study  of  Kinosternon 
soroniense  LeConte  (Chelonia:  Kinosternidae ).  PhD 
dissertation,  Dep.  Zool.,  Arizona  State  University, 


Tempe.  105pp. 

JOHNSON,  R.R.  1978.  The  lower  Colorado  River:  A  west- 
ern system.  Pages  41-55  in  Johnson,  R.R.  and  J.F. 
McCormick,  tech.  coords.  Strategies  for  Protection 
and  Management  of  Floodplain  Wetlands  and  Other 
Riparian  Ecosystems:  Proc.  of  the  Symposium.  U.S. 
Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep.  WO-12.,  Wash- 
ington, DC.  4 10pp. 

and  S.W.  CAROTHERS.  1981.  Southwestern  riparian 

habitats  and  recreation:  Interrelationships  and  impacts 
in  the  Rocky  Mountain  region.  Eisenhower  Consor- 
tium Bulletin.  U.S.  Dep.  Agric,  For.  Serv.  Rocky  Moun- 
tain For.  Range  Exp.  Sta.,  Fort  Collins,  CO. 

and  D.A.JONES,  tech.  coords.  1977.  Importance, 

preservation  and  management  of  riparian  habitat: 
A  symposium.  U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech. 
Rep.  RM-43.  217pp. 

and  J.F.  MCCORMICK,  tech.  coords.  1978.  Strate- 
gies for  protection  and  management  of  floodplain 
wetlands  and  other  riparian  ecosystems.  Proc.  Sym- 
posium. U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep. 
WO-12.  4 10pp. 

,  CD.  ZIEBELL,  D.R.  PATTON,  P.F.  FFOLLIOTT,  and 


R.H.  HAMRE,  tech.  coords.  1985.  Riparian  ecosystems 
and  their  management:  reconciling  conflicting  uses. 
Proc.  First  North  Am.  Riparian  Conf.  U.S.  Dep.  Agric, 
For.  Serv.  Gen.  Tech.  Rep.  RM-120.  523pp. 

JUDD,  J.B.,  J.M.  LAUGHLIN,  H.R.  GUENTHER,  and  R.  HAN- 
DERGRADE.  1971.  The  lethal  decline  of  mesquite 
on  the  Casa  Grande  National  Monument.  Great  Basin 
Nat.  31:153-159. 

KARR,  JR.  and  I.J.  SCHLOSSER.  1977.  Impact  of  near- 
stream  vegetation  and  stream  morphology  on  water 
quality  and  stream  biota.  Env.  Res.  Lab.  EPA-600/3-77- 
097,  U.S.  EPA,  Athens,  GA.  91pp. 

and .  1978.  Water  resources  and  the  land- 
water  interface.  Science  201:229-234. 

KELLERT,  S.R.  1980.  What  do  North  Americans  expect  of 
wildlife  management  agencies?  Proc.  70th  Meeting 
Int.  Assoc.  Fish,  and  Wildl.  agencies,  Washington,  DC. 
9pp. 

KELLY,  G.M.  1977.  Fisher  (Martes pennanti)  biology  in 
the  White  Mountains  National  Forest  and  adjacent 
areas.  Ph.D.  dissertation,  Univ.  Massachusetts,  Am- 
herst. 178pp. 

KENNEDY,  CE.  1977.  Wildlife  conflicts  in  riparian  man- 
agement: grazing.  Pages  49-51  in  Johnson,  R.R.  and 
D.A.  Jones,  tech.  coords.  Importance,  Preservation  and 
Management  of  Riparian  Habitat:  A  Symposium.  U.S. 
Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep.  RM-43-  217pp. 

KNIGHT,  A.W.  and  R.L.  BOTTORFF.  1984.  The  importance 
of  riparian  vegetation  to  stream  ecosystems.  Pages 
160-167  in  Warner,  RE.  and  KM.  Hendrix,  eds.  Cali- 
fornia Riparian  Systems:  Ecology,  Conservation,  and 
Productive  Management.  LIniv.  California  Press,  Berke- 
ley. 1035pp. 

KNOPF,  F.L.  1985.  Significance  of  riparian  vegetation  to 
breeding  birds  across  an  altitudinal  cline.  Pages  105- 
111  in  Johnson,  R.R.,  CD.  Ziebell,  D.R.  Patton,  P.F. 
Ffolliott,  and  R.H.  Hamre,  tech.  coords.  Riparian  Eco- 
systems and  Their  Management:  Reconciling  Conflict- 
ing Uses.  Proc.  First  North  Am.  Riparian  Conf.  U.S. 
Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep.  RM-120. 
523pp. 

and  R.W.  CANNON.  1981.  Structural  resilience  of  a 

willow  riparian  community  to  change  in  grazing  prac- 
tices in  Peek,  J.,  ed.  Livestock- Wildlife  Relationships 


196 


Riparian  Habitats 


Symposium.  Univ.  Idaho,  Moscow. 

KRUEGER,  HO.  and  S.H.  ANDERSON.  1985.  The  use  of 
cattle  as  a  management  tool  for  wildlife  in  shrub- 
willow  riparian  systems.  Pages  300-304  in  Johnson, 
R.R.,  CD.  Ziebell,  DR.  Patton,  P.F.  Ffolliott,  and  R.H. 
Hamre,  tech.  coords.  Riparian  Ecosystems  and  Their 
Management:  Reconciling  Conflicting  Uses.  Proc.  First 
North  Am.  Riparian  Conf.  U.S.  Dep.  Agric,  For.  Serv. 
Gen.  Tech.  Rep.  RM-120.  523pp. 

LANG,  G.E.  and  R.K.  WEIDER.  1984.  The  role  of  beaver  in 
vegetation  patterning  and  development  in  Sphagnum- 
dominated  wetlands  in  West  Virginia.  Bull.  Ecol.  Soc. 
Am.  65(2):243. 

LAURENZI,  A.W.,  R.D.  OHMART,  and  V.  HINK.  1983. 
Classification  of  mixed  broadleaf  riparian  forest  in 
Tonto  National  Forest.  Pages  72-81  in  Moir,  W.H.  and 
L.  Hendzel,  tech.  coords.  Proc.  of  the  Workshop  on 
Southwestern  Habitat  Types.  U.S.  Dep.  Agric,  For. 
Serv.,  Southwest  Reg.,  Albuquerque,  NM. 

LI,  R.M.  and  W.H.  SHEN.  1973.  Effects  of  tall  vegetation 
and  flow  sediment.  J.  Hydraul.  Div.,  ASCE,  Vol.  9,  No. 
HY5,  Proc.  Paper  9748. 

LORZ,  H.H.  1974.  Ecology  and  management  of  brown 

trout  in  Little  Deschutes  River.  Oregon  Dep.  Fish  and 
Wildl.  Fish.  Res.  Rep.  8.  49pp. 

LOWE,  C.H.,  ed.  1964.  The  vertebrates  of  Arizona.  Univ. 
Arizona  Press,  Tucson.  270pp. 

LYONS,  JK.  and  R.L.  BESCHTA.  1983.  Land  use,  floods,  and 
channel  changes,  Upper  Middle  Fork  Williamette 
River,  Oregon  (1936-1980).  Water  Resour.  Res. 
19:436-471. 

MACARTHUR,  R.H.  and  J.W.  MACARTHUR.  1961.  On  bird 
species  diversity.  Ecology  42:594-598. 

MAHONEY,  D.L.  and  D.C.  ERMAN.  1984.  The  role  of 
streamside  buffer  strips  in  the  ecology  of  aquatic 
biota.  Pages  168-176  in  Warner,  R.E.  and  KM.  Hen- 
drix,  eds.  California  Riparian  Systems:  Ecology,  Con- 
servation, and  Productive  management.  Univ. 
California  Press,  Berkeley.  1035pp. 

MANNAN,  R.W.,  E.C.  MESLOW,  and  H.M  WIGHT.  1980. 
Use  of  snags  by  birds  in  Douglas  fir  forests,  western 
Oregon.  J.  Wildl.  Manage.  44:787-797. 

MARCUM,  C.L.  1975.  Summer-fall  habitat  selection  and  use 
by  a  western  Montana  elk  herd.  Ph.D.  dissertation, 
Univ.  Montana,  Missoula.  188pp. 

MARCUSON,  P.E.  1977.  The  effects  of  cattle  grazing  on 
brown  trout  in  Rock  Creek,  Montana.  Project  F-20-R- 
21-1  la.  Montana  Game  and  Fish,  Helena.  28pp. 

MARLOW,  C.B.  and  T.M.  POGACNIK  1985.  Time  of  graz- 
ing and  cattle-induced  damage  to  streambanks.  Pages 
279-284  in  Johnson,  R.R.,  CD.  Ziebell,  D.R.  Patton, 
P.F.  Ffolliott,  and  R.H.  Hamre,  tech.  coords.  Riparian 
Ecosystems  and  Their  Management:  Reconciling  Con- 
flicting Uses.  Proc.  First  North  Am.  Riparian  Conf. 
U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep.  RM-120. 
523pp. 

MARTIN,  KE.  1984.  Recreation  planning  as  a  tool  to  re- 
store and  protect  riparian  systems.  Pages  748-757  in 
Warner,  R.E.  and  KM.  Hendrix,  eds.  California  Ripar- 
ian Systems:  Ecology,  Conservation,  and  Productive 
Management.  Univ.  California,  Berkeley.  1035pp. 

MAY,  B.  and  B.  DAVIS.  1982.  Practices  for  livestock  grazing 
and  aquatic  habitat  protection  on  western  rangelands. 
Pages  271-278  in  Wildlife-Livestock  Relationships 
Symposium:  Proc.  10,  Univ.  Idaho,  Forest,  Wildl., 
Range  Exp.  Sta.,  Moscow. 

MCLEAN,  A.,  H.H.  NICHOLSON,  and  A.L.  VAN  RYSWYK 


1963-  Growth,  productivity  and  chemical  composi- 
tion of  a  subalpine  meadow  in  interior  British  Colum- 
bia. J.  Range  Manage.  16.235-240. 

MEENTS,  JK,  B.W.  ANDERSON,  and  R.D.  OHMART.  1982. 
Vegetation  relationships  and  food  of  sage  sparrows 
wintering  in  honey  mesquite  habitat.  Wilson  Bull. 
94:129-138. 

MEGANK,  R.A.  and  KC  GIBBS.  1979.  A  methodology 
applied  to  the  analysis  of  selected  grazing  manage- 
ment strategy  and  dispersed  recreation.  Final  report 
to  U.S.  Dep.  Agric,  For.  Serv.,  Pacific  Northwest  For. 
Range  Exp.  Sta.  under  Coop.  Aid  Prog.  PNW-40, 
Suppl.  57.  Oregon  State  Univ.,  Corvallis.  67pp. 

MERRIAM,  C.H.  1890.  Results  of  a  biological  survey  of  the 
San  Francisco  Mountains  region  and  desert  of  the 
Little  Colorado  in  Arizona.  U.S.  Dep.  Agric,  North  Am. 
Fauna  3:1-136. 

MORING,  JR.  1975.  The  Alsea  watershed  study:  Effects  of 
logging  on  the  aquatic  resources  of  the  headwater 
streams  of  the  Alsea  River,  Oregon.  Part  II.  Fishery 
Res.  Rep.  9,  Oregon  Dep.  Fish  and  Wildlife,  Corvallis. 
39pp. 

MORRIS,  R.H.  1984.  Planning  recreation  development  and 
wildlife  enhancement  in  a  riparian  environment  at 
Oristimba  Creek.  Pages  767-772  in  Warner,  R.E.  and 
KM.  Hendrix,  eds.  California  Riparian  Systems:  Ecol- 
ogy, Conservation,  and  Productive  Management.  Univ. 
California,  Berkeley.  1035pp. 

MOSCONI,  S.L.  and  R.L.  HUTTO  1982.  The  effects  of 
grazing  on  land  birds  of  a  western  Montana  riparian 
habitat.  Pages  221-233  in  Wildlife-Livestock  Relation- 
ships Symposium:  Proc.  10,  Univ.  Idaho  For.  Wildl. 
Range  Exp.  Sta.,  Moscow. 

MOULTON,  M.  1978.  Small  mammal  associations  in  grazed 
versus  ungrazed  Cottonwood  riparian  woodland  in 
eastern  Colorado:  A  symposium.  Pages  133-140.  Colo- 
rado Chap.  Wildl.  Soc.  and  Colorado  Audubon  Coun- 
cil, Greeley.  195pp. 

MULLIS,  C.  1985.  Habitat  utilization  by  fisher  {Martes 

pennanti)  near  Hayfork  Bally,  California.  M.S.  Thesis, 
Humboldt  State  Univ.,  Areata,  CA. 

MURPHY,  ML.  and J.D.  HALL.  1981.  Varied  effects  of 
clearcut  logging  on  predators  and  their  habitat  in 
small  streams  of  the  Cascade  Mountains,  Oregon.  Can. 
J.  Fish.  Aquat.  Sci.  38:137-145. 

NELSON, JR.  and  D.G.  BURNELL.  1975.  Elk-cattle  compe- 
tition in  central  Washington.  Pages  71-83  in  Range 
Multiple  Use  Management.  Univ.  Idaho,  Moscow. 

NELSON,  L,  Jr.  and  JK.  HOOPER.  1976.  California  fur 

bearers  and  their  management.  Univ.  California,  Coop. 
Ext.  Serv.,  Leaflet  2721,  Berkeley. 

NEWBOLD, J.D.,  DC.  ERMAN,  and  KB.  RUBY.  1980.  Ef- 
fects of  logging  on  macroinvertebrates  in  streams 
with  and  without  bufferstrips.  Can.  J.  Fish.  Aquat  Sci. 
37:1076-1085. 

NUSSBAUM,  R.A.  1974.  The  distributional  ecology  and  life 
history  of  the  Siskiyou  Mountain  salamander  {Pletho- 
don  stormi),  in  relation  to  the  potential  impact  of  the 
proposed  Applegate  Reservoir  on  this  species.  Rep. 
submitted  to  U.S.  Army  Corps  of  Engineers,  Portland, 
OR.  52pp. 

OHMART,  R.D.  and  B.W.  ANDERSON.  1974.  Vegetation 
management:  Annual  report.  Submitted  to  U.S.  Dep. 
Inter.,  Bureau  of  Reclamation,  Boulder  City,  NV. 

and .  1982.  North  American  desert  riparian 

ecosystems.  Pages  433-479  in  Bender,  G.L.,  ed.  Refer- 
ence Handbook  on  the  Deserts  of  North  America. 


Riparian  Habitats 


197 


Greenwood  Press,  Westport,  CT.  594pp. 

OLSON,  R.W.  and  C.L.  ARMOUR.  1979.  Economic  consid- 
erations for  improved  livestock  management  ap- 
proaches for  fish  and  wildlife  in  riparian/stream  areas. 
Pages  67-71  in  Cope,  O.B.,  ed.  Forum — Grazing  and 
Riparian/Stream  Ecosystems.  Trout  Unlimited,  Inc. 
Vienna,  VA.  94pp. 

OREGON-WASHINGTON  INTERAGENCY  WILDLIFE 

COUNCIL.  1978.  Managing  riparian  zones  for  fish  and 
wildlife  in  eastern  Oregon  and  eastern  Washington. 
Unpubl.  rep. 

ORR,  R.T.  1940.  The  rabbits  of  California.  Occ.  Papers. 
California  Acad.  Sci.  19:1-227. 

PARKER,  M.,  F.  WOOD,  Jr.,  B.H.  SMITH,  and  R.G.  ELDER. 
1985.  Erosional  downcutting  in  lower  order  riparian 
ecosystems:  Have  historical  changes  been  caused 
by  removal  of  beaver?  Pages  35-38  in  Johnson,  R.R., 
CD.  Ziebell,  DR.  Patton,  P.F.  Ffolliott,  and  R.  H. 
Hamre,  tech.  coords.  Riparian  Ecosystems  and  Their 
Management:  Reconciling  Conflicting  Uses.  Proc.  First 
North  Am.  Riparian  Conf.  U.S.  Dep.  Agric,  For.  Serv. 
Gen.  Tech.  Rep.  RM-120.  523pp. 

PAULSEN,  H.A.,  Jr.  1969.  Forage  value  on  a  mountain 

grassland  aspen  range  in  western  Colorado.  J.  Range 
Manage.  22:102-107. 

PEDERSON,  R.J.,  AW.  ADAMS,  and  J.  SKOVLIN.  1979.  Elk 
management  in  Blue  Mountain  habitats.  Oregon  Dep. 
Fish  and  Wildl.,  Portland.  27pp. 

PETERSON,  AW.  and  T.C.  GRUBB,  Jr.  1983-  Artificial  trees 
as  a  cavity  substrate  for  woodpeckers.  J.  Wildl.  Man- 
age. 47:790-798. 

PLATTS,  W.S.  1979.  Livestock  grazing  and  riparian/stream 
ecosystems — An  overview.  Pages  39-45  in  Cope,  O.B., 
ed.  Proc.  of  the  Forum — Grazing  and  Riparian/Stream 
Ecosystems.  Trout  Unlimited,  Inc.,  Vienna,  VA.  94pp. 

.  1984.  Riparian  system/livestock  grazing  interaction 

research  in  the  intermountain  West.  Pages  424-429 
in  Warner,  R.E.  and  KM.  Hendrix,  eds.  California 
Riparian  Systems:  Ecology,  Conservation,  and  Produc- 
tive Management.  Univ.  California,  Berkeley.  1035pp. 

,  KA.  GEBHARDT,  and  W.L.  JACKSON.  1985.  The 

effects  of  large  storm  events  on  basin-range  riparian 
stream  habitats.  Pages  30-34  in  Johnson,  R.R.,  CD. 
Ziebell,  DR.  Patton,  P.F.  Ffolliott,  and  R.  H.  Hamre, 
tech.  coords.  Riparian  Ecosystems  and  Their  Manage- 
ment: Reconciling  Conflicting  Uses.  Proc.  First  North 
Am.  Riparian  Conf.  U.S.  Dep.  Agric,  For.  Serv.  Gen. 
Tech.  Rep.  RM-120.  523pp. 

and  R.L.  NELSON.  1983-  Population  and  generic 

differentiation  in  the  Humboldt  cutthroat  trout  of 
Gance  Creek,  Nevada.  Pages  15-19  in  California- 
Nevada  Wildl.  Trans. 

,  O.  CASEY,  and  V.  CRISPIN.  1983.  Riparian-stream 

habitat  conditions  on  Tabor  Creek,  Nevada,  under 
grazed  and  ungrazed  conditions.  Pages  10-14  in  Proc. 
Ann.  Conf.  Western  Assoc.  Fish  and  Wildl.  Agencies. 
63:10-14. 

REID,  EH.  and  G.D.  PICKFORD.  1946.  Judging  mountain 
meadow  range  condition  in  eastern  Oregon  and  east- 
ern Washington.  U.S.  Dep.  Agric,  Circular  748. 

RICE,  J.,  B.W.  ANDERSON,  and  R.D.  OHMART.  1984.  Com- 
parison of  the  importance  of  different  habitat  attri- 
butes of  avian  community  organization.  J.  Wildl. 
Manage.  48:895-911. 

ROATH,  L.R.  and  W.C  KRUEGER.  1982.  Cattle  grazing 
influence  on  a  mountain  riparian  zone.  J.  Range  Man- 
age. 35:100-104. 


RUBY,  KB.,  DC.  ERMAN,  and  J.D.  NEWBOLD.  1977.  Bio- 
logical assessment  of  timber  management  activity 
impacts  and  bufferstrip  effectiveness  on  National  For- 
est streams  of  northern  California.  Earth  Resources 
Monogr.  1,  U.S.  Dep.  Agric,  For.  Serv.,  Reg.  5,  San 
Francisco,  CA.  170pp. 

SALWASSER,  H.  and  K  SHIMAMOTO.  1984.  Pronghorn, 
cattle  and  feral  horse  use  of  wetland  and  upland  habi- 
tats. Pages  210-213  in  Warner,  RE.  and  KM.  Hendrix, 
eds.  California  Riparian  Systems:  Ecology,  Conserva- 
tion, and  Productive  Management.  Univ.  California, 
Berkeley.  1035pp. 

SCOTT,  LB.  1984.  Economic  values  of  three  furbearers 
inhabiting  California  riparian  systems.  Pages  731-738 
in  Warner,  RE.  and  KM.  Hendrix,  eds.  California 
Riparian  Systems:  Ecology,  Conservation,  and  Produc- 
tive Management.  Univ.  California,  Berkeley.  1035pp. 

SCOTT,  V.E.  1979.  Bird  responses  to  snag  removal  in  pon- 
derosa  pine.  J.  Forestry.  77:26-28. 

,  KE.  EVANS,  DR.  PATTON,  and  CP.  STONE.  1977. 

Cavity-nesting  birds  of  North  American  forests.  U.S. 
Dep.  Agric,  For.  Serv.  Agric.  Handbook  511,  Washing- 
ton, DC.  112pp. 

SEVERSON,  KE.  and  CE.  BOLDT.  1978.  Cattle,  wildlife, 
and  riparian  habitats  in  the  western  Dakotas.  Pages 
94-103  in  Management  and  Use  of  Northern  Plain 
Rangeland.  Reg.  Rangeland  Symp.,  Bismarck,  ND. 

SIEKERT,  RE.,  Q.D.  SKINNER,  MA.  SMITH,  J.L.  DODD, 
and  J.D.  RODGERS.  1985.  Channel  response  of  an 
ephemeral  stream  in  Wyoming  to  selected  grazing 
treatments.  Pages  276-278  in  Johnson,  R.R.,  CD. 
Ziebell,  DR.  Patton,  P.F.  Ffolliott,  and  R.  H.  Hamre, 
tech.  coords.  Riparian  Ecosystems  and  Their  Manage- 
ment: Reconciling  Conflicting  Uses.  Proc.  First  North 
Am.  Riparian  Conf.  U.S.  Dep.  Agric,  For.  Serv.  Gen. 
Tech.  Rep.  RM-120.  523pp. 

SKOVLIN,  J.M.  1967.  Fluctuations  in  forage  quality  on 

summer  range  in  the  Blue  Mountains.  U.S.  Dep.  Agric, 
For.  Serv.  PNW-Res.  Paper  44. 

.  1984.  Impacts  of  grazing  on  wetlands  and  riparian 

habitat:  A  review  of  our  knowledge.  Pages  1001- 
1 1 04  in  Committee  on  Developing  Strategies  for 
Rangeland  Management.  Natl.  Resour.  Council/Natl. 
Acad.  Sci.,  Westview  Press,  Boulder,  CO. 

,  R.W.  HARRIS,  G.S.  STRICKLER,  and  G.A.  GARRI- 
SON. 1976.  Effects  of  cattle  grazing  methods  on  pon- 
derosa  pine-bunchgrass  range  in  the  Pacific 
Northwest.  U.S.  Dep.  Agric,  Tech.  Bull.  1531.  Wash- 
ington, DC.  40pp. 

W.R.  MEEHAN,  J.C  BUCKHOUSE,  and  M.  VAVRA. 


1977.  A  method  of  study  for  determining  the  influ- 
ence of  grazing  on  riparian  and  aquatic  habitats  in  the 
Blue  Mountains  of  Oregon.  Pages  164-169  in  Live- 
stock Interactions  with  Wildlife,  Fisheries,  and  Their 
Environments.  Proceedings  of  the  Symposium.  U.S. 
Dep.  Agric,  For.  Serv.  Pacific  Southwest  For.  Range 
Exp.  Sta.,  Berkeley,  CA. 

SMITH,  F.  1 977.  A  short  review  of  the  status  of  riparian 
forests  in  California.  Pages  1-2  in  Sands,  A.,  ed.  Ripar- 
ian Forests  in  California:  Their  Ecology  and  Conserva- 
tion. Inst.  Ecol.,  Publ.  15,  Univ.  California,  Davis. 
122pp. 

STAMP,  N.  and  R.D.  OHMART.  1979.  Rodents  of  desert 
shrub  and  riparian  woodland  habitats  in  the  Sonoran 
Desert.  Southwestern  Nat.  24:279-289- 

SULLIVAN,  B.K  1981.  Distribution  and  relative  abundance 
of  snakes  along  a  transect  in  California.  J.  Herpert. 


198 


Riparian  Habitats 


15:247-248. 

SWANSON,  F.J.,  S.V.  GREGORY,  JR.  SEDELL,  and  AG. 

CAMPBELL.  1982.  Land-water  interactions:  The  ripar- 
ian zone.  Pages  267-291  in  Analysis  of  Coniferous 
Forest  Ecosystems  in  the  Western  United  States.  US/ 
IBP  Synthesis  Series  14,  Hutchinson  Ross  Publ.  Co., 
Stroudsburg,  PA. 

SZARO,  R.C.,  S.C.  BELFIT,  J.K  AITKIN,  and  J.N.  RINNE. 
1985.  Impact  of  grazing  on  a  riparian  garter  snake. 
Pages  359-363  in  Johnson,  R.R.,  CD.  Ziebell,  DR. 
Patton,  P.F.  Ffolliott,  and  R.  H.  Hamre,  tech.  coords. 
Riparian  Ecosystems  and  Their  Management:  Recon- 
ciling Conflicting  Uses.  Proc.  First  North  Am.  Riparian 
Conf.  U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep. 
RM-120.  523pp. 

THOMAS,  J.W.,  C.  MASER,  and  J.E.  RODIEK  1979.  Wildlife 
habitat  in  managed  rangelands — the  Great  Basin  of 
southeastern  Oregon-riparian  zone.  U.S.  Dep.  Agric, 
For.  Serv.  Gen.  Tech.  Rep.  RNW-80.  18pp. 

,  R.J.  MILLER,  H.  BLACK,  J.E.  RODIEK  and  C. 

MASER.  1976.  Guidelines  for  maintaining  and  enhanc- 
ing wildlife  habitat  in  forest  management  in  the  Blue 
Mountains  of  Oregon  and  Washington.  Trans.  North 
Am.  Wildl.  Nat.  Resour.  Conf.  41:452-476. 

U.S.  COUNCIL  ON  ENVIRONMENTAL  QUALITY.  1978. 
Environmental  quality.  The  ninth  report  of  the  Coun- 
cil on  Environmental  Quality.  U.S.  Govt.  Printing 
Office,  Washington,  DC.  (Stock  No.  041-011-00040- 
8).  599pp. 

U.S.  DEPARTMENT  OF  AGRICULTURE,  FOREST  SER 
VICE.  1972.  Western  regional  working  conference — 
results  of  work  group  sessions — delegate  meetings. 
National  Program  of  Research  for  Forest  and  Associ- 
ated Rangelands.  U.S.  Dep.  Agric,  For.  Serv.  Inter- 
mountain  For.  Range  Exper.  Sta.,  Ogden,  UT.  39pp. 

U.S.  DEPARTMENT  OF  THE  INTERIOR,  BUREAU  OF 
LAND  MANAGEMENT.  1978.  Recreational  carrying 
capacity  in  the  California  Desert.  U.S.  Dep.  Inter.,  Bur. 
Land  Manage.,  Sacramento,  CA.  115pp. 

.  1980.  The  California  desert  conservation  area  plan. 

U.S.  Govt.  Printing  Office,  San  Francisco,  CA. 

U.S.  DEPARTMENT  OF  THE  INTERIOR,  BUREAU  OF 
RECLAMATION.  1979.  Pecos  River  Basin  water  sal- 
vage project:  Final  environmental  statement.  U.S.  Dep. 
Inter.,  Bur.  Rec,  Southwest  Reg.  Off.,  Amarillo,  TX. 

VAN  VELSON,  R.  1979.  Effects  of  livestock  grazing  upon 
rainbow  trout  in  Otter  Creek,  Nebraska.  Pages  53-55 
in  Cope,  O.B.,  ed.  Proc.  of  the  Forum — Grazing  and 


Riparian/Stream  Ecosystems.  Trout  Unlimited,  Inc., 
Vienna,  VA.  94pp. 

WARNER,  RE.,  recorder.  1979a.  Fish  and  wildlife  resource 
needs  in  riparian  ecosystems:  Proc.  of  a  workshop. 
Nat.  Water  Resour.  Analysis  Group,  Eastern  Energy 
and  Land  Use  Team,  U.S.  Dep.  Inter.,  Fish  Wildl.  Serv., 
Kearneysville,  WV.  53pp. 

.  1979b.  California  riparian  study  program:  Back- 
ground information  and  proposed  study  design.  Plan- 
ning Branch,  California  Dep.  Fish  and  Game, 
Sacramento.  177pp. 

and  KM.  HENDRIX,  eds.  1984.  California  riparian 


systems:  Ecology,  conservation,  and  productive  man- 
agement. Univ.  California  Press,  Berkeley.  1035pp. 

WAUER,  R.H.  1977.  Significance  of  Rio  Grande  riparian 

systems  upon  the  avifauna.  Pages  165-174  in  Johnson, 
R.R.  and  D.A.  Jones,  tech.  coords.  Importance,  Preser- 
vation and  Management  of  Riparian  Habitat:  A  Sym- 
posium. U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep. 
RM-43.  217pp. 

WEBSTER'S  NEW  COLLEGIATE  DICTIONARY.  1979. 
G.&C.  Merriam  Co.,  Springfield,  MA.  1532pp. 

WILLIAMS,  D.F.  1985.  Mammalian  species  of  special  con- 
cern in  California.  California  Dep.  Fish  and  Game, 
Nongame  Wildl.  Invest.,  Final  Rep.,  Project  E-W-4, 
IV-14.1.  Sacramento.  184pp. 

and  KS.  KILBURN.  1984.  Sensitive,  threatened,  and 

endangered  mammals  of  riparian  and  other  wetland 
communities  in  California.  Pages  950-957  in  Warner, 
R.E.  and  KM.  Hendrix,  eds.  California  Riparian  Sys- 
tems: Ecology,  Conservation,  and  Productive  Manage- 
ment. Univ.  California  Press,  Berkeley.  1035pp. 

WINEGAR,  H.H.  1977.  Camp  Creek  channel  fencing- 
plant,  wildlife,  soil,  and  water  response.  Rangeman's  J. 
4:10-12. 

WINTERNITZ,  B.L.  1973.  Ecological  patterns  in  a  montane 
breeding  bird  community.  Ph.D.  dissertation,  Univ. 
Colorado,  Boulder. 

1976.  Temporal  change  and  habitat  preference  of 

some  montane  breeding  birds.  Condor  78:383-393. 

and  H.  CAHN.  1983.  Nestholes  in  live  and  dead 


aspen.  Pages  102-106  in  Davis,  J.W.,  G.A.  Goodwin, 
and  R.A.  Ockenfels,  tech.  coords.  Snag  Habitat  Man- 
agement: Proceedings  of  the  Symposium.  U.S.  Dep. 
Agric,  For.  Serv.  Gen.  Tech.  Rep.  RM-99.  226pp. 
YOUNG,  R.A.,  T  HUNTRODS,  and  W.  ANDERSON.  1980. 
Effectiveness  of  vegetated  bufferstrips  in  controlling 
pollution  from  feedlot  runoff.  J.  Environ.  Qual.  9:483- 
487. 


Riparian  Habitats 


199 


10 


MARSHES 


Milton  W.  Weller 

Department  of  Wildlife  &  Fisheries  Sciences 
Texas  A&M  University 
College  Station,  TX      77843 


Editor's  Note:  Wetlands  compose  a  small  percentage 
of  lands  in  the  ivestern  U.S.,  but  their  importance 
for  wildlife  far  outweighs  their  acreage.  In  this  pub- 
lication, marshes  are  considered  to  be  wetlands  as- 
sociated with  standing  water.  Among  the  most 
complex  habitats,  marshes  are  therefore  most  diffi- 
cult to  inventory  or  monitor.  As  the  author  empha- 
sizes, evaluation,  impact  assessment,  and  manage- 
ment decisions  require  first-hand,  mud-on-your- 
boots  experience  in  the  marsh. 


INTRODUCTION 


"Greater  familiarity  with  marshes  on  the  part  of 
more  people  could  give  man  a  truer  and  a  more 
wholesome  view  of  himself  in  relation  to  Nature. 
. . .  Marshes  comprise  their  own  form  of  wilderness. 
They  have  their  own  life-rich  genuineness  and  reflect 
forces  that  are  much  older,  much  more  permanent, 
and  much  mightier  than  man." 


-Paul  L.  Errington,  from  Of  Men  and 
Marshes 


Marshes,  the  dynamic  meeting  places  of  land 
and  water,  are  among  the  most  productive  and  excit- 
ing of  all  wildlife  habitats.  Not  only  do  they  attract  a 
great  diversity  of  wildlife,  but  numbers  of  some  spe- 
cies can  be  overwhelming.  As  a  result,  the  value  of 
marshes  for  recreation  such  as  hunting  or  bird 
watching  can  outweigh  those  activities  on  far  larger 
tracts  of  other  habitats.  Moreover,  their  values  in 
damping  flood  surges,  in  water  purification,  and  in 
water  recharge  make  marshes  highly  important.  In- 
ventory and  monitoring  are  essential  to  identify-  criti- 
cal habitats  and  resource  values.  With  this  knowl- 
edge biologists  can  develop  management  alternatives 
or  mitigation  plans  on  these  numerically  restricted 
but  highly  valued  habitats. 

Wetlands  are  characterized  by  hydric  soils  and 
water-loving  plants  (Cowardin  et  al.  1979).  Of  the 
many  diverse  types  of  wetlands,  marshes  are  the 
most  widely  distributed  and  the  best-known  form. 
They  are  dominated  by  emergent  plants  such  as  cat- 
tail (Typha  sp. ),  bulrushes  (Scirpus  sp. ),  sedges 
{Carex  sp. ),  and  water-tolerant  grasses.  The  general 
height  and  structure  of  such  plant  communities  dis- 
tinguish marshes  from  swamps  (wooded  and  shrub- 
dominated  wetlands )  or  from  bogs  ( moss  and  heath- 
covered  wet  organic  soils ).  Whereas  swamps  and 
bogs  are  restricted  in  distribution  by  climatic  re- 
gime, marshes  may  be  found  anywhere.  They  are  the 
characteristic  wetland  of  intermountain  lowlands  of 
the  western  U.S.  where  they  form  a  true  oasis  for 
wildlife,  both  aquatic  and  terrestrial. 

Marshes  often  are  complete  entities,  found  in 
shallow  basins.  The  term  may  also  be  used  for  any 
emergent  hydrophyte  community.  The  emphasis 
here  is  on  shallow  basins  that  are  marshes  in  the 
geomorphic  as  well  as  biological  sense. 

Water  is  the  driving  force  in  determining  wet- 
land type  and  habitat  quality.  Water  permanency  and 
associated  vegetation  are  key  factors  in  classifying 
wetlands.  Because  water  cycles  are  variable,  marshes 
are  rarely  constant.  These  fluctuations  induce  "boom 
and  bust"  in  wildlife  numbers,  but  are  essential  to 
nutrient  recycling.  In  shallow  marshes  such  as 


Marshes 


201 


ephemeral  prairie  potholes,  dense  vegetation  pro- 
duced during  natural  drawdowns  is  the  major  food 
for  detritus-feeding  macroinvertebrates  (those  larger 
aquatic  invertebrates  visible  to  the  unaided  eye) 
which  are  in  turn  important  food  resources  for  wa- 
terbirds  and  fish  (McKnight  and  Low  1969;  Krapu 
1974).  In  deeper  marshes  or  shallow  lakes,  drying  al- 
lows decomposition  of  bottom  organic  deposits  that 
provide  nutrients  for  a  new  surge  of  vegetation  that 
germinates  under  the  low  water  conditions. 

The  character  of  marshes — water,  sediment,  or- 
ganic matter,  and  dense  vegetation — makes  walking 
or  boating  difficult.  The  tendency  is  to  circle  around 
the  area  rather  than  go  through  it.  But  evaluation, 
impact  assessment,  and  management  decisions  re- 
quire first-hand  experience  in  the  marsh.  It  is  hoped 
that  this  chapter  will  provide  the  inducement  to  gain 
this  experience. 


Bureau  of  Biological  Survey  employees,  1935. 


Because  of  the  dynamics  of  water  and  season, 
observations  of  conditions  or  populations  in  a  single 
year  have  relatively  little  meaning  except  when 
viewed  in  relation  to  the  following:  ( 1 )  past  records 
of  the  same  marsh,  or  (2)  a  predicted  potential 
based  on  knowledge  of  similar  wetlands  for  which 
data  are  available.  Long-term  monitoring  is  vital  to 
understanding  the  quality,  importance,  and  condition 
of  a  marsh. 


Losses  of  wetlands  occur  in  many  ways.  Drain- 
age for  agriculture  is  the  major  cause  of  loss  of 
marshes  (Frayer  et  al.  1983).  Filling  of  small  marshes 
for  agriculture  and  other  uses  occurs  occasionally, 
but  even  large  marshes  have  been  filled  for  urban 
and  industrial  development.  Because  marshes  that 
have  constant  water  levels  become  lake-like,  perma- 
nent flooding  may  have  serious  detrimental  impacts 
on  marsh  wildlife  (Weller  1981b).  The  impact  of  to- 
tal loss  is  obvious.  Less  measurable  negative  influ- 
ences include  modified  water  flow,  increased  eutro- 
phication  or  turbidity,  and  introduction  of  exotic 
animals  or  plants.  These  impacts  are  both  long-term 
and  widespread,  but  they  are  difficult  to  assess  with- 
out ecological  and  biological  baseline  data  for  mak- 
ing comparisons. 

A  number  of  wildlife  species  inhabiting  western 
marshes  are  classified  as  endangered,  mostly  because 
of  disappearing  habitats.  Three  subspecies  of  Clapper 
Rail — Yuma  (Rallus  longirostris  yumaensis),  Califor- 
nia (R.  I.  obsoletus),  and  Light-footed  (R  I.  levipes) — 
are  endangered  due  to  loss  of  river  or  coastal  marsh. 
The  salt  marsh  harvest  mouse  (Reithrodontomys  ra- 
viventris)  now  occurs  in  only  a  few  scattered 
marshes  in  San  Francisco  Bay.  Wetlands  are  impor- 
tant not  only  for  local  wildlife  populations  but  also 
for  migratory  birds,  a  national  and  international  re- 
source. Protection,  restoration,  and  management  of 
marshes  are  essential  to  restore  these  species. 


Yuma  clapper  rail. 


202 


Marshes 


CLASSIFICATION  SYSTEMS  AND 
DESCRIPTORS  FOR  MARSHES 

Interest  in  wetlands  has  been  dominated  by  bi- 
ologists studying  semiaquatic  animal  and  plant  life. 
Acquisition  of  marshes  was  promoted  by  waterfowl 
hunters,  fur  trappers,  and  fishery  managers.  Different 
species  of  wildlife  were  associated  with  different 
types  of  marshes,  and  efforts  to  preserve  species 
through  acquisition  or  management  required  a 
method  of  choosing  different  wetlands  and  assessing 
how  many  of  each  existed.  Hence,  biologists  devel- 
oped classification  systems  to  broadly  define  and  in- 
ventory wetlands  (e.g.,  Martin  et  al.  1953). 

A  classic  document  by  Shaw  and  Fredine  ( 1956) 
outlined  20  types  of  wetlands  across  the  U.S.,  de- 
scribed their  use  by  wildlife,  and  estimated  the  ex- 
tent of  losses  since  the  earliest  surveys.  A  significant 
advancement  was  made  in  classification  precision  by 
a  detailed  regional  study  in  the  North  Dakota  prairie 
pothole  region  (Stewart  and  Kantrud  1971 ).  This 
system  is  strongly  influenced  by  data  on  water  per- 
manence and  vegetation  patterns.  The  system  de- 
vised by  Millar  ( 1979)  for  western  Canada  also  used 
data  on  basin,  area,  watershed,  and  other  limnologi- 
cal  variables  to  develop  a  precise  classification  sys- 
tem that  could  be  used  when  wetlands  were  dry  or 
devoid  of  vegetation.  Data  on  wildlife  generally  have 
not  been  used  as  a  means  of  defining  or  clarifying 
types  of  wetlands  except  by  Bergman  et  al.  ( 1977) 
for  Alaskan  tundra  wetlands. 

With  the  advent  of  widespread  interest  in  wet- 
land values  for  food  chain  support  systems,  flood  and 
erosion  control,  and  energy  and  sewage  treatment,  a 
broader  classification  system  was  needed.  An  inter- 
agency task  force,  headed  by  the  U.S.  Fish  and  Wild- 
life Service,  devised  a  system  ranging  from  the  gen- 
eral to  more  and  more  specific.  Known  as  the 
"Classification  of  Wetlands  and  Deepwater  Habitats 
of  the  United  States"  (Cowardin  et  al.  1979),  it  is 
being  used  for  the  National  Wetland  Inventory.  It  is 
hierarchical,  like  a  taxonomic  key.  All  wetlands  are 
first  placed  into  one  of  five  "systems"  by  means  of 
ecological  terms  of  broad  usage: 


Palustrine 

vegetated  wetlands  such  as 
marshes 

Estuarine 

brackish  water  which  may  include 
marshes  or  swamps 

Lacustrine 

lake-like 

Riverine 

stream-related  but  not  including 
marshes  or  oxbow  lakes 

Marine 

deepwater  habitats 

These  systems  are  then  divided  into  subsystems 
except  for  Palustrine  wetlands.  The  Palustrine  Sys- 
tem is  divided  into  units  called  Classes  which  are 
based  on  dominant  vegetative  type.  A  marsh  is 
termed  an  Emergent  Wetland  as  opposed  to  a  Moss- 
lichen  Wetland  (bog)  or  a  Forested  Wetland 
(swamp).  Further  subdivisions  allow  more  and  more 
precise  descriptors,  so  that  Emergent  Wetlands  can 
be  divided  into  subclasses  by  the  presence  of  persist- 
ent (lasting  over  winter  or  more)  plants  such  as  cat- 
tail or  bulrushes,  to  the  nonpersistent  (rapidly  de- 
composing) plants  such  as  arrowhead  (Sagittaria 
sp.)  or  dock  {Rutnex  sp.).  The  precise  plant  species 
or  community  is  referred  to  as  dominance  type, 
which  allows  a  rather  precise  taxonomic  description 
of  the  community.  However,  the  habitat  significance 
of  these  descriptors  is  not  evident  from  the 
classification. 


On  a  local  basis,  some  of  the  earlier  classifica- 
tion systems  are  still  in  use  and  may  be  extremely 
helpful  in  wildlife  evaluations.  They  should  be  used 
as  an  addition  to  rather  than  replacement  for  the 
interagency  system. 

For  the  western  U.S.,  especially  the  glaciated 
pothole  region,  the  Stewart  and  Kantrud  (1971)  sys- 
tem is  useful  because  of  the  use  of  common  terms 
related  to  hydroperiods.  Classes  I  through  V  reflect 
pattern  of  flooding  or  mean  water  depth  and  are  self 
explanatory:  Ephemeral,  Temporary,  Seasonal, 
Semipermanent,  and  Permanent.  Special  wetland 
classes  include  VI  Alkali  ponds  and  shallow  lakes 
rich  in  submergents  and  VII-Fens  (or  alkaline  bogs). 
Subclasses  were  used  to  denote  five  ranges  from 
fresh  to  subsaline. 

Because  wetlands  are  so  dynamic  in  water  level, 
a  wetland  of  the  same  class  can  range  from  bare  mud 
flat  to  emergent  vegetation  through  the  action  of 
water  levels  and  herbivores  (Weller  and  Spatcher 
1965;  Weller  and  Fredrickson  1974).  Special  termi- 
nology reflecting  those  short-term  successional 
stages  was  used  by  Stewart  and  Kantrud  ( 1971 )  to 
assist  in  detailed  descriptions:  emergent  phase,  open- 
water  phase,  drawdown  phase,  and  bare-soil  phase — 
the  last  two  inducing  germination  conditions  that 


Marshes 


203 


ultimately  result  in  reestablishment  of  emergent 
vegetation. 

Another  set  of  descriptive  terms  denotes  the 
typical  combination  of  plants  found  under  mean  pre- 
vailing water  conditions  as  well  as  topography.  Mil- 
lar's ( 1979)  system  uses  traditional  terms  for  vegeta- 
tion zones  of  large  wetlands,  or  for  classifying  a 
marsh  by  the  vegetation  of  the  central  part  of  the 
marsh: 

1.  wet  meadow 

2.  shallow  marsh 

3.  emergent  deep  marsh 

4.  open  water  marsh 

5.  shallow  open  water 

6.  open  alkali  wetland 

7.  disturbed  areas  (due  to  drought  or  agricultural 
activities). 

A  wet-meadow  community  is  made  up  mostly  of 
nonpersistent  annuals  and  is  commonly  flooded  only 
in  spring.  A  deep  marsh  is  dominated  by  those  per- 
sistent perennials  in  the  center  that  do  well  with 
regular  flooding,  but  the  edge  will  have  species  char- 
acteristic of  wet  meadow  or  shallow  marsh. 

Most  classification  and  inventory  systems  deal 
with  one  basin  or  a  part  of  one  basin.  Yet  values  for 
wildlife  and  fish  are  enhanced  by  the  proximity  of 
several  different  wetlands.  The  importance  of  conti- 
guity has  not  been  documented,  but  birds  and  mam- 
mals move  from  area  to  area.  Adult  fish  that  move 
from  lakes  to  wetlands  to  spawn,  or  fry  that  mature 
in  nursery  marshes,  must  have  connecting  links  (For- 
ney 1968).  Such  relations  sometimes  can  be  de- 
tected from  aerial  photographs  or  detailed  maps,  but 
the  extra  value  to  be  placed  on  such  spatial  relation- 
ships is  uncertain.  Data  need  to  be  gathered  and  ana- 
lyzed from  this  perspective  because  acquisition,  pro- 
tection, and  management  of  marshes  would  be 
strongly  influenced  by  this  information. 

One  of  the  most  difficult  descriptive  decisions  of 
wetlands  is  the  demarcation  of  the  wetland  bound- 
ary. Marsh  edges  often  grade  imperceptibly  into  the 
uplands,  and  plants  at  the  marsh  edge  may  be  toler- 
ant of  both  wet  and  dry  conditions.  Moreover,  the 
hydrologic  influence  of  a  wetland  may  affect  upland 
vegetation  since  increased  water  availability  near  the 
wetland  may  encourage  or  maintain  terrestrial  vege- 
tation that  might  otherwise  not  exist. 

Coastal  wetlands  may  have  the  additional  com- 
plicating variables  of  lunar  tidal  regime,  wind  and 
storm  surges,  and  salinity.  These  too  are  dynamic 
processes  with  tides  varying  seasonally,  and  fresh- 
water inflow  varying  with  rainfall.  At  times,  estuarine 
marshes  may  be  totally  fresh  due  to  high  rates  of  in- 
flow, whereas  marine  storm  tides  may  modify  and 


even  reverse  the  situation  periodically.  As  a  result,  a 
much  reduced  flora  dominated  by  grasses  such  as 
salt  marsh  cordgrasses  {Spartina  sp.)  and  saltgrass 
{Distichlis  spicata)  characterizes  vast  areas  of 
coastal  marsh.  Mean  tide  ranges  have  produced 
zones  of  high  ( less  commonly  flooded )  and  low 
(wet)  marsh  based  on  the  relative  degree  of  flooding 
and  the  influence  of  salt  water  (Nixon  1982;  Zedler 
1982). 


MAJOR  SPECIES  GROUPS 

The  availability  of  water,  diverse  plants,  and  rich 
invertebrate  life  make  marshes  ideal  for  many  spe- 
cies of  wildlife  and  fish.  Some  species  evolved  with 
more  permanent  water  and  are  equipped  with  gills 
for  total  water  emersion  throughout  life  (fish  and 
some  salamanders )  or  for  part  of  their  life  cycle 
(most  salamanders,  frogs,  and  toads).  Others  are 
semiaquatic,  being  air  breathers  throughout  life  as  a 
secondary  adaptation  to  fluctuating  water  regimes 
found  in  marshes.  Some  species  have  highly  modified 
feet  (beaver,  Castor  canadensis,  and  river  otters,  Lu- 
tra  canadensis)  or  flattened  tails  for  swimming 
(muskrats,  Ondatra  zibethicus).  Some  avian  species 
have  long  legs  for  wading  (herons  and  storks),  or 
adaptions  for  swimming  and  diving  (ducks  and 
grebes ).  Still  other  species  retain  greater  range  of 
habitat  by  means  of  less  inhibiting  adaptations  to 
water,  but  are  common  around  marshes  ( mink,  Mus- 
tela  vision,  and  raccoons,  Procyon  lotor).  All  de- 
grees of  variation  are  found  between  these  extremes. 
Some  species  frequent  marsh  areas  and  exploit  their 
resources  without  being  intimately  associated.  Swal- 
lows and  bats,  for  example,  catch  insects  there  but 
do  not  use  marshes  for  nesting  or  cover.  Some  ter- 
restrial species  may  invade  during  drawdown  stages 
to  seek  out  stranded  invertebrates  or  fish.  Red  foxes 
(Vulpes  vulpes)  take  advantage  of  muskrats  unpro- 
tected by  water  or  lodges. 

A  review  of  some  common  taxonomic  groups 
will  facilitate  a  view  of  the  more  complex  communi- 
ties of  vertebrates  common  to  marshes  of  the  west- 
ern U.S.  For  further  detail,  see  Weller  (  1979a). 

Fish 

Considering  both  freshwater  (about  95)  and  es- 
tuarine (about  109)  species,  wetlands  serve  over 
200  species  on  a  national  basis  (Adamus  1983)-  Inte- 
rior palustrine  wetlands  may  have  several  species  or 
none  at  all.  The  most  abundant  are  small  minnow- 
like fishes  that  feed  on  tiny  invertebrates,  such  as 
mosquito  fish  (Gambusia  sp. ).  Other  species  include 
herbivores  or  omnivores  like  crappie  (Pomoxis  sp.) 
and  bluegill  (Lepomis  macrochirus),  herbivore-detri- 
tivores  like  carp  (Cyprinus  carpio);  and  carnivores 
like  walleye  (Stizostedium  vitreum),  pike  (Esox  In- 
dus), and  largemouth  bass  (Macropterus  sal- 


204 


Marshes 


moides ).  The  exact  complement  varies  with  size  and 
permanence  of  the  water.  Association  with  source 
streams  may  contribute  to  pioneering.  Oxygen  short- 
age is  perhaps  the  most  important  limiting  influence 
because  marshes  may  be  oxygen  depleted  during  the 
summer. 

Amphibians 

Dominant  amphibians  range  from  the  most 
aquatic  sirens  {Siren  sp. )  that  have  gills  throughout 
life,  to  frogs  that  breathe  air  as  adults.  Some  frogs 
stay  in  or  near  water  always  (bullfrogs,  Rana  cates- 
beiana ).  Others  stay  near  water  but  feed  in  wet 
meadows  and  grassy  areas  (northern  leopard  frogs, 
Rana  pipiens ).  A  few  are  highly  terrestrial  species 
that  associate  with  moist  forest  most  of  the  year 
(tree  frogs,  Hyla  sp. ).  The  most  terrestrial  amphibi- 
ans are  toads  (Bufo  sp. )  that  range  in  habitat  from 
moist  uplands  to  near-desert.  Amphibians  are  impor- 
tant as  food  for  various  fish,  reptile,  avian,  and  mam- 
malian predators. 


Reptiles 


Although  the  species  vary  locally,  turtles  are  the 
most  widespread  and  often  abundant  reptilian  preda- 
tor or  omnivore  in  marshes.  Some  like  the  musk  tur- 
tles (Sternotherns  sp. ),  mud  turtles  (Kinostemon 
sp),  cooters  and  sliders  (Pseudemys  sp. )  loaf  on  logs 
or  vegetation  in  the  sun  and  are  conspicuous.  A  few 
are  true  marsh  turtles,  found  nowhere  else:  Bind- 
ing's turtle  {Emydoidea  blandingii)  and  chicken  tur- 
tles {Deirochelys  reticularia).  Others  like  the  com- 
mon ( Chelydra  sp. )  and  alligator  snapping  turtles 
(Macroclemys  temminckii)  may  be  inconspicuous 
until  they  move  into  uplands  during  egg-laying  or 
are  exposed  during  droughts.  Snakes,  especially 
water  snakes  of  the  genus  Nerodius,  are  widespread. 
The  massasauga  (Sistrurus  catenatus)  occurs  in 
either  swamps  or  wet  prairies.  Some  species  like  rat 
snakes  (Elapha  sp. )  and  garter  snakes  ( Thamnophis 
sp. )  go  into  any  habitat,  including  marshes,  for  birds 
and  their  eggs. 


Flycatchers 
Warblers 


Geese  •* 

iVxi  Dabblers 
.  ^  Sparrows  ^ 
W  ^ 

Blackbirds 
(Red  Winged)  (Yellow-Headed) 


Redheads,  Ruddy  Ducks 
Swans 


-+■  Mergansers 


*WWl 

Sedge)  (Marsh) 


■^w  Wrens 


Gulls,  Terns 


Rails 
Egrets  -  Herons 
Bitterns 
Short-Eared  Owls 


Gallinules 
Snail  Kites 


Moorhens,  Coots 
Grebes 


MUD  FLAT 


Harriers 
-Cranes 
Shorebirds         SniPes 


Loons,  Pelicans 


Shorebirds 
Waders 


WET 
MEADOW 


SHALLOW 
MARSH 


UNVEGETATED 
SHALLOWS/ 
FLATS 


Figure  1.  Waterbirds  of  wetlands — with  emphasis  on  habitat  selection  for  breeding  and  feeding. 


Marshes 


205 


Birds 

Over  1 40  species  of  birds  are  associated  with 
wetlands  intimately  or  casually  (Adamus  1983).  Like 
fish,  but  with  still  greater  flexibility,  birds  may  use 
marshes  only  part  of  the  year  to  fulfill  certain  life 
functions.  These  include  summer  breeding  and  feed- 
ing sites  (see  Figure  1 ),  molting  sites  (ducks),  feed- 
ing sites  away  from  the  nest  (pelicans  and  waders), 
migration  stops  (shore  birds,  ducks,  and  geese),  food 
exploration  zones  (most  waders  moving  north  after 
breeding),  and  wintering  (all  migratory  species).  Al- 
though breeding  sites  tend  to  be  valued  more  highly, 
each  area  plays  a  vital  role  in  the  life  cycle  of  mobile 
species.  All  are  essential. 

Mammals 


Compared  with  other  classes  of  vertebrates,  few 
mammals  are  semiaquatic  and  very  few  are  adapted 


to  marshes.  Lack  of  species  numbers  is  compensated 
for  by  numbers  of  individuals.  Several  small  rodents 
characteristic  of  damp  areas  swim  easily  and  do  not 
avoid  water  or  flooded  areas:  marsh  rice  rat  (Oryzo- 
mys palustris)  in  Gulf  Coast  marshes  (Esher  et  al. 
1978),  and  California  vole  (Microtus  californicus)  in 
tidal  marshes  (Fisler  1961).  Both  are  omnivorous. 
Beavers  may  create  their  own  marsh  by  damming  of 
streams,  but  also  use  large  marshes  and  build  lodges 
of  willows  and  cattail. 

Communities  of  Vertebrates 

Species  associations  and  numbers  are  influenced 
by  ( 1 )  plant  structure,  (  2  )  size  and  diversity  of  the 
unit,  and  (3)  water  depth  and  permanence.  Hence,  a 
wet  meadow  dominated  by  nonpersistent,  temporary 
plants  and  water  cover  is  likely  to  be  used  for  breed- 
ing by  smaller  species  requiring  only  low  cover 
(sparrows,  mice,  rats,  shore  birds,  rails)  or  by  tran- 
sients using  the  area  for  food  (muskrats,  shorebirds, 


HABITAT  LAYERS 


REPRESENTATIVE  GUILDS 

F       B  F      B  F       B 


Midstory 


Terrestrial  surface 


Bottom  of  water  column 


F  -  Feeding 

B-  Breeding  (Nesting] 


1  2  3 

Mallard        Redhead      Muskrat 


Pintail 


Canvas- 
back 

Ruddy 
Duck 


Relationship  between  habitat  layers  and  guilds. 


206 


Marshes 


waders,  dabbling  ducks ).  Shallow  marshes  may  har- 
bor nesting  king  rails  (Rallus  elegans),  purple  gallin- 
ules  (Porpbyrula  martinica)  or  common  moorhens 
(Gallinula  chloropus),  and  species  of  blackbirds 
that  vary  by  geographic  area.  More  persistent  and 
taller  vegetation  of  deep  marshes  support  larger  spe- 
cies that  need  cover,  support  for  nests,  and  the  foods 
induced  by  more  permanent  water  (diving  ducks, 
swans,  cranes,  some  geese,  American  coots  [Fulica 
atra],  bitterns,  and  wader  colonies). 

Because  many  vertebrates  may  use  the  general 
structural  layer  of  vegetation,  ecologists  often  use 
the  term  guild  to  reflect  taxonomically  unrelated 
species  that  use  the  same  habitat  in  the  same  way 
(Short  and  Burnham  1982).  In  marshes,  bird  nests 
may  be  built  at  three  layers:  ( 1 )  in  tall  and  robust 
emergents,  (2)  in  short  and  nonpersistent  emergents, 
and  (  3 )  at  the  water's  surface.  For  feeding,  birds  may 
forage  in  all  three  of  the  nesting  layers,  and  divers 
may  seek  foods  in  the  water  column  or  on  the 
bottom. 

Surface  feeding  birds  that  specialize  on  foliage 
include  gadwall  (Anas  strepera)  and  American  wid- 
geon (Anas  americana).  Northern  shovelers  (Anas 
clypeata)  can  feed  in  the  open  because  they  can 
strain  plankton  or  in  shallow  water  where  they  may 
eat  snails  or  seeds.  Phalaropes  prey  on  plankton 
forms  by  swimming  and  picking.  Blue-winged  teal 
(Anas  discors),  however,  are  typical  of  many  dab- 
blers that  have  generalized  bill  structure  and  food 
habits.  These  species  can  take  invertebrates  in  sum- 
mer when  hens  need  protein  for  eggs  and  young 
need  it  for  body  growth  (Swanson,  et  al.  1979),  or 
seeds  and  foliage  in  nonbreeding  periods.  Predators 
include  mergansers,  grebes  and  loons,  and  others 
that  feed  on  fish  in  open  marshes.  Mink  may  occur 
around  or  in  any  of  these  wetland  types,  but  dens 
are  most  common  by  semipermanent  water,  presum- 
ably because  of  the  food  supply. 


CRITICAL  HABITAT  FEATURES 

To  appreciate  the  features  of  a  marsh  that  make 
it  valuable  to  wildlife,  it  is  essential  to  relate  these 
components  to  wildlife  needs.  Regardless  of  the  cli- 
matic conditions  or  the  season,  wildlife  often  seek 
shelter  from  sun,  wind,  rain,  and  hail.  Cover  also  is 
essential  for  protection  from  predators.  No  wonder 
predatory  mink  and  weasels  are  long-bodied,  and 
aerial  predators  like  the  northern  harrier  (Circus  cy- 
aneus )  hover  to  follow  and  catch  their  prey. 

The  concentrated  nutrients  of  the  marsh  pro- 
duce high  primary  productivity  in  plants.  The  sec- 
ondary' production  of  invertebrates  produces  an  ar- 
ray of  potential  food  exceeded  in  few  if  any  habitats. 

The  major  herbivores  in  marshes  are  mammals 


like  muskrats  which  not  only  eat  leaves  and  tubers  of 
various  emergent  plants  but  build  lodges  for  winter 
use,  thereby  cutting  vast  quantities  of  emergents. 
Geese,  especially  Snow  Geese  (Anser  coerulescens), 
also  may  utilize  tubers  and  rootstocks  in  wintering 
areas.  Swans,  widgeons,  gadwalls,  and  coots  devour 
huge  quantities  of  submergent  vegetation.  Addition- 
ally, many  bottom  (benthic)  invertebrates  utilize  de- 
composing plant  matter,  or  feed  on  tiny  organisms 
like  bacteria  and  fungi  that  live  on  old  stems.  Inver- 
tebrate abundance  must  be  directly  related  to  plant 
productivity  at  an  earlier  stage,  but  there  has  been 
little  effort  at  quantitating  such  energy  flow  in  fresh- 
water marshes. 

Marsh  vegetation  is  vital  for  many  species  of 
birds  that  nest — often  colonially — in  marshes:  ibis, 
egrets,  herons,  grebes,  ducks,  gulls,  etc.  Diverse  veg- 
etation, whether  it  be  for  food,  cover,  or  nest  sub- 
strate, clearly  influences  the  potential  richness  of 
species  found  in  the  system  by  its  various  heights, 
coverage,  foods  potential,  and  food-chain  support  for 
other  lower  organisms. 

Water  Chemistry  and  Plant  Communities 

Water  that  enters  a  marsh  is  a  product  of  pre- 
cipitation and  hydrologic  processes.  Much  of  the 
water  may  enter  the  system  from  surface  runoff  and 
slow  moving  underground  flows  at  various  depths. 
The  timing  of  water  availability  (hydroperiod)  and 
depth  determine  the  character  and  type  of  marsh 
through  influence  on  chemical  characteristics  of  the 
water  and  subsequently  on  vegetation  and  inverte- 
brates used  by  wildlife. 

Typically,  rainwater  carries  with  it  few  dissolved 
chemicals  except  where  it  collects  wind-driven  soil 
particles  or  salt  spray.  Water  that  flows  over  soil  or 
rock  strata  or  through  porous  substrates  may  dis- 
solve and  carry  a  wide  range  of  dissolved  solids 
which  may  serve  as  plant  nutrients  or,  in  extreme  in- 
stances, be  so  concentrated  as  to  be  toxic  to  life. 
Thus,  general  characters  of  the  water  in  marshes  and 
lakes  reflect  the  character  of  the  watershed  as  shown 
by  Moyle  (1945)  in  Minnesota  and  Metcalf  (1931) 
in  North  Dakota.  Moyle's  work  spanned  more  diverse 
water  chemistry  and  plant  types,  which  he  classified 
as  ( 1 )  soft  waters  (low  dissolved  salts)  of  northeast- 
ern Minnesota  Laurentian  Shield  country,  (2)  hard 
water  morainic  lakes  of  central  and  southern  areas, 
and  (  3  )  alkali  lakes  of  western  Minnesota.  He 
pointed  out  the  plant  community  that  reflects  each 
of  these  types  is  distinctive  (Table  1 ),  but  it  is  a 
product  of  individual  tolerance  to  chemical  charac- 
teristics, as  shown  by  Bourn  (1935)  and  others  on 
salt  tolerance.  Significant  measured  features  were  to- 
tal alkalinity,  sulfate  ions,  and  hydrogen  ions  (pH). 

Soft  water  flora  are  limited  due  to  poor  nu- 
trients and  are  characterized  by  underwater  leaves 


Marshes 


207 


Table  1.  Some  examples  of  lake  waters'  influence  on  aquatic  plants  of  Minnesota  (selected  from  Moyle  1945). 


Soft  water:     alkalinity       <  40  ppm 
sulfates          <    5  ppm 
pH                 <    7.5 

►           reflect  poor  nutrients 

Lobelia  dortmanna                             Scirpus  subterminalis 
Sparganium  sp.                                    Eriocaulon  septanquatare 
Lsoetes  braunii                                     Sagittaria  latifolia 

Hard  water:     alkalinity     <  1 50  ppm 
sulfates        <    40  ppm 
pH                     8.0-8.8 

►            richer,  attractive  to 
waterfowl 

Chara  sp.                                               Ceratophylum  demersum 
Phragmites  communis                        Lemna  minor 
Potamogen  pectinatus                        Myrsophyllum  exalbescens 
Scirpus  validus                                    Lemna  trisulca 
Potamogen  natans                              Nuphar  variegatum 
Scirpus  acutus                                      Spirodela  polyrhiza 
Elodea  canadensis                               Numphaea  tuberosa 
Typha  latifolia                                     Utricularis  vulgaris 

Alkali  water:     sulfates      <  50  ppm 
pH                    8.4-9.0 

►           reduced  species,  but 
dense  growth 

Ruppia  sp.                                             Juncus  torreyi 
Typha  augustifolia                              Scirpus  paludosus 
Najas  marina                                       Potamogeton  pectinatus 

and  rosettes.  Floating  plants  are  rare.  Hard  water 
floras  are  richer,  more  abundant,  and  are  attractive 
to  waterfowl.  Alkali  floras  have  reduced  species  but 
often  have  dense  growth  of  a  few  well-adapted 
species. 

Stewart  and  Kantrud  (  1972)  found  that  specific 
conductivity  measured  dissolved  solids  simply  and 
reliably  to  indicate  alkalinity  or  salinity  of  prairie 
marshes.  Means  ranged  from  fresh  (about  300  mi- 
cromhos  per  cm  at  25°  C)  to  brackish  (6,300)  to  sa- 
line (37,500)  in  their  North  Dakota  ponds.  Plant  spe- 
cies composition  in  relation  to  conductivity  readings 
are  reproduced  in  modified  form  in  Figures  2 
through  5  for  wet  meadow,  shallow  marsh,  and  deep 
marsh  communities.  Despite  broad  ranges  of  toler- 
ance, common  emergent  hydrophytes  tended  to  be 
associated  with  the  following  mean  dissolved  solids 
measured  by  Stewart  and  Kantrud  (1972):  Slender 
bulrush  {Scirpus  heterochaetus)  350  mh/cm,  cattail 
1000  mh/cm,  and  alkali  bulrush  (Scirpus paludosus) 
3,500-32,000  mh/cm  (Figure  6). 

Salinity  is  clearly  one  of  the  more  important 
chemical  factors  of  coastal  and  certain  western 
marshes.  For  example,  the  effects  of  salinity  in  estab- 
lishment and  growth  of  widgeongrass  (Ruppia  mar- 


tina)  are  well  known  (Bourn  1935;  Joanen  and  Glas- 
gow 1966).  Inland  saline  waters  support  many  salt- 
tolerant  species  (halophytes),  and  their  salt  concen- 
trations strongly  influence  plant  species  composition 
and  growth  as  shown  by  Rawson  and  Moore  (1944). 
Experimental  studies  of  vegetation  in  the  Great  Salt 
Lake  basin  and  efforts  to  freshen  such  areas  to  in- 
duce more  suitable  plants  for  waterfowl  have  pro- 
vided excellent  insights  into  plant  species  and  water 
quality  influences  in  these  extreme  conditions  (Nel- 
son 1954;  Christiansen  1970).  Additionally,  observa- 
tions of  marshes  and  marsh  hydrophytes  associated 
with  warm,  salt  springs  in  Utah  provide  excellent 
data  on  species  composition  and  the  role  of  water  ta- 
bles in  salt  deposition  so  important  to  wetland  plants 
(Bolen  1964). 


Key  to  Figures  2  thru  5: 


Width  of  bars  indicates  relative  abundance. 

Thick  bars  =  frequently  common  or  abundant. 

Thinner  bars  =  common,  occasionally  abundant 
to  occasionally  fairly  common. 


208 


Marshes 


PRIMARY  SPECIES: 

False-aster 

Boltonia  latisquama 

Fowl  bluegrass 

Poa  palustris 
Silver  sedge 

Carex  praegracilis 

Lowland  white  aster 

Aster  simplex 

Northern  reedgrass 

Calamagrostis  inexpansa 

Prairie  cordgrass 

Spartina  pectinata 

Baltic  rush 

Juncus  balticus 
Foxtail  barley 

Hordeum  jubatum 
Inland  saltgrass 

Distichlis  stricta 

1 

t 
1 

Fresh 

Slightly 
brackish 

Moderately 
brackish 

Brackish 

Subsaline 

1 

1 

1 

, 

1- 

1 

- 

1 

i 

1 

1 

Figure  2.  Characteristic  plant  species  of  wet  meadows  arranged  from  fresh  to  salt  water  tolerance. 


PRIMARY  SPECIES: 

Fresh 

Slightly 
brackish 

Moderately 
brackish 

Brackish 

Subsaline 

Tall  mannagrass 

Glyceria  grandis 

Giant  burreed 

Sparganium  eurycarpum 

American  waterplantain 

Alisma  triviale 

Marsh  knotweed 

Polygonum  coccineum 

American  sloughgrass 

Beckmannia  syzigachne 

Awned  sedge 

Carex  atherodes 

Whitetop  rivergrass 

Scolochloa  festucacea 

1 

Common  spikerush 

Eleocharis  palustris 

1 



Narrowleaf  waterplantain 

Alisma  gramineum 

1 

American  bullrush 

i                                  i 

Scirpus  americanus 

i 1 

Nuttall  alkaligrass 

_i                                  I 

Puccinellia  nuttalliana 

1      ■ 

1 

Samphire 

Salicornia  rubra 

1 

Figure  3.  Characteristic  plant  species  of  shallow  marsh  emergent  vegetation. 


Marshes 


209 


PRIMARY  SPECIES: 
Slender  bulrush 

Scirpus  heterochaetus 

River  bulrush 

Scirpus  fluviatilis 

Hybrid  cattail 

Typha  latifolia  x  angustifolia 

Tule  bulrush 

Scirpus  acutus 

Alkali  bulrush 

Scirpus  paludosus 

SECONDARY  SPECIES: 
Softstem  bulrush 

Scirpus  validus 
Common  cattail 

Typha  latifolia 
Common  reed 

Phragmites  communis 

Narrowleaf  cattail 

Typha  angustifolia 

Fresh 

Slightly 
brackish 

Moderately 

brackish 

Brackish 

Subsaline 

, 

1 

Figure  4.  Characteristic  plant  species  of  deep-marsh  emergent  vegetation. 


210 


Marshes 


PRIMARY  SPECIES: 
Aquatic  liverwort 

Riccia  fluitans 
Grassleaf  pondweed 

Potamogeton  gramineus 

Common  waterstarwort 

Potamogeton  gramineus 
Aquatic  liverwort 

Ricciecarpus  natans 
Common  marestail 

Hippuris  vulgaris 
Baby  pondweed 

Potamogeton  pusillus 
Star  duckweed 

Lemna  trisulca 
Common  bladderwort 

Utricularia  vulgaris 
Richardson  pondweed 

Potamogeton  richardsonii 
Common  hornwort 

Ceratophyllum  demersum 

Spiked  watermilfoil 

Myriophyllum  exalbescens 
Aquatic  moss 

Drepanocladus  sp. 
White  buttercup 

Ranunculus  trichophyllus 
Common  duckweed 

Lemna  minor 
Western  widgeonweed 

Ruppia  occidentalis 
Horned  poolmat 

Zannichellia  palustris 
Muskgrass 

Charu  sp. 
Fennelleaf  pondweed 

Potamogeton  pectinatus 
Common  widgeonweed 

Ruppia  maritima 


Fresh 

Slightly 
brackish 

Moderately 
brackish 

Brackish 

Subsaline 

Saline 

1                              1 

1 

3 

| 1 

1 1 

1                                                       1                           I 
i 1 

1                            1                                                       1 

i 

1                                                       1                            1 

1                                                                                                               i 

Figure  5.  Characteristic  plant  species  of  submerged  and  floating  vegetation. 


Marshes 


211 


PLANT  COMMUNITY 
(in  order  of  increasing 
specific  conductance) 

Scirpus  heterochaetus 


Scirpus  fluviatilis 


Typha  sp. 


Typha  sp. 

—  Scirpus  acutus 

Scirpus  acutus 


Scirpus  acutus 

—  Scirpus  paludosus 

Scirpus  paludosus 

—  Scirpus  acatus 

Scirpus  paludosus 


SPECIFIC  CONDUCTANCE  IN  MICROMHOS  PER  CENTIMETER  AT  25  DEGREES  CENTIGRADE 
100  316  1000  3160  10,000  31,600  100,000 


1 

EXPLANATION 


Geometric  mean 


Range 
±  Standard  Deviation 

Standard  Error  of  the  mean 


2.00         225  2.50         2.75  3.00  3.25  3.50         3  75         4.00  4.25         4.50  4.75         5.00 

SPECIFIC  CONDUCTANCE  (LOG  ,0  SCALE) 


Figure  6.  Relationship  of  dominant  deep-marsh  vegetation  to  specific  conductance. 


Many  water  characteristics  are  influenced  by 
processes  within  a  basin.  These  include  the  years  of 
alternate  flooding  and  drying  which  may  concentrate 
deposits,  the  rate  of  water  flow  through  the  system 
that  carries  away  dissolved  materials  (including  nu- 
trients), and  the  processes  of  storage  and  decomposi- 
tion of  dissolved  chemicals.  Marshes  also  vary  sea- 
sonally in  concentrations  of  dissolved  solids. 

Community  Structure 

Life  Form  and  Zonation.  Wildlife  of  marshes  re- 
spond visually  to  plant  structure  and  by  trial  and  er- 
ror to  the  associated  cover  and  food.  Diversity  of 
structure  is  produced  by  various  species  of  plants 
adapted  to  marshes  and  also  by  zonation  resulting 
from  basin  configuration. 

Plants  of  marshes  have  Life  forms,  growth,  and 
reproductive  strategies  that  result  from  adaptation  to 
water  depth  and  hydroperiod.  Their  mechanisms  of 
germination,  growth,  and  spread  also  influence  how 
they  are  used  as  food  and  cover  by  wildlife.  The 
dominant  life  form,  rooted  emergents,  represents  the 
easiest  adaptation  of  hydrophytes  to  the  semiaquatic 
environment.  Seeds  typically  germinate  on  mud  flats 
or  in  very  shallow  water,  deriving  their  nutrients 


( 1 )  emergents 


212 


Marshes 


from  soil  or  organic  substrate.  Emergents  arc  di- 
verse and  complex.  The  continuum  can  be  subdi- 
vided several  ways.  Nonpersistent,  soft  and  often 
short  grasses,  sedges,  and  forbs  decompose  usually  in 
less  than  one  growth  season.  Larger,  more  robust, 
and  persistent  emergents  like  cattail,  some  bulrushes, 
and  reeds  may  stand  for  several  years.  This  category 
normally  does  not  include  woody  plants  like  willow 
(Salix  sp. )  which  are  classified  as  shrubs  or  trees. 

Still  more  truly  aquatic  plants  are  the  submer- 
gents,  most  of  which  are  rooted.  They  have  much 
dissected  or  linear  leaves  efficient  at  tapping  reduced 
sunlight  at  considerable  depths.  Such  plants,  like  the 
pondweeds  (Potamogen  sp. ),  may  have  flowers  and 
seedheads  that  reach  the  water's  surface.  Others  like 
stonewort  {Cham  sp. )  are  bottom  plants  with  totally 
submerged  fruiting  bodies. 

Plants  adapted  to  deeper  water  may  have  float- 
ing leaves  rising  from  long  stems  resting  on  the 
water's  surface  ("pad"  plants).  They  reproduce  by 
means  of  large  seeds  or  well-established  tubers  (ex- 
ample: water  lily ).  Such  floating-leaf  plants  can  tol- 
erate considerable  water  fluctuation  in  spite  of  their 
general  preference  for  deeper  water. 


The  most  aquatic  plants  are  floating  plants  that 
may  have  roots  dangling  in  the  water  for  nutrients, 
such  as  duckweed  (Lemna  sp. )  or  the  introduced 
water-hyacinth  (Eichornia  crassipes).  Another  spe- 
cies of  Lemna,  star  duckweed  (L.  trisnlca),  drifts  in 
the  water  column  like  the  many  species  of  micro- 
scopic plankton. 

Marshes  are  characterized  by  concentric  zones 
of  vegetation,  often  intergrading  or  changing  in  a 
continuum.  Sometimes  they  change  abruptly,  which 
adds  diversity  of  structure  horizontally  as  well  as 
plant  species  richness.  Thus,  a  deep-marsh  which  re- 
sults from  semipermanent  water  has  a  shallow-marsh 
zone  around  the  perimeter.  It  may  grade  into  a  wet- 
meadow  or  damp  situation  where  water  occurs  in 
spring  but  dries  out  in  early  summer  (Figure  7).  The 
shallow-marsh  or  wet-meadow  plants  are  those  one 
finds  in  smaller  basins  which  hold  shallow  water 
only  temporarily  in  spring,  but  these  lack  the  vegeta- 
tive diversity  in  structure  and  layering  found  in  a 
deep  marsh  with  gradual  shorelines  (Figure  7). 
Nonetheless,  some  concentric  zonation  is  expected 
in  any  marsh,  providing  a  horizontal  change  in  struc- 
ture which  also  induces  vertical  or  layering  diversity. 
Few  marsh  basins  have  perfect  symmetry,  and  other 
diversity  of  pattern  is  created  by  flow-through  chan- 


Life  forms  of  marshes:  (1)  emergents,  (2)  submergents,  (3)  floating-leaf  plants,  (•*)  floating  plants. 


f?a&*    £>tv4t*'f(d- 


(2)  submergents 


(3)  floating-leaf  plants 


(4)  floating  plants 


Marshes 


213 


A.  Three  examples  of  small  marshes  with  water  regimes  varying  from  ephemeral  to  shallow  to  deep, 
resulting  in:  wet-meadow,  shallow-marsh,  and  deep-marsh  vegetation. 


,~^^Alk^ljlW),   ,\lUr\)W~   ••-*-« 


&-.-'   ?1 


B.  Typical  deep  marsh  displaying  zonation  with  deep-marsh  vegetation  in  center,  shallow-marsh 
peripherally,  and  wet-meadow  at  the  edge. 


C.  Open-water  marsh  resulting  from  loss  of  vegetation  (due  to  continued  high  water  or  muskrats). 


■U  i    III. I  .1    yij 


D.  Germination  phase  at  low  water  with  a  mix  of  annuals  and  perennials,  persistent,  and 
nonpersistent  plants. 


•j**#m**ulM«^r.\^J)\j(,M 


ikuJk 


ii     i  i/i 


^wmmmmmmwJl^d 


Figure  7.  Common  vegetational  patterns  of  marshes. 


214  Marshes 


Concentric  zones  of  marsh  vegetation. 


nels,  islands  of  soil,  clumps  of  trees,  or  other  unique 
vegetation  established  at  a  lower  water  level. 

Patterns  of  Vegetation  and  Water.  The  above  com- 
ments might  suggest  that  the  more  abundant  and  di- 
verse the  vegetation,  the  better;  but  this  is  not  al- 
ways true  for  breeding  birds.  Many  bird  species  that 
are  attracted  to  vegetation  also  need  open  water  for 
landing  and  swimming.  Openings  in  emergents  cre- 
ate edge,  providing  access  to  vegetation.  Openings 
also  allow  sunlight  penetration  essential  to  submer- 
gents  and  some  animals.  Thus,  interspersion  of  cover 
and  water  is  best.  Various  species  respond  to  differ- 
ent cover-water  ratios.  Weller  and  Fredrickson 
(1974)  showed  an  increase  in  species  richness  of 
marsh  birds  and  numbers  of  individuals  with  increas- 
ing numbers  of  openings  or  pools  in  nearly  pure  cat- 
tail up  to  the  point  where  too  little  vegetation  re- 
duced nest  sites  and  food  resources.  Various  cover- 
water  ratios  supported  the  idea  that  ducks  were 
more  attracted  to  50-50  cover- water  patterns  (Ka- 
minski  and  Prince  1981;  Murkin  et  al.  1982),  al- 
though the  reasons  are  more  complex  than  earlier 
thought. 


Marsh  Size  and  Wetland  Complexes 

Wetland  size  is  an  important  influence  of  the 
presence  or  absence  of  wildlife  species  and  richness 
only  because  larger  areas  tend  to  have  more  habitat 
types  (i.e.,  diversity).  Certain  birds  (red-winged 
blackbirds  Agelaiiis  phoniceus,  blue-winged  teal,  and 
some  rails)  are  well  adapted  to  small  units  or  are 
ubiquitous,  and  use  these  as  well  as  perimeters  of 
large  marshes.  Other  species  (ruddy  ducks,  Osyura 
jamaicensis,  yellow-headed  blackbirds,  X.  xanthoce- 
phalus)  occur  mainly  in  larger,  deeper  areas  with 
persistent  emergents  and  submergents.  Thus,  clusters 
of  diverse  wetlands  may  function  like  larger  and 
more  complex  wetlands  in  having  more  bird  species 
in  the  community  (Smith  1971;  Weller  1981a). 

Marsh  Succession  and  Vegetation  Dynamics 

Because  marshes  are  shallow,  dynamic,  water 
bodies,  they  undergo  changes  in  depth,  chemical 
characteristics  of  the  water,  vegetation  (species,  den- 
sity, and  coverage),  and  associated  wildlife.  Water  re- 
gimes over  long  periods  dictate  most  or  all  of  these 


Marshes 


215 


changes;  and  the  presence,  richness,  success,  and  du- 
ration of  the  plant  community  is  in  response  to  this 
regime.  Hence,  it  is  vital  to  have  records  of  water 
depth  and  extent  of  water  coverage  of  the  marsh 
bottom,  and  to  document  seasonal  or  year-to-year 
variation  in  hydroperiods  to  interpret  responses  by 
vegetation  and  wildlife.  Moreover,  the  seed  bank  of 
long-lived  seeds  is  a  product  of  this  water  history 
(Van  der  Valk  and  Davis  1978).  The  lack  of  vegeta- 
tive response  in  a  management  setting  may  mean 
that  seeds  are  lacking  (Pederson  1981 ).  Logically, 
shallow  or  temporary  marshes  have  seeds  of  pioneer- 
ing, annual,  or  short-lived  plants.  They  are  adapted  to 
quick  response  and  maturation  that  enhance  the 
seed  bank.  Ducks  and  other  seed  eaters  respond  to 
this  stage  as  well,  but  sufficient  seeds  survive  that 
the  species  is  not  eliminated  locally.  When  condi- 
tions such  as  mud  flats  are  again  available,  these 
seeds  will  germinate.  In  deep,  open  waters  domi- 
nated by  submergents,  there  may  be  fewer  seeds  of 
emergents,  although  this  situation  varies  by  site. 
More  importantly,  terrestrial  sites  flooded  artificially 
may  lack  the  seed  bank  entirely,  and  establishment 
of  vegetation  may  take  several  years. 

The  usual  pattern  in  natural  shallow  marshes  is 
that  diverse  seeds  of  annuals  and  perennials  are  pres- 
ent. With  natural  drought  at  various  times  of  the 
year,  different  seeds  respond  (Van  der  Valk  1981 ). 
Harris  and  Marshall  ( 1963)  showed  that  early  draw- 
down with  fairly  moist  soil  induces  germinations  of 
persistent,  perennial  emergents  (e.g.,  cattail).  Later 
and  drier  drawdown  conditions  may  result  in  quick- 
maturing  and  short-lived  annuals  like  smartweed. 
They  still  produce  seeds  that  survive  over  winter — 
and  often  for  many  years — until  conditions  are 
suitable  (see  Weller  1981a). 

If  water  levels  in  subsequent  years  remain  low, 
vegetation  tends  to  be  annual  plants  or  a  few  peren- 
nials characteristic  of  shallow  marshes:  sedges, 
rushes,  and  softstem  bulrush.  If  the  water  level  in- 
creases, it  is  the  perennials  that  prosper.  Cattail,  tule 
bulrush,  common  reed,  and  other  plants  tolerant  of 
deep  water  survive.  In  addition  to  the  development 
of  masses  of  seeds,  some  floating-leaf  and  submergent 
plants  reproduce  by  tubers  or  bulbs.  They  store  en- 
ergy in  one  year  that  allows  vegetative  growth  in  the 
next  season — before  the  availability  of  suitable  seed 
germination  substrate  such  as  mudflats.  These  peren- 
nials may  go  for  years  without  germination  of  seeds, 
as  an  influence  on  the  local  population  of  the  plant. 
However,  seeds  spread  by  wind  or  birds  may  influ- 
ence the  establishment  of  these  species  in  other 
areas. 

Several  studies  have  been  cited  demonstrating 
the  impacts  of  successional  changes  on  wildlife  of 
single  marshes.  Because  of  long-term  water  cycles  or 
random  changes,  wetland  habitats  in  large  regions 
are  sometimes  adversely  affected  for  wildlife.  Yeager 


and  Swope  (  1956),  Smith  (  1971 ),  and  Pospahala  et 
al.  (1974)  all  observed  major  population  responses 
to  droughts.  Several  workers  have  shown  how  water- 
fowl populations  in  wetter  or  more  stable  areas  in- 
crease when  drought  strikes  the  interior  grassland 
marshes  (Derksen  and  Eldridge  1980). 

Macroinvertebrates 

Diverse  and  often  abundant  invertebrates  pro- 
vide needed  protein  for  reproduction  and  growth  of 
many  fish,  amphibians,  birds,  and  some  mammals.  In- 
vertebrates occupy  every  habitat  in  marshes  and  ex- 
ploit every  food  resource.  They  have  evolved  life  his- 
tory strategies  attuned  to  dynamic  water  regimes, 
seasonal  temperature  cycles,  and  differing  salinities 
and  acidities.  Life  cycles  are  intimately  tied  to  water 
regimes.  Marshes  characterized  by  deeper  and  more 
constant  water  levels  will  have  populations  of  dra- 
gonflies  and  other  long-lived  insects  that  may  require 
several  years  to  mature.  Shallow  and  more  temporary 
bodies  have  annual  crops  of  invertebrates  (fairy 
shrimps,  mosquitoes,  some  midges)  that  hatch  from 
drought-tolerant  eggs  and  mature  rapidly,  to  leave 
eggs  for  another  time  (Wiggins  et  al.  1980). 

Population  fluctuations  among  invertebrates  are 
dramatic.  The  emergences  of  annual  crops  of  insects 
like  midges  and  caddisflies,  or  the  "blooms"  of  plank- 
tonic  crustaceans  such  as  Daphnia  or  Cyclops  repre- 
sent discrete  seasonal  energy  flow  patterns  as  well  as 
reproductive  cycles.  The  abundance  and  species 
richness  of  invertebrates  vary  with  marsh  type  or 
successional  stage  (Voigts  1973),  and  are  influenced 
by  nutrient  availability  and  shifts  in  food  abundance. 
Vertebrate  consumers  respond  promptly.  An  emer- 
gence of  midges  in  mid  to  late  summer  feeds  fish, 
amphibians  and  ducks  in  the  water,  ducks  and  terns 
at  the  water's  surface,  plus  gulls  and  swallows  in  the 
air  column  up  several  hundred  feet.  Plankton  blooms 
are  exploited  by  shovelers  with  a  straining  bill,  phal- 
aropes  that  pick  out  individual  targets,  and  fish  and 
amphibians  in  the  water  column.  The  predators  and 
scavengers  too  must  be  adaptive,  because  emer- 
gences or  blooms  may  last  a  few  days  to  a  few 
weeks,  and  other  foods  may  appear  that  require  dif- 
ferent feeding  methods. 


216 


Marshes 


Wetland  biologists  now  recognize  the  vital  role 
that  macroinvertebrates  play  in  dictating  the  pres- 
ence or  absence  of  wildlife  that  need  these  for  food. 
Some  managers  are  even  manipulating  water  levels 
to  induce  high  invertebrate  populations  to  attract 
waterfowl  and  shorebirds.  Most  biologists  who  have 
worked  with  invertebrates  use  them  as  indicators  of 
marsh  condition,  but  few  techniques  have  been  de- 
veloped to  monitor  populations  or  interpret  such 
data  for  rapid  assessments  in  the  field. 


INVENTORY  AND  MONITORING  SYSTEMS 
FOR  MARSH  HABITATS 
AND  THEIR  WILDLIFE 

As  discrete,  often  small  and  self-contained  eco- 
systems, marshes  usually  are  more  easily  inventoried 
than  more  extensive  terrestrial  habitats.  But  evalua- 
ting them  or  monitoring  changes  can  present  diffi- 
cult problems.  First,  biologists  are  often  not  as  famil- 
iar with  the  plants  and  invertebrates  or  their  use  by 
wildlife  as  in  terrestrial  ecosystems.  Second,  biolo- 
gists lack  technology,  manpower,  and  precise  evalua- 
tion systems.  To  examine  these  problems  and  possi- 
ble solutions,  I  will  first  define  the  systems.  I  will 
then  outline  the  ways  in  which  one  might  want  to 
use  these  systems  or  devise  procedures  to  meet  par- 
ticular goals. 

Inventory 

Inventory  is  the  determination  of  number,  size, 
and  distribution  of  marshes.  Mapping  involves  the 


Wetland  maps,  such  as  these,  are  available  from  the  U.S. 
Fish  and  Wildlife  Service. 


geographic  aspect  of  location  and  distribution  pat- 
terns. The  most  extensive  national  wetland  inventory 
ever  attempted  is  in  progress.  It  is  a  multiagency 
Federal  effort  to  record  the  number,  size,  and  distri- 
bution of  all  wetlands  of  the  contiguous  United 
States.  Much  of  the  survey  is  already  complete.  Some 
maps  have  been  published  and  are  available  from  the 
U.S.  Fish  and  Wildlife  Service.  Other  areas  are  still 
being  mapped  or  have  not  been  mapped  in  detailed 
scale.  Vegetation  zones  may  be  gross  or  lacking  on 
these  maps.  A  general  summary  of  wetland  changes 
over  time  has  resulted  from  these  inventories 
(Frayer  et  al.  1983).  They  suggest  a  loss  in  vegetated 
palustrine  wetlands  (marshes)  of  94,770  ha  (234,000 
a.)  per  year  from  the  1950s  to  the  1970s.  The  data 
on  wetland  number,  area,  and  modifications  thereof 
are  stored  in  computer  data  bases,  and  more  detailed 
analysis  can  be  expected  in  the  future. 

On  a  regional  or  local  scale,  the  inventory  pro- 
cess may  produce  greater  detail,  not  only  in  identify- 
ing smaller  areas  but  in  perceiving  changes  due  to 
dynamic  water  regimes.  Additionally,  on-the-ground 
vegetation  and  water  depth  measurements  can  be 
taken. 

On  a  very  gross  scale,  the  number  of  wetlands 
has  been  related  to  waterfowl  populations  over  vast 
areas  of  the  northern  U.S.  and  Canada.  In  addition  to 
the  standard  annual  aerial  waterfowl  censuses,  ob- 
servers have  recorded  the  number  of  prairie  wet- 
lands to  reflect  year-to-year  climatic  conditions.  Cor- 
relations of  mallard  population  size  and  distributions 
with  pond  numbers  reflect  the  impact  of  habitat 
quality  on  breeding  conditions  for  mallards  (Pospa- 
hala  et  al.  1974).  On  a  smaller  scale,  a  simple  posi- 
tive correlation  between  pond  numbers  and  canvas- 
back  pairs  was  noted  by  Sugden  ( 1978)  over  a  series 
of  years  of  observations. 

Cover  Mapping 

Cover  mapping  produces  a  detailed  analysis  of 
the  vegetation  patterns  found  within  a  marsh  or 
other  wetland.  Coupled  with  these  is  classification, 
which  describes  vegetation,  basin,  and  water  regimes 
to  place  marshes  or  other  wetlands  into  identifiable 
types  or  classes.  Vegetation  mapping  is  a  vital  first 
step  in  understanding  potential  for  wildlife  use.  It  es- 
tablishes baseline  information  necessary  for  monitor- 
ing. Mappers  gain  a  familiarity  with  the  area  that  as- 
sists in  devising  census  methods  or  observation  sites. 
On  a  national  scale,  the  National  Wetlands  Inventory 
is  using  aerial  photographs  almost  exclusively  with 
ground  truthing  in  various  wetland  types  (Montanari 
and  Townsend  1977).  Periodic  black-and-white 
photo  series  by  the  Soil  Conservation  Service  usually 
are  available  for  baseline  maps.  Current  color  photos 
have  excellent  potential  for  indicating  significant 
characteristics  in  wildlife  habitat  evaluation  (Cowar- 
din  et  al.  1981 ).  Local  overflights  can  provide  views 


Marshes 


217 


of  marshes  that  add  great  appreciation  and  under- 
standing as  well  as  an  opportunity  for  photographs. 
Out-of-the-window  obliques  are  useful  for  records  of 
vegetation  and  water  patterns,  but  are  very  mislead- 
ing for  mapping  due  to  the  dramatic  distortion  that 
even  a  small  angle  produces.  A  plane  with  a  "belly" 
port  is  ideal  for  low-level  (9 15- 1,830m;  3,000-6,000 
ft)  35mm  pictures.  Where  this  equipment  is  not 
available,  out-of-the-window  verticals  can  be  taken 
with  an  externally-mounted  camera  (Meyer  and 
Grumstrup  1978). 

A  series  of  aerial  photographs  and  the  resulting 
maps  can  be  used  to  demonstrate  seasonal  or  yearly 
changes  in  vegetation  and  coverage,  cover-water  ra- 
tios, and  numbers  of  muskrat  lodges  in  a  given  wet- 
land. Vegetation  surveys  are  necessary  in  the  marsh 
to  provide  detailed  data  on  species  richness  or  diver- 
sity, stem  density,  water  depth,  submergent  plant  dis- 
tribution, etc.  Moreover,  work  in  the  marsh  is  vital 
to  understanding  wildlife-plant-water  relationships. 
There  is  no  substitute  for  a  good  pair  of  hip  boots  or 
chest  waders! 


Monitoring  Habitat  and  Wildlife 

Monitoring  implies  biological  assessment  with  a 
series  of  timed  observations  for  a  stated  purpose. 
Objectives  can  include  assessment  of  habitat  and 
wildlife  values  to  prevent  habitat  deterioration,  mea- 
surement of  fluctuation  in  populations  or  vegetation 
as  natural  entities  of  the  system,  or  establishment  of 
a  baseline  for  evaluating  the  consequences  of  human 
activities  on  wetland  quality.  All  of  these  activities 
are  essential  for  making  sound  resource  decisions. 
Ordinarily,  the  process  requires  measurement  of  hab- 
itat and  wildlife  populations.  Many  of  these  tech- 
niques are  covered  elsewhere.  Here  it  is  important 
only  to  consider  the  combination  of  individual  tech- 
niques to  form  a  "system"  that  will  satisfy  your  par- 
ticular needs  or  goals.  The  following  are  possible  fea- 
tures of  the  habitat,  and  use  by  wildlife  and  fish  that 
might  be  measured. 

Water  Depth  and  Fluctuation.  Water  level  changes 
are  important  to  understanding  vegetation  germina- 
tion, growth,  and  loss.  Water  levels  also  affect  nests 
of  birds  or  litters  of  muskrats,  potential  predation  by 
terrestrial  predators,  and  the  survival  of  duck  broods. 
Where  practical,  gauges  should  read  in  tenths  and 
hundredths  of  a  foot  in  relation  to  known  mean  sea 
level.  This  relation  is  an  engineering  standard  that 
has  practical  mathematical  and  communication  ad- 
vantages when  working  with  managed  or  artificial 
impoundments. 

Water  depth  and  extent  of  coverage  in  a  marsh 
determines  wildlife  use  and  seasonally  influences 
plant  germination,  growth,  and  seed  production. 
Mud  flats  may  be  viewed  negatively  at  times  of  nest- 
ing, but  may  be  beneficial  in  the  long  run  because 
the  resulting  germination  creates  emergent  vegeta- 
tion suitable  for  nesting  in  subsequent  years. 


Vegetation.  The  species  composition,  density, 
height,  distribution,  and  cover-water  interspersion 
patterns  by  water  depth  are  among  the  features  of 
the  plant  community  that  might  be  assessed.  How- 
ever, time  and  financial  constraints  prevent  measure- 
ment of  all  these  details  unless  the  goals  of  the  moni- 
toring demand  them.  Moreover,  there  is  significant 
vertical  layering  in  marsh  vegetation  as  a  result  of 
the  four  dominant  life  forms  of  aquatic  plants:  rooted 
emergcnts,  floating-leaf,  rooted  submergents,  and 
free-floating  (surface  or  water  column).  There  also 
are  some  epiphytic  filamentous  algaes  that  can  be 
abundant,  suitable  food  for  fish  and  ducks,  and  sub- 
strates for  macroinvertebrates.  Sampling  systems 
must  be  suitable  for  this  variety. 

Point-counts  are  especially  useful  because  they 
are  rapid  and  can  be  used  to  assess  all  life  forms  of 
plants  above,  at,  and  below  water  level  as  well  as 
water  depth  (Weller  and  Voigts  1983).  However, 


218 


Marshes 


. 


they  provide  only  a  erude  index  to  density  of  less 
common  plants.  Additional  sampling  systems  such  as 
quadrats  or  releves  (Mueller-Dumbois  and  Ellenberg 
1978)  may  be  necessary.  Several  studies  have  dem- 
onstrated dramatic  year-to-year  changes  in  vegetation 
that  strongly  influenced  wildlife  populations  (Weller 
and  Fredrickson  1974). 

Wildlife  Census.  Investigators  censusing  wildlife  in 
marshes  use  strategies  similar  to  those  for  terrestrial 
wildlife,  but  are  complicated  by  access  and  visibility. 
This  is  why  aerial  census  of  waterfowl  and  large 
wading  birds  of  open  water  developed  early,  and  is 
used  in  place  of  an  index  to  population  change 
(Henny  et  al.  1972).  Boats,  canoes,  air  boats,  and 
marsh  buggies  also  have  been  used  when  aerial  sur- 
vey is  either  financially  impossible  or  the  size  of  the 
area  is  too  small.  Air  boats  and  marsh  buggies  are  ex- 
pensive to  buy  and  maintain,  and  can  only  be  justi- 
fied in  large  marsh  units.  Boats  often  will  not  go 
through  emergents  or  submergents  and  thus  do  not 
perfect  survey  techniques.  Canoes  have  the  advan- 
tage of  being  quiet,  but  are  slow  and  sometimes  awk- 
ward, and  like  walking,  visibility  may  be  impaired. 
Hence,  biologists  often  use  subsamples,  indexes,  and 
indirect  estimates.  Certainly  no  simple  technique  can 
be  advised  for  all  areas. 

Censusing  wildlife  in  marshes  requires  prelimi- 
nary experimentation  because  each  area  differs.  Con- 
stancy in  technique  will  then  provide  an  index  even 
when  abundance  or  density  figures  cannot  be  ob- 
tained. No  one  technique  will  work  for  all  birds  or 
mammals.  Fish,  amphibians,  and  reptiles  are  still 
more  difficult  because  of  poor  visibility,  and  require 
subsampling  techniques  that  will  produce  different 
levels  of  success. 

Fish.  Seines  are  not  useful  in  marshes  because 
of  the  vegetation  and  organic  debris.  Kushlan  ( 1974) 
used  lift-traps,  drop-traps,  and  bottom  nets  to  get  es- 
timates of  species  richness  and  species  composition. 
Rotenone  and  electrofishing  also  can  be  useful,  but 
no  technique  is  equally  suitable  for  all  species  and 
all  vegetation  types.  Submergents  are  a  particular 
hindrance.  Some  minnows  and  smaller  species  in 
prairie  marshes  have  been  studied  by  trapping 
(Payer  and  Scalet  1978).  Much  needs  to  be  learned 
about  census  techniques  that  assess  communities  of 
fish  in  marshes. 

Amphibians.  I  know  of  no  accurate  method  of 
censusing  various  kinds  of  frogs  in  various  habitats. 
Bullfrogs  are  conspicuous  at  night  with  a  flashlight, 
but  leopard  frogs  and  small  peepers  are  difficult  to 
count.  Incidence  observed  per  unit  time  by  habitat 
category  is  possibly  the  best  available  index. 

Birds.  I  have  been  well  satisfied  with  counts  of 
territorial  male  songbirds  on  quadrats  from  observa- 
tion towers  or  raised  shorelines.  Red-winged  and  yel- 


low-headed blackbirds  are  especially  conspicuous. 
Even  less  visible  but  vocal  species  like  marsh  wrens 
can  be  pinpointed  by  triangulation  and  repeated  sur- 
veys. At  least  three  early  morning  surveys  at  the 
peak  of  breeding  activity  are  required  for  statistical 
reasons  as  well  as  to  compensate  for  variation  in  ac- 
tivity of  birds.  In  small  marshes  and  ponds  where  ac- 
cess is  enhanced  by  walking  on  knolls,  biologists 
were  able  to  survey  blue-winged  teal  and  other  pairs 
of  ducks  and  get  fairly  consistent  results  on  each  of 
three  annual  pair  counts  (Weller  1979c). 

Rails  represent  one  of  the  most  difficult  of  all 
birds  to  census  in  any  habitat.  Except  for  vocaliza- 
tions, and  an  occasional  nest  accidentally  discovered, 
rails  can  be  present  in  large  numbers  and  remain  un- 
detected. Some  workers  have  been  successful  in 
stimulating  calling  by  playing  back  recorded  vocali- 
zations (Zimmerman  1977).  Nest  density  of  larger 
species  has  been  recorded  as  a  product  of  very  in- 
tensive effort  (Tanner  and  Hendrickson  1956).  Least 
Bitterns  belong  in  the  same  difficult-to-census  class. 

The  above  examples  demonstrate  the  difficulty 
of  finding  one  technique  that  works  for  species  using 
several  different  strata  of  vegetation  and  several  dif- 
ferent reproductive  strategies.  The  line  transect 
method  of  sighting  and  vocalization  is  now  in  fairly 
widespread  use  and  holds  promise  for  providing  in- 
dexes to  various  species  (Mikol  1980).  Whether  the 
method  is  equally  functional  in  a  marsh  in  early-sea- 
son low  vegetation  and  after  a  full  season's  growth  is 
still  uncertain. 

Nest  counts  are  favored  for  conspicuous  colony 
birds  such  as  ibises,  egrets,  and  herons.  They  have 
also  been  used  for  ducks  where  chain  drags  and 
other  devices  are  used  to  arouse  hens  from  their  in- 
cubation (Higgins  et  al.  1969).  In  some  instances, 
post-breeding  season  surveys  have  been  used  to  cor- 
rect for  partial  counts  or  counts  in  dense  vegetation. 

Mammals.  Censuses  in  marshes  have  focused 
around  sign  or  dens  as  opposed  to  visual  count, 
since  many  species  are  nocturnal  or  hidden  in  dense 
vegetation.  Muskrats  are  censused  by  numbers  of 
"lodges"  with  a  fall  per-house  number  of  five  com- 
monly used  to  estimate  totals  (Dozier  1948).  Num- 
bers and  home  range  size  of  swamp  rabbits  in  south- 
east Missouri  were  derived  by  driving  with  men  and 
dogs  and  by  grid  trapping  (Toll  et  al.  I960). 

Evaluation 

Goals  of  Evaluation.  Evaluation  of  marshes  is  com- 
plex. It  involves  some  method  of  correlating  charac- 
teristics of  the  marsh  with  a  value  in  which  there  is 
special  interest,  such  as  wildlife  populations  or  vari- 
ety. Qualities  that  will  attract  an  endangered  species 
are  also  evaluated.  There  are  no  well-defined  proce- 
dures. Biologists  are  usually  dependent  on  simple 


Marshes 


219 


correlations  or  associations  rather  than  cause-and-ef- 
fect  relationships  described  in  quantitative  terms. 
Dependent  upon  objectives,  such  evaluations  may  be 
simplified  by  comparing  two  wetlands  on  qualitative 
rather  than  quantitative  terms.  Thus,  assessing  the 
quality  of  a  marsh  for  wildlife  is  a  skill  that  is  still 
evolving.  Unfortunately,  we  have  a  small  data  base, 
and  our  parameters  usually  are  qualitative  rather 
than  quantitative.  Hence,  it  is  imperative  that  we 
clarify  the  objectives  and  the  parameters  of  an 
evaluation. 

Often  we  limit  our  interest  and  evaluation  to  a 
single  species  (e.g.,  whooping  cranes,  as  an  endan- 
gered species)  or  a  group  of  recreational  importance 
(waterfowl)  or  a  nuisance  group  (blackbirds).  Using 
individuals  per  hectare  (or  square  mile  on  vast 
areas),  or  duck-use  days,  we  identify  marshes  that 
are  highly  productive  or  much  used  versus  those 
with  low  numbers.  If  we  use  a  wetland  classification 
system,  we  associate  number  of  species  and  individu- 
als with  types  of  wetlands  as  did  Stewart  and  Kan- 
trud  ( 1973)  in  North  Dakota.  Or  we  may  map  the 
cover  of  zones  or  areas  of  a  single  marsh  to  denote 
where  nesting  occurs,  where  feeding  occurs,  etc.  We 
can  then  determine  how  much  of  any  one  habitat  is 
needed  to  attract  and  maintain  a  single  species. 

There  are  dangers  in  species  or  group-oriented 
assessments  in  that  we  ignore  the  community  of  ver- 
tebrates as  a  whole  and  bias  our  effort  toward  spe- 
cies of  our  choosing.  Other  species  may  be  of  impor- 
tance to  other  persons  for  other  reasons.  Another 
danger  rests  with  a  comparative  approach  of  choos- 
ing between  two  marshes  because  two  wetlands  in 
the  same  area  may  have  different  water  regimes  and 
be  in  different  phases  of  the  drawdown-reflooding 
cycle.  Their  potential  to  produce  may  be  just  as 
great  but  each  is  on  a  different  time  schedule.  With- 
out long-term  observations,  or  experience  in  a  cer- 
tain climatic-geomorphic  area,  a  highly  erroneous 
conclusion  might  be  reached.  Generally,  such  rela- 
tions between  maximum  wildlife  populations  and 
some  habitat  feature  are  based  on  one  or  few  stud- 
ies, so  confidence  levels  may  be  low.  Clearly,  addi- 
tional work  is  needed. 

Tools  for  Evaluation.  For  marsh  wildlife,  we  look 
positively  on  certain  features  as  being  important  en- 
hancements in  marshes  and  use  them  as  parameters 
of  a  quality  habitat. 

Complexity  or  diversity  of  vegetative  structures 
is  measured  by  cover-water  ratio,  vertical  layering, 
and  possibly  plant  species  richness.  Cover-cover 
edges  or  cover-water  edges  were  identified  as  attrac- 
tive to  nesting  birds  by  Beecher  ( 1942),  and  maxi- 
mum species  richness  of  birds  was  noted  with  a  ratio 
of  30%  cover  to  70%  open  water  ( Weller  and  Fred- 
rickson  1974).  Using  experimental  approaches, 


workers  found  the  50-50  cover-water  ratio  most  at- 
tractive to  feeding  dabbling  ducks  as  well  (Kaminski 
and  Prince  1981;  Murkin  et  al.  1982).  Vertical  layer- 
ing (strata)  has  been  suggested  as  important  for  nest- 
ing marsh  birds  by  Weller  and  Spatcher  ( 1965)  and 
Short  (1982).  These  measures  of  structural  diversity 
suggest  that  increased  variety  and  numbers  of  birds 
would  be  expected  with  increased  water,  cover  ra- 
tio, or  number  of  strata,  but  that  low  numbers  and 
different  species  may  be  expected  in  the  extremes  of 
either  a  dense  monoculture  or  sparse  vegetation 
(open  marsh). 

A  wetland  has  characteristic  water  regimes  due 
to  local  climate,  surface  drainage  patterns,  and  sub- 
surface hydrology.  Determining  what  is  a  favorable 
regime  is  a  moot  issue,  and  dependent  on  the  values 
and  functions  attributed  to  the  wetland.  For  wildlife, 
water  level  stability  during  breeding  seasons  is  ideal 
to  prevent  flooding  of  duck  nests  (Low  1945)  or 
drowning  of  young  muskrats  (Errington  1937).  Peri- 
odic drawdowns  are  essential  to  establishment  of  di- 
verse wetland  plants  in  reestablishment  of  vegetation 
after  a  flood-out  or  an  eat-out  by  herbivores.  Musk- 
rats  or  nutria  are  more  vulnerable  to  predators  in 
dry  or  shallow  marshes  (Errington  1961 ).  Permanent 
changes  producing  higher  water  level  may  be  nearly 
as  detrimental  to  wildlife  as  is  drainage  of  a  wetland. 
Such  "flooding"  reduces  submergent  growth  (Robel 
1962),  availability  of  benthic  organisms  to  some  spe- 
cies (Boyer  and  Psejek  1977),  and  covers  established 
beds  of  emergents  (Weller  1981b). 

Shoreline  gradient  and  vegetative  cover  can  be 
very  important  to  suitability  of  an  area  for  wildlife. 
Steep  shores  reduce  emergent  growth  and  induce 
more  terrestrial  plants.  Such  shorelines  may  be  ideal 
loafing  sites  for  ducks  or  muskrats,  but  are  less  suita- 
ble for  invertebrates  that  favor  shallow  water  and 
aquatic  plants. 

Turbidity  reduces  growth  of  important  submer- 
gents  like  sago  pondweed,  P.  pectinatus  (Robel 
1961 ).  Current  or  wave  action  tends  to  reduce  perim- 
eter emergents  and  takes  out  nutrients  from  the  sys- 
tem. Nutrient  levels  in  water  and  soil  can  be  mea- 
sured by  various  chemical  means.  Usually,  some  sim- 
ple measurements  of  alkalinity  are  used  to  reflect 
nutrient-rich  waters  (eutrophic  conditions)  as  op- 
posed to  less  enriched  (oligotrophic)  water.  Conduc- 
tivity, an  electrical  measure  of  dissolved  solids,  is  a 
common,  quick  technique  to  assess  whether  water  is 
extremely  pure,  highly  alkaline,  or  saline  due  to  con- 
centrations of  minerals.  Extreme  levels  of  salt  con- 
tent may  inhibit  plant  growth. 

Nutrient  cycling  also  is  involved  as  nutrients 
tend  to  be  tied  up  quickly  (seasonally)  by  plants, 
and  returned  to  water  and  substrate  reservoirs  by 
decomposition  of  plants.  Where  decomposition  is  ex- 
tensive in  late  stages  of  succession  or  following  eat- 


220 


Marshes 


outs,  extensive  surface  floating  plants  or  algal  blooms 
may  be  common.  Again,  this  is  not  easily  measured 
but  one  must  be  conscious  that  fluctuation  in  float- 
ing duckweed  or  other  seasonal  events  is  not  neces- 
sarily negative.  Floating  duckweed  can  be  valuable  as 
food  itself,  and  for  the  snails  and  amphipods  it  har- 
bors. But  continuous  spring  to  fall  crops  suggest  ex- 
cess nutrients  and  duckweed  will  screen  out  growth 
of  submergents  that  produce  seeds  and  tubers  valua- 
ble to  waterfowl  later  in  the  year.  Filamentous  algae 
can  have  the  same  effects  but  is  less  valuable  as  a 
food. 

Food  resources  are  a  product  of  water  regimes 
plus  plant  and  invertebrate  communities.  Generally, 
there  are  positive  correlations  between  vegetative 
complexity  and  bird  populations  (Weller  and  Fred- 
rickson  1974)  and  between  invertebrate  abundance 
and  duck  populations  (Whitman  1976;  Eriksson 
1978;Joyner  1980).  The  presence  of  seed-producing 
plants  like  smartweed  and  sago,  or  of  foliage-produc- 
ing plants  like  sago  and  widgeongrass,  is  of  high 
value  for  wildlife  (Fredrickson  and  Taylor  1982).  But 
data  are  not  available  for  all  foods  or  all  wildlife  spe- 
cies. What  causes  the  increase  in  populations  of  am- 
phibians like  frogs  and  salamanders  is  complicated 
by  complex  food  chains  of  algae,  phytoplankton,  zoo- 
plankton,  and  small  invertebrates,  and  cannot  be 
quickly  or  easily  assessed.  The  presence  of  fish  is  a 
strong  influence  on  quality  of  habitat  for  mergansers 
and  grebes,  but  large  predatory  fish  may  prey  heavily 
on  ducklings  (Lagler  1956). 

The  presence  of  macroinvertebrates  is  positive. 
Cover-water  ratios  do  not  always  enhance  inverte- 
brate numbers  (Murkin  et  al.  1982),  but  variation  in 
type  and  numbers  of  invertebrates  varies  with 
successional  stage  of  the  marsh  (Voigts  1973).  Be- 
cause invertebrates  are  not  easily  assessed  by  any 
simple  and  easily  obtained  measurement,  taxon  rich- 
ness is  being  explored  as  the  most  simple  way  to  as- 
sess relative  importance.  However,  one  must  be  cau- 
tious when  associating  abundance  with  availability  as 
deep  water  can  prevent  some  ducks  from  obtaining 
food  (Boyer  and  Psujek  1977). 

Nest  sites  and  materials  are  a  product  of  vegeta- 
tion types  and  patterns.  Plant  density  could  influence 
suitability  of  cover  for  nests — although  data  are  not 
available  to  clearly  support  this  assumption.  Sparse 
vegetation  may  serve  coots  and  grebes  that  carry 
dead  vegetation  for  nests  and  need  only  the  sparse 
material  to  hold  the  nest  in  a  wind.  But  ducks  do  not 
carry  vegetation  to  build  nests.  Dead  and  live  mate- 
rials on  the  site  must  be  sufficient  to  create  a  nest 
bowl  and  sometimes  a  canopy.  Songbirds  seem  to 
need  vertical  stalks  of  vegetation  that  are  sufficiently 
robust  that  the  weight  of  the  nest,  eggs,  female,  and 
young  do  not  cause  collapse.  The  stalks  must  be  suf- 
ficiently close  that  weaving  of  nest  materials  be- 


tween the  uprights  provides  support  and  cover 
(Weller  and  Spatcher  1965). 

Muskrats  carry  lodge  material,  but  build  large 
lodges  where  material  is  abundant,  and  small  plat- 
forms in  suboptimal  habitats.  Breeding  sites  for  fish, 
frogs,  and  most  reptiles  really  have  not  been  evalu- 
ated in  marshes;  it  is  believed  that  they  too  are  at 
peak  levels  when  the  entire  ecosystem  is  responding 
to  nutrient  availability,  successional  stages,  inverte- 
brate variety,  and  water  depths  that  allow  comple- 
tion of  life  cycles. 

Evaluation  Systems.  Evaluation  of  wetlands  for 
wildlife  is  not  new,  and  several  systems  are  currently 
in  use.  The  Habitat  Evaluation  Procedures  (HEP) 
developed  by  the  U.S.  Fish  and  Wildlife  Service 
(USFWS  1980)  utilizes  individual  species  models 
called  Habitat  Suitability  Index  ( HSI )  models. 
These  models  are  used  to  build  a  composite  of  key 
species  within  the  habitat  ( Edwards  and  Twomey 
1982;  Schroeder  1982).  HSI  models  are  not  available 
for  all  species  but  are  being  prepared  by  USFWS. 
Such  a  system  is  used  in  mitigation  but  also  has  wide 
applicability  to  biological  survey  work  in  marshes. 
The  system  has  been  criticized  for  being  species 
rather  than  community  oriented.  This  and  other  sys- 
tems are  described  in  the  chapter  on  habitat  evalua- 
tion systems. 

The  U.S.  Army  Corps  of  Engineers  has  a  more 
ecosystem-oriented  method  ( Habitat  Evaluation 
System),  but  it  has  not  reached  the  same  level  of 
development  and  now  deals  mainly  with  terrestrial 
systems  ( U.S.  Army  Corps  of  Engineers  1 980 ). 

The  need  for  evaluation  is  so  great  for  so  many 
diverse  fields  that  less  rigorous  approaches  are  being 
developed.  A  system  in  progress  by  the  Federal  High- 
way Administration  evaluates  wetlands  in  their  en- 
tirety for  all  values  and  functions,  with  qualitative 
ratings  of  high,  moderate,  and  low  for  some  75  pre- 
dictors (Adamus  1983;  Adamus  and  Stockwell  1983). 
Waterfowl  and  other  water  birds  are  used  as  the  ma- 
jor indicator  groups  so  that  much  of  the  wildlife 
community  is  not  represented,  but  there  are  several 
important  predictors  related  to  values  for  wildlife. 
These  include  contiguity  between  wetlands,  size, 
vegetation  form  and  density,  salinity-conductivity, 
pH,  flooding,  water  depths,  edge,  waterfowl  food  val- 
ues of  plants,  invertebrate  density,  alkalinity,  and  eu- 
trophication.  Collectively,  these  include  many  of  the 
important  factors  that  have  been  identified  as  contrib- 
uting to  quality  habitat  for  marsh  wildlife.  But  no 
one  has  derived  a  mechanism  for  rating  these  and 
developing  a  cumulative  rating  without  going  to  the 
detail  used  in  the  HEP. 

Several  states  (Michigan,  New  York,  and  Wiscon- 
sin) also  have  wetland  evaluation  or  rating  systems 
that  emphasize  wildlife  and  fish  values,  but  thev  tend 


Marshes 


221 


to  have  only  regional  applicability.  Numerous  indi- 
viduals and  groups  are  experimenting  with  group- 
specific  evaluation  systems  incorporating  current 
ecological  community  factors  such  as  species  rich- 
ness and  species  diversity  in  addition  to  population 
data  (Williams  1980).  Thus,  at  the  present  stage  of 
development,  the  options  are  -- 

•  make  qualitative  judgments  based  on  items  of 
special  interest  to  the  evaluator  (done  consis- 
tently and  comparatively), 

•  use  HEP  if  the  HSI  models  are  available  and 
seem  to  fit  the  needs,  or 

•  make  a  series  of  gross  to  detailed  surveys  and 
measurements  that  provide  indices  to  those 
items  of  special  interest  or  importance  at  the 
site  or  sites  in  question. 

DISCUSSION 

Compared  with  other  habitats,  marshes  are  so 
dynamic  they  create  difficult  situations  for  assess- 
ment as  wildlife  habitats.  Marshes  require  -  - 

•  long-term  observations  or  use  of  past  aerial  pho- 
tos to  aid  in  interpretation, 

•  careful  selection  of  parameters  that  fit  the  local 
situation, 

•  testing  of  measurements  for  reliability  as  indica- 
tors of  wildlife  habitats, 

•  standardization  of  times  and  techniques  when 
reassessing  habitats, 

•  experimental  efforts  to  improve  technologies, 
and 

•  data  gathering  on  localized  areas  to  reflect  local 
water  regimes  and  wildlife  populations. 


Marshes  are  exciting  because  of  their  diversity 
and  dynamics,  but  they  can  be  disappointing  and  de- 
ceiving as  well.  A  rich  habitat  once  lost  causes  out- 
cries from  an  uninformed  public  for  correction  of  a 
problem  that  may  not  exist.  Public  information  on 
the  "boom  and  bust"  dynamics  of  wetlands  is  essen- 
tial to  prevent  wasted  and  needless  action  on  areas 
best  left  alone.  Concurrently,  gradual  changes  can  be 
hidden  among  the  extremes,  and  only  long-term 
monitoring  will  indicate  trends  that  induce  concern 
and  investigation. 


The  accuracy  of  predictors,  qualitative  judg- 
ments, or  even  detailed  measurements  is  poorly 
known.  It  is  difficult  to  encourage  extensive  effort 


when  the  variability  of  the  system  may  mask  impor- 
tant events.  Yet,  there  is  no  option.  The  alternative 
of  no  evaluation,  census,  or  monitoring  is  too  likely 
to  result  in  irreparable  damage  to  the  system.  The 
level  of  accuracy  can  only  be  set  by  the  needs,  and 
even  detailed  evaluation  can  ignore  obvious  truths.  If 
goals  are  established  first,  and  the  level  of  detail  set 
by  need,  many  of  the  otherwise  difficult  decisions 
become  obvious. 

Above  all,  do  not  assume  someone  else  knows 
all  the  answers.  Instead,  learn  the  system  by  repeated 
visits  and  careful  observation,  search  for  the  obvious 
truths,  use  common  sense  in  devising  an  evaluation 
system,  then  stick  with  it  as  much  as  possible  to  pro- 
vide comparable  data  between  observations. 


LITERATURE  CITED 

ADAMUS,  PR.  1983.  A  method  for  wetland  functional  as- 
sessment, Vol.  II.  FHWA  assessment  method.  U.S.  Dep. 
of  Transportation.  Federal  Highway  Admin.  Rep. 
FHWA-IP-82-24.  138pp. 

PR.  and  L.T.  STOCKWELL.  1983-  A  method  for 

wetland  functional  assessment:  Vol.  I.  Critical  review 
and  evaluation  concepts.  U.S.  Dep.  of  Transportation. 
Federal  Highway  Admin.  Rep.  FHWA-IP-82-23.  181pp. 

BEECHER,  WJ.  1940.  Nesting  birds  and  the  vegetative 
substrate.  Chicago.  Chicago  Ornith.  Soc.  69pp. 

BERGMAN,  R.D.,  R.L.  HOWARD,  K.F.  ABRAHAM,  and  M.W. 
WELLER.  1977.  Waterbirds  and  their  wetland  re- 
sources in  relation  to  oil  development  at  Storkersen 
Point,  Alaska.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  Re- 
sour.  Publ.  129.  38pp. 

BOLEN,  E.G.  1964.  Plant  ecology  of  spring-fed  salt 

marshes  in  western  Utah.  Ecol.  Monogr.  34:143-166. 

BOURN,  W.S.  1935.  Sea-water  tolerance  of  Ruppia  mari- 
tima  Contrib.  Boyce  Thompson  Inst.  7:249-255. 

BOYER,  R.L.  and  MJ.  PSUJEK.  1977.  A  comparison  of  wet- 
land bird  aggregations  and  macrobenthos.  Trans.,  111. 
State  Acad.  Sci.  70:332-340 

CHABRECK,  R.H.  1972.  Vegetation,  water  and  soil  charac- 
teristics of  the  Louisiana  coastal  region.  Louisiana  St. 
Univ!  Bull.  66-4.  72pp. 

CHRISTIANSEN,  J.E.  1970.  Water  requirements  of  water- 
fowl marshlands  in  northern  Utah.  Utah  Div.  of  Fish 
and  Game.  Publ.  69-12.  108pp. 

COWARDIN,  L.M.  and  D.H.JOHNSON.  1973.  A  prelimi- 
nary classification  of  wetland  plant  communities  in 
north-central  Minnesota.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  Special  Sci.  Rep— Wildl.  168.  33pp. 

,  D.S.  GILMER,  and  L.M.  MECHLIN.  1981.  Character- 
istics of  central  North  Dakota  wetlands  determined 
from  sample  aerial  photographs  and  ground  study. 
Wildl.  Soc.  Bull.  9:280-288. 
-,  V.  CARTER,  F.C  GOLET,  and  E.T.  LAROE.  1979. 


Classification  of  wetlands  and  deepwater  habitats  of 
the  United  States.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv. 
FWS/OBS-79/31-  103pp. 
DERKSEN,  D.V.  and  WD.  ELDRIDGE.  1980.  Drought-dis- 
placement of  pintails  to  the  arctic  coastal  plain,  Alaska. 
J.  Wildl.  Manage.  44:224-229. 


222 


Marshes 


DOZIER,  H.L.  1948.  Estimating  muskrat  populations  by 
house  counts.  Trans.  North  Am.  Wildl.  Conf.  13:372- 
392. 

EDWARDS,  E.A.  and  K.  TWOMEY.  1982.  Habitat  suitability 
index  models:  common  carp.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  FWS/OBS-82/10.12.  28pp. 

ERIKSSON,  O.G.  1978.  Lake  selection  by  goldeneye  duck- 
lings in  relation  to  the  abundance  of  food.  Wildfowl 
29:81-85. 

.  1979.  Competition  between  freshwater  fish  and 

goldeneyes  {Bucephala  clangula )  for  prey.  Oceologia 
41:99-107. 

ERRINGTON,  P.L.  1937.  Drowning  as  a  cause  of  mortality 
in  muskrats.  J.  Wildl.  Manage.  18:497-500. 

.  1961.  Muskrats  and  marsh  management.  The  Stack- 
pole  Co.,  Harrisburg,  PA  and  The  Wildl.  Manage.  Inst., 
Washington,  DC.  183pp. 

and  T.G.  SCOTT.  1945.  Reduction  in  productivity 


of  muskrat  pelts  on  an  Iowa  marsh  through  depreda- 
tions of  red  fox.  J.  Agric.  Res.  71:137-148. 

ESHER,  R.J.,  R.L.  WOLFE,  and  J.H.  LAYNE.  1978.  Swimming 
behavior  of  rice  rats  ( Oryzomya  palustris )  and  cotton 
rats  (Sigtnodon  hispidis).  J.  Mammal.  59:551-558. 

FISLER,  G.F.  1961.  Behavior  of  salt  marsh  Microtiis  during 
winter  high  tides.  J.  Mammal.  42:37-43- 

FORNEY,  J. L.  1968.  Production  of  young  northern  pike  in 
a  regulated  marsh.  N.Y.  Fish  and  Game  J.  15:143-154. 

FRAYER,  W.E.,  DC.  BOWDEN,  FA.  GRAYGILL,  and  T.J. 
MONAHAN.  1983-  Status  and  trends  of  wetlands  and 
deepwater  habitats.  U.S.  Dep.  Inter.,  Fish  and  Wildl. 
Serv.  31pp 

FREDRICKSON,  L.H.  and  T.S.  TAYLOR.  1982.  Management 
of  seasonally  flooded  impoundments  for  wildlife.  U.S. 
Dep.  Inter.,  Fish  and  Wildl.  Serv.  Resour.  Publ.  148. 
29pp. 

GOLET,  F.C.  and  J.S.  LARSON.  1974.  Classification  of  fresh- 
water wetlands  in  the  glaciated  Northeast.  U.S.  Dep. 
Inter.,  Fish  and  Wildl.  Serv.  Resour.  Publ.  116.  56pp. 

HARRIS,  S.W.  and  W.H.  MARSHALL.  1963.  Ecology  of 
water-level  manipulations  on  a  northern  marsh.  Ecol. 
44:331-343. 

HENNY,  C.J.,  DR.  ANDERSON,  and  R.S.  POSPAHALA.  1972. 
Aerial  surveys  of  waterfowl  production  in  North 
America,  1955-71.  U.S.  Dep.  Inter.,  Fish  and  Wildl. 
Serv.  Special  Sci.  Rep.-Wildl.  160.  48pp. 

HIGGINS,  K.F.,  L.M.  KIRSCH,  and  I.J.  BALL.  1969-  A  cable- 
chain  device  for  locating  duck  nests.  J.  Wildl.  Manage. 
33:1009-1011. 

JOANEN,  J.T.  and  L.L.  GLASGOW.  1966.  Factors  influenc- 
ing the  establishment  of  widgeongrass  stands  in  Loui- 
siana. Proc.  Ann.  Conf.  SE  Assoc.  Game  and  Fish 
Comm.  19:78-92. 

JOYNER,  D.E.  1980.  Influence  of  invertebrates  on  pond  se- 
lection by  ducks  in  Ontario.  J.  Wildl.  Manage.  44:700- 
705. 

KAMINISKI,  R.M.  and  H.H.  PRINCE.  1981.  Dabbling  duck 
and  aquatic  macroinvertebrate  responses  to  manipu- 
lated wetland  habitat.  J.  Wildl.  Manage.  45:1-15. 

KRAPU,  G.  1974.  Feeding  ecology7  of  pintail  hens  during 
reproduction.  Auk  91:278-290. 

KUSHLAN,  J.A.  1974.  Quantitative  sampling  offish  popula- 
tions in  shallow,  freshwater  environments.  Trans.  Am. 
Fish.  Soc.  103:348-352. 

LAGLER,  K.F.  1956.  The  pike,  Esox  lucius  Linnaeus,  in  re- 
lation to  waterfowl  on  the  Seney  National  Wildlife 
Refuge,  Michigan.  J.  Wildl.  Manage.  20:114-124. 

LOW,  J.B.  1945.  Ecology  and  management  of  the  redhead, 
Nyroca  americana,  in  Iowa.  Ecol.  Monogr.  15:35-69. 


MARTIN,  AC,  N.  HOTCHKISS,  F.S.  UHLER,  and  W.S. 

BOURN.  1953.  Classification  of  wetlands  of  the  United 
States.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  Special 
Sci.  Rep.-Wildl.  20.  14pp. 

McILHENNY,  E.A.  1976.  The  alligators  life  history.  Soc.  for 
the  Study  of  Amphibians  and  Reptiles.  117pp. 

McKNIGHT,  D.  and  J.B.  LOW.  1969.  Factors  affecting  wa- 
terfowl production  on  a  spring-fed  salt  marsh  in  Utah. 
Trans.  North  Am.  Wildl.  Nat.  Resour.  Conf.  34:307- 
314. 

METCALF,  F.P.  1931.  Wild  duck  foods  of  North  Dakota 
lakes.  U.S.  Dep.  Agric.  Tech.  Bull.  221.  72pp. 

MEYER,  MP.  and  P.D.  GRUMSTRUP.  1978.  Remote  sensing 
applications  in  agriculture  and  forestry.  Univ.  of  Min- 
nesota, St.  Paul.  IAFHE  RSL  Res.  Rep.  78-1.  60pp. 

MIKOL,  S.A.  1980.  Field  guidelines  for  using  transects  to 
sample  nongame  bird  populations.  U.S.  Dep.  Inter., 
Fish  and  Wildl.  Serv.  FWS/OBS-80/58. 

MONTANARI,  J.H.  and  J.E.  TOWNSEND.  1977.  Status  of 
the  national  wetlands  inventory.  Trans.  North  Am. 
Wildl.  Nat.  Resour.  Conf.  42:66-72 

MOYLE,  J.B.  1945.  Some  chemical  factors  influencing  the 
distribution  of  aquatic  plants  in  Minnesota.  Am.  Midi. 
Naturalist.  34:402-420. 

MUELLER-DUMBOIS,  D  and  H.  ELLENBERG.  1974.  Aims 
and  methods  of  vegetation  ecology.  John  Wiley  & 
Sons,  Inc.  545pp. 

MURKIN,  H.R.,  R.M.  KAMINSKI,  and  R.D.  TITMAN.  1982. 
Responses  by  dabbling  ducks  and  aquatic  inverte- 
brates to  an  experimentally  manipulated  cattail  marsh. 
Can.  J.  of  Zool.  60:2324-2332. 

NELSON,  N.F.  1954.  Factors  in  the  development  and  resto- 
ration of  waterfowl  habitat  at  Ogden  Bay  Refuge,  We- 
ber County,  Utah.  Utah  Dep.  of  Fish  and  Game.  Publ. 
6.  87pp. 

NEXON,  S.W.  1982.  The  ecology  of  New  England  high  salt 
marshes:  a  community  profile.  U.S.  Dep.  Inter.,  Fish 
and  Wildl.  Serv.  FWS/OBS-81-55.  70pp. 

PAYER,  R.D.  and  C.G.  SCALET.  1978.  Population  and  pro- 
duction estimates  of  fathead  minnows  in  a  South  Da- 
kota prairie  marsh.  Prog.  Fish-Cult.  40:63-66. 

PEDERSON,  R.L  1981.  Seed  bank  characteristics  of  the 
Delta  Marsh,  Manitoba:  applications  for  wetland  man- 
agement. Pages  61-82  in  Richardson,  B.,  Selected  Pro- 
ceedings of  the  Midwest  Conference  on  Wetland  Val- 
ues and  Management.  660pp. 

PENFOUND,  W.T.  and  E.S.  HATHAWAY.  1938.  Plant  com- 
munities in  the  marshland  of  southeastern  Louisiana. 
Ecol.  Monogr.  8:1-56. 

POSPAHALA,  R.S.,  DR.  ANDERSON,  and  C.J   HENNY. 
1974.  Population  ecology  of  the  mallard:  II  Breeding 
habitat  conditions,  size  of  breeding  populations,  and 
production  indices.  U.S.  Dep.  Inter.,  Fish  and  Wildl. 
Serv.  Resour.  Publ.  1 1 5.  73pp. 

RAWSON,  D.S.  and  G.T.  MOORE.  1944.  Saline  lakes  of  Sas- 
katchewan. Can.  J.  of  Res.  22:141-201. 

ROBEL,  R.J.  1961.  The  effects  of  carp  populations  on  the 
production  of  waterfowl  food  plants  on  a  western  wa- 
terfowl marsh.  Trans.  North  Am.  Wildl.  Nat.  Resour. 
Conf.  26:147-159. 

ROBEL,  R.J.  1962.  Changes  in  submersed  vegetation  fol- 
lowing a  change  in  water  level.  J.  Wildl.  Manage. 
26:221-224. 

SCHROEDER,  R.L.  1982.  Habitat  suitability  index  models: 
yellow-headed  blackbird.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  FWS/OBS-82/ 10.26.  12pp. 

SHAW,  S.P.  and  C.G.  FREDINE.  1956.  Wetlands  of 
the  United  States:  their  extent  and  their  value  to 


Marshes 


225 


waterfowl  and  other  wildlife.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  Circular  39.  67pp. 

SHORT,  H.L.  1982.  Use  of  northern  prairie  and  wetland 
habitat  by  breeding  birds.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  27  pp. 

and  K.P.  BURNHAM.  1982.  Technique  for  structur 

ing  wildlife  guilds  to  evaluate  impacts  on  wildlife 
communities.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv. 
Spec.  Sci.  Rep.-Wildl.  244.  34pp. 

SMITH,  AG.  1971.  Ecological  factors  affecting  waterfowl 
production  in  the  Alberta  parklands.  U.S.  Dep.  Inter., 
Fish  and  Wildl.  Serv.  Resour.  Publ.  98.  49pp. 

STEWART,  R.F.  and  HA.  KANTRUD.  1971.  Classification  of 
natural  ponds  and  lakes  in  the  glaciated  prairie  region. 
U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  Resour.  Publ.  92. 
57pp. 

.  1972.  Vegetation  of  prairie  potholes,  North  Dakota, 

in  relation  to  quality  of  water  and  other  environmen- 
tal factors.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  and 
Geological  Survey.  Geological  Survey  Professional  Pa- 
per 585-D.  36pp.' 

.  1973.  Ecological  distribution  of  breeding  water- 
fowl populations  in  North  Dakota.  J.  Wildl.  Manage. 
37:39-50. 

SUGDEN,  L.G.  1978.  Canvasback  habitat  use  and  produc- 
tion in  Saskatchewan  parklands.  Can.  Wildl.  Serv.  Occ. 
Pap.  34.  32pp. 

SWANSON,  G.A.,  G.L.  KRAPU,  and  JR.  SERIE.  1979.  Foods 
of  laying  female  dabbling  ducks  on  the  breeding 
grounds.  Pages  47-57  in  Bookhout,  T.A.,  ed.  Proc. 
1977  Symp.  Waterfowl  and  Wetlands.  The  Wildlife  So- 
ciety. Madison,  WI. 

TANNER,  W.D.  and  G.O.  HENDRICKSON.  1956.  Ecology 
of  the  king  rail  in  Clay  County,  Iowa.  Iowa  Bird  Life. 
26:54-56. 

TOLL,  J.E.,  T.S.  BASKETT,  and  C.H.  CONAWAY.  I960. 
Home  range,  reproduction,  and  foods  of  the  swamp 
rabbit  in  Missouri.  Am.  Midi.  Naturalist.  63:398-412. 

U.S.  ARMY  CORPS  OF  ENGINEERS.  1980.  A  habitat  eval 
uation  system  for  water  resources  planning.  U.S.  Army- 
Corps  of  Engineers,  Lower  Mississippi  Valley  Division. 
1 58pp. 

U.S.  FISH  AND  WILDLIFE  SERVICE.  1980  Habitat  evalu 
ation  procedures  (HEP).  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  ESM  102.  27pp. 

VAN  DER  VALK,  AG.  1981.  Succession  in  wetlands:  a 
gleasonian  approach.  Ecol.  62:688-696. 

and  C.B.  DAVIS.  1978.  The  role  of  seed  banks  in 

the  vegetation  dynamics  of  prairie  glacial  marshes. 
Ecol.  59:322-335. 


VOIGTS,  D.K.  1973-  Food  niche  overlap  of  two  Iowa 
marsh  icterids.  Condor.  75:392-399. 

WELLER,  M.W.  1979a.  Wetland  habitats.  Pages  210-234  in 
Greeson,  P.E.,  J.R.  Clark  and  J.E.  Clark,  eds.  Wetland 
Functions  and  Values:  the  State  of  Our  Understanding. 
Am.  Water  Resour.  Assoc.  674pp. 

.  1979b.  Birds  of  some  Iowa  wetlands  in  relation  to 

concepts  of  faunal  preservation.  Proc.  la.  Acad.  Sci. 
86:81-88. 

.  1979c.  Density  and  habitat  relationships  of  blue- 
winged  teal  nesting  in  Northwestern  Iowa.  J.  Wildl. 
Manage.  43:367-374. 

.  1981a.  Freshwater  marshes:  ecology  and  wildlife 

management.  Univ.  Minn.  Press.  146pp. 

.  1981b.  Estimating  wildlife  and  other  wetland 

losses  due  to  drainage  and  other  perturbations.  Pages 
337-346  in  Richardson,  B.,  ed.  Selected  Proceedings  of 
the  Midwest  Conference  on  Wetland  Values  and  Man- 
agement. Minn.  Water  Planning  Board,  St.  Paul,  MN. 
660pp. 

and  L.H.  FREDRICKSON.  1974.  Avian  ecology  of  a 

managed  glacial  marsh.  Living  Bird.  12:269-291. 

and  C.E.  SPATCHER.  1965.  Role  of  habitat  in  the 


distribution  and  abundance  of  marsh  birds.  Iowa  State 
Univ.  Agric.  and  Home  Econ.  Exp.  Sta.  Spec.  Rep.  43. 
31pp 

and  D.K.  VOIGTS.  1983.  Changes  in  the  vegetation 


and  wildlife  use  of  a  small  prairie  wetland  following  a 
drought.  Proc.  Iowa  Acad.  Sci.  90:50-54. 

WHITMAN,  W.R.  1976.  Impoundments  for  waterfowl.  Can. 
Wildl.  Serv.  Occ.  Pap.  22.  22pp. 

WIGGINS,  G.B.,  R.J.  MACKAY,  and  I.M.  SMITH.  1980.  Evo- 
lutionary and  ecological  strategies  of  animals  in  annual 
temporary  pools.  Arch.  Hydrobiol.  Suppl.  Bd.  58. 
206pp. 

WILLIAMS,  G.  1980.  An  index  for  the  ranking  of  wildfowl 
habitats,  as  applied  to  eleven  sites  in  West  Surrey, 
England.  Biol.  Conserv.  18:93-99. 

YEAGER,  L.E.  and  H.M.  SWOPE.  1956.  Waterfowl  produc- 
tion during  wet  and  dry  years  in  north-central  Colo- 
rado. J.  Wildl.  Manage.  20:442-446. 

ZEDLER,  J.B.  1982.  The  ecology  of  southern  California 
coastal  salt  marshes:  a  community  profile.  U.S.  Dep.  In- 
ter., Fish  and  Wildl.  Serv.  FWS/OBS-81/54.  1 10pp. 

ZIMMERMAN,  J. L.  1977.  Management  of  migrator)'  shore 
and  upland  game  birds.  Pages  45-46  in  Sanderson, 
G.C.,  ed.  Virginia  Rail  (Rallus  limicola).  International 
Assoc.  Fish  and  Wildlife  Agencies,  Washington,  DC. 
358pp. 


224 


Marshes 


11 


STREAMS 


Paul  Cuplin 

U.S.  Bureau  of  Land  Management 
Service  Center 
Denver,  CO  80225 


"I  love  any  discourse  of  rivers,  and  fish,  and  fishing. 
— Izaak  Walton,  Compleat  Angler 


Editor's  Note:  This  chapter  is  the  first  of  two  on 
aquatic  habitats,  and  covers  moving  water  habitat 
(rivers  and  streams).  The  emphasis  is  on  the 
smaller  streams  typical  of  the  western  United  States, 
as  opposed  to  large  systems  such  as  the  Mississippi 
River.  As  with  all  other  natural  systems,  the  streams 
cannot  be  understood  or  measured  without  looking 
at  the  associated  wetlands  and  uplands.  Therefore, 
stream  inventory  and  monitoring  involves  measur- 
ing or  estimating  factors  such  as  stream  bank  sta- 
bility and  dominant  riparian  vegetation.  Never- 
theless, the  focus  of  a  study  is  often  the  stream  and 
its  potential  for  supporting  fish  and  other  aquatic 
vertebrates.  This  chapter  covers  the  basic  measure- 
ments and  measurement  systems  for  describing 
stream  habitat. 


INTRODUCTION 


Eighty-five  thousand  miles  of  fishing  streams 
occur  on  public  lands  managed  by  the  U.S.  Bureau  of 
Land  Management  ( BLM )  in  the  1 1  western  States 
and  Alaska  (Table  1 ).  Streams  provide  recreational 
opportunities  ranging  from  white-water  boating  to 
family  picnicking.  Important  wildlife  habitat  associ- 
ated with  streams  include  riparian  vegetation  which 
provides  food  and  cover  for  many  wildlife  species. 
Fresh-water  fishing  in  terms  of  sportsfishing  dollars  is 
estimated  to  be  $11  per  fisherman  per  day.  About 
36.4  million  Americans  made  620.5  million  fishing 
trips  in  1980.  Total  dollars  expended  annually  for 
fresh-water  fishing  in  the  U.S.  is  estimated  to  be  $7.8 
billion  (U.S.  Department  of  the  Interior,  Fish  and 
Wildlife  Service  and  U.S.  Department  of  Commerce 
1982).  Salmon  and  steelhead  originating  from  public 
land  streams  that  are  taken  in  commercial  fisheries  of 
Alaska  and  Oregon  totaled  58,358,000  pounds  in 
1983  (U.S.  Department  of  the  Interior,  Bureau  of 
Land  Management  1983).  Commercial  fisheries  catch 
statistics  can  be  related  to  dollar  values  by  fish 
species. 

Stream  habitat  management  requires  an  aware- 
ness of  land  uses  that  can  improve,  stabilize,  or  de- 
stroy stream  habitat.  A  starting  point  is  a  baseline 
description  of  existing  habitat  conditions.  Changes  in 
habitat  conditions  are  detected  by  monitoring  spe- 
cific attributes  of  the  stream  over  time.  Stream  habi- 
tat features,  i.e.,  stream  bank  stability,  streambed 
siltation,  stream  channel  stability,  shade,  water  qual- 
ity, stream  width  and  depth,  and  other  features  are 
monitored  to  detect  changes  in  stream  habitat 
conditions. 


STREAM  CLASSIFICATION 

Although  classification  of  streams  has  received 
some  attention  in  the  U.S.,  there  is  no  universally 
accepted  method  of  classification.  The  U.S.  Fish  and 


Streams 


225 


Wildlife  Service  has  developed  a  wetland  and  deep- 
water  classification  for  a  national  wetland  inventory 
(Cowardin  et  al.  1979)  to  comply  with  the  Clean 
Water  Act  of  1977.  This  classification  is  very  useful 
for  broad-base  wetland  inventory  and  mapping,  but  it 
does  not  provide  information  on  condition  that  is 
needed  by  the  public  land  resource  specialist  or 
manager. 

A  stream  classification  would  be  useful  for  com- 
paring streams  that  are  similar  in  size,  water  quality, 
elevation,  and  geomorphology.  Common  classifica- 
tion terminology  would  allow  for  predicting  aquatic 
habitat  potential  relative  to  aquatic  habitat 
improvement. 

Stream  order,  which  provides  a  concept  of 
stream  size  and  characteristics  for  the  biologist  and 
manager,  was  first  defined  by  Horton  (1945).  He 
designated  unbranched  tributaries  as  first-order 
streams;  streams  that  receive  first-order  tributaries,  as 
second-order  streams;  those  that  receive  second-  or 
first-  and  second-order  streams  as  third-order 
streams,  and  so  on  until  the  mouth  of  the  stream  is 
reached.  Strahler  (1952)  modified  Horton's  stream 
order  and  called  all  unbranched  tributaries  first-order 
streams;  two  first-order  streams  join  to  make  a  sec- 
ond-order stream,  and  so  on  downstream  to  the 
stream  mouth  (Figure  1). 


Figure  1.     Stream  order  by  A.N.  Strahler  ( 1952). 


Table  1.     Estimated  miles  of  existing  fishing  streams,  and  acres  of  lakes  and  reservoirs  on  public 
lands  administered  by  U.S.  Bureau  of  Land  Management  ( 1983). 


State 

Lakes1 

Reservoirs1 

Streams2 

Alaska 

3,874 

0 

65,000 

Arizona 

1 

26 

439 

California 

19 

2 

818 

Colorado 

1 

19 

1,822 

Idaho 

10 

39 

3,577 

Montana 

16 

22 

996 

Nevada 

23 

5 

1,134 

New  Mexico 

2 

3 

176 

Oregon, 

Washington 

29 

18 

7,136 

Utah 

1 

8 

2,300 

Wyoming 

7 

32 

2,537 

Total 

3,983 

174 

85,935 

1ln  thousands  of  acres. 

includes  only  miles  of  fishable  streams. 


226 


Streams 


Third-order  and  larger  streams  generally  have 
gradients  less  than  0.5%,  and  streambed  is  usually 
rubble,  gravel,  sand,  silt,  and  organic  detritus.  These 
streams  are  meandering  with  abundant  riparian 
shrubs  and  trees;  stream  stability  is  generally  fair. 
Sediment  discharge  is  higher  than  the  smaller  first- 
and  second-order  streams.  Fishery  habitat  is  good. 

Map  scale  influences  determination  of  stream  or- 
der. For  example,  a  l:500,000-scale  map  would  not 
show  most  first-  and  second-order  streams.  Although 
l:24,000-scale  maps  will  show  first-order  and  larger 
streams,  Boehne  and  House  (1983)  suggested 
1:12,000  as  a  minimum  scale  for  mapping  stream 
orders  of  Oregon  streams. 

Stream  order  descriptions  vary  within  physio- 
graphic regions.  Streams  in  mountainous  regions  are 
used  as  an  example  for  a  general  description  of 
stream  orders. 

In  my  experience,  first-  and  second-order 
streams  generally  have  steep  gradients  (  3%  or  more ) 
and  large  coarse  streambed  and  streambank  material. 
Stream  meander  is  moderate;  riparian  shrubs  and 
trees  are  present;  and  debris  jams  are  common.  Fish- 
ery habitat  is  suitable  in  some  reaches  for  spawning 
and  rearing. 


STREAM  HABITAT  FEATURES 

Stream  habitat  feature  is  a  convenient  term  used 
to  identify  the  variables  that  make  up  stream  habitat. 
Stream  habitat  features  are  streamflow  patterns, 
streambank  stability,  stream  channel  stability,  ripar- 
ian vegetation,  riffles,  pools,  streambed,  stream 
depth,  stream  width,  stream  gradient,  stream  diver- 
sity, stream  water  chemistry,  and  macroinvertebrates. 

Streamflow  Pattern 

Streamflow  pattern  significantly  affects  the  biotic 
life  of  the  stream.  The  most  desirable  streamflow 
pattern  for  aquatic  production  is  one  of  perennial 
flow  with  moderate  spring  runoff.  Perennial  streams 
can  provide  year-round  spawning,  rearing,  and  feed- 
ing requirements  for  fish.  However,  extreme  drought 
may  cause  drying  of  perennial  streams.  The  type  of 
spring  runoff  will  govern  habitat  stability. 

Intermittent  streams  flow  only  part  of  each  year. 
Generally  streamflow  occurs  during  the  rainy  season 
or  during  snow  melt;  otherwise  the  streambed  is  dry. 
Intermittent  streams  are  indicated  by  a  dashed  line 
on  U.S.  Geological  Survey  (USGS)  7.5-minute  quad 
maps.  Intermittent  streams  often  provide  spawning 
and  rearing  habitat  for  fish  before  the  streambed 
dries  up  during  the  summer  months. 


Ephemeral  streams  are  those  that  flow  only  in 
response  to  a  rain  shower.  The  ephemeral  stream 
provides  marginal  habitat  for  aquatic  life  but  it  may 
be  very  important  to  amphibians. 

Instream  Flow  Needs 

The  amount  of  water  needed  to  maintain  stream 
habitat  on  a  year-round  basis  is  termed  "instream 
flow  needs."  It  can  be  determined  by  several  meth- 
ods as  summarized  by  Cuplin  et  al.  ( 1979).  Where 
detailed,  precise,  legally  defensible  information  is 
required,  the  multiple  transect-incremental  flow 
method  is  recommended.  This  method  is  described 
in  detail  in  Bovee  and  Milhous  ( 1978). 

For  rapid  assessment,  a  single  transect  method  is 
available  (Cuplin  et  al.  1979).  This  technique  re- 
quires less  time  for  field  work  and  data  analysis  but 
it  produces  less  precise  results  than  the  incremental 
flow  method. 

Computer  programs  have  been  developed  for 
analyzing  stream  channel  cross-section  survey  data 
collected  in  conjunction  with  instream  flow  assess- 
ments (Parsons  and  Hudson  1985).  A  single  transect 
method  can  be  used  for  rapid  assessment.  A  com- 
puter analysis  program  is  available  for  BLM  offices 
through  the  Denver  Service  Center,  Division  of  Re- 
source Systems  (Parsons  and  Hudson  1985). 

The  flow  of  a  stream  is  measured  in  cubic  feet 
per  second  (cfs)  to  provide  information  on  stream 
water  volume.  Water  flow  is  estimated  to  the  nearest 
cfs  during  stream  inventory.  The  U.S.  Geological 
Survey  maintains  permanent  streamflow  measuring 
weirs  and  annually  publishes  surface  water  reports 
for  each  State.  These  reports  include  data  on  stream- 
flow  and  water  chemistry  of  selected  streams. 

Streambank  Stability 

Stable  streambanks  and  abundant  riparian  vege- 
tation provide  good  habitat  for  aquatic  and  terrestrial 
wildlife.  Aquatic  systems  depend  on  the  condition 
of  the  adjacent  land.  Removal  of  riparian  vegetation 
through  timber  harvest,  road  construction,  overgraz- 
ing, or  excessive  recreational  use  can  cause  unstable 
streambank  conditions.  Eroding  streambanks  allow 
sediment  to  enter  the  stream  and  degrade  the 
streambed.  Silt  and  sediment  fill  the  spaces  between 
streambed  gravel  and  rubble  and  prevent  the  flow 
of  water  and  oxygen  necessary  for  the  survival  of  fish 
eggs,  larval  fish,  and  macroinvertebrates. 

Stream  Channel  Stability- 
Stream  channel  stability  is  related  to  the  water 
flow  of  the  stream  and  the  sediment  load  carried  by 
the  stream.  Material  that  enters  the  stream  from  the 


Streams 


227 


watershed  or  streambanks  is  either  washed  down- 
stream or  becomes  part  of  the  bed  material  ( Bovee 
1982).  Schumm  and  Meyer  (1979)  related  the  shape 
of  the  stream  channel  to  the  amount  and  kind  of 
load  it  carries.  Channels  can  be  classified  as  sus- 
pended load  (silt  and  clay),  mixed  load,  or  bed  load 
channels  (sand  and  gravel).  Five  channel  patterns 
have  been  identified  which  correspond  to  each  of 
the  types  of  load  and  transitions  between  load  types 
(Figure  2;  Shen  et  al.  1981 ). 

Riparian  Vegetation 

Woody  riparian  vegetation,  shrubs  and  trees,  are 
most  important  to  smaller  streams.  Streamside 


woody  vegetation  provides  shade  that  prevents  ex- 
cessive warming  of  water  during  summer  months. 
Overhanging  trees  provide  cover  for  fish  and  the 
leaves  and  twigs  provide  stream  energy  through 
nutrient  cycling.  Riparian  vegetation  provides  a 
buffer  from  upland  activities  as  well  as  a  filter  for 
overland  soil  erosion. 

A  method  of  predicting  riparian  vegetation  po- 
tential has  been  developed  by  Crouse  and  Kindschy 
(1981).  This  method  provides  guidelines  for  predict- 
ing the  riparian  vegetation  that  can  exist  under  var- 
ious conditions  of  soil  and  water  flow  if  protection 
through  fencing  or  grazing  management  system  is 
applied  to  a  stream/riparian  area. 


CHANNEL  TYPE 

Suspended 
Load 

Mixed  Load 

Bed  Load 

h- 

> 

*, *»=!* 

X 
CD 

CO 

CD 
3 

T. 

< 

§  £ 

3; 

CO 

O 

DC 

o  o 

I 

h- 
co 

-J  _l 

CD 
Co 

-^£^*£ 

Z 
DC 

i((&£)  /£ 

\ 

CD     CO 
O     3 

2 

CD 
CO 

3 

m 

LU 
(- 

< 

CD 

o 

CO 
DC  _ 

ajw\ 

^1  fe 

^ 

o 

c 

a. 

CD 

r- 
> 

< 

LL 

DC 

_c   ^ 

'         IK         1  i 

m 

LU 

±3    CD 

1   Q)w.       J    1 

\    V 

_J 

Q 

m-O 

V <*ttB^  ) 

\  v- 

05 

LU 

Z 

< 

Ll2 

v^y 

V. 

o 

S 

> 

z 
< 

X 

o 

LU 
2 

L 

CD     ^t 

o 

c 

CO 

3 
Q. 
CD 

CO 

r~ 

H 

-< 

Q 

> 

O 

3 

I — 

^-——+.            ^-^-~"^ 

LU 

D 

< 

DC 
CD 

< 

c 

o' 

3 

3 
3 
CD 

CO 

3; 

O 

Small 

Sediment  Size 

Large 

Small 

Sediment  Load 

Large 

Low 

Flow  Velocity 

High 

""     ""*— —  Chanr 

el  Boundary 

Low 

Stream  Power 

High 

,-*•""  "*>    Flow 

(3%  >)  Low 

Bed  Load  Total  Load  Ratio                     High  (>  11%) 

tfTgg^  Bars 

LOW 

RELATIVE  STABILITY                         HIGH 

Figure  2.     Channel  classification  showing  relative  stability  and  types  of  hazards  encountered  with  each  pat- 
tern (from  Shen  et  al.  1981 ). 


228 


Streams 


Typical  riparian  vegetation  along  a  foothills'  stream. 


Low  gradient  stream. 


Riffles 

Riffles  are  the  steeper  gradient,  high-water  ve- 
locity stream  sections.  Macroinvertebrates  reproduce 
in  the  riffle,  and  fish  feed  on  them  at  the  base  of 
riffles  in  either  pools  or  runs.  Steep  gradient  streams 
that  are  largely  composed  of  riffles  provide  food 
and  some  spawning  area  but  not  the  necessary  rest- 
ing and  shelter  mix  to  provide  good  habitat  for  sal- 
monids.  Other  fishes,  e.g.,  certain  suckers  and 
minnows,  are  almost  exclusively  found  over  riffles. 

Pools 

Pools  are  the  deeper,  slow-moving-water  por- 
tions of  streams.  Pools  can  be  classified  by  width  and 
depth  (Duff  and  Cooper  1978;  Hamilton  and  Berger- 
sen  1984)  and  they  provide  resting  areas  for  fish. 

Runs 

Runs  are  shallow  troughs  that  generally  have 
sand  or  gravel  bottoms.  Runs  are  smooth,  laminar 
flows  of  slow  to  moderate  velocity,  intermediate 
between  a  pool  and  a  riffle. 

Streambed 

The  streambed  consists  of  various  amounts  and 
sizes  of  material.  These  are  classified  by  particle  size 
(Table  2).  The  amount  of  different-sized  particles 


that  compose/make  up  the  streambed  determines  the 
uses  of  the  stream  bottom  by  fish,  macroinverte- 
brates, and  other  aquatic  life.  For  example,  coarse 
gravel  (2.5-7.5  cm  [1-3  in.  diameter])  is  most  com- 
monly used  by  salmonids  for  spawning,  whereas 
small  rubble  (7.5-15.0  cm  [3-6  in.  diameter])  free  of 
silt  and  sand,  provides  good  habitat  for  macroinver- 
tebrates. Silt  and  sand  inhibit  the  flow  of  water  and 
oxygen  through  rocks  in  the  streambed  and  reduce 
the  success  of  fish  egg  incubation  and  aquatic  orga- 
nism survival. 

Table  2.     Classification  of  streambed  material  (Duff 
and  Cooper  1976). 


Material 

Particle  Size 

Bedrock 

Exposed  solid  rock 

Boulder 

12"  (305  mm) 

Large  rubble 

6-12"  (152-305  mm) 

Small  rubble 

3-6"  (76-152  mm) 

Coarse  gravel 

1  -3"  (25-76  mm) 

Fine  gravel 

0.1-1.0"  (2.5-25  mm) 

Sand 

(0.074-2.5  mm) 

Clay 

(0.074  mm) 

Sapropel 

Reduced  organic  matter 

Detritus 

Particulate  organic  matter 

Streams 


229 


Stream  Width  and  Depth 

The  average  width  of  a  stream  is  the  average 
distance  between  the  water's  edges,  perpendicular  to 
the  stream  flow  (Hays  et  al.  1981).  Stream  depth  is 
the  average  vertical  distance  from  the  surface  to  the 
bottom  (Hays  et  al.  1981). 

Wide  shallow  streams  provide  reduced  habitat 
for  aquatic  life.  They  usually  lack  resting  cover  for 
fish  and  allow  water  temperature  to  rise  beyond  the 
tolerance  limits  of  some  fishes.  Narrow  deep  streams, 
in  comparison,  maintain  lower  water  temperatures 
because  of  less  surface  area  exposure  to  solar  warm- 
ing, and  the  deep  water  provides  good  resting  cover 
for  fish. 

Stream  Gradient 

A  steep,  torrential  mountain  stream  provides 
very  little  habitat  for  fish.  A  moderate  gradient  will 
provide  some  of  the  requirements  for  food  produc- 
tion, spawning,  and  rearing  of  fish.  A  very  low  gra- 
dient, deepwater  stream  provides  the  greatest 
stability  and  warmer  water.  Large,  low  gradient 
streams,  free  of  pollution,  provide  the  greatest  habi- 
tat diversity  and  the  largest  number  of  aquatic 
species. 

Stream  Water  Chemistry 

Suspended  Solids  and  Turbidity.  Suspended 
solids  in  the  form  of  silt,  clays,  and  other  fine 


materials  can  cause  temporary  to  permanent 
turbidity  or  murkiness. 

Prolonged  turbidity  caused  by  runoff  or  distur- 
bance in  normally  clear  streams  can  cover  the 
streambed  with  silt  which  can  smother  macroinver- 
tebrates,  cover  spawning  areas,  and  reduce  photosyn- 
thetic  rates.  Turbidity  can  cause  feeding  problems 
for  sight-feeding  trout  and  gill  irritation  to  most 
fishes  with  the  exception  of  those  that  are  adapted 
to  year-round  turbidity. 

Turbidity  caused  by  suspended  solids  can  be 
judged  in  terms  of  visibility  of  objects  in  water,  such 
as  muddy  water  (0.5-1  foot  of  visibility),  murky 
water  (1-5  feet  of  visibility),  and  clear  water  (5  feet 
of  visibility).  Turbidity  can  be  measured  and  re- 
ported in  Nephelometric  Turbidity  Units  (NTUs), 
which  are  measures  of  light  transmittance  in  a  water 
sample.  Turbidity  in  excess  of  25  NTUs  is  judged 
to  be  too  high  for  sight-feeding  fish. 

Water  Temperature.  Water  temperature  range  is 
an  important  criterion  of  the  resident  fish  and 
aquatic  species  in  a  stream.  Species  that  are 
acclimated  to  various  maximum  water  temperatures 
in  their  life  cycle  are  listed  in  Table  3  (U.S. 
Environmental  Protection  Agency  1971).  An  increase 
in  temperature  of  a  few  degrees  can  upset  the 
natural  balance  in  a  stream.  Fish  and  other  organisms 
exposed  to  maximum  temperatures  for  long  periods 
may  suffer  from  unusual  stresses,  disease,  and 
reduced  appetite. 


Table  3.     Maximum  temperature  probably  compatible  with  the  well-being  of  various  species  of  fish  and  their 
associated  biota  in  °C  and  °F  (U.S.  Environmental  Protection  Agency  1971 ). 


Temperature 

Taxa 

34°C  (93.2°F) 

Growth  of  catfish  (Ictaluridae),  gar  (Lepisosteus  sp.),  white  bass  (Morone  chrysops), 
yellow  bass  (M.  mississippiensis),  spotted  bass  (Micropterus  punctulatus),  buffalo 
{Ictiobus  sp),  carpsucker  (Carpiodes  sp),  threadfin  shad  (Dorosoma  petenense),  and 
gizzard  shad  (D.  cepedianum) 

32°C  (89.6°F) 

Growth  of  large-mouth  bass  (M.  salmoides),  drum  {Aplodinotus  grunniens),  bluegill 
(Lepomis  macrochirus),  and  crappie  (Pomoxis  sp.) 

29°C  (84.2T) 

Growth  of  pike  (Esox  lucius),  perch  {Percina  sp),  walleye  {Stizostedion  vitreum),  small- 
mouth  bass  (M.  dolomieui),  sauger  (S.  canadense),  California  killifish  (Fundulus  parvi- 
pinnis),  and  topsmelt  (Atherinops  affinis) 

27°C  (80.6°F) 

Spawning  and  egg  development  of  catfish,  buffalo,  threadfin  shad,  gizzard  shad,  Cali- 
fornia grunion  (Leuresthes  tenuis),  opaleye  (Girella  nigricans),  and  northern  swellfish 
(Sphoeroides  maculatus) 

24°C  (75.2°F) 

Spawning  and  egg  development  of  largemouth  bass,  white  and  yellow  bass,  spotted 
bass,  sea  lamprey  (Petromyzon  marinus),  alewife  {Alosa  pseudoharengus),  and  striped 
bass  {Morone  saxitilis) 

19°C  (66.2°F) 

Growth  of  migration  routes  of  salmonids  and  for  egg  development  of  perch,  smallmouth 
bass,  winter  flounder  (Pseudopleuronectes  americanus),  and  herring  (Clupea  sp.) 

12°C  (53.6°F) 

Spawning  and  egg  development  of  all  salmonidae  (other  than  lake  trout  [Salvelinus 
namayacush]) 

9°C  (48°F) 

Spawning  and  egg  development  of  lake  trout,  walleye,  northern  pike,  sauger,  and  Atlan- 
tic salmon  {Salmo  salar) 

230 


Streams 


Chemical  Variables.  The  three  most  important 
chemical  measurements  in  stream  water  are 
dissolved  oxygen,  pH  (hydrogen  ion  concentration), 
and  specific  conductance  which  can  be  converted 
to  represent  total  dissolved  solids  in  milligrams  per 
liter  (mg/L). 

Dissolved  Oxygen.  The  dissolved  oxygen  con- 
tent of  water  is  an  indicator  of  the  biochemical  con- 
dition of  water.  It  also  indicates  the  balance  between 
oxygen-consuming  and  oxygen-producing  processes. 
Fish  and  other  desirable  clean-water  biota  require 
relatively  high  dissolved  oxygen  levels  at  all  times. 
Dissolved  oxygen  levels  below  6  mg/L  are  dangerous 
to  sensitive  or  higher  elevation  fish.  Desert  fishes 
can  survive  and  withstand  low  oxygen  tensions  to  1 
mg/L. 

Streams  with  large  loads  of  organic  material  may 
have  oxygen-consuming  and  inorganic  reactions 
that  deplete  oxygen  to  levels  unfavorable  for  the 
clean-water  species. 

pH.  The  pH  is  the  abbreviation  for  the  negative 
base  10  log  of  the  hydrogen  ion  concentration. 
Stream  water  not  influenced  by  pollution  has  a  pH 
between  6.5  and  9.0,  which  is  the  acceptable  range 
of  pH  for  fish. 

Specific  Conductance.  Specific  conductance  is 
the  measure  of  the  ability  of  water  to  conduct  an 
electrical  current  and  is  expressed  in  micromhos  per 
centimeter  at  25°C  (7°F).  It  can  be  used  for  approxi- 
mating the  dissolved  solids  in  water  by  using  the 
formula — 

Specific  conductance  x  (0.65  ±  0.05)  =  mg/L 
dissolved  solids  or  total  dissolved  solids 
(see  below). 

The  formula  should  be  verified  by  comparing 
specific  conductance  with  total  dissloved  solids  de- 
termined by  a  chemist. 

Total  Dissolved  Solids  Total  dissolved  solids 
(TDS)  are  all  of  the  dissolved  materials  present  in 
natural  waters  and  consist  of  carbonates,  bicarbon- 
ates,  chlorides,  sulfates,  phosphates,  and  other  sub- 
stances. Most  productive  fresh  water  has  a  TDS 
above  350  mg/L.  The  maximum  safe  level  is  about 
1,500  mg/L  TDS  in  fresh  water.  Chemical  laboratory 
assistance  will  be  needed  for  the  analysis  of  heavy 
metals,  TDS,  pesticides,  and  pollutants. 


Macroinvertebrates 

Macroinvertebrates  are  aquatic  organisms  that 
can  be  seen  with  the  naked  eye,  such  as  mayfly,  stone- 
fly,  and  caddis  fly  larvae;  Gammarus  sp.,  aquatic 
worms,  beetles,  clams,  and  snails.  These  animals  pro- 
vide an  important  food  supply  for  fish. 


Macroinvertebrate  production  in  streams  is  re- 
lated to  water  temperature,  water  chemistry,  and 
habitat  diversity.  A  stream  that  is  low  in  TDS  and  has 
low  average  temperatures  will  produce  few  macroin- 
vertebrates per  square  foot  of  stream  bottom. 

Rock  and  rubble  7.5-15  cm  (3-6  in.)  in  diameter 
provide  good  habitat  for  macroinvertebrates.  Smaller 
rocks  from  2.5-7.5  cm  (1-3  in.)  provide  less  desir- 
able habitat,  hence  fewer  numbers  of  macroinverte- 
brates. Sand  and  silt  provide  little  habitat  except 
for  tubificid  worms  if  nutrient  levels  are  high. 

Plant-eating  macroinvertebrates  consume  algae, 
detritus,  diatoms,  wood,  and  leaves.  Some  macroin- 
vertebrates are  carnivorous  and  prey  on  other 
stream  dwellers. 

Stream  Diversity 

The  pools,  riffles,  runs,  boulders,  streambed 
gravel  and  rubble,  shade,  downed  trees,  and  gradient 
all  provide  habitat  diversity  in  a  stream.  Streams 
with  good  habitat  diversity  provide  the  essential  re- 
quirements of  food,  cover,  spawning,  and  rearing  for 
fish. 

Stream  habitat  improvement  enhances  diversity 
and  provides  the  missing  habitat  features  such  as 
pools,  streambed  spawning  gravel,  cover,  and  stream- 
bank  stabilization. 

An  example  of  diversity  destruction,  which  was 
once  a  common  practice,  is  the  removal  of  all 
downed  trees  and  logs  from  a  stream.  It  is  now  rec- 
ognized that  the  natural  tree  fall  into  a  stream  adds 
to  the  diversity  and  to  the  flow  of  energy  to  support 
the  food  chain. 


STREAM  INVENTORY  AND  MONITORING 
SYSTEMS 

Stream  Habitat  Inventory 

Streams  are  inventoried  to  determine  existing 
and  potential  aquatic  and  terrestrial  wildlife  habitat 
related  to  existing  or  planned  land  uses.  A  general 
system  that  can  be  used  for  photointerpretation  and 
on-the-ground  assessment  is  the  stream  habitat  in- 
ventory developed  by  the  Oregon  State  Office,  BLM 
(Cuplin  1981).  This  method  uses  five  criteria  (shade, 
riparian  zone  condition,  streambank  stability,  stream 
channel  stability,  and  sedimentation  of  streambed)  to 
classify  a  stream  as  poor,  fair,  good,  or  excellent. 
Each  of  these  attributes  is  rated  in  one  of  three  or 
four  categories  and  the  ratings  are  then  summed  to 
give  an  overall  stream  condition  rating  (Figure  3). 
This  classification  has  limited  applicability  on  south- 
ern desert  erosional  streams. 


Streams 


231 


CO 

b 


c 
g 

S 
"cd 

Q. 

a> 
c 

o 

o 

.c 

Q. 

o 

o 
c 

CD 
> 

0 


■D 
CD 
CO 

CD 

JD 

o 

E 
o 

CO 


CD 

co 


T3 

CD 

CD 
CO 

E 
to 

CD 


-C 
Q. 
CO 

w_ 

o 


I  I  * 

>  en  .5? 

C-  c  >- 

3  CD  CD 

CO  -J  < 


2  O 

s    z 

CD 

■4— • 

CO 


E 

CO 
CD 

CO 


c 

cn 

c 
'c 
c 
jo 

CL 


>> 

CD 

3 

CO 

32 

CD 
LL 


CD 

CD 

CD 

CD 

CO 

>h_ 

k~ 

co 

O 

o 

o 

O 

0 

E 

E 

E 

E 

o 

o 

o 

o 

o 

0s 

^5 

•4> 

O 

LO 

f— 

CD 

CD 

«fr 

CM 

■"3" 

CNJ 

CM 

CM 

C\J 

CNJ 

CJ 

CO 

CO 

^5 

.9 

^5 

^5 

O 

ID 

k_ 

o 

LD 

LO 

CD 

CM 

^— 

CM 

■ 

1 

i 

' 

O 

CD 

O 

T 

T— 

"* 

■<t 

co 

CO 

co 

CO 

co 

CO 

CO 

CO 

CO 

sS 

^5 

_CD 

0 

o 
co 

ID 

o 

O 

o 

0^ 

o 

CD 

o 

CD 

O 

CD 

CM 

■<* 

■* 

^t 

■t 

CO 
CO 

co 

CO 

CO 

CD 

CD 

CO 

o^ 

+ 
o 

CO 

o 

o 

1 

o 

O 
ID 

CD 
-Q 

o 

1- 
c 
E 

3 

o 
O 

C 

£>_ 

| 

o 

-^  "*"— 

CO 

"6 

c 
o 

Is 

CO  § 

CO 

o 

1 —  > 

o 

CO  CO 

CD  O 

^E, 

c-2 

c 
o 

O  O 

C  (0 

2o 

c-  CO 

COQ 
CD^ 

CD 
TJ 

CO 

SZ 
CO 

NCO 
C   CD 

co  CD 

■9-^p 
QC2^ 

CO 

c 

CD  — 

^5 

CO  2^ 

co£^ 

co^ 

CD 
O 

co 

Q. 

w 

£ 

.5? 

Q. 

O 

Q. 

Q. 
CO 


c 

LLI 


CD 
00 

_g 
co 
> 

LU 

E 

CO 
CD 

co 

o 


c 

CD 


c 
ro 

DC 

c 
o 

t5 
c 
o 
o 

E 

CO 
CD 

00 


Is-  CD 

*~  V  co 

c  "*  V  cn 

fl  CJ  >_     (-) 


CO 


luOllq. 


cd 


CD 
CO 

jo 
o 

c 
£ 
o 


co 
co 

"3 

n 


cn 

c 

T3 
CO 

>> 

n 


co  2 

CD    (0 

.       CL 


0)     t 
"D     CD 

b  t! 

-I 

CM   2 

_T   "CO 
CO    ^^ 

^-   o 

i_     c 

o   c 

—  CO 

c    ° 

co   c 
■c  o 

Q_    CO 

E  ^ 

—  a> 

CO    CD 

—  > 

CD    c 

^■g 

§1 

CD    <D 

y  53 

CD   -C 


232 


Streams 


Habitat  Quality  Index  (HQI) 

Binns  (1979)  developed  a  Habitat  Quality  Index 
(HQI)  which  relates  habitat  quality  to  trout  biomass 
in  streams.  He  field-tested  1 3  physical  habitat  attri- 
butes, 5  water  chemistry  attributes,  and  4  biological 
attributes  and  chose  9  attributes  for  the  HQI  (Table 
4).  The  attributes  chosen  for  the  HQI  are  late  sum- 
mer streamflow,  annual  streamflow  variation,  maxi- 
mum summer  stream  temperature,  water  velocity, 
cover,  stream  width,  eroding  banks,  substrate,  and  ni- 
trate nitrogen. 

The  HQI  is  given  by  the  expression: 

log10  (Y+  1)  =  [(  -0.903)  +  (0.807)log,„  (X,  +  1 )  + 
(0.877)  log  10(X2  +   1)  +  (1.233) 
log10(X3  +  1)  +  (0.631)  log10(F  + 
1)  +  (0.182)  log10(S  +    1)] 
[1.12085] 

Where. 

Y  =  Predicted  trout  standing  crop  (biomass) 

X,  =  Late  summer  stream  flow 

X2  =  Annual  stream  flow  variation 

X3  =  Maximum  summer  stream  temperature 

F  =Food  index  =  X,  (X4)  (X9)  (X10) 

S  =  Shelter  index  =  X7  (X8)  (X,, ) 

X4  =  Nitrate  nitrogen 

X7  =  Cover 


X8  =  Eroding  stream  banks 
X9  =  Substrate 
Xio  =  Water  velocity 
Xn  =  Stream  width 

Binns  (1979)  found  a  close  correlation  between 
predicted  and  measured  standing  crop  of  trout  in  36 
Wyoming  streams. 


Riparian  Aquatic  Information  Data  System 
(RAIDS) 

A  computer-based  system  for  summarizing  ripar- 
ian and  aquatic  data  has  been  developed  by  the  U.S. 
Bureau  of  Land  Management,  Service  Center.  This 
system,  called  the  Riparian  Aquatic  Information  Data 
System  (RAIDS),  can  accommodate  most  stream 
inventory  methods.  It  is  available  to  BLM  offices; 
more  detailed  information  can  be  obtained  from  the 
BLM  Service  Center,  Division  of  Resource  Systems. 


DISCUSSION 

A  recurring  problem  in  BLM  is  the  failure  of 
biologists  to  record  their  stream  inventory  data  in 
the  data  base.  Monitoring  changes  in  a  stream  eco- 
system requires  that  baseline  data  are  available  in  an 
analyzed  and  recorded  format.  Inventory  data  must 
be  on  a  specific  area  and  have  repeatable  variables, 
otherwise  monitoring  will  be  of  little  value.  Variables 
that  can  be  measured  with  good  to  excellent  repeat- 
ability over  time  are  stream  width,  stream  depth, 


Table  4.     Stream  habitat  attributes  selected  for  field  testing  during  development  of  a  stream  Habitat  Quality 
Index  (HQI)  for  trout  (Binns  1979). 


Attribute  Class 

Physical 

Chemical 

Biological 

Late  summer  stream  flow1 

Nitrate  nitrogen 

Streambank  vegetation 

Annual  stream  flow  variation 

Total  alkalinity 

Fish  food  abundance 

Maximum  summer  stream 
temperature 

Total  phosphorous 

Fish  food  diversity 

Water  velocity 

Total  dissolved  solids 

Fish  food  type 

Turbidity 
Cover 

Hydrogen  ion 

Stream  width 

Stream  depth 

Stream  morphology 
Eroding  banks 

Substrate 

Bed  material 

Silt  deposition 

Attributes  underlined  were  ones  selected  for 

use  in  the  HQI. 

Streams 


233 


steambank  undercut,  and  streambank  angle.  Variables 
that  have  poor  repeatability  are  percentage  pools, 
pool  quality,  percentage  riffle,  bank-to-bank  width, 
high  stream  water  width,  and  stream  rock  content 
(Table  5). 


LAND-USE  IMPACTS  ON  STREAM  HABITAT 

Land  uses  such  as  livestock  grazing,  timber  har- 
vest, road  construction,  mining;  and  recreation  can 
adversely  affect  stream  habitat  if  good  management 
practices  are  ignored. 

Timber  harvest  impacts  are  reduced  by  leaving 
buffer  strips  of  uncut  trees  in  a  75-ft.  or  wider  zone 
on  each  side  of  the  stream.  Stream  habitat  can  be 
destroyed  by  snagging,  log  dragging,  crossing  with 
heavy  equipment,  and  slash  and  sawdust  dumping. 

Livestock  grazing  in  the  riparian  zone  can  cause 
streambank  sloughing,  reduced  shrub  and  tree  repro- 
duction, compacted  soils,  and  increased  bacterial 
count  in  the  water. 

Hydraulic  mining  affects  all  stream  habitat  fea- 
tures as  well  as  the  riparian  zone,  especially  if 
streamside  settling  ponds  are  required.  Mining  explo- 
ration should  be  conducted  a  safe  distance  away 
from  a  stream  to  prevent  erosion  and  degradation  of 
the  streambed. 


Recreation  activities  such  as  picnics  and  over- 
night camping  facilities  should  be  carefully  designed 
to  prevent  water  pollution  and  deterioration  of  the 
riparian  zone. 

Road  construction  should  be  confined  to  areas 
other  than  riparian  zones.  Culvert  design  must  con- 
form to  accepted  practices  for  fish  passage. 

Stream  channelization  destroys  stream  diversity 
and  causes  long-term  damage  to  stream  productivity. 
The  impact  of  each  of  these  activities  on  stream 
habitat  features  is  summarized  in  Table  6. 

Streams  are  extremely  important  on  arid  lands. 
The  variety  of  uses  received  in  stream/riparian  areas 
are  compounded  by  the  attractiveness  of  these  areas 
to  all  users.  Livestock  and  wildlife  congregate  in 
stream/riparian  areas  because  of  the  availability  of 
food,  cover,  and  water.  Recreationists  enjoy  boating 
and  aesthetic  values,  and  fishermen  enjoy  the  pursuit 
of  their  sport.  Road  construction  is  the  least  costly 
in  the  riparian  zone.  Hydraulic  mining  must  of  neces- 
sity be  carried  on  in  the  streambed. 

These  competing  uses  require  biologists  to  have 
a  data  base  on  stream  habitat  conditions  as  well  as 
a  knowledge  of  the  impacts  from  all  land  uses  on 
such  conditions. 


Unregulated  logging  can  destroy  stream  habitat. 


Hydraulic  mining  residues  can  pollute  main  stem  streams 
in  addition  to  side  streams. 


234 


Streams 


Table  5.     Average  expected  repeatability,  precision,  and  confidence  intervals  of  water  column  and  streambank 
measurement  means  from  six  selected  streams  in  Idaho  and  Nevada.  Confidence  intervals  are  at  the  95% 
level  and  expressed  as  percentage  of  the  mean  (from  Platts  1981). 


Confidence 

Repeatability 

Item 

Interval 

Precision 

over  Time 

Water  Column 

Stream  width 

5.4 

9 

9 

Stream  depth 

8.2 

9 

©-• 

Streamside  water  depth 

16.6 

e 

e-9 

Pool  (percent) 

10.3 

9 

o 

Pool  (quality) 

8.0 

9 

o-e 

Riffle  (percent) 

12.5 

e 

o 

Sun  arc  angle 

1.1 

• 

9 

Bank-to-bank  width 

Very  wide 

o 

O 

High  water  stream  width 

Very  wide 

o 

o 

Streambank 

Soil  alteration 

12.3 

e 

Q-9 

Vegetative  stability 

3.1 

• 

e-9 

Undercut 

18.5 

e 

9 

Angle 

4.4 

• 

9 

Rock  content 

Very  wide 

o 

o 

Excellent      9  =  Good      0  =  Fair      Q  =  Poor 


Table  6.     Land  uses  in  stream/riparian  zones  and  their  impacts  on  stream  habitat  features  (from  American 
Fisheries  Society,  Western  Division  1982). 


Stream  Habitat  Feature 

Livestock 
Grazing 

Mining 

Road 
Construction 

Stream 
Channeli- 
zation 

Timber 
Harvest 

Recreation 

(Fishing, 

Picnicking) 

Stream  flow  pattern, 
increased  runoff 

Vegetation  removal, 
shade  and  cover 
reduced 

Stream  temperature 
increased 

Streambank  stability 
reduced 

• 

Channel  stability 
reduced 

Streambed  siltation 
increased 

• 

Stream  width  increased 

Stream  depth 
decreased 

Bacterial  count 
increased 

• 

Suspended  solids 
increased 

• 

• 

• 

Stream  diversity 
reduced 

• 

• 

• 

Streams 


235 


LITERATURE  CITED 


AMERICAN  FISHERIES  SOCIETY,  WESTERN  DIVISION. 
1982.  The  best  management  practices  for  the  man- 
agement protection  of  western  riparian  stream  eco- 
systems. Riparian  Habitat  Committee,  Western 
Division,  Am.  Fisheries  Soc.  45pp. 

BINNS,  N.A.  1979.  A  habitat  quality  index  for  Wyoming 
trout  streams.  Fishery  Research  Report  Monograph 
Series,  2.  Wyoming  Game  and  Fish  Dep.,  Cheyenne. 
75pp. 

BOEHNE,  P.L.  and  R.A.  HOUSE.  1983.  Stream  ordering:  a 
tool  for  land  managers  to  classify  western  Oregon 
streams.  U.S.  Dep.  Inter.,  Bur.  Land  Manage.  TN  OR  3, 
Oregon  State  Office,  Portland.  6pp. 

BOVEE,  K.D.  1982.  A  guide  to  stream  habitat  analysis 
using  the  instream  flow  incremental  methodology. 
Instream  Flow  Information  Paper  12.  Cooperative  In- 
stream Flow  Service  Group,  U.S.  Dep.  Inter.,  Fish 
and  Wildl.  Serv.,  Ft.  Collins,  CO.  FWS/OBS-82/26. 
248pp. 

and  R.  MILHOUS.  1978.  Hydraulic  simulation  in 

instream  flow  studies:  theory  and  techniques.  IFIP  5, 
Cooperative  Instream  Flow  Service  Group,  U.S.  Fish 
and  Wildl.  Serv.,  Ft.  Collins,  CO.  1 30pp. 

COWARDIN,  L.M.,  V.  CARTER,  F.C  GOLET,  and  ET. 

LAROE.  1979.  Classification  of  wetlands  and  deepwa- 
ter  habitats  of  the  United  States.  U.S.  Dep.  Inter.,  Fish 
and  Wildl.  Serv.,  Washington,  DC.  103pp. 

CROUSE,  M.R.  and  R.R.  KINDSCHY.  1981.  A  method  for 
predicting  riparian  vegetation  potential  of  semiarid 
rangelands.  Pages  110-116  in  Acquisition  and  Utiliza- 
tion of  Aquatic  Habitat  Information.  Western  Div.  of 
Amer.  Fisheries  Soc.  Portland,  OR. 

CUPLIN,  P.  1981.  The  use  of  large  scale,  color-infrared 
photography  for  stream  habitat  and  riparian  vegeta- 
tion inventory.  U.S.  Dep.  Inter.,  Bur.  Land  Manage., 
Denver,  CO.  57pp. 

,  B.  VAN  HAVEREN,  R.  BOROUICKA,  J.  ERDMANN, 

L.  LEE,  R.  MCQUISTEN,  and  M.  WHITTINGTON. 
1979.  Instream  flow  guidelines.  U.S.  Dep.  Inter.,  Bur. 
Land  Manage.,  Denver,  CO.  57pp. 


DUFF,  DA.  and  J.L.  COOPER.  1978.  Techniques  for  con- 
ducting stream  habitat  surveys  on  national  resource 
land.  U.S.  Dep.  Inter.,  Bur.  Land  Manage.  Denver,  CO. 
Tech.  Note  283-  72pp. 

HAMILTON,  K.  and  E.P.  BERGERSEN.  1984.  Methods  to 
estimate  aquatic  habitat  variables.  U.S.  Dep.  Inter., 
Bur.  Reclamation.  Eng.  Res.  Center.  Denver,  CO. 
1984. 

HAYS,  R.L.,  C.  SUMMERS,  and  W.  SEITZ.  1981.  Estimating 
wildlife  habitat  variables.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  FWS/OBS-81/47.  111pp. 

HORTON,  RE.  1945.  Erosional  development  of  streams 
and  their  drainage  basins — hydrophysical  approach  to 
quantitative  morphology.  Bull.  Geol.  Soc.  Amer. 
56:275-370. 

PARSONS,  S.C  and  S.  HUDSON.  1985.  Stream  channel 

cross-section  surveys  and  data  analysis.  Pam.  228.  U.S. 
Dep.  Inter.,  Bur.  Land  Manage.  Denver,  CO.  48pp. 

PLATTS,  W.S.  1981.  A  plea  for  fishery  habitat  classification. 
Fisheries  5(  1  ):2— 6. 

SCHUMM,  S.A.  and  D.F.  MEYER.  1979.  Morphology  of 
alluvial  rivers  of  the  Great  Plains.  Pages  9-14  in 
Agric.  Council  Publ.  91.  U.S.  Dep.  Agric,  For.  Serv. 
Fort  Collins,  CO. 

SHEN,  H.W.,  S.A.  SCHUMM,  J.D.  NELSON,  DO.  DOEHR- 
ING,  MM.  SKINNER,  and  GL.  SMITH.  1981.  Assess- 
ment of  stream-related  hazards  to  highways  and 
bridges.  Federal  Highway  Admin.,  Washington,  DC. 
24  lpp. 

STRAHLER,  A.N.  1952.  Dynamic  basis  of  geomorphology. 
Bull.  Geol.  Soc.  Am.  63:923-938. 

U.S.  DEPARTMENT  OF  THE  INTERIOR,  BUREAU  OF 
LAND  MANAGEMENT.  1983.  Public  land  statistics, 
Vol.  168.  U.S.  Govt.  Print.  Off.  Washington,  DC. 

U.S.  DEPARTMENT  OF  THE  INTERIOR,  FISH  AND 

WILDLIFE  SERVICE  and  U.S.  DEPARTMENT  OF  COM- 
MERCE. 1982.  1980  National  survey  of  fishing,  hunt- 
ing, and  wildlife-associated  recreation.  U.S.  Govt. 
Print.  Off.  Washington,  DC. 

U.S.  ENVIRONMENTAL  PROTECTION  AGENCY.  1971. 
Freshwater  biology  and  pollution  ecology  training 
manual.  Training  Manual,  Water  Quality  Office,  Wash- 
ington, DC. 


236 


Streams 


12 


LAKES 


James  F.  LaBounty 

U.S.  Bureau  of  Reclamation 
Denver,  CO  80225 


"A  lake  . . .  forms  a  little  world  within  itself — a  mi- 
crocosm within  which  all  elemental  forces  are  at 
work  and  the  play  of  life  goes  on  in  full,  but  on  so 
small  a  scale  as  to  bring  it  easily  within  the  mental 
grasp." 

— Stephen  A.  Forbes 


Editor's  Note:  This  second  chapter  on  aquatic  habi- 
tats covers  bodies  of  standing  water,  including 
lakes,  ponds,  and  reservoirs.  These  habitats  are  as 
important  to  the  many  aquatic  and  terrestrial  spe- 
cies that  depend  on  them  as  flowing  water  is  in 
the  arid  western  States.  However,  the  techniques  for 
inventorying  and  monitoring  these  habitats  and 
their  fauna  are  often  different  from  those  used  for 
flowing  bodies  of  water.  Entire  books  and  complete 
courses  have  been  devoted  to  the  study  and  meas- 
urement of  aquatic  life  in  lakes.  This  chapter  sum- 
marizes this  complex  subject  and  provides 
guidance  for  inventorying  and  monitoring  of  lakes. 


INTRODUCTION 

In  a  very  simple  sense,  lakes  are  bodies  of  stand- 
ing fresh  water.  In  reality,  they  are  complicated  sys- 
tems of  living  and  nonliving  features  that  influence 
each  other  immensely.  Lakes  may  be  shallow  or 
deep,  large  or  small,  but  the  features  that  give  them 
their  most  distinctive  characteristics  are  all  related 
to  their  standing  water.  A  lake  basin  can  be  thought 
of  as  a  trap  or  catchment  area,  both  for  water  and 
materials  from  within  the  watershed.  Typically,  lakes 
have  inlets  and  outlets  so  that  some  of  the  water 
and  materials  that  enter  a  lake  are  discharged  from 
the  lake  at  some  future  time. 


Even  though  one  can  go  back  to  a  particular 
lake  year  after  year  and  it  appears  to  remain  the 
same,  it  is  constantly  changing.  Lakes  are  only  tem- 
porary features  on  the  landscape.  Most  lakes  in  the 
U.S.  were  formed  between  10,000  and  12,000  years 
ago  after  the  last  retreat  of  the  glaciers  ( Figure  1 ; 
Odum  1971).  Others  were  formed  as  a  result  of 
volcanic  activity  or  a  change  in  the  earth's  crust 
(e.g.,  earthquakes).  Lakes  will  eventually  fill  with  the 
materials  that  accumulate  and  become  a  portion  of 
the  terrestrial  landscape.  The  life  span  of  lakes  varies 
considerably,  but  the  outcome  is  always  the  same. 
The  stage  of  development  of  a  lake  is  very  important 
in  preparing  management  strategies.  Generally,  there 
is  an  increase  in  the  complexity  and  diversity  of  life 
within  a  lake  as  it  ages. 

Most  consider  reservoirs  (Figure  2)  to  be  just 
another  kind  of  lake,  but  there  are  two  major  differ- 
ences between  reservoirs  and  natural  lakes  that 
should  be  considered.  First,  reservoirs  are  man-made 
and  their  water  levels  are  controlled.  Many  lakes 
have  control  structures  placed  on  their  outlets.  How- 
ever, the  degree  of  control  between  reservoirs  and 
lakes  is  very  different.  The  deepest  point  in  a  reser- 
voir is  generally  at  its  outlet — the  dam.  Therefore 
it  could  be  drained.  On  the  other  hand,  the  deepest 
point  of  a  lake  is  somewhere  near  its  center.  The 
lake's  natural  outlet  is  logically  the  shallowest  point 


Lakes 


237 


\^*>    .  ..-  * 


Figure     1.     A  mountain  lake  at  9,200  feet  in  elevation  in  Colorado. 


Figure     2.     Seminoe  Reservoir  located  in  the  high  plains  of  southern  Wyoming. 
238  Lakes 


in  the  body  of  water.  Thus,  without  some  major 
excavation  or  catastrophic  event  such  as  an  earth- 
quake, a  lake  could  not  be  emptied.  These  two  dif- 
ferences are  very  important  considerations  in  putting 
together  management  strategies  for  a  body  of  water. 


Ponds  (Figure  3)  can,  in  a  small  way,  exhibit 
all  of  the  characteristics  of  a  lake  or  a  reservoir. 
They  play  important  roles  in  watershed  management 
in  the  western  U.S. 


Figure    3.     A  pond  located  along  Lake  Creek  in 
central  Colorado. 


CLASSIFICATION  SYSTEMS 

Lakes  (Figure  1 )  possess  a  great  variety  of  com- 
binations of  properties  (Odum  1971).  This  makes 
it  difficult  to  select  any  one  basis  to  classify  lakes. 
Therefore,  one  can  find  all  sorts  of  terminology  ap- 
plied to  lake  types.  To  illustrate  this,  Hutchinson 
(1957)  listed  at  least  75  lake  types  based  on  geo- 
morphology  and  origin.  In  this  chapter,  Odum's 


1971  use  of  three-lake  categorization  schemes  will 
introduce  the  most  commonly  encountered  termi- 
nology as  applied  to  lake  types: 


( 1 )  the  oligotrophic-eutrophic  series  of  ordinary, 
clear-water  lakes  based  on  productivity; 

(2)  special  lake  types;  and 

(3)  impoundments. 

Oligotrophic — Eutrophic  Series 

These  terms,  implying  productivity  or  fertility, 
are  probably  the  most  commonly  used  to  describe 
lakes.  The  classic  oligotrophic  lake  is  one  that  has 
low  primary  productivity  and  is  relatively  deep.  In 
contrast,  eutrophic  ("good  foods")  lakes  are  shal- 
lower and  have  a  greater  primary  productivity.  The 
term  mesotrophic  implies  moderate  productivity, 
meaning  the  lake  falls  in  a  category  somewhere  be- 
tween oligotrophic  and  eutrophic. 

Other  categories — ultraoligotrophic,  oligomeso- 
trophic,  and  hypereutrophic — modify  the  above- 
mentioned  terms.  However,  emphasis  should  remain 
on  using  the  oligotrophic,  mesotrophic,  and  eu- 
trophic terms  because  they  will  suffice  when  refer- 
ring to  a  lake's  productivity.  Many  times  the  term 
eutrophic  is  improperly  used  as  a  synonym  for  pol- 
luted. Pollution  does  not  always  correlate  with  the 
process  of  eutrophication  when  referring  to  increas- 
ing the  nutrient  (nitrogen  and  phosphorus)  content 
of  a  lake  from  sewage  effluent.  A  lake  can,  in  fact, 
be  "polluted"  by  a  heavy  metal  or  other  toxic  sub- 
stance and  become  more  oligotrophic.  Care  and 
thought  should  then  be  made  before  using  the  term 
eutrophic  in  describing  a  lake,  even  though  in- 
creased eutrophication  of  a  lake  frequently  results 
from  untreated  or  little-treated  effluent,  such  as  run- 
off from  a  cattle  feed  lot  or  nearby  town. 

In  relatively  natural  situations,  lakes  become 
more  eutrophic  as  they  age.  Another  general  rule  is 
that  there  is  a  trend  of  increasing  productivity  (or 
eutrophication)  with  decreasing  depth  of  a  lake. 
Therefore,  a  rule  of  thumb  is  that  eutrophic  lakes  are 
shallow  and  have  greater  primary  productivity,  and 
oligotrophic  lakes  are  deep  and  have  lower 
productivity. 

Special  Lake  Types 

Odum  ( 1971 )  listed  seven  special  lake  types: 


( 1 )    Dystrophic  lakes  generally  have  high  concen- 
trations of  humic  acid.The  water  appears 
brown  and  the  lakes  eventually  develop  into 
peat  bogs. 


Lakes 


239 


(2)    Deep,  ancient  lakes  contain  animals  that  are 
found  nowhere  else  on  earth. 


(3)    Desert  salt  lakes  occur  in  sedimentary  drain- 
ages in  arid  climates  where  evaporation  ex- 
ceeds precipitation.  They  are  sumps  for 
drainages  where  salts  build  up  that  have  no 
outlets  such  as  the  Great  Salt  Lake.  Communi- 
ties of  few  species  are  the  rule  for  these 
lakes. 


(4)    Desert  alkali  lakes  occur  in  igneous  drainages 
in  arid  climates.  Being  alkaline,  their  pH  val- 
ues and  concentrations  of  carbonates  are 
high.  Pyramid  Lake,  Nevada  is  an  example. 

(  5  )   Volcanic  lakes  occur  in  areas  with  active 
volcanoes  and  may  be  acid  or  alkaline. 

(6)  Chemically  stratified  meromictic  or  partly- 
mixed  lakes  become  permanently  stratified 
due  to  such  things  as  intrusion  of  saline 
water  or  salts  liberated  from  sediments.  This 
sets  up  a  permanent  density  difference  be- 
tween surface  and  bottom  waters.  The 
boundary  between  the  upper,  circulating 
waters  and  the  lower,  non-circulating  layer  is 
termed  a  chemocline.  The  bottom  layers  are 
free  of  any  organisms.  Big  Soda  Lake,  Nevada 
is  an  example  of  a  meromictic  lake.  In  addi- 
tion, Flaming  Gorge  Reservoir  on  the  Green 
River  of  Utah  and  Wyoming  had  a  chemo- 
cline from  the  time  it  was  filled  in  1967  until 
1982.  The  chemocline  resulted  from  salts 
being  liberated  from  the  sediments. 

(7)  Polar  or  alpine  lakes  have  surface  tempera- 
tures that  remain  below  4°C  (39°F)  except 
for  brief  periods  during  the  ice-free  summer. 
Plankton  populations  grow  rapidly  and  store 
fat  for  the  long  winter.  These  lakes  are  not 
productive  enough  to  obtain  optimum 
growth  of  fish. 

Impoundments 

Impoundments  (Figure  2)  are  artificial  lakes 
created  by  placing  a  dam  on  a  river  or  stream.  Bea- 
ver ponds  are  the  smallest  example  in  this  category. 
A  tall  concrete  dam  such  as  the  Hoover  and  Glen 
Canyon  dams  on  the  Colorado  River  are  examples  of 
the  other  extreme.  The  limnological  characteristics 
of  an  impoundment  vary  according  to  the  nature 
of  the  drainage,  the  climate,  and  probably  most  of 
all,  the  operation  of  the  reservoir.  Wright  (1967) 
and  Odum  ( 1971 )  listed  a  number  of  effects  of 
dams.  Some  of  the  more  notable  include  the  fact  that 
water  is  released  from  the  surface  of  natural  lakes 


and  from  the  bottom  of  most  reservoirs.  Therefore, 
the  releases  from  the  deep  layers  of  a  reservoir  will 
have  higher  salinities,  be  colder  during  the  ice-free 
period,  have  a  lower  concentration  of  dissolved  oxy- 
gen, be  higher  in  some  of  the  reducing  compounds 
(such  as  hydrogen  sulfide),  and  have  higher  nutrient 
concentrations  than  would  normally  be  stored  in  a 
natural  lake.  The  advent  of  using  multilevel  outlets 
on  dams  has  resulted  in  the  moderation  of  some 
of  the  above  effects.  Nevertheless,  water  levels  of 
reservoirs  can  fluctuate  greatly  whereas  lake  levels 
remain  relatively  stable.  This  operational  difference, 
along  with  all  the  ramifications,  puts  impoundments 
in  a  category  of  lakes  that  require  different  manage- 
ment techniques  than  natural  lakes. 


CRITICAL  HABITAT  FEATURES 

The  physical  feature  that  exerts  the  most  direct 
control  on  a  lake's  characteristics  is  its  thermal  strati- 
fication. The  maximum  density  of  pure  water  is  just 
under  4°C  (39°F).  All  fresh  water  at  any  other  tem- 
perature is  lighter  and  found  on  the  surface.  Thus, 
ice  floats  and  in  summer,  the  upper  layer  of  a  lake  is 
warmest.  The  thermal  regime  of  a  lake  definitely 
reflects  its  ambient  temperatures  as  they  fluctuate 
over  a  day  and  over  a  season. 

The  following  is  a  scenario  of  a  temperate  lake. 
The  example  will  also  be  used  to  define  some  terms. 
Consider  a  deep  lake  whose  temperature  in  late 
spring  is  fairly  uniform  from  top  to  bottom  at  about 
4°C  (39°F).  As  the  day  lengthens,  solar  radiation  is 
absorbed  by  the  water,  mostly  by  the  upper  layers  of 
the  lake.  Light  intensity  decreases  with  increasing 
depth  until  the  point  is  reached  where  complete 
darkness  occurs.  Wind  mixes  the  surface  water 
rather  thoroughly  but  does  not  reach  down  to  the 
deeper  water.  As  heating  and  mixing  of  the  surface 
layers  continue  to  midsummer,  a  typical  pattern 
of  temperature  distribution  develops  (Figure  4).  The 
warm  top  layer,  which  is  heated  by  the  sun  and 
mixed  by  the  wind  and  other  currents,  is  called  the 
epi  limn  ion.  The  bottom  layer  (neither  heated  by 
the  sun  nor  mixed  by  the  wind)  is  called  the  hypo- 
limnion.  The  transition  between  the  two  is  the 
metalimnion  or  thermocline.  The  thermocline  is 
generally  recognized  where  the  water  temperature 
drops  at  a  rate  of  1°C  per  m  of  water  depth,  al- 
though this  is  a  textbook  definition  and  may  in  real- 
ity be  better  described  as  a  biological  barrier. 

Once  a  thermocline  has  formed,  the  two  layers 
are  virtually  separated  and  no  water  is  exchanged 
between  them;  even  the  current  patterns  within 
them  are  different.  These  layers  are  maintained 
strictly  because  warmer  water  is  lighter  than  colder 
water.  As  fall  progresses,  heat  is  lost  from  the  epilim- 
nion  to  the  atmosphere  faster  than  it  is  absorbed 
and  the  temperature  drops.  At  the  same  time,  the 


240 


Lakes 


4°                                                  Temperature 

u 

CD 

5 

/   Epilimnion 

^            Metalimnion,  or 
^_^--~^^~^                        Thermocline 

o 

c 

a 
CD 

2 

Hypolimnion 

Figure    4.     Typical  pattern  of  temperature  distribu- 
tion in  a  deep  lake  in  the  summer  zone  during 
the  summer. 

thermocline  continues  to  sink.  The  difference  be- 
tween water  temperatures  at  the  surface  and  at  the 
bottom  steadily  decreases  until  the  thermocline  no 
longer  exists  and  the  lake  "turns  over"  as  the  wind 
mixes  it. 

As  winter  intensifies,  stratification,  in  reverse  of 
summer,  is  setup.  Ice  forms  on  the  surface  of  the 
water  at  about  0°C  (32°F)  and  temperatures  at  the 
bottom  are  about  4°C  (39°F).  In  typical  temperate- 
zone  lakes,  the  winter  stratification  is  not  as  strongly 
developed  as  the  summer  stratification;  and  in  some 
areas  ice  never  forms  and  the  lakes,  more  or  less, 
continually  turn  over  the  entire  winter  due  to  wind 
driving  their  circulation. 

In  spring,  the  lake  again  absorbs  heat  from  the 
sun  and  its  temperature  rises  until  it  is  uniform  at 
about  4°C  (39°F)  from  top  to  bottom.  The  cycle 
begins  again.  Lakes  that  undergo  two  turnovers  per 
year  are  called  dimictic.  Those  that  turn  over  only 
once  and  never  manage  to  form  ice  are  called  mon- 
omictic.  There  is  a  range  of  lakes  found  in  each  of 
the  two  categories. 

The  determining  factors  for  the  complexity  of 
life  in  a  lake  are  a  combination  of  the  physical  and 
chemical  factors  in  and  around  a  lake.  Factors  other 
than  temperature  include  the  basin  configuration, 
depth  of  lake,  water  clarity,  dissolved  oxygen  and 
dissolved  solids,  inflow-outflow  volume  versus  lake 
volume  (flushing  rate),  nutrient  (e.g.,  nitrogen,  phos- 
phorus, and  silica),  and  a  myriad  of  other  chemical 
and  physical  factors  including  many  caused  by  man. 


Lakes  are  ephemeral  features,  considering  the 
geologic  time  scale.  The  length  of  their  life  is  ex- 
tremely variable,  but  they  all  progress  toward  the  eu- 
trophic  state.  This  rate  of  change  is  influenced 
tremendously  by  man's  activities.  Water  is  a  precious 
commodity  in  the  western  U.S.  and  numerous  factors 
compete  for  the  limited  supply.  Therefore,  man  has 
varying  degrees  of  influence  on  all  the  lakes,  ponds, 
and  reservoirs  in  the  western  U.S.  The  watershed 
that  supplies  water  to  lakes  may  have  cattle  grazing, 
which  causes  added  nutrient  input  and  erosion  of 
streambeds  from  vegetation  loss.  Added  nutrient 
input  tends  to  increase  eutrophication  of  a  lake 
whereas  erosion  of  streambeds  adds  turbidity.  Min- 
ing activities  in  a  watershed  may  add  heavy  metals  to 
a  lake.  Acid  mine  pollution  can  essentially  kill  a  lake. 
In  some  areas,  acid  mine  effluent  from  old  mine 
shafts  or  tailing  piles  has  been  continually  polluting 
area  streams  and  lakes  for  over  100  years.  The  ulti- 
mate influence  of  this  acid  mine  effluent  is  to  ac- 
tually suppress  a  lake's  development,  keeping  its 
living  component  at  a  minimum. 

Proper  management  of  the  watershed  will  avoid 
disasters  and  reduce  problems  that  may  later  be 
too  expensive  or  impractical  to  fix.  The  paramount 
consideration  is  that  a  lake  is  only  part  of  the  entire 
watershed;  therefore,  proper  management  of  the  lake 
alone  will  not  be  enough. 

Different  lakes  and  even  different  parts  of  a  lake 
have  different  communities  of  animal  and  plant  life. 
These  differences  are  functions  of  many  things,  in- 
cluding lake  size  and  depth,  substrate,  light,  thermal 
stratification,  and  geographic  location.  All  of  these 
factors  should  be  considered  in  the  management 
scheme  of  a  lake. 

MAIN  LIFE  ZONES  OF  A  LAKE 

Figure  5  diagrams  the  main  life  zones  within  a 
lake.  All  open  water,  where  the  bottom  is  too  deep 
to  be  inhabited  by  living  plants,  is  termed  the  pe- 
lagic zone.  Within  the  pelagic  zone  there  are  two 
subzones:  the  limnetic,  where  there  is  enough  light 
for  photosynthesis  to  occur,  and  the  profundal, 
the  dark  bottom  layers.  The  layer  of  most  biological 
importance  in  the  pelagic  zone  is  above  the  compen- 
sation point  where  light  is  just  sufficient  to  produce 
exactly  what  is  used  up  (e.g.,  photosynthesis  =  de- 
composition). Above  this  level,  autotrophs  (such 
as  algae)  can  produce  food  through  photosynthesis. 
These  autotrophs  form  the  base  of  the  food  web 
as  they  are  the  main  source  of  primary  production 
for  the  lake.  Diversity  of  life  in  the  limnetic  zone 
is  greatest.  Below  the  compensation  level  is  the  pro- 
fundal zone.  Here  autotrophs  cannot  produce  food 
(lack  of  enough  light)  and  the  main  source  of  energy 
is  detritus  that  "rains"  out  of  the  limnetic  zone.  All 
organisms  in  the  profundal  zone  are  called  hetero- 
trophs  and  are  either  detritus  feeders  or  carnivores. 


Lakes 


241 


Eulittoral 


Infralittora 


Figure     5.     The  main  life  zones  of  a  lake. 

One  final  way  a  lake  is  partitioned  is  by  the 
bottom.  The  lake  bottom  may  be  of  original  rock  in 
younger  lakes;  in  older  lakes  the  bottom  is  covered 
with  sediments  to  form  a  uniform  substrate  of  mud 
or  sand.  The  area  of  bottom  below  the  compensation 
level  is  the  profundal  benthic  zone.  The  area  from 
the  lower  limit  of  rooted  vegetation  to  the  shore  is 
known  as  the  littoral  zone.  The  littoral  zone  has  the 
lake's  greatest  diversity  both  in  habitat  and  commu- 
nity. The  littoral  zone  is  further  broken  down  into 
the  infralittoral  zone,  which  is  always  under  water 
and  contains  rooted  vegetation;  and  the  eulittoral 
zone,  which  is  nearest  the  shore  and  may  be  cov- 
ered by  water  only  part  of  the  time.  Much  erosion 
can  occur  in  this  zone  as  agitation  is  greatest. 


MAJOR  SPECIES  GROUPS 

Lakes  are  used  for  many  forms  of  recreation. 
These  include  swimming,  boating,  water  skiing,  sail- 
ing and,  of  course,  fishing.  As  the  population  of  the 
western  U.S.  increases,  so  does  the  fishing  pressure 
on  its  limited  number  of  lakes,  ponds,  and  reservoirs. 
The  fishery  must  then  be  closely  managed  by  stock- 
ing and  other  techniques.  Above  all  other  forms  of 
life,  fish  are  generally  the  primary  concern.  To  put 
this  into  a  proper  perspective,  along  with  the  fish 
themselves,  other  aspects  about  a  particular  lake 
govern  the  fishery  and  must  be  considered  and 
known. 

The  variety  of  animal  and  plant  assemblages 
found  in  lakes  is  as  great  as  the  variety  of  different 
kinds  of  lakes.  In  addition,  different  parts  of  a  lake 
have  different  characteristics  and  therefore  have 
different  animal  and  plant  communities.  Probably  the 
most  influential  factor  in  determining  the  kind  of 
plant  and  animal  communities  that  occur  in  a  lake  is 


temperature.  One  commonly  encounters  the  terms 
cold-water  fishery,  cool-water  fishery,  and  warm- 
water  fishery.  The  term  two-story  fishery  is  also  com- 
monly used  by  fishery  biologists.  The  following  are 
rules  of  thumb:  A  cold-water  fishery  in  the  western 
U.S.  usually  means  a  trout  fishery.  Therefore,  it  could 
include  any  kind  of  native  or  introduced  salmonid 
such  as  rainbow  (Salmo  gairdneri),  brown  (5.  trutta), 
or  brook  trout  (Salvelinus  fontinalis),  or  a  variety  of 
salmon.  A  cool-water  fishery  includes  walleye  (Sti- 
nostedion  rutreum),  sauger  (5.  canadense),  yellow 
perch  (Perca  flavescens),  northern  pike  (Esox  lu- 
cius),  and  muskellunge  (Esox  masquinongy)  (Tran- 
dahl  1978).  A  warm -water  fishery  in  the  western  U.S. 
usually  includes  spiny-rayed  fishes,  such  as  sunfish 
(Centrarchidae)  and  bass  (Micropterus  sp.),  and  also 
catfish,  bullheads  (Ictaluridae),  and  carp  (Cyprinus 
carpio).  A  two-story  fishery  is  one  that  includes  both 
cold-water  and  warm-water  fishes. 

The  situation  for  a  two-story  fishery  is  special  in 
that  thermal  stratification  during  the  summer  is 
strong  enough  that  the  epilimnion  is  generally  above 
20°  to  25°C  (68°  to  77°F)  in  midsummer  while  the 
hypolimnion  (lower  layer)  is  around  10°  to  15°C 
(50°  to  59°F).  For  a  successful  two-story  fishery, 
thermal  conditions  during  winter  cannot  be  so  se- 
vere as  to  violate  the  ecological  requirements  of 
warm-water  species.  In  addition,  the  proper  kinds 
and  amounts  of  food  must  be  available  for  the  cold- 
water  and  warm-water  fish.  Thus,  considerable  infor- 
mation on  a  particular  lake  must  be  available  to  de- 
termine if  it  will  sustain  a  two-story  fishery.  All  of 
the  fisheries  mentioned  are  found  throughout  the 
western  U.S. 

Both  recreationally  and  commercially,  fish  are 
the  most  important  living  component  in  a  lake.  How- 
ever, other  living  components  within  the  lake  dictate 


242 


Lakes 


what  the  fishery  is  and  how  successful  it  is  (or  could 
be  if  managed  correctly).  The  differences  in  commu- 
nities between  lakes  and  within  any  particular  lake 
are  functions  of  many  other  physical  and  chemical 
factors  besides  temperature.  These  include  substrate, 
light,  cation-anion  concentrations,  and  geographic 
position  of  communities  within  the  lake.  The  com- 
munities within  the  different  life  zones  (e.g.,  lim- 
netic, profundal,  pelagic,  and  littoral )  of  a  lake  are 
not  completely  separate,  as  some  species  overlap 
two  or  more  zones. 

The  upper  portion  (limnetic  zone)  of  the  open 
water  (pelagic)  area  is  of  the  most  obvious  biologi- 
cal importance,  as  it  is  where  autotrophs  can  pro- 
duce food  through  photosynthesis.  The  simplest 
form  of  life  in  the  food  chain  of  a  lake  would  most 
likely  be  found  in  its  upper  zone.  The  autotroph 
is  algae  that  converts  the  sun's  energy  and  inorganic 
material  to  living  material.  There  are  many  kinds  of 
algae,  ranging  from  diatoms  dominating  in  lakes  of 
low  production  to  blue-greens  that  dominate  more 
productive  lakes  during  midsummer.  The  algae  are 
grazed  on  by  microscopic  animals  collectively  called 
zooplankton.  These  range  from  relatively  small  roti- 
fers (10  microns)  to  fresh-water  shrimp  (20-30  mm 
[1  in.]  long)  that  feed  on  the  smaller  zooplankton. 

Zooplankton  provide  the  food  base  for  fish. 
Some  species  of  fish  only  eat  zooplankton  for  a  short 
time  in  their  life  cycle,  while  they  are  very  small. 
Other  fish  (commonly  called  filter  feeders  because 
they  filter  the  water)  feed  on  zooplankton  their  en- 
tire life.  The  next  predator  in  this  simple  food  chain 
is  a  fish  that  feeds  on  other  fish.  An  example  is  a 
largemouth  bass  feeding  on  minnows.  Man,  other 
fish-eating  mammals,  or  some  fish-eating  birds  would 
be  the  final  link  in  this  food  chain.  The  following 
illustration  is  a  sample  food  chain. 

Inorganic  Material 

I 

Light  -►  —  Algae 

I 

Zooplankton  (small) 

I 

Zooplankton  (large) 

T 

Fish  (small) 

i 

Fish  (large) 

t 

Mammals  or  Birds 


In  reality,  the  community  of  a  lake  is  a  food 
web.  That  is,  there  are  many  kinds  and  species  of 
plants  and  animals,  and  a  variety  of  conditions  and 
factors  that  result  in  a  web  pattern  versus  the  simple 
chain  described  above.  The  following  illustration  is 
a  sample  food  web. 


"^.Inorganic   Material^ 


ALGAE 
(Diatom) 
7  species 


T 


ALGAE 

(Green) 

4  species 


ALGAE 

(Yellow-brown) 

2  species 


~1 
ALGAE 
(Blue-green) 
2  species 


*     i     ir 

ZOOPLANKTON 
(Rotifers) 


ZOOPLANKTON 
(Copepods) 


tTI! 


ZOOPLANKTON 
(Cladoceran) 


1    f 


ZOOPLANKTON 
(Shrimp) 
1   species 


-FISH- 


t\- 


-^FISH--- 


1 
-FISH 


(Kokanee) 


(Lake  Trout)- 


(Rainbow  Trout)—1 

I 


-MAN 


The  benthic  and  littoral  components  of  the 
community  also  play  significant  roles  in  the  food 
web.  The  benthic  community  may  consist  of  insect 
larvae,  true  worms,  and  clams.  These  animals  break 
down  the  organic  material  (e.g.,  dead  fish)  that  falls 
to  the  bottom  into  the  inorganic  material  that  reen- 
ters the  food  web  when  the  lake  circulates  at  spring 
and  fall  turnover.  The  littoral  zone  contains  an  even 
greater  diversity  of  insect  larvae,  worms,  clams,  and 
rooted  and  floating  aquatic  plants.  The  littoral  zone 
often  provides  the  most  diverse  aquatic  habitat 
within  a  lake;  it  also  may  have  the  most  complex 
food  web  of  a  lake. 


INVENTORY  AND  MONITORING  METHODS 
FOR  LAKES 

A  number  of  useful  manuals  and  books  can  be 
consulted  for  methodology  of  studying  lakes.  They 
range  from  sections  of  limnology  and  fishery  text 
books  to  entire  manuals  devoted  to  methods.  Some 
of  the  more  common  manuals  or  textbooks  that 
include  methods  are  listed  in  Appendix  1 . 

The  U.S.  Geological  Survey  publication  entitled 
National  Handbook  of  Recommended  Methods 
for  Water  Data  Acquisition  ( 1977)  is  considered 
to  be  the  "bible"  by  users  and  collectors  of  data 
relating  to  hydrology,  chemical  quality,  groundwater, 
sediment,  biological  quality  soil  water,  evapotranspi- 
ration,  and  hydrometeorology.  It  should  be  referred 
to  by  anyone  surveying  or  monitoring  lakes,  ponds, 
reservoirs,  or  streams.  Compiled  by  agencies  that 
acquire  information  on  or  related  to  water  in  the 
U.S.,  it  contains  standardized  methodology  to  assure 


Lakes 


243 


greater  comparability,  compatibility,  and  usability 
of  water  data.  It  is  easily  obtainable  from  the  Office 
of  Water  Data  Coordination,  U.S.  Geological  Survey, 
Department  of  the  Interior,  Reston,  VA  22092. 

Levels  of  Inventory  or  Monitoring 

There  are,  of  course,  several  ways  (and  levels  of 
effort)  to  conduct  limnological  or  fishery  surveys  of 
lakes.  The  goals  and  needs  should  first  be  identified, 
which  will  dictate  both  the  type  of  survey  and  the 
level  of  effort.  If,  for  example,  several  hundred  lakes 
need  to  be  characterized  as  eutrophic  versus  oligo- 
trophic,  a  remote  sensing  survey  with  a  minimum  of 
ground-truthing  during  midsummer  could  provide 
the  information  necessary.  It  should  be  cautioned, 
however,  that  in  almost  every  situation  there  would 
not  be  enough  information  provided  from  a  survey 
of  this  type  to  make  judgments  as  to  the  fishery. 
Also,  a  lake  one  year  could  be  classified  as  eutrophic, 
whereas  another  year  it  could  not. 

The  most  difficult  level  of  effort  would  include  a 
long-term  (5-year),  twice-monthly  to  monthly,  sur- 
vey of  the  physical,  chemical,  and  biological  limnol- 
ogy and  frequent  netting  or  creel  census  data 
collection  of  each  lake  or  lakes.  The  long-term  sur- 
vey could  involve  expenditures  of  more  than 
$200,000  per  year,  whereas  the  remote  sensing  sur- 
vey may  cost  only  a  few  hundred  dollars.  Although  a 
prudent  manager  is  likely  to  opt  for  the  remote  sen- 
sing survey,  the  needs  and  long-term  usefulness  of 
the  data  should  be  considered.  The  following  are 
descriptions  of  several  levels  of  effort  that  can  be 
chosen  or  blended  to  fit  a  needed  situation.  Level  I  is 
the  least  amount  of  effort;  Level  IV  the  most. 

The  first  thing  that  needs  to  be  determined  is 
why  a  survey  is  needed.  When  that  question  is  an- 
swered, the  tests  to  be  run  and  samples  to  be  col- 
lected can  be  planned.  Upon  completion  of  the 
reconnaissance  survey,  when  the  data  are  in  hand,  it 
is  important  for  the  manager  to  realize  the  limitation 
of  the  data.  That  is,  a  one-shot  collection  of  data 
during  midsummer  reveals  little  of  the  lake's  ecology 
during  winter  or  even  for  the  month  before  and 
after  the  collection  of  data.  Nevertheless,  remember 
that  by  doing  the  reconnaissance  survey,  managers 
can  know  things  about  the  lake  not  previously 
known.  As  long  as  the  data  are  kept  in  proper  per- 
spective, such  a  survey  is  useful. 

Level  I — Minimum  Effort  Inventory.  This  level, 
which  is  a  simple  inventory,  can  also  be  described  as 
a  reconnaissance  survey.  Reconnaissance  surveys 
are  usually  used  to  determine  the  general  ecological 
trend  of  a  system  and  should  be  conducted  in 
midsummer.  They  may  also  be  used  to  identify  areas 
in  which  more  detailed  or  synoptic  surveys  should 
be  done. 


A  Level  I  survey  can  be  approached  in  various 
ways,  depending  upon  the  need.  If  there  is  need 
to  simply  characterize  a  lake  as  to  cold  water  versus 
warm  water,  measuring  depth  and  surface  and  bot- 
tom temperatures  during  midsummer  may  be  suffi- 
cient. If  there  is  need  to  characterize  the  trophic 
status  of  a  lake  or  several  lakes,  remote  sensing  may 
be  used,  or  carefully  collected  and  properly  handled 
water  samples  may  be  analyzed  for  phosphorus  and 
nitrogen  concentrations. 

If  there  is  a  need  to  know  the  quality  of  a  lake's 
fish  food  base  (i.e.,  plankton),  a  top  to  bottom  plank- 
ton sample  needs  to  be  collected  with  a  proper 
plankton  net  and  analyzed.  If  some  information  on  a 
lake's  fishery  needs  to  be  known,  creel-censusing 
(interviewing  fishermen)  or  spot-netting  with  appro- 
priate gill  nets  is  necessary. 

All  of  the  above  are  examples  of  information 
that  would  be  obtained  at  a  "one-shot"  reconnais- 
sance level  for  a  specific  need.  It  may  be  that  several 
things  need  to  be  known  about  a  lake;  therefore,  a 
reconnaissance  survey  would  include  many  of  the 
above  tests.  It  must  be  remembered  that  there  are 
several  approaches  to  obtaining  information  for  any 
one  particular  need.  I  recommend  that  the  National 
Handbook  of  Recommended  Methods  for  Water- 
Data  Acquisition  (U.S.  Geological  Survey  1977)  be 
consulted  for  each  test. 

Level  II — Maximum  Level  Inventory.  This  level 
would  also  be  classified  as  a  reconnaissance  survey, 
with  several  tests  done  to  obtain  more  knowledge  of 
the  subject  lake.  For  example,  suppose  a  manager 
wanted  to  know  the  general  characteristics  of  four 
or  five  lakes  so  that  future  studies  could  be  planned 
to  develop  an  intensively  managed  fishery.  An 
approach  I  recommend  would  involve  not  more  than 
four  workdays  per  lake,  preferably  on  the  same  day. 
Each  lake's  survey  would  be  in  two  parts — fishery 
and  limnology.  The  limnology  would  involve 
choosing  the  deepest  part  (or  parts  if  it  is  relatively 
large )  of  the  lake  and  collecting  some  critical  profile 
data  and  samples  there.  A  profile  of  temperatures 
each  1-2  m  (3. 3-6.6  ft)  from  surface  to  bottom  is 
essential;  dissolved  oxygen  concentration,  at  least  at 
the  bottom,  is  desirable. 

Also,  some  information  on  other  water  quality 
variables  is  desirable.  A  determination  must  be  made 
as  to  what  kind  of  water  quality  data  are  needed 
and  what  is  to  be  done  with  them.  This  will  deter- 
mine how  many  samples  are  collected  and  how 
these  samples  are  handled.  At  a  minimum,  a  sample 
or  samples  of  water  should  be  collected  from  at  least 
three  depths — surface,  mid-depth  (thermocline), 
and  near-bottom. 

Nitrogen  and  phosphorous  concentrations  are 
always  important  data  to  collect  as  they  indicate 


244 


Lakes 


trophic  status  of  a  body  of  water.  The  samples  for 
these  nutrient  analyses  must  be  collected  in  clean, 
approved  plastic  bottles,  with  no  contamination  from 
handling.  Usually  0.5  L  (0.5  qt)  is  a  sufficient  sample 
for  these  analyses.  After  collection  of  the  samples, 
they  need  to  be  analyzed  by  a  qualified  laboratory  or 
preserved  by  freezing  until  they  are  sent  to  a 
laboratory. 

Next,  a  sample  of  the  lake's  plankton  should  be 
collected  with  a  sock-like  plankton  net  which  is 
pulled  from  bottom  to  top  or  within  layers  of  the 
lake  (e.g.,  5  m  [16  ft]  to  surface;  10  m  [33  ft]  to  sur- 
face), depending  on  the  equipment  available  and 
the  need  for  a  detailed  study.  These  samples  provide 
information  on  the  food  base  for  fish.  The  samples 
need  to  be  placed  in  acceptable  bottles  (0.5  L  [0.5 
qt])  and  preserved  with  several  drops  of  formalin  or 
other  preservative.  The  samples  are  then  analyzed 
qualitatively  and  quantitatively,  if  desired,  in  the 
laboratory.  A  binocular,  low-power  microscope, 
standard  counting  cell,  and  simple  pipette  are 
needed  to  do  these  analyses. 

Next,  benthic  samples  of  the  lake  could  be  col- 
lected. This  involves  using  one  of  several  types  of 
dredges,  a  standard  No.  30  sieve  bucket,  and  some 
forceps.  The  bottom  mud  is  collected,  washed  in  the 
sieve  bucket,  and  the  sample  plus  residue  preserved 
in  a  bottle  with  a  few  milliliters  of  formalin.  These 
samples  will  reveal  many  things  about  the  lake,  such 
as  the  benthos  food  base,  past  stratification  strength, 
hypolimnion  chemistry,  and  general  status  of  the 
lake.  These  samples  would  also  be  analyzed  in  a 
laboratory. 

Finally,  in  this  brief  reconnaissance  limnological 
survey,  some  indication  of  light  penetration  is  desira- 
ble. The  simplest  way  to  do  this  is  with  a  Secchi 
disc.  This  is  a  circular  metal  plate,  20  cm  (8  in.)  in 
diameter,  the  upper  surface  of  which  is  divided  into 
quarter  sections  painted  black  and  white.  It  is  at- 
tached to  a  line  by  a  center  eyebolt.  The  point  at 
which  the  eyebolt  can  no  longer  be  seen  upon  low- 
ering is  the  Secchi  depth.  This  simple  piece  of  infor- 
mation is  extremely  valuable  since  light  is  one  of  the 
important  qualities  of  the  lake's  limnology.  The  lake 
in  question  can  be  easily  compared  to  others  by  this 
method. 

The  second  portion  of  this  reconnaissance  is  to 
examine  the  fishery.  If  no  data  exist,  a  gill-net  survey 
is  a  good  place  to  start.  This  involves  first  having 
the  proper  collecting  permit,  which  should  be  ob- 
tained well  ahead  of  time  from  the  state  game  and 
fish  agencies.  Assistance  from  the  agencies  is  desira- 
ble and  often  essential.  They  can  assist  in  using 
proper  equipment,  such  as  gill  nets,  shocking  boats, 
and  beach  seines.  Without  their  involvement,  a  sur- 
vey of  the  fishery  will  most  likely  lack  some  impor- 
tant knowledge. 


Often,  a  one-night  set  of  experimental  gill  nets 
in  some  strategic  locations  will  yield  information 
that  will  help  managers  make  some  preliminary  deci- 
sions as  to  what  kind  of  a  fishery  exists  and  what 
needs  to  be  done.  Sometimes  beach  seining  during 
late  summer  or  early  fall  can  give  fishery  biologists 
and  ichthyologists  a  good  indication  of  the  type  of 
fishery  existing  in  a  lake  where  little  is  known.  The 
Level  II  survey  involves  much  preparation  and  fol- 
low-up to  be  done  properly. 

Level  HI — Minimum  Effort  Monitoring.  This 
level  of  effort  involves  part  or  all  of  the  tasks 
described  under  Level  II  except  they  are  performed 
on  a  seasonal  basis,  that  is,  four  times  a  year  at  mid- 
season.  Again,  details  of  the  study  depend  on  the 
questions  being  asked.  The  Level  III  studies  are  the 
first  stage  of  synoptic  or  detailed  studies.  Instead 
of  an  inventory,  a  monitoring  study  is  being 
performed.  These  synoptic  or  detailed  studies  are 
done  to  determine  ecological  baselines  from  which 
change  can  be  determined  in  the  future.  In  general, 
the  premise  on  which  synoptic  surveys  are  based 
is  that  an  evaluation  of  kinds,  numbers,  and  relative 
abundance  of  species  enables  one  to  establish  the 
biological  health  of  a  lake,  pond,  or  reservoir  (or  any 
other  ecological  system ).  As  with  the  other  levels  of 
effort,  this  monitoring  should  be  attempted  only 
after  careful  planning  and  preparation.  If  several 
lakes  are  involved,  some  coordination  is  beneficial  so 
that  inter-lake  comparisons  can  be  made. 

If  the  needs  dictate  a  Level  III  effort,  costs  can 
add  up  to  tens  of  thousands  of  dollars  per  lake  per 
year.  However,  if  there  is  only  a  need  to  monitor 
lake  temperatures,  all  that  is  needed  is  a  boat  and 
motor,  a  thermometer  probe  with  enough  cable  to 
reach  the  bottom,  and  a  readout.  If  plankton  species 
need  to  be  documented,  all  that  is  needed  is  a  boat 
and  motor,  a  plankton  net,  a  collection  bottle,  and 
some  preservative.  Seasonal  samples  of  plankton 
would  then  reveal  not  only  if  certain  kinds  required 
for  certain  species  of  fish  are  present,  they  would 
reveal  many  things  about  the  lake's  health. 

Level  IV — Maximum  Effort  Monitoring.  Level  IV 
should  be  chosen  only  when  a  highly  synoptic  set  of 
data  is  needed.  Monitoring  studies  at  this  level 
become  more  involved  and  more  frequent.  The  main 
thing  is  that  they  are  more  synoptic,  and  routine 
field  sampling  dates  need  to  be  selected.  An  example 
would  be  every  other  month,  or  monthly  during 
turnover,  and  every  third  month  at  other  times.  A 
resource  specialist  or  manager  needs  to  ask  why  the 
data  are  needed,  when  the  most  significant 
information  can  be  collected,  and  how  much  and 
what  kind  of  data  are  needed  to  solve  the  problem. 
It  cannot  be  overemphasized  that  the  studies  must 
constantly  be  adjusted  and  fine-tuned  to  keep  them 
in  a  proper  perspective.  It  is  common  for  a  resource 


Lakes 


245 


specialist  to  convince  a  manager  of  the  need  to 
perform  a  certain  study,  yet  the  manager  may  not 
understand  the  utility  of  collecting  a  lot  of  data.  Both 
resource  specialists  and  managers  should  agree  how 
the  data  are  to  be  used  by  their  agency. 


For  example,  the  need  may  exist  to  monitor  a 
lake  that  is  involved  in  a  water  resources  develop- 
ment scheme.  Suppose  that  its  ecosystem  is  likely  to 
be  modified  considerably.  For  this  example,  consider 
that  the  lake  is  a  natural  subalpine  lake  with  a  sur- 
face area  of  about  700  ha  (1,750  a.).  Furthermore, 
consider  that  the  lake  is  to  be  raised  behind  a  newly 
constructed  dam.  In  addition,  a  pumped  storage 
power  plant  will  be  constructed  that  will  pump  lake 
water  150  m  (495  ft)  up  to  a  newly  constructed 
reservoir.  During  the  generation  phase,  water  will 
pass  from  the  upper  reservoir  through  the  power 
plant  back  to  the  natural  lake.  Current  environmental 
laws  require  that  the  environmental  effects  be  recog- 
nized and  some  values  protected.  Therefore,  the 
ecosystem  must  be  understood.  Limnology  and  fish- 
ery investigations  would  be  necessary  before,  during, 
and  after  changes  were  made.  The  lake's  food  chain 
should  be  studied  in  enough  detail  to  be  able  to  first 
predict,  then  detail  and  quantify  changes  to  the  lake 
that  are  accountable  to  the  development  project. 

The  actual  level  of  detail  of  each  study  will  vary 
from  location  to  location  and  will  be  dictated  by 
needs  identified  by  resource  specialists.  The  manuals 
and  textbooks  listed  in  Literature  Cited  should  be 
referred  to  as  a  guide  in  determining  how  the  out- 
lined needs  will  be  met.  One  of  the  more  difficult 
tasks  for  a  manager  is  knowing  how  much  and  what 
kind  of  information  are  necessary  for  a  particular 
situation.  Many  times,  not  enough  information  is  col- 
lected to  provide  managers  or  resource  specialists 
with  the  evidence  they  need  to  make  a  critical  deci- 
sion. This  not  only  can  do  an  injustice  to  the  re- 
source, it  can  be  embarrassing  for  the  decisionmaker. 
On  the  other  hand,  some  studies  become  "gold- 
plated"  and  the  information  may  not  apply  to  the 
need.  An  example  may  be  that  an  adequate  amount 
of  chemistry  data  is  available  on  a  particular  lake  for 
the  manager  to  determine  if  the  inflow  actually  is 
contributing  to  the  lake  pollution.  However,  these 
data  need  summarization  by  a  resource  specialist  so 
that  managers  can  use  them  in  their  decisions.  Sup- 
pose the  investigators  selected  to  do  the  summariza- 
tion are  involved  deeply  in  computer  modeling  and 
their  product  is  a  detailed  and  highly  technical 
model  of  the  lake.  There  have  been  instances  when 
the  manager  has  not  been  able  to  use  this  product 
because  it  is  too  technical. 


The  more  practical  way  to  handle  this  situation 
would  be  to  have  the  summarization  done  by  a  re- 
source specialist  who  will  also  know  the  practical 


need.  The  manager  may  be  less  technically  qualified 
than  a  professional  statistician  or  modeler,  but  the 
manager's  grasp  of  the  situation  will  allow  the  data 
to  be  useful  and  allow  a  decision  to  be  made.  This 
does  not  mean  that  models  are  not  useful.  The  point 
is  that  there  will  always  be  difficulty  in  knowing 
exactly  how  much  information  is  needed.  A  consen- 
sus of  resource  personnel  and  management  should 
be  sought  as  to  how  much  information  is  necessary 
to  make  a  wise  decision  regarding  a  lake,  pond,  or 
reservoir. 


Field  Techniques 

This  section  includes  a  summarization  of  some 
of  the  accepted  field  methods  and  the  types  of  gear 
needed  for  the  various  studies  that  are  done  for  the 
levels  described  in  the  previous  section.  Three  as- 
pects of  a  survey  are  used  to  facilitate  this  presenta- 
tion: physical-chemical  limnology,  biological 
limnology,  and  fishery. 

For  all  the  studies  listed  below,  a  boat,  outboard 
motor,  field  book,  and  pen  or  pencil  are  required. 
At  all  times,  approved  life  jackets  or  vests  are  re- 
quired equipment  for  a  survey.  The  following  sum- 
marizes the  requirements  for  specific  tests. 

Physical-chemical  surveys.  Collecting  profile 
information  on  a  lake  requires  some  instrument  s) 
that  measures  the  desired  features.  The  most 
primitive  way  is  to  lower  a  water  bottle  (described 
below)  to  each  desired  depth;  collect  a  sample; 
bring  it  to  the  surface;  and  measure  the  temperature, 
conductance,  and  dissolved  oxygen  (DO)  of  water 
in  the  bottles  with  a  simple  hand-held  thermometer 
and  other  portable  gear.  This  method  is  acceptable  if 
errors  of  up  to  several  degrees  temperature  can  be 
tolerated.  Also,  the  methodology  for  DO 
measurement  requires  clear  glass  bottles  that  hold 
about  300  ml  (9  oz)  and  have  glass  stoppers  (called 
BOD  bottles),  and  several  chemicals.  In  many 
situations,  this  sampling  method  is  adequate,  but  in 
others,  more  sophisticated  gear  is  needed.  This  gear 
may  be  very  complex  electronic  equipment  that 
measures  five  or  six  variables  at  once  (temperature, 
pH,  DO,  conductance,  oxidation-reduction  potential, 
and  depth)  and  stores  the  data  on  internal  memory 
chips. 

Practically  speaking,  a  multiparameter  probe 
(see  Figure  6)  is  a  fairly  basic  piece  of  equipment 
where  routine  lake  surveys  are  being  done.  These  in- 
struments cost  from  $4,000  to  $8,000,  but  are  accu- 
rate and  save  much  time  in  the  field  and  laboratory. 

Water  chemistry  data  generally  require  that 
water  samples  be  collected,  placed  into  acceptable 


246 


Lakes 


Figure    6.     A  multiparameter  probe  is  used  to  measure  profile  data  (courtesy  of  Hydrolab  Corp.). 
( 1 )  Digital  Display  Unit,  (2)  Sohole  Unit,  (3)  Circulatory  Assembly,  (4)  Cables,  (5)  Battery  Pack, 
and  (6)  Digital  Field  Datalogger. 


containers,  preserved,  and  taken  to  a  qualified  labora- 
tory for  analyses.  Portable  water  chemistry  kits  are 
available  and  the  data  obtained  from  their  use  are 
acceptable  for  some  measurements  when  the  analyst 
is  qualified.  However,  for  synoptic  studies  or  where 
data  are  required  for  critical  decisions,  the  samples 
should  be  carefully  collected  and  analyzed  by  a  qual- 
ified laboratory. 


Ideally,  the  sample  bottles  should  be  provided 
by  the  laboratory  that  does  the  analyses.  The  labora- 
tory should  be  consulted  as  to  the  amount  of  water 
they  need  for  the  requested  analysis.  A  water  bottle 
of  the  Van  Dorn  or  Kemmerer  types  (see  Figure 
7)  is  necessary  if  more  than  surface  samples  are  to 
be  collected.  The  water  bottle  must  be  attached  to  a 
reel  of  cable  of  sufficient  length.  Along  with  the 
water  bottle,  a  messenger  or  something  to  trigger  or 
close  the  bottle  at  the  desired  sampling  depth  is 
needed.  Samples  for  chemical  analyses  should  in- 
clude both  inflow  and  outflow  water  to  determine 
quantitatively  the  chemical  elements  that  flow  in  and 
out  of  a  lake  or  reservoir  and  those  that  are  retained. 
Finally,  the  proper  preservative  is  needed.  For  nor- 
mal cation  and  anion  analysis,  only  refrigeration  is 
needed;  for  nitrogen  and  phosphorus  nutrients,  freez- 
ing is  best;  and  for  most  heavy  metal  analyses,  lower- 
ing the  pH  with  nitric  acid  below  a  value  of  2  is 
recommended. 


Figure    7.     A  Kemmerer  water  sample  bottle  in  the 
open  position. 


Lakes 


247 


The  amount  of  light  that  enters  a  lake  is  an  im- 
portant quality.  It  can  be  measured  easily  with  a 
Secchi  disc.  If  more  quantitative  information  on  light 
is  desired,  instrumentation  called  limnophotometers 
may  be  used.  These  cost  between  $3,000  and  $5,000 
and  actually  measure  light  in  units  as  the  photocell 
is  lowered  into  the  water. 


Secchi  disc 


The  following  list  of  equipment  will  be  sufficient 
to  perform  a  physical-chemical  survey: 

•  map  of  the  lake  (optional); 

•  reel  with  enough  cable  or  rope  to  reach  desired 
depth; 

•  thermometer,  multiparameter  probe,  or  what- 
ever is  appropriate  to  measure  temperature, 
dissolved  oxygen,  conductance,  pH,  redox  po- 
tential, and  depth; 

•  water  sample  bottle  (Kemmerer  or  Van  Dorn 
type); 

•  appropriate  containers  to  hold  samples  for 
water  analysis; 

•  appropriate  preservative;  and 

•  messenger  (heavy  metal  weight  made  especially 
to  trigger  sampling  gear). 

Biological  Limnology.  The  selection  of  equipment 
here  is  as  broad  as  the  kinds  of  data  that  need  to  be 
collected.  The  following  is  by  no  means  exhaustive. 
A  limnological  text  should  be  consulted  before 
proceeding.  Some  examples  are  nevertheless  given. 

Primary  production  is  commonly  measured  by 
two  methods — chlorophyll  or  light  versus  dark  bot- 
tle studies.  An  analysis  of  chlorophyll  pigments  in 
a  known  volume  of  lake  water  will  provide  an  indi- 
cation of  the  algae  biomass.  This  then  gives  direct 
evidence  of  the  amount  of  primary  products  in  the 
lake.  The  entire  process  involves  access  to  a  labora- 
tory with  a  filtration  system,  spectrophotometer,  test 
tubes,  and  acetone.  Water  is  collected  and  stored 
for  a  short  period  until  the  samples  are  filtered  (usu- 


ally with  a  pressure  filter )  through  recommended 
fiberglass  filters.  These  are  kept  frozen  until  they  are 
put  into  a  test  tube  of  acetone  and  crushed.  The 
test  tubes  are  centrifuged  and  the  decanted  sample  is 
put  into  a  spectrophotometer.  Light  penetration  at 
wavelengths  630,  645,  and  667  is  recorded  and  the 
results  put  into  equations  found  in  Wood  (1975) 
or  Maitland  (1978). 

An  actual  measurement  of  photosynthetic  rate 
may  be  done  in  several  ways.  Most  involve  compar- 
ing activity  in  light  versus  dark  BOD  bottles  left  in 
the  lake  at  desired  depths  for  specified  times.  One 
common  method  is  to  measure  the  DO  before  and 
after  a  predetermined  time  period  in  both  dark  and 
light  bottles.  The  difference  in  dissolved  oxygen 
between  the  light  and  dark  bottle  then  represents  an 
indication  of  the  amount  of  primary  production. 
Specific  methodology  for  this  is  available  in  Wood 
(1975).  Another  commonly  used  method  is  even 
more  direct  and  involves  spiking  the  light  and  dark 
bottles  with  a  known  concentration  of  radioisotope 
tagged  source  of  carbon  (NaH  ^C03).  The  theory 
is  that  algae  during  photosyntheses  use  carbon- 1 2 
and  by  putting  tagged  carbon- 14  in,  filtering  the 
algae  later,  and  measuring  the  amount  of  radioactiv- 
ity, the  amount  of  carbon  uptake  by  algae  will  be 
known.  The  amount  of  carbon  assimilation  then  indi- 
cates the  primary  productivity  rate  at  the  time  of 
the  survey.  The  carbon- 14  methodology  requires  ex- 
pensive laboratory  equipment,  a  radioactive  source, 
and  individuals  licensed  by  the  Nuclear  Regulatory 
Commission.  Nevertheless,  when  done  properly,  it  is 
the  best  indication  of  a  lake's  primary  production 
rate. 

The  next  step  is  to  collect  the  plankton  (both 
phytoplankton  and  zooplankton)  and  measure  densi- 
ties (i.e.,  number  per  volume).  Again,  there  are  nu- 
merous ways  to  accomplish  this  task.  Probably  the 
most  common  method  is  to  use  a  closing  net  (Figure 
8).  This  is  a  sock-like  net  that  is  lowered  into  the 
water  column  to  various  depths.  It  strains  the  water 
as  it  is  raised  through  the  water  column.  A  cup  at 
the  lower  end  is  emptied  into  a  collection  jar  and 
preserved  with  a  few  drops  of  formalin,  alcohol,  or 
other  preservative. 

A  subsample  (e.g.,  1  ml)  of  the  plankton  is  put 
into  a  counting  cell  for  counting  and  identification 
under  a  microscope.  The  densities  of  each  kind  of 
plankton  are  then  calculated.  Other  commonly  used 
methods  include  towing  a  metered  Clark-Bumpus 
plankton  net  horizontally  a  short  distance  across  the 
lake  or  collecting  a  known  volume  of  water  with 
the  water  bottle  and  pouring  it  through  a  net.  All  of 
these  involve  counting  and  identifying  the  samples  in 
the  laboratory  by  qualified  personnel.  A  note  of  cau- 
tion: some  of  the  zooplankton  may  not  be  there 
when  and  where  sampling  is  done.  For  example,  the 
freshwater  shrimp  (Mysis  relicta)  that  live  in  cold- 


248 


Lakes 


water  lakes  are  nocturnal  and  will  not  be  collected 
during  daylight  hours.  Therefore,  adequate  sampling 
techniques  would  include  spatial  and  temporal 
considerations. 

The  benthos  of  a  lake  are  collected  by  gathering 
a  known  area  of  bottom  mud  and  straining  it 
through  a  sieve.  Sampling  gear  includes  Ponar,  Peter- 
son, and  Ekman  dredges  (Figures  9  and  10).  They 
all  have  their  most  appropriate  uses — an  Ekman  for 
shallow  or  soft  mud,  Peterson  for  hard  lake  bottom, 
and  Ponar  for  deep  lakes  of  either  type. 


Figure    9.     Ponar  dredge  in  the  open  position. 


Figure    8.     The  closing  style  plankton  net — 


(b) 

detatchable  bucket 

(cs) 

canvas  sleeve 

(h) 

headpiece 

(lr) 

lower  ring 

(m) 

messenger 

(r) 

release 

(s) 

sleeve  of  silk  bolting  cloth  or  other 

suitable  material 

(ur) 

upper  ring 

Figure  10.     Ekman  dredge,  open  and  closed 
positions. 


Lakes 


249 


Samples  are  preserved  in  formalin  and  analyzed 
in  the  laboratory  (identification  and  enumeration). 
Qualified  personnel  are  required  for  the  laboratory 
procedure. 

One  aspect  of  a  lake's  limnology  that  is  fre- 
quently ignored  is  the  emergent  and  submergent 
plants  (macrophytes).  This  is  because  it  is  difficult  to 
collect  information  on  these,  especially  underwater. 
Usually  found  in  the  shallow  waters  of  littoral  zones 
of  lakes,  these  plants  may  be  free-floating  or  at- 
tached. They  serve  as  habitat  for  a  variety  of  other 
aquatic  organisms.  The  contribution  of  macrophytes 
to  the  primary  production  of  lakes,  ponds,  and  even 
reservoirs  has  been  grossly  underestimated  or  ig- 
nored in  many  synoptic  studies.  However,  rooted 
aquatic  plants  are  becoming  increasingly  significant 
in  the  study  of  aquatic  communities,  supplementing 
data  obtained  on  algae  (Patrick  and  Martin  1974). 
Collection  of  these  plants  is  normally  done  by  har- 
vesting. Species  identification,  biomass,  and  regrowth 
measurements  are  done  in  the  laboratory.  A  surface 
plot  is  harvested  and  the  regrowth  rate  is  deter- 
mined as  a  function  of  nutrients  and  productivity.  A 
qualified  aquatic  botanist  with  access  to  laboratory 
and  greenhouse  facilities  should  be  employed  to 
accurately  determine  the  macrophytes  of  a  lake. 

Bacterial  counts  may  be  important  in  some  sur- 
veys. They  give  an  indication  of  the  amount  and  type 
of  pollution  in  the  water.  These  surveys  are  particu- 
larly important  where  water  is  used  for  contact 


sports,  drinking  water,  or  water  in  which  fish  live. 
There  are  a  number  of  techniques  available  for  de- 
termining bacteria  counts  (see  Am.  Public  Health 
Assoc.  1971).  A  qualified  bacteriologist  should  be 
employed  to  perform  the  study  and  analyze  the  data. 

Figure  1 1  lists  the  equipment  or  alternatives 
that  will  be  sufficient  to  perform  biological  limnol- 
ogy studies.  Appendix  II  lists  sources  of  equipment. 

Fishery  Studies.  Fish  are  considered  by  many  to  be 
the  most  important  component  of  a  lake's  food 
chain.  This  may  or  may  not  be  true,  but  under  any 
circumstance,  the  entire  food  chain  needs  to  be 
considered.  Emphasis  is  usually  on  the  fish  because 
of  fishermen's  demands.  There  are  many  ways  to 
approach  a  fishery  survey  and,  therefore,  the 
equipment  needed  also  varies.  As  previously 
mentioned,  the  state  game  and  fish  agency  should  be 
aware  and  involved  in  many  of  these  studies.  In  all 
states,  a  collection  permit  is  required  before  any 
work  begins.  A  creel  census  can  take  many  hours 
because  fishermen  need  to  be  frequently 
interviewed.  On  the  other  hand,  a  netting  survey  can 
require  very  expensive,  sophisticated  gear. 

There  are  basically  two  categories  of  methods 
for  actually  collecting  fish — those  that  trap  and  hold 
the  fish  and  those  that  physically  incapacitate  them. 
Those  two  can  further  be  broken  down  by  the  var- 
ious habitats  in  which  each  individual  method  is 
applicable.  No  one  method  is  the  best  for  all  sizes 


Studies 

Reel  with 

Cable  or 

Rope 

Water 

Sample 

Bottle  and 

Messenger 

Sample 
Bottles 

Preservative 

Laboratory 

Facility 

Includes 

Other 
Equipment 

Primary 
Productivity 

• 

• 

• 

• 

—  NRC  licensed 
personnel  for 
C'4  techniques 

—  BOD  light  and 
dark  bottles 
with  clips 

—  Anchor  with 
float 

Plankton 

• 

• 

• 

• 

—  Binocular 
microscope 

—  Counting  cell 

—  Suitable 
plankton  net 

Chlorophyll 

• 

• 

• 

—  Filtration 

system 
-  Fiberglass 

filters 

—  Cooler 

—  Spectrophoto- 
meter 

Benthos 

• 

• 

• 

—  Binocular 
microscope 

—  Wash  pans 

—  Forceps 

—  No.  30  sieve 
wash  bucket 

—  Ponar,  or 
similar  dredge 

Macrophyte 

—  Greenhouse 

facilities 
—  Aquaria/bell 

jars 

—  Tape  measure 
and  plot  frame 

—  Plastic  bags 

—  Digging  device 

Figure  11.     Equipment  recommended  for  biological  limnology  studies. 


250 


Lakes 


. 


and  species  of  fish  in  all  locations.  Trawls  ( Figure 
12)  are  useful  where  water  depth  and  bottom  com- 
position are  favorable;  trawls  are  most  successfully 
used  at  night.  The  boat  must  tow  the  trammel  net 
efficiently,  so  sufficient  water  depth  is  needed.  The 
only  fish  not  sampled  with  a  trawl  are  the  very  fast 
swimmers  and  very  small  ones. 


A  variety  of  nets  or  seines  are  available  for  tak- 
ing fish.  Nets  are  usually  referred  to  as  fixed  or  pas- 
sive sampling  devices.  Seines  are  usually  indicative  of 
an  action  device,  one  that  is  moved  through  the 


water  and  traps  fish.  Gill  nets  (Figure  13)  are  station- 
ary and  fish  swim  into  them  and  become  captured; 
hoop  or  trap  nets  are  those  that  fish  swim  into  and 
are  held  in  the  enclosed  device  until  removed.  Hori- 
zontal gill  nets  are  most  commonly  used,  but  vertical 
gill  nets  are  also  used  to  determine  depth  of  water 
occupied  by  fish.  The  best  nets  for  reconnaissance 
surveys  are  termed  "experimental  gill  nets"  since 
they  include  a  variety  of  mesh  sizes  sewn  together. 
Gill  nets  are  most  commonly  fished  overnight.  After 
reconnaissance  is  done  and  the  desired  size  of  fish  to 
be  studied  is  determined,  the  size  mesh  to  be  used 
is  selected. 


Figure  12.     The  Beam  Fish  Trawl  (a)  and  the  Otter  Fish  Trawl  (b). 


Figure  13.     Gill  net  (a)  and  hoop  net  (b). 


Lakes 


251 


Seines  commonly  used  are  the  straight  or  min- 
now seine.  Two  individuals  wading  (or  one  wading 
and  one  in  a  boat)  pull  it  near  the  shore.  These  are 
good  for  littoral  zone  sampling.  The  minnow  (or 
haul)  seine  has  a  trailing  bag.  There  are  several 
methods  of  incapacitating  fish.  One  is  the  use  of 
piscicide  chemicals.  The  most  commonly  used  of 
these  is  rotenone.  Creol  and  cyanide  are  also  used. 
All  of  these  require,  in  some  states  at  least,  personnel 
that  are  licensed  piscicide  applicators.  Use  of  these 
piscicides  can  cause  serious  damage  to  fish  and  other 
biota.  Care  in  their  application  should  be  taken. 

Another  way  of  incapacitating  fish  is  the  use  of 
electrofishing  techniques.  This  method  works  well  in 
very  shallow  lakes  or  portions  of  lakes.  Where 
aquatic  vegetation  precludes  efficient  seining,  elec- 
trofishing gear  is  very  applicable.  The  gear  can  be 
either  portable  or  mounted  on  a  boat.  Night  is  the 
best  time  to  electrofish  in  lakes,  ponds,  and 
reservoirs. 

The  last  method  is  the  use  of  hook  and  line. 
This  method,  however,  is  highly  selective  to  pre- 
ferred game  species.  After  fish  are  collected,  species 
identification,  number  of  fish  captured  per  unit  ef- 
fort, length,  weight,  sex,  and  age  should  all  be  re- 
corded. These  data  can  usually  be  collected  without 
harm  to  the  fish.  Some  individuals  are  collected  and 
preserved  for  both  verification  purposes  and  also 
for  further  analysis  (e.g.,  food  habit  studies). 


DISCUSSION 

Each  lake  is  a  world  within  itself.  One  lake  may 
be  similar  to  another  lake,  but  each  is  still  unique. 
That  is,  two  lakes  may  both  have  rainbow  trout, 
freeze  every  winter,  be  too  cold  for  people  to  ever 
swim  in  comfortably,  contain  sparkling  clear  water, 
have  a  stream  flowing  in  and  a  stream  flowing  out, 
and  even  have  pine  trees  up  to  their  shoreline.  A 
closer  look  reveals,  however,  that  one  lake  is  twice 
as  deep  as  the  other  and  has  stronger  thermal  stratifi- 
cation; twice  the  abundance  of  plant  nutrients;  three 
times  the  densities  of  plankton;  a  relatively  large, 
shallow  littoral  area  in  its  upper  end;  and  almost  four 
times  more  fish  caught  in  any  year. 

From  the  above  facts,  a  biologist  can  see  that 
one  lake  probably  needs  a  little  closer  attention  and 


perhaps  a  good  management  plan,  whereas  the  other 
lake  with  richer  flora  and  fauna  should  be  left  alone. 
They  then  are  different  internally.  As  knowledge  is 
gained  from  their  similarities,  knowledge  also  is 
gained  from  their  differences. 

As  alternative  management  decisions  regarding  a 
resource  are  being  considered,  it  is  advisable  that  as 
many  concerned  resource  specialists  and  managers 
as  possible  be  consulted  for  their  opinions.  This  is,  of 
course,  a  difficult  task.  I  know  from  experience  that 
as  the  number  of  interests  (agencies  and  people) 
in  natural  resources  increases,  the  complexity  of  any 
decision  increases  and  the  reality  of  achieving  a  wise 
decision  seems  more  remote.  However,  biologists 
and  resource  managers  should  not  shy  away  from 
seeking  all  opinions  possible,  no  matter  how  diverse 
they  may  become,  so  that  decisionmakers  can  make 
decisions  based  on  the  best  information. 

The  questions  are  always  how  much  data  is 
needed  and  how  many  times  do  things  need  to  be 
measured?  These  questions  should  be  asked  and 
answers  sought  from  resource  personnel  from  all 
interested  resource  agencies.  The  decision  that  is  fi- 
nally made  may  strongly  influence  the  future  of  a 
particular  resource.  For  example,  some  decisions 
may  result  in  irreversible  changes  to  a  system  (e.g., 
introduction  of  a  new  species  of  game  fish ),  whereas 
others  may  be  reversed  (e.g.,  change  in  the  annual 
flow  regime  from  a  reservoir).  Having  the  most  and 
best  information  possible  adds  to  the  credibility  of 
these  important  decisions. 

The  final  important  step  that  is  neglected  most 
of  the  time  is  to  follow  up  on  the  actual  influence 
and  success  of  the  decision.  A  good  rule  for  the  re- 
source's sake  is  to  see  that  follow-up  monitoring 
investigations  are  performed.  For  example,  suppose 
it  had  been  decided,  based  on  scientific  evidence, 
that  a  forage  species  of  fish  was  desirable  in  a  certain 
lake  because  walleye  did  not  have  enough  to  eat 
after  they  became  15  cm  (6  in.)  long.  It  would  be 
desirable  to  monitor  whether  the  introduction  of 
forage  fish  was  successful  or  not,  also  to  know  if 
forage  fish  are  reproducing  naturally  in  sufficient 
numbers  to  provide  an  adequate  food  base  for  wall- 
eye in  the  future.  Certainly  monitoring  this  situation 
would  make  the  next  similar  decision  easier  to  make. 
Post-decision  monitoring  then  makes  each  future 
decision  easier  to  make. 


252 


Lakes 


LITERATURE  CITED 


AMERICAN  PUBLIC  HEALTH  ASSOCIATION  and  AMER- 
ICAN WATER  WORKS  ASSOCIATION.  1971.  Stand- 
ard methods  for  the  examination  of  water  and 
wastewater.  Water  Control  Bull.  1971.  13th  ed.  Am. 
Pub.  Health  Assoc.  Washington,  DC.  874pp. 

HUTCHINSON,  G.E.  1957.  A  treatise  on  limnology,  Vol- 
ume I,  geography,  physics  and  chemistry.  John  Wiley 
&  Sons,  Inc.  New  York,  NY.  1015pp. 

MAITLAND,  PS.  1978.  Biology  of  freshwaters.  John  Wiley 
&  Sons,  New  York,  NY.  244pp. 

ODUM,  E.P  1971.  Fundamentals  of  ecology.  W.B.  Saunders 
Co.  Philadelphia,  PA.  574pp. 


PATRICK,  R.  and  D.M.  MARTIN.  1974.  Biological  surveys 
and  biological  monitoring  in  freshwaters.  Number  5. 
Contributions  from  Dep.  Limnology,  Acad.  Nat.  Sci. 
Philadelphia,  PA.  63pp. 

TRANDAHL,  A.  1978.  Preface,  Page  x  in  Kendall,  R.L.  ed., 
Cool-water  Fishes  of  North  America.  Special  Publica- 
tion 11,  Am.  Fish.  Soc.  Washington,  DC.  437pp. 

U.S.  GEOLOGICAL  SURVEY,  Office  of  Water-Data  Coordi- 
nation. 1977.  National  handbook  of  recommended 
methods  for  water-data  acquisition.  Two  Volumes. 
U.S.  Dep.  Inter.  Reston,  VA. 

WOOD,  R.D.  1975.  Hydrobotanical  methods.  Univ.  Park 
Press.  Baltimore,  MD.  173pp. 

WRIGHT,  J.C.  1967.  Effects  of  impoundments  on  produc- 
tivity, water  chemistry,  and  heat  budgets  of  rivers. 
Pages  188-199  in  Reservoir  Fishery  Resources  Sym- 
posium. Am.  Fish.  Soc.  Washington,  DC. 


Lakes 


253 


APPENDIX  I. 


APPENDIX  II. 


Manuals  and  Textbooks  with  Inventory 
and  Monitoring  Methods  for  Lakes. 


Sources  of  Limnological  and  Fishery 
Equipment. 


CAIRNS,  J.,  Jr.  and  K.L.  DICKSON.  1973.  Biological  meth- 
ods for  the  assessment  of  water  quality.  ASTM  Special 
Publication  528.  Am.  Soc.  for  Test  &  Mat.  Philadel- 
phia, PA.  256pp. 

EDDY,  S.  and  J.  UNDERHILL.  1978.  The  freshwater  fishes. 
Wm.  C  Brown  Co.  Dubuque,  IA.  215pp. 

EVERHARDT,  W.H.,  AW.  EIPPER,  and  WD.  YOUNG. 
1975.  Principles  of  fishery  science.  Cornell  Univ. 
Press.  Ithaca,  NY.  288pp. 

HUTCHINSON,  G.E.  1967.  A  treatise  on  limnology,  Vol- 
ume II.  Introduction  to  lake  biology  and  the  limno- 
plankton.  John  Wiley  &  Sons,  Inc.  New  York,  NY. 
1115pp. 

LAGER,  K.F.  1956.  Freshwater  fishery  biology.  Wm.  C 
Brown  Co.  Dubuque,  IA.  421pp. 

LIETH,  H.  and  R.H.  WHITTAKER,  eds.  1975.  Primary  pro- 
ductivity of  the  biosphere.  Ecological  Studies  14, 
Springer-Verlag,  New  York,  NY.  339pp. 

LIND,  O.T.  1979.  Handbook  of  common  methods  in  lim- 
nology, Second  Edition.  The  C.V.  Mosby  Co.  St.  Louis, 
MO.  154pp. 

WELCH,  P.S.  1948.  Limnological  methods.  McGraw  Hill, 
Inc.  New  York,  NY.  381pp. 


This  list  implies  no  endorsement  of  product  or 
service.  A  more  complete  listing  is  "Sources  of  Lim- 
nological and  Oceanographic  Apparatus  and  Sup- 
plies" (special  publication  No.  1,  third  revision) 
available  from  the  American  Society  of  Limnology 
and  Oceanography,  1530  12th  Avenue,  Grafton, 
Wisconsin  53024. 

Test  kits 

Hach  Chemical  Co. 
P.O.  Box  389 
Loveland,  CO  80539 

General  Sampling  &  Field  Equipment 
Wildlife  Supply  Company 

(Wildco) 
2200  S.  Hamilton  St. 
Saginaw,  MI  48602 

Kalh  Scientific 

Instrument  Co. 
P.O.  Box  1166 
El  Cajon,  CA  92022 

Fish  Nets  and  Fish  Sampling  Gear 
Nylon  Net  Company 
7  Vance  Avenue 
P.O.  Box  592 
Memphis,  TN  38101 

Cofelt  Electronic  Co., 

Inc. 
3910  So.  Windermere  St. 
Englewood,  CO  80110 

Radio  Isotopes 

Amersham-Searle  Corp. 
2636  Clearbrook  Dr. 
Arlington  Heights,  IL 
60005 

New  England  Nuclear 
575  Albany  St. 
Boston,  MA  02118 

Multiparameter  Instruments  Submarine  Photometers 
Hydrolab 
P.O.  Box  50116 
Austin,  TX  78763 

Kalh  Scientific 

Instrument  Co. 
P.O.  Box  1 166 
El  Cajon  CA  92022 


254 


Lakes 


Ill  SPECIES  GROUPS 


13  Fish 


14  Amphibians  and  Reptiles 

i 

15  Songbirds 

16  Raptors 

17  Marsh  and  Shorebirds 

18  Waterfowl 

19  Colonial  Waterbirds 

20  Upland  Game  Birds 

21  Rodents  and  Insectivores 

22  Lagomorphs 

23  Carnivores 

24  Bats 

25  Ungulates 


'.  ■ 


13 
FISH 


Paul  Cuplin 

U.S.  Bureau  of  Land  Management 
Service  Center 
Denver,  CO  80225 


Editor's  Note:  This  chapter  is  the  first  of  a  series  in 
which  the  focus  is  on  species  groups.  In  this  book, 
species  are  lumped  into  groups  based  upon  similar- 
ity of  the  techniques  used  to  inventory  and  monitor 
their  habitat  or  populations.  Fish  are  of  course  a 
diverse  and  widespread  group,  and  fisheries  man- 
agement is  often  considered  as  a  separate  discipline 
from  wildlife  management.  However,  inventory 
and  monitoring  of  an  area  of  land  would  be  in- 
complete without  taking  into  consideration  the 
aquatic  resources,  of  which  fish  are  extremely  im- 
portant. One  chapter  cannot  cover,  except  in  the 
briefest  manner,  the  massive  field  of  fisheries  biol- 
ogy. However,  it  can  serve  as  a  starting  point  for 
the  biologist  with  little  experience  in  this  area  The 
references  cited  provide  further  guidance. 


INTRODUCTION 

The  purpose  of  this  chapter  is  to  assist  biologists 
in  identifying  information  needed  to  inventory  and 
monitor  fish-producing  waters  including  streams, 
lakes,  ponds,  and  reservoirs.  Fishery  biology  is  a 
well-developed  profession  with  academic  curricula 
and  degrees,  professional  societies,  and  scientific 
publications.  A  single  chapter  on  fish  cannot  cover 
the  entire  subject  of  fish  habitat  and  population  tech- 
niques in  any  depth.  This  chapter  emphasizes  the 
techniques  appropriate  for  inland  fisheries  investiga- 
tions in  the  western  U.S.,  particularly  for  salmonids 
and  for  smaller  streams  and  ponds  in  arid  lands.  For 
more  detailed  information,  consult  Nielsen  and  John- 
son (1983)  or  the  papers  cited  in  this  chapter.  Carl- 
son and  Gifford  (1983)  provide  guidance  on  locating 
and  accessing  fisheries  literature. 

State  fish  and  game  agencies  have  historically 
managed  fish  populations;  federal  land  agencies  man- 
age habitat  on  public  land.  There  must  of  necessity 
be  some  overlapping  of  responsibility  since  federal 
agencies  cannot  evaluate  the  habitats  properly  with- 
out knowing  something  about  existing  fish  popula- 
tions. Likewise,  State  agencies  need  to  know  habitat 
variables  in  order  to  manage  fish  populations.  Nor 
can  critical  habitat  for  threatened  or  endangered  fish 
species  be  preserved  without  knowledge  of  habitat 
variables  needed  to  sustain  listed  fish  species. 


AQUATIC  HABITAT  FEATURES 

Important  habitat  features  for  fish  are  described 
in  this  section.  Methods  for  measuring  many  of  the 
following  attributes  or  components  are  described  by 
Hamilton  and  Bergersen  (1984). 

Soils 

The  ability  of  the  resource  specialist  to  deter- 
mine potential  of  riparian  zones  through  the  analysis 


Fish 


257 


of  soil  and  water  flows  will  direct  rehabilitation  ef- 
forts to  riparian  areas  with  the  highest  potential. 
Streams  in  improved  riparian  habitat  will  support 
larger  fish  populations. 

Crouse  and  Kindschy  (1981)  described  soil 
conditions  required  by  various  riparian  plant  species 
along  streams  in  southeastern  Oregon.  In  the  west- 
ern U.S.,  highly  alkaline  soils  along  perennial  streams 
will  predictably  produce  alkali  bullrush  (Scirpus 
paludosus),  greasewood  (Sarcobatus  sp.),  buffalo- 
berry  (Shepherdia  sp.),  and  salt  cedar  (Tamarix  sp.); 
less  alkaline  soils  will  produce  sedges  (Carex  sp.), 
forbs,  grasses,  cattails  (Typha  sp.),  and  few  woody 
species.  Soil  that  is  extremely  rocky  will  support 
willow  (Salix  sp. ),  mock  orange  (Philadelphus  sp. ), 
chokecherry  (Prunus  sp.),  and  sparse  stands  of 
grasses  and  forbs.  Fine  textured  soil  will  support  tree 
willow,  cottonwood  (Populus  sp.),  alder  (Alnus 
sp.),  aspen,  dogwood  {Cornus  sp.),  mock  orange  and 
other  shrubs,  dense  stands  of  grasses,  sedges,  and 
forbs. 

Physical  Features 

Physical  features  of  streams  and  lakes  such  as 
stream  gradient,  may  be  correlated  with  the  capabil- 
ity of  the  water  to  support  or  produce  fish.  These 
are  described  in  more  detail  in  the  Lakes  and 
Streams  chapters.  Methods  for  measuring  such  attri- 
butes are  described  in  the  Aquatic  Physical  Features 
chapter. 

Vegetation 

Along  streams,  the  capability  of  the  stream  to 
produce  or  support  fish  populations  is  often  a  func- 
tion of  or  correlated  with  the  condition  of  the  ripar- 
ian vegetation.  Shrubs,  trees,  and  other  woody 
riparian  vegetation  are  needed  to  shade  smaller 
streams  and  control  high  summer  water  tempera- 
tures for  cold-water  fishes  such  as  the  Salmonidae, 
which  include  trout,  salmon,  and  grayling.  Stream- 
banks  are  strengthened  and  stabilized  by  the  root 
structure  of  trees  and  shrubs.  Streams  with  only 
grass  and  forb  cover  are  subject  to  streambank  ero- 
sion during  high  water  flows. 

Food  of  salmonids  sometimes  consists  of  as 
much  as  35%  terrestrial  insects  during  the  summer 
months.  If  woody  riparian  vegetation  is  absent,  this 
source  of  food  is  also  absent. 


Water  Quality 

The  chemical  and  physical  properties  of  water 
are  extremely  important  habitat  characteristics  for 
fish.  Three  of  the  most  important  attributes  are  pH, 
dissolved  materials,  and  dissolved  oxygen.  Other 
properties  are  described  in  the  chapters  on  Streams 
and  Lakes.  Methods  for  measuring  water  quality  are 
described  in  the  chapter  on  that  subject. 

pH.  Most  freshwater  fishes  can  accommodate  pH 
ranging  from  6.0  to  95,  although  some  species  can 
adapt  below  6.0  or  above  9.5  pH. 

Dissolved  Materials.  Total  dissolved  materials  in 
excess  of  1,500  mg/L  NaCl  cannot  be  tolerated  by 
most  fresh-water  fish  species.  Most  waters  in  the 
West  fall  in  a  range  of  less  than  400  mg/L  except  for 
the  lower  Colorado  River,  saline  seeps,  and  springs. 

Dissolved  Oxygen.  For  warm-water  fish  the 
dissolved  oxygen  (DO)  concentration  should  be  5 
mg/L  or  greater.  Cold-water  fishes  prefer  the  DO 
concentration  to  be  at  saturation  which  varies  with 
water  temperature  and  elevation.  Under  extreme 
conditions  the  DO  may  range  to  5  mg/L  or  lower 
during  the  winter.  The  critical  periods  for  DO  are 
during  late  summer  and  during  long  periods  of  ice 
and  snow  cover  during  winter  months. 


Macroinvertebrates 

Macroinvertebrates  can  be  used  to  characterize 
stream  water  quality  by  the  diversity  of  species, 
number  of  species,  and  by  the  presence  or  absence 
of  certain  species  ( see  the  Macroinvertebrates 
chapter). 

The  simplified  stream-aquatic  organism  relation- 
ship is  presented  in  Figure  1.  Macroinvertebrates 
play  a  major  role  in  consuming  aquatic  plants  and 
detritus  and  providing  fish  with  most  of  their  food 
supply.  A  wide  diversity  of  macroinvertebrates  is 
produced  in  the  stream  substrate  under  optimum 
conditions. 


Food  Supply 

The  flow  of  nutrients  into  the  stream  system 
governs  the  number  and  species  of  macroinverte- 
brates that  are  available  to  fish  for  food. 


Shrubs  and  trees  also  provide  much  of  the  en- 
ergy source  for  the  stream  from  leaf  fall  and  twigs. 
Tree  fall  across  the  stream  provides  resting  shelter 
and  stream  habitat  diversity.  Healthy  riparian  vegeta- 
tion acts  as  a  filter  to  prevent  soil  from  reaching 
the  stream  during  heavy  rainstorms  or  snowmelt 
(Odum  1978). 


Fish  feeding  habits  vary  widely  by  family,  spe- 
cies, young  of  the  year,  juvenile,  and  adult.  Examples 
of  planktivores,  plankton  feeding  species;  benthic  or 
bottom  feeders;  and  piscivores  or  carnivorous  fish 
are  shown  in  Table  1.  Not  listed  are  sockeye  or  ko- 
kanee  (non-migratory)  salmon  that  are  planktivores 
throughout  all  life  stages. 


258 


Fish 


^A  Sun  (Energy) 

Algae,  aquatic  plants 

Riparian  vegetation 
(leaves,  twigs,  wood,  detritis) 

\ 

Herbivorous  aquatic  invertebrates 
(mayfly,  caddis,  diptera  larvae,  snails) 

\ 

Fi! 

-h       ri 

Carnivorous  aquatic  insects 
(dragonflies,  caddis,  some  stoneflies) 

an  "* 

Figure  1.     Simplified  stream-aquatic  organism  relationships. 

Table  1.     Examples  of  families  and  species  of  planktivores,  benthic  feeding  fish,  and  piscivores  (from  Fritz  et 
al.  1980). 


Group 

Family 

Species 

Planktivores 

Clupeids 
Corregonids 

Alewife  (Alosa  pseudoharengus) 
Gizzard  shad  (Dorosoma  cepedianum) 

Lake  whitefish  (Coregonus  clupeaformis) 

Ciscoe  (Coregonus  sp.) 

Round  whitefish  (Prosopium  cylindraceum) 

Benthic  feeding  fish 

Catostomids 

Cyprinids 
Percopsids 
Cottids 
Ictalurids 

Percids 

White  sucker  (Catostomus  commersomi) 
Longnose  sucker  (C.  catostomus) 
Carp  (Cyprinus  carpio) 

Spottail  shiner  (Notropis  hudsonius) 

Troutperch  (Percopsis  omiscomaycus) 

Sculpins  (Cottidae) 

Catfish  (Ictalurus  sp.) 
Bullhead  (Ictalurus  sp.) 

Walleye  (Stizostedion  vitreum) 
Yellow  perch  (Perca  flavescens) 

Piscivores 

Salmonids 

Osmerids 
Esocids 

Lake  trout  (Salvelinus  namaycush) 

Brown  trout  (Salmo  trutta) 

Rainbow  trout  (Salmo  gairdneri) 

Chinook  salmon  (Oncorhynchus  tshawytscha) 

Coho  salmon  (0.  kisutch) 

Rainbow  smelt  (Osmerus  mordax) 

Northern  pike  (Esox  lucius) 

Fish 


259 


The  food  supply  may  be  an  important  habitat 
component  limiting  fish  populations  or  productivity. 
This  may  be  important  primarily  during  one  season 
of  the  year  or  during  one  life  stage.  Bowen  (1983) 
describes  methods  for  determining  fish  diets. 


HABITAT  OF  IMPORTANT  FISH  FAMILIES 

Families  of  fishes  have  developed  under  various 
ecological  conditions.  The  habitat  requirements  of 
each  family  of  fish  are  well  known,  and  habitat  stud- 
ies continue  to  add  to  our  knowledge. 

The  families  of  fishes  inhabiting  public  land 
waters  in  the  western  States  and  Alaska  can  be 
grouped  into  cold-water,  cool-water,  and  warm-water 
fishes. 

Cold-Water  Families 

Salmonidae.  Family  Salmonidae,  for  example,  is 
represented  by  trout,  salmon,  whitefish,  and  grayling. 
This  family  represents  important  game  and 
commercial  species.  Their  life  requirements  are 
similar  in  that  they  inhabit  cold  water,  and  most 
spawn  in  or  on  stream  substrate  in  the  fall  or  spring. 
Anadromous  members  (migrating  from  the  ocean 
up  a  fresh  water  stream  to  spawn ),  such  as  the 
Pacific  salmon  (Oncorhynchns  sp.),  migrate  to  the 
ocean  in  the  first  or  second  year  of  life  and  return  to 
fresh  water  to  spawn  and  die  after  spending  1  to  3 
years  in  saltwater.  Unlike  the  Pacific  salmon,  the  sea- 
run  rainbow  trout  (steelhead),  sea-run  cutthroat 
trout  (S.  clarki),  and  Dolly  Varden  trout  (Salvelinus 
malma)  return  from  the  ocean  to  spawn  in  fresh 
water  but  do  not  die  after  spawning.  They  can  repeat 
the  ocean — fresh  water  spawning  migration  cycle. 
Whitefish  and  grayling  (Thymallus  arcticus)  spend 
their  entire  life  as  residents  of  a  lake  or  stream. 


Cottidae.  The  family  Cottidae,  sculpins  or  cottids, 
lives  on  the  stream  bottom  of  cold-water  streams. 
Sculpins  or  cottids  inhabit  much  of  the  same  stream 
habitat  occupied  by  trout  and  provide  food  for  them. 

Cool-Water  Families 

Catostomidae.  The  family  Catostomidae,  suckers, 
uses  some  of  the  same  habitat  as  trout.  Suckers 
migrate  upstream  to  spawn  during  the  spring  months 
and  compete  with  trout  for  spawning  space.  They 
are  common  in  degraded  trout  streams  which  have 
silted  streambeds  and  higher  than  normal  summer 
water  temperatures. 

Escoidae.  The  family  Escoidae,  pike  and  pickerels, 
inhabits  streams,  lakes,  and  reservoirs  and  is  a  native 
east  of  the  Rocky  Mountains  to  Alaska.  Introductions 
have  been  made  in  western  waters.  The  best  known 
representative  is  the  northern  pike  (Esox  lucius). 
Preferred  habitats  are  weedy  areas  of  lakes  and 
reservoirs.  Spawning  takes  place  in  the  spring.  The 
eggs  adhere  to  the  stream  substrate. 

Percidae.  Family  Percidae,  yellow  perch  (Perca 
flavescens)  and  walleye  (Stizostedion  vitreum), 
originally  occurred  in  lakes  and  reservoirs  east  of  the 
Continental  Divide.  They  have  been  widely 
introduced  into  the  western  States.  They  prefer  the 
cool  water  in  lakes  and  reservoirs. 

Warm-Water  Families 

Centrarchidae.  The  family  Centrarchidae,  sunfishes, 
is  best  known  for  bluegill  (Lepomis  macrochirus), 
largemouth  bass  (Micropterus  salmoides),  and 
crappie  (Pomoxis  sp. ).  Most  of  these  species  prefer 
warm-water  lakes  and  reservoirs.  Smallmouth  bass 
(M.  dolomieui),  however,  inhabit  the  larger,  warmer 
streams.  Only  one  member  of  this  family,  the 


Grayling. 


Northern  pike  (top)  and  sauger  (bottom). 


260 


Fish 


Sacramento  perch  (Archoplites  interruptus),  is 
native  west  of  the  Continental  Divide,  primarily 
found  in  the  Sacramento  and  San  Joaquin  River 
Basins  in  California  (Simpson  and  Wallace  1978). 

Cyprinidae.  The  family  Cyprinidae,  minnows  and 
carps,  is  best  represented  by  the  common  carp 
(Cyprinus  carpio),  squawfish  (Ptychocheilns  sp. ), 
and  the  goldfish  (Carassius  auratus).  Carp, 
introduced  from  Asia,  thrive  in  warm,  shallow 
streams  and  lakes  with  abundant  aquatic  vegetation. 
The  rooting  habits  of  carp  contribute  to  the  high 
turbidity  of  water. 

Squawfish  are  found  in  a  variety  of  habitats  from 
reservoirs  to  the  largest  western  rivers.  Squawfish 
once  achieved  a  very  large  size — as  much  as  27  kg 
(60  pounds)  in  the  Colorado  River.  The  numbers  of 
Colorado  squawfish  have  been  severely  reduced 
due  to  dam  construction  which  altered  their  re- 
quired stream  habitat  and  water  temperatures  for 
spawning. 

Cyprinodontidae.  The  family  Cyprinodontidae, 
killifishes,  inhabits  ponds,  springs,  and  streams  of  the 
southwestern  deserts.  The  most  well  known  of  this 
group  is  the  Devil's  Hole  Pupfish  (Cyprinodon 
diabolis)  which  exists  in  the  most  restricted  fish 
habitat  on  earth,  a  spring-fed  pond  in  Nevada  with  a 
surface  area  not  much  larger  than  a  bathtub;  the 
total  population  is  300  to  350  individuals.  The 
habitat  for  this  and  other  similar  fishes  of  the  desert 
is  in  jeopardy  due  to  the  use  of  water  for  crop 
irrigation. 


LolonraxLo    oorM-euju~C*^Ks 


~£<n~2-/W   LlvuJr- 


hUxAM^i^AcA^-     iMjuJr- 


Ictaluridac.  The  family  Ictaluridae,  catfishes, 
originally  found  east  of  the  Continental  Divide,  can 
now  be  found  in  lakes,  reservoirs,  and  streams 
throughout  the  West.  The  family  is  represented  by 
black  bullhead  (Ictalurus  melas),  brown  bullhead  (/. 
nebulosus),  channel  catfish  (/.  punctatus),  flathead 
catfish  (Pylodictis  al warts),  and  blue  catfish  (/. 
furcatus).  The  channel,  flathead,  and  blue  catfish  are 
most  commonly  found  in  the  large  warmer  ponds 
and  reservoirs  and  large  warm-water  streams. 


Endangered  Fish  Species 

Twenty-seven  of  the  44  U.S.  fish  species  listed 
under  the  Endangered  Species  Act  of  1973  are  en- 
demic to  the  southwestern  U.S.  (Johnson  and  Rinne 
1982),  an  indication  of  the  rapidly  disappearing 
aquatic  habitat  of  the  southwestern  deserts. 

Dewatering  of  desert  springs  through  ground- 
water pumping  and  continued  drying  of  desert 
aquatic  habitats  has  seriously  reduced  habitats  for 
desert  fishes.  Water  impoundments  on  the  Colorado 
River  have  reduced  water  temperatures  and  there- 
fore the  reproduction  capabilities  of  the  four  native 
"large  river"  species,  Colorado  Squawfish,  razorback 
sucker  (Xyrauchen  texanus),  humpback  chub  {Gila 
cypha),  and  bonytail  chub  (Gila  elegans). 


POPULATION  MEASUREMENT  TECHNIQUES 

Fish  population  studies  are  conducted  to  relate 
fish  biomass  to  habitat  conditions,  observe  change  in 
relative  fish  numbers,  determine  habitat  preference 
by  species,  and  determine  use  periods  of  the  year  at 
various  life  stages.  A  method  of  sampling  should  be 
selected  to  provide  the  required  information  at  a 
minimum  cost. 

Johnson  and  Nielsen  (1983)  compared  relative 
costs  to  obtain  different  types  of  information  on 
fish  populations  and  communities  (Table  2).  The 
lowest  level  (easiest  to  get  but  least  valuable)  of  in- 
formation about  fish  communities  is  the  number  of 
species  present  and  susceptible  to  sampling  gear.  At 
the  other  end  of  the  scale  are  radio  tracking  and 
food  habit  information;  both  activities  are  very  ex- 
pensive, but  the  resultant  information  may  be  exten- 
sive. Between  these  extremes  are  a  series  of 
activities  that  will  provide  more  information  at  in- 
creasing cost.  Table  2  lists  these  activities  and  an 
estimate  of  their  relative  cost.  Each  activity  is  com- 
pared to  the  first  activity,  and  each  activity  is  af- 
fected by  sampling  bias.  Since  the  relative  cost  in 
dollars  and  personnel  of  different  types  of  studies 
vary  considerably,  the  biologist  should  be  careful  not 
to  collect  more  information  than  is  required  for  the 
problem  being  addressed. 


Fish 


261 


Table  2.     A  cost-benefit  hierarchy  for  information  about  fish  populations  and  communities  (from  Johnson  and 
Nielsen  1983). 


Level 

Activity 

Information 

Relative  Cost 

Comments 

1 

Species 
enumeration 

Number  of  species 
present. 

1 

Useful  in  sampling. 

2 

Numbers  of  fish 
caught  of  each 
species 

Relative  abundance  of  the 
species  present. 

x2 

Usually  the  minimal  level 
of  information  needed. 

3 

Length  of  fish 

Length  distributions  can 
give  relative  year-class 
strength,  growth,  and 
mortality,  especially  for 
young  fish. 

x4 

A  great  deal  of  helpful 
information  added,  partic- 
ularly for  fast  growing, 
temperate  latitude. 

4 

Weight  of  fish 

Length-weight  curves, 
condition  factors,  relative 
weight. 

x12 

Most  field  scales  are  not 
accurate  enough.  Must 
construct  special  shelter 
or  move  indoors.  Condi- 
tion factors  need  both 
accurate  lengths  and 
weights. 

5 

Age  determination 

More  accurate  than  length 
distributions  for  calculating 
year-class  strength,  age 
distribution,  growth  history, 
and  mortality. 

x120 

Must  have  accurate  length 
measurements;  requires 
extra  handling  of  fish  and 
laboratory  time  for  analy- 
sis. 

6 

Radio  or  sonar 
tagging 

Exact  information  about 
fish  location  and  mea- 
surements. May  be  com- 
bined with  information 
about  water  depth,  tem- 
perature. 

x1,200 

Much  information  gained 
about  movements  of  rela- 
tively few  fish.  Equipment 
cost  and  maintenance  can 
be  quite  high. 

Inventorying  a  small  stream  by  electrofishing. 


262 


Fish 


General  Considerations 

Life  Stages  of  Fishes.  The  terms  applied  to  the 
developmental  life  stages  of  fish  assist  in  describing 
and  understanding  fishery  reports  and  studies. 

Developmental  stage  terms  most  commonly 
used  are  as  follows: 

Egg — Fertilized  ovum  with  developing  embryo. 

Eyed  Egg — Embryo  developed  to  the  state 

where  the  eyes  can  be  readily  ob- 
served. 

Larva — Prolarva  still  bear  yolk  sacs,  sometimes 
called  sac  fry.  With  postlarva,  yolk  sacs 
have  been  completely  absorbed. 

Alevin — The  newly  hatched  salmon,  with  yolk 
sac  still  attached. 

Fry — The  same  as  postlarva;  yolk  sac  has  been 
completely  absorbed. 

Fingerling — A  fish  in  the  first  or  second  year  of 
life,  5  to  10  cm  (2  to  4  in.)  long. 

Young  of  the  Year — Young  fish  added  to  the  fish 
population  during  the  cur- 
rent year. 

Smolt — One-  or  2-year-old  anadromous  fish 

physiologically  prepared  for  migration 
into  saltwater. 

Grilse — Early  maturing  salmon,  mostly  male  fish. 

Adult — Mature  fish  capable  of  spawning. 

Kelt — Spent  salmon  or  sea-run  trout  after  spawn- 
ing, in  a  weak,  emaciated  condition. 

In  the  past,  most  fish  population  work  has  con- 
centrated on  mature  stages  of  fish  (from  fry  through 
kelt).  Awareness  is  growing  of  the  need  to  collect, 
identify,  and  gather  data  on  larval  fish.  In  many  cases, 
habitat  for  larval  fish  is  quite  different  from  that  of 
mature  stages;  it  may  be  geographically  separated 
also.  If  larval  habitat  is  deteriorating,  then  it  should 
be  identified  and  monitored. 

Laboratories  are  available  for  assisting  in  larval 
fish  sampling  and  identification.  Since  1977  a  larval 
fish  conference  has  been  held  annually  in  the  U.S.  or 
Canada.  Information  on  larval  fish  procedures  are 
described  by  Snyder  (1983);  more  information  on 
the  conferences  and  on  procedures  can  be  obtained 
from  the  Larval  Fish  Laboratory,  Colorado  State  Uni- 
versity, Fort  Collins,  CO  80523. 


Fish  Collection.  Fish  specimens  sometimes  need  to 
be  collected  and  stored  for  later  identification  or 
tagged  or  marked  for  population  surveys.  Biologists 
must  comply  with  State  and  federal  laws  including 
collection  permit  requirements  when  sampling  or 
collecting  fish.  Procedures  for  tagging  and  marking 
fish  are  described  by  Wydoski  and  Emery  (1983). 

Fish  Preservation.  Specimens  should  be  preserved 
in  4%  formaldehyde  solution  (10%  formalin)  con- 
taining about  3  g  of  borax  per  liter  (Greeson  et  al. 
1977).  Fish  more  than  3  inches  long  should  be  slit  on 
the  right  side  to  allow  preservatives  to  penetrate 
the  body  cavity. 

Identify  samples  with  date,  species,  location 
where  collection  was  made,  and  name  of  collector.  If 
individual  fish  from  separate  locations  are  to  be 
stored  in  the  same  container,  they  can  be  wrapped 
in  cheesecloth  with  an  identification  tag  attached. 

Capture  Methods.  Fisheries  surveys  beyond  Level  1 
(presence;  Table  2)  normally  require  fish  capture. 
Fish  capture  can  also  be  used  to  determine  presence. 
Many  methods  for  capturing  fish  are  available. 
Seining,  electrofishing,  fish  toxicants,  and  hook-and- 
line  sampling  are  commonly  used  on  small  inland 
waters. 


Seining  Small  seines  are  useful  in  small  ponds 
but  are  less  effective  in  small  streams.  Details  about 
seining  techniques  are  described  by  Hayes 
(1983.139-143). 

Electrofishing.  Electrofishing  can  be  used  to 
determine  fish  presence,  species  composition,  and  to 
make  fish  population  estimates.  Electrofishing  is  the 
most  effective  method  of  fish  population  sampling 
for  a  stream.  The  most  effective  time  for  electrofish- 
ing is  from  midsummer  to  the  fall  months  when 
the  streamflow  is  at  a  low  point. 

A  600-m  (1/10-mi. )  sampling  area  is  blocked 
with  block  nets  at  the  upper  and  lower  ends  of  the 
sampling  area.  Population  estimates  for  important 
game,  sensitive,  and  endangered  or  threatened  fishes 
are  made  by  the  Seber-LeCren  (1967)  method.  A 
fish  population  estimate,  relative  abundance  estimate, 
fish  length  distribution,  and  pounds  of  fish  can  be 
calculated  for  each  fish  population  sample.  Reynolds 
( 1983)  describes  in  detail  procedures  for 
electrofishing. 

Other.  Many  other  techniques  for  capturing 
fish  are  available  including  use  of  toxicants  and 
hook-and-line.  Hubert  (1983)  and  Hayes  (1983) 
describe  procedures  for  passive  and  active  capture 
techniques,  and  Davies  (1983)  describes  use  of  toxi- 
cants. 


Fish 


263 


Specific  Techniques 

Appropriate  techniques  that  correspond  with 
each  of  the  6  levels  of  survey  (Table  2)  are  de- 
scribed below. 


Presence  (Level  1). 


Counting  Weirs  and  Spawning  Ground 
Surveys,  Aerial  Redd  Counts.  Observations  of 
salmon  and  other  anadromous  fish  species  in 
spawning  runs  have  often  been  made  at  counting 
wiers  and  by  the  enumeration  of  dead  fish  on  the 
spawning  grounds.  The  latter  method  is  used  to 
determine  numbers  and  species  on  specific  streams. 
Redd  counts  can  be  made  during  the  spawning 
season  with  helicopter  or  fixed-wing  aircraft 
(Neilson  and  Green  1981). 

Visual  Observation  of  Stream  Fish.  Stream 
surveyors  can  recognize  species  by  observing,  deter- 
mining presence,  and  estimating  relative  numbers 
per  measured  distance  of  a  stream. 

Relative  Abundance — Population  Estimation 
(Level  2).  Relative  abundance  can  be  expressed  as 
the  numbers  of  fish,  by  species,  per  unit  of  capture 
effort,  e.g.,  number  of  brook  trout  (S.  fontinalis)  per 
miles  of  stream  shocked. 

For  important  species,  population  estimates  can 
be  made  using  one  of  the  techniques  described  be- 
low. Express  population  estimates  in  numbers  per 
linear  mile  for  streams  or  numbers  per  surface  acre 
for  lakes,  ponds,  and  reservoirs.  For  details  of  statisti- 
cal analyses  or  information  on  alternative  methods, 
see  Lagler  (1956),  Ricker  (1958),  Armour  and  Platts 
(1983),  Otis  et  al.  (1978),  or  White  et  al.  (1982). 

Two-Catch  Depletion  (Seber-LeCren)  Esti- 
mator. Depletion  methods  require  that  the  fish  from 
samples  not  be  re-placed  (put  back  into  the  water). 
In  the  two-catch  depletion  method,  a  first  sample 
is  taken  (without  re-placement)  followed  by  a  sec- 
ond sample.  Equipment  and  effort  must  be  standard- 
ized so  that  the  expected  proportion  of  the 
population  caught  is  the  same  for  each  sample. 

The  total  population  can  then  be  estimated  with 
the  formula  (Seber  and  LeCren  1967): 

C,  -  C2 


Otis  et  al.  ( 1978)  provide  a  formula  for  calculat- 
ing sampling  variance.  However,  White  et  al.  (1982) 
recommended  using  three  or  more  captures  to  test 
the  assumption  of  equal  capture  probability  and  to 
increase  precision.  They  describe  assumptions  and 
statistical  analyses  for  three  or  more  samples. 

Mark  and  Release.  Mark  and  release  tech- 
niques have  most  often  been  used  in  lakes,  ponds, 
and  reservoirs  to  study  fish  population  numbers.  Fish 
populations  can  be  estimated  by  capturing  and  re- 
leasing a  number  of  marked  fish  and  then  recaptur- 
ing marked  and  unmarked  fish. 

The  population  can  then  be  estimated  by  the 
formula  (Davis  and  Winstead  1980). 


N  = 


Mn 


m 


Where  N  =  estimated  number  of  fish  in  population 

Ci   =  number  of  fish  caught  in  first  sample 

C2  =  number  of  fish  caught  in  second 
sample 


Where  N  =  estimated  number  of  fish  in  population 

M  =  number  of  fish  marked 

n  =  total  number  of  fish  captured  in 
second  sample 

m  =  number  of  marked  fish  in  second  sample 

Assumptions  applied  to  this  type  of  population 
study  are  ( 1 )  fish  do  not  lose  their  marks,  (2) 
marked  fish  are  randomly  distributed  in  the  popula- 
tion, (  3 )  marked  and  unmarked  fish  are  equally  cap- 
turable,  and  (4)  numbers  of  unmarked  fish  are  not 
added  to  the  population  through  recruitment  or 
immigration  (Lagler  1956).  Many  variations  on  the 
mark-recapture  technique  have  been  developed.  For 
details  of  these  see  Lagler  (1956),  Ricker  (1958), 
Otis  et  al.  (1978),  or  White  et  al.  (1982). 

Length  of  Fish  (Level  3).  Rate  of  change  in  length 
of  individuals  and  length-frequency  distributions 
are  key  attributes  of  fish  populations  (Anderson  and 
Gutreuter  1983).  Fish  length  distributions  can  be 
developed  from  samples.  For  details  of  measurement 
and  analysis  techniques,  see  Anderson  and  Gutreuter 
(1983). 

Weight  of  Fish  (Level  4).  Weight  of  individual  fish 
and  fish  populations  are  also  key  attributes  of  fish 
populations  (Anderson  and  Gutreuter  1983).  The 
relationship  between  weight  and  length  can  also  be 
used  to  develop  condition  indexes.  For  details  of 
these  techniques  see  Anderson  and  Gutreuter 
(1983). 

Age  Determination  (Level  5).  Age  determination 
requires  a  substantial  increase  in  time  and  money 
(Table  2)  and  should  only  be  done  where  the 
problem(s)  being  addressed  require  such 
information.  See  Jearld  (1983)  for  details  on 
techniques  and  analysis. 


264 


Fish 


Radio  or  Sonar  Tagging  (Level  6).  Radio  or  sonar 
tagging  is  extremely  costly  (Table  2)  and  the 
biologist  should  be  sure  that  such  an  investment  is 
justified  before  beginning  a  study.  Thome  (1983) 
describes  the  use  of  sonar  or  hydroacoustics.  Winter 
et  ai.  (1978)  and  Winter  (1983)  describe 
underwater  biotelemetry  including  sonar  and 
radiotelemetry. 


HABITAT  INVENTORY  AND  MONITORING 
SYSTEMS 

Inventory  and  monitoring  fish  habitat  allow  the 
land  manager  to  identify  trends  in  habitat  conditions 
needed  for  fish  production  and  to  adjust  land  use 
to  stabilize  or  improve  habitat  conditions.  No  fish  in- 
ventory will  provide  all  of  the  data  needed  to  de- 
scribe habitat  conditions  for  specific  fish  species. 

A  number  of  systems  have  been  developed  to 
inventory  fish  habitat  conditions  and  to  relate  habitat 
conditions  to  actual  or  potential  fish  populations. 
Three  of  these  are  described  below. 

Habitat  Quality  Index  (HQI) 

Binns  (1979)  has  developed  a  habitat  quality 
index  (HQI)  for  predicting  biomass  of  trout  in 
streams  based  on  nine  measurable  habitat  attributes. 
The  index  is  described  in  more  detail  in  the  Streams 
chapter. 

Habitat  Evaluation  Procedures  (HEP)  and 
Habitat  Suitability  Index  (HSI)  Models 

The  U.S.  Fish  and  Wildlife  Service  has  developed 
Habitat  Evaluation  Procedures  (HEP)  which  provide 
numerical  ratings  of  habitat  quality  for  pre-  and  post- 
project  conditions  (U.S.  Department  of  Interior,  Fish 
and  Wildlife  Service  1976).  HEP  uses  one  or  more 


Habitat  Suitability  Index  (HSI)  models  for  such  anal- 
yses. These  models  use  one  or  more  habitat  variables 
to  generate  an  index  of  habitat  quality  between  0 
and  1  for  individual  fish  species.  Aquatic  applications 
of  HEP  and  HSI  models  are  described  in  detail  by 
Terrell  et  al.  (1982).  HSI  models  have  been  devel- 
oped and  published  for  many  fish  species.  The  biolo- 
gist may  find  these  useful  in  developing  an  inventory 
and  monitoring  system  or  program  other  than  HEP. 
Models  and  their  use  in  HEP  are  described  in  more 
detail  in  the  chapter  on  Habitat  Evaluation  Systems. 
More  information  on  HEP  and  HSI  models  and  on 
availability  of  HSI  models  for  fish  can  be  obtained 
from  "the  National  Ecology  Center,''  U.S.  Fish  and 
Wildlife  Service,  Drake  Creekside  Building  One, 
2627  Redwing  Road,  Fort  Collins,  CO  80526-2899. 

Riparian/Aquatic  Information  Data 
Summary  (RAIDS) 

Riparian/ Aquatic  Information  Data  Summary 
(RAIDS)  is  a  computer-based  system  for  storing  data 
on  aquatic  habitat  and  the  associated  fish  popula- 
tions. It  does  not  specify  how  such  data  must  be  col- 
lected. Details  on  the  system  can  be  obtained  from 
the  BLM's  Denver  Service  Center,  Division  of 
Resources. 


SUMMARY 

Inventory  of  fish  and  fish  habitat  is  needed  to 
provide  baseline  information  for  the  land  manager. 
After  baseline  conditions  have  been  assessed,  moni- 
toring helps  evaluate  land  management  actions  and 
provides  justification  for  leads  toward  adjusting  land 
use  where  necessary.  This  chapter  provides  guidance 
for  planning  and  organizing  inventories  and  monitor- 
ing studies  of  fish-producing  waters.  For  detailed 
descriptions  of  techniques,  equipment,  and  analyses, 
the  biologist  will  need  to  consult  the  Literature 
Cited  section  in  this  chapter. 


Fish 


265 


LITERATURE  CITED 


ANDERSON,  R.O.  and  S.J.  GUTREUTER  1983.  Length, 

weight,  and  associated  structural  indices.  Pages  283— 
300  in  Nielsen,  L.A.  and  D.L.  Johnson,  eds.,  Fisheries 
Techniques.  Am.  Fish.  Soc.  Bethesda,  MD.  468pp. 

ARMOUR,  C.L.  and  W.S.  PIATTS  1983.  Field  methods  and 
statistical  analyses  for  monitoring  small  salmonid 
streams.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.,  FWS/ 
OBS-83/33.  200pp. 

BINNS,  N.A.  1979-  A  habitat  quality  index  for  Wyoming 
trout  streams.  Fishery  Research  Report  Monograph 
Series  2.  Wyoming  Game  and  Fish  Dep.,  Cheyenne, 
82002.  75pp. 

BOWEN,  S.H.  1983.  Quantitative  description  of  the  diet. 
Pages  325—336  in  Nielsen,  LA.  and  D.L.  Johnson,  eds., 
Fisheries  Techniques.  Am.  Fish.  Soc.  Bethesda,  MD. 
468pp. 

CARLSON,  CA.  and  C.L.  GIFFORD  1983.  Finding  literature 
and  reports.  Pages  23—42  in  Nielsen,  LA.  and  D.L. 
Johnson,  eds.,  Fisheries  Techniques.  Am.  Fish.  Soc.  Be- 
thesda, MD.  468pp. 

CROUSE,  MR  and  RR.  KINDSCHY.  1981.  A  method  for 
predicting  riparian  vegetation  potential  of  semiarid 
rangelands.  Pages  1 10-1 14  in  Proc.  of  Symp.,  Acquisi- 
tion and  Utilization  of  Aquatic  Habitat  Inventory 
Information,  Western  Div.  Am.  Fish.  Soc.  Portland, 
OR.  Oct.  1981. 

DAVIES,  WD.  1983-  Sampling  with  toxicants.  Pages  199- 
214  in  Nielsen,  L.A.  and  D.L.  Johnson,  eds.,  Fisheries 
Techniques.  Am.  Fish.  Soc.  Bethesda,  MD.  468pp. 

DAVIS,  D.E.  and  R.L.  WINSTEAD.  1980.  Estimating  the 
numbers  of  wildlife  populations.  Pages  221-246  in 
Schemnitz,  S.D.,  ed.,  Wildlife  Management  Techniques 
Manual.  4th  ed.  revised,  The  Wildl.  Soc,  Washington, 
DC.  686pp. 

EISERMAN,  F.M.  1979.  Quantification  of  fluvial  trout  habi- 
tat in  Wyoming.  Trans.  Am.  Fish.  Soc.  108:215-228. 

FRITZ,  E.S.,  P.J.  RAGO,  and  LP.  MURARKA.  1980.  Strategy 
for  assessing  impacts  of  power  plants  on  fish  and 
shellfish  populations.  U.S.  Dep.  Inter.,  Fish  and  Wildl. 
Serv.,  Biological  Services  Program,  National  Power 
Plant  Team,  FWS/OBS-08/34.  68pp. 

GREESON,  P.E.,  T.A.  EHLKE,  GA.  IRWIN,  B.W.  LIUM,  and 
K.V.  SLACK.  1977.  Techniques  of  water  resources 
investigations  of  the  U.S.  Geological  Survey,  Chapter 
A4.  332pp. 

HAMILTON,  K.  and  E.P.  BERGERSEN.  1984.  Methods  to 
estimate  aquatic  habitat  variables.  Colorado  Coop. 
Fishery  Res.  Unit,  Colorado  State  University,  Ft.  Col- 
lins. 

HAYES,  ML.  1983.  Active  fish  capture  methods.  Pages 

123-146  in  Nielsen,  LA.  and  D.L.  Johnson,  eds.,  Fish- 
eries Techniques.  Am.  Fish.  Soc.  Bethesda,  MD. 
468pp. 

HUBERT,  W.A.  1983-  Passive  capture  techniques.  Pages 
95—122  in  Nielsen,  LA.  and  D.L.  Johnson,  eds.,  Fisher- 
ies Techniques.  Am.  Fish.  Soc.  Bethesda,  MD.  468pp. 

JEARLD,  A.  Jr.  1983-  Age  determination.  Pages  301-324  in 
Nielsen,  LA.  and  D.L.  Johnson,  eds.,  Fisheries  Tech- 
niques. Am.  Fish.  Soc.  Bethesda,  MD. 

JOHNSON,  D.L.  and  LA.  NIELSEN.  1983.  Sampling  consid- 
erations. Pages  1-21  in  Nielsen,  LA.  and  D.L  Johnson, 


eds.,  Fisheries  Techniques.  Am.  Fish.  Soc.  Bethesda, 
MD.  468pp. 

JOHNSON,  J.E.  and  J.N.  RINNE.  1982.  The  Endangered 

Species  Act  and  Southwest  fishes.  Pages  2-7  in  Fisher- 
ies Bulletin,  7(4),  Am.  Fish.  Soc. 

LAGLER,  K.R.  1956.  Freshwater  fishery  biology,  2nd  ed. 
Wm.  C.  Brown  Co.,  Dubuque,  LA.  421pp. 

NEILSON,  J.D.  and  GH.  GREEN.  1981.  Enumeration  of 
spawning  salmon  from  spawner  residence  time  and 
aerial  counts.  Trans.  Am.  Fish.  Soc.  110(4):554-556. 

NIELSEN,  L.A.  and  D.L.  JOHNSON,  eds.  1983.  Fisheries 
techniques.  Am.  Fish.  Soc.  Bethesda,  MD.  468pp. 

ODUM,  E.P.  1 978.  Ecological  importance  of  the  riparian 
zone,  strategies  for  protection  and  management  of 
floodplain  wetlands  and  other  riparian  ecosystems,  in 
Proc.  of  Symp.,  U.S.  Dep.  Agric,  For.  Serv.,  Washing- 
ton, DC. 

ORTH,  D.J.  1983-  Aquatic  Habitat  Measurements.  Pages 
61—84  in  Nielsen,  LA.  and  D.L.  Johnson,  eds.,  Fisher- 
ies Techniques.  Am.  Fish.  Soc.  Bethesda,  MD.  468pp. 

OTIS,  D.L.,  K.P.  BURNHAM,  G.C.  WHITE,  and  DR.  ANDER- 
SON. 1978.  Statistical  inference  from  capture  data 
on  closed  animal  populations.  Wildl.  Monogr.  62. 
135pp. 

REYNOLDS,  J.B.  1983.  Electronshing.  Pages  147-164  in 
Nielsen,  LA.  and  D.L.  Johnson,  eds.,  Fisheries  Tech- 
niques. Am.  Fish.  Soc.  Bethesda,  MD.  468pp. 

RICKER,  W.E.  1958.  Handbook  of  computations  for  bio- 
logical statistics  of  fish  populations.  Fish.  Res.  Board, 
Can.  Bull.  119.  300pp. 

SEBER,  G.A.F.  and  ED.  LECREN.  1967.  Estimating  popula- 
tion parameters  from  large  catches  relative  to  a  popu- 
lation. J.  Animal  Ecol.  36(3):631-643. 

SIMPSON,  J.C.  and  R.L.  WALLACE.  1978.  Fishes  of  Idaho. 
The  University  Press  of  Idaho.  Idaho  Research  Foun- 
dation, Inc.  Moscow.  237pp. 

SNYDER,  D.E.  1983.  Fish  eggs  and  larvae.  Pages  165-198 
in  Nielsen,  L.A.  and  D.L.  Johnson,  eds.,  Fisheries  Tech- 
niques. Am.  Fish.  Soc.  Bethesda,  MD.  468pp. 

TERRELL,  J.W.,  T.E.  MCMAHON,  P.D.  INSKIP,  R.F.  RA- 
LEIGH, and  K.L.  WILLIAMSON.  1982.  Habitat  suitabil- 
ity index  models:  Appendix  A.  Guidelines  for  riverine 
and  lacustrine  applications  of  fish  HSI  models  with 
the  habitat  evaluation  procedures.  U.S.  Dep.  Inter., 
Fish  and  Wildl.  Serv.,  FWS/OBS-82/10.A.  54pp. 

THORNE,  RE.  1983.  Hydroacoustics.  Pages  239-259  in 
Nielsen,  LA.  and  D.L.  Johnson,  eds..  Fisheries  Tech- 
niques. Am.  Fish.  Soc,  Bethesda,  MD.  468pp. 

U.S.  DEPARTMENT  OF  INTERIOR,  FISH  AND  WILDLIFE 
SERVICE.  1976.  Habitat  evaluation  procedures.  Divi- 
sion of  Ecological  Services.  Washington,  DC.  30pp. 

WHITE,  G.C,  DR.  ANDERSON,  K.P.  BURNHAM,  and  D.L. 
OTIS.  1982.  Capture-recapture  and  removal  methods 
for  sampling  closed  populations.  Los  Alamos  National 
Laboratory,  Los  Alamos,  NM.  LA-8787  NERP.  235pp. 

WINTER,  J.D.  1983-  Underwater  biotelemetry.  Pages  371- 
396  in  Nielsen,  LA.  and  D.L.  Johnson,  eds.,  Fisheries 
Techniques.  Am.  Fish.  Soc.  Bethesda,  MD.  468pp. 

,  V.B.  BUECHLE,  D.B.  SINIFF,  and  JR.  TESTER.  1978. 

Equipment  and  methods  for  radio  tracking  freshwater 
fish.  Univ.  Minnesota,  Agric.  Experiment  Sta.  Misc. 
Rep.  152-1978.  18pp. 

WYDOSKI,  R.  and  L  EMERY.  1983.  Tagging  and  marking. 
Pages  215-238  in  Nielsen,  LA.  and  D.L.  Johnson,  eds., 
Fisheries  Techniques.  Am.  Fish.  Soc,  Bethesda,  MD. 
468pp. 


266 


Fish 


, 


14 

AMPHIBIANS 
AND  REPTILES 

K.  Bruce  Jones 


U.S.  Bureau  of  Land  Management 
Phoenix  Training  Center 
Phoenix,  AZ  85015 


Editor's  Note:  Among  the  vertebrates,  amphibians 
and  reptiles  have  often  been  ignored.  "Traditional" 
biologists  trained  in  game  management  may  not 
be  familiar  with  studies  or  techniques  for  these 
species.  Very  little  has  been  published,  and  what's 
available  tends  to  be  located  in  hard-to-get  publi- 
cations— not  the  journals  or  manuals  familiar 
to  most  biologists. 

Nonetheless,  amphibians  and  reptiles  are  important 
components  of  many  ecosystems.  This  chapter 
brings  together  the  diffuse  techniques  for  studying 
these  species  and  their  habitats. 


INTRODUCTION 

Until  recently,  amphibians  and  reptiles  have 
been  largely  overlooked  in  land  management  consid- 
erations. Federal  and  State  agencies  have  spent  most 
of  their  time  and  money  on  commercially  important 
species,  and  the  public  is  generally  unaware  of  how 
important  these  animals  are  in  natural  ecosystems. 
Only  within  the  past  10  years  have  we  become  con- 
cerned about  conditions  of  these  animals  and  their 
habitats. 

Public  awareness  of  the  importance  of  the  entire 
wildlife  community  has  led  to  legislation  (e.g.,  the 
Federal  Land  Policy  and  Management  Act  of  1976) 
requiring  study  and  management  of  nongame  wildlife 
such  as  amphibians  and  reptiles.  Studies  have  demon- 
strated the  importance  of  these  animals  in  natural 
ecosystems.  Terrestrial  amphibians,  lizards,  and  bur- 
rowing snakes  are  excellent  indicators  of  the  relative 
amounts  of  microhabitats  in  ecosystems.  For  exam- 
ple, the  Sacramento  Mountain  salamander  (Aneides 
hardii)  is  entirely  dependent  on  moist,  downed  logs 
and  litter  on  north-  and  northeast-facing  slopes 
(Wiegman  et  al.  1980).  In  addition,  information  on 
amphibian  and  reptile  abundance  and  diversity  helps 
determine  the  relative  health  of  ecosystems.  For 
example,  lizard  abundance  and  diversity  fluctuate 
directly  with  changes  in  the  composition  and 
amount  of  microhabitats.  These  microhabitat  changes 
often  result  from  land  management  practices  (Jones 
1981a;  Ortega  et  al.  1982;  Tinkle  1982;  Luckenbach 
and  Bury  1983). 

Aquatic  amphibians  and  snakes  are  good  indica- 
tors of  the  health  of  aquatic  systems.  These  animals 
are  especially  sensitive  to  pollution  and  loss  of 
aquatic  habitat  (Hall  1980). 

Amphibians  and  reptiles  are  also  important  in 
food  chains,  and  they  make  up  large  proportions  of 
vertebrates  in  certain  ecosystems  (Bury  and  Raphael 


'Current  address:  Office  of  Endangered  Species,  U.S.  Fish  and 
Wildlife  Service,  Washington, DC  20240. 


Amphibians  and  Reptiles 


267 


1983).  Many  carnivorous  mammals  and  raptorial 
birds  rely  on  these  animals  for  food.  For  example,  up 
to  50%  of  the  spring  diet  of  common  black  hawks 
(Buteogallus  anthracinus)  consists  of  aquatic  herpe- 
tofauna  (Millsap  and  Harrison  1981).  Recent  studies 
have  also  demonstrated  the  economic  importance, 
both  consumptive  and  nonconsumptive,  of  these 
animals  (see  Bury  et  al.  1980). 

Because  of  recent  concern  for  nongame  wildlife, 
biologists  and  land  managers  find  themselves  faced 
with  studies  and  management  needs  for  a  group 
of  animals  they  know  little  about.  Only  a  few  biolo- 
gists and  virtually  no  land  managers  have  had  any 
formal  education  in  herpetology. 

There  is  a  wealth  of  information  on  the  ecology 
of  individual  amphibians  and  reptiles  in  literature. 
Unfortunately,  few  papers  deal  with  amphibian  and 
reptile  communities,  and  even  fewer  summarize 
information  needs,  data  collection,  and  management 
on  public  lands. 

This  chapter  is  designed  to  help  biologists  who 
have  little  or  no  education  or  experience  with  am- 
phibians and  reptiles  organize  and  run  inventories  or 
monitoring  studies  on  selected  species  or  major 
groups  (e.g.,  lizards). 


HABITAT  FEATURES 

The  most  important  factor  affecting  amphibian 
and  reptile  distribution  and  habitat  use  is  horizontal 
and  vertical  habitat  availability.  In  studying  popula- 
tion condition,  trend,  and  cause  and  effect  factors, 
biologists  must  collect  data  on  habitat. 

Sixteen  habitat  components  and  attributes  are 
important  determinants  of  amphibian  and  reptile 
abundance  (Table  1). 

Microhabitat  components  are  site-specific,  physi- 
cal entities  that  provide  environmental  conditions 
necessary  for  a  wide  variety  of  ecological  functions 
such  as  reproduction,  foraging,  predator  avoidance 
or  escape,  thermoregulation,  and  resting.  Amphibians 
and  reptiles  are  ectothermic  (cold-blooded);  body 
temperatures  are  not  derived  from  metabolic  pro- 
cesses but  rather  from  the  surrounding  environment. 
Therefore,  behavioral  adaptations  and  use  of  different 
microhabitats  by  amphibians  and  reptiles  are  diverse. 
These  animals  often  demonstrate  a  high  dependence 
on  certain  microhabitats  to  thermoregulate.  Removal 
or  reduction  of  microhabitats  necessary  for  thermo- 
regulation can  detrimentally  affect  all  other  ecologi- 
cal functions  because  internal  temperature  regulation 
determines  the  intensity  of  activity  (see  Brattstrom 
1965  for  examples).  For  example,  certain  basking 
turtles  can  be  extirpated  from  a  pond  if  floating  logs 
are  eliminated. 


Amphibians  are  even  further  restricted  to  moist 
or  wet  habitats  with  varying  degrees  of  free-standing 
water.  Many  larval  toads,  frogs,  and  salamanders  re- 
quire water  for  development  into  adult  stages  (Steb- 
bins  1966).  Additionally,  some  adult  frogs  and 
salamanders  require  water  to  avoid  desiccation  and 
to  respirate  cutaneously.  For  example,  many  Pletho- 
don  (lungless)  salamanders  require  moist,  rotting 
logs  and  litter  for  egg  development  and  adult  cuta- 
neous respiration  (Stebbins  1966).  Without  rotting 
logs  and  associated  litter,  there  is  insufficient  mois- 
ture for  egg  development  and  adult  survival. 

Although  most  biologists  and  some  land  man- 
agers understand  the  significance  of  vegetation  struc- 
ture and  composition  to  wildlife,  few  realize  how 
important  litter  (fallen  logs,  leaves),  soil,  plant  root 
structure,  and  horizontal  vegetation  structure  are  to 
most  forms  of  wildlife,  especially  amphibians  and 
reptiles.  Litter  provides  moisture  regimes  necessary 
for  development  of  terrestrial  salamanders,  but  it 
also  provides  cool  surface  and  below-the-surface 
temperature  regimes.  Often  this  is  the  only  microhab- 
itat with  moderate  environmental  temperatures  in 
otherwise  hot,  dry  regions.  Because  of  low  preferred 
body  temperature  ranges,  many  species  such  as  the 
Arizona  alligator  lizard  (Elgaria  kingii)  and  Gilbert 
skink  (Eutneces  gilberti)  are  entirely  restricted  to 
deciduous  riparian  habitats  within  desert  regions 
(Stebbins  1966).  In  addition,  litter  provides  feeding 
substrate  for  several  common  lizards,  especially 
members  of  the  genus  Sceloporus  (fence  or  spiny 
lizards).  Loss  of  fallen  logs  and  trees  can  drastically 
reduce  populations  of  these  lizards  (see  discussions 
of  Jones  1981a  &  b  and  Vitt  et  al.  1974,  for  depen- 
dence on  woody  vegetation ). 

Soil  depth,  texture,  and  diversity  are  extremely 
important  in  determining  amphibian  and  reptile  dis- 
tribution in  an  area.  For  example,  soil  depth  and 


Fallen  logs  provide  important  microhabitats  for  terrestrial 
salamanders. 


268 


Amphibians  and  Reptiles 


Table  1.     Important  components  and  attributes  of  amphibian  and  reptile  habitat. 


Habitat  Component 

Types 

Variables/ 
Factors 

Amphibian  and 
Reptile  Associations 

Relations  of  Species' 
Ecology  to 
Components 

CD 
CO 

c 

3 
"S3 
Q 

0 
> 

o 

CJ 

a> 

CL 
CO 

o 
en 

LU 

0) 
CO 

to 

n 

-D 
CO 

CD 

c 
■o 

CD 
CD 
LL. 

CD 
Q. 

O 

r> 
o 
o 

LL 

s 

t/5 

n 

CO 

to 
CD 

Z 

CO 

o 

CD 

o 
o 

CO 

>. 

-C 

0_ 

c 
o 

o 

13 

o 

Q. 
CD 
DC 

C 

g 
J5 

Z3 
CO 
CD 

6 

E 

CD 

.C 
1— 

Microhabitat  Component 

Rivers  and 
streams 

Riffle/run/pool 
ratios,  water 
temperature,  tur- 
bidity, DO,  or- 
ganic content, 
siltation,  pollu- 
tants 

AQAL,  SAAL 

• 

• 

• 

• 

• 

• 

Water 
(Lotic) 

Water 
(Permanent  Lentic) 

Ponds,  marshes, 
lakes,  reservoirs, 
natural  catch- 
ments, agricul- 
tural run-off 

Water  tempera- 
ture, DO,  organic 
content,  siltation, 
pollutants,  emer- 
gent vegetation, 
logs  and  other 
litter,  substrate 

AQAL,  SAAL 

• 

• 

• 

• 

• 

• 

Water 
(Temporary  Lentic) 

Temporary  rain 
pools,  irrigation 
ditches 

DO,  water  tem- 
peratures, silta- 
tion, pollutants, 
duration,  fre- 
quency, emer- 
gent vegetation, 
substrate 

SATD,  SASA 
SATU,  SASN 

• 

• 

• 

• 

• 

• 

Rock 

Talus  slopes, 
cliffs,  boulders, 
substrate 

Rock  size,  heter- 
ogeneity, inter- 
faces, origin, 
vertical  and 
horizontal  struc- 
ture 

AQAL,  SAAL  (Where 
interfaced  with  aquatic 
habitats),  TAL 

• 

• 

• 

• 

• 

• 

Vegetation — 
Litter/Debris 

Leaves,  logs, 
limbs 

Litter  size,  depth, 
heterogeneity, 
horizontal  struc- 
ture, type,  mois- 
ture retention, 
temperature 

AQAL,  SAAL  (Where 
interfaced  with  aquatic 
habitats),  TAL 

• 

• 

• 

• 

• 

Vegetation — 
Live 

All  vegetation 
including  roots 

Horizontal  and 
vertical  structure, 
interfaces  with 
abiotic  compo- 
nents, hetero- 
geneity 

AQAL,  SAAL  (Where 
interfaced  with  aquatic 
habitats)  TAL 

• 

• 

• 

• 

• 

• 

• 

Vegetation — 
Dead 

Standing  vegeta- 
tion, roots 

Size,  interface 
with  other  habi- 
tats, heterogene- 
ity, vertical  and 
horizontal  struc- 
ture, soils  (roots 
only) 

AQAL,  SAAL  (Where 
interfaced  with  aquatic 
habitats)  TAL 

• 

• 

• 

• 

• 

• 

Vegetation — 
Plant  Species 

Individual  plant 
species 

Individual  plant 
species  abun- 
dance and  oc- 
currence 

TL,  TSN 

• 

• 

• 

Amphibians  and  Reptiles 


269 


Table  1.     Important  components  and  attributes  of  amphibian  and  reptile  habitat  (concluded). 


Habitat  Component 

Types 

Variables/ 
Factors 

Amphibian  and 
Reptile  Associations 

Relations  of  Species' 
Ecology  to 
Components 

CD 

CO 

C 

£ 

CD 

Q 

CD 
> 

O 

o 

CD 
Q. 
CO 
O 
CO 
LU 

CD 

CO 
JD 

3 
CO 

a> 

c 

CD 
CD 

LL 

>> 

CD 

Q. 

O 

"O 

o 
o 

LL 

CD 
00 

n 

3 
CO 

to 
CD 

CO 
CJ 
O) 

o 
o 

CO 

>. 

-C 
Q. 

c 
o 

o 

3 
"D 
O 

Q. 
CD 
DC 

g 
to 

3 
O) 
CD 

O 

E 

CD 
-C 

Microhabitat  Component 
(cont.) 

Surface  and 

subsurface  soil 
types 

Types,  death, 
heterogeneity, 
horizontal  and 
vertical  structure, 
interfaces  with 
other  habitat 
components 

AQAL,  SAAL  (Where 
interfaced  with  aquatic 
habitats),  TAL 

• 

• 

• 

• 

• 

• 

Soil 

Macrohabitat 
Components/Factors 

%  angle  of  area 
from  horizontal 

%  slope,  mois- 
ture availability, 
thermal  regimes, 
vegetation  struc- 
ture 

SAAL,  TAL 

• 

• 

• 

• 

• 

Slope 

Aspect 

South,  north,  east, 
and  west  facing 

Direction,  mois- 
ture availability, 
temperature 
regimes,  vegeta- 
tion structure 

SAAL,  TAL 

• 

• 

• 

• 

• 

Elevation 

Vertical,  above  or 
below  sea  level 

Vertical  distance, 
moisture  avail- 
ability, thermal 
regimes,  vegeta- 
tion structure 

AQAL,  SAAL,  TAL 

• 

• 

• 

• 

• 

• 

Precipitation 

All  forms 

Quantity,  type, 
duration,  fre- 
quency, moisture 
availability,  ther- 
mal regimes, 
vegetation  struc- 
ture 

AQAL,  SAAL  TAL 

Ecotones/Habitat 
Juxtaposition 

Habitat  interfaces 
and  locations 

Heterogenity, 
interface  size 
and  cumulative 
numbers,  posi- 
tion of  habitats 

AQAL,  SAAL  TAL 

• 

• 

• 

• 

• 

• 

• 

Topography 

General 

Slope,  aspect, 
and  physical 
feature  mixture 
and  position 

SAAL,  TAL 

• 

• 

• 

• 

• 

• 

• 

Geographic  Location 

Major  geographic 
boundaries  and 
barriers 

Size,  location 
and  frequency, 
habitat  size  and 
disjunction 

AQAL,  SAAL,  TAL 

• 

Codes  for  Amphibian  and  Reptile  Associations: 
AQ — aquatic 
SA — semiaquatic 

T — terrestrial 
TU— turtles 
FR — frogs 


TD— toads 
SA — salamanders 
LI — lizards 
SN — snakes 
AL — all  amphibians  and  reptiles 


270 


Amphibians  and  Reptiles 


texture  determine  the  rate  of  percolation  and  soil 
layer  moisture. 

In  some  areas  with  heavy  clay  soils,  water  accu- 
mulates on  the  surface,  especially  during  rainy  sea- 
sons. Because  surface  water  may  be  available  for  up 
to  3  weeks,  many  semiaquatic  amphibians  occur  in 
these  regions.  For  example,  six  species  of  semi- 
aquatic  toads  were  verified  within  a  small,  hot  region 
of  the  Sonoran  Desert  that  had  deep  clay  soils  (Jones 
et  al.  1983). 

Although  moist,  clay  soils  enhance  the  distribu- 
tion of  toads  in  some  regions,  hard,  shallow,  or  rocky 
soils  may  prevent  species  colonization  of  an  area. 
For  example,  spadefoot  toads  {Scaphiopus  sp.)  are 
well  adapted  to  survive  in  dry,  desert  regions,  but 
without  soils  loose  enough  for  burrowing,  generally 
will  not  occur  in  an  area  (unless  rodent  burrows  are 
present ). 

The  diversity  of  soils  interfaced  with  rocks  and 
other  structures  often  determines  amphibian  and 
reptile  species  richness.  Diversity  provides  more 
niches  for  amphibian  and  reptile  colonization  and 
existence. 

Plant  root  structure  also  contributes  greatly  to 
amphibian  and  reptile  species  richness  by  providing 
avenues  into  subsurface  space.  These  plant-created 
avenues  often  account  for  greater  species  richness 
than  expected  in  what  are  thought  to  be  structure- 
less plant  communities.  For  example,  deserts  domi- 
nated by  creosotebush  (Larrea  sp. )  appear  very 
homogeneous  with  little  horizontal  or  vertical  struc- 
ture. However,  the  root  systems  of  creosote  are  ex- 
tensive, and  they  provide  a  wide  variety  of 
subsurface  opportunities  to  many  of  the  desert's 
nocturnal  amphibians  and  reptiles.  The  result  is  a 
diverse  amphibian  and  reptile  community. 

Generally,  horizontal  vegetation  structure  deter- 
mines amphibian  and  reptile  composition  more  often 
than  vertical  structure,  especially  in  lizards.  This 
results  primarily  from  foraging  and  thermoregulation 
which  are  generally  conducted  on  the  surface  (there 
are  exceptions — some  amphibians  and  reptiles  forage 
in  trees  and  below  the  surface).  For  example,  zebra- 
tailed  lizards  (Callisaurus  draconoides)  prefer 
brushy,  open  habitats  where  they  ambush  prey  while 
sitting  in  the  open  and  then  move  into  shade  to  ther- 
moregulate  (Pianka  1966;  Jones  1981a  &  b).  Some 
species  such  as  desert  iguanas  (Dipsosaurus  dor- 
salts)  are  common  in  very  open,  sparsely  vegetated 
creosotebush  habitats;  this  habitat  allows  them  to 
reach  high  preferred  body  temperatures  and  obtain 
specific  food  items  (creosotebush  buds).  Other  spe- 
cies that  forage  in  or  on  litter  and  have  moderate 
internal  temperature  ranges  prefer  densely  vegetated 
habitats  with  large  amounts  of  surface  litter  (e.g., 
alligator  lizard). 


Macrohabitat  components  are  physical  phenom- 
ena that  affect  environmental  conditions  of  microhab- 
itats  and  in  some  instances  entire  regions.  Different 
types  of  microhabitat  and  macrohabitat  components 
are  listed  with  corresponding  variables  or  factors 
that  affect  the  ecological  requirements  of  species 
(Table  1 ).  Amphibians  and  reptiles  are  listed  by  ma- 
jor taxonomic  group  (e.g.,  frogs,  lizards).  Table  1 
serves  primarily  as  a  checklist  of  habitat  components 
and  variables  that  biologists  should  consider  when 
conducting  habitat  and  amphibian  and  reptile  sur- 
veys or  studies.  Biologists  should  collect  data  on 
soils,  litter,  water  availability,  and  horizontal  vegeta- 
tion on  each  amphibian  and  reptile  sample  site.  This 
information  will  allow  the  biologist  to  accurately 
assess  the  relation  between  microhabitat  availability 
and  presence  or  absence  of  certain  species.  By  being 
aware  of  macrohabitat  conditions,  biologists  will 
better  understand  causes  of  microhabitat  variables, 
and  how  and  where  to  set  up  samples  (e.g.,  biolo- 
gists can  reduce  sample  variability  due  to  slope  by 
sampling  only  north-facing  slopes,  or  they  may  elect 
to  sample  all  different  aspects  to  determine  their 
effect  on  species  composition). 


Zebra-tailed  lizard. 


POPULATION  MEASUREMENT  TECHNIQUES 

Sources  of  Information 

Although  amphibian  and  reptile  data  collection 
generally  does  not  require  the  degree  of  expertise 
required  for  surveys  of  animals  such  as  birds,  a  biolo- 
gist with  little  or  no  previous  experience  will  need 
to  review  certain  literature  and,  when  possible,  con- 
tact local  or  regional  experts.  Field  guides  and  keys 
that  provide  characteristics  for  identifying  various 
herpetofauna  (Table  2)  are  available  for  most  regions 
of  the  U.S.  The  biologist  should  thoroughly  review 
species  characteristics  before  starting  field  surveys. 


Amphibians  and  Reptiles 


271 


Review  of  field  guides,  books,  and  amphibian  and 
reptile  journals  will  also  provide  ecological  back- 
ground on  various  species  that  will  help  biologists 
select  the  most  appropriate  sampling  methods. 

Museum  and  individual  collection  records  and 
publications  can  be  used  to  verify  amphibian  and 
reptile  species  occurrence  within  a  geographic  area. 
Searches  of  these  sources  should  be  made  before 
starting  censuses  to  avoid  duplication. 

Several  large  museums  in  the  U.S.  hold  vast  rec- 
ords of  amphibians  and  reptiles.  Edwards  ( 1975)  lists 
U.S.  collections  of  amphibians  and  reptiles,  their 
locations,  and  major  sources  of  records  (e.g.,  Desert 
Southwest).  This  publication  serves  as  an  excellent 
starting  point  for  record  searches. 

Expert  herpetologists  often  possess  personal 
locality  records  that  can  add  to  species  lists  in  spe- 


cific geographic  regions.  Herpetologists  with  regional 
expertise  should  be  consulted  before  field  work  is 
initiated. 

Many  journals  contain  valuable  amphibian  and 
reptile  occurrence  and  ecological  data  (Table  2). 

Problems  Affecting  Sampling 

Biologists  should  be  aware  of  several  factors  that 
may  affect  their  results  when  sampling  amphibians 
and  reptiles.  Perhaps  the  largest  problem  in  assessing 
amphibian  and  reptile  populations  is  that  these  ani- 
mals' activities  and  reproduction  vary  with  natural 
environmental  fluctuations,  such  as  precipitation  and 
temperature  (Whitford  and  Creusere  1977;  Gibbons 
and  Semlitsch  1981;  Vogt  and  Hine  1982).  There- 
fore, biologists  should  be  cautious  in  interpreting 
cause  and  effect  data  because  observed  differences 
may  result  from  natural  fluctuations  in  weather 


Table  2.     Examples  of  major  references  providing  preliminary  and  supplemental  data  for  amphibian  and 
reptile  surveys  and  studies. 


Journal  or  Publication 

Types  of  Information 

Geographic  Applicability 

05 

o 
> 

CO 
-C 
CD 
CO 

CD 

cz 
o 

JD 

CO 

b 

CO 

O 

en 
o 
o 
o 

LU 

00 
CD 

O    CO 
CD    Q. 

S  o 
■£ ' « 

c 
o 

00 

o 
c 

CD 
"O 

C 
CD 

E 

CD 
CD 
CO 

c 
CO 

CO 
O 
D) 
O 
O 
00 

x: 
CL 

00 
CD 

Q. 
CO 

O) 
O 

in 

u 

E 
£ 

00 

>^ 

in 

o 

E 
o 

O 

CO 

h- 

Herpetologica 

• 

• 

• 

• 

WH 

Copeia 

• 

• 

WH 

Journal  of  Herpetology 

• 

• 

WW 

Ecology 

• 

O 

WW 

Ecological  Monographs 

• 

WW 

Herpetological  Review 

• 

• 

• 

• 

Primarily  US 

American  Midland  Naturalist 

• 

o 

• 

Primarily  CUS 

Great  Basin  Naturalist 

• 

o 

• 

Primarily  WUS 

Southwestern  Naturalist 

• 

• 

O 

• 

Primarily  SW  and  ME 

Academy  of  Sciences  publica- 
tions (Arizona,  Utah,  Kansas, 
Texas,  etc.) 

• 

• 

• 

• 

• 

Limited  generally  to  regions 
around  applicable  states 

Bulletin  of  the  Maryland  Herpeto- 
logical Society 

• 

• 

• 

• 

• 

Primarily  US  and  ME 

American  Museum  Novitiates 

• 

WW 

University  of  Kansas  Publications 
in  Zoology 

• 

• 

• 

Primarily  CUS  and  WUS 

University  of  California  Publica- 
tions in  Zoology 

• 

• 

• 

Primarily  WUS 

Smithsonian  Herpetological  Infor- 
mation Service 

• 

• 

• 

• 

WW 

°  Limited 


272 


Amphibians  and  Reptiles 


. 


Table  2.     Examples  of  major  references  providing  preliminary  and  supplemental  data  for  amphibian  and 
reptile  surveys  and  studies  (concluded). 


Journal  or  Publication 

Types  of  Information 

Geographic  Applicability 

2 
o 
> 

CO 

.c 

03 
CD 

03 

c 
g 

"3 
n 

00 

b 

CO 

o 
D) 
O 
O 

o 

LU 

00 

cu 

CJ     00 

cd  o. 

"  "to 
■*->    d 

So 
5  ' « 
1* 

c 
g 

to 
o 

"c 

CD 
"O 

c 

CD 

E 

CD 
O) 
CO 
c 
CO 

2 

CO 
O 
O) 
O 

o 

00 

Q. 

00 
CD 

a 

CO 

a> 
o 

n 

CD 

o 
to 

E 

0) 

w 
>. 

o 

'E 
o 

c 
O 
X 
CO 

1— 

Examples  of  books  with  broad  geographic  coverage 

Turtles  of  the  United  States  by  C. 
H.  Ernst  and  R.  W.  Barbour 
(1973) 

• 

• 

US 

Handbook  of  Lizards  by  H.  M. 
Smith  (1946) 

• 

• 

US 

A  Field  Guide  to  Western  Reptiles 
and  Amphibians  by  R.  C.  Steb- 
bins(1966) 

• 

cus,  wus 

Handbook  of  Frogs  and  Toads  of 
the  United  States  and  Canada 
by  AH.  Wright  and  A. A.  Wright 

• 

US  and  CA 

Handbook  of  Snakes  of  the 
United  States  and  Canada  by 
A.H.  Wright  and  A.A.  Wright 
(1957) 

• 

US  and  CA 

Handbook  of  Salamanders  by 
S.C.  Bishop  (1947) 

• 

• 

• 

• 

• 

• 

• 

• 

US  and  CA 

A  Field  Guide  to  Reptiles  and 
Amphibians  of  Eastern  and  Cen- 
tral North  America  by  R.  Conant 
(1975) 

• 

• 

• 

• 

• 

• 

• 

• 

CUS  and  EUS 

Example  of  books  with  regional  geographic  coverage3 

Amphibians  and  Reptiles  of  the 
Pacific  Northwest  by  R.  Nuss- 
baum  et  al.  (1983) 

Be  sure  to  check  thoroughly  for  these.  There  are  many  regional-  and  area-specific  books. 


Codes  for  Geographic  Applicability: 

CA   —Canada 
CUS— Central  United  States 
EUS  —Eastern  United  States 
ME   — Mexico 


SW  — Southwestern  United  States 
US    —United  States 
WH  — Western  hemisphere 
WUS— Western  United  States 
WW  —Worldwide 


rather  than  man-caused  changes.  Gibbons  and  Sem- 
litsch  (1981)  recommend  long-duration  studies  (>  5 
years)  in  assessing  amphibian  and  reptile  popula- 
tions. Shorter-duration  studies  in  an  area  can  yield 
species  occurrences,  and  they  are  also  used  to  com- 
pare composition  of  herpetological  communities 
provided  sample  areas  are  roughly  adjacent  to  each 
other  and  climatic  factors  are  considered 
(Jones  1981a). 


Other  major  factors  affecting  amphibian  and 
reptile  sampling  are  differences  in  species  morphol- 
ogy, physiology,  and  behavior  such  as  activity  pat- 
terns and  movement.  For  example,  certain  lizards 
and  snakes  are  too  large  (e.g.,  desert  iguana)  to  cap- 
ture in  pitfall  traps.  Other  methods  should  be  used 
in  conjunction  with  pitfall  trapping  to  adequately 
census  large  lizards  and  snakes.  Populations  of  rela- 
tively sedentary,  microhabitat-specinc  animals  are 


Amphibians  and  Reptiles 


273 


also  underestimated  by  pitfall  and  other  trapping 
methods  because  the  probability  of  these  animals 
falling  into  traps  is  far  less  than  those  species  that 
move  over  larger  areas.  An  active  searching  method 
such  as  turning  logs,  rocks,  and  debris  (Bury  and 
Raphael  1983)  would  yield  more  accurate  estimates 
of  sedentary  species  occurrence  and  abundance  than 
traps. 

The  frequency  and  type  of  amphibian  and  rep- 
tile movement  will  also  affect  sampling.  For  example, 
there  are  many  burrowing,  nocturnal  snakes  in  the 
Sonoran  Desert.  Transects  run  during  daylight  hours 
rarely  verify  nocturnal,  burrowing  species  even 
when  rocks  and  debris  are  searched.  Addition  of 
pitfall  traps  and  drift  fences  to  a  diurnal  transect 
greatly  increases  the  accuracy  of  the  sample. 

Yearly  fluctuations  in  amphibian  and  reptile 
activity  also  affect  verification  of  species  occurrence 
in  an  area.  Toads  within  the  Desert  Southwest  are 
active  primarily  during  thunderstorms  in  July  and 
August.  Samples  taken  other  than  in  July  and  August 
yield  poor  results. 

Certain  microhabitat-specific  species  tend  to 
move  in  nonrandom  patterns  between  preferred  hab- 
itats (see  Pianka  1966  for  lizard  examples),  and  this 
affects  accuracy  of  census  methods  (Gibbons  and 
Semlitsch  1981 ).  Placement  of  traps  and  fences  is  ex- 
tremely important  when  considering  different  move- 
ment patterns  in  lizards.  For  example,  zebra-tailed 
lizards  move  between  large  bushes  along  drainages 
(Pianka  1966).  Pitfall  traps  and  drift  fences  placed  in 
open  areas  with  few  bushes  will  catch  few  zebra- 
tailed  lizards,  whereas  fences  and  traps  placed  be- 
tween bushes  will  catch  more  lizards. 

Daily  and  weekly  weather  patterns  also  affect 
amphibian  and  reptile  activity.  Species  with  narrow 
preferred  temperature  ranges  are  more  likely  to  be 
active  during  a  smaller  range  of  environmental  tem- 
peratures. Therefore,  samples  can  be  variable  during 
weeks  with  large  temperature  differences.  There  may 
also  be  considerable  daily  and  seasonal  differences 
in  movement  between  different  age  and  size  classes 
and  sexes. 

Abilities  of  the  biologist  will  also  affect  the  accu- 
racy of  samples,  especially  transect  and  tape-record- 
ing techniques  where  animals  are  not  closely 
viewed.  For  example,  a  high  degree  of  expertise  is 
needed  to  identify  lizards  on  a  walk-through  transect. 
Similarly,  a  biologist  must  be  familiar  with  toad  and 
frog  calls  if  auditory  data  are  to  be  used  for  species 
identification.  Conversely,  pitfall  and  other  trapping 
methods  do  not  require  specific  expertise  because 
amphibians  and  reptiles  can  be  closely  viewed. 

Other  life  history  limitations  may  also  affect  the 
interpretation  of  sampling  data.  Whereas  one  lizard 


may  produce  multiple  clutches  within  a  year,  an- 
other may  reproduce  only  every  other  year.  This 
type  of  result  could  be  easily  misinterpreted  if  the 
biologist  does  not  consider  differences  in  reproduc- 
tive strategies.  Similarly,  size  and  age  class  ratios 
can  be  misinterpreted  if  the  surveyor  is  not  familiar 
with  differences  in  species'  life  history.  For  example, 
juveniles  tend  to  move  less  and,  therefore,  would 
be  underestimated  by  techniques  such  as  pitfall 
trapping. 

Relatively  accurate  samples  can  be  obtained  for 
individual  species  and  entire  amphibian  and  reptile 
communities  provided  that  species'  life  histories  and 
sampling  method  limitations  are  clearly  understood. 


Sampling  Methods 

There  are  many  methods  used  to  sample  am- 
phibians and  reptiles.  The  method  chosen  by  a  biolo- 
gist will  depend  on  the  objectives.  For  example,  if 
biologists  need  to  develop  a  species  list  by  habitat 
type  for  a  planning  document,  then  they  should  se- 
lect some  combination  of  verification  techniques. 
However,  if  biologists  need  to  show  differences  in 
species  abundance  between  habitat  types,  then  they 
must  select  a  method  that  yields  relative  abundance. 
Generally,  data  on  species  presence  are  adequate 
for  assessing  species  richness  of  habitat  types.  Abun- 
dance and  density  techniques  provide  additional  data 
for  comparing  habitat  types  and  also  provide  infor- 
mation about  individual  species  fitness.  For  example, 
density  estimates  provide  biologists  with  data  on 
the  relative  fitness  of  special  status  reptiles  such  as 
the  desert  tortoise  (Gopherus  agassizii). 

Depending  on  the  biologist's  objectives,  there 
are  generally  two  ways  to  collect  species  informa- 
tion: direct  and  indirect.  Direct  sampling  of  amphibi- 
ans and  reptiles  involves  observation  of  animals 
occurring  on  a  sample  site.  Indirect  sampling  in- 
volves obtaining  species  information  on  a  sample  site 
without  observing  the  animal.  Before  initiating  any 
sampling,  the  biologist  should  contact  the  State  game 
and  fish  agency  to  obtain  the  necessary  regulations 
and  collecting  permits. 


Direct  Search  Methods 

One  of  the  simplest  ways  of  verifying  species 
occurrence  in  an  area  is  to  walk  or  drive  through 
the  area,  recording  all  amphibians  and  reptiles  seen. 
Because  species  are  separated  in  time  as  well  as 
space  (see  Creusere  and  Whitford  1982  for  an  exam- 
ple in  a  lizard  community),  searches  run  at  different 
times  of  the  day  will  yield  different  species  in  vary- 
ing numbers  (searches  run  during  the  daylight  hours 
verify  some  diurnal  species  and  those  run  at  night 
some  nocturnal  species). 


274 


Amphibians  and  Reptiles 


Searches  of  areas  are  either  random  and  oppor- 
tunistic or  systematic  within  a  defined  time  and  area. 
Many  records  of  amphibians  and  reptiles  result  from 
casual,  opportunistic  observations  made  during  field 
work.  Cumulatively,  these  observations  have  contrib- 
uted more  to  known  occurrence  of  amphibians  and 
reptiles  than  any  other  method.  Although  this 
method  often  yields  valuable  locality  records,  it  gen- 
erally involves  a  great  deal  of  search  time.  It  also 
verifies  only  those  species  that  spend  a  great  deal  of 
time  on  the  surface  or  under  rocks  and  debris  such 
as  logs.  Generally,  only  a  small  percentage  of  species 
that  occupy  an  area  are  verified  during  searches. 
Only  through  months  and  often  years  are  all  species 
in  an  area  verified  by  search  methods. 

Road  riding  (road  cruising)  is  one  of  the  most 
popular  search  methods  used  to  verify  and  collect 
amphibians  and  reptiles,  especially  nocturnal  species 
(see  discussion  by  Campbell  and  Christman  1982). 
Generally  road  riding  consists  of  cruising  secondary 
roads  at  speeds  of  35-55  km/h  (22-34  mph),  using 
low  headlight  beams.  Many  small  nocturnal  snakes, 
lizards,  and  frogs  not  normally  found  during  daytime 
field  searches  are  obtained  by  this  method  when 
conducted  between  dusk  and  2  to  3  hours  after 
dusk.  Use  a  spotlight  or  strong  flashlight  while  road 
riding  to  spot  animals  moving  off  of  roads.  Flashlights 
are  useful  for  locating  amphibians  and  reptiles  on 
sides  of  roads  (see  Vitt  and  Ohmart  1978  for  an  ex- 
ample of  road-riding  results).  Road  riding  is  most 
productive  when  run  without  moonlight.  Road-riding 
routes  can  be  systematically  established,  and  relative 
abundance  of  amphibians  and  reptiles  can  be  ob- 
tained by  standardizing  sampling  efforts  by  units  of 
time  in  each  habitat.  For  example,  biologists  can 
measure  the  distance  of  each  habitat  along  a  road, 
and  measure  the  amount  of  time  spent  in  each  seg- 
ment of  the  habitat.  They  can  then  express  individ- 
ual species  abundance  data  as  the  number  of  animals 
seen  per  mile  of  habitat  per  hour.  This  allows  for 
comparison  of  habitats  using  road-riding  procedures. 
Bury  and  Raphael  (1983)  refer  to  searches  con- 
ducted per  unit  effort  of  time  as  time-constraint 
procedures. 

Although  providing  occurrence  and  abundance 
data  on  some  nocturnal  and  secretive  herpetofauna, 
road  riding: 

( 1 )  is  time-consuming; 

(2)  yields  relatively  few  records; 

(3)  verifies  only  nocturnal  species  that  migrate 
across  roads; 

(4)  biases  samples  because  it  is  limited  to  areas 
with  roads;  and 

(5)  is  sometimes  dangerous  to  observers,  espe- 
cially on  well-traveled  routes. 


Tortoises  (especially  desert  tortoises)  and  other 
turtles  are  often  verified  by  systematic  searches. 
Burge  (1979)  used  a  series  of  1.6-km  and  4.8-km  (1- 
mi.  and  3-mi. )  search  transects  in  the  Arizona  Sono- 
ran  Desert  to  verify  occurrence  of  desert  tortoises. 
The  4.8-km  (  3-mi. )  transect  consisted  of  a  triangle 
(1.6  km  [1  mi.]  each  side)  set  on  a  random  compass 
line.  Along  each  line,  observers  look  for  live  tor- 
toises, scat,  burrows,  and  tracks.  Data  are  then  re- 
corded on  standard  forms. 

Systematic  searches  of  defined  areas  yield  spe- 
cies occurrence  data  and  allow  for  further  assess- 
ment of  populations  of  amphibians  and  reptiles.  This 
method  is  especially  valuable  when  the  biologist 
needs  population  estimates  of  special  status  amphibi- 
ans and  reptiles.  Bury  (1982)  used  2-ha  (4.9-a.) 
quadrats  to  verify  and  study  diurnal  reptile  popula- 
tions in  the  Mojave  Desert.  He  used  a  removal 
method  to  determine  densities  of  reptiles  on  his 
quadrats  (discussed  in  more  detail  later  in  this  chap- 
ter). Bury  and  Luckenbach  ( 1977)  successfully  cen- 
sused  tortoise  populations  with  a  quadrat  and  grid 
location  system.  They  established  25-  to  100-ha  (62- 
to  247-a. )  quadrats,  subdividing  each  into  1-ha  (2.47- 
a. )  sections.  Systematic  searches  of  each  section  ac- 
curately located  tortoises  and  served  as  permanent 
sites  for  monitoring  tortoise  populations.  Schneider 

(1981)  used  2.59-km    (1-mi.  )  plots  to  collect  popu- 
lation data  on  desert  tortoises  in  Arizona.  Both  Bury 

( 1982)  and  Schneider  ( 1981 )  used  systematic 
searches  of  these  grids  to  determine  occurrence  and 
population  structure. 

Systematic  search  procedures  in  defined  areas 
are  very  time-consuming  and  should  only  be  used  if 
the  biologist  needs  an  accurate  estimate  of  popula- 
tion density.  Generally  this  involves  only  special 
status  species  such  as  federally  threatened  and  en- 
dangered amphibians  and  reptiles  where  density  data 
are  needed  to  assess  populations. 

Another  systematic  search  method  commonly 
used  involves  scouring  known  habitats  of  certain 
amphibians  and  reptiles.  For  example,  canyon  tree 
frogs  (Hyla  sp.)  are  most  easily  observed  in  rocky, 
boulder-strewn  canyons  with  permanent  water,  pri- 
marily pools  (Jones  1981b).  Secretive  diurnal  and 
nocturnal  species  can  be  verified  along  systematic 
search  routes  by  lifting  rocks,  vegetative  debris,  and 
uprooting  animal  shelters  such  as  pack  rat  dens. 
Zweifel  and  Lowe  ( 1966)  obtained  desert  night  liz- 
ard specimens  (Xantusia  vigilis)  by  lifting  downed 
Joshua  tree  (Yucca  brevifolia)  limbs.  Without  search 
of  their  preferred  habitat  (surface  debris),  these  liz- 
ards cannot  be  verified.  Amphibians  and  reptiles  also 
use  a  variety  of  other  animal  cover  sites;  therefore, 
a  large  number  of  amphibians  and  reptiles  can  be 
verified  by  searching  mammal  dens  and  burrows 
(Lee  1968). 


Amphibians  and  Reptiles 


275 


These  microhabitat  specific  searches  are  gener- 
ally used  to  verify  and  collect  data  on  the  abundance 
of  a  few  species.  If  used  for  verification  only,  this 
method  can  be  quick  and  easy.  When  relative  abun- 
dance or  density  is  needed,  this  method  is  consider- 
ably more  time-consuming. 

There  are  several  small  tools  that  increase  effec- 
tiveness of  searches.  Mirrors,  in  particular,  locate 
and  identify  animals  in  burrows,  especially  turtles 
and  snakes.  Similarly,  flashlights  increase  light  in  and 
around  burrows  and  are  extremely  helpful  in  verify- 
ing species  such  as  tortoises. 

Several  types  of  poles  have  been  used  to  locate 
inactive,  partially,  or  completely  concealed  reptiles, 
especially  turtles.  Noodling  is  an  effective  way  to 
locate  aquatic  turtles  concealed  in  muddy  creek 
bottoms  (Lagler  1943).  The  technique  involves 
cruising  shallow  lakes  and  ponds  using  a  blunt  steel 
rod  to  protrude  through  mud  and  emergent  vegeta- 
tion. A  hollow  sound  is  given  off  when  a  turtle  is  hit. 
Noodling  can  also  be  used  to  protrude  through  ter- 
restrial litter  and  soil  to  verify  hidden  turtles 
(Carpenter  1955). 

Another  extremely  effective  method  of  deter- 
mining species  occurrence  in  an  area  is  to  search 
nests  of  predators  for  amphibian  and  reptile  remains. 
Large  proportions  of  certain  raptor  diets  consist  of 
reptiles  (Millsap  1981).  A  total  of  13  species  of  rep- 
tiles were  verified  in  and  below  red-tailed  hawks 
(Buteo  jamaicensis)  nests  by  Millsap  (1981).  Red- 
tailed  hawks  forage  over  relatively  large  areas  and, 
therefore,  reptile  occurrence  in  nests  cannot  be  cor- 
related with  specific  areas  or  habitats.  However, 
some  raptors  forage  over  extremely  limited  areas  or 
in  specific  habitats  (e.g.,  riparian  and  aquatic  habi- 
tats). For  example,  black  hawks  are  highly  dependent 
on  aquatic  prey  within  western  Arizona  (over  90% 
of  their  diets  are  aquatic  organisms).  Millsap  and 
Harrison  ( 1981 )  identified  six  species  of  reptiles  at 
black  hawk  nests,  including  a  mud  turtle  {Kinoster- 
non  sp. ).  Generally  raptor  nests  cannot  be  used  to 
identify  amphibian  species  because  entire  animals 
are  usually  consumed.  However,  prey  items  can  be 
identified  as  they  come  into  raptor  nests. 

Owl  pellet  dissections  also  verify  amphibian  and 
reptile  occurrence  in  an  area  (B.A.  Millsap,  pers. 
commun. ).  Positive  identifications  are  obtained  by 
comparing  samples  with  skeletal  references. 

Although  examination  of  predator  nests  and 
feces  provides  species  occurrence  for  a  region,  it 
does  not  consistently  indicate  relative  abundance  of 
amphibians  and  reptiles  because  predators  eat  what 
they  prefer  and  not  necessarily  what  is  most 
abundant. 


Black  hawk  feeding  nestling. 


Trapping  and  Collecting  Methods 

Trapping  methods  have  been  designed  around 
specific  life  histories  of  certain  amphibians  and  rep- 
tiles. For  example,  a  trout  fly  was  extended  over  a 
pitfall  trap  to  lure  zebra-tailed  and  fringe-toed  lizards 
(Uma  sp.)  into  traps  (Lannon  1962).  Both  of  these 
lizards  are  sit-and-wait  species  that  forage  out  in 
open  spaces  (see  Pianka  1966).  They  remain  rela- 
tively stationary  until  visual  contact  is  made  with  an 
invertebrate  and  then  attempt  to  capture  and  swal- 
low the  prey.  Although  Lannon's  method  may  attract 
lizards  that  forage  in  similar  ways  on  similar  prey,  it 
will  not  attract  other  amphibians  and  reptiles  that 
forage  in  different  ways. 

Two  basic  procedures  exist  for  collecting  or 
capturing  amphibians  and  reptiles:  direct  and  indi- 
rect capture.  Direct  capture  consists  of  actively  seek- 
ing animals  and  using  various  equipment  to  increase 
capture  success.  Indirect  capture  involves  use  of 
equipment  capable  of  securing  animals  without  ac- 
tive searching  by  the  biologist. 

Several  types  of  equipment  are  used  to  capture 
animals  during  active  pursuit.  I  will  highlight  some 
of  the  most  commonly  used  equipment. 

Large  samples  of  frogs,  toads,  salamanders,  tad- 
poles, and  turtles  can  be  obtained  because  their  hab- 
itats (water  and  water/land  ecotones)  are  limited  to 
relatively  small  areas  (as  compared  with  terrestrial 
habitats). 

Dip  nets  increase  effectiveness  of  capturing  tur- 
tles, frogs,  toads,  and  salamanders  during  active  pur- 
suit or  while  turning  stones  and  logs  in  streams. 
Svihla  (1959)  greatly  increased  capture  of  tailed  frog 
tadpoles  (Ascaphns  truei)  by  simultaneously  placing 
a  dip  net  directly  downstream  while  turning  rocks. 
Without  the  net,  many  tadpoles  escaped  unnoticed. 


276 


Amphibians  and  Reptiles 


Lagler  (1943)  also  described  dip  nets  of  various  sizes 
used  to  collect  small  tadpoles,  salamander  larvae, 
and  small  turtles. 


Gunning  and  Lewis  (1959)  reported  low  mortality 
rates  when  using  shockers  that  generated  direct  cur- 
rent fields. 


Seines,  placed  across  a  stream  or  creek,  greatly 
increase  capture  success  of  salamander  and  salaman- 
der larvae,  and  frog  and  toad  tadpoles  (Balgooyen 
1977).  With  a  seine  placed  downstream,  biologists 
move  toward  the  net,  turning  rocks,  debris,  and  veg- 
etation. Disturbed  animals  are  then  swept  down- 
stream into  the  net  by  the  stream's  current. 

Stationary,  polyethylene  plastic  dams  can  be 
used  to  corral  stream-dwelling  organisms  including 
shovel-nosed  salamander  larvae  ( Leurognathus  sp. ) 
(Martof  1963)-  The  damming  effect  causes  larval 
salamanders  to  congregate  at  the  bottom  of  the  plas- 
tic sheet. 

Electroshocking  is  an  effective  means  of  collect- 
ing and  censusing  aquatic  habitats.  Although  more 
traditionally  used  for  collecting  fish,  direct  current 
electroshockers  have  been  used  to  census  aquatic 
amphibians  and  reptiles  (Gunning  and  Lewis  1959; 
Williams  et  al.  1981 ).  Electroshocking  stuns  adult 
and  larval  salamanders,  adult  and  larval  frogs,  and  to 
a  lesser  degree,  small  turtles,  making  capture  easier. 
Several  combinations  of  aquatic  collecting  tech- 
niques were  tested  on  hellbenders  (Cryptobranchus 
alleganiensis).  Combinations  of  electroshocking 
and  dip  nets  were  most  effective  although  results 
varied  in  different  aquatic  habitats.  Electroshocking 
and  dip  nets  were  most  successful  in  pools  and  slow- 
moving  stream  sections  because  dip  nets  capture 
hellbenders  before  they  sink.  Conversely,  seines  cap- 
tured hellbenders  more  effectively  in  fast-moving, 
riffle-dominated  stream  sections  because  animals  are 
more  easily  swept  into  a  seine  than  into  a  dip  net. 
Williams  et  al.  ( 1981 )  found  conventional  dip-net- 
ting and  seining  to  be  inferior  to  netting  with  elec- 
troshocking. Both  Williams  et  al.  (  1981 )  and 


Abundance  and  density  of  aquatic  amphibians 
can  be  determined  for  a  reach  of  stream  by  using  the 
previously  described  equipment.  First,  an  area  of  the 
stream  is  selected  and  measured  in  square  meters. 
Second,  a  seine  is  stretched  across  the  stream  at  the 
station's  farthest  point  downstream.  Finally,  the  biol- 
ogist systematically  walks  downstream,  swooping 
the  electrical  field  across  the  stream.  Shocked  am- 
phibians are  then  placed  in  the  net  and  counted.  A 
second  swoop  is  conducted  to  shock  the  remaining 
amphibians.  The  biologist  then  obtains  a  number 
of  each  species  for  the  area  shocked. 

Seining  by  itself  provides  verification  of  species, 
but  is  too  slow  to  provide  an  absolute  count  (unless 
sampling  a  small  isolated  pool).  Electroshocking, 
although  more  effective,  is  generally  not  effective  on 
large  species  with  small  surface  to  volume  ratios. 
As  a  quantitative  method,  electroshocking  is  limited 
to  relatively  narrow,  shallow,  non-turbid  reaches. 

Anderson  and  Smith  (1950)  designed  and  tested 
an  electrical  apparatus  that  effectively  stunned  both 
aquatic  and  terrestrial  amphibians  and  reptiles  (Fig- 
ure 1 ).  Mortality  rates  of  shocked  animals  varied, 
usually  with  size.  Large  snakes  such  as  water  mocca- 
sins (Agkistrodon  sp. )  must  be  shocked  on  the  up- 
per third  of  their  body  to  obtain  paralysis,  whereas 
single  shocks  to  small  amphibians  (e.g.,  tree  frogs) 
are  often  fatal. 

In  water  with  thick  vegetation  and  organic  de- 
bris, dredges  can  increase  collecting  success.  Goin 
( 1942)  used  a  double-trough  dredge  held  together 
by  hinges  (Figure  2).  The  upper  trough  is  con- 
structed of  larger  mesh  hardware  cloth  than  the 


ELECTRODES 


kmd 


mm 

BATTERIES 


Figure  1.     Electrical  stunning  apparatus. 


SWITCH 


S^MI  — 


MMm 


m 


aSwstsS 

lli 


COIL 


PLUG 


Amphibians  and  Reptiles 


277 


Figure  2.     Double-trough  dredge. 

lower  trough.  Vegetation-dwelling  larvae  and  adult 
salamanders,  tadpoles,  frogs,  and  turtles  are  captured 
when  the  dredge  is  lifted  up  through  vegetation 
and  debris.  Different  sized  animals  can  be  collected 
by  varying  the  size  of  the  screen  mesh. 

Other  methods,  such  as  gaffing  and  setting  lines 
of  baited  fish  hooks,  have  been  used  to  verify  aquatic 
turtles  (Lagler  1943).  A  gaff  consists  of  a  sharp  hook 
placed  on  the  end  of  a  pole.  Turtles  are  caught  by 
spearing  the  hook  through  the  shell.  This  method,  al- 
though not  used  as  a  quantitative  procedure,  pro- 
duces records  of  large  turtles. 

A  snare  has  been  designed  and  successfully 
tested  to  capture  water  snakes  (Franklin  1947).  The 
snare  consists  of  a  three -prong  sterilizing  forcep 
attached  by  a  scissor  grip  to  the  end  of  a  3.7-m  (12- 
ft)  lightweight  cane  pole.  The  snare  is  particularly 
effective  in  capturing  large,  swimming  water  snakes 
(e.g.  water  moccasin)  and  provides  a  means  of  veri- 
fying certain  species. 

An  alternative  to  collecting  amphibians  and  rep- 
tiles with  a  gun  is  to  use  a  rubberband  mechanism. 
This  type  of  collecting  method  is  considerably  safer 
than  a  firearm  and,  with  practice,  yields  live  speci- 
mens. Neill  (1956)  and  Dundee  (1950)  designed 
effective  rubberband-launching  devices  that  they  re- 
ported using  successfully  to  immobilize  frogs  and 
lizards  (see  Figure  3). 

Hand  snares  are  used  to  secure  reptiles,  espe- 
cially lizards.  Eakin  (  1957)  designed  a  copper  wire 


Figure  3.     Rubberband-stunning  devices. 


slip  noose  (Am.  Stand.  Wire,  gauge  34,  [0.15  mm]) 
attached  to  a  light  pole.  Lizards  are  captured  by 
slowly  placing  the  noose  over  their  heads  and  jerk- 
ing the  wire  tight.  Stickel  (1944)  designed  and  suc- 
cessfully tested  a  snare  that  has  a  trigger  mechanism 
and  a  thread  (Figure  4).  He  stated  that  his  mecha- 
nism is  superior  to  hand-pulled  snares  because  of  the 
speed  of  the  snare  mechanism.  Fishing  line  (0.91- 
1.36  kg  [2-3  lb.]  test)  can  also  be  used  to  form  the 
noose  (R.  Bowker,  pers.  commun.). 

All  of  the  previously  described  equipment  such 
as  snares,  nooses,  and  dredges  help  biologists  verify 
occurrence  of  species  on  a  site  or  in  a  habitat.  Rarely 
is  this  equipment  used  to  determine  relative  abun- 
dances or  densities  unless  they  are  part  of  a  system- 
atic search  or  procedure. 

Perhaps  the  most  widely  and  successfully  used 
amphibian  and  reptile  collecting  and  censusing  tech- 
niques are  those  that  involve  indirect  capture,  specif- 
ically traps  and  funnels.  Traps  and  funnels  have  the 
advantage  of  capturing  animals  while  the  biologist  is 
not  present.  Generally  they  provide  a  greater  num- 
ber of  species  per  sampling  effort  than  active,  direct 
capture  methods.  Funnels  and  traps  are  used  widely 
for  both  terrestrial  and  aquatic  sampling.  Most  trap 
and  funnel  procedures  are  designed  to  maximize 
capture  of  certain  types  of  herpetofauna.  Their  place- 
ment in  certain  microhabitats  and  the  use  of  differ- 
ent baits  also  bias  sampling  toward  certain  species. 

Several  investigators  successfully  designed  and 
tested  hoop  funnel  traps  for  capturing  turtles  (Lagler 
1943,  I960;  Pirnie  1935;  Iverson  1979;  Feuer  1980). 


278 


Amphibians  and  Reptiles 


TRIGGER  STICK 


RUBBER   BAND 
.TRIGGER   NAIL 


Figure  4.     Trigger  snare. 

Generally,  traps  consist  of  mesh  netting  connected 
to  four  aluminum  hoops  or  guide  lines  (Figure  5).  A 
purse-string  regulated  opening  provides  turtles  with 
access  through  the  throat  to  a  piece  of  suspended 
bait.  Funnels  can  have  single  or  double  throats.  Iver- 
son's  (  1979)  funnel  trap  is  constructed  of  chicken 
wire  rather  than  netting.  These  funnel  traps  are  most 
effective  in  shallow  water  and  should  be  checked 
frequently  to  avoid  drowning  the  specimen. 

Funnel  traps  have  also  been  designed  to  capture 
larvae  and  adult  salamanders  and  frogs  (Carpenter 
1953;  Moulton  1954).  Carpenter's  trap  consists  of  a 


6.35-mm  (0.25-in.)  mesh  hardware  cloth  formed  into 
a  funnel  that  is  lowered  into  deep  pools  by  a  1 .8-  to 
3.0-m  (6-  to  10-ft)  rope.  This  method  collects  larval 
and  adult  salamanders  in  pools  too  deep  for  seines. 
Funnel  openings  should  face  toward  shores  of  ponds 
to  take  advantage  of  animal  movements. 

Moulton's  trap  is  slightly  different  than  Carpen- 
ter's, consisting  of  a  rectangular  wire  screen  box 
placed  on  a  wooden  frame  with  an  access  funnel 
placed  63.5  cm  (25  in.)  above  the  box  floor.  The 
trap  is  placed  in  a  shallow  depression  along  the 
shore  of  a  pond  or  lake  with  the  funnel  facing  to- 
ward land.  Salamander  and  anuran  prey  accumulate 
in  the  trap  and  lure  animals  toward  the  trap  during 
movements  between  land  and  water. 

Funnel  traps  have  been  used  extensively  in  cap- 
turing terrestrial  and  semiaquatic  frogs,  salamanders, 
lizards,  and  snakes.  Generally  these  traps  function 
similarly  to  aquatic  funnel  traps;  they  encourage 
entrance  into  traps  and  restrict  access  out.  These 
traps  are  usually  constructed  of  hardware  cloth, 
shaped  into  cylindrical  form,  and  accessed  at  one  or 
both  ends  by  funnels  (Vogt  1941;  Lagler  1943). 
Many  minor  variations  of  this  general  construction 
have  been  used.  Fitch  (1951)  used  a  transparent, 
cellulose  pivotal  door  at  the  entrance  to  the  trap, 
preventing  lizards  and  snakes  from  crawling  out  (Fig- 
ure 6).  Clark  (1966)  and  Hall  (1967)  further  modi- 
fied the  trap  by  using  mason  jars  and  0.45-kg  (1-lb) 
coffee  cans,  respectively,  as  the  collecting  area.  Both 
used  hardware  cloth  to  funnel  animals  into  the  trap 
(Figure  7).  Many  of  these  investigators  used  bait 
to  attract  animals. 

Many  investigators  have  also  combined  funnel 
trapping  with  other  techniques,  especially  use  of 
drift  fence.  Dargan  and  Stickel  (1949)  used  a  0.1 0-m 
(4-in.)  high,  7.6-m  (25-ft)  long  hardware  cloth  drift 
fence  to  guide  snakes  into  hardware  cloth  funnel 


*'  ***• 


Figure  5.     Standard  funnel  trap  for  capturing  turtles. 


Drift  fence  reptile  trap. 


Amphibians  and  Reptiles 


279 


traps.  Clark  (  1966)  found  that  rolled  aluminum 
(0.019  gauge)  and  board  drift  fences  (2.4  to  3.0  m 
[8  to  10  ft])  used  with  funnel  traps  increased  trap- 
ping success  of  small  snakes.  Migrating  salamanders 
and  frogs  can  be  surveyed  by  completely  enclosing 
ponds  with  6.35  mm  (1/4  in.),  0.25-m  (104n.)  high 
hardware  cloth,  placing  hardware  cloth  and  bronze 
screen  funnel  traps  inside  and  outside  the  enclosure 
along  the  fence  (Storm  and  Pimentel  1954). 

Pitfall  trapping  is  one  of  the  most  widely  used 
methods  for  collecting  data  on  amphibians  and  rep- 
tiles. Generally  this  method  involves  placing  a  square 
or  round  container  below  water  or  ground  with  the 
container  top  flush  with  the  surface.  Container  sizes 
and  shapes  vary  greatly,  depending  on  which  species 
is  trapped. 

Floating  pitfall  traps  can  be  used  in  water  for 
capturing  turtles.  Lagler  (1943),  Breen  (1949),  and 
Petokas  and  Alexander  (1979)  successfully  used 
these  traps  for  turtles  in  ponds  and  lakes.  Lagler 
( 1943)  and  Petokas  and  Alexander  ( 1979)  used  a 
floating  wood  frame  with  a  submerged  trap  to  cap- 
ture basking  turtles  (Figure  8).  Breen  (1949)  used  a 
submerged  wooden  barrel  with  the  top  above  the 
water  surface.  Wooden  planks  were  extended  from 
the  barrel  top  to  the  water  with  bait  placed  at  the 
top  of  each  plank.  Turtles  fell  into  the  barrel  while 
attempting  to  swallow  bait.  Pitfall  traps  capture 
mostly  basking  turtles  that  spend  time  on  or  above 
the  water's  surface. 


Pitfall  traps  have  been  used  extensively  in  ter- 
restrial amphibian  and  reptile  surveys  and  studies. 
Generally  pitfall  traps  are  various  sized  containers 
placed  in  the  ground  with  tops  flush  to  the  surface. 
There  are  a  number  of  variations  in  pitfall  traps  rang- 
ing from  wooden  box  traps  to  18.9-L  (5-gal.)  plastic 
containers. 

Rogers  (1939)  used  a  trapdoor  box-trap  buried 
with  the  top  flush  to  the  surface  to  capture  lizards. 
Banta  (1957)  used  4.7-L  (5-qt)  tin  cans  and  covered 
tops  (cardboard)  leaving  a  few  inches  between  tops 
and  pitfall  openings  (Figure  9).  Lizards  fell  into  traps 
while  attempting  to  seek  shade.  Whitaker  (1967) 
increased  pitfall  trapping  success  of  some  lizards  by 
using  fruit  and  honey  bait.  Besides  providing  occur- 
rence data,  pitfall  traps  obtain  specimens  for  popula- 
tion, reproductive,  and  taxonomic  studies.  Porzer 

(1982)  used  double-depth  18.9-L  (5-gal.)  plastic 
containers  to  capture  Gila  monsters  (Heloderma  sus- 
pectum)  for  radio  telemetry  studies. 

When  arranged  systematically  and  standardized 
per  unit  effort  (e.g.,  animals  caught  in  a  24-hour 
trapping  period),  pitfall  and  other  trapping  data  can 
be  quantified  and  compared.  For  example,  biologists 
can  compare  relative  abundance  of  certain  species 
between  two  or  more  habitat  types  if  they  set  out 
traps  in  a  standardized  way  (similar  arrangement  and 
size)  during  the  same  season.  Bury  and  Raphael 

(1983)  described  a  systematic  pitfall  trapping  system 


Figure  6.     Funnel  trap  for  capturing  terrestrial  amphibians  and  reptiles. 
280  Amphibians  and  Reptiles 


Figure  7.    Jar  funnel  trap. 

consisting  of  7.6-L  (2  gal.)  sized  pitfall  traps,  ar- 
ranged in  a  5-  x  13-cm  (2-  x  5-in. )  grid,  20  m  (66  ft) 
apart.  This  procedure  captured  seven  times  the  num- 
ber of  lizards  than  a  technique  of  searching  a  prede- 
fined area,  lifting  logs  and  rocks  (previously 
described  as  time-constraint  collecting),  but  pro- 
duced only  one-half  the  number  of  salamanders.  K.B. 
Jones  (unpubl.  data)  and  Price  (pers.  commun. ) 
used  a  large  (26-  x  30-cm  [10-  x  12-in.]  trap  grid) 
2-ha  (5-a.)  pitfall  grid  to  determine  relative  abun- 
dance of  lizards  in  riparian  and  Chihuahuan  Desert 
habitats,  respectively  (Figure  10).  Both  used  double- 
deep,  1.35-kg  (3-lb)  coffee  cans  spaced  15  m  (49.5 
ft)  apart  with  wooden  tops  placed  15.2-20.3  cm  (6-8 
in.)  over  the  traps  for  shade.  Jones  and  Price 
(unnubl.  data)  accumulated  large  samples  of  most 
lizards  in  a  period  of  less  than  a  month  during  the 
spring. 


By  marking  amphibians  and  reptiles  for  individ- 
ual identification  while  using  a  pitfall  grid,  biologists 
can  determine  species  home  ranges  and  population 
sizes  (mark  and  recapture). 


Drift  fences  have  been  widely  used  in  combina- 
tion with  pitfall  traps,  especially  in  recent  years. 


Figure  8.     Floating  pitfall  trap  for  capturing  turtles. 


Gibbons  and  Semlitsch  (1981 )  used  several  configu- 
rations of  pitfall  traps  and  drift  fences  for  capturing 
amphibians,  reptiles,  and  small  mammals.  They  found 
these  methods  yielded  a  large  number  of  species 
and  accurately  estimated  the  abundances  of  readily 
trapped  species.  They  also  provide  time  and  cost 
estimates  for  trapping.  Campbell  and  Christman 
(1982)  described  an  array-trapping  method  consist- 
ing of  eight  18.9-L  (5-gal.)  plastic  containers,  7.6- 
m  (25 -ft )  drift  fences,  and  hardware  cloth  funnel 
traps  ( Figure  11).  The  array  technique  yielded  a 
large  number  of  amphibian  and  reptile  species,  espe- 
cially secretive  species  not  normally  verified  by 
other  techniques.  Jones  (1981a  &  b)  modified  Camp- 
bell's and  Christman's  (1982)  array  system  by  using 
only  three  drift  fences  and  four  18.9-L  (5-gal.)  plastic 
containers  (Figure  12).  As  in  Campbell  and  Christ- 
man's  ( 1 982 )  studies,  Jones  verified  a  large  number 
of  lizards  and  snakes,  especially  secretive  forms,  but 
with  about  one-half  the  equipment  at  each  site. 


Amphibians  and  Reptiles 


281 


Figure  9.     Simple  pitfall  trap  for  terrestrial 
amphibians  and  reptiles. 


Vogt  and  Hine  (1982)  also  used  a  wide  variety 
of  pitfall/drift  fence  configurations  (Figure  13)  to 
sample  amphibians  and  reptiles.  These  configurations 
were  designed  to  maximize  trapping  success  in  each 
habitat.  Although  Vogt  and  Hine  (1982)  reported 
drift  fence  lengths  of  less  than  15  m  (49. 5  ft)  as 
generally  ineffective,  Jones  (1981a  &  b)  demon- 
strated that  a  7.6-m  (25-ft)  fence  length  was  very 
effective.  These  differences  may  reflect  different 
types  of  habitat  and  animals  sampled  by  the  two  in- 
vestigators (wet  woodland  versus  desert). 

The  sampling  duration  and  seasons  when  array 
or  pitfall/drift  fence  trapping  is  most  successful  de- 
pends on  general  climatic  features  of  an  area,  and 
the  species  and  habitats  to  be  sampled.  For  example, 
Vogt  and  Hine  (1982)  suggested  that  several  short 
sampling  periods,  staggered  over  an  activity  season 
(especially  after  rains),  yield  the  most  accurate  esti- 
mate of  species  composition  and  species  abundance. 
They  believed  this  procedure  takes  advantage  of 
increased  herpetofauna  activity  resulting  from  rain. 
Unlike  Wisconsin,  the  Desert  Southwest  receives 
infrequent  precipitation  and  has  a  far  greater  number 
of  habitats  where  temperature  and  moisture  availabil- 
ity vary.  In  Arizona,  desert  habitats  are  most  effec- 
tively trapped  in  April  and  May,  and  forest  and 
woodland  habitats  between  May  and  July.  I  have  also 
found  that  broken,  short-duration  sampling  suggested 
by  Vogt  and  Hine  (1982)  does  not  verify  all  species, 


v 

/  II   buckets 

o 

o 

o 

o     o     o     o 

o 

o 

o 

o     o     o 

o 

o 

o 

Grid    area  =  1  ha 

\1/ 

110  buckets 

10  CK 

o 

>  1  5m 

o 

oJ 

Double-deep  31b. 
coffee  can 

tape 


Figure  10.     Grid  pitfall  trapping  system. 


especially  many  secretive  desert  snakes,  present  in 
a  habitat.  I  attribute  this  finding  to  infrequent,  erratic 
rainfall  and  suggest  sample  periods  of  not  less  than 
30  days. 

To  obtain  differences  in  species  composition 
and  abundance,  biologists  should  place  arrays  in 
different  habitats,  provided  that  a  standard  configura- 
tion is  used.  Fences  located  so  that  amphibian  and 
reptile  migration  routes  are  dissected  will  increase 
capture  success.  Jones  (1981a  &  b)  used  arrays  for 
assessing  lizard  and  small  snake  relative  abundances 
in  a  variety  of  habitats  in  western  Arizona.  Array 
abundance  data  can  be  expressed  as  the  number  of 
animals  caught  in  24  hours  (Jones  1981a  &  b). 

Pitfall/drift  fence  or  array  trapping  most  accu- 
rately assesses  lizard  and  small  snake  composition 


282 


Amphibians  and  Reptiles 


and  abundance,  but  is  less  effective  in  estimating 
salamander,  frog,  and  toad  abundances,  although  it 
does  provide  some  excellent  records  for  these  her- 
petofauna.  This  method  is  not  recommended  for 
assessment  of  most  terrestrial  turtles  and  large 
snakes.  There  is  some  amphibian  and  reptile  mortal- 
ity, especially  salamanders,  frogs,  and  toads,  associ- 
ated with  most  trapping  methods  previously  listed. 
To  reduce  mortality,  biologists  should  check  traps  as 
often  as  possible,  generally  every  other  day.  Mortal- 
ity due  to  exposure  can  be  reduced  by  placing  traps 
in  shaded  areas,  by  covering  trap  openings  with 
shades,  or  by  placing  about  2  to  4  cm  (1  to  2  in. )  of 
soil  or  litter  at  the  bottom  of  the  trap.  These  precau- 
tions will  also  reduce  predation  on  trapped  animals. 

Water  accumulation  in  pitfall  traps,  especially  in 
wet  regions,  can  cause  significant  mortality.  Whereas 
punching  holes  in  the  bottom  of  traps  provides  little 
drainage,  floatable  objects  such  as  styrofoam  reduce 
mortality  due  to  drowning. 


Q 


FENCE 


FUNNEL  TRAP 


6— 


CAN   TRAP 


76M 


O 


-O  I5M O 


-o 


Q 


DO 


6 


Baited  snap  traps,  such  as  those  commonly  used 
to  census  small  mammals,  can  be  used  to  collect 
large  diurnal  lizards.  Heatwole  et  al.  (1964)  trapped 
several  large,  diurnal  lizards  by  baiting  museum  snap 
traps  with  live  insects,  meat,  fruit,  and  peanut  butter. 


Figure  11.     Array  pitfall  trapping  technique,  with 
funnels. 


Table  3.     Comparison  of  mean  relative  abundance  and  species  diversity  for  lizards  of  five  heavily  grazed 
vegetative  communities.  Tests  are  based  on  Student's  t-test  values  at  the  95%  confidence  interval.  Each 
heavily  and  lightly  grazed  site  is  based  on  seven  sample  sites. 


& 

it 

K 


H  =  Heavily  grazed    L  =  Lightly  grazed 


Vegetative  Community 

Mean  Relative 
Abundance  ±  SD 

Species 
Diversity  (H') 

Chaparral 

Grazed 

Array- 
Nights 

Lizards 
Trapped 

1.18  ±  32 
1.69  ±  .41a 

1.02 
1.09 

H 

L 

672 
623 

782 
1050 

Desert  grassland 

H 
L 

294 

224 

179 
176 

0.56  ±  .08 
0.77  ±  .13 

0.88 
1.01 

Mixed  riparian 
scrub 

H 
L 

392 

252 

295 

310 

0.73  ±  0.5 
1.20  ±  .06d 

0.69 
0.90 

Cottonwood-willow 
riparian 

H 
L 

658 
175 

419 
201 

0.64  ±  .23 
1.13  ±  .15a 

0.59 
0.86 

Sonoran  desert 
scrub 

H 
L 

714 
238 

757 
244 

1.06  ±  .10 
1.03  ±  .12 

0.93 
0.93 

aThe  comparison  of  heavily  grazed  with  lightly  grazed  is  significantly  different  at  the  95%  confidence  interval  in  three  vegetative 
communities. 


Amphibians  and  Reptiles 


285 


Large,  man-made  developments  such  as  canals 
act  as  large  traps  and  can  provide  valuable  distribu- 
tional records  of  amphibians  and  reptiles.  Hawken 
( 1951 )  captured  large  numbers  of  amphibians  and 
reptiles  by  a  flume  located  downstream  from  a  lake. 

A  total  of  582  specimens  (15  species)  of  amphibians 
and  reptiles  were  trapped  by  this  man-made  system. 

Because  some  records  obtained  during  trapping 
represent  significant  range  extensions,  biologists 
should  collect  voucher  specimens  of  amphibians  and 
reptiles  taken  out  of  their  published  geographic 
range.  Pisani  (1973)  discussed  field  notes  and  preser- 
vation techniques  for  amphibians  and  reptiles.  After 
preserving  range  extension  specimens,  biologists 
should  contact  regional  experts  to  verify  records. 
These  animals  should  then  be  deposited  in  a  univer- 
sity, regional,  or  national  museum.  For  example, 
range  extension  specimens  collected  by  the  BLM's 
Phoenix  District  were  placed  in  the  National  Mu- 
seum of  Natural  History  at  the  Smithsonian  Institute 
in  Washington,  DC. 


Certain  field  procedures  have  been  developed  to 
identify  amphibians  and  reptiles  where  direct  obser- 
vations are  not  possible.  Many  frogs  and  toads  pro- 
duce specific  calls  that  can  be  recorded  and  used  to 
identify  individual  species  (see  Michaud  1964; 
Brown  and  Littlejohn  1972;  Fouquette  1980  for 
examples). 

Certain  species  of  reptiles,  especially  lizards  and 
snakes,  leave  characteristic  tracks  and  when  com- 
pared to  standards  developed  for  each  species,  can 
be  used  to  identify  species  occurrence  in  an  area. 
Lillywhite  (1982)  used  tracks  left  in  fine-textured 
soils  and  on  roads  to  identify  common  snakes  such 
as  sidewinders  (Crotalus  cerastes)  and  whipsnakes 
{Masticophis  sp. ).  Although  this  procedure  provides 
some  indication  of  species  and  habitat  use,  track 
identification  does  not  provide  accurate  estimates  of 
abundance,  and  it  may  not  be  practical  to  use  be- 
cause of  precipitation  and  wind.  A  biologist  may 
want  to  use  this  method  when  movement  and  habi- 
tat use  data  are  needed  for  priority  species  that  leave 
easily-recognized  tracks. 


bushes 


7.6  M 


drift  fence   placed  to 
cut   off   migrational 
routes  between  vege- 
tation 


Figure  12.     Modified  array  pitfall  trapping  technique. 
284  Amphibians  and  Reptiles 


Species  and  Population  Measurements 

Depending  on  the  specific  objectives  of  a  survey 
or  study,  a  number  of  morphological,  behavioral, 
and  ecological  measurements  can  be  taken.  Only  a 
few  measurements  such  as  general  age  class  and  size 
are  taken  from  opportunistic  and  systematic  observa- 
tions when  animals  are  not  captured. 

Capture  methods  allow  biologists  to  obtain  a 
number  of  measurements  such  as  total  length  (e.g., 
snakes),  snout-vent  length  (lizards),  weight,  sex 
(from  coloration  or  sex  probes),  reproductive  condi- 
tion (swollen  testes  or  coloration  that  indicates 
breeding),  and  limb  and  other  morphological  traits 
(used  to  tie  species  to  habitats)  such  as  jaw  width/ 
length  ratios.  All  of  these  measurements  provide 
biologists  with  indicators  of  species'  conditions,  es- 
pecially when  animals  are  released  and  recaptured. 

Stomach  analysis  of  captured  animals  can  also  be 
obtained,  especially  on  lizards  and  snakes.  Pietruszka 
( 1981 )  found  that  stomach-flushing  devices  provided 
data  on  lizard  prey.  Kephart  and  Arnold  ( 1982)  de- 
termined garter  snake  (Thamnophis  sp. )  diets  by 
running  their  hands  tightly  up  snakes,  forcing 
regurgitation. 

Diet  information  of  lizards  and  snakes  can  be 
used  to  determine  these  animals'  food  needs,  and  it 
also  provides  information  on  the  quantity  and  trends 
of  certain  prey  species  (see  Kephart  and  Arnold 
1982). 

Fecal  analyses  of  certain  herbivorous  reptiles 
can  be  conducted  to  determine  dietary  preferences. 


■  / 


^ /////. 
///_///, 

////ft//. 
^-///^■//. 

nil 

//////'?". 


15m    ALUMINUM    VALLEY 
2) — 3.8m— (in  —  7.4m -(IF)— 3.8m— (2' 


///////////////// 
///////////////// 
///////////////// 

'///////// 
'/////// 


'///? 


22.5m   ALUMINUM    SCREEN 
5j— 3m  — (5)  —  6m (2)  —  9m  —  (  I)  — 4.5m  — (5' 


15m    ALUMINUM   VALLEY 
5) 7.5m (IT) 


15m- 


FUNNEL  TRAP 


FUNNELTRAP 


30m    ALUMINUM    VALLEY 


■7.5m- 


ShUNNtL 
TRAP 
FUNNEL 
TRAP 


15m   ALUMINUM    VALLEY 


Figure  13.     Pitfall/drift  fence  trapping  schemes  in  different  habitats. 


Hansen  et  al.  (1976)  used  a  microhistological  analy- 
sis of  fecal  droppings  to  determine  dietary  prefer- 
ences of  the  desert  tortoise. 

When  individual  amphibians  and  reptiles  are 
marked,  released,  and  recaptured  at  a  later  date,  a 
number  of  statistical  treatments  can  be  used  to  de- 
termine species'  density,  movement,  and  activity 
patterns.  Capture  of  amphibians  and  reptiles  allows 
biologists  to  mark  individual  animals  for  future  iden- 
tification. A  wide  variety  of  marking  techniques, 
ranging  from  tagging  (including  radioactive)  to  se- 
quential toe-clipping,  have  been  used.  Pough  (1970) 
attached  plastic  plug  tags  to  identale  clipped  just 
anterior  to  the  cloaca  (Figure  14).  Shell  notching  is 
perhaps  the  most  common  method  of  marking  tur- 
tles, although  tagging,  branding,  and  painting  are  also 
used  for  individual  identification  (Ferner  1979; 
Figure  14). 

Heckel  and  Roughgarden  (1979)  used  a  paint 
spray  gun  to  mark  lizards.  Although  the  marking 
technique  did  not  provide  individual  lizard  recogni- 
tion necessary  to  determine  movements,  it  provided 
quick  and  accurate  data  necessary  to  determine 
home  range  and  population  size. 


Although  more  commonly  used  for  fish,  removal 
methods  yield  data  that  can  be  submitted  to  certain 
statistical  treatments  to  estimate  amphibian  and  rep- 
tile population  size.  Bury  (1982)  used  a  removal 
method  to  estimate  population  size  of  diurnal  lizards. 
He  verified  all  resident  diurnal  lizards  after  2  days  of 
sampling  and  used  the  total  number  of  lizards  re- 
moved over  the  2-day  period  as  the  population  size. 

Trailing  devices  and  radiotelemetry  provide  data 
on  individual  reptile  movement  and  activity.  Scott 
and  Dobie  (1980)  used  an  aluminum-canister, 
thread-trailing  device  to  determine  movements  of 
turtles  (Figure  15). 

All  of  the  previously  described  procedures  and 
measurements  can  be  used  by  biologists  to  deter- 
mine the  relative  health  of  amphibians  and  reptile 
species  and  communities.  Most  methods  involved 
with  marking,  tracking,  and  recapture  are  time-con- 
suming and,  in  some  instances,  expensive  (e.g.,  te- 
lemetry). These  methods  should  be  used  when 
detailed  information  on  the  condition  of  a  species 
(e.g.,  threatened  or  endangered)  or  a  group  of  spe- 
cies in  a  specific  habitat  (e.g.,  riparian)  is  needed. 


Amphibians  and  Reptiles 


285 


*   Js 


Gila  monster  leaving  tracks  in  the  sand. 


Community  Surveys 

Broader,  more  extensive  surveys  of  entire  am- 
phibian and  reptile  faunas  in  a  variety  of  habitat 
types  do  not  allow  for  intensive  study  of  individual 
species.  In  a  community  survey,  biologists  attempt  to 
determine  species  composition  of  each  major  habitat 
type;  some  rough  estimate  of  abundance;  and,  if  pos- 
sible, species'  uses  of  microhabitats. 


Because  individual  amphibian  and  reptile  mor- 
phology, behavior,  and  ecology  vary,  biologists 
should  use  several  censusing  methods  for  determin- 
ing herpetofauna  community  composition.  Campbell 
and  Christman  (1982)  and  Gibbons  and  Semlitsch 
(1981)  emphasized  the  need  to  use  combinations  of 
opportunistic  observations,  transects,  pitfall  and  fun- 
nel traps,  drift  fence,  and  road  riding  for  deriving 
complete  species  lists  (occurrences)  within  specific 
areas  or  habitats.  Jones  (1981b)  used  several  meth- 
ods to  determine  amphibian  and  reptile  composition 
and  abundance  in  14  major  habitat  types.  Conducted 
during  spring,  summer,  and  fall  of  a  year,  the  survey 
generated  27  significant  range  extensions  for  the 
region.  Information  gained  from  the  survey  also  con- 
tributed heavily  to  management  decisions  made  for 
the  area.  Bury  and  Raphael  ( 1983)  recommend  a 
combination  of  time-constraint  and  pitfall  array  tech- 
niques for  generating  data  useful  in  making  most 
types  of  management  decisions. 


Herpetofauna  richness  and  abundance  data  can 
help  determine  the  effects  of  land  use  on  these  ani- 
mals. Busack  and  Bury  (1974)  and  Jones  (1981a) 
provided  examples  of  studies  describing  land-use  im- 
pacts, including  grazing  effects  on  lizards.  Data  such 
as  in  Table  3  can  be  used  to  assess  impacts  of  land 
use. 

There  are  other  statistical  analyses  that  assess 
amphibian  and  reptile  community  structure.  These 
analyses,  which  include  Horn's  Overlap  Index  (Horn 
1966)  and  several  similarity  coefficients,  are  treated 
in  a  later  chapter. 


DISCUSSION 

The  primary  difference  between  inventory  and 
monitoring  of  amphibians  and  reptiles  and  their  habi- 
tat is  the  objective  established  by  the  biologist.  An 
inventory  and  monitoring  study  may  involve  similar 
data  collection  methods.  However,  inventories  usu- 
ally verify  what  is  there  and  how  habitat  resources 
are  being  used,  whereas  monitoring  determines  how 
individual  species  or  communities  change  as  a  result 
of  specific  types  of  land  use.  An  inventory  can  suc- 
cessfully verify  most  amphibians  and  reptiles  in  rep- 
resentative habitat  types  within  a  year,  provided 
that — 

( 1 )    there  are  multiple  samples  in  each  habitat; 

(  2 )    samples  are  taken  during  all  peak  activity 
periods;  and 

(3)    a  variety  of  methods  are  used. 

If  funds  prohibit  complete  community  samples,  the 
biologist  must  then  decide  which  species  or  habitats 
will  produce  the  most  useful  information  for  the 
money. 

Monitoring  generally  requires  sampling  over 
several  years  so  that  species  and  community  health 
can  be  accurately  estimated.  This  is  especially 
needed  in  sampling  amphibians  and  reptiles  because 
populations  fluctuate  greatly  from  year  to  year  with 
environmental  changes,  particularly  precipitation. 
Multiyear  data  collection  allows  the  biologist  to  de- 
termine which  population  trends  are  due  to  naturally 
fluctuating  environmental  conditions  and  which  ones 
are  due  to  land-use  practices. 

Most  agency  budgets  will  not  permit  long-term, 
intensive,  multiyear  samples  of  individual  species 
or  entire  communities.  To  offset  budget  limitations, 
biologists  should  concentrate  on  long-term  changes 
in  species  richness  and  important  microhabitats, 
especially  when  losses  are  involved.  To  determine 
these  changes,  the  biologist  does  not  need  to  use 
expensive,  time-consuming  species  trend  techniques, 


286 


Amphibians  and  Reptiles 


1000 
900 


900 


t==l 


No. 718 


Figure  14.     Marking  techniques  for  amphibians  and  reptiles.  A  &  B.  Turtles  C.  Snakes  D.  Lizards 


Anterior  anchor  wire 


Film      cam  ster 


Canister  lid  (inside 
view) 


Eye  ring 


Washer 
Machine  screw 

Machine  screw  nut 

Film      canister 


ter    lid 


Machine      Posterior  Spool   and   thread 

screw  anchor  (inside   canister) 

wires 


^    Washe  r 

Cani  ster     lid           //' 

fS^ 

JX\      Machine    screw 

Posterior       /\\ 

Jl\        Posterior 

anchor  — _/       x; 

;//    \-—'  anchor 

wire         / 

^\  \           wire 

Anterior 

Film    canister 

anchor  ' 

wire 

Figure  15.     Trailing  device  to  determine  turtle 
movement. 


Short-term  method  of  marking  desert  tortoises. 


Amphibians  and  Reptiles 


287 


but  rather  species  verification  and  habitat  measure- 
ment techniques.  Data  can  also  be  obtained  intermit- 
tently over  3-  to  5-year  intervals  to  reduce  costs. 
Generally  these  types  of  data  will  provide  biologists 
with  adequate  information  for  management  deci- 
sions, although  some  management  situations  may  re- 
quire intensive,  multiyear  monitoring. 

Another  way  of  reducing  costs,  while  obtaining 
community  data  on  amphibians  and  reptiles,  is  to 
sample  indicator  species.  There  are  two  general 
types  of  indicator  species: 

•  amphibians  and  reptiles  that  represent  species 
assemblages  that  use  habitats  in  similar  ways 
(species  guilds),  or 

•  species  that  use  specific  habitat  components. 

For  example,  population  trend  data  can  be  collected 
on  one  of  several  species  that  require  downed  litter. 
The  trend  of  the  species  sampled  should  reflect 
trends  of  the  entire  species  assemblage  because  they 
require  the  same  habitat  component.  However,  some 
tests  have  revealed  that  species  within  delineated 
guilds  respond  differently  to  land-use  practices.  Man- 
nan  et  al.  ( 1984)  found  that  67%  of  birds  within 
guilds  responded  differently  to  timber  harvest.  The 
problem  with  guilds  is  that  all  species  occupy  sepa- 
rate niches,  some  more  specific  than  others.  There- 
fore, many  amphibians  and  reptiles  within  the  same 
guild  as  the  indicator  may  respond  differently  to 
habitat  changes.  Mannan  et  al.  (1984)  pointed  out 
that  guilding  works  either  when  the  investigator 
defines  very  specific  guilds  or  the  land-use  impact  is 
so  severe  that  several  habitat  components  are  lost. 


The  other  use  of  indicator  species  is  to  deter- 
mine conditions  of  a  habitat  component  based  on 
condition  and  trend  of  an  amphibian  or  reptile.  Al- 
though certain  microhabitat  specific  species  could  be 
used  in  this  manner,  greater  accuracy  and  less  sam- 
ple cost  may  be  obtained  by  sampling  the  habitat 
components  directly. 


In  selecting  indicator  amphibians  and  reptiles,  a 
biologist  must  again  be  fully  aware  of  sampling 
method  and  species  life  history  limitations.  Amphibi- 
ans and  reptiles  whose  populations  respond  to  cli- 
matic conditions  should  not  be  used  as  indicator 
species,  especially  in  short-duration  studies.  How- 
ever, it  may  be  possible  to  use  these  species  in  long- 
term  studies  where  climatic  effects  can  be  separated. 

Presence  and  absence  can  be  used  to  assess 
gross  species  assemblage  and  habitat  component 
trends  provided  sampling  techniques  are  adequate. 
For  example,  absence  of  a  salamander  reflects  the 
lack  of  required  habitat  components  on  one  site,  and 
presence  of  the  species  on  a  similar  site  reflects  the 
presence  of  required  habitat  components.  The  key 
phrase  in  the  previous  sentence  is  "on  a  similar  site." 
To  be  compared,  these  sites  must  have  similar  topog- 
raphy, rainfall,  temperature,  and  vegetation.  If  these 
variables  are  not  similar  on  each  site,  the  biologist 
cannot  determine  whether  lack  of  habitat  compo- 
nents on  one  site  limits  the  salamander's  occurrence. 
Biologists  should  select  species  that  rely  heavily  on 
a  specific  microhabitat.  This  will  increase  the  accu- 
racy of  predictions  made  about  microhabitat 
conditions. 


288 


Amphibians  and  Reptiles 


LITERATURE  CITED 


ANDERSON,  K  and  C.L.  SMITH.  1950.  An  electrical  appa- 
ratus for  herpetological  collecting.  Copeia  1950:322. 

BALGOOYEN,  T.G.  1977.  Collecting  methods  for  amphibi- 
ans and  reptiles.  U.S.  Dep.  Inter.,  Bur.  Land  Manage. 
Tech.  Note  299.  Denver,  CO. 

BANTA,  B.H.  1957.  A  simple  trap  for  collecting  desert 
reptiles.  Herpetologica  13:174-176. 

BRATTSTROM,  B.H.  1965.  Body  temperatures  of  reptiles. 
Am.  Midi.  Nat.  73:376-422. 

BREEN,  J.F.  1949.  Reptiles:  their  habits  and  care.  All-Pets 
Mag. 

BROWN,  L.E.  and  M.J.  LITTLEJOHN.  1972.  Male  release 
call  in  the  Bufo  americanns  group.  Pages  310-323  in 
W.F.  Blair,  ed.  Evolution  in  the  Genus  Bufo.  Univ. 
Texas  Press,  Austin. 

BURGE,  B.L.  1979.  Survey  of  the  present  distribution  of 
the  desert  tortoise,  Gopherus  agassizii,  in  Arizona. 
U.S.  Dep.  Inter.,  Bur.  Land  Manage.  Contract  YA-512- 
CT8-108.  Denver,  CO. 

BURY,  R.B.  1982.  Structure  and  composition  of  Mojave 
Desert  reptile  communities  determined  with  a  re- 
moval method.  Herpetological  Communities,  U.S.  Dep. 
Inter.,  Fish  and  Wildl.  Serv.  Wildl.  Res.  Rep.  13. 

,  H.W.  CAMPBELL,  and  N.J.  SCOTT,  Jr.  1980.  Role 

and  importance  of  nongame  wildlife.  Pages  197-207  in 
Trans.  45th  North  Am.  Wildl.  Nat.  Res.  Conf.  Washing- 
ton, DC. 

and  R.A.  LUCKENBACH.  1977.  Censusing  desert 

tortoise  populations  using  a  quadrat  and  grid  location 
system.  Desert  Tortoise  Council  Symp.  Proc. 
1977:169-178. 

and  M.G.  RAPHAEL.  1983.  Inventory  methods  for 


amphibians  and  reptiles.  Proc.  Int.  Conf.  Renewable 
Resour.  Inventories  for  Monitoring  Changes  and 
Trends.  Oregon  State  Univ.,  Corvallis. 

BUSACK,  S.D.  and  R.B.  BURY.  1974.  Some  effects  of  off- 
road  vehicles  and  sheep  grazing  on  lizard  populations 
in  the  Mojave  Desert.  Biol.  Consev.  6:179-183. 

CAMPBELL,  H.W.  and  S.P.  CHRISTMAN.  1982.  Field  tech- 
niques for  herpetofaunal  community  analysis.  Herpe- 
tological Communities,  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  Wildl.  Res.  Rep.  13. 

CARPENTER,  C.C.  1953-  Trapping  technique  for  aquatic 
salamanders.  Herpetologica  8(4):  183- 

.  1955.  Sounding  turtles:  a  field  location  technique. 

Herpetologica  1 1:120. 

CLARK,  DR.,  Jr.  1966.  A  funnel  trap  for  small  snakes. 
Trans.  Kansas  Acad.  Sci.  69(  1  ):91-95. 

CREUSERE,  F.M.  and  W.G.  WHITFORD  1982.  Use  of  time 
and  space  by  lizards.  Herpetological  Communities, 
U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  Wildl.  Res.  Rep. 
13. 

DARGAN,  L.M.  and  W.H.  STICKEL.  1949.  An  experiment 
with  snake  trapping.  Copeia  1949:264-268. 

DUNDEE,  HA.  1950.  An  improved  method  for  collecting 
living  lizards  and  frogs.  Herpetologica  6(3):78-79. 

EAKIN,  R.M.  1957.  Use  of  copper  wire  in  noosing  lizards. 
Copeia  1957:148. 

EDWARDS,  S.R.  1975.  Collections  of  preserved  amphibians 
and  reptiles  in  the  United  States.  SSAR  Misc.  Publ. 
Herpetological  Circular  3- 

FERNER,  J.W.  1979.  A  review  of  marking  techniques  for 
amphibians  and  reptiles.  SSAR.  Herpetological  Circular 
9. 


FEUER,  R.C.  1980.  Underwater  traps  for  aquatic  turtles. 
Herp.  Review  11(4):107-108 

FITCH,  H.S.  1951.  A  simplified  type  of  funnel  trap  for 
reptiles.  Herpetologica  7:77-80. 

FOUQUETTE,  M.J. ,  Jr.  1980.  Effect  of  environmental  tem- 
peratures on  body  temperature  of  aquatic-calling 
anurans.J.  Herpetol.  l4(4):347-352. 

FRANKLIN,  M.A.  1947.  An  inexpensive  snare  for  water 
snakes.  Copeia  1947:143- 

GIBBONS,  J.W.  and  R.D.  SEMLITSCH.  1981.  Terrestrial 
drift  fences  with  pitfall  traps:  an  effective  technique 
for  quantitative  sampling  of  animal  populations.  Brim- 
ley  ana  7:1-16. 

GO  IN,  C.J.  1942.  A  method  of  collecting  the  vertebrates 
associated  with  water  hyacinths.  Copeia  (3):  183- 184. 

GUNNING,  G.E.  and  W.M.  LEWIS.  1957.  An  electrical 

shocker  for  the  collection  of  amphibians  and  reptiles 
in  the  aquatic  environment.  Copeia  1957:52. 

HALL,  R.J.  1967.  A  simplified  live-trap  for  reptiles.  Trans. 
Kansas  Acad.  Sci.  70(3):402-404. 

.  1 980.  Effects  of  environmental  contaminants  on 

reptiles:  a  review.  U.S.  Dep.  Inter.,  Fish  and  Wildl. 
Serv.  Spec.  Sci.  Rep.  228.  Washington,  DC. 

HANSEN,  R.M.,  M.K.  JOHNSON,  and  T.R.  VAN  DEVENDER. 
1976.  Foods  of  the  desert  tortoise,  Gopherus  agassi- 
zii, in  Arizona  and  Utah.  Herpetologica  32(3):247- 
251. 

HAWKEN,  J.L.  1951.  Water  system  acts  as  reptile  and 
amphibian  trap.  Herpetologica  7:81-83- 

HEATWOLE,  H„  A.  MALDONADO,  and  J.  OJASTI.  1964.  A 
trapping  method  for  capturing  lizards.  Herpetologica 
20(3):212-213. 

HECKEL,  D.G.  and  J.  ROUGHGARDEN.  1979.  A  technique 
for  estimating  the  size  of  lizard  populations.  Ecology 
60(5):966-975. 

HORN,  H.S.  1966.  Measurement  of  overlap  in  comparative 
ecological  studies.  Am.  Nat.  100:419-424. 

IVERSON,  J.B.  1979.  Another  inexpensive  turtle  trap. 
Harp.  Review  10(2):55. 

JONES,  KB.  1981a.  Effects  of  grazing  on  lizard  abundance 
and  diversity  in  western  Arizona.  Southwest  Nat. 
26(2):107-115. 

.  1981b.  Distribution,  ecology,  and  habitat  manage- 
ment of  the  reptiles  and  amphibians  of  the  Hualapai- 
Aquarius  planning  area,  Mohave  and  Yavapai  Counties, 
Arizona.  U.S.  Dep.  Inter.,  Bur.  Land  Manage.  Tech. 
Note  353-  Denver,  CO. 

,  LP.  KEPNER,  and  W.G.  KEPNER.  1983.  Anurans  of 


Vekol  Valley,  Central  Arizona.  Southwest  Nat. 

28(4):469-470. 
KEPHART,  D.G.  and  S.J.  ARNOLD.  1982.  Garter  snake 

diets  in  a  fluctuating  environment:  a  seven-year  study. 

Ecology  63(5):  1232- 1236. 
LAGLER,  KF.  1943.  Methods  of  collecting  freshwater 

turtles.  Copeia  1943:21-25. 
LANNON,  JR.,  Jr.  1962.  A  different  method  of  catching  the 

desert  lizards,  Callisaurus  and  Uma.  Copeia 

1962:437-438. 
LEE,  D.S.  1968.  Herpetofauna  associated  with  central  Flor- 
ida mammals.  Herpetologica  24(  1  ):83-84. 
LEGLER,  J.M.  I960.  A  simple  and  inexpensive  device  for 

trapping  aquatic  turtles.  Utah  Acad.  Sci.  Proc.  37:63- 

66. 
LILLYWHITE,  H.B.  1982.  Tracking  as  an  aid  in  ecological 

studies  of  snakes.  Herpetological  Communities,  U.S. 

Dep.  Inter.,  Fish  and  Wildl.  Serv.  Wildl.  Res.  Rep.  1 3- 
LUCKENBACH,  R.A.  and  R.B.  BURY.  1983-  Effects  of  off- 


Amphibians  and  Reptiles 


289 


road  vehicles  on  the  biota  of  the  Algodones  Dunes, 
Imperial  County,  California.  J.  Appl.  Ecol.  20:265-286. 

MANNAN,  R.W.,  ML.  MORRISON,  and  EC.  MESLOW. 

1984.  Comment:  The  use  of  guilds  in  forest  bird  man- 
agement. The  Wildl.  Soc.  Bull.  12(4):426-430. 

MARTOF,  B.S.  1963.  An  effective  technique  for  capturing 
stream -dwelling  organisms.  Copeia  1963:439-440. 

MICHAUD,  T.C.  1964.  Vocal  variation  in  two  species  of 
chorus  frogs,  Pseudacris  nigrita  and  Pseudacris 
clarki,  in  Texas.  Evolution  18:498-506. 

MILLSAP,  B.A.  1981.  Distributional  status  of  Falconiformes 
in  west-central  Arizona:  with  notes  on  ecology,  repro- 
ductive success,  and  management.  U.S.  Dep.  Inter., 
Bur.  Land  Manage.  Tech.  Note  355.  Denver,  CO. 

MILLSAP,  B.A.  and  W.  HARRISON.  1981.  Food  and  forag- 
ing habitats  of  common  black  hawks  (Buteogallus 
anthracinus)  in  western  Arizona.  U.S.  Dep.  Inter., 
Bur.  Land  Manage.  Phoenix  District,  AZ.  Unpubl. 

MOULTON,  J.M.  1954.  Notes  on  the  natural  history,  col- 
lection, and  maintenance  of  the  salamander  Ambys- 
toma  maculatum  Copeia  1954:64-65. 

NEILL,  W.T.  1956.  Another  device  for  collecting  lizards. 
Copeia  1956:124-125. 

ORTEGA,  A.,  ME.  MAURY,  and  R.  BARBAULT.  1982.  Spa 
tial  organization  and  habitat  partitioning  in  a  moun- 
tain lizard  community  of  Mexico.  Ecol.  Gen. 
3(3):323-330. 

PETOKAS,  P.J.  and  MM.  ALEXANDER.  1979.  A  new  trap 
for  basking  turtles.  Herp.  Review  10(3):90. 

PIANKA,  E.R.  1966.  Convexity,  desert  lizards,  and  spatial 
heterogeneity.  Ecology  47(6):1055-1059. 

PEETRUSZKA,  R.D.  1981.  An  evaluation  of  stomach  flush- 
ing for  desert  lizard  diet  analysis.  Southwest.  Nat. 
26(2):101-105. 

PIRNIE,  M.D.  1935.  Michigan  waterfowl  management. 
Michigan  Dep.  Conserv.,  Lansing,  ML 

PISANI,  G.R.  1973.  A  guide  to  preservation  techniques  for 
amphibians  and  reptiles.  SSAR  Misc.  Publ.  Herpetolog- 
ical  Circular  1. 

PORZER,  L.M.  1982.  Movements,  behavior,  and  body 

temperatures  of  the  Gila  monster  {Heloderma  suspec- 
turn  Cope )  in  the  vicinity  of  Queen  Creek,  Arizona. 
M.S.  Thesis,  Arizona  State  Univ.,  Tempe. 

POUGH,  F.H.  1970.  A  quick  method  for  permanently 

marking  snakes  and  turtles.  Herpetologica  26:428-430. 

ROGERS,  T.L.  1939.  A  lizard  live-trap.  Copeia  1939:51. 

SCHNEIDER,  P.B.  1981.  A  population  analysis  of  the  de- 
sert tortoise,  Gophenis  agassizii,  in  Arizona.  U.S.  Dep. 


Inter.,  Bur.  Land  Manage.  Phoenix  District  Office. 
Contract  AZ-950-CT9-0014. 

SCOTT,  A.F.  and  J.L.  DOBIE.  1980.  An  improved  design  for 
a  thread-trailing  device  used  to  study  terrestrial 
movements  of  turtles.  Herp.  Review  1 1(4):  106- 107. 

STEBBINS,  R.C.  1966.  A  field  guide  to  western  reptiles 
and  amphibians.  Houghton  Mifflin  Co.  Boston,  MA. 
27pp. 

STICKEL,  W.H.  1944.  A  simple  and  effective  lizard  snare. 
Copeia  1944:251-252. 

STORM,  R.M.  and  R.A.  PIMENTEL.  1954.  A  method  of 

studying  amphibian  breeding  populations.  Herpetolo- 
gica 10:161-166. 

SVIHLA,  A.  1959-  A  simple  method  of  collecting  Ascaphus 
truei  tadpoles.  Copeia  1959:72. 

TINKLE,  D.W.  1982.  Results  of  experimental  density  ma- 
nipulation in  an  Arizona  lizard  community.  Ecology 
63(  1  ):57-65. 

VITT,  L.J.,  AC.  HULSE,  and  R.D.  OHMART.  1974.  Repro- 
duction and  ecology  of  a  Colorado  River  population 
of  See iopous  magister.  Herpetologica  30:410-417. 

VITT,  L.J.  and  R.D.  OHMART.  1978.  Herpetofauna  of  the 
lower  Colorado  River:  Davis  Dam  to  the  Mexican 
border.  Proc.  West.  Found.  Vert.  Zool.  2(2>35-72. 

VOGT,  R.C.  and  R.L  HINE.  1982.  Evaluation  of  techniques 
for  assessment  of  amphibian  and  reptile  populations 
in  Wisconsin.  Pages  201-217  in  Scott,  N.J.,  Jr.  ed. 
Herpetological  Communities.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  Wildl.  Res.  Rep.  13- 

VOGT,  W.  1941.  A  practical  lizard  trap.  Copeia  1941:115. 

WHITAKER,  AH.  1967.  Baiting  pitfall  traps  for  small  liz- 
ards. Herpetologica  23(4):309-310. 

WHITFORD,  W.G.  and  F.M.  CREUSERE.  1977.  Seasonal  and 
yearly  fluctuations  in  Chihuahuan  Desert  lizard  com- 
munities. Herpetologica  33:54-65. 

WDEGMAN,  D.L.,  M.  HAKKILA,  K.  WHITMORE,  and  R.A. 
COLE.  1980.  Survey  of  the  Sacramento  Mountain 
salamander  (Aneides  hardyi)  habitat  on  the  Cloud- 
croft  and  Mayhill  Districts  in  the  Lincoln  National 
Forest.  U.S.  Dep.  Agric,  For.  Serv.  Contract  OM-40- 
7512-0-632. 

WILLIAMS,  R.D.,  J.E.  GATES,  and  C.H.  HOCUTT.  1981.  An 
evaluation  of  known  and  potential  sampling  tech- 
niques for  hellbender,  Cryptobranchus  alleganiensis. 
J.  Herpetol.  15(l):23-27. 

ZWEIFEL,  R.G.  and  C.H.  LOWE.  1966.  The  ecology  of  a 
population  of  Xantusia  vigilis,  the  desert  night  lizard. 
Am.  Mus.  Novit.  2247:1-57. 


290 


Amphibians  and  Reptiles 


15 

SONGBIRDS 


Ronald  A.  Ryder 

Department  of  Fishery  &  Wildlife  Biology 
Colorado  State  University 
Fort  Collins,  CO  80523 


Editor's  Note:  This  chapter  on  Songbirds  is  the  first 
of  six  chapters  on  birds.  The  chapter  covers  the  true 
songbirds  (Order  Passeriformes)  and  also  other 
orders  that  require  similar  techniques  for  habitat 
and  population  surveys.  Techniques  for  surveying 
songbirds  are  relatively  well-described  and  formal- 
ized, and  this  chapter  provides  an  overview  of  these 
along  with  references  to  the  appropriate  publica- 
tion for  step  by  step  procedures. 

In  addition,  the  chapter  covers  general  sources  of 
information  on  birds.  In  North  America,  birds  are 
actively  observed,  counted,  and  recorded  by  mil- 
lions of  avid,  amateur  birdwatchers  as  well  as 
more  limited  numbers  of  professional  ornitholo- 
gists. The  observations  of  these  people,  as  summa- 
rized in  publications  such  as  local  bird  lists  and 
periodicals  such  as  American  Birds,  are  valuable 
sources  of  background  information  on  distribution 
and  habitat  use,  for  not  only  songbirds  but  for  all 
bird  species.  To  avoid  repetition,  these  sources  of 
information  are  covered  in  this  chapter  and  not  in 
the  others. 


INTRODUCTION 

Songbirds  will  be  considered  to  include  not 
only  birds  of  the  order  perching  birds  (Passeri- 
formes), but  also  cuckoos  (Cuculiformes),  night- 
hawks  (Caprimulgiformes),  swifts  and  hummingbirds 
(Apodiformes),  trogons  (Trogoniformes),  kingfishers 
(Coraciiformes),  and  woodpeckers  (Piciformes), 
constituting  over  350  species  in  North  America, 
north  of  Mexico.  With  few  exceptions,  most  are  ex- 
tremely mobile,  and  many  are  migratory,  moving 
latitudinally  or  altitudinally  between  breeding 
grounds  and  wintering  areas  each  year.  Most  have 
relatively  high  breeding  potentials,  and  their  num- 
bers and  densities  may  fluctuate  greatly  from  season 
to  season  due  to  recruitment  and  climatic  changes. 
Some  even  shift  their  centers  of  abundance  based  on 
vegetation  and  prey  densities — especially  in  the 
West — depending  on  seasonal  and  regional  variations 
in  precipitation. 

Although  it  is  possible  to  obtain  comparatively 
accurate  estimates  of  absolute  densities  for  a  few 
endangered  forms  (such  as  Kirtland's  warbler  [Den- 
droica  kirtlandii]  and  dusky  seaside  sparrow  [Am- 
modramus  maritimus  nigrescens]),  for  most  species 
it  is  neither  physically  possible  nor  economically 
reasonable  to  obtain  absolute  densities.  For  most 
management  purposes,  estimates  of  relative  density 
are  adequate. 

The  literature  concerning  inventorying  and 
monitoring  songbirds  is  vast,  if  not  overwhelming. 
Ralph  and  Scott  (1981)  edited  symposium  proceed- 
ings which  are  probably  the  best  available  references 
on  methods  for  estimating  numbers  of  terrestrial 
birds.  Franzeb  (1977)  and  Call  (1981 )  discussed 


Songbirds 


291 


songbirds  in  their  U.S.  Bureau  of  Land  Management 
(BLM)  Technical  Notes  on  terrestrial  wildlife  inven- 
tories. Davis  (1982)  and  Davis  and  Winstead  (1980) 
also  provided  useful  references;  they  gave  specific 
inventory  methods  for  many  species  of  songbirds. 
The  U.S.  Forest  Service  has  published  proceedings  of 
several  symposiums  and  workshops  relating  songbird 
numbers  to  habitats  and  their  management  (Smith 
1975;  DeGraff  1978a,  b;  DeGraff  and  Evans  1979; 
DeGraff  and  Tilghman  1980).  Verner  (1985)  criti- 
cally evaluated  many  counting  techniques  suitable 
for  songbirds.  American  Birds  (and  earlier  Audubon 
Field  Notes  and  Bird  Lore)  regularly  summarizes 
Christmas  Bird  Counts  (CBCs)  in  its  July -August  is- 
sue, and  winter  bird-population  studies  and  breeding 
bird  censuses  in  its  January-February  issue.  Not  only 
do  the  main  ornithological  journals  such  as  Auk, 
Condor,  Wilson  Bulletin,  and  the  Journal  for  Field 
Ornithology  regularly  contain  songbird-habitat  pa- 
pers, but  a  multitude  of  state  and  regional  bird  peri- 
odicals are  also  useful  sources  of  such  information. 
Many  of  these  periodicals  are  indexed  in  Wildlife 
Review  and  Wildlife  Abstracts. 

This  chapter  describes  habitat  characteristics, 
such  as  cliffs  and  other  physical  features,  as  well  as 
vegetative  structure  and  composition  as  they  relate 
to  the  over  20  families  of  songbirds  found  in  the 
West.  Throughout,  key  references  to  major  avian 
groups  are  given,  which  can  be  consulted  for  specific 
habitat  requirements  (See  Table  1). 

The  major  methods  or  techniques  used  to  moni- 
tor or  inventory  songbirds  are  described,  with  criti- 
cal evaluations  made  of  their  applicability,  especially 
in  the  western  U.S.  (See  Table  2). 

Ornithologists  and  land-use  managers  believed 
that  numbers  of  bird  species  and  individuals  might 
serve  as  a  "litmus  test  of  the  environment."  It  was 
thought  that  a  decline  in  species  richness  and  diver- 
sity, as  determined  by  routine  monitoring,  would 
serve  as  an  early  warning  of  environmental  degrada- 
tion. More  specifically,  overgrazing,  overcutting,  or 
too  heavy  recreational  use  of  forest  and  range  eco- 
systems would  be  revealed  by  changes  in  bird 
populations. 

Efforts  to  use  songbird  trends  as  indicators  of 
habitat  quality  and  habitat  changes,  however,  have 
not  been  as  successful  as  many  had  hoped.  Songbirds 
are  highly  mobile  and  many  individuals  may,  at  most, 
spend  only  3  to  4  months  per  year  on  a  given  area. 
Migratory  species  are  probably  more  affected  by 
habitat  destruction  and  pesticide  use  on  their  winter- 
ing grounds  in  the  Neotropics,  where  they  spend 
more  of  their  annual  life  cycle,  than  they  are  by 
changes  on  their  breeding  grounds  in  the  western 
U.S.,  where  they  may  spend  one  third  or  less  of  the 
year. 


Lark  bunting. 


The  skill,  time,  and  dedication  required  to  thor- 
oughly inventory  and  monitor  bird  populations  is 
probably  more  than  the  average  land-use  manager 
can  provide.  Identification  of  the  many  species,  by 
sight  and  sound,  and  determination  of  sex  and  age 
classes  of  songbirds  to  be  found  in  a  BLM  District  re- 
quire more  knowledge  and  skill  than  most  typical 
"wildlifers"  possess  or  can  acquire  without  consider- 
able training  and  practice. 

Also,  most  songbirds  are  not  primary  consumers 
but  insectivores,  at  least  during  the  breeding  season. 
Thus,  cause-and-effect  relationships  with  vegetative 
changes  are  not  as  direct  as  with  resident  primary 
consumers,  such  as  most  rodents. 

With  these  reservations  in  mind,  some  monitor- 
ing approaches  that  might  be  used  to  ascertain 
trends  in  bird  numbers  follow. 


HABITAT  FEATURES  CORRELATED  WITH 
SPECIES  GROUPS 

Physical  Features 

Most  songbirds  are  associated  with  vegetation  of 
a  variety  of  types,  but  a  few  species,  especially  for 
nesting,  are  largely  dependent  on  cliffs  (cliff  swal- 
lows [Hirundo  pyrrhonota]  and  white-throated 
swifts  [Aeronautes  saxatalis],  for  example).  Others 
use  earthen  banks  (bank  swallows  [Riparia  riparia] 
and  belted  kingfishers  [Ceryle  alcyon]).  These  special 
features  are  described  in  Chapters  27  and  31  and 
listed  in  Table  1. 

An  increasing  number  of  songbirds  are  adapting 
to  nesting  and  roosting  in  or  on  structures,  such  as 
buildings  and  bridges,  as  well  as  utilizing  artificial 
nest  boxes  and  platforms. 


292 


Songbirds 


Table  1.     Selected  taxa  of  birds  of  western  North  America,  structural  features  used,  and  some  key  references 
to  avian  ecology. 


Taxa 

Common  Names 

Structural  Features  Used 
(mainly  for  nesting) 

Key  References 

03 
T3 
03 

E 

C 

CO 

E 

■o 

c 

O 

i_ 
O) 

d) 

I— 
CO 
JD 

CO 
CD 

c 
> 

CO 

c 
co 

to 

CO 

CO 
CO 

.Q 
o 
w 

CO 
CO 

h_ 

J3 

\— 
CO 

o 

CD 
CD 

i_ 

0 
0 

CO 

PASSERI- 
FORMES 

Perching  birds 

Anderson  (1979),  Balda 
(1975),  Call  (1981),  Ralph  & 
Scott  (1981),  Trimble  (1975) 

Tyrannidae 

Tyrant  flycatchers 

• 

• 

• 

• 

• 

• 

Bent  (1942),  Forest  Service 
(1982),  Graberetal.  (1974), 
Maser  et  al.  (1984) 

Alaudidae 

Larks 

• 

• 

• 

• 

Beason  &  Franks  (1973, 
1974),  Boyd  (1976),  Pickwell 
(1931) 

Hirundinidae 

Swallows 

Bent  (1942),  Graberetal. 
(1972),  Harrison  (1975) 

Corvidae 

Jays,  magpies,  &  crows 

• 

• 

• 

• 

Balda  &  Bateman  (1972), 
Bent  (1946),  Goodwin 
(1976) 

Paridae 

Chickadees  &  titmice 

• 

• 

• 

Bailey  &  Niedrach  (1965), 
Bent  (1946),  Dixon  (1961) 

Sittidae 

Nuthatches 

• 

Bent  (1948),  Bock  &  Lep- 
thien  (1972),  Norris  (1958) 

Certhiidae 

Creepers 

• 

Bent  (1948),  McClelland  & 
Frissell  (1975) 

Troglodytidae 

Wrens 

• 

• 

• 

• 

• 

• 

Anderson  &  Anderson 
(1973),  Armstrong  (1955), 
Bent  (1948) 

Cinclidae 

Dippers 

• 

Bakus  (1959a,  1959b),  Price 
&Bock  (1983) 

Sylviinae 

Kinglets  &  gnatcatchers 

• 

• 

• 

Bent  (1949),  Root  (1967), 
Trimble  (1975) 

Turdinae 

Solitaires,  thrushes,  & 
allies 

Bent  (1949),  Graber  et  al. 
(1971),  Jackman  &  Scott 
(1975),  Young  (1955),  Ze- 
leny  (1976) 

Mimidae 

Mockingbirds  & 
thrushes 

Bent  (1948),  Brazier  (1964), 
Graber  et  al.  (1970) 

Motacillidae 

Wagtails  &  pipits 

• 

• 

Bent  (1950),  Verbeek  (1970) 

Bombycillidae 

Waxwings 

• 

• 

• 

Bent  (1950),  Lea  (1942), 
Putnam  (1949) 

Ptilogonatidae 

Silky  flycatchers 

• 

• 

• 

Bent  (1950),  Verner  &  Boss 
(1980) 

Laniidae 

Shrikes 

• 

• 

• 

• 

Bystrak  (1983),  Graber  et  al. 
(1973),  Miller  (1931),  Porter 
et  al.  (1975) 

Sturnidae 

Starlings 

DeHaven  &  DeHaven 
(1973),  Feare  (1984),  Kessel 
(1957) 

Songbirds 


293 


Table  1.     Selected  taxa  of  birds  of  western  North  America,  structural  features  used,  and  some  key  references 
to  avian  ecology  (concluded). 


Taxa 

Common  Names 

Structural  Features  Used 
(mainly  for  nesting) 

Key  References 

0 
-o 

03 

E 

c 

03 

E 

■o 

c 

o 

I— 

en 
0 
co 

CO 

CD 

c 
> 

CO 

k_ 

"D 

c 

CO 
C/3 

CO 
CO 

co 

k_ 

o 

CO 

co 

O) 

X3 

k_ 
JZ 
CO 

o 

CD 

0 

I— 
CO 

0 
0 

sz 

CO 
03 

Vireonidae 

Vireos 

• 

• 

• 

Barlow  (1962),  Bent  (1950), 
Verner  &  Boss  (1980),  Win- 
ternitz(1976) 

Parulinae 

Wood-warblers 

• 

• 

• 

Bent  (1953),  Graber  et  al. 
(1983),  Griscom  &  Sprunt 
(1957),  Harrison  (1984) 

Thraupinae 

Tanagers 

• 

• 

Bent  (1958),  Verner  &  Boss 
(1980) 

Cardinalinae 

Cardinals,  grosbeaks, 
&  allies 

Austin  (1968),  Verner  & 
Boss  (1980) 

Emberizinae 

Emberizine  finches 

Austin  (1968),  Forest  Ser- 
vice (1982) 

Icterinae 

Blackbirds  &  allies 

Bent  (1958),  Nero  (1984), 
Orians  (1980) 

Fringillinae 

Fringilline  finches 

Austin  (1968),  Maser  et  al. 
(1984),  Trimble  (1975) 

Carduelinae 

Cardueline  finches 

Austin  (1968),  Trimble 
(1975),  Verner  &  Boss 
(1980) 

Passeridae 

Old  world  sparrows 

• 

• 

Packard  (1966),  Summers- 
Smith  (1963,  1967) 

CUCULIFORMES 

Cuckoos,  roadrunners, 
&  anis 

• 

• 

• 

• 

Bent  (1940),  Ohmart  (1973), 
Preble  (1957) 

Cuculidae 

CAPRIMULGI- 
FORMES 

Goatsuckers 

• 

Bent  (1940),  Caccamise 
(1974),  Verner  &  Boss 
(1980) 

Caprimulgidae 

APODIFORMES 

Swifts 

• 

• 

Bent  (1940),  Knorr(1961) 

Apodidae 

Trochilidae 

Hummingbirds 

• 

• 

• 

• 

• 

Calder  (1973),  Johnsgard 
(1983) 

TROGONI- 
FORMES 

Trogons 

• 

• 

• 

Bent  (1940),  Forest  Service 
(1975),  Phillips  et  al.  (1964) 

Trogonidae 

CORACII- 
FORMES 

Kingfishers 

• 

Bent  (1940),  Cornwell 
(1963),  Eipper(1956) 

Alacedinidae 

PICIFORMES 

Woodpeckers  and 
allies 

• 

• 

• 

Bent  (1939),  Bock  (1970), 
Jackson  &  Scott  (1975) 

Picidae 

294 


Songbirds 


Table  2.     Some  methods  for  monitoring  songbirds — advantages  and  disadvantages.  (See  text  for  more  details.) 


Method 

Season  of  Use 

Advantages 

Disadvantages 

Checklist 

All 

Quick,  relatively  cheap  to 
compile.  Use  literature  and 
volunteers.  Can  visually 
portray  arrival  and  depar- 
ture dates  and  relative 
abundance. 

Mainly  qualitative,  only 
relative  abundance  and 
gross  habitat  preferences. 

Atlas 

Mainly  stress-breeding 
season  (can  show  other 
seasons  also) 

Gives  "big  picture."  Fairly 
understandable  to  the  pub- 
lic. 

Usually  rather  large  scale 
(1°  blocks  of  latitude  and 
longitude).  Requires  con- 
siderable organization  and 
cooperation  of  many  ob- 
servers. 

Christmas  Bird 
Count  (CBC) 

Winter 

Many  areas  covered  in 
past.  Large  sample  size 
nationally. 

Only  one  coverage/area/ 
year.  Area  of  special  inter- 
est may  not  have  been 
covered.  Urban  areas 
stressed.  Only  indexes,  no 
absolute  densities. 

Winter  bird-popu- 
lation studies 

Winter 

More  coverages/season 
than  CBC.  Replications, 
hence  greater  possibility  of 
statistical  analysis. 

Smaller  area  and  number 
of  areas  covered  than 
CBCs.  More  time  and  effort 
required. 

Spot-mapping 

Mainly  breeding  season 
(some  winter  territorial 
mapping  possible) 

Usually  considered  to  give 
good  estimates  of  breeding 
densities. 

Requires  repeated  cover- 
ages by  skilled  observer. 
Need  good  vegetative  map 
and  vegetative  measure- 
ments to  be  most  useful. 

Transects  (mainly 
walked) 

Any  season 

Quicker,  cheaper  to  con- 
duct than  spot-mapping. 
More  random  samplings 
possible. 

Accuracy  highly  variable 
species  to  species,  habitat 
to  habitat.  Less  accurate 
in  nonbreeding  season. 

North  American 
Breeding  Bird 
Survey  (BBS) 

Breeding  season  (has  been 
used  for  other  seasons) 

Large  sample  size  nation- 
ally. Relatively  cheap  and 
quick  to  conduct. 

Only  single  coverage  of  a 
given  route  each  year. 
Habitat  data  not  regularly 
gathered.  Only  indexes 
of  abundance.  Area  must 
be  accessible  by  road. 

Point  count  (IPA) 

Breeding  season 

Good  statistical  reliance 
reported  for  Europe  (not 
well  tested  in  U.S.). 

Only  two  coverages/plot/ 
season,  normally.  Consid- 
erable time  (20  min)  per 
stop.  Only  frequency  data. 

Density-fre- 
quency relation- 
ship (EFP) 

Breeding 

As  above.  Said  to  give 
density  correlations. 

More  time-consuming  than 
point  count. 

Maryland  winter 
bird  survey 

Winter 

Can  be  used  where  roads 
are  lacking.  Comparatively 
quick  and  cheap. 

Walked  (unlike  roadside 
count),  requires  more  time 
and  effort  than  BBS.  Not 
well  tested  in  western  U.S. 

Variable-circular 
plots 

All  seasons  (but  mainly 
used  in  breeding  season) 

Can  cover  large  geographi- 
cal areas,  compare  differ- 
ent habitats,  and  work  in 
rugged  and  remote  terrain. 

Requires  considerable 
training  in  estimating  or 
time  in  measuring  dis- 
tances (aural  as  well  as 
visual  distances). 

Songbirds 


295 


Table  2.     Some  methods  for  monitoring  songbirds 
(concluded). 


-advantages  and  disadvantages.  (See  text  for  more  details.) 


Method 

Season  of  Use 

Advantages 

Disadvantages 

Playback  of  tape 
recordings 

Mainly  breeding  season 

Quick  way  to  get  male 
responses.  Speeds  up 
spot-mapping. 

Only  certain  species  re- 
spond. Can  alter  behavior 
(interfere  with  nesting  suc- 
cess?). 

Mark  and  recap- 
ture 

Any  season 

Can  use  similar  population 
estimates  as  used  with 
game  species. 

Not  highly  efficient.  Consid- 
erable time  and  effort  re- 
quired. Special  permits 
required.  Many  biases  in- 
volved (trap-happy  vs.  trap- 
shy  individuals). 

Nest  monitoring 

Breeding  season 

Relates  to  actual  nesting 
efforts. 

Very  time-consuming. 
Highly  variable  with  species 
and  habitats  involved.  Can 
increase  predation  and 
desertion. 

Vegetation 

Structure.  Vegetative  structure  and  habitat 
configuration  (physiognomy)  are  generally 
considered  more  important  in  distribution  and 
abundance  of  birds  than  plant  species  composition 
(floristics — Anderson  and  Shugart  1974;  Hilden  1965; 
James  1971;  Wiens  1969).  However,  some  believe 
floristics  may  be  more  important  within  gross  habitat 
types  (Rotenberry,  pers.  commun.).  Rotenberry's 
analysis  of  International  Biological  Program  data 
from  grasslands  disclosed  that  55%  of  the  variation 
in  bird  community  composition  was  associated  with 
floristic  variation  and  only  35%  was  associated  with 
physiognomy. 


Others  have  shown  patterns  of  avian  distribution 
strongly  correlated  with  vegetational  structure 
(MacArthur  and  MacArthur  1961;  Willson  1974  et 
al. ).  Structural  features  commonly  measured  include 
density,  mean  distance  to  nearest  plant  neighbors, 
height,  diameter  at  breast  height  (DBH)  for  trees, 
relative  dominance,  and  canopy  cover.  Readers  of 
American  Birds  are  urged  to  follow  "A  Quantitative 
Method  of  Habitat  Description"  by  James  and  Shu- 
gart (1970).  Numerous  review  papers  on  bird-habitat 
structure  studies  are  included  in  Abbott  ( 1976), 
Anderson  ( 1979),  Anderson  and  Shugart  (1974), 
Balda  (1975),  Cody  (1981),  Hilden  ( 1965),  Holmes 
et  al.  (  1979),  Holmes  and  Robinson  (1981 ),  James 
(1971),  Karr  (1980),  Meents  et  al.  (1983),  Rice  et 
al.  (1983),  Rotenberry  and  Wiens  (1980a,  b),  Roth 
(1981),  Taylor  and  Littlefield  (1984),  Wiens  (1973, 
1983),  and  Willson  (1974).  Capen  (1981 )  discusses 
the  use  of  multivariate  statistics  in  describing  avian 
habitats.  Principal  component  analysis  and  discrimi- 
nate function  analysis  are  frequently  used  to  examine 
avian  habitat  selection  (Sedgwick  1981). 


DENSITY 

HEIGHT 

3  °® 

Mr 

9  wL             \, 

n 

DBH 

CANOPY  COVER 

Species  Composition.  Although  some  researchers 
have  recently  shown  floristics,  or  species  of  plants 
making  up  songbird  habitat,  to  be  important 
(Rotenberry,  pers.  commun.),  most  past  studies  have 
placed  more  importance  on  vegetative  structure. 
Notable  exceptions  are  the  close  dependence  of 
Kirtland's  warbler  on  young-growth  jack  pine  (Pinus 
banksiana )  and  the  red-cockaded  woodpecker 
(Picoides  borealis)  on  southern  pines  (P.  palustris, 
P.  elliottii,  P.  taeda,  and  P.  echinata )  that  are 
infected  with  red  heart  (Formes pint). 


296 


Songbirds 


STRUCTURE 


■ 


FLORISTICS 


Canopy  Cover  by  Type 


38%  Trees 
32%  Shrubs 
20%  Grass 

5%  Forbs 

5%  Bare  ground 


Species  Composition 


5%  Ponderosa  pine 

10%  True  mountain 
mahogany 

30%  Arizona  fescue 

30%  Bluebunch  wheatgrass 

25%  Other 


POPULATION  MEASUREMENT  TECHNIQUES 

Presence 

Sometimes  mere  lists  of  species  or  maps  show- 
ing species  distribution  are  adequate  for  making 
management  decisions  (for  example,  the  location  of 
red-cockaded  woodpecker  colonies  and  decisions 
to  log  or  not  log).  Most  state  ornithological  societies 
have  official  state  lists  of  birds  and  often  have  "offi- 
cial records"  committees  and  procedures  for  decid- 
ing the  validity  of  records.  Many  local  chapters  of 
the  National  Audubon  Society  have  checklists  of 
birds  for  local  areas,  as  do  various  land  use  agencies 
(e.g.,  BLM,  U.S.  Forest  Service,  U.S.  National  Park 
Service,  U.S.  Fish  and  Wildlife  Service,  and  state  con- 
servation agencies). 

Breeding  Bird  Atlases 

Breeding  bird  "atlasing"  has  become  a  popular 
pastime  among  bird  observers  in  Europe  during  the 
past  decade.  Atlases  showing  breeding  distribution 
maps  for  each  species  have  recently  been  published 
for  the  British  Isles,  France,  and  Denmark.  Similar 
atlases  are  in  progress  in  16  other  European  nations 
(Sharrock  1975),  Australia  (Serventy  1980),  and 
New  Zealand  (Gibb  1980). 

The  original  purpose  of  preparing  a  breeding 
bird  atlas  was  to  correlate  bird  distribution  with  that 
of  plants  as  shown  in  the  Atlas  of  British  Flora  (Per- 
ring  and  Walters  1962).  In  a  government-sponsored 
program  carried  out  through  the  British  Trust  for 


Western  tanager. 


Ornithology  and  the  Irish  Wildbird  Conservancy,  ob- 
servers visited  every  one  of  the  3,862  10-km  squares 
(  100  km"  [38.6.mi.~]  each)  of  land  throughout  the 
British  Isles  during  a  5-year  period  and  reported  the 
presence  or  absence  of  each  bird  species  (Sharrock 
1976).  Twelve  transparent  overlays,  ordered  sepa- 
rately, facilitate  correlation  of  bird  distribution  with 
selected  environmental  factors.  Sampling  blocks  have 
been  different  sizes  in  other  countries,  depending 
on  the  size  of  the  area  to  be  sampled  and  the  stand- 
ard maps  available.  In  France,  the  sampling  unit  was 
20  by  27  km  ( 12.4  by  16.8  mi.;  Yeatman  1976)  and 
in  Denmark  it  was  5-km  (3-1 -mi.)  square  (Dybbro 
1976).  Several  countries  modified  the  method  to 
include  some  indication  of  abundance  rather  than 
merely  presence  or  absence. 

No  large-scale  atlas  has  been  attempted  in  the 
U.S.  because  the  U.S.  Fish  and  Wildlife  Service's 
Breeding  Bird  Survey  (BBS)  provides  an  annual  sam- 
ple of  changing  abundance  of  each  species  and  also 
gives  a  density  of  coverage  roughly  comparable  with 
that  of  the  projected  European  atlas  of  the  1980s. 
Nevertheless,  atlas  studies  have  been  initiated  in 
several  states  (Laughlin  1982;  Laughlin  et  al.  1982). 
These  state  atlases  will  provide  information  on  pres- 
ence or  absence  of  the  various  species  in  many  for- 
ested areas.  In  the  Rocky  Mountains,  they  are  often 
termed  "Latilong"  studies,  a  word  originally  coined 
by  Skaar  ( 1969)  to  denote  1°  of  latitude  and  longi- 
tude blocks  (Figure  1 ). 

Although  western  state  atlases  use  grids  larger 
than  the  British  10-km  (6.1 -mi.)  model,  in  most  east- 
ern states  a  5-km  (3.1  -mi.)  grid  (six  blocks  on  a 
7  '/j-minute  topographic  map)  is  used.  The  Maryland 
Ornithological  Society,  however,  uses  2  '/2-km  ( 1 
l/2-mi.)  "quarter  blocks"  (about  600  ha  [1,500  a.]). 
The  quarter  blocks  make  it  possible  to  pinpoint  the 
location  of  rare  species  and  others  of  special  interest 
and  are  much  better  for  outlining  areas  where  a 


Songbirds 


297 


particular  species  is  not  present  (Klimkiewicz  and 
Robbins  1974).  This  accuracy  is  particularly  impor- 
tant where  commercial  or  residential  communities 
are  expanding  or  where  habitats  are  being  lost  to 
other  types  of  development.  In  areas  where  large  for- 
ests are  being  destroyed  by  changes  in  land  use, 
quarter-block  atlas  data  have  been  of  immense  value 
in  showing  the  degree  various  breeding  species  dis- 
appear when  forested  areas  are  fragmented  into 
smaller  tracts.  Raynor  (1983)  proposed  a  method  for 
evaluating  atlas  coverages. 


Christmas  Bird  Count  (CBC) 


The  CBC  is  the  best  known  and  probably  most 
used  source  of  information  on  geographical  distribu- 
tion of  nongame  birds  in  winter  (Table  3).  Started 
in  1900,  the  counts  now  involve  over  1,300  circles, 
each  24  km  (15  mi. )  in  diameter,  which  are  covered 
by  varying  numbers  of  birders  in  an  8-hour  period, 
sometime  between  December  20  and  January  2  each 
year.  Dawn-to-dusk  (or  longer)  counts  are  preferred. 


Cedar  Waxwing 

Breedinq 

Miqr. 

Winter 

W  b  b  R 

b  N  b   Habitat: 

R,Sb 
U 

R.Ag.PJ 
FC 

U,R,Ag 
Irr 

W   R  N 

W   b 

B  b  N 

W  N 

R   W 

N  W  W  Abundance: 

Figure  1.     Example  of  bird  distribution  by  Latilong  blocks  in  Colorado. 
298  Songbirds 


Counts  less  than  8  hours  long,  except  in  arctic  areas 
or  at  sea,  are  not  acceptable.  Results  are  summarized 
on  standardized  forms  which  include  entries  for 
weather,  habitat  description,  and  quality  of  the  food 
supply.  Summaries  are  published  annually  in  Ameri- 
can Birds  (Arbib  1978)  (Figure  2).  In  1980,  1,320 
counts  were  included,  covering  a  vast  area  from 
Newfoundland  to  Alaska,  south  to  Panama  and  Ha- 
waii. In  1982,  36  countries  and  6  continents  were 
involved  (Arbib  1982).  Popular  evaluations  of  the 
CBC  have  been  written  by  Robbins  ( 1966),  Bock 
(1979),  and  Wing  and  Jenks  (1939). 

Individual  birds  reported  on  CBCs  in  1979 
ranged  from  92  at  Churchill,  Manitoba,  to 
22,352,044  at  Pine  Prairie,  Louisiana.  Sixty-five 
counts  had  1 50  or  more  species.  The  most  meaning- 
ful data  are  expressed  in  number  of  birds  observed 
per  party-hour.  Other  bases,  such  as  birds  per  mile 
and  birds  per  observer,  have  proven  unreliable 
(Bock  1979). 

Winter  Bird  Population  Study 

In  1948,  the  National  Audubon  Society  inaugu- 
rated an  annual  Winter  Bird  Population  Study  which, 
like  the  CBC,  is  published  in  American  Birds.  The 
purpose  of  this  type  of  count  is  to  estimate  the  aver- 
age number  of  birds  using  a  particular  habitat.  Many 
observers  use  the  same  plot  in  which  they  have  con- 
ducted a  breeding  bird  census  by  spot-mapping. 
Most  plots  range  from  6  to  20  ha  (14.8  to  49.4  a.). 
Plots  are  visited  6  to  10  times  in  midwinter,  and  the 
totals  for  each  species  are  averaged  (Kolb  1965). 
Results  are  expressed  in  terms  of  birds  per  kilometer 
squared  [  1  km  ]  and  birds  per  40.5  ha  (100  a. ).  In 
1979-80,  92  counts  were  reported  in  American 
Birds  (Cink  and  Boyd  1981 ),  only  a  fraction  of  the 
1,300  Christmas  counts  reported  that  winter.  Several 
visits  per  winter  are  necessary  because  populations, 
and  even  the  species  present  in  a  given  plot,  vary 
from  day  to  day  and  from  morning  to  afternoon. 
Finches  (Carpodacus  sp. ),  robins  (Turdus  sp. ),  and 
waxwings  {Bombycilla  sp. ),  for  example,  often  range 
over  many  square  kilometers  in  the  course  of  2  or  3 
days.  Other  species,  in  mixed  flocks,  may  range  in 
and  out  of  a  study  plot.  Such  flocks  tend,  especially 
on  cold  days,  to  favor  sunny  exposures,  concentrate 
in  better  feeding  sites,  and  avoid  windy  areas 
(Robbins  1978). 

Difficulties  notwithstanding,  the  Winter  Bird 
Population  Study  does  enable  one  to  compare  popu- 
lations of  different  habitats  and,  to  a  lesser  degree, 
to  follow  population  trends  over  a  period  of  years. 
Webster  ( 1966)  analyzed  the  results  of  248  winter 
studies  in  forest  habitats  and  25  studies  in  grasslands. 
By  plotting  species  richness  against  population  den- 
sity, he  found  that  southern  pine  forests  mixed  with 
oaks  or  gums  tended  to  have  a  higher  species  rich- 
ness than  other  eastern  forests.  Webster,  however, 


Table  3.     Key  references  to  uses  of  Christmas  Bird 
Counts  in  North  America. 


Topic  Considered 

References 

Evaluation  of 
method 

Arbib  (1967),  Bock  &  Lep- 
thien  (1975b),  Burtt  &  Burtt 
(1982),  Bystrak(1974), 
Preston  (1958),  Tramer 
(1974) 

General  application 

Bock  (1979,  1984) 

Nuthatches 

Bock  &  Lepthien  (1972, 
1974,  1976d) 

Woodpeckers 

Bock  &  Lepthien  (1975a) 

Finches 

Bock  &  Lepthien  (1976a, 
1976b) 

Various  species 

Bock  &  Lepthien  (1972, 
1976b,  1976c),  Bocketal. 
(1977),  Bock&  Root 
(1981),  Bock&  Smith 
(1971) 

1176.  Aspen,  Colo.  39°15'N  106°54'W, 
center  n.e.  corner  Sec.  29,  R85W,  T9S,  as  de- 
scribed 1982;  elevation  7000  to  8200  ft;  habi- 
tat coverage:  residential  35%,  oak/ 
serviceberry/chokecherry  21%,  willow/ 
cottonwood  riverbottom  15%,  spruce  fir  12%, 
aspen  grove  9%,  sage/ rabbi tbrush  3%,  open 
meadow  3%,  pinyon  juniper  2%. — Dec.  19;  8 
a.m.  to  5  p.m.  Clear.  Temp.  9°  to  46°F.  Wind 
SE-W,  0-5  m.p.h.  Snow  cover  0  to  24  in.  Wa- 
ter mostly  frozen.  Wild  food  crop  excellent. 
Nineteen  observers,  11-14  in  8  parties,  5-8  at 
feeders.  Total  party-hours,  37  (18  on  foot,  5 
by  car,  14  on  skis)  plus  19  hours  at  feeders; 
total  party-miles,  119.5  (13.5  on  foot,  93  by 
car,  13  on  skis). 

•  Canada  Goose  5 1 ;  Mallard  304;  Am.  Wi- 
geon  1;  Wood  Duck  1;  Goshawk  1;  Cooper's 
Hawk  2;  Great  Horned  Owl  1;  Com.  (Red- 
sh.)  Flicker  10;  Hairy  Woodpecker  5;  Downy 
Woodpecker  4;  Gray  Jay  4;  Steller's  Jay  47; 
Scrub  Jay  12;  Black-billed  Magpie  121;  Com. 
Raven  5;  Black-capped  Chickadee  116; 
Mountain  Chickadee  27;  White-breasted 
Nuthatch  7;  Brown  Creeper  1;  Dipper  7;  Am. 
Robin  121;  Townsend's  Solitaire  3;  Cedar 
Waxwing  67;  Starling  95;  House  Sparrow  356; 
Red-winged  Blackbird  54;  Evening  Grosbeak 
46;  Cassin's  Finch  37;  Pine  Grosbeak  6; 
(Gray-crowned)  Rosy  Finch  24;  (Brown- 
capped)  Rosy  Finch  4;  Pine  Siskin  1;  Rufous- 
sided  Towhee  1;  Dark-eyed  (Slate-col.)  Junco 
23;  Dark-eyed  (Oregon)  Junco  36;  Dark-eyed 
(Gray-headed)  Junco  13. 

Total,  33  species  (3  additional  races);  1615 
individuals.  (In  count  area  count  week  but  not 
seen  count  day:  Bald  Eagle  ad.,  Com.  Snipe, 


Blue  Jay  9 
Crow  1;  Bh 
tain  Chicka 
in  19;  Wat 
1761;  Hous 
128;  Red-u 
Blackbird  2 
Cowbird  1 
170;  Pine  S 
eyed  (Slate 
gon)  Junco 
Junco  1;  T 
Sparrow  25 
Total,  54 
individuals 
Barry  Knap 
W.  9th  PI 
Ward,  Judy 

1178.     I 

104°38*W,  < 
described  1 
habitat  covi 
19;  7:30  a.r 
to  60°F.  Wii 
to  6  in.  Wat 
fair.  Fifteer 
party-hours, 
party-miles, 
•  Coopei 
Rough-legg< 
Marsh  Hawl 
2;  Com.  Sni[ 
Owl  1;  Coi 
Woodpeckei 
Horned  Lar 
3;  Black-bill 
Com.  Crow 


Figure  2.     Example  of  winter  bird  population  study 
taken  from  American  Birds,  January/February 
1984. 


Songbirds 


299 


did  not  list  actual  species  involved  but  merely  cited 
17  years  of  data  published  originally  in  Audubon 
Field  Notes. 

This  method  can  be  used  to  obtain  indexes  of 
winter  use.  However,  because  winter  bird  popula- 
tions vary  enormously  from  year  to  year  in  any  loca- 
tion, a  minimum  of  2  years  is  required  for  a 
meaningful  study.  In  a  critique  of  this  method,  Rob- 
bins  (1972)  showed  that  in  forest  habitats  six  traps 
were  sufficient  to  obtain  a  stable  minimum  estimate 
of  the  total  wintering  bird  population,  but  that  at 
least  8  to  10  trips  were  required  to  obtain  such  esti- 
mates for  individual  species.  Brewer  (  1972)  and 
Engstrom  and  James  (1981)  also  evaluated  the 
technique. 

Spot-Mapping  Census 

The  spot-mapping  technique,  also  called  the  plot 
census  or  simply  mapping  census,  was  first  used  in 
North  America  by  Williams  (1936)  and  in  Sweden 
by  Enemar  ( 1959).  Basically,  this  count  involves  8  to 
10  census  trips  through  an  area  of  specified  size  and 
preferably  uniform  habitat,  which  has  been  surveyed 
and  mapped  with  a  grid  system  (Figure  3).  On  each 
visit,  the  position  of  each  bird  seen  or  heard  is  re- 
corded on  the  plot  map.  Kendeigh  ( 1944),  Lack 
(1937),  and  Udvardy  (1957)  gave  good  historical 
accounts,  described  the  method  in  detail,  and  in- 
cluded comprehensive  bibliographies.  An  important 
feature  of  spot-mapping  is  to  designate,  with  appro- 
priate symbols,  those  individual  birds  that  are  heard 
singing  at  the  same  time  (simultaneous  registra- 
tions). These,  in  conjunction  with  clusters  of  single 
registrations,  make  it  possible  to  outline  the  approxi- 
mate territorial  limits  of  each  male  bird  and  make 
reasonable  estimates  of  the  total  number  of  territo- 
rial males  of  each  species  present  in  the  area. 

Spot-mapping  has  been  widely  adopted  in 
Europe  and  North  America.  Procedures  have  been 
standardized  by  the  International  Bird  Census  Com- 
mittee (1969)  so  that  results  obtained  in  different 
countries  can  be  compared.  Spot-mapping  has  been 
widely  used  in  England  since  1962  to  monitor  bird 
population  changes  (Batten  and  Marchant  1977). 
The  British  Trust  for  Ornithology  refers  to  such 
counts  as  the  Common  Bird  Census  (their  CBS,  not 
to  be  confused  with  CBC,  Christmas  Bird  Count). 
A  quantitative  description  of  habitat  (James  and  Shu- 
gart  1970)  is  now  a  standard  feature  of  many  of  the 
breeding  bird  censuses  in  forest  habitats  published  in 
American  Birds. 

Results  from  over  200  spot-mapping  censuses 
were  published  in  American  Birds  in  1981  (Van 
Velzcn  1981 ).  Many  private  consultants  and  govern- 
ment biologists  regularly  use  the  spot-mapping  tech- 
nique, although  their  results  are  not  published  in 
American  Birds.  Dickson  (  1978)  compared  the  tech- 


nique with  winter  bird-population  count  techniques 
(also  referred  to  as  "mean  detections  per  count") 
and  Palmgren's  (1930)  "summation  method"  and 
concluded  that  spot-mapping  gave  the  highest  popu- 
lation estimates. 

Robbins  (1978)  considered  spot-mapping  to  be 
the  most  accurate  of  the  various  bird  census  meth- 
ods because  it — 


( 1 )  gives  the  greatest  opportunity  to  record  all 
species  breeding  in  the  area, 

(2)  most  closely  approximates  the  absolute  num- 
ber of  breeding  pairs, 

(  3  )    is  more  accurate  in  estimating  whether  the 
birds  recorded  are  inside  or  outside  plot 
boundaries,  and 

(4)    involves  less  observer  bias. 

The  chief  disadvantage  of  the  technique  is  the 
amount  of  time  required  to  set  up  the  plot  and  con- 
duct a  minimum  of  eight  census  trips. 

The  results  of  228  breeding  bird  censuses  con- 
ducted in  Canada  were  summarized  in  three  catalogs 
by  Erskine  (1971,  1972,  1976).  Censuses  published 
in  American  Birds  since  1937  are  on  magnetic  tape 
at  the  Migratory  Bird  and  Habitat  Research  Labora- 
tory, U.S.  Fish  and  Wildlife  Service.  They  provide 
a  valuable  resource  for  comparing  habitat,  bird  spe- 
cies, or  other  variables  included  in  the  computer 
record.  The  files  contain  information  on  1,000  plots 
(Robbins  1978). 


Ladder-backed  woodpecker. 


300 


Songbirds 


1A 

1B 

1C 

1D 

IE 

1F 

1G 

1H 

?A 

7" 

X 

t 

3A 

yV 

X 

t 

X 

\    / 

\   / 
\  / 

i  / 
I  / 
1/ 

X            * 

** 

XI 

4A 

\  X 

• 

• 
t- 

\     ^ 

X^ 

5A 

>w      * 

1 
\    'I 

— *y / x 

«" 

X        ^ 

6A 

/  * 

X 

* 

'  ^\ 

7A 

v      * 

* 

•* 

• 

8A 

•  ^^^ 

Figure  3.     Spot-map  for  the  yellow-rumped  warbler  (Dendroica  coronata)  in  a  mixed-coniferous  forest. 
(Blocks  are  50  m  by  50  m  producing  a  grid  plot  of  12.25  ha.  Total  breeding  bird  density  is  3  territories 
x  2  =  6  breeding  yellow-rumped  warblers  per  12.25  ha  or  a  density  of  19.6/40  ha  [19.6/  100a.]). 


Songbirds 


301 


TRANSECT  LINE 


•  =  detected 
o  =  undetected 


TRANSECT  LINE 


Figure  4.     In  line  transect  sampling,  the  observer 
records  either  the  right  angle  distance  (x),  or 
both  the  flushing  distance  (r)  and  the  sighting 
angle  (O)  for  each  bird  detected.  Bird  density  is 
estimated  mathematically  using  the  distribution 
of  distances  (x  or  r). 


Figure  5.     In  belt  transect  sampling,  all  birds  are 
assumed  to  be  detected  within  a  specified,  fixed 
distance  (w)  from  the  transect  line  of  length  L. 
Birds  outside  the  belt  are  not  recorded.  Bird 
density  =  n/2Lw,  where  n  is  the  number  of 
birds  recorded. 


Transects 

Transect  methods  (Figures  4  and  5)  involve 
counting  birds  on  one  or  both  sides  of  a  line 
through  one  or  more  habitats.  Usually  either  the 
width  of  the  transect  is  defined  or  the  distance  to 
each  bird  encountered  is  estimated.  The  transect 
method  was  first  used  extensively  in  the  U.S.  in 
1906-09  by  S.A.  Forbes  and  A.O.  Gross  (Graber  and 
Graber  1963)  and  in  Finland  in  1941-56  by  Merikal- 
lio  (1958).  In  both  studies,  the  transect  results  were 
used  to  estimate  total  populations  by  habitat  for  an 
entire  state  or  nation.  The  Forbes  and  Gross  study 
was  later  repeated  by  Graber  and  Graber  ( 1963)  to 
show  bird  population  changes  over  a  50-year  period 
in  Illinois. 

One  advantage  of  the  transect  method  is  that  a 
relatively  large  area  can  be  sampled  in  a  short  time. 
Flack  (1976),  for  example,  used  45-m  (150-ft)-wide 
and  1,097-m  (l,200-yd)-long  transects  to  compare 
breeding  bird  populations  in  4 1  aspen  forests  in 
9  western  states  and  provinces  in  1966-69. 

The  transect  method  may  be  used  throughout 
the  year,  but  the  results  are  less  accurate  outside  the 
breeding  season.  However,  the  chief  disadvantage  of 
the  transect  method  is  that  a  single  coverage  of  a 
transect  does  not  permit  a  good  estimate  of  the 
number  of  birds  missed.  Anderson  and  Pospahala 
(1970),  using  data  from  2,574  km  (1,600  mi.)  of 
transects,  generated  a  curvilinear  (quadratic)  equa- 


tion to  show  the  fraction  of  waterfowl  nests  missed 
at  various  distances  from  the  center  line  of  the  tran- 
sect. However,  they  pointed  out,  that  to  adequately 
correct  for  the  number  of  fixed  objects  missed,  one 
needs  a  large  sample  and  must  also  assume  that  all 
the  objects  closest  to  the  center  of  the  transect  are 
detected. 

Emlen  (1971)  also  considered  how  the  error  in 
the  transect  count  may  be  reduced  by  first  estimat- 
ing the  lateral  distance  to  each  bird  encountered  and 
then  deriving  a  coefficient  of  detectability.  He  as- 
sumed that  no  bird  close  to  the  observer  goes  unde- 
tected. In  actual  practice,  a  large  number  of  birds 
within  a  few  meters  of  the  transect  line  may  be  un- 
detected (Jarvinen  and  Vaisanen  1976),  especially  in 
a  mature  forest  habitat.  Even  during  the  height  of  the 
breeding  season,  there  are  enormous  differences  in 
the  singing  behavior  and  conspicuousness  of  birds.  A 
noisy,  active  species  such  as  the  tufted  titmouse 
(Parus  bicolor)  may  be  recorded  on  68%  of  the  vis- 
its if  within  50  m  (54.7  yd)  of  the  observer.  Species 
such  as  the  ruby-throated  hummingbird  (Archilochus 
colubris),  worm-eating  warbler  (Helmitheros  vermi- 
vorus),  and  even  the  American  redstart  (Setophaga 
ruticilla )  may  be  recorded  only  36%  to  39%  of  the 
time  (Stewart  et  al.  1952).  Emlen  (1977)  estimated 
the  number  of  unrecorded  males  of  common  species 
by  running  each  transect  five  times,  plotting  all  sing- 
ing birds  on  maps,  outlining  the  territory  of  each, 
and  determining  what  European  workers  call  the  "ef- 
fectivity"  of  a  single  trip  for  each  species.  He  then 


302 


Songbirds 


used  the  computed  effectivity  for  correcting  his 
breeding  season  transect  results. 

The  transect  method  is  effective  for  comparing 
the  abundance  of  a  given  species  in  two  or  more 
plots  of  similar  habitat.  Also,  unless  visibility  is 
strongly  influenced  by  the  structure  of  the  habitat, 
the  transect  method  may  be  used  to  compare  abun- 
dance of  a  given  species  from  one  habitat  to  another 
(Robbins  1978).  It  is  not,  however,  a  desirable 
method  for  comparing  abundance  of  two  species 
that  may  not  be  equally  conspicuous,  unless  appro- 
priate corrections  are  made,  species  by  species. 
These  corrections  can  be  made  by  taking  a  series  of 
transect  counts  through  plots  where  the  population 
has  been  estimated  by  other  methods  (Ferry  and 
Frochot  1970;  Enemar  and  Sjostrand  1970). 

Although  most  transect  workers  record  birds 
per  unit  of  distance  or  area — kilometer  (or  mile)  or 
square  kilometer  (or  100  acres) — a  few  have  pre- 
ferred to  use  units  of  time,  such  as  birds  per  10 
hours  (Colquhoun  1940).  Conner  and  Dickson 
(1980)  discussed  sampling  design  and  statistical 
treatment  of  data  gathered  by  transect  censusing. 

North  American  Breeding  Bird  Survey 
(BBS) 

The  North  American  Breeding  Bird  Survey 
(BBS)  (Figure  6)  was  developed  by  the  U.S.  Fish  and 


Figure  6.     Breeding  Bird  Survey  summaries  are 
published  in  American  Birds. 


Wildlife  Service  to  monitor  bird  population  changes 
in  North  America  over  a  period  of  years  (Bystrak  and 
Robbins  1978;  Erskine  1978;  Robbins  and  Van  Vel- 
zen  1967,  1969,  1974).  Each  survey  route  is  a  series 
of  50,  3-minute  point  counts  at  800-m  (V^-mi.)  inter- 
vals along  a  39.4-km  (24  V^-mi.)  roadside  transect 
selected  randomly.  Coverage  extends  from  half  an 
hour  before  sunrise  to  about  4  hours  after  sunrise, 
embracing  the  period  of  greatest  bird  activity.  At 
each  of  the  50  points  (counts),  all  birds  heard  and 
all  birds  seen  within  400  m  ( XA  mi. )  of  the  counting 
position  are  tallied.  The  BBS,  which  now  embraces 
the  populated  areas  of  Canada  and  all  of  the  U.S. 
except  Hawaii,  provides  an  annual  sample  from 
1,700  or  more  roadside  transects  (Robbins  1978). 

The  BBS  results  are  used  primarily  for  statistical 
analysis  of  population  changes  over  the  years  and 
for  mapping  relative  breeding  densities  throughout 
the  North  American  range  of  a  species.  A  1 5-year 
summary  of  the  BBS  results  is  presented  by  Robbins 
etal.  (1986). 

For  intensive  local  studies,  BBS  routes  can  be 
laid  out  in  a  non-random  way  so  that  all  or  most 
secondary  roads  within  the  area  of  interest  are  in- 
cluded in  a  sample.  The  route  is  termed  "mini-route" 
because  25  instead  of  50  stops  are  made.  By  using  a 


Songbirds 


303 


shorter  route,  observers  are  able  to  cover  the  route 
before  working  hours  in  the  morning.  Covering  each 
route  twice  (once  in  each  direction)  and  combining 
the  results  of  the  two  counts  eliminated  most  of 
the  difference  in  activity  resulting  from  time  of  day. 
D.  Bystrak  and  others  (Klimkiewicz  and  Solem  1974) 
used  mini-routes  to  map  relative  abundance  of 
breeding  birds  throughout  two  Maryland  counties  as 
part  of  a  Breeding  Bird  Atlas  program  for  these 
counties.  Although  the  mini-route  technique  was 
designed  for  roadside  use,  with  slight  modifications, 
it  could  be  used  to  cover  forested  areas  by  horse- 
back or  off-road  vehicles.  Such  application  would 
make  it  possible  to  map  distribution  of  breeding 
birds  over  a  wide  area  in  a  relatively  short  time.  Dif- 
ferences in  bird  populations  could  then  be  corre- 
lated with  differences  in  vegetation  obtained  from 
aerial  surveys  or  by  ground  survey  methods. 


The  Indexes  Ponctuels  d'Abondance  (IPA) 
or  Point  Count  Method 

The  point  count  or  IPA  method  was  developed 
in  France  by  Ferry  and  Frochot  ( 1970)  as  a  means  of 
obtaining  indexes  of  abundance  for  comparing  bird 
populations  of  different  habitats  (or  of  the  same  hab- 
itat in  different  locations)  during  the  breeding  sea- 
son (Robbins  1978). 

The  IPA  counts  by  the  French  ornithologists 
consist  of  establishing  a  network  of  points  regularly 
distributed  through  the  habitat  to  be  studied.  An 
observer  then  stands  at  each  designated  spot  for  20 
minutes  in  the  early  morning  in  good  weather  and 
notes  all  birds  seen  or  heard.  Each  spot  is  censused 
twice  during  the  breeding  season.  The  higher  of 
the  two  counts  of  pair  numbers  is  used  as  an  index 
of  abundance  for  each  species.  Each  singing  male, 
occupied  nest,  or  family  of  birds  out  of  the  nest 
counts  as  one  pair,  whereas  a  bird  merely  seen  or 
heard  calling  counts  as  half  a  pair.  The  efficiency  of  a 
20-minute  stop  seemed  satisfactory  to  the  French 
investigators  because  during  the  last  5  of  the  20 
minutes  only  3%  more  species  and  9%  more  individ- 
uals were  recorded  in  forest  habitat. 

In  Denmark,  Jorgensen  (  1974)  conducted  81 
IPA  censuses  on  eight  mornings  from  mid-May  to 
mid-June.  The  1 3  V2  hours  of  effective  field  work 
was  about  50%  less  than  would  have  been  needed 
for  covering  one  census  plot  by  the  spot-mapping 
method.  Jorgensen  compared  the  density  of  each 
species  in  different  habitats,  using  the  Mann-Whitney 
U-test.  He  concluded  that  the  IPA  method  was  well- 
suited  to  a  study  of  forest  succession  in  which  statis- 
tical comparisons  are  necessary  or  desirable.  He 
summarized  habitat  use  (based  on  15  to  18  counts 
in  each  habitat )  in  terms  of  a  list  of  dominant  spe- 


cies, each  making  up  5%  or  more  of  the  registra- 
tions, and  subdominants  (2%  to  5%  ).  Then,  using 
only  the  dominant  species,  he  computed  similarity 
indexes  for  the  various  habitats  using  the  formula: 


S  = 


2c 


(a  +  b) 

Where. 

S  =  the  index. 

a  and  b  =  the  numbers  of  species  in  each  sample. 

c  =  the  number  of  species  common  to  the 
two  samples. 

Density-Frequency  Relationship.  Blondel  (1975) 
introduced  a  further  modification  of  the  IPA  method, 
Echantillonnage  Frequentiel  Progressif  (EFP).  The 
EFP  method  uses  the  presence  or  absence  of  a 
species  on  each  of  the  20-minute  IPA  counts  to 
determine  frequency  of  that  species  in  each  plot. 
Comparison  of  the  IPA  with  EFP  figures  allows  one 
to  determine  for  each  species  the  relationship 
between  its  density  and  frequency.  The  frequency  of 
a  species  is  closely  correlated  with  the  logarithm  of 
its  density;  the  lower  the  frequency,  the  better  the 
correlation.  Rotenberry  and  Wiens  (1976)  found 
a  similar  correlation  between  density  and  frequency, 
using  BBS  roadside  transect  data. 

Blondel  claimed  that  the  EFP  method,  which  is 
highly  standardized,  is  very  useful  for  a  rigorous 
statistical  interpretation  of  data.  He  used  the  EFP 
method  to  calculate  ecological  profiles  and  niche 
breadth  for  each  species  and  analyzed  the  structure 
of  bird  communities  according  to  the  structure  of 
the  vegetation.  For  each  community,  he  determined 
the  species  richness,  the  species  diversity  index  (H'), 
the  equitability  (J'),  and  the  level  of  fit  to  Galton's 
log-normal  model.  He  also  discussed  the  influence  of 
reforestation  on  bird  communities.  Blondel  con- 
cluded that  the  EFP  method  is  "very  well  adapted  to 
solve  problems  of  theoretical  and  applied  ecology 
at  the  community  level,  and  can  be  used  fruitfully 
for  environmental  monitoring"  (Robbins  1978).  Un- 
fortunately, the  method  has  not  been  tested  in  the 
West  but  seems  worthy  of  consideration. 

Comparison  with  Spot-Mapping.  In  comparing 
point  counts  (IPA)  with  mapping  census  in  the 
Bialowieza  Forest  in  Poland,  Tomialojc  et  al.  (1978) 
found  that  point  counts  overestimated  the 
population  when  the  density  was  low  and 
underestimated  it  when  the  density  was  high.  IPA 
counts  also  require  better  trained  observers  and 
involve  more  problems  in  the  separation  of  migrants 
or  other  non-breeding  birds  from  breeding 
individuals  than  mapping  censuses. 


304 


Songbirds 


Transect  and  Point  Count  Combination.  Bond 
(1957)  used  a  method  that  was  essentially  a 
combination  of  the  transect  and  point  count 
methods  to  compare  bird  populations  in  64  upland 
hardwood  stands  in  Wisconsin.  After  entering  a 
woodland,  he  walked  about  50  m  (55  yds)  along  a 
transect  line.  At  this  point,  he  stopped  for  5  minutes 
and  counted  all  birds  seen  or  heard  ahead  of  him.  He 
then  walked  ahead  slowly  for  5  minutes,  averaging 
150  to  175  m  (164  to  191  yds).  He  repeated  this 
procedure  until  he  had  made  five  10-minute  counts 
from  each  forest  interior.  Two  early  morning  visits 
were  made  to  each  woodlot,  and  the  highest  count 
for  each  species  was  used.  In  these  counts,  he 
detected  76% ,  78% ,  and  70%  as  many  pairs  as  he 
found  by  spot-mapping  censuses  in  three  of  the  same 
woodlands.  Palmgren  (1930)  and  Kendeigh  (1944) 
found  that  81%  and  63%  of  their  birds  in  spot- 
mapping  censuses  were  detected  on  the  first  two 
visits. 


Winter  Transects 

The  Maryland  Ornithological  Society  (MOS) 
devised  a  winter  bird  survey  in  central  Maryland, 
using  walked  transects  (Robbins  1970).  Preliminary 
tests  showed  that  the  winter  bird  survey  method  was 
not  practical  because  of  heavy  traffic  on  many  roads 
during  the  first  few  hours  after  sunrise.  Also,  because 
the  birds  were  not  singing,  those  that  could  not  be 
seen  from  the  roadside  could  not  be  detected.  The 
walked  transects,  away  from  roads,  would  reduce  the 
problem  caused  by  traffic  noises. 

The  winter  bird  survey  method  involves  tran- 
sects of  8  km  (5  mi. )  that  are  covered  on  foot  dur- 
ing the  first  4  hours  after  sunrise.  One  route  was 
established  at  the  center  of  each  7  '/2-minute  U.S. 
Geological  Survey  map  of  central  Maryland,  repre- 
senting an  1 1-  by  14-km  (6.8-  by  8.7-mi. )  grid.  An  ef- 
fort was  made  to  lay  out  each  route  in  the  form  of  a 
square,  2  km  (1.2  mi.)  to  a  side.  Many  routes  could 
not  be  formed  into  a  square  because  of  streams, 
ponds,  buildings,  high  fences,  or  other  obstacles;  de- 
spite changes  in  shape,  the  total  length  of  8  km  (  5 
mi. )  was  maintained.  By  timing  their  walking  speed 
for  the  first  quarter  of  the  route,  observers  were  able 
to  return  to  the  starting  point  within  a  few  minutes 
of  the  prescribed  4-hour  period.  Separate  counts 
were  kept  for  each  hour,  and  birds  identified  at  a  dis- 
tance greater  than  402  m  ( 'A  mi.)  were  recorded  in 
a  separate  column  on  the  form  (Robbins  1978).  The 
winter  survey  should  be  applicable  for  key  wintering 
areas  in  the  Southwest. 

Variable-Circular  Plots 


in  studies  of  endangered  species  in  dense  Hawaiian 
forests  (Ramsey  and  Scott  1979;  Scott  and  Ramsey 
1981;  Scott  et  al.  1981),  oak-pine  woodlands  (Verner 
and  Ritter  1985),  ponderosa  pine  forest  (Szaro  and 
Balda  1982),  and  desert  habitats  (Szaro  and  Jakle 
1982).  De  Sante  (1981)  found  that  the  variable  cir- 
cular plot  underestimated  the  densities  of  breeding 
birds  in  California  coastal  scrub  by  2%  to  70%  (de- 
pending on  the  species  of  bird )  when  compared 
with  estimates  derived  from  a  combination  of  color 
banding,  spot-mapping,  and  nest  monitoring.  De 
Sante  (1985)  found  similar  discrepancies  in  lodge- 
pole  (Pinus  contorta)  populations  of  birds  in  the 
Sierras. 

For  many  species,  tape  recordings  can  be  played 
back  to  induce  songs  from  silent  territorial  males 
(Down  1970).  The  technique  may  increase  census 
accuracy  (especially  for  species  with  low  song  activ- 
ity) with  little  expenditure  of  time  and  can  be  used 
to  determine  territorial  boundaries.  Repeated  use 
of  tape  recordings  during  the  breeding  season,  how- 
ever, can  bias  results  because  birds  may  alter  their 
habits  or  their  territorial  boundaries  if  they  believe  a 
competing  member  of  the  same  species  has  a  terri- 
tory nearby  (Robbins  1978;  McNicholl  1981).  Re- 
cordings have  proven  useful  in  censusing  various 
songbirds  (Falls  1981;  Johnson  et  al.  1981;  Marion  et 
al.  1981).  Although  recordings  are  probably  most 
useful  for  non-songbirds  such  as  rails  and  owls,  they 
have  been  used  successfully  to  monitor  Kirtland 
and  golden-cheeked  warblers  as  well  as  trogons. 


OBSERVATION   POINT 


•    BIRDS  PRESENT 
'    BIRDS   OBSERVED 


Reynolds  et  al.  (1980)  described  a  count,  using 
a  combination  of  a  transect  and  circular  plots  (Fig- 
ure 7).  This  count  method  has  been  widely  used 


Figure  7.     Variable  circular-plot  method  is  based  on 
principles  of  the  variable-strip  transect  method. 


Songbirds 


305 


Mark  and  Recapture  Methods 

Several  investigators  in  the  U.S.,  Sweden,  and 
France  have  marked  and  recaptured  birds  as  a  means 
of  studying  the  effectiveness  of  other  census  meth- 
ods (De  Sante  1981,  1985;  Robbins  1978).  Capture, 
mark,  and  recapture  methods  are  often  called  Lin- 
coln or  Peterson  indexes  (Davis  and  Winstead 
1980). 

Trapping  and  banding  in  itself  is  neither  an  effi- 
cient nor  a  highly  accurate  way  of  measuring  entire 
breeding  bird  populations.  It  is,  however,  a  very 
effective  way  in  determining  how  many  pairs  of  cer- 
tain species  are  present  and  in  distinguishing  mi- 
grants from  summer  residents  and,  to  some  degree, 
non-breeding  from  breeding  individuals.  Color-band- 
ing or  radiotelemetry  can  be  used  to  define  the 
ranges  of  individual  birds  and  also  can  point  out 
errors  in  judgment  that  occasionally  occur  when  the 
observer  relies  entirely  on  the  mapping  method 
(Robbins  1978). 

In  France,  Frochot  et  al.  (1978)  compared  spot- 
mapping,  IPA,  and  capture-recapture  methods  in  a 
100-ha  (247-a.)  oak  forest  plot.  Comparing  the  re- 
sults for  1 2  common  species,  he  estimated  49. 9 
breeding  pairs  per  10  ha  (24.7  a.)  by  the  mapping 
method  and  47.0  by  the  IPA  method  (with  appropri- 
ate corrections  for  conspicuousness ).  Banding  data 


sufficient  for  computation  of  population  estimates 
were  available  for  only  4  of  the  1 2  species,  but  for  3 
of  these  4,  the  estimate  from  capture-recapture  was 
higher  than  obtained  by  either  of  the  other  two 
methods.  Frochot  et  al.  (1978)  reported  that  while 
the  IPA  census  required  less  than  10  hours  of  prime 
time  in  the  early  morning,  the  mapping  census  re- 
quired 43  hours  and  the  banding  study,  400.  They 
emphasized  three  major  advantages  of  the  capture- 
recapture  method: 

( 1 )  It  permits  a  census  of  females  and  young  as 
well  as  of  singing  males; 

(2)  It  can  be  used  for  testing  mapping  and  IPA 
methods;  and 

(3)  It  gives  additional  information  about  daily 
range  and  habitat  use  of  the  individual  birds. 

In  Colorado,  Porter  (1973)  also  used  capture-recap- 
ture methods  to  check  other  census  methods  in 
grasslands. 

Hewitt  (1967)  devised  a  roadside  count  to  esti- 
mate breeding  populations  of  red-winged  blackbirds 
(Agelaius  phoeniceus),  which  basically  is  a  Lincoln 
Index  without  having  to  capture  and  mark.  It  should 
work  well  for  other  blackbirds  found  in  western 
marshes. 


^^ 


Western  bluebird. 


Red-breasted  nuthatch. 


306 


Songbirds 


Nest  Monitoring 

Perhaps  nest  monitoring  would  seem  an  obvious 
way  to  accurately  determine  breeding  bird  popula- 
tions. In  actual  practice,  however,  it  is  seldom  possi- 
ble to  find  enough  active  nests  of  a  species  to  use 
this  method  for  measuring  the  breeding  population, 
especially  in  forest  habitats. 

A  good  example  of  a  nesting  study  of  a  single 
species  is  an  investigation  of  wood  thrushes  (Hylo- 
cichla  mustelina)  made  by  Longcore  and  Jones 
(1969)  in  a  14.4  ha  (35.6  a.)  Delaware  woodlot. 
Systematic  nest  searches  were  conducted  during  a 
3-month  period  in  1965  and  1966.  Grid  lines  located 
at  45.7-m  (150-ft)  intervals  were  traversed  at  least 
once  every  3  days,  except  for  a  2 -week  period  in 
July  1965.  A  total  of  142  wood  thrush  nesting  at- 
tempts were  documented,  of  which  54  (  38%  )  were 
successful.  Nest  finding  was  supplemented  by  inten- 
sive banding  efforts  every  2  to  3  weeks  during  the 
1966  breeding  season,  which  resulted  in  the  banding 
of  46  wood  thrushes.  The  aim  of  this  study  was  to 
determine  reproductive  success  rather  than  to  meas- 
ure the  population,  but  it  gives  an  idea  of  how  much 
effort  would  be  required  to  gather  enough  informa- 
tion about  a  single  species  to  get  an  accurate  meas- 
urement of  the  breeding  population.  Similar 
approaches  should  work  for  western  thrushes. 

De  Sante  (1981)  used  nest  monitoring  in  con- 
junction with  color-banding  and  spot-mapping  to 
estimate  breeding  densities  of  eight  species  of  Cali- 
fornia coastal  scrub  at  Point  Reyes  National  Seashore. 
De  Sante  (1985)  used  a  similar  approach  to  count 
individuals  of  19  species  of  breeding  birds  in  a  subal- 
pine  lodgepole  pine  forest  in  the  Sierras. 

Conclusion 

Many  techniques  are  available  to  inventory  and 
monitor  songbird  numbers  and  the  habitats  they 
utilize.  For  breeding  birds,  the  best  overall  method 
involves  spot-mapping  territorial  males,  but  consider- 
able time  and  effort  are  required.  Eight  to  10  early 
morning  counts  of  10-  to  20-ha  (24.7-  to  49.4-a.) 
plots  are  recommended.  Additional  time  is  required 
to  analyze  the  habitat.  Walked-line  transects  or  vari- 
able-circular plots  along  transects  can  also  be  used 
to  estimate  avian  populations.  These  usually  require 
fewer  coverages  than  spot-mapping,  but  greater  skill 
and  training  are  needed  to  estimate  distances  to 
birds  detected.  Playback  of  recorded  calls  can  be 
used  to  elicit  responses  from  more  elusive  species 
but  must  be  used  consistently  or  not  at  all.  Roadside 
counts,  such  as  the  North  American  Breeding  Bird 
Survey  (BBS),  can  sample  larger  areas  but  only  give 
indexes  to  bird  numbers.  BBS  results  can  be  corre- 
lated to  habitat  if  such  data  are  recorded  for  each 
stop.  A  variety  of  point-counts  and  capture,  mark, 
and  recapture  methods  are  available  but,  especially  if 


trapping  is  involved,  require  considerable  effort. 
The  results  of  Christmas  Bird  Counts  and  winter  bird 
population  studies  sponsored  by  the  National  Audu- 
bon Society  are  published  regularly  in  American 
Birds  and  are  valuable  sources  of  information  on 
winter  distribution  and  trends  in  bird  numbers.  Ta- 
ble 2  summarizes  the  comparative  advantages  and 
disadvantages  of  several  monitoring  methods. 

Rather  than  attempting  to  monitor  all  species  of 
birds  in  many  different  habitats,  one  might  want  to 
concentrate  on  just  a  few  species  in  selected  habi- 
tats. Often  threatened  or  endangered  species  in  "crit- 
ical" habitats  are  stressed.  Szaro  and  Balda  (1982) 
recommended  selecting  bird  species  for  use  as  "indi- 
cator" species  by  considering: 

(1)    residency  status  (i.e.,  summer,  winter,  or 
permanent  residence), 

(  2  )  foraging  and/or  nesting  substrate, 

(3)  adequate  data  base, 

(4)  ease  of  monitoring, 

(5)  sensitivity  to  habitat  perturbations,. 

(6)  sensitivity  to  environmental  fluctuations, 

(7)  the  condition  or  range  of  conditions  a  given 
species  will  indicate,  and 

(8)  biogeographic  considerations  based  on  frag- 
mentation of  the  habitat. 

Graul  et  al.  (1976)  and  Graul  and  Miller  (1984) 
recommended  an  ecological-indicator  approach  for 
management,  where  the  species  in  an  area  having 
the  most  exacting  ecological  requirements  are  moni- 
tored. This  "steno"  (as  opposed  to  "eury")  species 
approach  is  based  on  the  idea  that  by  managing  to 
create  habitat  conditions  for  the  species  with  the 
narrowest  requirements,  a  manager  will  provide  for 
all  species.  Other  related  approaches  have  stressed  a 
featured-species  concept  where  a  single  species, 
not  necessarily  stenotypic,  is  selected,  and  a  unit  of 
habitat  managed  primarily  to  benefit  or  feature  that 
species.  Graul  and  Miller  (1984)  reviewed  still  other 
ecosystem  management  approaches  in  which  song- 
birds might  be  emphasized.  An  example  of  a  steno- 
typic songbird  that  might  be  monitored  as  an  early 
warning  of  excessive  human  disruption  of  its  envi- 
ronment would  be  the  loggerhead  shrike  {Laniiis 
ludovicianus;  Bystrak  1983).  The  species  has  dem- 
onstrated a  considerable  population  decline,  which 
suggests  a  narrow  range  of  tolerance,  thus  qualifying 
it  as  an  indicator  species.  The  BBS  provides  good 
population  trends  for  loggerhead  shrike  (Robbins  et 
al.,  in  press). 


Songbirds 


307 


LITERATURE  CITED 


ABBOTT,  I.  1976.  Comparisons  of  habitat  structure  and 
plant,  arthropod,  and  bird  diversity  between  mainland 
and  island  sites  near  Perth,  Western  Australia.  Austra- 
lian J.  Ecol.  1:275-280. 

ANDERSON,  AH.  and  A.  ANDERSON.  1973-  The  cactus 
wren.  Univ.  Arizona  Press,  Tucson.  226pp. 

ANDERSON,  DR.  and  R.S.  POSPAHALA.  1970.  Correction 
of  bias  in  belt  transect  studies  of  immobile  objects. 
J.  Wildl.  Manage.  34:141-146. 

ANDERSON,  S.H.  1979.  Habitat  structure,  succession  and 
bird  communities.  Pages  9-21  in  Proceedings  of 
Workshop:  Management  of  North  Central  and  North- 
eastern Forests  for  Nongame  Birds,  U.S.  Dep.  Agric, 
For.  Serv.  Gen.  Tech.  Rep.  NC-51.  Minneapolis,  MN. 

and  H.H.  SHUGART,  Jr.  1974.  Habitat  selection 

of  breeding  birds  in  an  east  Tennessee  deciduous  for- 
est. Ecology  55:828-837. 

ARBIB,  R.S.,  Jr.  1967.  Considering  the  Christmas  count. 
Audubon  Field  Notes  21(l):39-42. 

.  1978.  The  National  Audubon  Society  and  amateur 

ornithologists.  Pages  26-29  in  McCrimmon, Jr.,  DA. 
and  A.  Sprunt,  IV,  eds.  Proceedings  of  a  Conference 
on  the  Amateur  and  North  American  Ornithology. 
Natl.  Audubon  Soc.  Res.  Dep.  Ithaca,  NY.  80pp. 

.  1982.  The  Christmas  bird  count  and  the  world. 


Am.  Birds  36:365-368. 

ARMSTRONG,  E.A.  1955.  The  wren.  Macmillan  Co.  New 
York,  NY.  312pp. 

AUSTIN,  O.L.,  Jr.,  ed.  1968.  Life  histories  of  North  Ameri- 
can cardinals,  grosbeaks,  buntings,  towhees,  finches, 
sparrows,  and  allies.  U.S.  Natl.  Mus.  Bull.  237,  3  vols. 
1889pp. 

BAILEY,  A.M.  and  R.J.  NIEDRACH.  1965.  Birds  of  Colo- 
rado. Denver  Mus.  Nat.  Hist.,  2  vols.  895pp. 

BAKUS,  G.J.  1959a.  Observations  on  the  life  history  of  the 
dipper  in  Montana.  Auk  76:190-207. 

.  1959b.  Territoriality,  movements,  and  population 

density  of  the  dipper  in  Montana.  Condor  61:410-425. 

BALDA,  R.P.  1975.  Vegetation  structure  and  breeding 
bird  diversity.  Pages  59-80  in  Smith,  DR.,  ed.  Pro- 
ceedings of  the  Symposium  on  Management  of  Forest 
and  Range  Habitats  for  Nongame  Birds.  U.S.  Dep. 
Agric,  For.  Serv.,  Gen.  Tech.  Rep.  WO-1,  Washington, 
DC. 

and  G.C.  BATEMAN.  1972.  The  breeding  biology  of 

the  pinon  jay.  Living  Bird  1 1:5-42. 

BARLOW,  J.C.  1962.  Natural  history  of  the  Bell  vireo  ( Vi- 
reo  bellii)  Audubon.  Univ.  Kansas  Publ.,  Mus.  Nat. 
Hist.  12:241-296. 

BATTEN,  LA.  and  J.H.  MARCHANT.  1977.  Bird  population 
changes  for  the  years  1975-76.  Bird  Study  24:159- 
164. 

BEASON,  R.C.  1974.  Breeding  behavior  of  the  horned  lark. 
Auk  91:65-74. 

and  E.C.  FRANKS.  1973-  Development  of  young 

horned  larks.  Auk  90:359-363. 

BENT,  AC.  1939.  Life  histories  of  North  American  wood- 
peckers. U.S.  Natl.  Mus.  Bull.  174:1-334. 

.  1940.  Life  histories  of  North  American  cuckoos, 

goatsuckers,  hummingbirds,  and  their  allies.  U.S.  Natl. 
Mus.  Bull.  176:1-506. 


— .  1942.  Life  histories  of  North  American  flycatchers, 

larks,  swallows,  and  their  allies.  U.S.  Natl.  Mus.  Bull. 

179:1-555. 
— .  1946.  Life  histories  of  North  American  jays,  crows, 

and  titmice.  U.S.  Natl.  Mus.  Bull.  191:1-495. 
— .  1948.  Life  histories  of  North  American  nuthatches, 

wrens,  thrashers,  and  their  allies.  U.S.  Natl.  Mus.  Bull. 

195:1-475. 
— .  1949.  Life  histories  of  North  American  thrushes, 

kinglets,  and  their  allies.  U.S.  Natl.  Mus.  Bull.  196:1- 

454. 
— .  1950.  Life  histories  of  North  American  wagtails, 

shrikes,  vireos,  and  their  allies.  U.S.  Natl.  Mus.  Bull. 

197:25-38. 
— .  1953-  Life  histories  of  North  American  wood  war- 
blers. U.S.  Natl.  Mus.  Bull.  203:1-734. 
— .  1958.  Life  histories  of  North  American  blackbirds, 


orioles,  tanagers,  and  allies.  U.S.  Natl.  Mus.  Bull.  211:1- 
549. 

BLONDEL,  J.  1975.  L'analyses  des  peuplements  d'oiseaux, 
elements  d'un  diagnostic  ecologique.  La  Terre  et  la 
Vie.  29:533-589. 

BOCK,  C.E.  1970.  The  ecology  and  behavior  of  the  Lewis 
woodpecker  (Asyndesmus  lewis).  Univ.  California 
Publ.  in  Zoology  92:1-100. 

.  1979.  Christmas  bird  count.  Nat.  Hist.  88(10):7-11. 

.  1 984.  Geographical  correlates  of  abundance  versus 

rarity  in  some  North  American  winter  landbirds.  Auk 
101:266-273- 

,  J.H.  BOCK,  and  L.W.  LEPTHIEN.  1977.  Abundance 

patterns  of  some  bird  species  wintering  on  the  Great 
Plains  of  the  U.S.A.  J.  Biogeography  4:101-110. 

and  L.W.  LEPTHIEN.  1972.  Winter  eruptions  of  red- 
breasted  nuthatches  in  North  America,  1950-1970. 
Am.  Birds  26:558-561. 

and .  1975a.  A  Christmas  count  analysis  of 

woodpecker  abundance  in  the  United  States.  Wilson 
Bull.  87:355-366. 

and .  1975b.  Patterns  of  bird  species  diver- 
sity revealed  by  Christmas  counts  versus  breeding 
bird  surveys.  Western  Birds  6.95-100. 

and .  1976a.  Growth  in  the  eastern  house 

finch  population,  1962-1971.  Am.  Birds  30:791-792. 

and .  1976b.  Synchronous  eruptions  of  bo- 
real seed-eating  birds.  Am.  Nat.  110:559-571. 

and .  1976c.  Geographical  ecology  of  the 


common  species  of  Buteo  and  Parabuteo  wintering  in 
North  America.  Condor  78:554-557. 
—  and  T.L.  ROOT.  1981.  Winter  abundance  patterns 
of  landbirds  in  the  U.S.  and  southern  Canada.  Am. 
Birds  35:891-897. 

and  R.B.  SMITH.  1971.  An  analysis  of  Colorado 


Christmas  counts.  Am.  Birds  25:945-947. 

BOND,  R.R.  1957.  Ecological  distribution  of  breeding  birds 
in  the  upland  forests  of  southern  Wisconsin.  Ecol. 
Monogr.  27:351-384. 

BOYD,  R.L.  1976.  Behavioral  biology  and  energy  expendi- 
ture in  a  horned  lark  population.  Ph.D.  dissertation, 
Colorado  State  Univ.,  Fort  Collins.  194pp. 

BRAZIER,  F.H.  1964.  Status  of  the  mockingbird  in  the 
northern  Great  Plains.  Blue  Jay  22:63-75. 

BREWER,  R.  1972.  An  evaluation  of  winter  bird  popula- 
tion studies.  Wilson  Bull.  84:261-277. 

BURTT,  HE.  and  B.P.  BURTT.  1982.  Reliability  of  the 
Christmas  bird  count.  Redstart  49:90-93- 

BYSTRAK,  D.,  ed.  1974.  Wintering  areas  of  bird  species 
potentially  hazardous  to  aircraft.  Nad.  Audubon  Soc,  NY. 


308 


Songbirds 


— .  1983.  Loggerhead  shrike  (Lanius  ludovicianm). 
Pages  301-310  in  Armbruster,  J.S.,  ed.  Impacts  of  Coal 
Surface  Mining  on  25  Migratory  Bird  Species  of  High 
Federal  Interest.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv. 
FWS/OBS-83/35.  348pp. 

and  C.S.  ROBBINS.  1978.  Bird  population  trends 


detected  by  the  North  American  breeding  bird  sur- 
vey. Polish  Ecol.  Studies  3(4):13M43. 

CACCAMISE,  D.F.  1974.  Competitive  relationships  of  the 
common  and  lesser  nighthawks.  Condor  76:1-20. 

CALDER,  W.A.  1973-  Microhabitat  selection  during  nesting 
of  hummingbirds  in  the  Rocky  Mountains.  Ecol. 
54:127-134. 

CALL,  M.W.  1981.  Terrestrial  wildlife  inventories:  Some 
methods  and  concepts.  U.S.  Dep.  Inter.,  Bur.  Land 
Manage.  Tech.  Note  349:1-171. 

CAPEN,  D.E.,  ed.  1981.  The  use  of  multivariate  statistics  in 
studies  of  wildlife  habitat.  U.S.  Dep.  Agric,  For.  Serv. 
Gen.  Tech.  Rep.  RM087.  249pp. 

CINK,  C.L.  and  R.L.  BOYD   1981.  Thirty^third  winter  bird- 
population  study.  Am.  Birds  35:21-45. 

CODY,  ML.  1981.  Habitat  selection  in  birds:  The  roles  of 
vegetation  structure,  competitors,  and  productivity. 
BioScience  31:107-113. 

COLQUHOUN,  M.K.  1940.  Visual  and  auditory  conspicu- 
ousness  in  a  woodland  bird  community:  A  quantita- 
tive analysis.  Proceedings,  Zoological  Soc.  London 
110:129-148. 

CONNER,  R.N.  and  J.G.  DICKSON.  1980.  Strip  transect 
sampling  and  analysis  for  avian  habitat  studies.  Wildl. 
Soc.  Bull.  8:4-10. 

CORNWELL,  G.W.  1963.  Observations  on  the  breeding 
biology  and  behavior  of  a  nesting  population  of 
belted  kingfishers.  Condor  65:426-431. 

DAVIS,  D.E.,  ed.  1982.  CRC  handbook  of  census  methods 
for  terrestrial  vertebrates.  CRC  Press,  Inc.  Boca  Raton, 
FL.  397pp. 

and  R.L.  WINSTEAD.  1980.  Estimating  the  numbers 

of  wildlife  populations.  Pages  221-245  in  Schemnitz, 
S.D.,  ed.  Wildlife  Management  Techniques  Manual. 
The  Wildlife  Soc.  Washington,  DC.  686pp. 

DEGRAFF,  R.M.,  tech.  coord.  1978a.  Proceedings  of  the 
workshop  on  management  of  southern  forests  for 
nongame  birds.  U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech. 
Rep.  SE-14. 

1978b.  Proceedings  of  workshop  on  nongame  bird 

management  in  the  coniferous  forests  of  the  western 
U.S.  U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep.  PNW- 
64. 

and  K.E.  EVANS,  compilers.  1979.  Management  of 

north-central  and  northeastern  forests  for  nongame 
birds.  U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep.  NC- 
51. 

and  N.G.  TILGHMAN,  eds.  1980.  Workshop  pro- 
ceedings, management  of  western  forests  and  grass- 
lands for  nongame  birds.  U.S.  Dep.  Agric,  For.  Serv. 
Gen.  Tech.  Rep.  INT-86. 

DEHAVEN,  R.W.  and  P.J.  DEHAVEN.  1973.  A  contribution 
toward  a  bibliography  on  the  starling.  U.S.  Dep.  Inter., 
Fish  and  Wildl.  Service.  Denver,  CO.  92pp. 

DESANTE,  D.F.  1981.  A  field  test  of  the  variable  circular- 
plot  censusing  technique  in  a  California  coastal  scrub 
breeding  bird  community.  Pages  177-185  in  Ralph, 
C.J.  and  V.M.  Scott,  eds.  Estimating  the  numbers  of 
terrestrial  birds.  Studies  Avian  Biol.  6. 

1985.  A  field  test  of  the  variable  circular-plot  cen- 
susing technique  in  a  Sierran  subalpine  forest  breed- 


ing bird  community.  Abstract  of  paper  presented  at 
fourth  joint  meeting  of  Wilson  and  Cooper  Ornithol- 
ogical Societies,  Univ.  Colorado,  Boulder,  June  4-9. 
Upp. 

DICKSON,  J.G.  1978.  Comparison  of  breeding  bird  census 
techniques.  Am.  Birds  32:10-13- 

DLXON,  K.L.  1961.  Habitat  distribution  and  niche  relation- 
ships in  North  American  species  of  Parus.  Pages  179- 
216  in  Blair,  W.F.,  ed.  Vertebrate  Speciation.  Univ. 
Texas  Press,  Austin. 

DOWN,  D.D.  1970.  Indexing  population  densities  of  the 
cardinal  with  tape-recorded  calls.  Wilson  Bull.  86:83- 
91. 

DYBBRO,  T.  1976.  De  danoke  gorglefulgles  udbredelse. 
Dansk  Ornith  Forening,  Copenhagen.  293pp. 

EIPPER,  AW.  1956.  Differences  in  vulnerability  of  the 
prey  of  nesting  kingfishers.  J.  Wildl.  Manage.  20:177- 
183- 

EMLEN,  J.T  1971.  Population  densities  of  birds  derived 
from  transect  counts.  Auk  88:323-342. 

.  1977.  Estimating  breeding  season  bird  densities 

from  transect  counts.  Auk  94:455-468. 

ENEMAR,  A.  1959.  On  the  determination  of  the  size  and 
composition  of  a  passerine  bird  population  during  the 
breeding  season.  Var  Fagelvard,  Supplement  2:1-114. 

and  B.  SJOSTRAND.  1970.  Bird  species  densities 

derived  from  study  area  investigations  and  line  tran- 
sects. Swedish  Nat.  Sci.  Res.  Council,  Ecol.  Res. 
Comm.  Bull.  9:33-37. 

ENGSTROM,  R.T.  and  F.C.  JAMES.  1981.  Plot  size  as  a 
factor  in  winter  bird-population  studies.  Condor 
83:34-41. 

ERSKINE,  A.J.  1971.  A  preliminary  catalogue  of  bird  cen- 
sus plot  studies  in  Canada.  Canadian  Wildl.  Serv.  Prog- 
ress Notes  20:1-78. 

.  1972.  A  preliminary  catalogue  of  bird  census  plot 

studies  in  Canada.  Part  2,  Canadian  Wildl.  Serv.  Prog- 
ress Notes  30:1-42. 

.  1976.  A  preliminary  catalogue  of  bird  census  plot 

studies  in  Canada.  Part  3,  Canadian  Wildl.  Serv.  Prog- 
ress Notes  59:1-24. 

.  1978.  The  first  10  years  of  the  cooperative  breed- 
ing bird  survey  in  Canada.  Canadian  Wildl.  Serv.  Rep., 
Serial  42:1-59. 

FALLS,  J. B.  1981.  Mapping  territories  with  playback:  An 
accurate  census  method  for  songbirds.  Pages  86-9 1  in 
Ralph,  C.J.  and  J.M.  Scott,  eds.  Estimating  the  numbers 
of  terrestrial  birds.  Studies  Avian  Biol.  6. 

FEARE,  C.  1984.  The  starling.  Oxford  Univ.  Press.  New 
York,  NY.  315pp. 

FERRY,  C.  and  B.  FROCHOT.  1970.  Lairfaune  nidofrea- 
trice  d'une  foret  de  chenes  pedoncules  en  bourgogne 
etude  de  deux  successions  ecologiques.  La  Terre  et 
la  Vie  24:153-250. 

FLACK,  J.A.D.  1976.  Bird  populations  of  aspen  forests  in 
western  North  America.  Ornith.  Monogr.  19.  97pp. 

FRANZEB,  K.E.  1977.  Inventory  techniques  for  sampling 
avian  populations.  U.S.  Dep.  Inter.,  Bur.  Land  Manage. 
Tech.  Note  307.  17pp. 

FROCHOT,  B.,  D.  REUDET,  and  Y  LERUTH.  1978.  A  com 
parison  of  three  different  methods  of  census  applied 
to  the  same  population  of  forest  birds.  Polish  Ecol. 
Studies  3(4>.71-76. 

GIBB,  J.A.  1980.  New  Zealand  ornithology  during  the  past 
50  years.  Bull.  British  Ornith.  Club  100:93-96. 

GOODWIN,  D.  1976.  Crows  of  the  world.  Cornell  Univ. 
Press.  Ithaca,  NY.  354pp. 


Songbirds 


309 


GRABER,  R.R.  1971.  Illinois  birds:  Turdidae.  Illinois  Nat. 

Hist.  Surv.  Biol.  Notes  75:1-44. 
.  1972.  Illinois  birds:  Hirundenidae.  Illinois  Nat. 

Hist.  Surv.  Biol.  Notes  80:1-36. 
.  1973-  Illinois  birds:  Laniidae.  Illinois  Nat.  Hist. 

Surv.  Biol.  Notes  83:1-18. 
.  1974.  Illinois  birds:  Tyrannidae.  Illinois  Nat.  Hist. 

Surv.  Biol.  Notes  86:1-56. 
and  J.W.  GRABER.  1963.  A  comparative  study  of 

bird  populations  in  Illinois,  1906-1909  and  1956- 

1958.  Illinois  Nat.  Hist.  Surv.  Bull.  28:377-528. 

-,  and  E.L.  KIRK.  1970.  Illinois  birds:  Mimi- 


del.  Illinois  Nat.  Hist.  Surv.  Biol.  Notes  68:1-38. 

GRABER,  V.W.,  R.R.  GRABER,  and  E.L.  KIRK  1983.  Illinois 
birds:  Wood  warblers.  Illinois  Nat.  Hist.  Surv.  Biol. 
Notes  118.  144pp. 

GRAUL,  WD.  and  G.C.  MILLER.  1984.  Strengthening  eco- 
system management  approaches.  Wildl.  Soc.  Bull. 
12:282-289- 

,  J.  TORRES,  and  R.  DENNEY.  1976.  A  species-eco- 
system approach  for  nongame  programs.  Wildl.  Soc. 
Bull.  4:79-80. 

GRISCOM,  L.  and  A.  SPRUNT,  Jr.  1957.  The  warblers  of 
America.  Devin-Adair  Co.  New  York,  NY.  356pp. 

HARRISON,  H.H.  1975.  A  field  guide  to  birds'  nests. 
Houghton  Mifflin  Co.,  Boston,  MA.  257pp. 

.  1984.  Wood  warbler's  world.  Simon  &  Schuster, 

New  York,  NY.  335pp. 

HEWITT,  D.H.  1967.  A  road-count  index  to  breeding  pop- 
ulations of  red-winged  blackbirds.  J.  Wildl.  Manage. 
31:39-47. 

HILDEN,  O.  1965.  Habitat  selection  in  birds.  Ann.  Zoology 
Fennici  2:53-75. 

HOLMES,  R.T.,  RE.  BONNEY,  and  S.W.  PACALA.  1979. 
Guild  structure  of  the  Hubbard  Brook  bird  commu- 
nity: A  multivariate  approach.  Ecol.  60:512-520. 

and  S.K  ROBINSON.  1981.  Tree  species  preference 

of  foraging  insectivorous  birds  in  a  northern  hard- 
woods forest.  Oecologia  48:31-35. 

INTERNATIONAL  BIRD  CENSUS  COMMITTEE.  1969 
Recommendations  for  an  international  standard  for  a 
mapping  method  in  bird  census  work.  Bird  Study 
16:249-255  (also  1970,  Audubon  Field  Notes  24:723- 
726.) 

JACKMAN,  S.M.  and  J.M.  SCOTT.  1975.  Literature  review 
of  23  selected  forest  birds  of  the  Pacific  Northwest. 
U.S.  Dep.  Agric,  For.  Serv.  Portland,  OR.  382pp. 

JAMES,  F.C.  1971.  Ordinations  of  habitat  relationships 
among  breeding  birds.  Wilson  Bull.  83:215-236. 

and  H.H.  SHUGART,  Jr.  1970.  A  quantitative 

method  of  habitat  description. 

JARVTNEN,  O.  and  R.A.  VAISANEN.  1976.  Finnish  line 
transect  censuses.  Ornis  Fenn  53:115-118. 

JOHNSGARD,  PA.  1983.  The  hummingbirds  of  North 
America.  Smithsonian  Inst.  Press,  Washington,  DC. 
303pp. 

JOHNSON,  R.R.,  B.T.  BROCON,  L.T.  HAIGHT,  and  J.M. 
SIMPSON.  1981.  Playback  recordings  as  a  special 
avian  censusing  technique.  Pages  68-75  in  Ralph,  C.J. 
and  J.M.  Scott,  eds.  Estimating  the  numbers  of  terres- 
trial birds.  Studies  Avian  Biol.  6. 

JORGENSEN,  OH.  1974.  Results  of  IPA-censuses  on  Dan- 
ish farmland.  Acta  Ornithologica  14:310-321. 

KARR,  JR.  1980.  History  of  the  habitat  concept  in  birds 
and  the  measurement  of  avian  habitats.  Pages  991-997 
in  Nohring,  R.,  ed.  Acta  XVII  Int.  Congress  Ornitholo- 
gica, Berlin. 


KENDEIGH,  S.C.  1944.  Measurement  of  bird  populations. 
Ecol.  Monogr.  14:67-106. 

KESSEL,  B.  1957.  A  study  of  the  breeding  biology  of  the 
European  starling  (Sturnus  vulgaris  L. )  in  North 
America.  Am.  Midi.  Nat.  58:257-331. 

KLIMKIEWICZ,  M.K  and  C.S.  ROBBINS.  1974.  The  breed- 
ing bird  atlas  of  Montgomery  County,  Maryland,  USA. 
Acta  Ornithologica  14:446-458. 

and  J.K  SOLEM.  1974.  First  year  of  breeding  bird 

atlas,  Howard  County,  Maryland.  Maryland  Birdlife 
30:27-35. 

KNORR,  O.A.  1961.  The  geographical  and  ecological  dis- 
tribution of  the  black  swift  in  Colorado.  Wilson  Bull. 
73:155-170. 

KOLB,  H.  1965.  The  Audubon  winter  bird-population 
study.  Audubon  Field  Notes  19:432-434. 

LACK,  D.  1937.  A  review  of  bird  census  work  and  bird 
population  problems.  Ibis  79:369-395. 

LAUGHLIN,  S.D.,  ed.  1982.  Proceedings  of  the  northeast- 
ern breeding  bird  atlas  conference.  Vermont  Inst, 
of  Nat.  Sci.,  Woodstock,  VT. 

,  D.F.  KIBBE,  and  P.F.J.  EAGLES.  1982.  Atlasing  the 

distribution  of  the  breeding  birds  of  North  America. 
Am.  Birds  35:6-19. 

LEA,  R.B.  1942.  A  study  of  the  nesting  habits  of  the  cedar 
waxwing.  Wilson  Bull.  54:225-237. 

LONGCORE, JR.  and  R.E.JONES.  1969.  Reproductive 
success  of  the  wood  thrush  in  a  Delaware  woodlot. 
Wilson  Bull.  81:396-406. 

MACARTHUR,  R.W.  and  J.W.  MACARTHUR.  1961.  On  bird 
species  diversity.  Ecol.  42:594-598. 

MARION,  W.R.,  T.E.  OMEARA,  and  D.S.  MAEHR.  1981.  Use 
of  playback  recordings  in  sampling  elusive  or  secre- 
tive birds.  Pages  81-85  in  Ralph,  C.J.  and  J.M.  Scott, 
eds.  Estimating  the  Numbers  of  Terrestrial  Birds. 
Studies  Avian  Biol.  6. 

MASER,  C,  J.W.  THOMAS,  and  R.G.  ANDERSON.  1984. 
The  relationship  of  terrestrial  vertebrates  to  plant 
communities  and  structural  conditions.  U.S.  Dep. 
Agric,  For.  Serv.  Gen.  Tech.  Rep.  PNW-172,  2  parts. 
237pp. 

MCCLELLAND,  B.R.  and  S.S.  FRISSELL.  1975.  Identifying 
forest  snags  for  hole-nesting  birds.  J.  Forestry  73:414- 
417. 

MCNICHOLL,  M.K  1981.  Caution  needed  in  use  of  play- 
backs to  census  bird  populations.  Am.  Birds  35:235- 
236. 

MEENTS, J.K, JR.  RICE,  B.W.  ANDERSON,  and  R.D. 
OHMART.  1983-  Nonlinear  relationships  between 
birds  and  vegetation.  Ecol.  64:1022-1027. 

MERIKALLIO,  E.  1958.  Finnish  birds,  their  distribution 
and  numbers.  Fauna  Fennica  5:1-181. 

MILLER,  A.H.  1931.  Systematic  revision  and  natural  history 
of  the  American  shrikes.  Univ.  California  Publ.  in 
Zoology  38:11-242. 

NERO,  R.W.  1984.  Redwings.  Smithsonian  Institute  Press, 
Washington,  DC.  160pp. 

NORRIS,  R.A.  1958.  Comparative  biosystematics  and  life 
history  of  the  nuthatches  {Sitta  pygmaoa  and  Sitta 
pusilla).  Univ.  California  Publ.  Zoology  56:1 19-300. 

OHMART,  R.D.  1973.  Observations  on  the  breeding  adap- 
tations of  the  roadrunner.  Condor  75:140-149. 

ORIANS,  G.H.  1980.  Some  adaptations  of  marsh-nesting 
blackbirds.  Princeton  Univ.  Press,  Princeton,  NJ. 
295pp. 

PACKARD,  G.C.  1966.  Evolution  of  North  American  house 
sparrows  in  relation  to  altitude  and  aridity.  Ph.D. 


310 


Songbirds 


1 


thesis,  Univ.  Kansas,  Lawrence.  86pp. 
PALMGREN,  P.  1930.  Quantitative  Untersuchesogen  uber 

die  Vogelfauna  in  den  Waldern  Sudfinnlands  mit  be- 

sonderer  Berucksichtigring  Alands.  Acta  Zod.  Fenn. 

7:1-218. 
PEERING,  F.H  and  S.M.  WALTERS.  1962.  Atlas  of  the 

British  flora.  Botanical  Soc.  British  Isles.  Nelson  Press, 

New  York,  NY.  32pp. 
PHILLIPS,  A.R.,  J.  MARSHALL,  and  G.  MONSON.  1964.  The 

birds  of  Arizona.  Univ.  Arizona  Press,  Tucson.  220pp. 
PICKWELL,  G.A.  1931.  The  prairie  horned  lark.  Trans. 

Acad.  Sci.,  St.  Louis,  MO.  27:1-153. 
PORTER,  D.K.  1973.  Accuracy  in  censusing  breeding 

passerines  on  the  short -grass  prairie.  M.S.  thesis,  Co- 

loado  State  Univ.,  Fort  Collins.  107pp. 
,  MA.  STRONG, J.B.  GIEZENTANNER,  and  R.A. 

RYDER.  1975.  Nest  ecology,  productivity,  and  growth 

of  the  loggerhead  shrike  on  the  short-grass  prairie. 

Southwestern  Nat.  19:429-436. 
PREBLE,  N.A.  1957.  Nesting  habits  of  the  yellow-billed 

cuckoo.  Am.  Midi.  Nat.  57:474-482. 
PRESTON,  F.W.  1958.  Analysis  of  the  Audubon  Christmas 

counts  in  terms  of  the  lognormal  curve.  Ecol.  39:620- 

624. 
PRICE,  F.E.  and  C.E.  BOCK.  1983.  Population  ecology  of 

the  dipper  {Cinclus  mexicanus)  in  the  Front  Range 

of  Colorado.  Cooper  Ornith.  Soc,  Studies  Avian  Biol. 

7.  84pp. 
PUTNAM,  L.S.  1949.  Life  history  of  the  waxwing.  Wilson 

Bull.  61:141-181. 
RALPH,  C.J.  and  J.M.  SCOTT,  eds.  1981.  Estimating  num- 
ber of  terrestrial  birds.  Studies  Avian  Biol.  6.  630pp. 
RAMSEY,  F.L.  and  J.M.  SCOTT.  1979.  Estimating  popula- 
tion densities  from  variable  circular  plot  surveys. 

Pages  155-181  in  Cormack,  R.M.,  G.P.  Patel,  and  D.S. 

Rolson,  eds.  Sampling  Biological  Populations.  Int. 

Co-op  Publ.  House,  Fairfield,  MD. 
RAYNOR,  G.S.  1983.  A  method  for  evaluating  quality  of 

coverage  in  breeding  bird  atlas  projects.  Am.  Birds 

37:9-13. 
REYNOLDS,  R.T.,  J.M.  SCOTT,  and  R.A.  NUSSBAUM.  1980. 

A  variable  circular -plot  method  for  estimating  bird 

numbers.  Condor  82:309-313- 
RICE,J.C,  R.D.  OHMART,  and  B.W.  ANDERSON.  1983. 

Habitat  selection  attributes  of  an  avian  community:  A 

discriminant  analysis  investigation.  Ecol.  Monogr. 

53:263-290. 
ROBBINS,  C.S.  1966.  The  Christmas  count.  Pages  154-163 

in  Stefferud,  A.,  ed.  Birds  in  Our  Lives.  U.S.  Dep.  In- 
ter., Fish  and  Wildl.  Serv.  Washington,  DC.  56  lpp. 
.  1969.  The  breeding  bird  survey  of  1967  and  1968. 

U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  Spec.  Sci.  Rep. — 

Wildl.  124:1-107. 
.  1970.  Winter  bird  survey  technique  tested  in 

Maryland.  Maryland  Birdlife.  26:11-20. 
.  1972.  An  appraisal  of  the  winter  bird-population 

study  technique.  Am.  Birds  26:688-692. 
.  1974.  Progress  report  on  the  North  American 

breeding  bird  survey.  Acta  Ornithologica  14:170-191. 
.  1 978.  Census  techniques  for  forest  birds.  Pages 

142-163  in  Proceedings  of  Workshop:  Management  of 

southern  forests  for  nongame  birds.  U.S.  Dep.  Agric, 

For.  Serv.  Gen.  Tech.  Rep.  SE-14. 
,  D.  BYSTRAK,  and  PH.  GEISSLER.  1986.  The  breed- 
ing bird  survey:  its  first  15  years,  1965-1979.  U.S. 

Dep.  Inter.,  Fish  and  Wildl.  Serv.,  Resour.  Publ.  157:1- 

154. 


and  W.T.  VAN  VELZEN.  1967.  The  breeding  bird 

survey,  1966.  U.S.  Dep.  Inter.,  Fish  &  Wildl.  Serv. 

Spec.  Sci.  Rep— Wildl.  102:1-43- 
ROOT,  R.B.  1967.  The  niche  exploitation  pattern  of  the 

blue-gray  gnatcatcher.  Ecol.  Monogr.  37:317-350. 
ROTENBERRY,  J.T.  1980a.  Temporal  variation  in  habitat 

structure  and  shrub-steppe  bird  dynamics.  Oecologia 

47:1-9. 
.  1980b.  Habitat  structure,  patchiness,  and  avian 

communities  in  North  American  steppe  vegetation:  A 

multivariate  analysis.  Ecol.  61:1228-1250. 

and  J.A.  WIENS.  1976.  A  method  for  estimating 


species  dispersion  from  transect  data.  Am.  Midi.  Nat. 
95:64-78. 

ROTH,  R.R.  1981.  Vegetation  as  a  determinant  in  avian 
ecology.  Pages  162-174  in  Proceedings  First  Welder 
Wildl.  Foundation  Symp. 

SCOTT,  J.M.  and  F.L.  RAMSEY.  1981.  Effects  of  abundant 
species  on  the  ability  of  observers  to  make  accurate 
counts  of  birds.  Auk  98:610-612. 

, ,  and  C.B.  KEPLER.  1981.  Distance  estima- 
tion as  a  variable  in  estimating  bird  numbers.  Pages 
366-371  in  Ralph,  C.J.  and  Scott,  J.M. ,  eds.  Estimating 
the  Numbers  of  Terrestrial  Birds.  Studies  Avian  Biol. 
6. 

SEDGWICK,  J.A.  1981.  Breeding  bird  and  small  mammal 
habitat  relationships  in  northwestern  Colorado.  PhD 
dissertation,  Colorado  State  Univ.,  Fort  Collins.  1 39pp. 

SERVENTY,  D.L.  1980.  Developments  in  Australian  orni- 
thology. Bull.  British  Ornith.  Club.  100:89-93- 

SHARROCK,  J.T.R.  1975.  Dot  distribution  mapping  of 
breeding  birds  in  Europe.  Ardeola  21:797-810. 

.  1976.  The  atlas  of  breeding  birds  in  Britain  and 

Ireland.  British  Trust  for  Ornith.,  Tring.  Herts.  477pp. 

SKAAR,  P.D.  1969.  Birds  of  Bozeman  Latilong.  Published 
by  author,  501  S.  Third,  Bozeman,  MT.  131pp. 

SMITH,  DR.  1975.  Proceedings  of  the  symposium  on 

management  of  forest  and  range  habitats  for  nongame 
birds.  U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep. 
WO-1. 

STEWART,  RE.,  J.B.  COPE,  C.S.  ROBBINS,  and  J.W.  BRAI- 
NERD.  1952.  Seasonal  distribution  of  bird  populations 
at  the  Patuxent  Research  Refuge.  Am.  Midi.  Nat. 
47:257-363. 

SUMMERS-SMITH,  J.D.  1963.  The  house  sparrow.  Collins. 
London.  269pp. 

.  1 967.  Bibliography  of  the  genus  Passer.  Intl.  Stud- 
ies of  Sparrows.  Warsaw,  Poland.  82pp. 

SZARO,  R.C.  and  R.P.  BALDA.  1982.  Selection  and  moni- 
toring of  avian  indicator  species:  An  example  from 
a  ponderosa  pine  forest  in  the  Southwest.  U.S.  Dep. 
Agric,  For.  Serv.  Gen.  Tech.  Rep.  RM-89.  8pp. 

and  M.D.  JAKLE.  1982.  Comparison  of  variable 

circular-plot  and  spot-mag  methods  in  desert  riparian 
and  scrub  habitats.  Wilson  Bull.  94:546-550. 

TAYLOR,  DM.  and  CD.  LITTLEFIELD.  1984.  Willow  fly- 
catcher and  yellow  warbler  responses  to  reduced 
cattle  use  of  willow  habitat.  Abstract  of  paper  pre- 
sented at  54th  Ann.  Meeting,  Cooper  Ornith.  Soc. 
Humboldt  State  Univ.,  Areata,  CA.  June  19-23-  26pp. 

TOMIALOJC,  L.,  W.  WALANKIEWICZ,  and  T  WESO- 

LOWSKI.  1978.  Methods  and  preliminary  results  of 
the  collective  bird  census  work  in  the  primeval  forest 
of  the  Bialowieza  National  Park.  Polish  Ecol.  Studies 
3(4):215-224. 

TRAMER,  E.J.  1974.  An  analysis  of  the  species  density  of 
U.S.  landbirds  during  the  winter  using  the  1971 


Songbirds 


311 


Christmas  bird  count.  Am.  Birds  28:563-567. 
TRIMBLE,  S.  1975.  Non-game  birds  of  the  West:  An  anno- 
tated bibliography.  U.S.  Dep.  Inter.,  Bur.  Land  Manage. 

Tech.  Note  269.  320pp. 
UDVARDY,  M.D.F.  1957.  An  evaluation  of  quantitative 

studies  in  birds.  Cold  Spring  Harbor  Symp.  Quantita- 
tive Biol.  22:301-311. 
U.S.  DEPARTMENT  OF  AGRICULTURE,  FOREST  SER 

VICE,  Rocky  Mountain  Region.  1982.  Wildlife  and  fish 

habitat  relationships.  2  vols.  Denver,  CO. 
,  Southwestern  Region.  1975.  Endangered  and 

unique  fish  and  wildlife  of  the  Southwestern  National 

Forests.  Albuquerque,  NM.  203pp. 
VAN  VELZEN,  W.T.,  ed.  1981.  Forty -fourth  breeding  bird 

census.  Am.  Birds  35:46-112. 
VERBEEK,  N.A.M.  1970.  Breeding- ecology  of  the  water 

pipit.  Auk  87:425-451. 
VERNER,  J.  1985.  Assessment  of  counting  techniques. 

Chapter  8  in  Current  Ornithology  2:247-302. 
and  A.S.  BOSS.  1980.  California  wildlife  and  their 

habitats:  western  Sierra  Nevada.  U.S.  Dep.  Agric,  For. 

Serv.  Gen.  Tech.  Rep.  PSW-37.  439pp. 
and  J.V.  RITTER.  1985.  A  comparison  of  transects 

and  point  counts  in  oak-pine  woodlands  of  California. 

Condor  87:46-68. 
WEBSTER,  J.D.  1966.  An  analysis  of  winter  bird-population 

studies.  Wilson  Bull.  78:456-461. 


WIENS,  J.A.  1969.  An  approach  to  the  study  of  ecological 
relationships  among  grassland  birds.  Ornith.  Monogr. 
8:1-93. 

.  1973-  Pattern  and  process  in  grassland  bird  com- 
munities. Ecol.  Monogr.  43:237-270. 

1983-  Avian  community  ecology:  An  iconoclastic 


view.  Pages  355-403  in  Brush,  AH.  and  G.A.  Clark,  Jr., 

eds.  Perspectives  in  Ornithology.  Cambridge  Univ. 

Press.  New  York,  NY.  560pp. 
WILLIAMS,  A.B.  1936.  The  composition  and  dynamics  of  a 

beech-maple  climax  community.  Ecol.  Monogr.  6:317- 

408. 
WILLSON,  M.F.  1974.  Avian  community  organization  and 

habitat  structure.  Ecol.  55:1017-1029. 
WING,  L.  and  M.  JENKS.  1939.  Christmas  censuses:  The 

amateur's  contribution  to  science.  Bird  Lore  41:343- 

350. 
WINTERNITZ,  B.L.  1976.  Temporal  change  and  habitat 

preference  of  the  montane  breeding  birds.  Condor 

78:382-393. 
YEATMAN,  L.  1976.  Atlas  des  oiseaux  nicheur  de  France 

de  1970  a  1975.  Soc.  Ornithology  de  France,  55  Rue 

de  Buffon,  Paris.  283pp. 
YOUNG,  H.  1955.  Territorial  behavior  and  nesting  of  the 

eastern  robin.  Am.  Midi.  Nat.  53:329-352. 
ZELENY,  L  1976.  The  bluebird:  How  you  can  help  its  fight 

for  survival.  Indiana  Univ.  Press,  Bloomington.  170pp. 


312 


Songbirds 


16 
RAPTORS 


Michael  N.  Kochert 

U.S.  Bureau  of  Land  Management 

Snake  River  Birds  of  Prey  Research  Project 

Boise,  ID  83705 


Editor's  Note:  The  importance  of  raptors  is  widely 
recognized.  Interest  in  these  species  has  resulted  in 
many  studies  and  much  literature.  At  times,  biolo- 
gists can  have  trouble  wading  through  this  mate- 
rial to  find  what  they  need.  This  chapter 
summarizes  the  available  information  on  raptors 
and  refers  the  biologist  to  major  studies  and 
reports. 


INTRODUCTION 

Raptors  are  difficult  to  accurately  count  because 
they  often  nest  in  inaccessible  areas  at  relatively 
low  densities.  They  can  be  secretive  and  difficult  to 
find,  and  are  wide-ranging  and  rapid-moving.  Because 
of  these  difficulties,  a  census  (complete  count)  is 
rarely  possible.  In  most  raptor  surveys,  population 
and  productivity  estimates  are  based  on  samples. 
Because  data  are  sometimes  hard  to  obtain,  it  may 
be  impractical  to  obtain  adequate  sample  sizes  and 
results  may  differ  depending  on  how  the  data  were 
collected.  These  problems  can  be  compounded  by 
poor  sampling  design  and  methods  of  interpretation 
(Steenhof  and  Kochert  1982),  and  may  result  in  a 
waste  of  time  and  money. 

Surveys  should  provide  adequate  information  to 
determine  population  changes.  The  type  and  inten- 
sity of  an  inventory  or  monitoring  effort  may  be 
dictated  by  personnel,  funding,  or  logistics.  Because 
of  these  restrictions,  biologists  may  be  required  to 
use  low  resolution  surveys.  However,  statistically 
sound  sampling  procedures  must  be  used  regardless 
of  the  resolution. 

When  designing  an  inventory  or  monitoring 
effort,  biologists  should  obtain  general  information 
about  the  raptor  species  or  group  they  intend  to 
survey  and  the  habitats  in  which  it  occurs  (Fuller 
and  Mosher  1981 ).  They  also  must  consider  the 
biases  that  influence  survey  results.  Observer  compe- 
tence, experience  in  identifying  raptors  and  conduct- 
ing surveys,  and  knowledge  of  raptor  behavior  are 
important  factors  influencing  survey  results  and  can 
be  sources  of  variability.  This  variability  is  com- 
pounded when  several  species  in  many  different  hab- 
itats are  surveyed  simultaneously. 

This  chapter  provides  information  necessary  to 
design  a  raptor  inventory  and  monitoring  effort.  The 
first  section  presents  the  principal  habitat  features 
related  to  the  major  species  and  the  use  of  these 
features.  The  second  section  briefly  describes  the 
major  techniques  for  measuring  raptor  populations 
and  evaluates  their  limitations,  biases,  and  utility 
in  an  inventory  or  monitoring  program. 


' 


Raptors 


313 


HABITAT  FEATURES  CORRELATED  WITH 
SPECIES  GROUPS 

Raptors,  like  other  animals,  are  associated  with 
specific  habitat  features.  Often  the  occurrence  of  a 
raptor  species  can  be  predicted  by  the  presence 
of  certain  habitat  characteristics.  Knowledge  of  the 
relationships  between  raptors  and  habitat  features 
can  assist  biologists  in  designing  and  conducting  an 
inventory  or  monitoring  effort.  This  section  summa- 
rizes the  main  habitat  features  for  the  major  raptor 
species  in  the  western  United  States  (Table  1 )  and 
discusses  how  biologists  should  apply  these  features 
in  an  inventory  or  monitoring  effort.  Although  west- 
ern habitats  are  emphasized,  eastern  habitats  are 
described  if  the  species  also  occurs  in  the  eastern 
United  States. 

Vegetation  and  physical  (topographic  and  phys- 
iographic) features  are  the  two  most  important 
habitat  features  that  influence  the  abundance  and 
distribution  of  raptors  (Newton  1979).  These  fea- 
tures influence  availability  of  nest  sites  and  foraging 
habitat.  Some  raptors  are  adapted  to  forests  (e.g., 
accipiters)  whereas  others  are  adapted  to  open  grass- 
lands and  shrublands  (e.g.,  prairie  falcons  and  ferru- 
ginous hawks;  Table  1 ).  Some  species  occur  in  a 
wide  variety  of  topographic  and  physiographic  situa- 
tions and  others  require  a  relatively  narrow  set  of 
physical  features.  Great-horned  owls,  for  example, 
nest  in  numerous  types  of  topographic  and  physio- 
graphic situations,  but  nest  dispersion  of  cliff-nesting 


species,  such  as  prairie  falcons,  is  directly  influenced 
by  the  distribution  of  rock  outcroppings  and  escarp- 
ments. Other  raptor  species  are  closely  associated 
with  water  or  riparian  areas  (Table  1 ). 

Nesting  habitats  are  emphasized  in  Table  1  be- 
cause more  is  known  about  nesting  requirements 
than  requirements  during  other  portions  of  raptors' 
life  cycles,  and  successful  nesting  is  critical  in  main- 
taining population  stability  (Newton  1979).  Nesting 
habitat  features  in  Table  1  are  presented  in  the  form 
of  a  hierarchy.  The  vegetation  category  includes 
the  major  vegetation  types  (Kuchler  1964;  Brown  et 
al.  1980)  in  which  a  species  occurs  throughout  its 
range.  Physical  features  represent  specific  physio- 
graphic and  topographic  situations  within  the  vegeta- 
tion types  which  are  associated  with  a  particular 
species.  Habitat  characteristics  reflect  special  features 
of  the  vegetation  where  the  species  nests,  and  can 
be  related  to  age  class  (e.g.,  old-growth)  or  vegeta- 
tive structure  (e.g.,  closed,  dense  canopy,  or  open 
low  shrubs).  For  example,  northern  goshawks  and 
Cooper's  hawks  both  nest  in  conifer  forests  in  the 
west;  however,  goshawks  occur  in  dense  mature 
stands  with  a  multi-layered  canopy  and  Cooper's 
hawks  occur  in  even-aged  second  growth  stands 
with  a  dense  canopy  (Table  1).  Associations  with 
farmland  and  urban  development  are  included  in  this 
category  of  habitat  characteristics.  Farmland  habitat 
includes  cultivated  lands  or  pasture  lands,  whereas 
urban  habitat  consists  of  small  towns,  large  cities, 
and  suburban  areas. 


Cliffs  furnish  nest  sites  and  overlook  hunting  areas. 


314 


Raptors 


Within  these  areas  of  specific  habitat  characteris- 
tics, different  raptors  use  different  substrates  for 
nesting.  Some  species  have  a  very  narrow  range  of 
requirements  whereas  others  nest  on  many  different 
substrates.  Knowledge  of  these  requirements  aids 
in  locating  nests  and  predicting  the  occurrence  of 
nesting  birds.  Also,  certain  populations  can  be  re- 
stricted by  the  availability  of  nesting  sites  (Newton 
1979),  especially  for  tree  nesting  raptors  in  the 
Great  Plains  and  those  that  use  snags. 

Availability  of  food  is  the.  second  factor  that 
regulates  raptor  abundance  and  distribution  (Newton 
1979).  Knowledge  of  foraging  habitat  requirements 
can  help  biologists  predict  raptor  occurrence.  This 
knowledge  is  important  in  managing  and  maintaining 
foraging  habitat  that  can  be  critical  to  raptor  survival 
and  nesting  success.  The  ranges  in  foraging  habitat 
presented  in  Table  1  are  based  on  a  few  studies  in 
the  literature  and  should  be  considered  guidelines 
instead  of  absolute  rules. 

Wintering  habitat  may  be  as  important  as  nest- 
ing habitat  because  habitat  quality  may  influence 
juvenile  survival  and  the  condition  of  breeding  fe- 
males the  following  spring  (Newton  1979).  Winter- 
ing habitat  is  less  predictable  than  nesting  habitat 
because  raptors  usually  use  more  varied  habitats  in 
the  winter.  Biologists  should  be  aware  that  species 
such  as  bald  eagles,  northern  harriers,  and  long-eared 
owls  communally  roost  during  the  winter,  and  pro- 
tection and  management  of  these  roosts  are  impor- 
tant in  maintaining  wintering  habitat  for  these 
species. 

Little  is  known  about  migration  routes  and  habi- 
tat use  during  migration.  Certain  raptor  species, 
however,  use  migration  corridors  dictated  by  geo- 
graphic features  or  barriers  (e.g.,  mountain  ranges  or 
water  bodies)  which  cause  migrating  birds  to  con- 
gregate (Newton  1979).  In  the  western  United 
States,  migrating  raptors  tend  to  follow  north-south 
mountain  ranges  where  birds  use  deflective  air  cur- 
rents caused  by  escarpments.  Surveys  of  migrating 
raptors  are  of  little  value  for  local  inventory  or  moni- 
toring efforts  because  the  origin  of  the  birds  is 
mostly  unknown  and  migration  counts  are  subject  to 
many  biases  (Fuller  and  Mosher  1981).  These 
counts,  however,  can  have  value  on  a  regional  basis 
in  determining  general  population  trends. 

Biologists  should  use  the  information  presented 
in  Table  1  as  guidelines  for  predicting  possible  pres- 
ence of  species  and  for  developing  inventory  and 
monitoring  designs,  especially  those  using  stratified 
sampling  procedures.  However,  biologists  should  not 
use  these  features  as  a  sole  means  of  assessing  rela- 
tive raptor  abundance  or  habitat  quality  because 
many  environmental  factors  can  influence  actual 
population  numbers. 


POPULATION  MEASUREMENT  TECHNIQUES 

This  section  describes  the  major  techniques 
for  surveying  raptor  populations  that  can  be  applied 
in  a  BLM  inventory  and  monitoring  program.  How- 
ever, before  designing  a  survey,  biologists  should 
consult  Steenhof  ( 1986)  and  Fuller  and  Mosher 
(1986)  for  further  discussion  of  biases,  limitations, 
and  application  of  these  techniques.  Discussions 
in  this  section  focus  on  the  utility  of  each  technique 
in  assessing  relative  population  changes,  including 
evaluations  of  accuracy  and  precision.  Accuracy  is 
the  closeness  of  the  value  computed  or  measured  by 
the  technique  to  the  true  or  known  value  and  is 
expressed  as  a  percentage  of  the  true  or  known 
value.  Precision  is  the  closeness  of  repeated  meas- 
urements of  the  same  variable  and  is  the  amount  of 
variation  expressed  as  a  percentage  of  the  mean. 
Therefore,  measurements  with  greater  precision  have 
less  variation  in  relation  to  the  mean  and  thus 
smaller  precision  values.  I  used  coefficients  of  varia- 
tion (Sokal  and  Rohlf  1981 )  to  measure  precision. 
When  that  was  not  possible,  and  if  95%  confidence 
intervals  were  reported,  I  followed  Postovit  (1979) 
and  used  the  following  formula  which  gave  a  meas- 
ure roughly  double  the  coefficient  of  variation. 


Percent  precision : 


±  95%  Confidence  Interval 
Mean 


X100 


Presence 

A  survey  to  determine  presence  is  often  the  first 
step  in  evaluating  potential  impacts  of  land  use  and 
is  also  used  to  describe  geographical  range  of  a  spe- 
cies and  changes  in  distribution  (Fuller  and  Mosher 
1981).  Specific  techniques  for  assessing  presence  are 
not  discussed  in  this  section  because  they  are  essen- 
tially the  same  as  those  used  to  assess  relative  abun- 
dance, which  I  describe  in  a  later  section.  Use  of 
information  about  presence  of  a  species  does  not  re- 
quire careful  enumeration  of  birds,  and  data  collec- 
tion and  analysis  requirements  are  more  rigorous  for 
surveys  to  determine  relative  abundance. 

The  method  used  to  determine  presence  will 
depend  on  funding  and  logistics  as  well  as  the  raptor 
species  to  be  surveyed.  Although  data  collection 
requirements  are  the  least  rigorous,  methods  are 
subjected  to  biases  with  detectability  being  the  main 
bias  (Fuller  and  Mosher  1981).  The  literature,  habi- 
tat models,  and  knowledgeable  people  are  initial 
information  sources  in  assessing  presence.  Aerial  sur- 
veys are  an  expeditious  way  to  determine  presence 
of  conspicuous  raptors  over  large  areas.  Surveys  from 
boats  and  land  vehicles  can  be  more  efficient  than 
those  from  aircraft  because  of  less  cost  and  more 
opportunity  to  see  inconspicuous  raptors.  Vocaliza- 
tions provide  a  quick  means  to  detect  the  presence 


Raptors 


315 


Table  1.     Habitat  features  correlated  with  selected  Falconiform  species  in  the  United  States. 


SPECIES 

NESTING  HABITAT 

Vegetation 

Physical  Features 

Habitat  Characteristics 

Black  vulture 
(Coragyps  atratus) 

Shrubland;  grassland; 
eastern  and  southern  de- 
ciduous, coniferous,  and 
mixed  forest  types. 

Varied.  Flatlands  to  rolling 
hills  and  foothills,  can- 
yons, and  ravines. 

In  or  near  open  to  semi-open 
habitat  including  farmland. 

Turkey  vulture 
(Cathartes  aura) 

Numerous  shrubland, 
grassland,  coniferous, 
and  deciduous  forest 
types. 

Varied.  Flatlands  to 
mountainous  to  3750  m 
(12000  ft).  Canyons, 
ravines,  and  hillsides. 

In  or  near  open  to  semi-open 
habitat  including  farmland. 

California  condor 

(Gymnogyps 

californianus) 

Primarily  chaparral  vege- 
tation type.  One  nest 
located  in  a  redwood 
{Sequoia  sp.)  type. 

Rugged  mountainous 
terrain  between  610-1370 
m  (2000-4500  ft). 

Open  to  semi-open  habitats. 

Mississippi  kite 

(Ictinia 

mississippiensis) 

Eastern  deciduous  forest, 
grassland,  and  shrubland 
types. 

Level  to  rolling  uplands; 
also  floodplain. 

Tree  groves,  shelter  belts; 
gallery  and  riparian  forests 
near  open  habitat  including 
farmland  and  urban  areas. 

Black-shouldered  kite 
(Elanus  caeruleus) 

Grassland  and  oak 
(Quercus  sp.)  savannas. 

Coastal  valleys  and 
grassy  foothills. 

Open  to  semi-open  habitats 
including  farmland. 

Northern  goshawk 
{Accipiter  gentilis) 

Coniferous,  deciduous, 
and  mixed  forest  types. 

Gentle  slopes  or  flat 
areas  near  a  steep  incline 
or  canyon  often  near 
water.  Northerly  exposure 
in  western  U.S.  Southerly 
exposure  in  Alaska. 

Dense  mature  stands  with  a 
multilayered  canopy  and 
dense  to  open  understory. 

Cooper's  hawk 
(Accipiter  cooperii) 

Coniferous,  deciduous, 
and  mixed  forest  types. 
Also  shrubland  and 
grassland  types. 

Flat  terrain  or  gentle 
slopes  (30°),  sometimes 
near  water.  Low  eleva- 
tions. Sea  level  to  2700  m 
(8700  ft)  in  western  U.S. 
on  northerly  exposures. 

Wooded  areas  exceeding  6-8 
ha  (15-20  a.);  occasionally 
smaller  areas  or  isolated 
trees.  Mature  deciduous,  con- 
iferous, or  mixed  forest  with 
dense  understory  in  the  East. 
In  western  mountainous,  old 
or  even-aged  second  growth 
stands  (30-80  years)  with  a 
dense  canopy  and  sparse 
ground  cover.  Also  oakwood- 
land  habitat.  Riparian  forests 
in  grassland  and  shrubland 
types. 

Common  barn-owl 
(Tyto  alba) 

Nearly  every  shrubland, 
grassland,  and  forest 
type  except  in  montane 
areas  and  north  of  50°. 

Variable.  Absent  in  moun- 
tainous terrain. 

Open  to  semi-open  habitat. 
Common  in  farmland  and 
urban  habitats. 

Western  screech-owl 
(Otus  kennicottii) 

Most  forest  associations, 
shrubland,  and  grassland 
types. 

Variable  below  2470  m 
(8000  ft).  Also  riparian 
areas. 

Woodland  habitat  near  open 
areas.  Also  farmland  and 
urban  habitats. 

Whiskered  screech- 
owl 
{Otus  trichopsis) 

Oak  and  oak-pine  forests 
in  the  Southwest. 

Usually  canyons. 

Usually  riparian  habitat  and 
dense  stands. 

316 


Raptors 


FORAGING  HABITAT 

WINTERING  HABITAT 

SOURCES 

Nesting  Substrate 

Ground,  hollow  logs,  and 
stumps;  occasionally 
caves  and  old  buildings. 

Open  to  semi-open  areas 
where  carrion  is  available. 

Similar  to  nesting; 
also  urban  habitat. 

Scott  et  al.  (1977);  Am. 
Ornithol.  Union  (1983). 

Cliff  cavities  and  caves, 
ground,  hollow  logs,  and 
trees  near  ground  level. 

See  black  vulture. 

Similar  to  nesting. 

Scott  et  al.  (1977);  Call 
(1978). 

Cavities,  caves  and  crev- 
ices of  cliffs,  and  rarely 
hollow  trees. 

Grassland  and  oak  savanna 
foothills  up  to  60-70  km  (35- 
45  mi.)  from  nests  where 
large  mammal  carrion  is 
available. 

Same  as  nesting. 

Wilbur  (1978);  Studer 
(1983). 

Trees. 

Open  areas  near  colonies  or 
above  the  forest  canopy  for 
insects. 

N/A* 

Parker  (1975,  pers.  com- 
mun.);  Parker  and  Ogden 
(1979). 

Trees. 

Open  pasturelands,  grass- 
lands, and  forest  canopy  for 
insects. 

Same  as  nesting. 

Waian  and  Stendall 
(1970);  Bammann  (1975). 

Tall  trees,  greater  than  8 
m  (25  ft)  high. 

Same  as  nesting.  Also  eco- 
tone  between  forest  and 
openings.  Areas  containing 
rodents  and  large  birds  up  to 
2.8  km  (1.8  mi.)  from  nests. 

Varied  forested  and 
open  habitats.  More 
open  than  nesting 
habitat. 

McGowan  (1975);  Rey- 
nolds et  al.  (1982);  Rey- 
nolds (1983). 

Medium  high  trees, 
greater  than  6  m  (20  ft) 
high. 

Same  as  nesting.  Use  areas 
containing  medium  to  large 
passerines  within  2.4  km  (1.5 
mi.)  of  nests. 

Diverse  habitats. 
Woodland,  grass- 
land, shrubland, 
farmland,  and  urban. 
More  open  than 
nesting. 

Jones  (1979);  Millsap 
(1981);  Titus  and  Mosher 
(1981);  Reynolds  et  al. 
(1982);  Reynolds  (1983); 
Fuller  and  Partelow 
(1983). 

Cavities  in  trees,  cliffs, 
and  banks.  Wide  variety 
of  man-made  structures. 

Open  habitats  containing 
small  rodents  within  5.6  km 
(3.5  mi.)  of  nests. 

Similar  to  nesting. 

Scott  et  al.  (1977);  Call 
(1978);  C.  Marti  (pers. 
commun.);  Hegdal  and 
Blaskiewicz  (1984). 

Cavities,  woodpecker  or 
flicker  holes  in  trees, 
saguaro  cactus,  and  oc- 
casionally snags.  Man- 
made  structure. 

Open  habitats  containing 
rodents  and  insects  usually 
within  0.5  km  (0.3  mi.)  of  nest. 

Similar  to  nesting. 

Scott  et  al.  (1977);  Call 
(1978);  Johnson  et  al. 
(1979);  Marti  (1979). 

Cavities  and  flicker  holes 
in  trees. 

Areas  near  nest  containing 
insects. 

N/A* 

Marshall  (1967);  Scott  et 
al.  (1977). 

Raptors 


317 


Table  1.     Habitat  features  correlated  with  selected  Falconiform  species  in  the  United  States  (continued). 


SPECIES 

NESTING  HABITAT 

Vegetation 

Physical  Features 

Habitat  Characteristics 

Flammulated  owl 
(Otus  flammeolus) 

Coniferous  forest  and 
mixed  coniferous-decid- 
uous forest  types. 

Mountainous  areas  be- 
tween 1000-3100  m 
(3250-10000  ft).  Usually 
on  dry  hillsides  or  on 
ridge  tops. 

Usually  mature  forest  with 
dense  canopies  and  a  brush 
understory  or  intermixture 
of  oaks.  Occasionally  in  sec- 
ond growth  forest.  Rarely  in 
logged-over  areas. 

Northern  hawk-owl 
(Surnia  ulula) 

Northern  boreal  forests. 

Open  woods,  parklands, 
and  barrens  with  low 
shrubs  and  scattered 
trees. 

Forest  edge  areas  or  open 
canopy  (20-60%  canopy 
cover)  stands. 

Northern  pygmy-owl 
(Glaucidium  gnoma) 

Most  coniferous  associa- 
tions throughout  the 
West. 

Mountainous  up  to  3750 
m  (12000  ft). 

Open  to  semi-open  forests. 

Elf  owl 

(Micrathene  whitneyi) 

Palo  verde-saguaro  asso- 
ciations and  occasionally 
coniferous-deciduous 
forests  at  higher 
elevations. 

Steep  canyons,  riparian 
areas,  hillsides,  and  flats. 

Variable. 

Sharp-shinned  hawk 
(Accipiter  striatus) 

Coniferous,  deciduous,  or 
mixed  forest  types. 

Canyons,  valleys,  and 
riparian  areas.  In  can- 
yons, nest  50-100  m 
(162-325  ft)  upslopefrom 
a  stream. 

Dense  stands  with  well-devel- 
oped canopy  (60%)  and 
dense  understory.  Dense, 
young  (25-50  year),  even- 
aged,  second  growth  stands 
with  single-layered  canopy 
in  western  coniferous  forests. 
Also  in  conifers  within  dense 
mixed  stands. 

Osprey 

(Pandion  haliaetus) 

Numerous  shrubland, 
grassland,  coniferous  for- 
est, and  deciduous  forest 
types. 

Proximity  to  water  (can 
be  up  to  1 1  km  or  7  mi. 
away). 

Varied  habitats  including 
farmland  and  urban.  Old 
growth  stands  better  than 
even-age  second  growth 
stands. 

Bald  eagle 

(Haliaetus 

leucocephalus) 

Numerous  shrubland, 
deciduous  forest,  and 
mixed  forest  types. 

Proximity  to  water  (usually 
within  3.2  km  or  2  mi.). 

Uneven-aged,  multi-layered 
forests  with  an  availability 
of  large  trees  with  sturdy 
branches. 

Golden  eagle 
(Aquila  chrysaetos) 

Shrubland  and  grassland 
types.  Also  tundra,  coni- 
ferous forests,  and  wood- 
land-brushland. 

Canyons,  buttes,  and 
mountainous  escarp- 
ments between  600-3100 
m  (1950-10000  ft)  in 
elevation. 

Open  to  semi-open  habitats. 
Sometimes  riparian  and  farm- 
land habitats. 

318 


Raptors 


FORAGING  HABITAT 

WINTERING  HABITAT 

SOURCES 

Nesting  Substrate 

Woodpecker  or  flicker 

1 7-27  ha  (43-63  a.)  areas 

N/A* 

Bull  and  Anderson 

holes  in  trees  or  snags. 

around  nests  containing 

(1978);  Franzreb  and 

Also  nest  boxes. 

insects. 

Omart  (1978);  Winter 
(1979);  Reynolds  and 
Linkhart  (1984). 

Cavities  in  trees,  some- 

Open areas  containing  micro- 

Same  as  nesting. 

Grossman  and  Hamlet 

times  snags.  Occasion- 

tines, small  rodents,  and 

Sometimes  open 

(1964);  Scott  etal. 

ally  corvid  nests  in  trees 

birds. 

habitat  in  temperate 

(1977);  Meehan  and 

or  on  cliffs. 

areas. 

Ritchie  (1982). 

Woodpecker  holes  in 

Areas  near  nests  containing 

Open  to  semi-open 

Scott  etal.  (1977);  Call 

trees  or  snags. 

insects  and  small  rodents. 

habitats. 

(1978);  Hayward  (1983). 

Woodpecker  holes  in 

Small  area  around  nest  con- 

N/A* 

Scott  etal.  (1977);  Call 

saguaro  cactus  or  snags. 

taining  insects,  particularly 

(1978);  Johnson  etal. 

Also  cavities  in  dead 

beetles,  moths,  and  crickets. 

(1979);  Millsap  (pers. 

branches  of  living  trees. 

commun.). 

Small  trees  (less  than  6  m 

Same  as  nesting  areas  con- 

Diverse habitats 

Piatt  (1973);  Jones 

[20  ft]  high). 

taining  small  passerine  birds 

including  riparian, 

(1979);  Millsap  (1981); 

within  approximately  1.5  km 

woodland,  farmland, 

Clarke  (1982);  Reynolds 

(0.9  mi.)  of  nests. 

and  urban  habitats. 

et  al.  (1982);  Reynolds 

More  open  than 

(1983). 

nesting.  Perches 

have  substantial  ar- 

boreal cover. 

Tops  of  trees,  rock  and 

Slow-moving  water  usually 

N/A* 

Johnson  and  Melquist 

dirt  pinnacles,  cactus, 

within  a  few  km  of  the  nest. 

(1973);  Zarn  (1974c); 

and  numerous  man-made 

Can  be  up  to  1 1  km  (7  mi.) 

Melquist  (1975). 

structures  (snags  or  trees 

away. 

with  dead  tops 

preferred). 

Primarily  trees,  occasion- 

Primarily slow-moving  water 

Same  as  nesting,  but 

Call  (1978);  Fraser 

ally  cliffs.  Nest  below 

within  a  few  km  of  the  nest. 

also  in  farmland 

(1978);  Steenhof  (1978); 

crown  in  super  canopy 

During  winter,  bodies  of 

and  urban  habitats. 

Steenhof  et  al.  (1980); 

trees.  Occasionally  man- 

water;  occasionally  upland 

Open  water  most 

Anthony  etal.  (1982); 

made  structures. 

habitats.  Winter  foraging 

important;  however, 

Keister  and  Anthony 

areas  can  be  up  to  27  km  (17 

shrubland  also  used. 

(1983);  U.S.  Fish  and 

mi.)  from  roost. 

Diurnal  perches 
primarily  in  trees 
close  to  a  food 
source.  Night  roosts 
more  protected  than 
day  roosts.  Stands 
generally  have  large, 
old,  open-structured 
trees  with  horizontal 
branches. 

Wildlife  Service  (1983). 

Primarily  cliffs,  sometimes 

Open  habitats  where  rabbits 

Same  as  nesting.  In 

Snow  (1973);  Wrakestraw 

trees,  and  occasionally 

and  small  rodents  are  avail- 

the West  below  1400 

(1973);  U.S.  Bureau  of 

ground  and  man-made 

able,  usually  within  7  km  (4.4 

m  (4500  ft)  in  eleva- 

Land Management 

structures. 

mi.)  of  nests. 

tion.  In  the  East 

(1979);  Kochert  (1980); 

coastal  plains  and 

Millsap  and  Vana  (1984). 

wetlands. 

Raptors 


319 


Table  1.     Habitat  features  correlated  with  selected  Falconiform  species  in  the  United  States  (continued). 


SPECIES 

NESTING  HABITAT 

Vegetation 

Physical  Features 

Habitat  Characteristics 

Northern  harrier  (Cir- 
cus cyaneus) 

Grassland  and  shrubland 
associations;  also  large 
open  expanses  in  decid- 
uous forests. 

Relatively  flat  and  open 
terrain  from  sea  level 
to  3200  m  (10,400  ft)  in 
elevation. 

Open  habitat  with  short  vege- 
tation. Mainly  wetland  and 
riparian  habitats,  but  also 
shrub  and  grass  upland  and 
farmland  habitats. 

Red-tailed  hawk 
(Buteo  jamaicensis) 

Nearly  every  open  to 
semi-open  vegetation 
type.  Absent  in  tundra 
and  uncommon  in  dense 
forests. 

Canyons,  hillsides,  and 
level  terrain.  In  East  oc- 
cupy forested  uplands, 
ridges,  and  steep  slopes. 
Riparian  floodplain  in 
Alaska. 

In  or  near  open  to  semi-open 
habitats.  Nests  situated  for 
unobstructed  access  in  ripar- 
ian and  dense  forested  areas. 

Swainson's  hawk 
{Buteo  swainsoni) 

Shrubland  and  grassland 
types. 

Low  to  moderate  eleva- 
tions, valleys,  foothills, 
and  level  uplands.  In 
areas  moist  enough  to 
support  trees. 

Small  tree  groves  in  open 
habitats,  including  farmland. 

Ferruginous  hawk 
{Bueto  regal  is) 

Grassland,  shrubland, 
and  juniper-pinyon  (Juni- 
perus-Pinus)  woodland 
types. 

More  xeric  sites  than 
other  sympatric  buteos. 
Avoid  mountainous  areas, 
steep  canyons,  and  high 
cliffs. 

Open  habitat.  Avoid  heavily 
farmed  areas. 

Red-shouldered 

hawk 

(Buteo  lineatus) 

Deciduous  forest  associ- 
ations in  East.  Riparian 
and  sycamore  (Platanus) 
woodland  in  West. 

Mosaic  lowlands,  riparian 
areas,  floodplains,  and 
valleys. 

Mature  forests  with  fairly 
closed  canopy  and  large 
amount  of  floodplain  edge  in 
the  East.  In  the  West,  riparian 
habitat  with  taller-than-aver- 
age  trees. 

Broad-winged  hawk 
(Buteo  platypterus) 

Deciduous  and  mixed 
deciduous  forest  types. 

Wet,  poorly  drained 
areas.  More  upland  than 
red-shouldered  hawk 
areas. 

Much  like  red-shouldered 
hawk.  Nest  close  to  small 
streams,  lakes  or  swampy 
areas. 

Rough-legged  hawk 
(Buteo  lagopus) 

Tundra  and  taiga. 

Foothills  and  mountain 
valleys  with  bluffs  and 
rock  formations. 

Open  habitat. 

Zone-tailed  hawk 
(Buteo  albonotatus) 

Various  shrubland  types 
in  the  Southwest;  also  the 
Arizona  Pine  Forest  type. 

Rugged  terrain.  Deep, 
broad  canyons  and  table- 
lands up  to  2750  m  (9000 
ft)  in  elevation. 

Large  trees  required.  Usually 
in  perennial  and  intermittent 
streams;  occasionally  in  table- 
lands away  from  water. 

Gray  hawk 
(Buteo  nitidus) 

Riparian  forests  in  mes- 
quite  and  shrubland 
types. 

Floodplain  areas. 

Presently  mature  gallery  trees. 
Historically  mesquite  trees. 

320 


Raptors 


FORAGING  HABITAT 

WINTERING  HABITAT 

SOURCES 

Nesting  Substrate 

Ground.  In  rank 
vegetation. 

Open  habitats  up  to  18  km 
(11.2  mi.)  from  the  nest. 

Same  as  nesting,  but 
at  lower  elevations. 
Use  uplands  more 
often.  Will  commun- 
ally roost  in 
depressions. 

Call  (1978);  Apfelbaum 
and  Seelbach  (1983); 
Thompson-Hanson 
(1984). 

Trees  preferred,  also 
cliffs  and  man-made 
structures.  Higher  and  in 
taller  trees  than  other 
sympatric  Buteo  species. 

Open  habitats  containing 
lagomorphs,  small  rodents, 
and  snakes,  usually  within  3 
km  (1.9.  mi.)  of  nest. 

Same  as  nesting. 
Absent  from  high 
elevations. 

Brown  and  Amadon 
(1968);  U.S.  Bureau  of 
Land  Management 
(1979);  Schmutz  et  al. 
(1980);  Cottrell  (1981); 
Millsap  (1981);  Titus  and 
Mosher  (1981);  Bednarz 
and  Dinsmore  (1982). 

Primarily  trees  and  small 
shrubs,  occasionally 
man-made  structures. 
Closer  to  the  ground  than 
other  sympatric  Buteo 
species. 

Open  habitats  with  short  veg- 
etation containing  small  ro- 
dents and  insects  within  3  km 
(1 .9  mi.)  of  the  nest  and  a 
home  range  of  approximately 
9  km2  (3.5  mi.2). 

N/A* 

Fitzner  (1980);  Schmutz 
et  al.  (1980);  Cottrell 
(1981);  Bechard  (1982); 
Gilmer  and  Stewart 
(1984). 

Low  rocky  outcrops  (pre- 
ferred), trees  in  small 
groves,  on  cut  banks, 
and  a  variety  of  man- 
made  structures. 

Open  habitats  with  short  veg- 
etation containing  ground 
squirrels,  pocket  gophers, 
and  rabbits  usually  within  2-3 
km  (1.2-1.9  mi.)  of  nest. 

Same  as  nesting 
habitat,  also 
pastureland. 

Wakely  (1978);  Schmutz 
et  al.  (1980);  Cottrell 
(1981);  Millsap  (1981); 
Evans  (1983). 

Trees,  larger  than  those 
used  by  broad-winged 
hawks.  Farther  below 
canopy  than  red-tailed 
hawks. 

Closed  to  semi-closed  wood- 
land in  and  near  wet  areas 
containing  small  rodents, 
reptiles,  amphibians,  and 
crayfish  in  an  area  0.6-0.9 
km   (0.2-3.5  mi.2)  around 
nest.  Up  to  1.0-2.3  km  (0.6- 
1.4  mi.)  from  nest. 

Similar  to  nesting; 
also  forest  edge  and 
openings. 

Brown  and  Amadon 
(1968);  Wiley  (1975); 
Titus  and  Mosher  (1981); 
McCrary  (1981);  Bednarz 
and  Dinsmore  (1982). 

Trees  (see  red-shoul- 
dered hawk). 

Closed  to  semi-open  habitat 
in  an  approximately  13  km2 
(5.1  mi.2)  range  containing 
small  mammals,  reptiles, 
amphibians,  and  large 
insects. 

N/A* 

Brown  and  Amadon 
(1968);  Matray  (1974); 
Fuller  (1979);  Titus  and 
Mosher  (1981). 

Primarily  river  bluffs, 
lesser  degree  upland 
outcroppings  and  es- 
carpments. Occasionally 
the  ground,  sometimes 
trees. 

Wet  meadows,  bogs, 
marshes,  open  riparian  areas, 
pastures.  Also  shrub  and 
grass  uplands  containing 
small  rodents. 

Open  shrub  and 
grass  habitats  in 
temperate  areas  (see 
foraging  habitat). 

Zarn  (1975);  Bildstein 
(1978);  BLM  (unpubl. 
data). 

Primarily  trees,  rarely 
cliffs. 

Upland  habitats,  open  shrub- 
land  containing  an  abun- 
dance of  lizards  up  to  27  km 
(16.0  mi.)  from  nests. 

N/A*  Rare  winter 
sightings  in  south- 
west U.S. 

Call  (1978);  Millsap 
(1981,  pers.  commun.); 
Fuller  (1983). 

Trees. 

Thorn  shrub-mesquite  bosque 
with  an  abundance  of  lizards. 

N/A*  Rare  winter 
sightings  in  south- 
west U.S. 

Millsap  (pers.  commun.). 

Raptors 


321 


Table  1.     Habitat  features  correlated  with  selected  Falconiform  species  in  the  United  States  (continued). 


SPECIES 

NESTING  HABITAT 

Vegetation 

Physical  Features 

Habitat  Characteristics 

Common  black-hawk 

{Buteogallus 

anthracinus) 

Shrubland  types  of  the 
Southwest. 

Riparian  and  floodplain 
between  430-1850  m 
(1400-6000  ft)  in  eleva- 
tion. Perennial  streams 
essential. 

Mature  riparian  habitat.  Tree 
groves  preferred  over  single 
trees. 

Harris'  hawk 

(Parabuteo 

unicinctus) 

Mesquite  (Prosopis  sp.) 
and  Palo  verde  (Cerci- 
dium  sp.)  vegetation 
types. 

Flatlands  with  washes 
and  ravines  that  give  way 
to  low  hills. 

Structurally  complex  and 
diverse  vegetation.  Multilay- 
ered  and  semi-closed  shrub. 

Gyrfalcon 
(Falco  rusticolus) 

Tundra,  occasionally  in 
taiga. 

Foothills  and  mountain 
valleys  with  rock  forma- 
tions, bluffs,  and  rock 
outcroppings. 

Varied.  Vast  expanses  of 
open  habitat. 

Peregrine  falcon 
(Falco  peregrinus) 

Nearly  all  shrubland, 
grassland,  and  forest 
types.  Rare  in  alpine 
areas. 

Flat  terrain  to  rugged 
canyons.  Usually  on  high 
cliffs  (62  m  or  200  ft),  in 
areas  of  open  expanses 
and  high  topographic 
relief,  and  near  (within  4.8 
km  or  3.0  mi.)  permanent 
or  semi-permanent 
sources  of  water. 

Open  habitats,  including 
farmland  and  urban  habitats. 

Prairie  falcon 
(Falco  mexicanus) 

Arid  and  semi-arid  shrub- 
land and  grassland 
types.  Sometimes  the 
open  parklands  of  coni- 
ferous forest  types. 

Canyons,  buttes,  moun- 
tainous escarpments, 
outcroppings  to  eleva- 
tions of  3100  m  (10,000 
ft). 

Open  habitat  with  short 
vegetation. 

Merlin 

(Falco  columbarius) 

Shrubland,  grassland, 
boreal  forest,  and  coastal 
coniferous  forest  types. 

Varied.  Uncommon  in 

rugged-mountainous 

terrain. 

Near  open  habitat.  Forests 
broken  by  expanses  of  open 
country.  Openings  in  dense 
forests.  Tree  groves,  riparian 
forests. 

American  kestrel 
(Falco  sparverius) 

Nearly  all  shrubland, 
grassland,  and  forest 
types. 

Variable  terrain  associ- 
ated with  open  areas  up 
to  3750  m  (12,000  ft). 

Wide  variety  of  open  to  semi- 
open  habitats,  including  farm- 
land and  urban  habitats. 

Spotted  owl     Northern 
(Strix               spotted 
occidentalis)  owl 

Douglas-fir  (Pseudotsuga 
menziesii),  cedar  (Thuja 
sp.),  and  hemlock  (Tsuga 
sp.)  forests. 

Usually  below  1700  m 
(5500  ft)  elevation  on 
75%  slopes  half  way 
down  hillside. 

Multilayered  old  growth  (100- 
200+  years)  forests  with 
deep  ravines  and  50-90% 
canopy  closure. 

California 

spotted 

owl 

Fir-pine/Douglas-fir  for- 
ests and  oak  woodlands. 

Steep  slopes. 

Heavily  timbered  old  growth. 

Mexican 

spotted 

owl 

Spruce,  Douglas-fir,  and 
pinyon-juniper  forests. 
Aspen  and  maple  stands. 

Deep  narrow  timbered 
canyons  below  2770 
m  (9000  ft)  in  elevation 

Timbered  riparian  habitat. 

322 


Raptors 


FORAGING  HABITAT 

WINTERING  HABITAT 

SOURCES 

Nesting  Substrate 

Trees. 

Riparian  habitat  (mostly  in  the 
water)  containing  fish,  am- 
phibians, reptiles,  and  small 
mammals  within  0.5  km  (3.1 
mi.)  of  nests. 

N/A* 

Schnell  (1979);  Millsap 
(pers.  commun). 

Palo  verde,  mesquite, 
other  trees,  and  saguaro 
cactus  (Cereus 
giganteus). 

Semi-closed  areas  (more 
closed  than  those  used  by 
red-tailed  hawks)  containing 
rabbits,  medium  to  small 
rodents  and  birds  (quail)  in  a 
5  km2  (2.0  mi.2)  area  around 
the  nest. 

Same  as  nesting, 
also  occur  in 
farmland. 

Call  (1978);  Mader  (1976, 
1978);  Millsap  (1981). 

Primarily  cliffs  (6-90  m) 
occasionally  trees  in 
taiga. 

Open  habitats  with  an  abun- 
dance of  hare  and  ptarmigan. 

Same  as  nesting. 
Sometimes  in  the 
northern  contermi- 
nous U.S. 

Cade  (1982);  Mindell 
(1983). 

Cliffs  (60  m),  rarely  in 
trees. 

Open  habitats.  Usually  ripar- 
ian and  shoreline  areas.  Also 
upland  habitats  containing 
shorebirds  and  passerines. 
Usually  within  5  km  (3.1  mi.) 
of  nests,  but  up  to  24  km  (15 
mi.). 

N/A* 

Porter  and  White  (1973); 
Cade  (1982);  Ellis  (1982). 

Cliffs  and  rock  outcrop- 
pings,  rarely  abandoned 
raptor  nests  in  trees. 

Open  areas  of  low  vegetation 
containing  ground  squirrels 
and  passerine  birds  up  to  24 
km  (15  mi.)  from  nests. 

Similar  to  nesting, 
also  farmland 
habitats. 

U.S.  Bureau  of  Land 
Management  (1979); 
Cade  (1982);  Becker  and 
Ball  (1983a). 

Abandoned  corvid  nests 
in  trees,  occasionally  in 
tree  cavities  or  on  the 
ground. 

Open  shrub  and  grass  habi- 
tats containing  passerine 
birds  up  to  9  km2  (5.6  mi.2) 
from  nests.  Also  forage  for  in- 
sects above  forest  canopy. 

Coastal,  marsh, 
grassland,  farmland, 
and  urban  habitats 
where  passerine  and 
shore  birds 
concentrate. 

Trimble  (1975);  Scott  et 
al.  (1977);  Cade  (1982); 
Becker  and  Ball  (1983b); 
Becker  (1984);  Millsap 
(pers.  commun.). 

Old  flicker  holes  or  cavi- 
ties in  banks  and  cliffs, 
buildings,  abandoned 
magpie  nests,  and  nest 
boxes. 

Open  habitats  mostly  in 
marshland,  grassland,  sa- 
vanna, shrubland,  open  for- 
est, farmland,  and  urban 
habitats. 

Similar  to  nesting. 
Absent  in  high 
elevations. 

Scott  et  al.  (1977);  Cade 
(1982). 

Tree  cavities  or  other 
raptor  nests. 

Old  growth  stands,  primarily 
second  growth  stands  sec- 
ondarily containing  flying 
squirrels  (Glaucomys  sp.)  and 
woodrats  {Neotoma  sp.)  in  a 
0.5-3.4  km2  (0.2-1.3  mi.  )  area 
around  nest. 

Similar  to  nesting. 

Forsman  (1983a);  Fors- 
man  et  al.  (1984). 

Tree  cavities,  other  raptor 
nests,  cavities  in  cliffs; 
occasionally  on  the  ground. 

Same  as  nesting. 

Lower  elevations. 
Tend  to  be  more 
open. 

Zarn  (1974a);  Layman 
(pers.  commun). 

Tree  cavities,  cliff  ledges 
and  cavities,  empty  rap- 
tor nests. 

Same  as  nesting. 

Similar  to  nesting. 

Zarn  (1974a);  C.  Marti 
(pers.  commun.). 

Raptors 


323 


Table  1.     Habitat  features  correlated  with  selected  Falconiform  species  in  the  United  States  (concluded). 


SPECIES 

NESTING  HABITAT 

Vegetation 

Physical  Features 

Habitat  Characteristics 

Barred  owl 
(Strix  varia) 

Eastern  deciduous  and 
mixed  coniferous  forests; 
boreal  and  montane  coni- 
ferous forests. 

In  lowlands  in  East.  In 
West,  features  similar  to 
spotted  owl  habitats. 

In  the  West,  resembles  north- 
ern spotted  owl  habitat;  how- 
ever, less  affiliation  with  old- 
growth  forests.  In  East,  dense 
stands  in  swamps  and  river 
bottoms. 

Boreal  owl 
(Aegolius  funereus) 

Boreal  and  montane 
coniferous  forests. 

Variable  from  flat  boreal 
uplands  to  mountainous 
terrain. 

Dense,  closed  canopy  stands. 

Northern  saw-whet 

owl 

(Aegolius  acadicus) 

Most  forest  types  up  to 
3400  m  (10,000  ft)  in 
elevation. 

Variable. 

Dense,  closed  canopy  stands 
with  smaller  trees,  fewer 
openings;  greater  canopy 
than  those  used  by  boreal 
owls. 

Great  horned  owl 
(Bubo  virginianus) 

Nearly  every  grassland, 
shrubland,  and  forest 
type.  Absent  in  tundra. 

Variable. 

Wide  variety  of  open  to  semi- 
open  habitats  including  farm- 
land and  urban  habitats. 

Great  gray  owl 
(Strix  nebulosa) 

Northern  boreal  forests 
and  montane  coniferous 
forests. 

Flat  lowlands  to  moun- 
tainous terrain  between 
1850-2500  m  (6000-8100 
ft)  in  elevation. 

Old  growth  conifer  or  mixed 
deciduous  stands  in  montane 
forests.  Mature  poplar,  and 
to  lesser  degree,  conifer 
stands  in  boreal  forests.  Nests 
usually  near  a  wet  meadow 
or  marsh. 

Long-eared  owl 
(Asio  otus) 

Most  shrubland,  grass- 
land, and  forest  types. 

Variable,  level,  to  moun- 
tainous terrain. 

Semi-open  to  dense  wooded 
areas,  tree  groves,  and  dense 
shrubs  near  open  areas  in- 
cluding farmlands.  Nests 
usually  within  a  few  meters  of 
the  woodland  edge. 

Burrowing  owl 
(Athene  cunicularia) 

Arid  and  semi-arid  shrub- 
land and  grassland 
types. 

Level  to  rolling  hills. 

Open  habitats  including  farm- 
land and  urban  areas  with 
short  vegetation  and  an  avail- 
ability of  burrows. 

Short-eared  owl 
(Asio  flammeus) 

Most  shrubland  and 
grassland  types  north  of 
37°  north  latitude.  Also 
in  tundra. 

Open  terrain.  Absent  from 
mountainous  areas. 

Open  habitat  including  grass- 
land, shrubland,  meadow, 
marsh,  and  farmland  habitat. 

Snowy  owl 
(Nyctea  scandiaca) 

Tundra. 

Open  flats  or  slopes. 
Nests  usually  on  highest 
point;  occasionally  on 
a  rocky  ledge. 

Open  habitat.  Nests  some- 
times near  marshes  or  small 
lakes. 

324 


Raptors 


FORAGING  HABITAT 

WINTERING  HABITAT 

SOURCES 

Nesting  Substrate 

Large  hollow  trees;  occa- 
sionally corvid  and  raptor 
nests.  Sometimes  snags. 
Also  nest  boxes. 

Wooded  areas  with  an  open 
understory  in  a  home  range 
between  0.8-3.6  km2  (0.3-1.4 
mi.2). 

Similar  to  nesting. 

Nicholls  and  Warner 
(1972);  Scott  et  al. 
(1977);  Apfelbaum  and 
Seelbach  (1983);  Taylor 
and  Forsman  (1976); 
Boxall  and  Stepney 
(1982). 

Woodpecker  holes  in 
snags. 

Semi-open  areas  containing 

microtines. 

Similar  to  nesting. 

Scott  et  al.  (1977);  Mee- 
han  and  Ritchie  (1982); 
Hayward  (1983). 

Woodpecker  holes  in 
snags  and  trees.  Also 
nest  boxes. 

Areas  containing  small  ro- 
dents and  insects  in  a  range 
up  to  0.5  km2  (0.2  mi.2). 

Similar  to  nesting. 

Forbes  and  Warner 
(1974);  Scott  etal. 
(1977);  Hayward  (1983). 

Abandoned  large  bird 
nests  in  trees  and  on 
cliffs.  Also  cliff  ledges 
and  cavities. 

Open  to  semi-open  areas 
containing  medium  to  small 
mammals  in  a  1 .3-3.7  km2 
(0.5-1.4  mi.2)  area. 

Similar  to  nesting. 

Call  (1978);  Fuller  (1979); 
Am.  Ornithol.  Union 
(1983). 

Abandoned  raptor  or 
corvid  nests  in  trees. 

Marsh  and  wet  meadows 
containing  microtines. 

Similar  to  nesting. 
Occasionally  farm- 
land and  urban 
habitat. 

Winter  (1979);  Nero 
(1980). 

Abandoned  corvid  nests 
in  small  scrubby  trees 
or  large  bushes. 

Open  habitats  containing 
microtines  and  other  small 
rodents  in  1.8-3.7  km2  (0.7-1.4 
mi.2)  range. 

Similar  to  nesting. 

Call  (1978);  Marti  (1979); 
Wijnandts  (1984);  Marks 
(1984). 

Abandoned  mammal 
burrows.  Occasional  cav- 
ities in  small  basalt 
outcroppings. 

Open  habitat  with  short  vege- 
tation containing  small  ro- 
dents, insects,  lizards,  and 
passerine  birds. 

Similar  to  nesting. 

Zarn  (1974b);  Marks  and 
Ball  (1983). 

Ground 

Open  habitat  (see  character- 
istics) containing  microtines  in 
a  0.2-1.2  km2  (0.07-0.5  mi.2) 
range. 

Similar  to  nesting. 

Clark  (1975);  Call  (1978). 

Ground 

Open  uplands  and  meadows 
with  a  supply  of  small 
rodents. 

Same  as  nesting. 
Also  coastal,  grass- 
land, farmland,  and 
urban  habitats. 

Grossman  and  Hamlet 
(1964);  Lein  and  Weber 
(1979). 

"Migrates  outside  the  United  States 


Raptors 


325 


of  secretive  raptors  over  large  areas,  and  walking 
surveys  are  useful  for  locating  raptors  that  are  diffi- 
cult to  detect. 


Relative  Abundance — Non-Nesting  Surveys 
Road  Counts. 


Description.  Road  surveys  can  be  conducted 
by  either  the  continuous  count  or  point  count 
methods  (Fuller  and  Mosher  1986).  Routes  either 
wander  through  different  habitats  (Craighead  and 
Craighead  1956),  circle  or  cut  through  an  area 
(Craig  1978),  or  consist  of  transects  arranged  to 
completely  cover  an  area  (Marion  and  Ryder  1975). 

When  conducting  continuous  counts,  observers 
repeatedly  drive  specified  routes  in  a  vehicle  at 
speeds  between  16-40  km/h  (10-25  mph)  and  count 
all  birds  seen  on  either  side  of  the  road.  They  record 
the  perpendicular  distance  from  the  transect  for 
each  bird  seen.  Because  birds  are  counted  on  either 
side  of  the  road  two  observers  are  necessary. 

For  the  point  count  method,  observers  stop  at 
systematically  placed  stations  (usually  0.8  km  or  0.5 
mi.  apart)  along  the  road  (Fuller  and  Mosher  1986). 
During  a  prescribed  time  period  (usually  3  min), 
they  count  all  raptors  seen  or  heard  within  a  prede- 
termined radius  of  the  station.  The  radius  at  stations 
can  be  fixed  (usually  0.4  km  [0.25  mi.]),  variable, 
or  unlimited.  In  either  situation,  biologists  must  re- 
cord the  distance  from  the  station  of  each  bird  seen 
or  heard.  An  advantage  of  this  method  is  that  each 
station  (or  circular  plot)  can  be  treated  as  an  inde- 
pendent sample  with  its  own  habitat  characteristics, 
thus  increasing  sample  size  and  enhancing  statistical 
analysis. 

Before  conducting  road  surveys,  biologists 
should  list  raptor  species  and  habitat  types  they  need 
to  survey.  They  should  delineate  habitat  types  on 
maps  and  compute  the  linear  distance  of  each  type 
along  the  road.  Transects  should  be  run  at  least  once 
to  delineate  the  habitat  types,  to  assess  effective 
detection  distances,  and  to  identify  any  biases  that 
may  affect  the  count.  Routes  should  be  selected  so 
that  all  major  communities  are  represented  in  the 
sample  area  (Craighead  and  Craighead  1956;  Millsap 
1981). 

Road  counts  should  be  made  under  similar 
weather  conditions  because  weather  conditions  can 
affect  raptor  activity  (Craighead  and  Craighead 
1956).  The  time,  weather  (at  least  precipitation, 
wind,  cloud  cover,  and  temperature),  and  vehicle 
odometer  reading  (read  to  0.16  km  [0.1  mi]  and  es- 
timated to  nearest  0.08  km  [0.05  mi.])  should  be 
recorded  at  the  beginning  and  end  of  the  survey.  For 


each  sighting,  observers  should  record  the  species, 
age  class,  sex  (if  possible),  and  activity  (perched  or 
flying)  of  the  raptor  as  well  as  the  time,  habitat  type, 
vehicle  odometer  reading  (read  to  0.16  km  [0.1  mi.] 
and  estimated  to  nearest  0.08  km  [0.05  mi.]),  and 
perpendicular  distance  from  the  bird  to  the  road 
(Millsap  1981). 

All  road  count  methods  have  inherent  biases  for 
which  biologists  must  compensate  (Fuller  and 
Mosher  1986).  The  season,  time  of  day,  weather,  and 
activity  of  the  birds  can  bias  counts  (Fuller  and 
Mosher  1981).  To  compensate  for  this,  researchers 
have  alternated  the  direction  of  each  survey  run, 
restricted  runs  to  specific  times  of  the  day,  or  run 
transects  at  several  times  of  the  day  (Craig  1978; 
Thiollay  1978;  Millsap  1981).  Repeated  runs  of  the 
transect  in  a  sampling  period  increase  accuracy,  and 
at  least  three  repeated  runs  are  desirable.  The  esti- 
mated number  of  birds  per  transect  for  each  survey 
period  is  the  maximum  number  of  birds  seen  on  a 
single  run.  Man-made  structures  such  as  fences,  pow- 
erlines,  and  towers  provide  perches  that  draw  birds 
along  census  routes  and  cause  inflated  population 
estimates.  If  possible,  routes  adjacent  to  these  struc- 
tures should  be  avoided,  especially  when  contrasting 
abundance  among  areas.  This  bias  may  not  be  critical 
when  assessing  population  changes  in  the  same  area 
over  time.  Temporary  influxes  of  birds,  concentra- 
tions of  birds,  and  interspecific  interactions  may  also 
bias  counts  and  should  be  considered  in  data  analysis 
(Fuller  and  Mosher  1981). 

Because  the  detectable  distance  of  a  raptor  var- 
ies with  species  and  habitat  type,  substantial  biases 
occur  when  comparisons  are  made  among  areas 
or  habitats.  To  compensate  for  this  problem,  survey- 
ors can  reduce  strip  width  and  only  count  birds 
within  a  narrow  width  which  reflects  detection 
(Millsap  1981 ).  A  better  alternative  is  to  record  all 
birds  seen  and  their  perpendicular  distances  from 
the  transect  center  line  regardless  of  their  distance, 
and  to  analyze  these  data  with  the  line  transect  com- 
puter program  TRANSECT  (Burnham  et  al.  1980). 
This  program  compensates  for  variable  detection 
distances  and  is  available  to  all  BLM  employees  on 
the  computer  at  the  BLM  Service  Center.  The  main 
problem  with  program  TRANSECT  is  that  the  ideal 
sample  size  is  2s  40  individuals  (Burnham  et  al. 
1980;  Mikol  1980).  Transects  can  be  analyzed  with 
less  than  40  individuals,  but  resolution  is  greatly 
reduced.  The  problem  can  be  compensated  for  by 
(1)  increasing  transect  length,  (2)  increasing  the 
number  of  repetitions,  or  (  3 )  pooling  species  and 
habitats  (Millsap  pers.  commun.). 

A  criticism  of  road  surveys  is  that  transects  usu- 
ally are  not  random  and  data  are  not  collected  for 
statistical  analysis.  However,  appropriate  methods 
can  reduce  this  problem.  If  many  roads  exist  in  a 
large  area  of  homogeneous  habitat,  investigators  can 


326 


Raptors 


randomly  select  transects  from  the  array  of  available 
roads.  Often  these  conditions  do  not  exist,  and  divid- 
ing the  transects  into  segments  provides  a  viable 
alternative.  Transects  should  be  established  in  habitat 
and  land  use  types  that  need  to  be  surveyed.  Tran- 
sect segments  can  be  delineated  by  the  actual  dis- 
tance of  specific  habitat  types  along  the  transect, 
resulting  in  segments  of  unequal  length.  These  seg- 
ments of  specific  habitats  are  conceptually  linked 
together  and  treated  as  one  long  transect.  Data  are 
analyzed  with  line  transect  programs,  and  compari- 
sons among  years  and  habitat  types  can  be  made 
(Mikol  1980). 

Another  method  divides  transects  into  equal  1.6 
km  (1.0  mi.)  segments  (Koplin  pers.  commun.).  Each 
segment  is  treated  as  an  independent  sample  plot 
with  specific  habitat,  land  use,  and  .visibility  charac- 
teristics. The  entire  transect  is  surveyed,  and  the 
exact  location  of  each  bird  is  recorded  during  re- 
peated runs.  The  mean  number  (and  variance)  of 
birds  per  sample  plot  is  calculated;  comparisons  are 
made  among  years  within  and  among  habitats,  as 
well  as  among  habitats  within  years.  To  test  for 
clumping,  one  can  compare  the  number  of  plots 
with  various  numbers  of  birds  with  a  Poisson  Model. 
More  simply,  biologists  can  compute  the  coefficient 
of  dispersion  (Sokal  and  Rohlf  1981 )  of  the  plots. 


in  woodland  habitat,  and  5  to  26%  in  heavily  for- 
ested habitat.  However,  when  effective  transect 
width  was  adjusted  to  habitat  specific  detection  dis- 
tances within  each  vegetation  type,  accuracy  aver- 
aged 90%  (Millsap  pers.  commun.). 

Craighead  and  Craighead  (1956:49)  and  Fuller 
and  Mosher  ( 1981 )  discussed  factors  that  affect  ac- 
curacy of  road  surveys.  Results  are  more  reliable  for 
determining  relative  densities  of  conspicuous  species 
(buteos,  vultures,  and  eagles)  with  similar  detectabil- 
ity  in  open  country  of  homogeneous  habitat  types. 
Accuracy  greatly  decreases  when  many  raptor  spe- 
cies are  sampled  in  habitats  that  differ  greatly  in 
vegetation  and  terrain.  This  gives  relatively  inflated 
density  estimates  in  open  areas  compared  with  esti- 
mates in  the  more  vegetated  areas. 

Although  road  surveys  can  be  inaccurate,  they 
are  precise.  Coefficients  of  variation  calculated  from 
seven  surveys,  which  counted  medium-large  raptors 
in  a  0.4  km  (0.25  mi.)  strip  in  open  to  semi-open 
habitats,  ranged  from  6  to  14%  (Craighead  and 
Craighead  1956;  Schnell  1967;  Kochert  et  al.  1975, 
1976;  Craig  1978).  Millsap  and  LeFranc  (pers.  com- 
mun. )  found  that  their  road  transect  data  were  more 
precise  when  they  adjusted  effective  transect  width. 


CD  =   Jl 
Y 

This  value  will  be  near  1  if  the  distribution  is 
essentially  Poisson  (random),  >1  in  clumped  samples, 
and  <  1  in  uniformly  distributed  samples.  Habitat-spe- 
cific sample  plots  can  also  be  linked  to  form  a  long 
conceptual  transect  and  analyzed  with  the  line  tran- 
sect program  (Mikol  1980).  The  advantage  of  this 
method  is  that  each  segment  can  be  considered  a 
replication,  thus  increasing  sample  size.  In  addition, 
sample  plots  can  be  randomly  selected  on  a  stratified 
basis. 

Accuracy  and  Precision.  Road  counts  appear 
to  have  variable  accuracy  which  may  be  affected 
by  transect  width.  Craighead  and  Craighead  (1956) 
made  repeated  vehicle  surveys  during  the  winter 
in  a  partially  forested  area  in  Michigan.  They  be- 
lieved that  92%  of  the  buteos,  43%  of  the  northern 
harriers,  33%  of  the  Cooper's  hawks,  and  33%  of  the 
American  kestrels  were  located  by  using  a  belt  of 
0.4  km  (0.25  mi.)  on  both  sides  of  the  road.  In  Vir- 
ginia, B.  Millsap  and  M.  LeFranc  (pers.  commun.) 
found  much  lower  accuracy  from  preliminary  results 
of  one  summer-fall  field  season.  Models  of  perched 
raptors  representing  various  size  classes  were  ran- 
domly distributed  within  a  0.4-km  (0.25-mi.)  strip  on 
either  side  of  the  road.  By  using  this  belt,  the  pro- 
portion of  models  detected  ranged  (depending  on 
size)  from  11  to  64%  in  grassland  habitat,  2  to  32% 


Equipment.  No  special  equipment  is  required 
except  binoculars,  spotting  scope  (minimum  20x), 
thermometer,  wind  meter,  watch,  field  forms,  and 
a  vehicle.  Biologists  should  use  a  range  finder  to 
assist  in  estimating  perpendicular  distances  from  the 
transect. 


Cost.  Road  transects  are  inexpensive,  and  large 
areas  can  be  surveyed  by  two  people  in  a  short  time. 
Two  people  in  the  Snake  River  Birds  of  Prey  Area 
covered  112  km  (70  mi.)  of  transect  in  4  hours 
(1  person-day).  This  is  a  reasonable  estimate  for  a 
survey  day  considering  travel  time  to  and  from  the 
survey  site. 


Raptors 


327 


Training.  At  least  one  observer  must  be 
trained  in  raptor  identification  and  should  be  able  to 
identify  color  phases  and  age  classes  of  those  species 
in  the  survey  area.  Short  courses  in  raptor  identifica- 
tion are  recommended. 

Discussion.  Road  counts  are  a  fast,  inexpensive 
means  of  surveying  raptors  over  large  areas.  The 
method,  however,  is  ineffective  for  nest  surveys  and 
monitoring  owls  and  forest  raptors.  It  is  a  reliable 
means  for  assessing  occurrence,  species  composition, 
and  age  structure  of  populations  of  suitable  species. 
The  technique  is  useful  for  assessing  long-term  popu- 
lation trends  through  relative  measures;  however,  it 
is  plagued  with  many  biases  (Fuller  and  Mosher 
1981,  1986).  For  meaningful  surveys,  biologists  must 
work  within  the  limitation  of  those  biases,  primarily 
detectability  and  observer  biases,  and  use  only  expe- 
rienced observers  on  surveys. 

Aerial  Surveys. 

Essentially  two  types  of  aerial  surveys  can  be 
used  to  census  large  raptors.  Random  transects  are 
used  to  survey  raptors  that  are  scattered  throughout 
a  survey  area.  Searches  of  specific  habitats  (e.g.,  lakes 
and  riparian  areas)  are  used  to  survey  species  that 
congregate  in  specific  habitats.  These  surveys  are 
used  primarily  to  census  wintering  birds. 

Description.  Random  aerial  transects  have 
been  used  primarily  for  golden  eagles  and  occasion- 
ally for  other  large  raptors  (Fuller  and  Mosher 
1981).  Starting  from  a  randomly  selected  section 
line,  investigators  randomly  select  sample  transects 
from  all  possible  transects  that  can  be  placed  at  fixed 
0.8-km  (0.5-mi.)  intervals  in  the  survey  area  (Wrak- 
estraw  1972;  BLM  unpubl.  data).  Transects  should  be 
of  equal  length  (usually  80  km  [50  mi.])  and  sepa- 
rated enough  to  avoid  duplicate  sightings  (at  least 
6.4  km  [4.0  mi.]).  However,  because  of  irregular 
survey  area  boundaries,  some  investigators  have  used 
transects  of  unequal  lengths  (Boeker  and  Bolen 
1972;  Craig  1974).  They  considered  all  transect  lines 
collectively  as  a  large  transect  which  was  divided 
into  80-km  (50-mi.)  segments.  This  approach  is  valid 
only  if  the  transects  and  the  beginning  point  for  the 
first  segment  are  randomly  selected;  however,  biolo- 
gists should  avoid  this  approach  because  segments 
may  not  be  totally  independent. 

Transects  should  be  flown  at  about  15  m  (50  ft) 
to  92  m  (300  ft)  above  ground  level  (AGL)  at 
speeds  between  160-190  km/h  (100-120  mph) 
(Boeker  and  Bolen  1972;  Boeker  1974).  All  birds 
seen  on  each  side  of  the  plane,  regardless  of  dis- 
tance, should  be  counted  and  their  perpendicular 
distance  from  the  transect  should  be  recorded.  This 
increases  analytical  flexibility,  and  data  can  be  ana- 
lyzed within  a  fixed  belt  (Boeker  and  Bolen  1972), 
or  by  using  line  transect  analysis  (Mikol  1980). 


Because  birds  are  counted  on  both  sides  of  the 
aircraft,  two  observers  and  a  pilot  are  necessary.  The 
right  side  observer  sits  in  the  right  front  seat  and 
searches  from  the  center  line  out  to  the  right  side  of 
the  aircraft,  scanning  both  the  sky  and  the  ground. 
The  left  side  observer  sits  in  the  left  rear  seat  of  the 
aircraft  and  searches  the  left  side  of  the  aircraft.  To 
reduce  biases,  surveys  must  be  flown  in  as  few  con- 
secutive days  as  possible,  but  days  should  not  be 
excessively  long.  After  3-4  hours  of  flying,  fatigue 
and  eyestrain  affect  a  person's  ability  to  see  birds 
(Roseneau  1972);  a  survey  day  should  be  6-7  hours 
maximum  of  flying,  interrupted  with  a  necessary 
1-hour  rest  period.  A  1600-km  (1000-mi.)  survey  is 
most  efficiently  flown  in  3  consecutive  days. 

Kochert  et  al.  (1983)  analyzed  3  years  of  golden 
eagle  transect  data  by  using  both  the  belt  transect 
method  and  line  transect  analysis  (program  TRAN- 
SECT, Burnham  et  al.  1980).  The  belt  method  gave 
consistently  lower  estimates  of  wintering  population 
levels  than  TRANSECT — they  believed  the  results 
from  line  transect  analysis  yielded  a  more  realistic 
view  of  the  actual  population  size.  Belt  transects 
typically  underestimate  population  densities  when 
the  size  of  the  belt  is  too  wide  to  detect  all  animals 
within  the  belt  (Franzbeb  1981);  however,  effective 
transect  width  varies  from  survey  to  survey  with 
program  transect,  depending  on  probabilities  of  de- 
tection (Burnham  et  al.  1980).  Although  TRANSECT 
can  be  compromised  by  small  sample  sizes  (see 
Road  Count  section),  it  appears  to  be  a  reliable  ana- 
lytical method. 

Aerial  surveys  of  specific  habitats  are  used  pri- 
marily for  bald  eagles,  and  biologists  can  either  com- 
pletely search  habitat  patches  or  randomly  sample 
them.  In  designing  the  initial  inventory,  they  should 
identify  habitat  areas  from  aerial  photographs  where 


Aerial  surveys  can  easily  identify  bald  eagle  nests  in  forest 
habitats. 


328 


Raptors 


birds  are  known  to  concentrate  and  should  then 
chart  survey  routes  on  maps.  U.S.  Fish  and  Wildlife 
Service  (1983)  presents  detailed  guidelines  for  win- 
tering bald  eagle  surveys  and  stresses  the  need  for 
subsequent  surveys  throughout  the  winter.  Areas  can 
be  surveyed  by  either  helicopters  or  fixed-wing  air- 
craft at  a  recommended  height  of  31-93  m  ( 100-300 
ft)  AGL  and  speed  between  96-120  km/h  (60-70 
mph).  The  best  time  to  survey  feeding  activity  is  1-3 
hours  after  daylight.  Entire  habitat  areas  should  be 
flown  to  count  all  birds  present;  large  areas  can  be 
covered  by  systematic  parallel  transects  or  concen- 
tric contours  of  the  habitat  border  (Hancock  1964). 
Where  potential  bald  eagle  foraging  area  is  vast,  sur- 
veys along  systematic  transect  lines  should  be  spaced 
2.4  km  (1.5  mi.)  apart  and  extend  no  more  than  27 
km  (15  mi.)  from  roosts  (U.S.  Fish  and  Wildlife 
Service  1983). 

Often  areas  are  too  large  to  survey  all  habitats. 
In  this  situation,  biologists  should  consider  surveying 
a  random  sample  of  quadrats  of  the  area  (King  et  al. 
1972;  Grier  1977;  Grier  et  al.  1981).  This  method 
provides  population  estimates  with  confidence  limits 
and  allows  statistical  comparison  of  subsequent  sur- 
veys (Hodges  et  al.  1979).  If  the  objective  is  to  sur- 
vey a  large  amount  of  linear  habitat  (i.e.,  a  river), 
surveyors  should  segment  the  linear  habitat  and  ran- 
domly sample  the  segments,  preferably  on  a  stratified 
basis. 

Gilmer  et  al.  (1981)  and  Grier  et  al.  (1981) 
described  aircraft  certification,  preflight  preparations, 
safety  precautions,  and  special  considerations  for 
aerial  surveys.  Surveys  should  be  flown  on  clear  days 
with  no  precipitation.  Though  not  mandatory,  winds 
should  be  less  than  32  km/h  (20  mph).  Safety  is  a 
prime  concern,  and  only  experienced  pilots  and 
aircraft  with  sufficient  power  should  be  used.  Heli- 
copters should  not  be  used  in  canyons  and  areas 
with  many  powerlines.  Pilot  attitude  and  experience 
can  also  influence  error  (Hoskinson  1976). 


Accuracy  and  Precision.  Caughley  (1974) 
discussed  biases  of  aerial  transect  surveys.  He  be- 
lieved that  the  method  had  moderate  to  low  accu- 
racy because  many  animals  were  not  seen;  accuracy 
deteriorated  progressively  with  increasing  transect 
width,  cruising  speed,  and  altitude.  Although  inaccur- 
ate, transect  surveys  from  fixed-wing  aircraft  are 
precise.  Precision  (based  on  95%  confidence  inter- 
vals) of  fixed-wing  transect  surveys  in  Wyoming  and 
Idaho  for  golden  eagles  ranged  between  14.0  and 
16.5%  (Wrakestraw  1972,  1973;  BLM  unpubl.  data). 

Fixed-wing  aircraft  surveys  of  bald  eagles  in 
specific  habitats  on  flat  terrain  appear  to  be  fairly 
accurate.  Steenhof  ( 1976)  surveyed  an  area  for  bald 
eagles  from  both  the  ground  and  a  fixed-wing  aircraft 


on  the  same  day  and  obtained  nearly  the  same  re- 
sults. Also,  helicopter  surveys  may  be  more  accurate 
than  fixed-wing  surveys  (U.S.  Fish  and  Wildlife 
Service  1983). 

Slight  biases  may  occur  in  assessing  age  struc- 
ture of  bald  eagles  from  airplanes.  Hancock  ( 1964) 
believed  that  perched  adult  bald  eagles  were  more 
conspicuous  than  immatures  and  assumed  that  he 
underestimated  immatures  by  20-35%,  compared 
with  only  10-15%  for  adults.  These  biases  may  not 
be  significant  for  golden  eagle  surveys  because  plu- 
mage difference  between  adults  and  sub-adults  are 
not  striking,  and  all  perched  birds  must  be  flushed 
and  inspected  at  close  range  before  they  are  aged. 


Perched  adult  bald  eagles  are  mice  as  easy  to  see  as  im- 
mature ones. 


Equipment.  No  special  equipment  is  required 
except  for  the  aircraft.  Certain  aircraft  may  have 
advantages  over  others.  Readers  should  see  the 
Equipment  section  of  Aerial  Nesting  Surveys  dis- 
cussed in  this  chapter.  Some  observers  prefer  to  use 
a  small  cassette  tape  recorder  to  record  data.  Others 
carry  binoculars;  however,  I  have  found  them  of 
limited  value  for  aerial  surveys.  Also,  hearing  protec- 
tors can  reduce  fatigue. 

Training.  Observers  must  be  able  to  identify 
eagles  and  other  raptors  to  species  and  age  class.  See 
Training  in  the  Road  Transect  section. 

Cost.  Cost  can  vary  considerably  depending  on 
type  of  survey  and  aircraft.  In  the  Birds  of  Prey  Area 
it  has  taken  20  hours  (including  ferry  time)  for  two 
observers  to  survey  20  transects  80  km  (50  mi.) 
long  (BLM  unpubl.  data).  The  same  survey  in  a  heli- 
copter would  take  about  30  hours  for  two  people. 
Costs  of  surveying  specific  habitats  can  vary  greaUy 
depending  on  the  amount  of  ferry  time  and  aircraft 
used. 


Raptors 


329 


Discussion.  Both  random  aerial  transects  and 
aerial  surveys  of  specific  habitats  provide  a  reliable 
and  expeditious  means  of  assessing  raptor  presence, 
including  smaller  conspicuous  species  over  large 
areas.  Random  aerial  transects  are  an  effective  means 
to  count  large  raptors  over  broad  areas  as  demon- 
strated by  the  statewide  survey  of  golden  eagles 
in  Wyoming  (Wrakestraw  1972,  1973).  This  method 
is  also  reliable  for  assessing  age  structure  of  golden 
eagles  for  population  dynamics  studies.  Although  the 
method  has  inherent  biases,  they  affect  accuracy 
more  than  precision  (Caughley  1974),  making  it  a 
reliable  means  of  assessing  relative  densities  and 
monitoring  long-term  population  trends.  Aerial  sur- 
veys of  wintering  habitat  can  be  an  accurate  and 
reliable  means  of  surveying  bald  eagle  concentra- 
tions, and  randomly  sampled  quadrats  could  be  a  re- 
liable way  to  monitor  numbers  of  wintering  raptors 
in  large  areas. 

Weather,  raptor  activity,  time  of  day,  foliage 
condition,  and  snow  cover  can  affect  raptor  detecta- 
bility  and  bias  surveys.  Biases  can  be  reduced  by 
standardizing  survey  times,  season,  and  weather  con- 
ditions. Because  observer  competence  is  a  major 
source  of  variability  in  survey  results,  the  same  expe- 
rienced observers )  should  conduct  all  surveys  in  a 
particular  area  with  the  same  pilot  (U.S.  Fish  and 
Wildlife  Service  1983). 

Counts  at  Roosts  and  Colonies. 


All  kite  species  in  North  America  nest  either 
colonially  or  semi-colonially,  and  many  falconiforms 
and  strigiforms  use  communal  roosts  (Newton  1979; 
Allen  and  Young  1982).  This  section  focuses  on 
techniques  used  to  locate  colonies  or  communal 
roosts  and  to  estimate  bird  numbers.  Often  the  same 
techniques  are  used  to  survey  roosts  and  colonies. 

Description.  Before  searching  for  roosts  or 
colonies,  biologists  should  identify  potential  nesting 
or  roosting  habitat  from  aerial  photographs  and  vege- 
tation maps.  A  common  method  of  locating  roosts 
or  colonies  employs  non-random  ground  or  aerial 
searches  of  these  potential  areas  for  concentrations 
of  raptors  (Bildstein  1979;  Fuller  and  Mosher  1986; 
Glinski  and  Ohmart  1983).  A  few  investigators,  how- 
ever, have  searched  for  roosts  or  colonies  by  run- 
ning systematically  spaced  transects  (Sykes  1979). 
Roost  searches  are  most  effective  during  the  last  90 
minutes  of  daylight  (Keister  1981).  Eagle  roosts  may 
be  identified  by  low  level  aerial  photography.  Roosts 
or  colonies  can  also  be  located  by  following  birds 
back  from  feeding  areas  to  the  roost  or  colony  by  di- 
rect observation  or  with  the  aid  of  radio  telemetry 
(Keister  1981).  The  location  of  suspected  roosts  or 
colonies  can  be  obtained  by  observing  birds  from 
a  vantage  point,  and  the  exact  location  and  relative 
size  of  roosts  can  be  determined  from  the  distribu- 
tion of  feces  and  castings  (Keister  1981). 


The  most  common  method  of  enumerating  indi- 
viduals in  a  roost  or  colony  is  by  directly  counting 
the  number  of  birds  (Parker  1975;  Steenhof  1976; 
Keister  1981).  Night  roost  counts  for  diurnal  raptors 
should  be  made  between  60  minutes  before  sunset 
to  30  minutes  after  sunset  (Keister  1981).  Evening 
counts  are  usually  more  accurate  than  morning 
counts  (Hein  1961;  U.S.  Fish  and  Wildlife  Service 
1983).  Most  Mississippi  kite  nesting  colonies  are 
small  enough  that  investigators  can  enter  the  colony 
and  count  the  number  of  nesting  attempts  (Parker 
pers.  commun.);  however,  caution  should  be  taken 
to  minimize  disturbance. 


Bald  eagles  are  easily  counted  at  winter  roost  concentra- 
tions. 


Sometimes  roosts  are  inaccessible,  or  birds  can- 
not be  seen  (e.g.,  northern  harriers),  and  biologists 
must  estimate  numbers  of  individuals  by  counting 
birds  as  they  fly  to  and  from  roosts  (Hein  1961;  Bild- 
stein 1979;  U.S.  Fish  and  Wildlife  Service  1983). 
This  technique  has  an  advantage  in  that  it  minimizes 
disturbance.  Hein  (1961)  gave  detailed  procedures 
for  conducting  roost  flight  counts.  Roosting  flight 
counts  should  be  made  from  90  minutes  before  sun- 
set to  30  minutes  after  sunset,  with  a  subsequent 
count  the  next  morning  from  30  minutes  before  sun- 
rise to  90  minutes  after  sunrise.  The  maximum  num- 
ber of  the  two  counts  should  be  used.  Sometimes 
several  counters  stationed  around  the  roost  are  re- 
quired when  birds  depart  in  different  directions 
(Weller  et  al.  1955). 

Biologists  sometimes  use  flight  counts  to  census 
nesting  colonies  (Fuller  and  Mosher  1981 ).  Glinski 
and  Ohmart  (1983)  estimated  Mississippi  kite  num- 
bers in  colonies  by  counting  birds  during  courtship, 
foraging,  and  predator-mobbing  flights.  They  used 
these  numbers  and  counts  of  nesting  attempts  to 
estimate  the  nonbreeding  component.  Accurate 
counts,  however,  can  only  be  made  by  many  peri- 
odic visits  to  the  colony  or  roost  starting  early  in  the 
season  (Bildstein  1979;  Parker  and  Ogden  1979). 
Although  long-eared  owls  can  be  counted  at  roosts, 


330 


Raptors 


distribution,  and  relative  size  of  colonies  and  roosts 
are  more  important  for  most  monitoring  and  inven- 
tory programs  than  absolute  counts  or  changes  in 
absolute  numbers  of  individuals.  Changes  in  numbers 
of  individuals  may  not  reflect  changes  in  the  popula- 
tion. Many  colonial  raptors  tend  to  be  nomadic 
(Newton  1979),  and  changes  in  bird  numbers  in  a 
colony  or  roost  may  only  reflect  a  local  or  regional 
shift.  To  account  for  any  inter-roost  shifts,  investiga- 
tors should  periodically  make  simultaneous  counts  at 
all  roosts  in  an  area. 

Because  many  factors  influence  counts  at  roosts 
(e.g.,  weather  and  shifts  in  prey  abundance),  counts 
should  be  considered  indexes  (Bildstein  1979;  Fuller 
and  Mosher  1981;  Keister  1981;  U.S.  Fish  and  Wild- 
life Service  1983).  Although  direct  counts  are  in- 
dexes, and  indirect  counts  (e.g.,  flight  counts)  may 
underestimate  the  number  of  birds  (U.S.  Fish  and 
Wildlife  Service  1983),  both  methods  provide  a  reli- 
able means  of  assessing  relative  abundance.  Counts 
at  night  roosts  provide  a  more  accurate  index  of 
wintering  bald  eagle  abundance  than  day  counts  on 
feeding  areas  (U.S.  Fish  and  Wildlife  Service  1983). 
Compared  with  other  census  techniques,  roost  and 
colony  counts  are  very  reliable.  As  with  any  census 
technique,  there  are  problems  with  detectability,  and 
when  communally  roosting  birds  bunch  up  they  are 
difficult  to  count.  However,  accuracy  of  counting 
methods  can  be  increased  by  taking  the  average  of 
numerous  repeated  counts  made  during  a  sampling 
period. 


Long-eared  owl. 


Millsap  (pers.  commun. )  found  that  morning  counts 
of  birds  flying  to  the  roost  were  very  similar  to  di- 
rect counts  of  birds  in  the  roost.  This  method  may 
also  work  for  short-eared  owls.  Some  investigators 
have  counted  these  owls  by  walking  through  roosts 
and  flushing  individuals  (Clark  1975),  but  this  is  not 
desirable  because  of  disturbance. 

Keister  (1981)  discussed  other  methods  of 
enumerating  roosting  eagles  such  as  thermal  infrared 
remote  sensing,  radar,  and  infrared  photography. 
These  methods  have  limited  utility  for  a  BLM  moni- 
toring or  inventory  program.  However,  in  the  Birds 
of  Prey  Area,  biologists  have  experimented  with 
standard  photography  using  1000  ASA  print  film  to 
count  communally  roosting  ravens  on  transmission 
towers  and  have  had  encouraging  results  (BLM  un- 
publ.  data). 

Discussion.  This  method  is  effective  in  assess- 
ing presence  of  birds  in  specific  areas.  It  is  also  use- 
ful in  population  dynamics  studies,  and  combined 
with  other  surveys  over  a  large  area,  it  may  also  be  a 
means  of  assessing  long-term  trends.  The  number, 


Relative  Abundance — Nesting  Surveys 
General  Considerations. 


Survey  Design.  When  designing  an  inventory, 
investigators  should  gather  all  available  relevant 
historical  information  on  the  area  to  be  surveyed 
(Fuller  and  Mosher  1981).  This  information  can  be 
obtained  through  the  literature,  records  (e.g.,  Federal 
and  State  agencies,  museums,  or  the  Laboratory  of 
Ornithology  at  Cornell  University),  and  inquiries  of 
local  residents,  falconers,  wildlife  managers  and 
biologists,  and  amateur  ornithologists.  Because  many 
species  have  relatively  restricted  nest  site  require- 
ments and  use  structures  that  last  many  years  (e.g., 
ledges),  historic  data  can  lead  one  to  specific  nest 
sites.  Historic  information  can  provide  data  on 
species  occurrence  and  distribution. 

Inventory  or  monitoring  objectives  will  deter- 
mine the  sampling  approach.  In  certain  unique  habi- 
tats or  areas  of  special  concern,  like  the  Birds  of 
Prey  Area,  biologists  may  be  required  to  design  sys- 
tematic searches  of  the  entire  area  staggered  over 
the  nesting  period  (U.S.  Bureau  of  Land  Management 
1979).  However,  in  most  situations,  investigators 


Raptors 


331 


will  not  be  able  to  survey  the  entire  area,  and  the 
most  efficient  approach  in  terms  of  time  and  people 
is  to  sample  the  area  in  a  simple  random  or  stratified 
random  manner. 

Biologists  should  consider  quadrat  sampling  to 
estimate  populations  over  large  areas  (King  et  al. 
1972;  Grier  1977;  Postovit  1979;  Grier  et  al.  1981; 
Gilmer  and  Stewart  1984).  Quadrats  are  randomly 
selected  from  grids  on  maps  or  aerial  photos.  Selec- 
tions can  be  simple  random  or  stratified  to  habitat 
characteristics  or  quality.  Investigators  can  survey 
either  the  entire  quadrat  or  just  the  primary  nesting 
habitat  (e.g.,  shorelines  or  cliffs)  within  the  quadrat. 
Quadrat  sampling  provides  population  estimates  with 
confidence  limits  (Grier  1977),  and  quadrats  can  be 
easily  monitored  with  statistical  comparisons  among 
surveys  (Hodges  et  al.  1979).  Surveys  that  include 
more  than  primary  habitat  may  be  more  precise 
(Leighton  et  al.  1979). 

To  survey  long  stretches  of  linear  habitat  (e.g., 
cliff,  shoreline,  or  riparian  areas),  biologists  can  di- 
vide the  habitat  into  equal  segments  and  select  seg- 
ments to  survey  on  a  stratified  random  basis.  Each 
segment  can  be  treated  as  a  sample  plot  for  statisti- 
cal comparisons.  Random  sampling  of  segments  gives 
an  inference  of  the  entire  habitat  stretch  sampled, 
but  non-randomly-sampled  segments  give  informa- 
tion only  on  the  segment  sampled. 

Locating  Nests  and  Nesting  Areas.  Successful 
location  of  raptor  nests  or  nesting  areas  requires 
knowledge  of  the  distribution,  habitat  relationships, 
nesting  chronology,  spacing,  and  nesting  behavior  of 
the  species  or  group  of  species  to  be  surveyed.  This 
knowledge  and  understanding  can  be  derived  from 
several  sources  (Craighead  and  Craighead  1956;  Call 
1978;  Newton  1979;  and  various  technical  notes 
available  from  the  BLM  Service  Center). 


Typical  ferruginous  hawk  nest  on  rock  outcrop. 


For  a  territory  to  be  classified  "occupied,"  it 
must  satisfy  the  criteria  for  occupancy  (see  Produc- 
tivity section).  There  are  a  variety  of  visible  signs 
that  advertise  raptor  occupancy  and  nesting  attempts 
(Craighead  and  Craighead  1956;  Call  1978;  Jones 
1979;  Fuller  and  Mosher  1981,  1986;  Reynolds  1982; 
Forsman  1983b).  These  signs  differ  with  species  and 
nesting  habitat.  In  cliff  areas,  rocks  stained  with  feces 
(whitewash)  are  the  primary  feature.  Decorated  or 
repaired  nests,  perches,  plucking  perches,  and  regur- 
gitated pellets  are  also  important.  In  forested  areas, 
stick  nests,  perches,  molted  feathers,  plucking 
perches,  and  prey  remains  below  a  nest  are  common 
signs  of  occupancy. 

Behavioral  observations  are  also  important  for 
locating  nests.  These  include  courtship  displays,  food 
carries,  food  transfers,  copulation,  nest  building,  ac- 
tive defense  and  calling  by  adults,  and  food  begging 
by  young  (Craighead  and  Craighead  1956;  Call  1978; 
Jones  1979;  Fuller  and  Mosher  1981;  Reynolds  1982; 
Mindell  1983)-  A  single  observation  of  an  unrepaired 
nest  does  not  mean  the  site  is  vacant  (Fuller  and 
Mosher  1981).  Nesting  areas  should  be  visited  more 
than  once  to  confirm  vacancy. 

Biases.  Many  factors  influence  detection  of 
nests  and  nesting  areas  (Fuller  and  Mosher  1981, 
1986).  Visibility  biases  are  influenced  by  light  condi- 
tions, time  of  day,  weather,  type  of  nesting  habitat, 
seasonal  changes  in  the  habitat,  seasonal  changes  in 
the  behavior  of  raptors,  and  observer  competence 
and  experience  (Craighead  and  Craighead  1956; 
Henny  et  al.  1974,  1977;  Postovit  1979;  Grier  et  al. 
1981).  Raptor  behavior  can  bias  observations;  secre- 
tive raptors  that  stay  away  from  nests  are  easily 
missed,  whereas  extremely  aggressive  birds  are  easily 
found  (Craighead  and  Craighead  1956;  Call  1978; 
Postovit  1979).  Also,  pairs  arriving  late  and  those 
that  have  multiple  broods  (e.g.,  barn  owls)  or  renest 
can  influence  accuracy  because  they  are  often 
missed. 


Aerial  Surveys. 


Description.  Boeker  (1970),  Hickman  (1972), 
Roseneau  (1972),  Grier  (1977),  Fraser  (1978), 
Leighton  et  al.  (1979),  and  Grier  et  al.  (1981) 
described  techniques  for  surveying  cliff  and  tree 
nesting  raptors  with  fixed-wing  aircraft.  They 
discussed  types  of  aircraft,  safety  precautions,  and 
costs,  and  stressed  the  need  to  retain  only 
experienced  pilots.  Observers  fly  predetermined 
routes  and  search  for  nests,  presence  of  adults,  or 
other  signs  of  occupancy.  Surveys  for  conspicuous 
bald  eagle  and  osprey  nests  are  flown  relatively  fast 
at  speeds  between  112  and  160  km/h  (70-100  mph) 
at  about  100  m  (325  ft)  AGL  (Grier  et  al.  1981). 
Cliff  nesting  surveys  tend  to  be  flown  slower 


332 


Raptors 


(between  72  and  112  km/h  [45  and  70  mph])  at 
distances  of  15-62  m  (50-200  ft)  from  the  cliff 
(Boeker  1970;  Hickman  1972;  Roseneau  1972). 
Observers  fly  slightly  below  the  rim,  scan  the  cliff, 
and  search  for  evidence  of  occupancy.  Often  a 
second  pass  is  required  to  determine  the  nest 
contents,  and  these  approaches  should  be  made 
going  "down  hill"  after  checking  air  movement 
(Boeker  1970).  Additional  passes  in  a  tiered  fashion 
descending  down  the  cliff  are  necessary  to 
completely  cover  large  cliff  faces. 

White  and  Sherrod  (1973)  and  Carrier  and 
Melquist  (1976)  described  survey  techniques  and 
safety  precautions  for  surveying  from  a  helicopter. 
Techniques  for  locating  nests  are  similar  to  those  for 
fixed-wing  aircraft.  A  desired  speed  to  survey  cliffs 
is  48-70  km/h  (30-40  mph),  and  open  nests  on  pow- 
erline  structures  can  be  adequately  surveyed  at  80- 
96  km/h  (50-60  mph).  Once  a  suspected  nest  is  lo- 
cated, observers  can  hover  or  slow  to  speeds  less 
than  32  km/h  (20  mph)  to  provide  a  better  view  of 
the  nest.  I  have  found  that  glare  on  the  helicopter 
bubble  reduces  one's  ability  to  see  nests;  this  can  be 
remedied  by  removing  the  door.  White  and  Sherrod 
(1973)  advised  approaching  nests  from  upwind  to 
avoid  hitting  flushed  birds;  however,  Carrier  and 
Melquist  (1976)  did  not  believe  this  was  a  problem. 
Many  times  downwind  approaches  are  necessary  in 
canyon  surveys  for  hovering  or  maneuvering  at  slow 
speeds.  Biologists  and  pilots  should  always  be  alert 
because  occasionally  birds  will  attack  the  aircraft  or 
fly  in  its  path  (Nelson  pers.  commun.). 


Surveys  can  be  systematic  grids  of  an  area,  ran- 
dom quadrat  samples  of  an  area,  or  random  or  non- 
random  surveys  of  nesting  habitat  (see  General  Con- 
siderations). Systematically  spaced  transects  are  used 
primarily  to  survey  entire  areas  for  tree-nesting  spe- 
cies (Petersen  1979).  Both  fixed-wing  and  helicopter 
nest  surveys  should  be  made  with  at  least  two  ob- 
servers on  calm  (winds  less  than  24  km/h  [15  mph]) 
days  (Carrier  and  Melquist  1976;  Grier  et  al.  1981). 
Helicopters,  however,  are  more  stable  and  effective 
in  winds  greater  than  24  km/h  (15  mph)  (Call 
1978).  Surveys  can  be  conducted  at  any  time  of  day, 
but  mornings  are  usually  the  best  time  for  canyon 
surveys  because  of  the  reduced  wind  conditions. 
Although  Grier  et  al.  ( 1981 )  found  no  difference  in 
visibility  due  to  season,  if  possible,  observers  should 
survey  tree-nesting  raptors  before  foliation  in  decid- 
uous forests.  Also,  possible  errors  can  be  avoided 
by  observers  recording  observations  independently. 


Accuracy  and  Precision.  Two  types  of  error 
characterize  aerial  nesting  surveys:  failure  to  locate 
occupied  sites  and  failure  to  properly  classify  them. 
Estimates  are  conservative  at  best  because  some 
birds  and  nests  are  not  seen,  and  visibility  biases  are 


common  and  often  unavoidable  (Caughley  1974; 
Grier  1977). 

Accuracy  varies  with  species  surveyed,  observ- 
ers, and  number  of  flights  in  each  survey.  Henny 
et  al.  (1974,  1977)  calculated  that  only  about  64% 
of  the  occupied  osprey  sites  on  various  substrates 
were  located  in  a  single,  fixed-wing  flight.  Accuracy 
for  similar  single-flight  surveys  of  cliff-nesting  golden 
eagles  was  estimated  at  81%  (Phillips  et  al.  1984). 
Accuracy  of  single  flight  fixed-wing  surveys  for  occu- 
pied breeding  areas  of  tree  nesting  bald  eagles  in 
three  areas  ranged  between  84  and  85%  (Leighton 
et  al.  1979;  Grier  et  al.  1981;  Fraser  et  al.  1983). 
Accuracy  increased  to  98%  by  adding  a  second  flight 
(Grier  et  al.  1981).  Although  differences  were  not 
significant,  the  proportion  of  bald  eagle  tree  nests 
found  by  experienced  observers  ranged  from  67  to 
87%  (Grier  et  al.  1981 ).  Specific  data  are  lacking  for 
the  smaller,  less  conspicuous  raptors  (gyrfalcons, 
rough-legged  hawks,  and  red-tailed  hawks).  Accuracy 
of  surveys  for  these  species  may  be  lower  than  those 
for  eagles,  considering  accuracy  in  locating  osprey 
pairs  (Henny  1974,  1977)  and  raven  tree  nests 
(12%;  Grier  et  al.  1981). 

Little  has  been  published  on  accuracy  of  nest 
surveys  from  helicopters.  In  a  single-flight  survey, 
Wier  ( 1982)  located  seven  of  the  eight  (88%  ) 
golden  eagle,  rough-legged  hawk,  and  gyrfalcon  nest- 
ing attempts  on  a  cliff  in  Alaska.  Birds  of  Prey  Area 
personnel  located  96%  of  a  sample  of  48  golden- 
eagle  nesting  attempts  during  a  single  flight  (BLM 
unpubl.  data).  Accuracy  of  occupancy  classifications 
may  be  related  to  the  proportion  of  pairs  laying  eggs. 
In  a  single-flight  occupancy  survey  for  cliff-nesting 
golden  eagles  during  incubation,  Birds  of  Prey  Area 
personnel  correctly  classified  88%  of  77  sites  when 
only  67%  of  the  pairs  laid  eggs;  however,  all  sites 
were  correctly  classified  during  a  year  when  all  pairs 
bred  (BLM  unpubl.  data).  Most  of  the  misclassified 
nesting  areas  contained  nonbreeding  pairs. 

Timing  and  classification  criteria  can  also  influ- 
ence accuracy  of  aerial  surveys.  Surveys  for  occu- 
pancy are  most  accurate  after  completion  of  all 
clutches,  but  before  the  first  brood  hatches  (Fraser 
et  al.  1983;  Steenhof  1986).  Fraser  (1978)  increased 
accuracy  in  classifying  bald  eagle  breeding  areas  for 
occupancy  from  85%  to  89%  when  criteria  required 
that  only  one  adult  instead  of  two  be  seen  in  the 
breeding  area.  However,  sites  where  one  bird  is  ob- 
served must  satisfy  the  criteria  for  occupancy  (see 
Productivity  section),  and  biologists  must  note 
whether  one  or  two  birds  were  seen  on  their  sur- 
veys so  data  can  be  compared  with  other  studies.  If 
they  find  no  evidence  of  egg  laying,  no  freshly  deco- 
rated nests,  and  only  a  single  adult,  the  site  should 
be  revisited,  preferably  on  the  ground,  to  verify  oc- 
cupancy (Steenhof  1986). 


Raptors 


333 


Fixed-wing  aerial  surveys  are  fairly  precise  for 
locating  occupied  bald  eagle  territories  (Grier  1977; 
Leighton  et  al.  1979).  However,  the  precision  of 
extrapolating  population  estimates  from  random  aer- 
ial survey  blocks  appears  low.  Leighton  et  al.  (1979) 
calculated  that  their  aerial  surveys  of  random  blocks 
could  not  detect  population  changes  of  less  than 
27% .  Surveyors  can  refine  precision  by  stratified  ran- 
dom sampling  (Grier  1977).  However,  Grier  and 
Hamilton  (1978)  observed  no  reduction  in  variance 
when  they  subsampled  clusters  of  random  samples 
on  a  stratified  basis  using  optimum  allocation  of  sam- 
ples. Fraser  (1978)  found  that  when  he  included  a 
single  adult  bald  eagle  at  the  nest,  or  single  adults  or 
pairs  away  from  the  nest  in  his  criteria  for  site  occu- 
pancy, it  increased  accuracy  but  reduced  precision. 

Equipment.  Often  local  availability  will  deter- 
mine the  type  of  aircraft  to  be  used.  Smaller  fixed- 
wing  aircraft  (e.g.,  Piper  Super  Cub)  may  be  better 
for  raptor  nest  surveys  than  larger  ones  (Cessna  180 
or  182)  because  smaller  craft  are  slower,  more  ma- 
neuverable,  and  offer  better  observer  visibility  (Hick- 
man 1972;  Roseneau  1972;  Grier  et  al.  1981).  For 
helicopter  surveys,  the  smaller  jet-turbine  aircraft 
(e.g.,  Hiller/Soloy)  may  be  more  effective  because  of 
lower  hourly  cost  and  the  ability  to  hover  in  unsta- 
ble air  conditions.  However,  if  surveys  involve  a 
large  amount  of  ferry  distance,  the  faster  medium- 
sized  helicopters  (e.g.,  Bell  Jet  Ranger)  are  more 
efficient. 

Cost.  Depending  on  the  aircraft,  fixed  wing 
surveys  cost  1/3  (Cessna  180)  to  1/10  (Piper  Super 
Cub)  of  an  equivalent  ground  survey.  A  Super  Cub 
took  25%  less  survey  time  per  linear  mile  (1.6  km) 
of  cliff  to  locate  golden  eagle  nests  than  did  the 
Cessna  180  and  98%  less  time  than  ground  surveys 
(Boeker  1970;  Hickman  1972).  Quadrat  surveys 
for  tree  nesting  bald  eagles  covered  150  km"  (58 
mi.  )/h  using  a  Cessna  180  and  71  km    (28  mi.  )/h 
with  a  Piper  Super  Cub  (King  et  al.  1972;  Grier 
1977;  Hodges  et  al.  1979).  By  cluster  sampling,  Grier 
and  Hamilton  (1978)  reduced  their  flight  time  by 
15%.  Considering  salaries,  travel,  and  per  diem,  heli- 
copter surveys  cost  about  5%  less  than  that  of  equiv- 
alent ground  surveys,  and  they  were  completed  10 
times  faster  (Carrier  and  Melquist  1976;  BLM  un- 
publ.  data). 

Discussion.  Aerial  surveys  are  most  successful 
with  large  conspicuous  raptors  that  have  high  visibil- 
ity, limited  nesting  habitat,  and  synchronized  nesting 
cycle  (Henny  et  al.  1977;  Call  1978;  Leighton  et  al. 
1979).  Although  aerial  nesting  surveys  have  been 
successful  for  ospreys,  gyrfalcons,  rough-legged 
hawks,  and  goshawks  (McGowan  1975;  Call  1978; 
Fuller  and  Mosher  1 98 1 ),  they  are  most  accurate  for 
bald  and  golden  eagles.  Accuracy  appears  to  de- 
crease for  smaller,  less  conspicuous  raptors.  Al- 
though peregrine  falcons,  snowy  owls,  great  horned 


owls,  and  red-tailed  hawks  have  also  been  surveyed 
by  aircraft,  these  surveys  have  usually  supplemented 
ground  searches  (Fuller  and  Mosher  1981).  The 
technique  has  limited  value  on  cavity-nesting  raptors 
(e.g.,  prairie  falcons,  owls;  Call  1978). 

Aerial  surveys  can  be  more  effective  than 
ground  surveys  because  they  cost  less,  they  are  an 
expeditious  means  of  surveying  large  areas,  and  they 
provide  ready  access  to  remote  country.  Investiga- 
tors, however,  must  consider  many  factors  when 
choosing  an  aircraft.  Although  more  costly  in  com- 
parison, helicopters  provide  more  visibility  (they  can 
hover  to  allow  a  better  view  of  nests),  better  maneu- 
verability and  stability,  and  in  many  instances, 
greater  accuracy  than  fixed-wing  aircraft.  Further- 
more, helicopters  can  land  in  the  field  for  ground  in- 
spection of  nests. 

Aerial  surveys  have  many  biases  and  limitations 
(Fuller  and  Mosher  1981,  1986).  Detectability  and 
classification  biases  are  most  important,  and  can  be 
reduced  by  supplementing  flights  with  follow-up 
ground  searches  or  flying  additional  surveys.  Also, 
problems  with  precision  can  be  reduced  with  larger 
sample  sizes.  The  technique  is  fairly  repeatable,  and 
if  biologists  work  within  the  limitations  of  these 
biases,  it  can  be  a  reliable  means  of  surveying  large 
areas. 

Calling  Surveys. 

Detection  of  calls  has  been  used  primarily  to 
survey  owls,  but  recently  the  technique  has  been  ap- 
plied to  some  falconiforms  (Fuller  and  Mosher 
1 98 1 ).  Every  North  American  owl  species,  except 
the  common  barn-owl,  snowy  owl,  short-eared  owl, 
and  northern  hawk-owl,  has  responded  to  human 
broadcasted  owl  vocalizations  (Foster  1965;  Martin 
1973;  Call  1978;  Fuller  and  Mosher  1981;  Hayward 
1983).  Falconiforms  presently  known  to  respond 
to  these  calls  are  the  red-shouldered  hawk,  red-tailed 
hawk,  broad-winged  hawk,  goshawk,  Cooper's  hawk, 
sharp-shinned  hawk,  common  black-hawk,  gray  hawk, 
and  zone-tailed  hawk  (Fuller  and  Mosher  1981;  Ro- 
senfield  et  al.  1985;  Mosher  et  al.  unpubl.  ms.;  Mill- 
sap  pers.  commua). 

Description.  Two  basic  methods  are  used  to 
locate  raptors  by  their  calls.  The  first  is  to  listen  for 
calls  and  note  their  location.  This  method,  used 
mainly  to  supplement  other  survey  techniques,  en- 
tails walking  or  driving  through  an  area  and  periodi- 
cally listening  for  calls.  The  other  method  is  to  elicit 
responses  either  by  imitating  calls  or  broadcasting 
recordings  of  calls;  responses  can  be  either  return 
calling  or  the  silent  approach  of  a  bird.  This  method 
involves  periodic  broadcasting  of  calls  followed  by 
periods  of  silence  during  which  surveyors  look  and 
listen  for  responses  (Forsman  et  al.  1977;  Fuller  and 
Mosher  1981;  Forsman  1983b). 


334 


Raptors 


Raptors  can  be  surveyed  by  the  continuous  tran- 
sect method  (Forsman  et  al.  1977;  Forsman  1983b) 
or  the  calling  station  method  (Fuller  and  Mosher 
1981).  In  the  continuous  transect  method,  investiga- 
tors walk  along  a  road  and  play  recorded  calls  at 
fixed  intervals  (often  every  15-20  sec).  After  a  bird 
responds,  surveyors  move  a  fixed  distance  (often  0.8 
km  or  0.5  mi.)  down  the  road  (presumably  out  of 
the  responding  owl's  territory)  before  calling  again. 
A  relative  density  estimate  of  the  number  of  pairs 
per  linear  distance  is  calculated  from  this  method. 


Is   it    live 
or 
I  s  it  Memorex  ? 


The  calling  station  method  involves  broadcast- 
ing calls  from  established  stations.  Stations  can  be 
non-randomly  placed  in  specific  habitat  types  or 
placed  at  fixed  distances  to  ensure  full  coverage  of 
an  area  (Siminski  1976;  Smith  1978;  Springer  1978). 
Stations  can  be  systematically  placed  along  a  road 
or  transect  (sometimes  called  point  count  transects; 
Simpson  1972;  Ellison  1980;  Mosher  et  al.  unpubl. 
ms.),  along  parallel  transects  to  form  a  grid  (Cink 
1975),  or  in  quadrats  randomly  selected  from  a  grid 
(Nowicki  1974).  Calls  are  usually  broadcast  for  15 
seconds  followed  by  a  1 5-60  second  listening  period, 
after  which  the  sequence  is  repeated.  Investigators 
usually  call  at  each  stop  for  4  to  1 5  minutes.  Mosher 
et  al.  (unpubl.  ms.)  recommended  that  observers 
look  and  listen  for  an  additional  5  minutes  before 
proceeding  to  the  next  stop.  Also  at  each  stop  they 
broadcast  to  one  side  of  the  road,  turned  the  speaker 
180  degrees  and  broadcast  again,  and  repeated  this 
process  until  three  recordings  had  been  broadcast  to 
each  side  of  the  road.  In  contrast,  Balding  and  Dibble 
(1984)  played  calls  through  four  speakers  each 
pointed  in  a  different  direction.  Relative  densities 
derived  by  this  method  can  be  reported  as  the  num- 
ber of  birds  per  calling  station,  number  of  birds  per 
linear  distance,  or  number  of  birds  in  an  area. 


Call  surveys  should  be  conducted  under  rela- 
tively calm  conditions  with  no  precipitation  or  fog 
(Simpson  1972;  Forsman  et  al.  1977;  Smith  1978; 
Springer  1978;  Forsman  1983b).  Minimum  data  to  be 
recorded  at  each  calling  stop  are  weather  (at  least 
temperature,  precipitation,  wind  speed,  and  cloud 
cover),  moon  phase,  terrain,  vegetation,  and  raptor 
response.  If  a  response  is  elicited,  investigators 
should  record  species,  approximate  location,  time  of 
response,  type  of  response  (vocal  or  visual),  habitat 
type,  and  sex  (if  distinguishable).  However,  inexperi- 
enced observers  may  not  be  able  to  distinguish  be- 
tween sexes. 

Most  researchers  call  owls  at  night,  usually  be- 
tween 30  minutes  after  sunset  to  1  hour  before  sun- 
rise (Bell  1964;  Simpson  1972;  Nowicki  1974; 
Forsman  et  al.  1977;  Hayward  1983);  however,  Sim- 
inski (1976)  and  Springer  (  1978)  successfully  called 
great  horned  owls  during  daylight.  Little  information 
exists  on  when  owl  calling  is  most  successful.  Some 
workers  believe  time  of  night  has  no  effect  (Cink 
1975);  however,  nesting  barred  owls  appeared  most 
responsive  near  the  middle  of  the  night  (Smith 
1978),  and  wintering  eastern  screech-owls  re- 
sponded more  towards  dawn  (Beatty  1977). 

Calling  for  nesting  diurnal  raptors  is  usually 
conducted  between  30  minutes  after  sunrise  to  mid- 
morning,  and  birds  may  be  more  responsive  toward 
mid-morning  (Balding  and  Dibble  1984;  Mosher  et 
al.  unpubl.  ms.;  Millsap  pers.  commun.).  Responses  to 
calls  appear  to  be  immediate,  mostly  within  5  min- 
utes of  broadcasting  (Rosenfield  et  al.  1985;  Mosher 
et  al.  unpubl.  ms.).  Although  hawks  respond  to  calls 
during  all  stages  of  the  nesting  cycle  ( Mosher  et 
al.  unpubl.  ms.),  Cooper's  hawks  appeared  to  be  less 
responsive  during  incubation  (Rosenfield  et  al. 
1985). 

If  the  objective  is  to  estimate  the  number  of 
birds  present  in  an  area,  it  is  important  to  know  the 
range  over  which  the  sample  is  being  taken.  How- 
ever, biologists  should  avoid  using  the  absolute 
measure,  number  of  birds  per  unit  area,  because  not 
all  birds  respond  to  calls.  However,  some  researchers 
have  calculated  number  of  birds  responding  per 
unit  area  by  dividing  the  area  contained  in  an  arbi- 
trarily delineated  study  area  by  the  number  of  birds 
responding  to  calls  (Siminski  1976;  Springer  1978; 
Smith  1978).  This  measure  is  meaningless  if  the  en- 
tire area  is  not  covered  by  vocalizations.  Conversely, 
it  may  overestimate  the  density  if  it  draws  birds  in 
from  outside  the  study  area.  To  estimate  the  number 
of  birds  in  a  large  area,  Nowicki  (1974)  randomly 
selected  quadrats  from  the  area  and  calculated  a 
density  based  on  the  number  of  owls  responding, 
divided  by  the  area  contained  in  all  sample  quadrats. 
Total  number  of  owls  were  estimated  by  multiplying 
this  figure  by  the  size  of  the  total  study  area.  This 
method  is  valid  if  quadrats  are  randomly  selected, 


Raptors 


335 


the  entire  area  of  the  quadrat  is  covered  by  the  calls, 
and  no  owls  from  adjacent  quadrats  are  drawn  in 
and  respond  to  the  calls. 

Some  researchers  estimated  the  distance  their 
broadcast  calls  carried  and  used  this  "effective  listen- 
ing distance"  to  determine  the  size  of  their  sampling 
area  (Cink  1975).  This  approach  has  many  biases. 
Vegetation  and  topography  influence  the  distance 
that  calls  carry,  and  effective  listening  distances 
based  on  human  detection  may  be  inaccurate  be- 
cause of  differences  between  raptor  and  human  hear- 
ing (Nowicki  1974;  Fuller  and  Mosher  1981). 
However,  one  should  know  the  audio  range  and 
effectiveness  of  equipment.  Mosher  et  al.  (unpubl. 
ms.)  measured  sound  levels  from  their  equipment 
and  suggested  that  vocalizations  would  be  audible  to 
about  800  m  ( 2640  ft ),  at  which  distance  sound 
levels  were  similar  to  background  noise  (30-40  dB); 
however,  they  believed  calls  could  carry  up  to  1.6 
km  (1.0  mi.). 

Sometimes  owls  are  difficult  to  locate,  and  when 
many  birds  simultaneously  respond  to  the  calls,  indi- 
viduals or  pairs  are  difficult  to  differentiate.  Also 
many  owls,  particularly  flammulated  owls,  have  ven- 
triloquial  abilities  which  make  it  extremely  difficult 
to  pinpoint  the  source  of  the  call  without  considera- 
ble effort  (Millsap  pers.  commun. ).  To  more  pre- 
cisely locate  calling  barred  owls,  Bell  (1964) 
triangulated  from  two  stations.  When  western 
screech-owl  and  elf  owl  nests  were  concentrated  ( 50 
m  [165  ft]  apart)  in  limited  riparian  zones,  Johnson 
et  al.  (1979)  placed  at  least  three  people  50-100 
m  (165-330  ft)  apart  to  differentiate  among  simulta- 
neously responding  pairs.  At  best,  locations  of  re- 
sponding owls  are  estimates,  and  often  a  reasonable 
estimate  is  extremely  difficult.  This  problem  is  com- 
pounded because  birds  responding  to  calls  may  ap- 
proach the  caller  beyond  the  limits  of  their  normal 
home  range  or  territory  (Reynolds  pers.  commun.). 
This  can  impose  great  limitations  when  assigning 
habitat  types  based  on  call  responses. 


Accuracy  and  Precision.  Although  certain 
raptors  respond  to  calls  year-round  (Foster  1965), 
vocalization  surveys  are  most  reliable  during  the 
breeding  season  (Fuller  and  Mosher  1981).  Not  all 
individuals  will  respond  to  calls  (Nowicki  1974; 
Beatty  1977;  Johnson  et  al.  1979;  Balding  and  Dibble 
1984).  Siminski  (1976),  Forsman  et  al.  (1977),  and 
Springer  ( 1978)  found  that  from  75  to  85%  of  the 
nesting  spotted  owls  and  great  horned  owls  in  their 
study  areas  responded  to  broadcasted  calls.  Male 
owls  appear  to  be  more  responsive  to  broadcasted 
calls  than  females  (Siminski  1976;  Springer  1978). 
Results  on  accuracy  of  this  method  on  diurnal  rap- 
tors are  preliminary  (Mosher  pers.  commun.).  How- 
ever, Mosher  et  al.  (unpubl.  ms.)  found  a  significant 
positive  correlation  between  the  number  of  nesting 


diurnal  pairs  and  the  number  of  responses  to  re- 
corded calls  of  conspecifics.  To  increase  accuracy, 
they  recommended  calling  on  ten  different  occasions 
during  the  nesting  season  and  establishing  more  and 
longer  transects  when  surveying  large  areas. 

Equipment.  Necessary  equipment  consists  of  a 
cassette  tape  player,  amplifier,  speaker,  and  cassette 
tape  recordings  of  calls.  Surveyors  have  used  many 
kinds  of  tape  recorders,  amplifiers,  and  speakers  that 
have  ranged  from  simple  units  to  sophisticated  com- 
ponent systems.  Examples  are  the  Model  600  BM 
Game  Tape  Caller  (Springer  1978),  Sony  Model  TC 
100  Tape  Recorder  (Smith  1978),  and  Marantz 
C-205  Tape  Recorder  with  10-in.  (25-cm),  8-ohm 
trumpet  speakers  (Mosher  et  al.  unpubl  ms.).  Regard- 
less of  equipment,  its  audible  range  should  be  no 
less  than  0.4  km  (1/4  mi.).  Mosher  et  al.  (unpubl. 
ms. )  recommended  a  tape  player  with  a  frequency  of 
about  40  Hz  to  12  kHz  and  a  power  output  of  about 
1.2  watts  at  1  kHz.  Some  owls  have  local  dialects 
which  could  make  them  less  responsive  to  calls  of 
the  same  species  from  a  different  geographic  region 
(Siminski  1976;  Fuller  and  Mosher  1981).  Surveyors 
should  try  to  use  calls  recorded  in  the  geographic 
region  where  the  survey  is  being  conducted. 

Taped  vocalizations  are  available  from  the  fol- 
lowing sources: 


Laboratory  of  Ornithology 
Cornell  University 
Sapsucker  Woods 
Ithaca,  NY  14850 

National  Audubon  Society 
Western  Education  Center 
376  Greenwood  Beach  Road 
Tiburon,  CA  94920 

National  Geographic  L.P.  Records 
National  Geographic  Society 
17th  &  M  Streets  NW 
Washington,  DC  20036 

Oregon  State  Office 

Bureau  of  Land  Management 

P.O.  Box  2965 

Portland,  OR  97208  (for  spotted  owl  tapes) 


Training.  Investigators  must  be  thoroughly 
familiar  with  the  repertoire  of  calls  ( including  any 
differences  between  sexes )  of  the  species  they  are 
surveying.  Training  for  both  new  and  experienced 
surveyors  is  essential.  Listening  to  recorded  calls 
from  the  many  field  guide  series  is  the  simplest  rep- 
ertoire of  the  species  (Johnson  et  al.  1979).  It  may 
be  necessary  to  obtain  taped  calls  from  an  expert  on 
a  particular  species. 


336 


Raptors 


Cost.  Vocalization  surveys  are  relatively  inex- 
pensive to  conduct,  and  fairly  large  areas  can  be 
surveyed  in  a  short  period  of  time.  Forsman  et  al. 
(1977)  surveyed  an  average  of  12.8  km  (8.0  mi.)  of 
road  per  night  using  the  continuous  transect 
method.  Ellison  (1980)  averaged  12  calling  stations 
per  night.  Using  the  procedure  recommended  by 
Mosher  et  al.  (unpubl.  ms.),  it  is  possible  for  one 
person  to  complete  16  calling  stations  or  cover  12.8 
km  (8.0  mi.)  of  transect  in  4  hours  if  all  stations 
are  easily  accessible. 

Discussion.  Vocalization  surveys  result  in  sig- 
nificantly more  bird  contacts  than  simply  looking 
and  listening  for  inconspicuous  raptors.  They  greatly 
reduce  the  time  required  to  find  birds  and  may  be 
the  only  reliable  means  to  locate  owls  and  forest- 
inhabiting  hawks.  Although  responsiveness  varies 
with  species,  the  method  is  reliable  for  assessing  spe- 
cies occurrence  (Mosher  pers.  commun.).  Results 
suggest  that  calling  from  point  counts  can  be  useful 
in  developing  indexes  to  local  populations  and  per- 
haps for  estimates  of  density  (Mosher  et  al.  unpubl. 
ms.).  Accuracy,  however,  is  reduced  when  the 
method  is  used  to  count  individuals  in  a  given  area. 
Estimates  of  birds  in  an  area  should  only  be  used  if  it 
is  certain  that  the  broadcast  calls  covered  the  sample 
area.  Otherwise,  a  more  realistic  measure  of  relative 
density  would  be  number  of  birds  per  linear  distance 
(Forsman  et  al.  1977)  or  contacts  per  station 
(Ellison  1980). 

Although  the  method  is  used  to  assess  relative 
abundance  of  nesting  raptors,  it  is  valuable  in  supple- 
menting road  counts  and  other  types  of  nest 


searches  (Fuller  and  Mosher  1981).  Biologists  use 
nocturnal  calling  to  obtain  the  general  location  of 
spotted  owl  territories  and  diurnal  calling  to  specifi- 
cally locate  roosts  and  nests  (Forsman  1983b).  To 
locate  diurnal  raptor  nests,  biologists  concentrate 
their  searches  in  areas  where  most  contacts  occur 
(Mosher  et  al.  unpubl.  ms.)  or  search  in  the  direc- 
tion of  departing  birds  after  a  contact  has  been  made 
(Rosenfield  et  al.  1985).  Whenever  possible,  survey- 
ors should  supplement  call  surveys  with  ground 
searches  for  birds  or  nests.  These  searches  are  essen- 
tial to  confirm  breeding  attempts,  and  they  also  serve 
to  assess  survey  accuracy. 

Investigators  should  be  cautious  when  using 
broadcast  calls  alone  to  assess  relative  densities  be- 
cause several  factors  contribute  to  the  difficulty  in 
estimating  numbers,  and  many  variables  can  influ- 
ence the  reliability  of  a  vocal  census.  Fuller  and 
Mosher  (1981)  discussed  the  main  variables,  includ- 
ing time  of  year,  time  of  day,  behavioral  differences, 
and  differences  in  vocalizations  and  in  responses 
between  sexes.  Both  sexes  of  the  great  horned  owl, 
spotted  owl,  screech-owl,  and  fiammulated  owl  re- 
spond to  calls,  and  there  may  be  sexual  differences 
in  call  responses  (Marshall  1939;  Siminski  1976; 
Forsman  et  al.  1977;  Springer  1978;  Smith  1978; 
Johnson  et  al.  1979).  They  demonstrate  sexual  differ- 
ences in  the  pitch  of  their  calls;  the  male  call  is  soft- 
er and  lower.  Also,  one  cannot  assume  if  a  male 
responds  to  a  call  that  a  pair  is  occupying  a  territory, 
because  many  responding  males  may  not  be  paired, 
and  unpaired  male  spotted  owls  and  fiammulated 
owls  appear  more  responsive  during  the  nesting 
season  (Reynolds  pers.  commun.). 


Spotted  owl. 


Great  horned  owl. 


Raptors 


337 


Equipment  and  environmental  variables  can  also 
affect  response  rates  (Fuller  and  Mosher  1981). 
Background  noise  produced  by  the  equipment  or  the 
environment  (e.g.,  traffic  noise)  may  interfere  with 
efficient  transmission  and  should  be  avoided  if  possi- 
ble. Effects  of  wind  on  sound  transmission  and  the 
observers  ability  to  hear  the  calls  vary  with  the  habi- 
tat type  but  have  less  of  an  influence  in  open  habi- 
tats. Steep  terrain,  thick  vegetation,  and  running 
water  can  greatly  restrict  the  audible  distance  of  the 
call  or  the  observer's  ability  to  hear  a  response.  Re- 
sponse rates  may  be  affected  by  the  lunar  cycle,  and 
owls  may  be  more  responsive  during  a  bright,  wax- 
ing moon  (Johnson  et  al.  1979).  Ellison  (1980)  ad- 
justed his  systematic  calling  sequence  monthly  to 
avoid  synchronizing  calling  station  visits  with  the 
moon  phase. 

Surveyors  need  to  understand  ranges  and  disper- 
sion of  target  species  pairs.  If  call  stops  are  too  far 
apart,  some  birds  may  be  missed,  and  if  too  close, 
the  same  pair  may  be  counted  twice.  Mosher  et  al. 
(unpubl.  ms.)  recommended  stations  spaced  at  0.8 
km  (0.48  mi.)  for  larger  raptors;  however,  points  can 
be  only  0.25  km  (790  ft)  for  smaller  raptors  such  as 
screech-owls.  It  is  worthwhile  to  test  the  audio 
range  of  equipment  on  raptors  in  a  known  location 
in  the  vegetational  and  topographic  areas  to  be 
sampled. 


Certain  factors  should  be  considered  when 
broadcasting  calls.  Johnson  et  al.  ( 1981 )  believed 
that  if  western  screech-owls  were  censused  too  often 
they  could  become  habituated  to  broadcasts  and  be 
less  responsive;  however,  Mosher  et  al.  (unpubl.  ms.) 
saw  no  indication  of  this  with  diurnal  raptors.  When 
calling  different  species  from  the  same  station,  Call 
(1978)  recommended  that  calls  of  the  smallest  spe- 
cies be  played  first  because  response  behavior  of  the 
smaller  species  may  be  inhibited  by  the  vocalization 
of  a  larger  owl.  However,  some  researchers  do  not 
recommend  calling  different  species  from  the  same 
station  because  of  the  possible  biases  (Mosher  pers. 
commun.). 

Ground  Surveys. 

I  include  searches  by  foot,  horseback,  boat,  and 
land  vehicle  (e.g.,  automobile,  motorcycle,  bicycle) 
in  this  category.  The  methods  involve  similar  tech- 
niques, allow  time  for  close  inspection,  and  are  often 
used  to  complement  each  other. 

Description.  Biologists  should  be  concerned 
with  collecting  two  types  of  data  during  ground 
surveys:  the  general  location  of  a  pair  (i.e.,  nesting 
territory)  and  the  exact  nest  location.  The  latter 
requires  more  time  and  effort,  and  slightly  different 


338 


Raptors 


techniques.  In  most  monitoring  programs  it  is  not 
necessary  to  determine  the  exact  nest  location. 
Therefore,  biologists  should  only  be  concerned  with 
obtaining  general  locations  of  pairs  unless  specific 
management  needs  require  the  exact  locations  of 
nests.  Biologists  should  use  multiple  techniques;  one 
negative  observation  does  not  necessarily  mean  that 
a  site  is  vacant.  Subsequent  visits  and  checks  for 
signs  of  occupancy  should  be  made  before  conclud- 
ing vacancy. 


General  Considerations)  or  defending  adults  is  re- 
quired to  determine  the  exact  nest  location.  To  lo- 
cate scrapes,  some  workers  have  flushed  falcons  by 
clapping  hands  or  throwing  rocks.  These  techniques 
should  be  used  with  caution  because  they  disturb 
birds,  and  they  may  not  be  totally  reliable  because 
some  birds  do  not  flush  when  disturbed  (BLM  un- 
publ.  data).  Sometimes  the  only  way  to  locate  nests 
involves  many  hours  of  observing  the  cliff  and 
watching  for  adults  or  young. 


Surveyors  must  be  aware  of  the  potential  haz- 
ards from  disturbance  caused  by  nest  searches.  Pre- 
cautions are  discussed  by  Fyfe  and  Olendorff  ( 1976), 
Call  (1978),  and  Steenhof  and  Kochert  (1982).  Al- 
though investigators  need  to  locate  nests  before 
hatching,  they  must  observe  them  from  a  safe  dis- 
tance and  avoid  entering  nests  and  disturbing  parent 
birds  until  the  young  can  thermoregulate  (Steenhof 
and  Kochert  1982). 

Craighead  and  Craighead  (1956),  Call  (1978), 
Jones  (1979),  Reynolds  (1982),  and  Forsman 
(1983b)  described  survey  techniques  for  forest- 
nesting  raptors.  To  determine  general  nest  locations 
in  deciduous  forests,  biologists  often  canvass  wooded 
areas  before  foliation  in  the  spring  and  look  for  stick 
nests.  These  searches  are  often  supplemented  by 
aerial  surveys  (Petersen  1979).  Observations  of  aer- 
ial courtship  displays  of  raptors  above  the  canopy 
yield  the  general  nest  location  for  signs  of  occu- 
pancy (see  General  Considerations)  and  defending 
adults  are  required  to  locate  the  nest.  Workers  some- 
times locate  pairs  and  follow  them  to  nests  (Village 
1984).  Searches  for  forest-nesting  raptors,  especially 
cavity  nesting  owls,  are  facilitated  by  calling  surveys 
(Forsman  1983b;  Fuller  and  Mosher  1986).  Occu- 
pancy of  suspected  cavities  by  smaller  raptors  may 
be  confirmed  by  rapping  the  tree  and  observing  the 
bird  at  the  cavity  entrance  (Call  1978);  however, 
this  does  not  necessarily  mean  there  is  a  nest  in  it. 
Also,  this  method  may  not  be  reliable  because  cer- 
tain species,  such  as  flammulated  owls,  are  seldom 
brought  to  the  cavity  entrance  by  rapping  (Reynolds 
pers.  commun. ).  Exact  nest  locations  of  spotted  owls 
have  been  determined  by  baiting  the  birds  with  a 
tethered  mouse  and  following  the  owl  with  prey 
back  to  its  nest  (Forsman  1983b). 

Call  (1978)  and  Mindell  (1983)  discussed  sur- 
vey techniques  for  cliff-nesting  raptors;  decorated 
stick  nests,  whitewash  on  the  cliff,  and  behavior  of 
adult  birds  are  the  most  commonly  used  clues.  Cliffs 
should  be  surveyed  early  in  the  nesting  season  to 
obtain  general  locations  of  breeding  pairs  and  to 
locate  stick  nests.  These  searches  are  often  later  aug- 
mented by  aerial  surveys.  If  the  nest  or  scrape  is 
not  located  during  the  early  survey,  a  thorough 
search  of  the  general  location  for  field  signs  (see 


Searches  for  ground  nesters  sometimes  require 
unique  techniques.  General  nesting  locations  of  har- 
riers and  other  ground-nesting  raptors  are  obtained 
by  observing  consistent  nesting  behavior  (i.e.,  court- 
ship flights,  territorial  defense,  copulation,  prey  ex- 
changes, and  nest  building)  in  an  area;  however, 
behavior  must  be  interpreted  with  caution  because 
unmated  harriers  sometimes  exhibit  courtship  behav- 
ior (Call  1978;  Thompson-Hanson  1984).  Exact  nest 
locations  can  be  determined  by  observing  the  female 
flying  consistently  to  the  same  spot  with  prey  during 
the  brood-rearing  period  or  by  systematically  search- 
ing a  general  nesting  location,  sometimes  with  a 
trained  dog  (Call  1978;  Thompson-Hanson  1984). 
Burrowing  owl  nests  are  located  by  observing 
perched  owls,  prey  remains,  and  perches  in  areas 
with  burrows  (Call  1978). 


Accuracy  and  Precision.  Little  published  in- 
formation exists  on  accuracy  and  precision  of 
ground  surveys  for  nesting  raptors;  however,  accu- 
racy varies  with  species,  habitat,  and  methods.  In 
most  instances,  foot  surveys  are  most  accurate  be- 
cause observers  are  able  to  see  inconspicuous  nests 
(Fuller  and  Mosher  1981),  and  accuracy  increases 
with  each  subsequent  survey  (Craighead  and  Craig- 
head 1956;  Postovit  1979).  As  with  the  other  meth- 
ods, surveys  are  most  accurate  after  completion  of 
all  clutches,  but  before  the  first  brood  hatches.  Vil- 
lage (1984)  showed  a  20-30%  decrease  in  accuracy 
when  European  kestrel  (Falco  tinnunculns)  nests 
were  visited  on  foot  late  in  the  season  because  he 
missed  nonbreeding  pairs  and  those  which  failed 
early.  Craighead  and  Craighead  (1956)  used  both 
foot  and  vehicle  surveys  of  a  deciduous  forest  area; 
they  believed  that  85%  of  all  raptor  nests  were  eas- 
ily found  and  the  remaining  1 5%  required  numerous 
repeated  searches.  Olendorff  (1975)  believed  that 
by  using  repeated  vehicle  and  foot  surveys  he  lo- 
cated over  95%  of  the  raptor  pairs  nesting  on  trees 
and  cliffs  in  a  grassland  area.  Large  conspicuous  nests 
are  easily  found;  Hodges  and  Robards  (1982)  be- 
lieved that  a  trained  observer  could  locate  90%  of 
the  bald  eagle  nests  within  200  m  (660  ft)  of  the 
shoreline  in  a  single  pass  in  a  boat  and  accurately  as- 
sess nesting  status  of  90%  of  the  nests  as  well.  Accu- 
racy for  boat  surveys  of  cliff-nesting  falcons  would 
probably  be  much  less. 


Raptors 


339 


Village  (1984)  presented  two  sources  of  error 
in  assessing  nesting  density:  (1)  study  areas  are  nqt 
chosen  at  random,  and  (2)  nests  may  be  more  easily 
missed  in  large  study  areas  than  small  ones.  This 
problem  is  compounded  because  clusters  of  sight- 
ings, normally  recorded  as  one  pair,  may  be  two 
pairs.  Biases  can  be  so  large  that  they  may  have 
masked  any  variation  in  breeding  density. 

Precision  of  ground  surveys  can  vary  greatly 
depending  on  size  of  the  sampling  unit,  sampling  in- 
tensity, and  number  of  species  surveyed.  Postovit 
(1979)  refined  precision  of  raptor  surveys  of  a  223 
km    (93  mi.  )  area  2.6  times  (from  63  to  24%  based 
on  95%  confidence  intervals)  by  reducing  sample 
quadrat  size  from  2.56  km    to  0.65  km    (1.02  mi.    to 
0.26  mi.  )  and  increasing  sampling  intensity  from 
10  to  33%  of  the  study  area.  However,  surveys  still 
lacked  sensitivity,  and  population  changes  less  than 
39%  could  not  be  detected.  Because  precision  may 
be  increased  by  sampling  a  large  proportion  of  an 
area  in  a  short  period  using  a  correction  factor  for 
visibility  bias  (Postovit  1979),  other  more  intensive 
surveys  may  be  more  precise.  Stratified  random  sam- 
pling apparently  does  not  increase  precision  when 
surveying  a  collection  of  raptor  species  (Postovit 
1979),  but  it  increases  precision  when  sampling  for 
a  single  species  that  occurs  in  restricted  habitats 
(Leighton  et  al.  1979;  Grier  et  al.  1981 ). 

Equipment.  Binoculars  (at  least  7x35)  and  a 
spotting  scope  (minimum  20x)  with  a  sun  shade  are 
essential. 

Training.  Observer  competence  is  one  of  the 
many  factors  that  influence  accuracy  of  nesting  sur- 
veys. Survey  manuals  (e.g.,  Call  1978)  and  short 
courses  are  helpful;  however,  novice  biologists 
should  survey  with  an  experienced  biologist  until 
they  have  developed  a  satisfactory  "search  image." 

Cost.  Cost  will  vary  depending  on  species  and 
accessibility,  and  ruggedness  and  remoteness  of  the 
terrain.  Village  (1984)  believed  that  one  full-time 
worker  could  properly  survey  about  100  km2  (39 
mi.  )  of  reasonably  accessible  habitat  for  European 
kestrels  during  a  nesting  season.  When  travel  costs 
and  salaries  are  considered,  ground  surveys  cost 
nearly  the  same  as  equivalent  helicopter  surveys  and 
more  than  fixed-wing  surveys. 


be  surveyed  quickly  by  automobile,  and  boats  pro- 
vide ready  access  to  remote  areas.  However,  these 
surveys  have  some  error,  and  questionable  nesting 
areas  should  be  reinspected  on  foot. 

Although  ground  surveys  may  be  the  most  accu- 
rate survey  method  for  nesting  raptors,  they  have 
certain  limitations  and  biases  (Postovit  1979;  Steen- 
hof  and  Kochert  1982;  Steenhof  1986;  Fuller  and 
Mosher  1986).  Observer  and  visibility  biases  are 
most  significant,  and  because  of  possible  poor  preci- 
sion, especially  with  sampling  of  large  areas,  there 
may  be  problems  using  single  surveys  to  monitor 
nesting  populations.  Biases  can  be  reduced  and  pre- 
cision increased  by  repeated  surveys  of  sample  areas, 
increased  sample  size,  and  by  using  only  experi- 
enced surveyors.  However,  ground  surveys  are  time- 
consuming,  and  time  is  one  of  the  main  limitations 
of  the  technique.  This  problem  can  be  reduced  by 
augmenting  ground  surveys  with  more  expedient 
methods  (e.g.,  aerial  surveys). 

Productivity  Surveys 

Data  Collection  and  Analysis. 


General  Considerations.  Data  on  raptor 
reproductive  rates  are  useful  to  managers  for 
assessing  the  status  and  reproductive  health  of  a 
raptor  population.  Information  required,  biases,  and 
timing  of  productivity  surveys  are  briefly  discussed 
in  this  section.  Postupalsky  (1974),  Steenhof  and 
Kochert  (1982),  Fraser  et  al.  (1983),  and  Steenhof 
(1986)  discussed  them  in  more  detail. 

Productivity  Data.  Clutch  size  and  number  of 
young  hatched  usually  are  not  necessary  for  produc- 
tivity surveys;  however,  the  following  basic  informa- 
tion is  used  for  estimates  of  reproductive  success 
and  production  (Steenhof  1986): 


( 1 )   Number  of  pairs  in  the  area  This  refers  to 
the  number  of  occupied  "breeding  areas"  or 
"nesting  territories"  (Postupalsky  1974; 
Steenhof  1986).  Occupancy  classification 
requires  at  least  one  of  the  following  observa- 
tions during  the  nesting  season: 


Discussion.  Ground  surveys  are  most  effective 
for  inconspicuous  secretive  raptors.  This  method, 
especially  on  foot,  is  useful  for  intensive  searches  of 
small  areas  of  specialized  habitat;  however,  these 
searches  are  the  most  time-consuming  of  all.  Surveys 
on  foot  or  horseback  may  be  more  accurate  than 
those  by  boat  or  automobile  because  the  former  al- 
low for  closer  inspection.  These  techniques,  how- 
ever, should  complement  each  other.  Large  areas  can 


(a)   Evidence  that  an  egg  was  laid  (i.e.,  obser- 
vation of  an  incubating  bird,  eggs,  egg- 
shell fragments,  young,  or  a  decorated 
nest  with  fresh  prey  and  feces). 


(  b  )  Observation  of  two  breeding- age  birds 
that  appear  to  be  paired  within  a  nesting 
territory. 


340 


Raptors 


(c)  Observations  of  one  or  more  birds  at- 
tending a  nest,  engaging  in  reproductive 
behavior,  or  defending  an  area. 

Some  researchers  disagree  with  the  last  crite- 
rion, but  observation  of  one  adult  attending  a  nest  or 
defending  an  area  usually  constitutes  enough  evi- 
dence that  a  pair  is  present  (Steenhof  1986).  Also 
the  presence  of  a  freshly  decorated  nest  usually  is 
evidence  of  occupancy  for  species  such  as  golden  ea- 
gles and  red-tailed  hawks.  Caution  must  be  used, 
however,  because  of  the  difficulty  in  distinguishing 
new  nesting  material  from  old,  and  because  winter- 
ing birds  in  some  areas  decorate  nests  and  then 
move  elsewhere  (Steenhof  1986).  Areas  where  only 
decorated  nests  are  observed  should  be  revisited 
on  the  ground  to  confirm  occupancy. 


(2)   Number  of  breeding  pairs.  This  is  the  num- 
ber of  pairs  that  lay  eggs  (see  1  above);  how- 
ever, this  information  is  not  mandatory. 
Useful  information  can  be  obtained  without 
knowing  the  number  of  breeding  pairs 
(Steenhof  and  Kochert  1982).  Biologists 
should  remember  that  observations  of  raptors 
near  a  nest  are  not  always  evidence  of  a 
breeding  attempt  (Fuller  and  Mosher  1981); 
nonbreeding  pairs  sometimes  decorate  or 
repair  nests.  Some  birds  may  also  assume  in- 
cubation posture  without  having  laid  an  egg; 
but  because  of  the  difficulty  of  distinguishing 
these  situations,  birds  observed  in  incubation 
position  should  be  considered  "breeding" 
for  analysis  (Steenhof  1986). 


(3)   Number  of  successful  pairs.  These  are  pairs 
that  raise  at  least  one  young  to  "acceptable" 
fledging  age  which  is  defined  as  80%  of  the 
average  age  that  most  young  leave  the  nest 
(Steenhof  and  Kochert  1982;  Steenhof  1986). 
Full  counts  of  broods  at  fledging  are  not 
necessary. 


(4)   Number  of  young  to  reach  acceptable  fledg- 
ing age.  The  number  of  young  fledged  per 
pair  is  the  most  important  measure  of  pro- 
ductivity (Steenhof  1986).  This  value  can  be 
calculated  directly  from  all  known  pairs  or 
a  random  sample  of  preselected  pairs  in  an 
area.  It  can  also  be  estimated  indirectly  by 
combining  independent  estimates  of  percent- 
age breeding,  percentage  success,  and  num- 
ber fledged  per  successful  nest  (Steenhof  and 
Kochert  1982;  Steenhof  1986). 


Red-tailed  hawk  on  nest. 


Biases.  As  with  other  surveys,  two  broad  types 
of  error  can  bias  productivity  surveys:  measurement 
error  and  sampling  error  (Steenhof  1986).  Measure- 
ment error  occurs  when  data  are  incorrectly  col- 
lected or  interpreted  because  of  the  technique, 
visibility  problems,  or  observer  incompetence. 
Proper  timing  of  surveys  can  reduce  measurement 
error  (Fraser  et  al.  1983).  Also,  by  calculating  a 
measurement  error  rate  and  an  estimated  standard 
error,  true  differences  in  productivity  can  be  tested 
where  all  breeding  pairs  are  known  and  are  repeat- 
edly sampled  in  a  given  year  (Fraser  et  al.  1984). 

Sampling  error  occurs  when  the  pairs  sampled 
do  not  represent  the  population.  Biases  can  be  re- 
duced by  randomly  selecting  pairs  (either  preselect- 
ing known  pairs  or  sampling  areas),  using  only  pairs 
found  during  or  before  incubation,  or  using  the  May- 
field  Model  (Mayfield  1961)  to  calculate  success 
(Steenhof  and  Kochert  1982).  An  unbiased  estimate 
of  percentage  of  pairs  breeding  can  be  obtained  only 
from  samples  of  traditional  pairs  selected  before  the 
nesting  season  (Steenhof  and  Kochert  1982).  Only 
these  nests  and  those  found  during  incubation 
should  be  used  for  calculating  reproductive  success 
and  productivity;  inclusion  of  pairs  found  late  yields 
inflated  estimates.  The  Mayfield  Model  allows  work- 
ers to  maximize  use  of  data  sets  with  small  sample 
sizes  by  including  nests  found  late  in  the  season 
(Steenhof  1986).  A  computer  program  is  available  to 
handle  the  tedious  calculations  required  by  the  May- 
field  approach  (contact  the  Birds  of  Prey  Research 
Staff,  Boise  District,  BLM  for  details). 


Raptors 


341 


Timing  of  Surveys.  For  single  species  surveys, 
a  minimum  of  two  surveys  are  required  (Fraser  et 
al.  1983),  and  more  may  be  necessary  (Steenhof 
1986).  The  first  survey,  to  count  occupied  territories 
and  pairs  with  eggs,  should  be  done  during  incuba- 
tion, ideally  after  completion  of  all  clutches  but  be- 
fore hatching  of  the  first  brood.  A  middle  survey  may 
be  necessary  if  there  is  a  wide  latitude  in  nesting 
chronology.  It  should  be  conducted  after  the  oldest 
broods  reach  acceptable  fledging  age  but  have  not 
left  the  nest.  The  function  of  this  survey  is  to  assess 
fledging  of  the  oldest  broods  and  to  age  the  remain- 
ing young  to  establish  the  optimal  time  of  the  third 
survey.  The  last  or  "productivity"  survey  should 
be  conducted  after  the  youngest  brood  reaches  ac- 
ceptable fledging  age  but  before  the  oldest  chicks, 
who  were  less  than  acceptable  fledging  age  on  the 
middle  survey,  leave  the  nest.  For  multiple-species 
surveys  in  the  Birds  of  Prey  Area,  investigators  found 
it  necessary  to  conduct  a  minimum  of  four  surveys. 
Because  of  the  wide  range  in  chronology,  a  second 
occupancy  survey  is  necessary  to  compensate  for  the 
wide  latitude  in  starting  dates  (BLM  unpubl.  data). 


Survey  Methods 
Aerial  Surveys. 


Description.  Techniques  for  assessing 
productivity  from  aircraft  are  the  same  as  those  used 
during  aerial  nesting  surveys;  however,  the  former 
requires  a  closer  inspection  of  the  nest  contents. 

Accuracy  and  Precision.  Aerial  productivity 
surveys  are  fairly  accurate  for  large  conspicuous 
raptors.  An  average  82%  of  the  breeding  pairs  were 
correctly  classified  and  86%  of  the  young  were  seen 
during  fixed-wing  surveys  of  nesting  bald  eagles 
(Fraser  et  al.  1983). 

In  the  Birds  of  Prey  Area,  accuracy  of  helicopter 
surveys  for  cliff  nesting  golden  eagles  was  86%  for 
assessing  breeding  status  and  95%  for  counting 
young  (BLM  unpubl.  data).  Accuracy  can  be  en- 
hanced by  adjusting  survey  timing.  By  properly  tim- 
ing surveys,  Fraser  et  al.  (1983)  increased  accuracy 
for  both  estimates  by  about  12%.  Accuracy  in  the 
Birds  of  Prey  Area  was  also  increased  with  properly 
timed  helicopter  surveys. 

Poole  ( 1981 )  found  fixed-wing  surveys  of  os- 
prey  breeding  pairs  and  fledged  young  were  as  accu- 
rate as  ground  counts.  However,  accuracy  may  be 
lower  for  smaller,  less  conspicuous  raptors.  During 
helicopter  surveys  in  the  Birds  of  Prey  Area,  investi- 
gators have  been  unable  to  obtain  accurate  brood 
counts  at  some  red-tailed  and  ferruginous  hawk  nests 
because  the  parents  covered  their  young  (BLM  un- 
publ. data). 


Discussion.  Aircraft  provide  a  fast  and  expedi- 
tious way  to  collect  productivity  data  for  large  con- 
spicuous raptors,  especially  if  surveys  are  optimally 
timed.  Multiple  productivity  surveys  are  often  re- 
quired and  additional  flights  further  increase  accu- 
racy (Grier  et  al.  1981 ).  Aerial  surveys  can  be 
reliable  for  monitoring  eagle  productivity,  especially 
if  all  breeding  sites  are  known  and  surveyed  at  the 
proper  time  (Fraser  et  al.  1984).  Lastly,  helicopters 
cost  more  than  fixed-wing  aircraft,  but  they  are  usu- 
ally more  accurate  because  they  can  hover  to  allow 
more  opportunity  to  correctly  count  broods. 

Ground  Surveys. 


Description.  Techniques  for  assessing 
productivity  from  ground  surveys  are  the  same  as 
those  used  for  nesting  surveys  except  exact  location 
of  the  nest  and  inspection  of  its  contents  are 
required.  Inspections  can  be  accomplished  by  distant 
observations  with  the  aid  of  telescopes  or  binoculars; 
close  observations  from  a  nearby  vantage  point,  or 
with  the  aid  of  a  mirror  attached  to  a  pole  (Parker 
1972);  or  by  entering  nests  (Call  1978;  Steenhof 
1986). 

Discussion.  The  usefulness  of  methods  for 
inspecting  nests  depends  on  survey  purpose  and  en- 
vironmental conditions  (Steenhof  1986).  Distant 
observations  are  necessary  and  adequate  to  confirm 
an  incubating  bird,  but  less  useful  for  brood  counts, 
especially  if  the  entire  nest  cannot  be  seen.  Climbing 
into  the  nest  is  undoubtedly  the  most  accurate 
method  to  obtain  productivity  data.  However,  it  is 
extremely  time-consuming  and  causes  disturbance  to 
the  birds.  Close  inspections  and  distant  observations 
require  less  time  but  are  subject  to  error.  Foliage 
and  topographic  features  often  do  not  allow  full  view 
of  the  nest  or  complete  brood  counts;  Millsap  (pers. 
commun.)  minimized  this  error  and  used  a  15  m 
(49.5  ft)  range  pole  and  mirror  to  count  eggs  and 
young  in  1 50  tree  nests.  He  spent  an  average  of  3 
minutes  at  each  nest  with  very  little  disturbance  and 
high  accuracy. 


SYSTEMS  FOR  CORRELATING  HABITAT 
VARIABLES  WITH  POPULATION 
MEASUREMENTS 

Habitat  models  provide  useful  information  for 
management  biologists.  They  can  be  used  to  predict 
potential  species  occurrence  and  to  evaluate  poten- 
tial habitat  quality.  The  models  are  also  useful  in 
designing  an  inventory.  Biologists  are  cautioned  from 
using  these  models  alone  as  a  population  measure 
or  as  a  management  tool.  They  still  need  to  inven- 
tory the  area  of  concern  by  using  the  techniques 
described  in  this  chapter  to  gather  the  information 
necessary  for  management  decisions. 


342 


Raptors 


Climbing  into  raptor  nests  is  the  most  accurate  method  of  obtaining  productivity  information. 


Few  habitat  models  have  been  published.  I  will 
discuss  those  that  are  available  or  being  developed. 
The  U.S.  Fish  and  Wildlife  Service  (FWS),  Office  of 
Biological  Services'  Habitat  Evaluation  Procedures 
(HEP)  Group  in  Fort  Collins,  Colorado,  has  devel- 
oped Habitat  Suitability  Index  Models  for  the  ferrugi- 
nous hawk,  prairie  falcon,  sharp-shinned  hawk, 
golden  eagle,  northern  spotted  owl,  short-eared  owl, 
burrowing  owl,  and  northern  harrier.  The  only  pub- 
lished model  (Jasikoff  1982)  is  on  ferruginous  hawks; 
others  are  in  draft  form.  Model  narratives  describe 
habitat  relationships  for  the  raptor.  The  FWS's  Divi- 
sion of  Ecological  Services  in  Anchorage,  Alaska,  has 
also  developed  a  draft  habitat  model  for  bald  eagles. 
Ellis  (1982)  developed  a  system  for  evaluating  the 
potential  for  peregrine  falcon  habitat  in  Arizona. 
It  provides  a  good  means  for  predicting  possible  oc- 
currence, and  it  may  be  applied  to  other  habitats 
in  the  arid  West.  Mosher  et  al.  ( 1985)  developed  a 
management  model  to  predict  nesting  habitat  for 
eastern  woodland  hawks.  Lastly,  Stalmaster  (1983) 
developed  a  detailed  energetics  model  for  bald  ea- 
gles wintering  in  the  Pacific  Northwest  that  deter- 
mines eagle  carrying  capacity  in  salmon  spawning 


drainages  and  estimates  use  of  salmon  by  eagles.  This 
model  also  provides  broad  scale  management 
recommendations. 


DISCUSSION 

Raptor  inventory  and  monitoring  programs  are 
frequently  limited  because  raptors  are  often  difficult 
to  detect  and  count.  Raptors  often  occur  at  relatively 
low  densities  in  a  diversity  of  habitats,  and  it  is  ex- 
tremely laborious  to  obtain  complete  counts  of  nest- 
ing pairs.  Estimates  usually  must  be  obtained  through 
sampling,  and  adequate  sampling  is  often  costly  and 
arduous. 

Design  and  implementation  of  a  successful  rap- 
tor inventory  or  monitoring  program  is  affected  by 
sampling,  measurement,  and  interpretation  problems. 
Many  sampling  problems  affect  estimates  of  numbers 
and  population  changes.  Although  reliable  for  assess- 
ing abundance,  the  usefulness  of  many  indirect  tech- 
niques (e.g.,  calling  surveys)  is  limited  if  studies  do 
not  relate  the  sample  to  the  actual  population  size 


Raptors 


343 


(Fuller  and  Mosher  1981).  Often  relative  indexes  are 
of  limited  value  because  surveys  are  not  conducted 
in  a  standardized  manner  or  samples  are  not  ade- 
quately collected  for  statistical  analysis.  However, 
these  problems  can  be  eliminated  by  proper  sam- 
pling procedures.  Real  problems  occur  when  logis- 
tics prevent  implementation  of  a  proper  sampling 
design,  when  sampling  intensity  is  inadequate  to 
allow  the  necessary  comparisons  to  be  made,  or 
when  samples  are  biased  and  do  not  represent  the 
population. 

Accurate  measurement  of  raptor  populations  is 
also  compromised  by  many  biases.  The  greatest  bias 
is  detectability;  not  all  birds  can  be  found  or  seen. 
Next  is  observer  competence.  Every  technique  dis- 
cussed here  is  affected  by  these  two  biases.  Their 
effects  can  be  reduced  by  adequate  sampling  inten- 
sity and  by  using  only  experienced  observers.  Envi- 
ronmental factors  (e.g.,  weather)  and  biological 
variability  (e.g.,  changes  in  food  or  bird  behavior) 
also  cause  error,  and  each  technique  has  its  own 
inherent  measurement  biases. 

Interpretation  problems  usually  stem  from  in- 
complete knowledge  of  how  raptors  interact  with 
their  environment.  For  example,  lack  of  knowledge 
of  the  effects  of  climate  and  food  supply  hampers 


one's  ability  to  distinguish  true  population  changes 
from  regional  shifts  in  distribution.  Interpretation 
errors  also  result  from  a  poor  selection  of  estimators. 
Some  variables  may  remain  the  same  when  the  popu- 
lation actually  changes.  For  example,  production  of 
successful  pairs  often  will  show  no  change  when 
productivity  of  the  population  actually  has  changed 
(BLM  unpubl.  data).  To  conduct  a  meaningful  inven- 
tory and  monitoring  program,  biologists  must  use 
those  features  that  best  reflect  the  population  (Fuller 
and  Mosher  1986;  Steenhof  1986). 

These  problems  can  be  reduced  by  using  effi- 
cient, standardized  techniques  suitable  for  use  with  a 
diversity  of  raptor  species  and  habitats.  Standardiza- 
tion of  estimated  variables,  including  measures  of 
variability  (e.g.,  confidence  limits,  standard  error),  is 
as  important  as  standardization  of  survey  techniques 
(Fraser  1978).  This  allows  valid  comparisons  to  be 
made  among  areas. 

Through  an  understanding  of  the  inherent  biases 
of  the  methods  used  to  measure  raptor  populations, 
surveys  can  be  designed  and  implemented  to  give 
results  of  optimal  accuracy  and  precision.  By  recog- 
nizing the  limitations  of  the  resultant  data  and  the 
necessity  for  careful  interpretation,  it  is  possible  to 
reach  conclusions  that  are  useful  and  effective  inputs 
to  management  decisions. 


344  Raptors 


LITERATURE  CITED 


ALLEN,  H.L.  and  L.S.  YOUNG.  1982.  An  annotated  bibliog- 
raphy of  avian  communal  roosting.  Washington  Dep. 
Game,  Olympia.  177pp. 

AMERICAN  ORNITHOLOGISTS  UNION.  1983-  Check  list 
of  North  American  birds,  6th  ed.  Allen  Press. 
Lawrence,  KS.  877pp. 

ANTHONY,  R.G.,  R.L.  KNIGHT,  G.T.  ALLEN,  BR.  MCCLEL- 
LAND, and  J.I.  HODGES.  1982.  Habitat  use  by  nesting 
and  roosting  bald  eagles  in  the  Pacific  Northwest. 
Trans.  North  Am.  Wildl.  Nat.  Resour.  Conf.  47:332- 
342. 

APFELBAUM,  S.I.  and  P.  SEELBACH.  1983.  Nest  tree,  habi- 
tat selection  and  productivity  of  seven  North  Ameri- 
can raptors  based  on  the  Cornell  University  nest 
record  card  program.  Raptor  Res.  17:97-113- 

BALDING,  T.  and  E.  DIBBLE.  1984.  Responses  of  red- 
tailed,  red-shouldered,  and  broad-winged  hawks  to 
high  volume  playback  recordings.  Passenger  Pigeon 
46:71-75. 

BAMMANN,  A.R.  1975.  Ecology  of  predation  and  social 
interactions  of  wintering  white-tailed  kites.  M.S.  The- 
sis, Humboldt  State  Univ.  Areata,  CA.  81pp. 

BEATTY,  W.H.  1977.  Attracting  screech  owls.  Redstart 
44:102-104. 

BECHARD,  M.J.  1982.  Effect  of  vegetative  cover  on  forag- 
ing site  selection  by  Swainson's  hawk.  Condor  84:153- 
159. 

BECKER,  DM.  1984.  Reproductive  ecology  and  habitat 
utilization  of  Richardson's  merlins  in  southeastern 
Montana.  M.S.  Thesis,  Univ.  Montana,  Missoula.  62pp. 

and  I.J.  BALL.  1983a.  Prairie  falcon  {Falco  tnexi- 

canus).  Pages  138-153  in  Armbruster,  J.S.  ed.  Impacts 
of  Coal  Surface  Mining  on  25  Migratory  Bird  Species 
of  High  Federal  Interest.  FWS/OBS-83/85.  U.S.  Dep. 
Inter.,  Fish  and  Wildl.  Serv.  Fort  Collins,  CO. 

and  I.J.  BALL.  1983b.  Merlin  (Falco  columbarius). 


Pages  124-137  in  Armbruster,  J.S.  ed.  Impacts  of 
Coal  Surface  Mining  on  25  Migratory  Bird  Species  of 
High  Federal  Interest.  FWS/OBS-83/85.  U.S.  Dep.  In- 
ter., Fish  and  Wildl.  Serv.,  Fort  Collins,  CO. 

BEDNARZ,  J.C.  and  J.J.  DINSMORE.  1982.  Nest-sites  and 
habitat  of  red-shouldered  hawks  and  red-tailed  hawks 
in  Iowa.  Wilson  Bull.  94:31-45. 

BELL,  R.E.  1964.  A  sound  triangulation  method  for  count- 
ing barred  owls.  Wilson  Bull.  76:292-294. 

BILDSTEIN,  KL.  1978.  Behavioral  ecology  of  red-tailed 
hawks  (Buteo  jamaicensis),  rough-legged  hawks  (B. 
lagopus),  northern  harriers  (Circus  cyaneus),  Ameri- 
can kestrels  (Falco  sparverius )  and  other  raptorial 
birds  wintering  in  south  central  Ohio.  Ph.D.  disserta- 
tion. Ohio  State  Univ.,  Columbus.  364pp. 

1979.  Fluctuations  in  the  number  of  northern 

harriers  (Circus  cyaneus  hudsonius)  at  communal 
roosts  in  south  central  Ohio.  Raptor  Res.  1 3:40-46. 

BOEKER,  EL.  1970.  Use  of  aircraft  to  determine  golden 
eagle  (Aquila  chrysaetos)  nesting  activity.  Southwest. 
Nat.  15:136-137.  " 

1974.  Status  of  golden  eagle  surveys  in  the  west- 
ern states.  Wildl.  Soc.  Bull.  2:46-49- 

and  E.B.  BOLEN.  1972.  Winter  golden  eagle  popu- 
lations in  the  Southwest.  J.  Wildl.  Manage.  36:477- 
484. 


BOXALL,  PC.  and  P.H.R.  STEPNEY.  1982.  The  distribution 
and  status  of  the  barred  owl  in  Alberta,  Can.  Field- 
Nat.  96:46-50. 

BROWN,  D.,  C.  LOWE,  and  C.  PASE.  1980.  Digitized  sys- 
tematic classification  for  ecosystems  with  an  illus- 
trated summary  of  the  natural  vegetation  of  North 
America.  Gen.  Tech.  Rep.  RM-73-  U.S.  Dep.  Agric,  For. 
Serv.,  Fort  Collins,  CO.  93pp. 

BROWN,  L.H.  and  D.  AMADON.  1968.  Eagles,  hawks  and 
falcons  of  the  world.  McGraw-Hill.  New  York,  NY. 
945pp. 

BULL,  EL.  and  R.G.  ANDERSON.  1978.  Notes  on  flammu- 
lated  owls  in  northeastern  Oregon.  Murrelet  59:26-27. 

BURNHAM,  KP.,  DR.  ANDERSON,  and  J.L  LAAKE.  1980. 
Estimation  of  density  from  line  transect  sampling 
of  biological  populations.  Wildl.  Monogr.72.  202pp. 

CADE,  T.J.  1982.  The  falcons  of  the  world.  Cornell  Univ. 
Press.  Ithaca,  NY.  188pp. 

CALL,  M.W.  1978.  Nesting  habitats  and  surveying  tech- 
niques for  common  western  raptors.  Tech.  Note  316. 
U.S.  Dep.  Inter.,  Bur.  Land  Manage.  Serv.  Cen.  Denver, 
CO.  115pp. 

CARRIER,  WD.  and  WE.  MELQUIST.  1976.  The  use  of  a 
rotor-winged  aircraft  in  conducting  nesting  surveys  of 
ospreys  in  northern  Idaho.  Raptor  Res.  10:77-83- 

CAUGHLEY,  G.  1974.  Bias  in  aerial  survey.  J.  Wildl.  Man- 
age. 38:921-933. 

CINK,  C.L.  1975.  Population  densities  of  screech  owls  in 
northeastern  Kansas.  Kansas  Ornithol.  Soc.  Bull. 
26:13-16. 

CLARK,  R.J.  1975.  A  field  study  of  the  short-eared  owl 
(Asio flammeus pontoppidan)  in  North  America. 
Wildl.  Monogr.  No.  47.  67pp. 

CLARKE,  R.G.  1982.  Nest  site  selection  by  sharp-shinned 
hawks  in  interior  Alaska.  Pages  155-162  in  Ladd,  W.N. 
and  P.F.  Schempf,  eds.  Raptor  Management  Biology 
in  Alaska  and  Western  Canada.  FWS/AK/PROC-82.  U.S. 
Dep.  Inter.,  Fish  and  Wildl.  Serv.  Anchorage,  AK 

COTTRELL,  M.J.  1981.  Resource  partitioning  and  repro- 
ductive success  of  three  species  of  hawks  (Buteo  sp. ) 
in  an  Oregon  prairie.  M.S.  Thesis,  Oregon  State  Univ., 
Corvallis.  72pp. 

CRAIG,  G.R.  1974.  Raptor  populations  and  characteristics 
studies.  Job  Prog.  Rep.  No.  W- 1 24-R- 1 .  Colorado  Div. 
Wildl.,  Denver.  36pp. 

CRAIG,  T.H.  1978.  A  car  survey  of  raptors  in  southeastern 
Idaho  1974-1976.  Raptor  Res.  12:40-45. 

CRAIGHEAD,  J.J.  and  EC.  CRAIGHEAD.  1956.  Hawks, 
owls,  and  wildlife.  Wildl.  Manage.  Inst.  Washington, 
DC.  443pp. 

ELLIS,  D.H.  1982.  The  peregrine  falcon  in  Arizona:  habitat 
utilization  and  management  recommendations.  Rep.  1. 
Inst.  Raptor  Stud.  Oracle,  AZ.  24pp. 

ELLISON,  P.T.  1980.  Habitat  use  by  resident  screech  owls 
(Otus  asio).  M.S.  Thesis,  Univ.  Massachusetts,  Am- 
herst. 86pp. 

EVANS,  D.L.  1983.  Ferruginous  hawk  (Buteo  regalis). 
Pages  109-123  in  Armbruster,  ed.  Impacts  of  Coal 
Surface  Mining  on  25  Migratory  Bird  Species  of  High 
Federal  Interest.  FWS/OBS-83/85.  U.S.  Dep.  Inter.,  Fish 
and  Wildl.  Serv.  Fort  Collins,  CO. 

FITZNER,  R.E.  1980.  Behavioral  ecology  of  the  Swainson's 
hawk  (Buteo  swainsoni)  in  Washington.  PNL-2754 
UC-11.  Battelle  Pac.  Northwest  Lab.  Richland,  WA. 
15pp. 


Raptors 


545 


FORBES,  J.E.  and  D.W.  WARNER.  1974.  Behavior  of  a 
radio-tagged  saw  whet  owl.  Auk  91:783-795. 

FORSMAN,  ED.  1983a.  Spotted  owl  (Strix  occidentalism 
Pages  243-255  in  Armbruster,  J.S.  ed.  Impacts  of  Coal 
Surface  Mining  on  25  Migratory  Bird  Species  of  High 
Federal  Interest.  FWS/OBS-83/85.  U.S.  Dep.  Inter.,  Fish 
and  Wildl.  Serv.  Fort  Collins,  CO. 

.  1983b.  Methods  and  materials  for  locating  and 

studying  spotted  owls.  U.S.  Dep.  Agric,  For.  Serv.  Gen. 
Tech.  Rep.  PNW-162.  8pp. 

,  EC.  MESLOW,  and  M.J.  STRUB.  1977.  Spotted  owl 

abundance  in  young  versus  old-growth  forests,  Ore- 
gon. Wildl.  Soc.  Bull.  5:43-47. 

and  H.M.  WIGHT.  1984.  Distribution  and  biology 


of  the  spotted  owl  in  Oregon.  Wildl.  Monogr.  87. 
64pp. 

FOSTER,  F.  1965.  An  early  reference  of  the  technique  of 
owl  calling.  Auk  82:651-653. 

FRANZREB,  KE.  1981.  The  determination  of  avian  densi- 
ties using  the  variable-strip  and  fixed-width  transect 
survey  methods.  Stud.  Avian  Biol.  6:139-145. 

and  R.D.  OHMART.  1978.  The  effects  of  timber 

harvesting  on  breeding  birds  in  a  mixed  coniferous 
forest  type.  Condor  80:431-441. 

FRASER,  J.D.  1978.  Bald  eagle  reproductive  surveys:  accu- 
racy, precision,  and  timing.  M.S.  Thesis.  Univ.  Minne- 
sota, Minneapolis.  82pp. 

,  L.D.  FRENZEL,  J.E.  MATHISEN,  F.  MARTIN,  and 

M.E.  SHOUGH.  1983.  Scheduling  bald  eagle  reproduc- 
tion surveys.  Wildl.  Soc.  Bull.  11:13-16. 
-,  F  MARTIN,  L.D.  FRENZEL,  and  J.E.  MATHISEN. 


1984.  Accounting  for  measurement  errors  in  bald 
eagle  reproduction  surveys.  J.  Wildl.  Manage.  48:595- 
598. 

FULLER,  M.R.  1979-  Spatiotemporal  ecology  of  four  sym- 
patric  raptor  species.  Ph.D.  dissertation.  Univ.  Minne- 
sota, Minneapolis.  220pp. 

.  1983-  Zone-tailed  hawk  (Buteo  albonotatus). 

Pages  98-108  in  Armbruster,  J.S.  ed.  Impacts  of  Coal 
Surface  Mining  on  25  Migratory  Bird  Species  of  High 
Federal  Interest.  FWS/OBS-83/85.  U.S.  Dep.  Inter.,  Fish 
and  Wildl.  Serv.  Fort  Collins,  CO. 

and  J. A.  MOSHER.  1981.  Methods  of  detecting  and 

counting  raptors:  a  review.  Stud.  Avian  Biol.  6:235- 
246. 

and .  1986.  Raptor  survey  techniques  in 

Millsap,  B.A.  and  K.W.  Kline,  eds.  Raptor  Management 
Techniques  Manual.  Vol.  1.  Natl.  Wildl.  Fed.  Washing- 
ton, DC. 

and  JR.  PARTELOW.  1983.  Cooper's  hawk.  (Accipi- 

ter  cooperii).  Pages  74-97  in  Armbruster,  J.S.  ed. 
Impacts  of  Coal  Surface  Mining  on  25  Migratory  Bird 
Species  of  High  Federal  Interest.  FWS/OBS-83/85. 
U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.,  Fort  Collins, 
CO. 

FYFE,  R.W.  and  R.R.  OLENDORFF.  1976.  Minimizing  the 
dangers  of  nesting  studies  to  raptors  and  other  sensi- 
tive species.  Can.  Wildl.  Serv.  Occas.  Pap.  23.  Can. 
Wildl.  Serv.  Ottawa,  Ont.  17pp. 

GILMER,  D.S.,  L.M.  COWARDIN,  R.L.  DUVAL,  L.M.  MECH 
UN,  C.W.  SHAIFFER,  and  V.B.  KUECHLE.  1981.  Proce- 
dures for  the  use  of  aircraft  in  wildlife  bio-telemetry 
studies.  Resour.  Publ.  140.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.,  Washington,  DC.  19pp. 

and  R.E.  STEWART.  1984.  Swainson's  hawk  nesting 

ecology  in  North  Dakota.  Condor  82:12-18. 


GLINSKI,  R.L.  and  R.D.  OHMART.  1983-  Breeding  ecology 
of  the  Mississippi  kite  in  Arizona.  Auk  85:200-207. 

GRIER,  J.W.  1977.  Quadrat  sampling  of  a  nesting  popula- 
tion of  bald  eagles.  J.  Wildl.  Manage.  41:438-443. 

,  J.M.  GERRARD,  G.D.  HAMILTON,  and  PA.  GRAY. 

1981.  Aerial  visibility  bias  and  survey  techniques 
for  nesting  bald  eagles  in  northwestern  Ontario.  J. 
Wildl.  Manage.  45:83-92. 

and  G.D.  HAMILTON.  1978.  Aerial  census  of  bald 


eagles  in  the  West  Patricia  Planning  Area.  West  Patri- 
cia Land  Use  Plan.  Wildl.  Tech.  Rep.  7.  Ontario  Min. 
Nat.  Resour.,  Ottawa.  25pp. 

GROSSMAN,  ML.  and  J.  HAMLET.  1964.  Birds  of  prey  of 
the  world.  Bonanza  Books,  New  York,  NY.  496pp. 

HANCOCK,  D.  1964.  Bald  eagles  wintering  in  the  south- 
ern Gulf  Islands,  British  Columbia.  Wilson  Bull. 
76:111-120. 

HAYWARD,  G.D.  1983  Resource  partitioning  among  six 
forest  owls  in  the  River  of  No  Return  Wilderness, 
Idaho.  M.S.  Thesis.  Univ.  Idaho,  Moscow.  127pp. 

HEGDAL,  P.L.  and  R.W.  BLASSKIEWICZ.  1984.  Evaluation 
of  the  potential  hazard  to  barn  owls  of  TALON  (bro- 
difacoum  bait)  used  to  control  rats  and  house  mice. 
Environ.  Toxicol.  Chem.  3:167-179. 

HEIN,  E.  1961.  Wood  duck  roosting  flights  at  Paint  Creek, 
Iowa.  Proc.  Iowa  Acad.  Sci.  68:264-270. 

HENNY,  C.J.,  MM.  SMITH,  and  V.D.  STOTTS.  1974.  The 
1973  distribution  and  abundance  of  breeding  ospreys 
in  the  Chesapeake  Bay.  Chesapeake  Sci.  15:125-133. 

,  MA.  BYRD,  J.A.  JACOBS,  P.D.  McLAIN,  MR.  TODD, 

and  B.F.  HALLA.  1977.  Mid-Atlantic  Coast  osprey 
population:  present  numbers,  productivity,  pollutant 
contamination,  and  status.  J.  Wildl.  Manage.  41:254- 
265. 

HICKMAN,  G.L.  1972.  Aerial  determination  of  golden 
eagle  nesting  status.  J.  Wildl.  Manage.  36:1289-1292. 

HODGES,  J.I.,  J.G.  KING,  and  F.C.  ROBARDS.  1979.  Re- 
survey  of  the  bald  eagle  breeding  populations  in 
southeast  Alaska.  J.  Wildl.  Manage.  43:219-221. 

and  F.C.  ROBARDS.  1982.  Observations  of  3850 

bald  eagle  nests  in  southeast  Alaska.  Pages  37-46  in 
Ladd,  W.N.  and  P.F.  Schempf,  eds.  Raptor  Management 
and  Biology  in  Alaska  and  Western  Canada.  U.S.  Dep. 
Inter.,  Fish  and  Wildl.  Serv.  FWS/AK/Proc.  82.  Anchor- 
age, AK 

HOSKINSON,  R.L.  1976.  The  effect  of  different  pilots  on 
aerial  telemetry  error.  J.  Wildl.  Manage.  40:137-139. 

JASIKOFF,  T.M.  1982.  Habitat  suitability  index  models: 
ferruginous  hawk.  U.S.  Dep.  Inter.,  Fish  and  Wildl. 
Serv.  FWS/BS-82/10.10.  Washington,  DC.  18pp. 

JOHNSON,  DR.  and  WE.  MELQUIST.  1973.  Unique,  rare, 
and  endangered  raptorial  birds  of  northern  Idaho: 
nesting  success  and  management  recommendations. 
Publ.  No.  Rl -73-021.  Univ.  Idaho  and  U.S.  Dep.  Agric, 
For.  Serv.  Moscow,  ID.  42pp. 

JOHNSON,  R.R.,  B.T.  BROWN,  L.T.  HAIGHT,  and  J.M. 

SIMPSON.  1981.  Playback  recording  as  a  special  avian 
censusing  technique.  Stud.  Avian  Biol.  6:68-75. 

,  L.T.  HAIGHT,  and  J.M.  SIMPSON.  1979.  Owl  popu- 
lations and  species  status  in  the  southwestern  United 
States.  Pages  40-59  in  Schaeffer,  P.P.  and  S.M.  Ehlers, 
eds.  Owls  of  the  West:  Their  Ecology  and  Conserva- 
tion. Natl.  Audubon  Soc.  Tiburon,  CA. 

JONES,  S.  1979.  Habitat  management  series  for  unique  or 
endangered  species.  Report  No.  17  -  The  accipiters 
goshawk,  Cooper's  hawk,  sharp-shinned  hawk.  Tech. 


346 


Raptors 


Note  335.  U.S.  Dep.  Inter.,  Bur.  Land  Manage.  Denver, 
CO.  51pp. 

KEISTER,  G.P.,  Jr.  1981.  An  assessment  of  bald  eagle  com- 
munal roosting  in  northwestern  Washington.  Wash. 
Dep.  Game,  Olympia.  39pp. 

and  R.G.  ANTHONY.  1983-  Characteristics  of  bald 

eagle  communal  roosts  in  the  Klamath  Basin,  Oregon 
and  California.  J.  Wildl.  Manage.  47:1072-1079. 

KING,  J.,  F.C  ROBARDS,  and  C.J.  LENSINK  1972.  Census 
of  the  bald  eagle  breeding  population  in  southeastern 
Alaska.  J.  Wildl.  Manage.  36:1292-1295. 

KOCHERT,  M.N.  1980.  Golden  eagle  reproduction  and 
population  changes  in  relation  to  jackrabbit  cycles: 
implications  to  eagle  electrocutions.  Pages  71-86 
in  Howard,  R.P.  and  J.F.  Gores,  eds.  A  Workshop  on 
Raptors  and  Energy  Development.  Idaho  Chapter 
Wildl.  Soc,  Boise. 

,  A.R.  BAMMANN,  R.P.  HOWARD,  J.H.  DOREMUS,  M. 

DELATE,  and  D.  DONAHUE.  1975.  Reproductive 
performance,  food  habits  and  population  dynamics  of 
raptors  in  the  Snake  River  Birds  of  Prey  Natural  Area. 
Pages  1-50  in  Snake  River  Birds  of  Prey  Res.  Proj. 
Ann.  Rep.  U.S.  Dep.  Inter.,  Bur.  Land  Manage.  Boise, 
ID. 

,  A.R.  BAMMANN,  J.H.  DOREMUS,  M.  DELATE,  and  J. 

WYATT.  1976.  Reproductive  performance,  food  habits 
and  population  dynamics  of  raptors  in  the  Snake 
River  Birds  of  Prey  Natural  Area.  Pages  1-57  in  Snake 
River  Birds  of  Prey  Res.  Proj.  Ann.  Rep.  U.S.  Dep. 
Inter.,  Bur.  Land  Manage.,  Boise,  ID. 
-,  J.H.  DOREMUS,  K  STEENHOF,  DR.  DUNCAN, 


M.Q.  MORITSCH,  AC.  DOLDE,  DM.  RAMIREZ,  S.A. 
ADAMS,  and  D.  DELSORDO.  1983.  Density  and  repro- 
ductive performance  of  raptors  in  the  Snake  River 
Birds  of  Prey  Area.  Pages  6-15  in  Snake  River  Birds  of 
Prey  Res.  Proj.  Ann.  Rep,  U.S.  Dep.  Inter.,  Bur.  Land 
Manage.  Boise,  ID. 

KUCHLER,  A.W.  1964.  Potential  natural  vegetation  of  the 
conterminous  United  States.  Am.  Geogr.  Soc.  Spec. 
Publ.  36.  Am.  Geogr.  Soc.  New  York,  NY.  1 16pp. 

LEIGHTON,  FA.,  J.M.  GERRARD,  P.  GERRARD,  D.W.A. 

WHITFIELD,  and  W.J.  MAHER  1979.  An  aerial  census 
of  bald  eagles  in  Saskatchewan.  J.  Wildl.  Manage. 
43:61-69. 

LEIN,  MR.  and  G.A.  WEBER.  1979.  Habitat  selection  by 
wintering  snowy  owls  (Nyctea  scandiaca).  Can.  Field- 
Nat.  93:176-178. 

MADER,  W.J.  1976.  Biology  of  Harris'  hawks  in  southern 
Arizona.  Living  Bird  14:59-85. 

.  1978.  A  comparative  study  of  red-tailed  hawks  and 

Harris'  hawks  in  southern  Arizona.  Auk  95:327-337. 

MARION,  W.R.  and  R.A.  RYDER.  1975.  Perch-site  prefer- 
ences of  four  diurnal  raptors  in  northeastern  Colo- 
rado. Condor  77:350-352. 

MARKS,  J.S.  1984.  Nest  site  characteristics,  reproductive 
success  and  food  habits  of  long-eared  owls  in  south- 
western Idaho.  M.S.  Thesis,  Univ.  Montana,  Missoula. 
91pp. 

and  I.J.  BALL.  1983-  Burrowing  owl  {Athene  cuni- 

cularia).  Pages  227-242  in  Armbruster,  J.S.,  ed.  Im- 
pacts of  Coal  Surface  Mining  on  25  Migratory  Bird 
Species  of  High  Federal  Interst.  U.S.  Dep.  Inter.,  Fish 
and  Wildl.  Serv.  FWS/OBS-83/85.  Fort  Collins,  CO. 

MARSHALL,  J.R.,  Jr.  1939.  Territorial  behavior  of  the  flam- 
mulated  screech  owl.  Condor  41:71-78. 


MARSHALL,  J.T.  1967.  Parallel  variation  of  North  and 
Middle  American  screech  owls.  Monogr.  1 .  West 
Found.  Vert.  Zool.  Los  Angeles,  CA.  72pp. 

MARTI,  CD.  1979.  Status  of  barn  owls  in  Utah.  Pages  29- 
35  in  Schaeffer,  P.  and  S.M.  Ehlers,  eds.  Owls  of  the 
West:  Their  Ecology  and  Conservation.  Natl.  Audubon 
Soc.  Tiburon,  CA. 

MARTIN,  D.J.  1973.  Selected  aspects  of  burrowing  owl 
ecology  and  behavior.  Condor  75:446-456. 

MATRAY,  P.F.  1974.  Broad-winged  hawk  nesting  and  ecol- 
ogy. Auk  91:307-324. 

MAYFffiLD,  E.G.  1961.  Nest  success  calculated  from  expo- 
sure. Wilson  Bull.  73:255-261. 

MEEHAN,  R.H.  and  R.J.  RITCHIE.  1982.  Habitat  require- 
ments of  boreal  and  hawk  owls  in  interior  Alaska. 
Pages  188-196  in  Ladd,  W.N.  and  P.F.  Schempf,  eds. 
Raptor  Management  and  Biology  in  Alaska  and  West- 
ern Canada.  FWS/AK/Proc.  82.  U.S.  Dep.  Inter.,  Fish 
and  Wildl.  Serv.,  Anchorage,  AK 

MELQUIST,  W.E.  1975.  Eagles  and  osprey:  abstracts  and 
discussion.  Page  142  in  Murphy,  J. R.,  CM.  White,  and 
B.E.  Harrell,  eds.  Population  status  of  raptors.  Proc. 
Conf.  Raptor  Conserv.  Tech.  Raptor  Res.  Rep.  3-  Rap- 
tor Res.  Found.  Vermillion,  SD. 

MIKOL,  S.A.  1980.  Field  guidelines  for  using  transects  to 
sample  nongame  bird  populations.  FWS/OBS-80/58. 
U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  Fort  Collins,  CO. 
26pp. 

MILLSAP,  B.A.  1981.  Distributional  status  of  falconiforms 
in  west  central  Arizona:  with  notes  on  ecology,  repro- 
ductive success  and  management.  Tech.  Note  355. 
U.S.  Dep.  Inter.,  Bur.  Land  Manage.  Serv.  Cen.  Denver, 
CO.  102pp. 

and  S.L.  VANA.  1984.  Distribution  of  wintering 

golden  eagles  in  the  eastern  United  States.  Wilson 
Bull.  96:692-701. 

MINDELL,  D.P.  1983.  Nesting  raptors  in  southwestern 
Alaska:  status,  distribution,  and  aspects  of  biology. 
Tech.  Rep.  8.  U.S.  Dep.  Inter.,  Bur.  Land  Manage. 
Anchorage,  AK  59pp. 

MOSHER,  J.A.,  K  TITUS,  and  MR  FULLER.  1985.  Develop- 
ing a  practical  model  to  predict  nesting  habitat  of 
woodland  hawks  in  Verner, J.,  ML  Morrison,  and  C.J. 
Ralph,  eds.  Modeling  Habitat  Relationships  of  Terres- 
trial Vertebrates.  Univ.  Wisconsin  Press,  Madison. 

,  MR  FULLER,  and  M.  KOPENY.  (Unpubl.  ms.).  Sur- 
veying woodland  raptors:  I.  Responsiveness  to  broad- 
cast of  conspecific  vocalizations.  (To  be  submitted 
to  J.  Wildl.  Manage.). 

-,  and  M.  KOPENY.  (Unpubl.  ms).  Surveying 


woodland  raptors:  II.  Density  and  distribution  of 
nesting  hawks.  (To  be  submitted  to  J.  Wildl. 
Manage. ). 

McCRARY,  M.D.  1981.  Space  and  habitat  utilization  by 
red-shouldered  hawks  {Buteo  lineatns  elegans)  in 
southern  California.  M.S.  Thesis,  California  State  Univ., 
Long  Beach.  85pp. 

McGOWAN,  J.D.  1975.  Distribution,  density  and  productiv- 
ity of  goshawks  in  interior  Alaska.  Fed.  Aid  Wildl. 
Restoration  Proj.  Rep.  W  17-3,  4,  5,  6.  Alaska  Dep. 
Fish  &  Game,  Juneau.  55pp. 

NERO,  R.W.  1980.  The  great  gray  owl.  Smithsonian  Inst. 
Press,  Washington,  DC.  167pp. 

NEWTON,  I.  1979.  Population  ecology  of  raptors.  Buteo 
Books.  Vermillion,  SD.  399pp. 


Raptors 


347 


NICHOLLS,  T.H.  and  D.W.  WARNER.  1972.  Barred  owl 
habitat  use  as  determined  by  radiotelemetry.  J.  Wildl. 
Manage.  36:213-224. 

NOWICKI,  T.  1974.  A  census  of  screech  owls  using  tape- 
recorded  calls.  Jack-Pine  Warbler  52:98-101. 

OLENDORFF,  R.R.  1975.  Population  status  of  large  raptors 
in  northeastern  Colorado  1970-1972.  Pages  185-295 
in  Murphy,  J.R.,  CM.  White,  and  BE.  Harrell,  eds. 
Proc.  Conf.  Raptor  Conserv.  Tech.  Raptor  Res.  Rep.  3. 
Raptor  Res.  Found.  Vermillion,  SD. 

PARKER,  J.W.  1972.  A  mirror  and  pole  device  for  examin- 
ing high  nests.  Bird-Banding  43:216-218. 

.  1975.  Populations  of  the  Mississippi  kite  in  the 

great  plains.  Pages  159-172  in  Murphy,  JR.,  CM. 
White,  and  B.E.  Harrell,  eds.  Proc.  Conf.  Raptor  Con- 
serv. Tech.  Raptor  Res.  Rep.  3-  Raptor  Res.  Found. 
Vermillion,  SD. 

and  J.C  OGDEN.  1979.  The  recent  history  and 


status  of  the  Mississippi  kite.  Am.  Birds  33:119-130. 

PETERSEN,  L.  1 979.  Ecology  of  great  horned  owls  and 
red-tailed  hawks  in  southeastern  Wisconsin.  Tech. 
Bull.  111.  Wisconsin  Dep.  Nat.  Resour.  Madison. 
63pp. 

PHILLIPS,  R.L.,  T.P.  McENEANEY,  and  A.E.  BESKE.  1984. 
Population  densities  of  breeding  golden  eagles  in 
Wyoming.  Wildl.  Soc.  Bull.  12:269-273. 

PLATT,  J.B.  1973-  Habitat  and  time  utilization  of  a  pair  of 
nesting  sharp-shinned  hawks  (Accipiter  striatus 
velox) — a  telemetry  study.  M.S.  Thesis,  Brigham 
Young  Univ.  Provo  UT.  42pp. 

POOLE,  A.  1981.  The  effects  of  human  disturbance  on 
osprey  reproductive  success.  Colonial  Waterbirds 
4:20-27. 

PORTER,  R.D.  and  CM.  WHITE.  1973.  The  peregrine 

falcon  in  Utah,  emphasizing  ecology  and  competition 
with  the  prairie  falcon.  Brigham  Young  Univ.  Sci.  Bull. 
Biol.  Ser.  18:1-74. 

POSTOVIT,  H.R.  1979.  Population  estimates  of  breeding 
raptors  in  North  Dakota  Badlands.  M.S.  Thesis,  North 
Dakota  State  Univ.,  Fargo.  55pp. 

POSTUPALSKY,  S.  1974.  Raptor  reproductive  success: 
some  problems  with  methods,  criteria,  and  terminol- 
ogy. Pages  21-31  in  Hamerstrom,  F.N.  Jr.,  B.E.  Harrell, 
and  R.R.  Olendorff,  eds.  Management  of  Raptors.  Rap- 
tor Res.  Found.  Vermillion,  SD. 

REYNOLDS,  RT.  1982.  North  American  accipiter  hawks. 
Pages  288-289  in  Davis,  D.E.,  ed.  Handbook  of  Census 
Methods  for  Terrestrial  Vertebrates.  CRC  Press.  Boca 
Raton,  FL. 

.  1983.  Management  of  western  coniferous  forest 

habitat  for  nesting  accipiter  hawks.  U.S.  Dep.  Agric, 
For.  Serv.  Gen.  Tech.  Rep.  RM-102.  7pp. 

and  B.D.  LINKHART.  1984.  Methods  and  materials 

for  capturing  and  monitoring  flammulated  owls.  Great 
Basin  Nat.  44:49-51. 

,  EC.  MESLOW,  and  H.M.  WIGHT.  1982.  Nesting 

habitat  of  coexisting  accipiters  in  Oregon.  J.  Wildl. 
Manage.  46:124-138. 

ROSENEAU,  D.G.  1972.  Summer  distribution,  numbers, 
and  food  habits  of  the  gyrfalcon  (Falco  rusticolus  L. ) 
on  the  Seward  Peninsula,  Alaska.  M.S.  Thesis,  Univ. 
Alaska,  Fairbanks.  124pp. 

ROSENFIELD,  R.M.,  J.  BIELEFELDT,  R.K  ANDERSON,  and 
W.A.  SMITH.  1985.  Taped  calls  as  an  aid  in  locating 
Cooper's  Hawk  nests.  Wildl.  Soc.  Bull.  1 3:62-63. 


SCHMUTZ,  J.K,  S.M.  SCHMUTZ,  and  DA.  BOAG.  1980. 

Coexistence  of  three  species  of  hawks  (Buteo  sp.)  in 

the  prairie  parkland  ecotone.  Can.  J.  Zool.  58:1075- 

1089. 
SCHNELL,  G.D.  1967.  Population  fluctuations,  spatial  dis- 
tribution, and  food  habits  of  rough-legged  hawks  in 

Illinois.  Kansas  Ornithol.  Soc.  Bull.  18:21-28. 
SCHNELL,  J.  1979.  Black  hawk  (Buteogallus  anthraci- 

nus).  Tech.  Note  TN-329-  U.S.  Dep.  Inter.,  Bur.  Land 

Manage.  Serv.  Cen.  Denver,  CO.  25pp. 
SCOTT,  V.E.,  KE.  EVANS,  DR.  PATTON,  and  CP.  STONE. 

1977.  Cavity  nesting  birds  of  North  American  forests. 

Agric.  Handb.  511.  U.S.  Dep.  Agric,  For.  Serv.  Wash- 
ington, DC.  122pp. 
SIMINSKI,  DP.  1976.  A  study  of  great  horned  owl  (Bubo 

virginianus)  population  density  with  recorded  calls 

in  northwestern  Ohio.  M.S.  Thesis,  Bowling  Green 

State  Univ.,  Bowling  Green,  OH.  35pp. 
SIMPSON,  MB.  Jr.  1972.  Saw-whet  owl  population  of 

North  Carolina's  southern  Great  Balsam  Mountains. 

Chat.  36:47. 
SMITH,  CF.  1978.  Distributional  ecology  of  barred  and 

great  horned  owls  in  relation  to  human  disturbance. 

M.S.  Thesis,  Univ.  Connecticut,  Storrs.  104pp. 
SNOW,  C.  1973.  Golden  eagle.  Tech.  Note  239.  U.S.  Dep. 

Inter.,  Bur.  Land  Manage.  Serv.  Cen.  Denver,  CO. 

52pp. 
SOKAL,  R.R.  and  F.J.  ROHLF.  1981.  Biometry.  The  princi- 
ples and  practice  of  statistics  in  biological  research, 

2nd  ed.  W.H.  Freeman,  San  Francisco,  CA.  859pp. 
SPRINGER,  M.A.  1978.  Foot  surveys  versus  owl  calling 

surveys:  a  comparative  study  of  two  great  horned  owl 

censusing  techniques.  Inland  Bird-Banding  News 

50:83-92. 
STALMASTER,  M.V.  1983  An  energetics  simulation  model 

for  managing  wintering  bald  eagles.  J.  Wildl.  Manage. 

47:349-359. 
STEENHOF,  K  1976.  The  ecology  of  wintering  bald  eagles 

in  southeastern  South  Dakota.  M.S.  Thesis,  Univ. 

Missouri,  Columbia.  147pp. 
.  1 978.  Management  of  wintering  bald  eagles.  FWS/ 

OBS/78-79-  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv. 

Washington,  DC.  59pp. 
.  1986.  Assessing  raptor  reproductive  success  and 

productivity  in  Millsap,  B.A.  and  KW.  Kline,  eds. 

Raptor  Management  Techniques  Manual.  Vol.  1.  Natl. 

Wildl.  Fed.  Washington,  DC. 
,  S.S.  BERLINGER,  and  L.H.  FREDRICKSON.  1980. 

Habitat  use  by  wintering  bald  eagles  in  South  Dakota. 

J.  Wildl.  Manage.  44:798-805. 
and  M.N.  KOCHERT.  1982.  An  evaluation  of  meth- 


ods used  to  estimate  raptor  nesting  success.  J.  Wildl. 
Manage.  46:885-893- 

STUDER,  CD.  1983.  Where  is  condor  habitat?  Outdoor 
California  44:17-18. 

SYKES,  P.W.  Jr.  1979-  Status  of  the  Everglade  kite  in  Flor- 
ida 1968-1978.  Wilson  Bull.  91:495-511. 

TAYLOR,  A.L.  Jr.  and  ED.  FORSMAN.  1976.  Recent  range 
extension  of  the  barred  owl  in  western  North  Amer- 
ica, including  the  first  records  for  Oregon.  Condor 
78:560-561. 

THIOLLAY,  J.M.  1978.  Population  structure  and  seasonal 
fluctuations  of  the  falconiforms  in  Uganda  national 
parks.  East  Afr.  Wildl.  J.  16:145-151. 


348 


Raptors 


THOMPSON-HANSON,  PA.  1984.  Nesting  ecology  of 
northern  harriers  on  the  Hanford  site,  south-central 
Washington.  M.S.  Thesis,  Washington  State  Univ., 
Pullman.  99pp. 

TITUS,  K.  and  J.A.  MOSHER.  1981.  Nest  site  habitat  se- 
lected by  woodland  hawks  in  the  central  Appalachi- 
ans. Auk  98:270-281. 

TRIMBLE,  S.A.  1975.  Habitat  management  series  for 

unique  or  endangered  species,  merlin  (Falco  colum- 
barius).  Tech.  Note  TN-271.  U.S.  Dep.  Inter.,  Bur. 
Land  Manage.  Serv.  Cen.  Denver,  CO.  4  lpp. 

U.S  BUREAU  OF  LAND  MANAGEMENT.  1979.  Special 
research  report  to  the  Secretary  of  the  Interior.  Bur. 
Land  Manage.,  Boise,  ID.  14 lpp. 

U.S.  FISH  AND  WILDLIFE  SERVICE.  1983-  Northern 
States  bald  eagle  recovery  plan.  U.S.  Dep.  Inter.,  Fish 
and  Wildl.  Serv.  Denver,  CO.  1 1 7pp. 

VILLAGE,  A.  1984.  Problems  in  estimating  kestrel  breed- 
ing density.  Bird  Study  31:121-125. 

WAIAN,  LB.  and  R.C.  STENDALL.  1970.  The  white-tailed 
kite  in  California  with  observations  of  the  Santa  Bar- 
bara population.  Calif.  Fish  Game.  56:188-189- 

WAKELEY,  J.S.  1978.  Factors  affecting  the  use  of  hunting 
sites  by  ferruginous  hawks.  Condor  80:316-326. 

WELLER,  M.W.,  I.C.  ADAMS,  Jr.,  and  B.J.  ROSE.  1955.  Win- 
ter roosts  of  marsh  hawks  and  short-eared  owls  in 
central  Missouri.  Wilson  Bull.  67:189-193- 

WHITE,  CM.  and  S.  SHERROD.  1973.  Advantages  and 
disadvantages  of  the  use  of  rotor-winged  aircraft  in 
raptor  surveys.  Raptor  Res.  7:97- 1 04. 

WIER,  D.N.  1982.  Cliff  nesting  raptors  of  the  Kisaralik 
River,  western  Alaska.  Pages  138-152  in  Ladd,  W.N. 
and  P.F.  Schempf,  eds.  Raptor  Management  and  Biol- 


ogy in  Alaska  and  Western  Canada.  WS/AK  Proc.  82. 
U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  Anchorage,  AK. 

WflNANDTS,  H.  1984.  Ecological  energetics  of  the  long- 
eared  owl  (Asio  otus).  Ardea  72:1-92. 

WILBUR,  S.R.  1978.  The  California  Condor,  1966-76:  a 
look  at  its  past  and  future.  North  Am.  Fauna  72.  U.S. 
Dep.  Inter.,  Fish  and  Wildl.  Serv.  Washington,  DC. 
136pp. 

WILEY,  J. W.  1975.  The  nesting  and  reproductive  success 
of  red-tailed  hawks  and  red-shouldered  hawks  in 
Orange  County,  California,  1973-  Condor  77:133-139. 

WINTER,  J.  1979-  Status  and  distribution  of  the  great  gray 
owl  and  the  flammulated  owl  in  California.  Pages 
60-85,  in  Schaffer,  P.P.  and  S.M.  Ehlers,  eds.  Owls  of 
the  West:  Their  Ecology  and  Conservation.  Nat.  Audu- 
bon Soc.  Tiburon,  CA. 

WRAKESTRAW,  G.F.  1972.  1972  Wyoming  bald  and 

golden  eagle  survey.  Job  Completion  Rep.  W-50-4-21. 
Wyoming  Game  Fish  Comm.,  Cheyenne.  7pp. 

.  1973-  The  1973  Wyoming  bald  and  golden  eagle 

survey.  Am.  Birds  27:716-718. 

ZARN,  M.  1974a.  Spotted  owl  (Strix  occidentalis).  Tech. 
Note  TN-242.  U.S.  Dep.  Inter.,  Bur.  Land  Manage.  Serv. 
Cen.  Denver,  CO.  22pp. 

.  1974b.  Burrowing  owl  {Speotyto  cunicularia 

hypugaea).  Tech.  Note  TN-250.  U.S.  Dep.  Inter.,  Bur. 
Land  Manage.  Serv.  Cen.  Denver,  CO.  25pp. 

.  1974c.  Osprey  {Pandion  haliaetus  carolinensis). 

Tech.  Note  TN-254.  U.S.  Dep.  Inter.,  Bur.  Land  Man- 
age. Serv.  Cen.  Denver,  CO.  4 lpp. 

.  1975.  Rough-legged  hawk  (Buteo  lagopus  sancti- 

johannis).  Tech.  Note  TN-270.  U.S.  Dep.  Inter.,  Bur. 
Land  Manage.  Serv.  Cen.  Denver,  CO.  23pp 


Raptors 


349 


17 

MARSH  AND 
SHOREBIRDS 

Peter  G.  Connors 


Bodega  Research  Associates 

P.O.  Box  247 

Bodega  Bay,  CA  94923 


Editor's  Note:  Marsh  and  shorebirds,  because  of  the 
habitat  they  occupy  and  the  unique  adaptations  for 
feeding  and  breeding,  require  inventory  and  moni- 
toring techniques  quite  different  from  those  used 
for  songbirds,  for  example.  Although  marsh  and 
shorebirds  are  not  a  homogeneous  group  in  terms 
of  habitat,  size,  taxonomic  affinity,  or  life  history, 
they  nevertheless  require  methodology  that  differs 
significantly  from  any  other  species  group,  includ- 
ing waterfowl  with  which  they  often  coexist.  This 
chapter  describes  these  specialized  techniques  and 
their  application. 


INTRODUCTION 

Few  general  statements  concerning  habitat  rela- 
tionships will  apply  accurately  to  all  the  groups  of 
birds  addressed  in  this  chapter.  Marsh  and  shorebirds 
include  two  families  in  the  order  Gruiformes  and 
five  families  in  the  order  Charadriiformes  (Table  1). 
Typical  nesting  habitats  range  from  arctic  tundra 
to  Colorado  River  marshes  and  from  western  range- 
lands  to  the  rocky  California  coastline.  As  if  that 
habitat  diversity  were  not  enough,  species  range 
from  permanent  residents  to  some  of  the  longest  dis- 
tance migrants  in  the  world.  Shorebirds  are  more 
consistently  and  emphatically  migratory  than  any 
other  taxonomic  group  of  North  American  birds,  and 
the  habitats  occupied  in  different  seasons  and  differ- 
ent areas  often  contrast  sharply.  Finally,  species 
range  in  size  from  least  sandpipers  (Calidris  minu- 
tilla)  to  sandhill  cranes  (Grus  canadensis)  and  in 
behavior  and  accessibility,  from  visually  prominent, 
frequently  flocking  species  seen  in  open  landscapes 
to  some  of  the  world's  most  secretive  denizens  of 
densely  vegetated  marshes,  the  rails. 


Table  1.     Families  of  birds  addressed  in  this  chapter 
occurring  regularly  in  Alaska  and  western  States. 


Order 
Family 

Number  of 
Species 

Gruiformes — marsh  birds 
Rallidae — rails  and  coots 
Gruidae — cranes 

7 
2 

Charadriiformes — shorebirds 
Charadriidae — plovers 
Haematopodidae — 

oystercatchers 
Recurvirostridae — stilts  and 

avocets 
Scolopacidae — sandpipers, 

phalaropes,  and  close 

relatives 

7 
1 

2 

35 

Marsh  and  Shorebirds 


351 


This  great  diversity  of  habitats,  life  cycles,  and 
behaviors  requires  a  divided  approach  to  this  discus- 
sion of  habitat  relationships  and  population  estima- 
tion techniques.  In  the  interest  of  brevity,  however, 
not  all  species  can  be  individually  addressed.  Instead 
I  consider  four  major  categories  of  habitat,  attempt- 
ing to  discuss  the  most  general  features  of  species 
group  habitat  use  within  each  category.  This  ap- 
proach will  unavoidably  omit  some  species  that  do 
not  share  characteristics  with  other  species  dis- 
cussed below;  some  of  these  will  be  mentioned  only 
briefly,  with  appropriate  references  if  available. 
Others  can  best  be  investigated  by  methods  dis- 
cussed for  other  species  groups  with  similar  habitat 
use,  behavior,  and  size.  For  example,  American  coots 
(Fulica  americana)  in  breeding  marshes  can  be 
studied  with  grebes  (Podiceps  sp. )  and,  in  winter, 
with  small  waterfowl;  sandhill  cranes  can  be  partly 
addressed  with  large  arctic  breeding  waterfowl. 

The  four  habitat  categories  addressed  in  this 
chapter  are  tundra,  rangeland,  shorelines  (coastal 
and  inland),  and  marshes.  Table  2  outlines  the  spe- 
cies groups  treated  under  each  habitat  heading  and 
emphasizes  the  generalizations  applied  to  efficiently 
treat  the  diversity  of  this  chapter's  subject.  These 
combinations  of  habitats  and  species  groups  define 
the  sections  and  limit  the  scope  of  the  discussion  in 
this  chapter. 


Table  2.     Major  habitats  and  associated  species 
groups. 


Habitat 

Species  Group 

Tundra — 
Arctic  and 
subarctic 

Breeding  shorebirds  and 
cranes 

Rangeland — 
Upland 
and 
wetlands 

Breeding  shorebirds  and 
cranes 

Shorelines — 
Coastal 
and 
inland 

Non-breeding  shorebirds 

Marshes — 
Coastal 
and 
inland 

Breeding  and  non-breeding 
rails  and  coots 

TUNDRA— BREEDING  SHOREBIRDS 

At  most  arctic  and  many  subarctic  sites,  shore- 
birds  are  a  major,  conspicuous  element  of  the  avi- 
fauna. On  the  arctic  coastal  plain  of  Alaska  near 


Prudhoe  Bay,  Troy  et  al.  ( 1983)  found  13  of  24  nest- 
ing species  to  be  shorebirds,  including  8  of  the  10 
species  with  nesting  densities  above  1  nest/knT 
(0.4  mi.2). 

Comparable  results  have  been  recorded  at  Bar- 
row (Myers  and  Pitelka  1980)  and  Point  Storkersen 
(Bergman  et  al.  1977)  as  well  as  at  other  Alaska 
sites.  Total  shorebird  densities  are  usually  higher  on 
tundra  with  numerous  small  wetlands  than  on  well- 
drained  sites  and,  therefore,  higher  on  most  coastal 
plain  tundra  than  in  inland  elevated  areas,  although 
densities  of  some  species  reverse  this  trend  because 
of  habitat  preferences.  All  tundra-breeding  shorebirds 
are  migratory,  arriving  at  nesting  sites  in  May  or 
early  June  in  Alaska  and  departing  soon  after  nesting 
duties  are  finished,  from  late  June  to  September. 
Timing  of  breeding  seasons  matches  the  hatching  of 
insects  on  which  adults  and  newly  hatched  chicks 
depend  (Holmes  and  Pitelka  1968).  In  many  areas, 
shorebirds  move  from  breeding  tundra  areas  to 
nearby  coastal  areas  in  late  summer  before  beginning 
southward  migration  (Connors  et  al.  1979). 

Habitat  Features 

Many  studies  have  related  tundra  habitat  fea- 
tures to  shorebird  populations  in  Alaska,  but  few 
have  dealt  quantitatively  with  fine-scale  variation  in 
habitat  types  or  with  basic  features  of  habitat  struc- 
ture. The  most  detailed  studies  dealing  principally 
with  shorebirds  were  conducted  by  Myers  and  Pi- 
telka ( 1 980 )  at  Barrow  and  Atkasook  and  by  Troy  et 
al.  (1983)  at  Prudhoe  Bay.  Many  other  studies  de- 
scribe habitat  characteristics  somewhat  more  gener- 
ally, often  relating  shorebird  densities  to  different 
classes  of  habitat.  These  include  Spindler  (1978), 
Martin  and  Moiteret  ( 1981 ),  Bergman  et  al.  (  1977), 
Connors  et  al.  (1979),  Williamson  et  al.  ( 1966), 
and  Holmes  and  Black  ( 1973).  Several  other  govern- 
ment reports  and  a  few  graduate  theses  also  address 
this  topic,  and  many  papers  deal  with  habitat  prefer- 
ences of  particular  species.  The  most  quantitatively 
derived  and  most  generally  applicable  conclusions 
from  these  sources  will  be  discussed  here. 

Physical  Features 

Myers  and  Pitelka  (1980)  measured  several 
physical  features  of  shorebird  habitat  use  at  Barrow 
and  Atkasook,  Alaska.  They  measured  average  and 
maximum  microtopographic  relief  on  transect  units, 
average  slope  of  unit  diagonals,  average  water  depth, 
and  proportion  of  sampling  points  with  different 
degrees  of  saturation  or  surface  water  present.  Com- 
bining these  variables  with  vegetation  measurements, 
they  used  multivariate  factor  analysis  to  identify  ma- 
jor gradients  underlying  the  habitat  variation  at  these 
sites  and  found  corresponding  differences  in  shore- 
bird  densities  across  these  gradients,  as  well  as  spe- 
cies differences  in  habitat  use.  Other  researchers 


352 


Marsh  and  Shorebirds 


have  used  classification  schemes  based  principally  on 
differences  in  topography,  wetness,  and  vegetation 
to  identify  differences  among  species  and  among 
habitats  (Bergman  et  al.  1977;  Spindler  1978;  Martin 
and  Moiteret  1981;  Troy  et  al.  1983).  Because  both 
these  approaches  relate  bird  use  to  integrated  de- 
scriptions of  habitat,  it  is  difficult  to  identify  individ- 
ual features  of  prime  importance,  but  topographic 
relief  and  the  occurrence  of  surface  water  certainly 
influence  shorebird  populations. 


that  provide  a  variety  of  nesting  and  foraging  habi- 
tats. However,  this  habitat  mix  is  not  well-repre- 
sented at  the  arctic  sites  where  detailed  shorebird- 
habitat  relationships  have  been  studied.  Directly 
censusing  shorebird  populations  is,  therefore,  often 
preferable  to  habitat  measurements,  except  when 
habitat  measurements  are  obtained  primarily  by  aer- 
ial photography  and  are  compared  with  bird-habitat 
studies  done  in  similar  terrain  within  the  same 
region. 


Individual  species  differ  in  habitat  preferences, 
but  total  shorebird  densities  are  usually  highest  dur- 
ing the  breeding  season  on  tundra  of  variable  topog- 
raphy (such  as  the  polygons  of  ridges  and  troughs 
produced  by  frost  action),  with  numerous  shallow 
wet  troughs  or  small  ponds.  This  combination  pro- 
vides habitat  for  nesting  and  foraging  for  several 
of  the  most  common  tundra  shorebirds.  Well-drained 
sites  without  ponds,  including  highly  polygonized 
areas,  support  lower  densities,  although  some  species 
such  as  golden  plovers  (Pluvialis  dominica)  and 
buff-breasted  sandpipers  (Tryngites  subruficollis) 
may  select  these  habitats.  Areas  of  continuous  marsh, 
lacking  elevated  dry  sites,  are  also  less  used  during 
the  breeding  season  because  of  the  scarcity  of  nest 
sites,  but  species  such  as  pectoral  sandpiper  (Cali- 
dris  melanotos)  use  these  areas  after  chicks  hatch. 

These  generations  do  not  apply  strictly  to  all 
sites  and  to  all  seasons.  However,  on  the  Canning 
River  Delta,  Martin  and  Moiteret  ( 1981 )  measured 
higher  breeding  densities  on  a  mosaic  of  wet  and  dry 
polygonized  tundra  than  on  sites  of  more  homoge- 
neous wet  or  dry  habitat.  Spindler  ( 1978)  found 
similar  results  at  the  Okpilak  Delta,  although  densi- 
ties in  low  wetlands  were  also  high.  At  Barrow,  a 
region  of  little  topographic  relief  but  extensive  wet- 
lands, Myers  and  Pitelka  (1980)  measured  highest 
densities  of  breeding  season  shorebirds  in  areas  of 
low  to  intermediate  polygonization,  with  many  small 
ponds,  and  in  areas  of  high  polygonization,  with 
fewer  ponds.  Inland  on  the  coastal  plain  at  Atkasook, 
where  well-drained  tundra  is  more  widespread,  they 
found  most  shorebirds  nesting  in  low,  non-polygon- 
ized  areas  having  a  wide  range  in  density  of  ponds. 
These  patterns  changed  seasonally,  however,  as  shor- 
ebird use  in  all  three  studies  became  heavier  in  low 
wetland  sites  after  shorebirds  finished  nesting  duties. 

Thus  assessment  of  habitat  value  for  shorebirds 
is  not  a  simple  matter  of  measuring  features  such 
as  density  of  ponds  or  height  of  polygons,  for  exam- 
ple. Furthermore,  regional  differences  in  species 
present  or  in  gross  geomorphology  between  sites 
may  override  the  influences  of  particular  habitat 
feature  studies  at  one  site.  For  example,  the  highest 
shorebird  nesting  densities  on  the  Yukon  Delta  oc- 
cur in  areas  of  heath  tundra  interspersed  with  marsh, 
delta  channels,  and  ponds  (Holmes  and  Black  1973) 


Vegetation 

On  most  arctic  tundra,  vegetation  structure  is 
relatively  monotonous  compared  with  temperate 
ecosystems.  Tundra  ranges  from  almost  unvegetated 
soils  to  dense  stands  of  grasses,  sedges,  and  forbs, 
with  areas  of  low  woody  shrubs  such  as  birches  and 
willows,  often  lower  than  the  surrounding  grasses. 
Plant  species  composition  and,  therefore,  vegetation 
structure  are  to  a  great  extent  determined  by  the 
physical  features  of  the  tundra  within  a  region.  Most 
habitat  classification  schemes  have  defined  tundra 
types  by  combining  these  factors.  For  some  species, 
vegetation  height  or  density  may  strongly  influence 
bird  use,  and  gross  differences  in  plant  species  com- 
position (as  between  shrubby  heath  tundra  and 
graminoid  tundra,  for  example)  may  separate  some 
species.  In  general,  however,  vegetation  is  not  a  sim- 
ple predictor  of  total  shorebird  populations  except 
as  it  corresponds  with  the  differences  in  physical 
features  discussed  above. 


Population  Measurement  Techniques 

Direct  measurement  of  tundra  shorebird  densi- 
ties may  often  be  preferred  over  indirect  population 
evaluation  based  on  habitat  measurements,  primarily 
because  tundra  shorebirds  are  easier  to  census  than 
most  other  non-colonial  nesting  birds.  In  areas  of 
tundra,  such  as  the  North  Slope  and  arctic  coastal 
areas  of  Alaska,  low  vegetation  reduces  visibility 
problems  that  plague  most  censuses  of  birds  in  habi- 
tats with  complex  vegetation  structure,  such  as 
woodlands.  Censusing  shorebirds  does  not  require 
singing  birds  or  estimates  of  distances  to  unseen 
singing  birds. 

On  arctic  coastal  plain  tundra  in  Alaska,  most 
shorebird  individuals,  except  some  incubating  sand- 
pipers and  secretive  species  such  as  long-billed  dow- 
itcher  (Limnodromus  scolopaceus)  and  common 
snipe  (Gallinago  gallinago),  can  be  located  by  an 
experienced  observer  on  a  walking  transect.  The 
relative  tameness  of  most  shorebirds  allows  a  closer 
approach  by  the  census  taker  than  is  possible  for 
more  elusive  species  such  as  waterfowl  and  cranes. 
Reliable  estimates  of  absolute  density  are  often  possi- 
ble with  an  acceptable  amount  of  effort.  Measure- 
ment of  simple  presence  or  absence  of  shorebirds  is 


Marsh  and  Shorebirds 


353 


therefore  seldom  justified,  because  relative  density 
estimates,  at  least,  can  usually  be  obtained  with  little 
extra  effort. 

Sandhill  cranes,  breeding  in  some  of  the  same 
tundra  habitats,  require  a  very  different  approach 
because  of  their  lower  densities  and  greater  flushing 
distances.  Their  large  size,  however,  permits  aerial 
censusing  for  relative  or  absolute  density  measure- 
ments (Riley  1982).  This  species  is  not  addressed  in 
the  discussions  below  of  techniques  designed  to 
assess  densities  of  groups  of  shorebirds  species. 


Relative  Density — Encounter  Rate  Method. 

For  a  simple  estimate  of  relative  density  of  shore- 
birds,  by  species,  an  observer  walks  through  a  site 
recording  all  birds  seen  per  distance  or  time  trav- 
eled. The  census  path  can  be  straight  or  meandering, 
provided  that  it  does  not  incorporate  any  strong 
bias  toward  habitat  of  a  particular  type  that  is  not 
equally  abundant  at  all  sites  compared.  A  continuous, 
roughly  circular  route  from  a  point  of  departure 
such  as  an  airplane  landing  site  may  be  most  efficient 
when  comparing  several  remote  sites.  Distances  are 
best  estimated  from  topographic  maps;  alternatively, 
time  elapsed  at  a  reasonably  constant  hiking  rate 
can  be  used. 

This  method  has  relatively  low  accuracy  and 
precision,  but  it  quickly  and  cheaply  obtains  a  list  of 
most  or  all  common  species  at  a  site  and  a  rough 
estimate  of  their  relative  abundances.  Reliability  of 
this  estimate  depends  on  distance  traveled  (I  suggest 
4  to  12  km  [2  to  7  mi.]);  consistency;  and  experi- 
ence of  the  observer,  weather  conditions,  hiking 
conditions,  and  season.  For  example,  windy  weather 
reduces  the  number  of  birds  heard  and  then  ob- 
served; hiking  rates  through  tussocky  or  very  marshy 
tundra  will  be  slower  than  in  easier  terrain;  and 
birds  differ  in  visibility  seasonally,  depending  on 
their  behaviors,  such  as  displaying,  incubating,  or 
mobbing.  Furthermore,  because  species  differ  in 
detectability  within  a  season,  relative  densities  of 
certain  species  detected  by  this  method  incorporate 
a  somewhat  predictable  bias. 

For  best  results,  the  census  should  be  along 
straight-line  paths  chosen  from  topographic  maps, 
during  the  same  period  of  the  breeding  season,  at  all 
sites  or  in  all  years.  Early  breeding  season  is  best 
for  estimating  relative  breeding  densities,  about  mid- 
June  at  arctic  locations. 

This  method  requires  only  binoculars,  notebook, 
and  hip  boots,  plus  an  experienced  observer  familiar 
with  displays,  vocalizations,  behavior,  and  identifica- 
tion of  the  species  that  may  be  encountered.  Cost 
is  low  except  for  transportation  to  the  sites.  The 
method  should  only  be  used  when  time  and  funds 
are  very  limited;  when  many  sites  will  be  visited 


briefly,  usually  only  one  time  each;  or  when  bird  ob- 
servations must  be  supplemental  to  other  observer 
duties.  In  these  circumstances,  it  becomes  a  very 
efficient  method  of  obtaining  relative  density  data  of 
limited  accuracy.  When  more  time  is  available  or 
greater  accuracy  is  needed,  the  following  methods 
should  be  used. 


Absolute  Density. 


Method  1 — Territory  Spot  Mapping.  On  a 

marked  study  area,  an  observer  records  the  location 
of  all  birds,  noting  especially  locations  of  territorial 
displays  and  nests.  Study  plots  and  data  sheets  should 
be  correspondingly  marked  with  grid  lines  indicating 
50  m  (165  ft)  intervals.  Censuses  should  take  place 
on  several  dates  during  the  early  and  middle 
breeding  season.  I  suggest  three  to  eight  censuses 
per  study  plot.  Combining  censuses  for  each  species 
using  overlay  maps,  the  observer  plots  territories 
or  nest  sites.  Densities  can  be  calculated  only  from 
fractional  portions  of  territories  that  cross  plot 
boundaries  within  the  study  area. 

This  method  is  reasonably  accurate  and  precise. 
Successive  censuses  add  to  the  information  base, 
increasing  accuracy  each  time.  Data  from  all  cen- 
suses are  combined  to  estimate  the  actual  resident 
population.  This  process  requires  judgment  in  the 
analysis  and  should  be  done  by  the  field  observer,  re- 
vising the  map  as  required  after  each  census.  The 
observer  must  know  displays,  nesting  behavior,  and 
identification  of  all  the  local  species,  and  must  be 
familiar  with  nest  location  techniques  for  shorebirds. 
Study  plots  should  be  large  enough  to  include  sev- 
eral territories  of  common  species  (typically  20  to 
100  ha  [50  to  250  a.]). 

Little  equipment  is  needed,  as  in  the  preceding 
method.  One  observer  can  map  30  to  60  ha  (75 
to  150  a.)  of  study  plot  per  day.  Many  of  the  inter- 
fering factors  discussed  above  apply  also  to  this 
method.  However,  the  marked  boundaries,  measured 
area,  and  repeated  censuses  greatly  increase  the  ac- 
curacy of  the  density  estimate.  In  addition,  densities 
correspond  to  known  or  strongly  suspected  breeding 
birds  only,  providing  what  is  often  the  most  useful 
population  measure.  Multiple  censuses,  however, 
require  much  greater  personnel  time  and,  therefore, 
increase  costs. 

Method  2 — Rope-Drag  Nest  Mapping.  The 

rope-dragging  technique  can  be  used  independently 
or  as  a  supplement  to  the  previous  method.  Two 
observers  drag  a  rope  (about  50  m  [165  ft]  long) 
across  the  tundra,  while  a  third  observer  follows 
behind,  noting  locations  of  any  birds  that  flush  from 
nests  as  the  rope  passes  over  or  near  them.  This 
simple  method  locates  incubating  birds  that  "sit 


354 


Marsh  and  Shorebirds 


Rope-drag  technique  using  vehicles. 


close,"  not  flushing,  unless  the  observer  almost  steps 
on  them.  These  are  the  most  difficult  of  shorebird 
nests  to  locate  by  usual  mapping  techniques  and  are 
frequent  among  some  species  such  as  the  long-billed 
dowitcher. 

Method  3 — Marked  Plot  Strip  Transect.  To 

measure  changes  in  total  bird  populations  (not  just 
nesting  pairs)  within  a  season,  between  seasons, 
or  between  sites,  many  researchers  have  used  perma- 
nently marked  transects  treated  as  study  plots  (Con- 
nors et  al.  1979;  Myers  and  Pitelka  1980;  Troy  et 
al.  1983)-  Numbered  wooden  stakes  or  similar  mark- 
ers are  placed  in  a  straight  line  at  50  m  ( 165  ft) 
intervals,  indicating  the  center  line  of  two  adjacent 
rows  of  50  x  50  m  (165  x  165  ft)  plots.  A  single 
observer  walks  a  rectangular  zigzag  path  across  all 
transect  plots,  walking  within  25  or  18  m  (82  or  59 
ft)  of  all  points  on  the  plots.  Alternatively,  two  ob- 
servers walk  the  center  lines  of  the  parallel  rows  of 
plots. 

The  plot  markers  help  observers  estimate  lateral 
distances,  and  additional  stakes  can  be  placed  on 
outer  transect  edges  to  aid  the  estimation.  Time 
required  averages  about  1  observer-hour/kilometer 
(1/2  mile)  of  transect  on  arctic  plain  tundra.  In  areas 
of  taller  shrubby  vegetation,  visual  obstructions  may 
require  increasing  the  amount  of  zigzagging  and 
the  time  required  per  transect.  All  birds  are  re- 
corded, and  the  observer  must  note  locations  and 
movements  to  avoid  counting  individuals  more  than 
once.  Censuses  can  be  repeated  at  regular  intervals 


and  averaged  to  obtain  mean  densities  and  changes 
in  bird  use  with  time.  Transects  can  be  placed 
within  single  habitats  or  randomly  crossing  all  habi- 
tats, depending  on  sampling  design. 

These  plot  transects  work  well  because  of  the 
relatively  high  visibility  of  tundra  shorebirds,  but 
they  are  subject  to  errors  caused  by  differential  visi- 
bility of  species.  Experienced  observers  are  confident 
that  almost  all  shorebirds  are  recorded  by  this 
method  Q.P.  Myers,  F.A.  Pitelka,  D.  Troy,  pers.  com- 
mun. ).  The  most  difficult  application  of  the  tech- 
nique occurs  after  chicks  hatch  and  some  shorebird 
parents  engage  in  mobbing  activities,  flying  across 
several  transect  plots  to  display,  near  the  observer,  a 
potential  predator.  The  observer  must  note  the 
origins  of  these  birds  to  avoid  overcounting  during 
this  period. 

Precision  and,  to  some  extent,  accuracy  of  this 
technique  can  be  assessed  by  replicate  censusing 
of  control  plots  or  of  all  plots.  This  might  be  done 
by  comparing  different  observers  or  a  single  ob- 
server over  time  intervals  of  one  hour,  several  hours, 
or  single  days.  This  assessment  greatly  aids  the  inter- 
pretation of  data  collected  from  single  counts  at 
comparable  sites.  A  similar  approach  of  replicate 
censuses  is  useful  for  all  the  techniques  mentioned 
in  this  chapter. 

Method  4 — Unmarked  Strip  Transect.  When 
several  sites  must  be  evaluated,  usually  with  single 
visits,  an  unmarked  strip  transect  of  fixed  width  ( I 


Marsh  and  Shorebirds 


355 


recommend  50  m  [  165  ft]  total  width)  can  be  used. 
This  is  an  upgraded  version  of  the  Encounter  Rate 
Method  for  relative  density.  By  controlling  the  width 
of  the  census  path,  absolute  densities  can  be  calcu- 
lated. Density  accuracy  depends  on  factors  men- 
tioned earlier,  plus  any  error  in  estimating  the  25  m 
(82  ft)  width  on  either  side  of  the  observer's  path. 
Accuracy  also  suffers  compared  with  Marked  Plot 
Transects  by  the  absence  of  replicate  estimates,  but 
Method  4  has  an  obvious  advantage  by  saving  time 
and  money. 

Method  5— Variable  Width  Line  Transect. 

Several  variations  of  this  method  have  been  devel- 
oped, principally  for  passerine  censusing  in  wooded 
habitats.  Corrections  have  been  applied  to  account 
for  differential  detectability  of  species  (Emlen  1971, 
1977;  Jarvinen  and  Vaisanen  1975;  Ralph  and  Scott 
1981).  Briefly,  all  birds  are  recorded  on  either  side 
of  a  central  transect  line,  along  with  the  perpendicu- 
lar distance  from  each  bird  to  the  transect  line.  Tun- 
dra shorebirds  do  not  pose  the  same  visibility 
problems  as  woodland  songbirds,  so  the  additional 
complication  of  estimating  horizontal  distances  to  all 
birds  is  probably  not  warranted. 

Discussion 

Tundra  shorebird  habitat  relationships  have 
been  investigated  at  several  arctic  sites.  The  closest 
scrutiny  focused  on  individual  species  preferences 
and  seasonal  changes  in  habitat  use.  Nevertheless,  no 
clear  definition  of  habitat  features  that  can  predict 
shorebird  abundances  has  yet  emerged.  The  most 
widely  applicable  generalization  from  these  studies 
suggests  that  total  shorebird  densities  will  usually  be 
highest  in  areas  of  numerous  small  wetlands  with 
moderate  microtopographic  relief,  providing  a  fine- 
grained mosaic  of  nest  sites  and  foraging  sites. 

The  low  tundra  vegetation,  moderate  conspicu- 
ousness,  and  non-elusive  character  of  most  shorebird 
species  at  most  tundra  sites  make  shorebird  census- 
ing on  breeding  grounds  relatively  easy.  For  this 
reason,  measures  of  simple  presence  or  absence  are 
not  recommended,  because  they  are  easily  replaced 
with  estimates  of  relative  or  absolute  density.  All 
these  techniques  require  at  least  the  ability  to  easily 
identify  all  local  species  in  a  variety  of  plumages,  and 
all  are  more  reliable  when  observers  know  the  be- 
havioral peculiarities  and  vocalizations  of  shorebird 
species.  Methods  should  be  chosen  on  the  basis  of 
available  time  and  personnel,  number  of  sites  to  be 
monitored,  duration  of  monitoring  planned,  and  the 
information  needs  of  the  monitoring  program. 

RANGELAND— BREEDING  SHOREBIRDS 

In  contrast  to  their  prominence  on  arctic  tun- 
dra, shorebirds  are  much  less  common  and  conspicu- 
ous on  western  rangelands,  but  several  species  nest 


there,  and  additional  species  are  found  throughout 
the  West  during  migration.  A  brief  summary  of  habi- 
tat affinities  and  census  techniques  follows. 

Habitat  Features 

Western  rangeland  encompasses  a  greater  vari- 
ety of  contrasting  habitat  types  than  can  readily  be 
treated  in  this  short  section,  from  prairies  and  plains 
to  mountain  slopes  and  valleys;  from  deserts  to  rich 
grasslands,  including  marshes,  streams,  fresh  and 
saline  lakes,  over  a  wide  range  of  elevations  and  win- 
ter climates.  Most  nesting  shorebird  species  can  be 
assigned  to  groups  associated  with  prairie  grasslands, 
marshes,  or  stream  and  lake  shores,  although  these 
categories  overlap  for  some  species. 

Grassland-nesting  shorebird  species  include 
mountain  plover  (Charadrius  montanus),  a  long- 
billed  curlew  (Numenius  americanus),  upland  sand- 
piper (Bartramia  longicauda),  and  sometimes  kill- 
deer  (Charadrius  vociferus).  Physical  characteristics 
of  different  grasslands  can  be  described  by  measur- 
ing such  factors  as  mean  vegetation  height;  percent- 
age vegetation  cover  types  (grass,  forb,  shrub, 
cactus);  slope  and  topographic  diversity;  and  dis- 
tance to  water. 

Intensive  studies  of  long-billed  curlews  in  Idaho 
(Jenni  et  al.  1982;  Redmond  et  al.  1981;  Redmond 
1984)  have  demonstrated  that  curlew  densities  in 
that  area  are  highest  in  areas  of  low  topographic 
slope,  low  vegetation  height,  and  low  vertical  vegeta- 
tive cover.  Mountain  plovers  in  Colorado  also  se- 
lected low  slope,  shortgrass  habitats  (Graul  1975). 
Upland  plovers  occur  in  tallgrass  or  mixed  grass 
prairie  habitats  (Wiens  1973).  Since  these  species 
are  conspicuous  enough  in  these  habitats,  direct 
censusing  of  birds  during  the  breeding  season  is 
more  effective  than  attempts  to  assess  habitat 
suitability. 

Marsh  and  pond  nesting  shorebirds  occurring 
within  western  rangelands  are  American  avocet  (Re- 
curvirostra  americana),  black-necked  stilt  (Himan- 
topus  mexicanus),  willet  (Catoptrophorus 
semipalmatus),  marbled  godwit  (Limosa  fedoa), 
common  snipe,  and  Wilson's  phalarope  (Phalaropus 
tricolor).  Distributions  and  habitats  of  these  species 
vary  widely  and  have  not  all  been  the  subjects  of 
intensive  habitat  studies.  Furthermore,  habitat  use  by 
a  single  species  may  change  regionally  as  wetland 
types  change  over  large  distances  of  western  North 
America.  Wetlands  in  many  of  these  areas  are  sea- 
sonal or  semipermanent  and  have  been  classified 
according  to  permanence  by  Stewart  and  Kantrud 
(  1971 )  for  the  northern  prairie  region. 

Black-necked  stilts  and  American  avocets  nest  in 
similar  habitats,  usually  along  shorelines  and  grassy 
or  unvegetated  flats  near  fresh  water,  brackish  or 


356 


Marsh  and  Shorebirds 


Black-necked  stilt  turning  eggs. 


alkaline  marshes,  and  ponds.  Foraging  habitats,  usu- 
ally shallow  water  areas  or  wet  mud  flats,  may  be 
continuous  with  or  separated  from  nest  territories 
(Gibson  1971;  Hamilton  1975).  In  prime  areas,  nest 
densities  may  be  very  high,  clustered  in  semi-colon- 
ial situations.  This  distribution  of  nests  and  the  usu- 
ally low  vegetation  facilitate  direct  spot-mapping 
techniques  for  nesting  birds. 

Willets  and  marbled  godwits  nest  near  prairie 
marshes  and  ponds  surrounded  by  grasslands.  Both 
species  prefer  permanent  or  semipermanent  pond 
types,  either  fresh  or  brackish,  and  nest  in  adjacent 
grassy  areas  (Ryan  1985).  Godwits  use  surrounding 
uplands  in  addition  to  marshes  for  foraging  and  pre- 
fer short  to  medium  grass  heights  for  nesting.  Willets 
use  taller  grass  areas  (Ryan,  unpubl.  ms).  Wilson's 
phalaropes  nest  in  ponds  and  marshy  wetlands  of 
prairie  and  basin  areas,  both  fresh  and  saline,  and 
common  snipes  nest  in  marshes  of  the  northern 
States. 


Shoreline  nesting  species  include  killdeer, 
snowy  plover  (Charadrius  alexandriniis),  spotted 
sandpiper  (Actitis  macularia),  and  in  some  northern 
prairie  regions,  piping  plover  (Charadrius  melodus). 
These  are  usually  associated  with  lakes,  ponds,  or 
streams  (except  that  killdeer  may  nest  well  away 
from  water)  and  nest  in  open  areas  or  areas  of  sparse 
or  short  vegetation  (Bent  1929;  Renaud  1979;  Page 
et  al.  1983;  Oring  et  al.  1983;  Weseloh  and  Weseloh 
1983). 

Sandhill  cranes  also  nest  in  western  rangeland 
areas,  in  or  near  marshes  or  ponds  (Bent  1926;  Lit- 
tlefield  and  Ryder  1968;  Drewien  1973). 

As  with  shorebirds  breeding  on  tundra,  many 
rangeland  species  are  relatively  conspicuous  in  the 
sparsely  vegetated  or  shortgrass  habitats  they  occupy 
in  western  rangelands.  This  circumstance  facilitates 
direct  measurement  of  shorebird  or  crane  densities. 
For  species  associated  with  wetlands,  however,  locat- 
ing and  mapping  areas  of  wetlands  by  wetland  type 
may  provide  a  large-scale  estimate  of  relative  abun- 
dance of  these  species.  This  approach  requires  cali- 
bration from  bird  studies  done  within  the  same 
region.  Information  relating  presence,  absence,  or 
relative  or  absolute  densities  of  species  to  wetland 
type  or  wetland  size  class  can  then  be  applied  to 
areawide  inventories  of  wetlands.  Weber  et  al. 
(1982)  provided  an  example  of  the  kind  of  calibra- 
tion information  required  on  northern  prairies. 

Population  Measurement  Techniques 

Census  methods  must  be  chosen  to  match  the 
species  or  habitat  characteristics  within  this  diverse 
group  of  species.  Many  of  the  methods  listed  for 
tundra  shorebirds  are  also  applicable  to  rangeland 
species  and  are  listed  here  briefly  with  the  habitats 
or  species  of  concern.  A  few  additional  approaches 
are  described  more  fully. 

Relative  Density — Encounter  Rate  Method. 

The  encounter  rate  method  is  generally  less  useful  in 
rangeland  shorebird  habitats  than  in  tundra,  where 
the  nesting  habitats  of  large  numbers  of  species  oc- 
cur in  a  fine-grained  mosaic.  Nevertheless,  walking 
around  wetlands  or  along  shorelines  during  the 
breeding  season,  or  traversing  grasslands  on  foot  or 
horseback,  can  provide  easy  but  limited  information 
on  relative  density  of  nesting  birds.  Large  areas  can 
be  covered  by  aircraft  searching  for  cranes,  but  the 
effort  involved  argues  for  attempting  absolute  den- 
sity estimates. 


Absolute  Density. 

Method  1 — Nest  Spot  Mapping.  Most  studies 
of  shorebird  populations  breeding  in  the  western 
U.S.  have  been  concentrated  on  locating  all  nests 


Marsh  and  Shorebirds 


357 


within  a  study  area.  Techniques  depend,  to  some 
extent,  on  the  species  involved  and  habitat  features, 
but  in  all  instances,  actual  ground  searching  is 
required.  This  is  usually  initiated  by  observing 
locations  of  birds  flushing  from  or  returning  to 
brooding  activities,  from  a  vantage  point  inside  or 
outside  the  studied  habitat.  For  large  birds  (sandhill 
crane,  long-billed  curlew,  marbled  godwit,  avocet), 
initial  observations  may  be  made  from  a  road,  dike, 
or  even  an  airplane.  For  smaller  birds,  initial 
observations  may  require  field  work  within  the 
habitat.  For  many  species,  thorough  hiking  may  be 
required  to  flush  brooding  birds  or  to  elicit 
antipredator  responses.  For  species  nesting  in  high 
densities  within  limited  habitats,  such  as  the  loosely 
colonial  breeders,  American  avocet  and  black-necked 
stilt,  spot-mapping  techniques  to  locate  all  nests  or 
territories  can  be  fairly  efficient.  In  other  situations, 
this  would  be  time-consuming. 

Method  2 — Rope-Drag  Nest  Mapping.  This 
method  is  usually  used  only  as  a  supplement  to 
Method  1  and  is  most  useful  for  species  that  nest  in 
dense  vegetation  and  those  that  may  not  flush  with- 
out a  close  approach,  such  as  the  marbled  godwit 
(Ryan  1985).  The  technique  is  described  in  the  tun- 
dra section. 

Method  3 — Strip  Transect.  This  method, 
which  requires  locating  birds  within  a  linear  transect 
of  a  fixed  width,  can  be  applied  as  a  walking  transect 
through  grasslands,  around  wetlands,  or  near  shore- 
lines in  a  manner  similar  to  the  tundra  transect  de- 
scribed in  the  preceding  section.  Alternatively,  to 
deal  with  the  low  densities  of  some  rangeland  spe- 
cies and  the  patchy  distribution  of  their  habitats,  the 
method  can  be  ( 1 )  a  driving  transect  along  a  road, 
counting  all  individuals  of  large  species  within  a 
standard  perpendicular  distance;  (2)  driving  along  a 
road  and  stopping  at  predetermined  stations  for  cen- 
susing;  or  (3)  a  road  transect  linking  wetland  census- 
ing  stations.  An  aircraft  flying  transect  lines  with 
observers  censusing  a  specified  width  along  the  tran- 
sect line  is  a  useful  variant  for  cranes. 

Method  4 — Variable  Width  Line  Transect. 

Variations  of  this  approach,  described  for  tundra,  can 
also  be  applied  to  solve  the  particular  problems  of 
rangeland  species  and  habitats.  Unlike  most  tundra 
and  shoreline  shorebirds,  long-billed  curlews  present 
special  difficulties,  in  both  size  and  breeding  habitat, 
which  affect  census  techniques.  Redmond  et  al. 
( 1981 )  compared  three  methods  of  censusing  cur- 
lew breeding  populations  in  Idaho  and  recom- 
mended a  modified  Finnish  Line  Transect  method. 
Observers  drive  a  route  along  existing  roads,  record- 
ing all  curlews  sighted  within  500  m  ( 1,650  ft)  of 
the  road,  and  estimate  perpendicular  distances  to  all 
birds  sighted.  The  census  is  repeated  five  times  on 
different  days,  and  pooled  data  are  corrected  for 


Avocet  turning  eggs. 


distance  effects  of  detectability  (Emlen  1971;  Jarvi- 
nen  and  Vaisanen  1975)  before  calculating  densities. 
Censuses  should  be  done  soon  after  curlews  arrive 
on  breeding  grounds  (late  March  or  early  April). 
Results  of  this  and  the  strip  transect  method  were 
fairly  accurate  when  compared  with  results  of  terri- 
tory mapping  studies  done  on  the  same  area  (Red- 
mond et  al.  1981).  As  with  tundra  shorebird 
censusing,  the  change  in  behavior  during  the  brood- 
rearing  period,  with  adults  flying  long  distances  to 
mob  potential  predators,  introduces  errors.  Census- 
ing early  in  the  breeding  season  avoids  this  problem. 

Discussion 

Western  rangeland,  including  such  diverse  habi- 
tats as  grasslands,  shrub  deserts,  mountain  slopes, 
marshes,  and  alkali  lakes,  to  name  a  few,  poses 
greater  problems  for  generalizing  conclusions  than 
other  major  habitat  areas  addressed  in  this  chapter. 
Most  rangeland  nesting  shorebirds,  however,  fall  into 
three  habitat  categories,  as  described  above,  and 
census  techniques  applicable  to  these  categories  and 
species  are  similar  to  those  for  comparable  situations 
elsewhere.  The  major  modifications  necessary  arise 
from  the  large  size  and  low  density  of  grassland 
shorebirds  or  cranes,  compared  with  most  tundra 
shorebirds,  and  the  patchy  distribution  of  western 
ponds,  marshes,  and  lakeshores.  On  a  large-scale 
area,  the  patchiness  of  gross  habitat  types,  such  as 
marshy  wetlands,  permits  successful  assessment  of 
probable  shorebird  locations,  but  assessments  of  fine 
habitat  distinctions  will  usually  be  less  useful  than 
direct  censuses  of  birds. 


358 


Marsh  and  Shorebirds 


SHORELINES— MIGRANT  SHOREBIRDS 

During  much  of  the  year,  from  late  summer  to 
late  spring,  shorebirds  are  most  common  in  or  near 
shoreline  habitats,  predominantly  coastal  but  includ- 
ing inland  shores  of  lakes  and  marshes.  This  descrip- 
tion encompasses  a  great  variety  of  habitats  ranging 
from  muddy  freshwater  ponds  and  flooded  fields, 
to  sandy  coastal  beaches;  from  tidal  mudflats  and 
saltmarshes,  to  exposed  rocky  coastlines;  and  from 
arctic  Alaska,  to  southern  California  and  beyond. 
Naturally,  shorebird  use,  habitat  features,  and  census 
methods  vary  over  this  range,  but  some  habitat  char- 
acteristics are  common  and  these  guide  the  choice 
of  census  techniques. 

First,  most  shoreline  habitats  of  importance  to 
shorebirds  are  open  and  free  of  much  vegetation 
structure  or  topographic  relief.  Shorebirds  in  these 
circumstances  are  even  more  readily  located  than  on 
tundra  or  rangelands.  Second,  most  of  these  habitats 
occur  where  land  meets  water,  in  areas  where  acces- 
sible invertebrate  prey  in  high  densities  cause  forag- 
ing shorebirds  to  concentrate  locally.  This  factor  also 
facilitates  shorebird  censusing.  However,  the  tidal 
nature  of  most  coastal  shoreline  habitats  and  the 
flocking  habits  of  most  migrant  and  wintering  shore- 
birds  inject  elements  of  instability  into  shorebird 
density  measurements. 


Migrant  marsh  and  shorebirds  move  frequently 
to  exploit  food  sources,  escape  predators,  or  con- 
tinue migrations.  This  instability  can  usually  be  ac- 
commodated in  at  least  one  of  several  ways.  For 
example,  biologists  might  census  when  shorebirds 
are  most  predictably  concentrated  and  stable,  as 
at  high  tide  roosts  in  coastal  areas.  They  might  al- 
ways census  under  the  same  set  of  conditions  that 
relate  to  movements,  such  as  tide  height,  time  of 
day,  and  calendar  date.  They  might  census  regularly 
and  repeatedly  throughout  a  period  of  migration 
or  simultaneously  census  an  entire  local  population, 
using  multiple  observers.  The  following  discussion 
addresses  the  most  generally  important  habitat  fea- 
tures and  most  widely  applicable  census  methods  for 
inland  and  coastal  shoreline  habitats  throughout 
Alaska  and  the  western  States. 

Habitat  Features 

Most  migrating  and  wintering  shorebirds  do  not 
require  vegetative  cover  for  nests  or  chicks  or  for 
their  own  protection.  This  frees  them  to  use  habitats 
where  their  food,  mainly  small  invertebrates  of  many 
orders,  are  most  abundant.  In  coastal  locations,  these 
habitats  are  usually  bare  or  sparsely  vegetated,  the 
exceptions  being  mainly  algae-covered  rocks  and 
mudflats,  and  saltmarshes  with  low  or  sparse 
vegetation. 


Greater  sandhill  cranes,  a  western  rangeland  nester. 


Marsh  and  Shorebirds 


359 


The  primary  features  of  habitat  that  relate  to 
shorebird  densities  must  usually  be  those  features 
that  result  in  the  highest  densities  of  available  shore- 
bird  prey  in  an  area,  because  shorebirds  in  winter 
are  generally  distributed  in  relation  to  availability  of 
prey  (Goss-Custard  1970;  Bryant  1979;  Myers  et  al. 
1979;  Rands  and  Barkham  1981).  In  arctic  Alaska, 
these  conditions  prevail  in  two  separate  situations  in 
late  summer  after  shorebird  breeding  has  been  com- 
pleted on  the  tundra.  Densities  of  some  species  (red 
phalaropes  [Phalaropus  fulicaria],  red-necked  phal- 
aropes  [P.  lobatus],  and  sanderlings  [Calidris  alba]) 
are  highest  along  beaches  or  spits,  bars,  and  barrier 
islands.  Numbers  of  several  other  species  (especially 
dunlin  [Calidris  alpina],  semipalmated  [C.  pusilla], 
western  [C.  mauri]  and  pectoral  sandpipers,  and 
long-billed  dowitchers)  are  highest  on  mudflats,  at 
edges  of  sloughs,  and  around  shallow  pools  in  short 
vegetation  saltmarshes  (Connors  et  al.  1979;  Con- 
nors 1984).  These  habitats  usually  support  higher 
shorebird  use  than  other  arctic  shorelines,  but  densi- 
ties are  highly  variable  from  one  site  to  another  and 
cannot  be  predicted  with  accuracy. 

On  the  Bering  Sea  coast,  similar  habitats,  espe- 
cially river  deltas,  lagoon  mudflats,  and  saltmarsh 
margins,  attract  highest  shorebird  densities  during 
post-breeding  movements,  from  July  through  Octo- 
ber (Shields  and  Peyton  1979;  Gill  et  al  1981;  Gill 
and  Handel  1981).  From  a  large-scale  view,  these 
coastal  habitats  in  arctic  and  subarctic  Alaska  are  lo- 
calized and  discrete,  and  their  occurrences  correlate 
well  with  post-breeding  shorebird  densities  through- 
out the  region.  Field  recognition  of  these  habitats  is 
straightforward  and  is  frequently  possible  from  maps 
and  aerial  photography.  Quantitative  characterization 
of  these  gross  habitat  features  is  easily  accomplished 
by  referencing  properties  of  water  depth,  slope,  dis- 
tance, or  extent  from  the  water's  edge;  substrate 
sediment  size;  substrate  penetrability;  vegetation 
density  and  height;  plant  species  composition;  salin- 
ity or  saltwater  influence;  and  surrounding  landform 
type. 

On  a  finer  scale,  little  quantitative  work  has 
been  done  to  isolate  those  features  within  mudflats 
or  saltmarsh,  for  example,  that  correlate  with  bird 
density  distributions  within  the  habitat.  As  in  other 
coastal  mudflat  environments,  density  of  available 
invertebrate  prey  is  probably  the  best  predictive 
feature  to  relate  to  shorebird  densities.  This  is  not 
always  easily  measured,  and  habitat  correlates  of 
prey  densities  are  unstudied  in  these  areas.  For  most 
management  purposes,  the  simple  presence  of  more 
than  trace  amounts  of  wet  mudflats,  shallow  muddy 
pools,  and  sparsely  vegetated  saltmarsh  with  wet 
mud  and  shallow  pools  is  sufficient  to  predict  ele- 
vated levels  of  seasonal  shorebird  use. 

A  few  physical  and  vegetative  features  can  be 
used  to  identify  and  separate  arctic  saltmarshes  from 


freshwater  tundra.  First,  because  lunar  tidal  fluctua- 
tions are  so  slight  (about  15  cm  [6  in.]  at  Barrow, 
Alaska,  for  example),  saltmarsh  areas  are  maintained 
by  occasional  flooding  during  storms  in  late  summer 
and  fall.  Therefore,  saltmarsh  is  always  located  in 
areas  with  access  routes  for  storm-flood  saltwater 
from  oceans  or  lagoons.  The  importance  of  storm 
flooding  leaves  another  clue  in  the  form  of  driftwood 
lines  on  coastal  land,  inland  from  any  saltmarshes. 
Depending  on  the  frequency  of  storm  flooding  repre- 
sented by  different  driftwood  lines,  vegetation  shore- 
ward of  the  driftwood  may  be  very  salt-tolerant, 
mildly  salt-tolerant,  or  fresh  tundra  pioneers  invading 
after  the  previous  salt-killing  of  tundra  vegetation. 

In  the  permanent  arctic  saltmarsh  zone  of  most 
interest,  two  salt-tolerant  plants  are  characteristic 
indicators:  Puccinellia  phryganodes,  a  prostrate 
grass,  and  Carex  subspathacea,  a  very  short  sedge. 
Both  have  a  pronounced  reddish  color  in  some  life 
stages  and,  together  with  the  reddish  iron  oxides 
in  saltmarsh  mud,  create  a  final  color  attribute  that  is 
helpful  in  identifying  and  delineating  arctic  salt- 
marshes. Within  these  vegetated  areas,  however,  the 
shallow  pools,  wet  mudflats,  and  muddy  margins 
show  the  heaviest  shorebird  use. 

For  phalaropes,  sanderlings,  and  lesser  numbers 
of  several  other  species  feeding  primarily  on  zoo- 
plankton  along  arctic  shorelines,  the  habitat  feature 
of  most  importance  seems  to  be  the  landform  on 
which  arctic  beaches  occur.  Spits  and  islands  gener- 
ally have  higher  densities  of  these  species  than  main- 
land beaches  (Johnson  and  Richardson  1981; 
Connors  1984).  Vegetation  and  substrate  characteris- 
tics are  less  important  than  landform  in  predicting 
these  bird  densities. 

Farther  south  in  Alaska,  coastal  shorebird  densi- 
ties are  usually  highest  in  areas  of  extensive  mudflats 
and,  for  some  species,  on  rocky  intertidal  beaches. 
The  Copper  River  Delta  is  an  important  habitat  area 
with  such  large  numbers  of  migrating  shorebirds, 
especially  during  spring,  that  total  population  num- 
bers can  only  be  estimated  rather  than  directly 
counted  (Isleib  1979;  Mickelson  et  al.  1980;  Senner 
1 979 ).  The  ample  prey  base,  principally  of  small 
molluscs  on  the  extensive  mudflats  and  sandflats  of 
that  area,  is  responsible  for  the  massive  shorebird 
concentrations  (Senner  1979).  In  Washington,  Ore- 
gon, and  California,  the  largest  numbers  of  shore- 
birds  occur  on  mudflats  and  sandflats  of  lagoons, 
estuaries,  and  fresh-  and  saltwater  marshes;  on 
coastal  beaches;  and  on  rocky  intertidal  shorelines, 
with  different  groups  of  species  in  different  habitats. 

Several  studies  have  addressed  species  habitat 
preferences  among  coastal  wetland  habitat  types  (for 
example,  Recher  1966;  Gerstenberg  1979;  Page  et 
al.  1979).  These  do  not,  however,  permit  assessment 
of  wildlife  values  of  different  examples  of  a  single 


360 


Marsh  and  Shorebirds 


wetland  type,  except  on  the  basis  of  the  extent  of 
habitat.  Indeed,  the  population  densities  of  some 
wetland  habitat  areas  seem  to  depend  on  the  loca- 
tions of  nearby  wetlands  of  other  types,  rather  than 
on  strictly  intrinsic  properties  (Page  et  al.  1979; 
Connors  et  al.  1 98 1 ).  Thus,  one  cannot  yet  profitably 
judge  shorebird  wetland  habitats  by  measuring  par- 
ticular habitat  variables,  except  in  a  few  specialized 
instances.  Sanderling  densities  on  beaches  in  central 
California,  for  example,  are  sometimes  correlated 
with  beach  slope  and  sediment  size,  because  these 
are  correlated  with  prey  densities  (Myers  et  al.  1979; 
Connors  et  al.  1981).  On  a  larger  scale,  however, 
one  can  assess  areas  by  the  presence  and  extent  of 
wetland  habitats  that  provide  foraging  areas  of  the 
types  selected  by  particular  species. 

The  situation  for  inland  areas  is  similar.  Wet- 
lands providing  appropriate  species  foraging  areas 
will  attract  shorebirds,  but  measurements  of  physical 
or  vegetational  features  within  those  habitats  do 
not  permit  assessment  of  relative  value.  This  is  not 
generally  a  handicap,  however,  because  direct  mea- 
surement of  the  shorebird  populations  in  different 
areas  gives  the  best  assessment  of  their  relative  val- 
ues, and  this  measurement  is  usually  not  prohibi- 
tively difficult  to  obtain. 

On  a  large  scale,  total  area  of  different  wetland 
types  sets  limits  on  migrant  marsh  or  shorebird  pop- 
ulations of  most  species  at  inland  sites.  For  example, 
most  migrant  sandpipers  will  occur  only  on  mudflats 
and  flooded  fields  or  the  beaches  and  shorelines  of 
ponds  and  lakes;  Wilson's  phalaropes  concentrate  in 
large  saline  lakes  during  fall  migration.  Sandhill 
cranes  roost  on  river  bars  and  forage  in  croplands. 
The  most  efficient  way  of  assessing  the  value  of  habi- 
tat to  shorebirds  of  different  habitats,  however,  re- 
mains the  direct  censusing  of  bird  populations  found 
there.  For  example,  recent  efforts  to  relate  numbers 
of  roosting  sandhill  cranes  on  Nebraska  river  bars 
to  habitat  features  such  as  water  depth  and  water  ve- 
locity have  established  significant  relationships. 
These  do  not  permit  model  predictions  of  crane 
numbers  at  other  river  locations  away  from  the  stud- 
ied sites  (Latka  and  Yahnke,  unpubl.  ms),  however. 

Except  for  the  influence  of  the  tidal  period  on 
bird  movements  and  habitat  availability,  most  aspects 
of  the  census  techniques  discussed  below  apply 
equally  well  to  coastal  and  inland  habitats.  In  the 
western  U.S.,  however,  most  inland  and  coastal  wet- 
lands differ  in  annual  seasonality  of  bird  use,  because 
inland  wetlands  away  from  the  coast  or  away  from 
low  elevations  in  the  Southwest  are  usually  frozen 
during  winter.  These  areas  are  used  by  shorebirds 
only  during  spring  and  fall  migration,  mainly  April 
through  May  and  July  through  September.  During 
these  periods,  however,  densities  of  migrant  shore- 
birds  may  be  very  high,  even  in  temporary,  seasonal 
wetlands.  Flooded  fields  in  springtime,  in  particular. 


may  attract  large  numbers  of  northbound  migrant 
shorebirds,  perhaps  because  of  elevated  densities  of 
invertebrate  prey  (Krapu  1974).  The  temporary  na- 
ture of  this  shorebird  habitat,  of  course,  requires  that 
any  habitat  evaluation  program,  like  shorebird  census 
studies,  focus  on  conditions  during  the  brief  migra- 
tion period. 


Population  Measurement  Techniques 


Presence — Aerial  Survey  Method.  Flying  in  a 
helicopter  or  fixed-wing  aircraft,  one  or  two 
observers  record,  on  tape  or  on  maps,  the  presence 
of  flocks  of  shorebirds  in  different  wetland  locations 
over  large  surveyed  areas.  Flight  elevation  should 
be  25  to  60  m  (80  to  200  ft)  to  promote  flushing  of 
roosting  or  foraging  birds.  Observers  can  record 
flock  size  or  identify  species  or  species  groups 
whenever  possible. 

This  method  is  useful  only  in  areas  where  al- 
most nothing  is  known  of  shorebird  distributions  for 
the  time  period  surveyed.  Except  in  unusual  circum- 
stances, aerial  surveys  cannot  provide  accurate 
estimates  of  population  size  because  of  the  low  relia- 
bility with  which  birds  are  sighted  or  counted  from 
the  air.  Compounding  this  problem,  only  well- 
marked  species  that  fly  under  the  aircraft  (dowitch- 
ers,  willets,  black-bellied  plovers  [Pluvialis  squata- 
rola\  or  phalaropes  sitting  on  calm  water,  etc.)  can 
be  identified  consistently.  Nevertheless,  if  large  inac- 
cessible areas  or  large  numbers  of  inaccessible,  small 
wetlands  must  be  surveyed  for  presence  of  migrant 
shorebird  flocks,  the  high  cost  of  aircraft  may  be 
justified. 


Relative  Density.  Most  techniques  of  ground 
censusing  for  migrant  shorebirds  can  be  targeted  to 
either  relative  or  absolute  density  estimates, 
depending  on  the  physical  arrangement  of  local 
habitat  areas  and  the  schedule  and  range  of 
shorebird  movements  among  different  areas.  In 
general,  relative  density  estimates  are  obtained  when 
shorebirds  are  patchily  distributed  over  the  available 
habitat  of  interest,  when  counts  are  too  limited  to 
sample  enough  of  the  patchiness,  or  when  parts  of 
local  populations  are  counted  in  special  circum- 
stances, such  as  at  roosting  sites.  Actual  censusing 
methods  are  discussed  only  under  Absolute  Density 
below,  but  the  applicability  of  census  results  must  be 
judged  by  biologists  on  the  basis  of  factors  just 
mentioned.  Although  actual  numbers  of  birds  in  an 
area  are  counted  in  either  situation,  local  population 
distributions  and  movements  and  local  habitat 
distributions  may  require  interpreting  census 
numbers  as  relative  rather  than  absolute  densities. 


Marsh  and  Shorebirds 


361 


Absolute  Density.  For  most  of  the  methods 
described  here,  important  choices  must  be  made 
beforehand,  to  set  some  conditions  for  the  census.  In 
coastal  areas,  the  most  important  decision  concerns 
tidal  conditions.  Birds  can  be  counted  at  high  or  low 
tides,  at  specified  time  intervals  before  or  after  high 
or  low  tides,  or  at  particular  tide  levels  on  either 
rising  or  falling  tides.  Setting  these  conditions  is 
central  to  intertidal  shorebird  censusing  because 
shorebirds  move  between  habitat  areas  on  a  tidal 
schedule,  and  the  area  of  available  habitat  changes 
on  a  tidal  schedule  (Connors  et  al.  1981). 

The  choice  of  conditions  should  be  coordinated 
with  the  census  method  chosen.  For  foraging  birds 
on  mudflats,  a  falling  tide  is  generally  best,  because 
birds  usually  forage  most  intensively  on  falling  tides, 
having  been  prohibited  from  access  to  feeding  areas 
during  the  previous  high  tide.  Selecting  a  particular 
interval  of  tide  levels  permits  the  observer  to  locate 
census  plots  conveniently.  If  very  low  tide  levels 
expose  extensive  mudflats,  birds  may  forage  too  far 
from  convenient  observation  points  along  shorelines; 
a  mid-tide  level  may  be  more  practical  (Page  et  al. 
1979).  For  censusing  birds  on  beaches  or  at  roosting 
sites,  a  very  high  tide  level  may  be  best,  because 
this  forces  birds  out  of  foraging  areas  on  nearby 
mudflats  (Connors  et  al.  1981).  In  nontidal  areas, 
time  of  day  assumes  greater  importance,  and  foraging 
birds  should  usually  be  censused  in  early  morning 
or  late  afternoon  when  most  are  active. 

Frequency  and  duration  of  censusing  also  must 
be  determined,  because  migrant  shorebird  popula- 
tions may  change  quickly.  To  determine  peak  migra- 
tion densities  or  timing  of  migration,  multiple  counts 
are  needed,  spaced  every  1  to  5  days.  Single  counts 
or  longer  intervals  between  counts  will  be  useful, 
but  could  miss  peak  movements  of  short  duration. 

Method  1 — Sample  Plot  Count.  In  the  sample 
plot  technique,  an  observer  counts  all  birds  within 
a  specified  area.  The  observer  can  select  areas  with 
natural  topographic  boundaries,  mark  plots  with 
stakes,  or  count  a  circular  plot  with  a  specified  ra- 
dius around  the  observing  station.  This  last  scheme, 
also  called  a  fixed-distance  point  count,  is  most  use- 
ful when  an  area  will  be  censused  only  once,  be- 
cause it  permits  sampling  many  points  without 
mapping  or  staking.  However,  it  requires  the  ob- 
server to  walk  through  the  open  shorebird  habitat, 
potentially  disturbing  the  birds,  and  its  accuracy 
depends  on  the  observer's  ability  to  estimate  the 
specified  radial  distance. 

Choosing  sample  plots  with  permanently 
marked  boundaries  avoids  this  last  problem  and,  in 
many  instances,  the  observer  can  count  the  entire 
plot  from  one  observing  station,  frequently  posi- 
tioned on  a  plot  boundary  (which  may  be  the  shore- 
line ).  Because  of  the  open  character  and  lack  of 
vegetation  or  topographic  diversity  in  many  migrant 


shorebird  habitats,  an  observer  with  binoculars  or 
a  telescope  can  count  and  identify  birds  at  great 
distances  (to  200  m  [660  ft])  under  good  viewing 
conditions.  This  permits  large  sample  plot  areas  to 
be  covered  from  a  single  station  or  from  a  series  of 
stations  along  a  shoreline.  This  method  is  well-suited 
to  regular  repetitive  censusing.  Densities  are  deter- 
mined by  field  measurement  or  map  estimation  of 
sample  plot  areas. 

Accuracy  of  the  sample  plot  method  is  generally 
high,  limited  mainly  by  difficulties  in  counting  birds 
that  may  move  appreciably  during  the  brief  count 
period.  Poor  visibility,  estimates  of  large  flocks,  diffi- 
culties with  identifications  at  long  distances,  and 
parallax  problems  of  estimating  plot  boundary  loca- 
tions may  also  affect  accuracy.  As  with  all  methods 
listed  here,  equipment  is  minimal,  training  is  easy  if 
observers  have  mastered  shorebird  identification,  and 
cost  depends  mainly  on  the  number  and  frequency 
of  sample  plot  counts. 

Method  2 — Marked  Plot  Strip  Transect.  This 
technique  is  essentially  identical  to  Method  3  for 
tundra-breeding  shorebirds.  It  is  especially  well- 
suited  to  marshy  areas  or  wetlands  with  open  pools 
and  mudflats,  interspersed  with  marsh,  because  the 
observer  moves  through  the  transect  close  enough 
to  all  points  to  discover  birds  partially  obscured 
by  vegetation.  The  division  of  the  transect  into  small 
plots  permits  association  of  birds  with  habitat  infor- 
mation for  each  plot,  an  advantage  in  some  studies. 
In  open  mudflat  habitats,  the  excellent  visibility  al- 
lows the  observer  to  walk  directly  along  the  center 
line  of  a  100  m  (330  ft)  wide  transect,  accelerating 
the  census.  As  with  the  preceding  method,  accuracy 
is  high  except  when  birds  move  on  or  off  the  tran- 
sect during  the  count.  The  time  required  to  mark 
the  transect  is  a  good  investment  only  when  a  series 
of  censuses  are  planned  at  the  same  site,  such  as 
censuses  throughout  a  migration  or  winter  period. 

Method  3 — Unmarked  Strip  Transect.  In  this 
variant  of  Method  2,  with  no  permanent  transect 
markers,  the  observer  must  estimate  a  census  width 
to  either  side  of  the  transect  line,  a  factor  that  re- 
duces the  accuracy  and  precision  of  the  census. 
However,  the  technique  requires  no  setup  time,  and 
is  useful  for  surveying  areas  when  only  single  cen- 
suses are  possible. 

Method  4 — Linear  Density  Transects.  Several 
variations  of  this  technique  all  obtain  a  linear  density 
estimate  along  shorelines,  rather  than  the  conven- 
tional density  measured  over  two-dimensional  areas. 
The  observer  moves  parallel  to  the  shoreline,  walk- 
ing or  riding  a  vehicle  (three-wheeled  motorcycles 
are  effective  on  arctic  beaches),  counting  birds  con- 
tinually. The  counting  width  may  be  constant,  cen- 
tered on  the  water's  edge  or  landward  from  the 
water's  edge,  or  it  may  be  variable,  including  for 


562 


Marsh  and  Shorebirds 


example,  all  of  a  beach  from  water  to  dunes  or  all  of 
a  specific  habitat,  such  as  the  wave-washed  zone  of 
a  sandy  beach  or  the  rocky  intertidal  zone  below 
a  cliff.  Densities  are  typically  expressed  as  birds  per 
kilometer  (0.6  mile).  Linear  density  transects  share 
the  same  properties  of  accuracy,  precision,  and  re- 
quired training  as  comparable  area  transect  methods. 

Method  5— Roosting  Flock  Counts.  If  con- 
sistent roosting  areas  are  known  from  prior  observa- 
tions, specific  areas  can  be  treated  as  sample  plots, 
or  general  areas  can  be  searched  for  the  daily  loca- 
tion of  total  roosting  flocks.  This  method  clearly 
depends  on  properly  choosing  sampling  times,  usu- 
ally dusk  or  midday  for  inland  areas  and  high  tide  for 
coastal  areas.  It  may  provide  an  absolute  density  for 
a  roosting  area,  but  the  real  values  of  such  a  count 
are  usually  ( 1 )  as  an  index  to  the  population  using 
nearby  areas  for  foraging  or  (  2  )  in  some  circumstan- 
ces, as  an  estimate  of  the  total  local  population  size. 
In  counting  roosting  flocks,  accuracy  and  precision 
are  sometimes  reduced  by  the  inability  to  visually 
separate  and  count  all  individuals  in  very  large  flocks 
of  small  birds.  A  vantage  point  providing  some  eleva- 
tion above  the  roosting  flock  is  helpful  in  these 
situations. 

Method  6 — Total  Population  Counts.  Be- 
cause migrating  shorebirds  are  so  concentrated  in 
specific,  limited  habitats,  total  population  size  esti- 
mates for  a  local  area  may  be  obtainable.  This  is 
unusual  compared  with  most  avian  censusing  situa- 
tions and  is  often  the  most  useful  of  population  vari- 
ables. This  technique  also  requires  prior  knowledge 
of  the  local  distributions  of  shorebird  habitats  or  the 
behavior  of  shorebird  species,  and  usually  more  than 
one  observer  censusing  different  areas  simultane- 
ously. It  has  been  used  to  estimate  total  populations 
of  a  single  estuary  (Page  et  al.  1979;  Connors  et  al. 
1981 )  or  of  hundreds  of  kilometers  of  a  coastline 
(J. P.  Myers,  pers.  commun.).  This  technique  is  costly 
in  personnel  and  its  accuracy  depends  on  the  physi- 
cal layout  of  the  area  censused  and  on  movements  of 
birds  during  the  census  period.  However,  the  num- 
bers obtained  are  frequently  the  most  useful  in  man- 
agement studies. 

Method  7 — Aerial  Counts.  As  discussed 
above,  transect  counts  from  aircraft  are  not  usually 
accurate  for  shorebirds  because  of  shorebirds'  small 
size  and  their  frequent  occurrence  in  large  flocks 
of  several  species.  For  cranes,  however,  the  large  size 
of  the  birds  makes  this  approach  feasible  and  effi- 
cient when  large  areas  must  be  sampled. 

Method  8 — Photographic  Counts.  For  very 
large  numbers  of  roosting  birds,  especially  shore- 
birds  or  sandhill  cranes  in  migration,  photographs 
taken  from  aircraft  or  from  an  elevated  fixed  point 
near  a  roost  may  provide  the  best  means  of  obtain- 
ing an  accurate  count.  A  photograph  can  be  taken 


quickly,  before  birds  flush  or  while  they  are  flying 
but  still  concentrated.  It  can  then  be  analyzed  me- 
thodically and  carefully  at  a  later  time. 


Discussion 

Shorebirds  in  migration  occupy  such  a  wide 
range  of  habitats  that  profitable  generalizations  are 
difficult.  However,  in  most  migrant  shorebird  situa- 
tions, the  open  nature  of  the  habitat  and  the  conspic- 
uousness  and  high  densities  of  the  birds  make 
shorebird  censusing  relatively  simple.  In  these  situa- 
tions, monitoring  shorebird  densities  directly  is  usu- 
ally preferable  to  measuring  habitat  features  as  a 
means  of  evaluating  probable  shorebird  use.  For  the 
same  reasons,  counting  actual  shorebird  totals  in 
monitored  areas  can  often  give  absolute  density  esti- 
mates at  little  extra  cost  over  simple  presence/ab- 
sence or  relative  density  estimates.  However, 
interpreting  census  results,  in  light  of  information  on 
local  habitat  distributions  and  bird  movements,  is 
more  centrally  important  than  for  most  other  bird- 
habitat  groups,  as  is  the  proper  choice  of  census 
conditions,  especially  with  tidal  fluctuations.  Because 
of  the  disjunct  distribution  of  most  migrant  and  win- 
tering shorebird  habitats,  estimates  of  total  size  of 
temporarily  discrete  local  populations  are  frequently 
possible  and  may  be  more  meaningful  than  estimates 
of  bird  densities  in  particular  habitats  under  particu- 
lar conditions. 

Finally,  dealing  with  populations  in  migration  at 
a  site  carries  certain  complications  that  differ  from 
breeding  population  studies.  Most  notable  is  the 
inherent  instability  of  the  population,  which  requires 
frequent  regular  censuses  to  estimate  peak  or  mean 
densities  during  a  migration  period.  Furthermore, 
unless  rates  of  individual  turnover  at  a  site  are 
known,  single  census  results  or  even  continuous 
census  results  do  not  provide  estimates  of  total  pop- 
ulation numbers  using  the  site — a  relationship  that 
is  markedly  different  from  that  which  applies  to 
breeding  bird  studies. 


MARSHES— RAILS  AND  COOTS 

Five  of  the  six  North  American  rail  species 
breed  and  winter  in  the  western  U.S.  where  they 
constitute  a  characteristic  segment  of  the  avifauna  of 
fresh,  brackish,  and  saltmarshes.  They  are  secretive 
and  inconspicuous,  spending  most  of  their  time  in 
wetlands  surrounded  by  dense  vegetative  cover, 
a  situation  that  has  led  to  the  development  of  spe- 
cialized census  techniques.  Coots  and  gallinules  nest 
in  some  of  the  same  marshes,  and  coots  in  winter 
are  widespread  in  open  water  situations,  both  fresh- 
and  saltwater,  where  waterfowl  are  common. 


Marsh  and  Shorebirds 


363 


Habitat  Features 

Rails  are  among  the  strictest  of  all  bird  groups 
in  their  consistency  of  habitat  use.  They  seldom  ven- 
ture out  of  marshes  or  the  muddy  borders  of 
marshes.  This  basic  requirement  of  vegetated  wet- 
lands provides  the  best,  albeit  rough,  guide  to  assess- 
ing rail  habitat.  Within  this  general  description, 
however,  suitable  marshlands  vary  from  freshwater 
to  brackish  to  saltwater;  from  edges  of  ponds,  lakes, 
or  rivers  to  large  estuaries;  and  from  moderately  low 
vegetation,  such  as  pickleweed  (Salicornia  sp.),  to 
tall  cattails  (Typha  sp.),  sedges  (Scirpus  sp.),  and 
rushes  (/uncus  sp.).  Species  habitat  preferences  and 
seasonal  habitat  use  patterns  differ  (Repking  and 
Ohmart  1977;  Gill  1979;  Glahn  1974;  Griese  et  al. 
1980;  Sayre  and  Rundle  1984).  Even  within  a  single 
species  and  a  single  region,  questions  remain  con- 
cerning the  differences  in  rail  population  densities 
between  superficially  similar  marshes,  as  with  popu- 
lations of  clapper  rail  (Rallus  longirostris)  in  San 
Francisco  Bay  (Gill  1979;  T.  Harvey,  pers.  commun.). 

Coots  are  abundant  and  fairly  conspicuous  at 
wetland  nesting  sites  in  western  North  America, 
where  they  use  a  wide  variety  of  marsh  and  pond 


types.  Habitat  requirements,  in  terms  of  pond  size, 
permanency,  and  characteristics  of  emergent  vegeta- 
tion at  nest  sites,  have  been  studied  by  Sugden 
(1979)  and  Nudds  ( 1982). 

Population  Measurement  Techniques 

Presence.  Because  of  the  difficulty  of  observing 
secretive  rails  in  dense  cover,  presence  and  absence 
is  sometimes  the  only  information  one  can  obtain. 
There  are  no  special  techniques  for  determining 
presence,  however,  other  than  direct  observation  of 
individuals,  usually  during  foraging  periods,  or  calls 
and  observations  recorded  by  the  methods  listed 
under  Relative  or  Absolute  Density. 

Relative  Density. 


Method  1 — Breeding  Vocalization.  The 

breeding  vocalization  technique  relies  on  naturally 
occurring  vocalizations  of  territorial  breeding  rails 
during  the  early  breeding  season  (April,  May,  and 
June  in  most  areas).  An  observer  familiar  with  rail 
calls  listens  at  selected  points  within  or  on  the 
periphery  of  a  marsh.  Listening  periods  should  be  of 


Coots  in  an  open-water  habitat. 


364 


Marsh  and  Shorebirds 


the  same  duration  and  similar  in  time  of  day,  date, 
and  weather  conditions  at  all  sites  to  be  compared. 
Relative  density  can  be  expressed  as  calling  birds  per 
time  of  observation  or  per  listening  station. 
Recordings  or  rail  vocalizations  are  available  for 
training  purposes.  Precision  and  accuracy  of  this 
technique  are  unknown,  but  may  be  low.  Costs  are 
also  low. 


recorded  within  the  local  region  of  the  species' 
range,  is  recommended  (Johnson  et  al.  1981 ).  A  con- 
sistent schedule  of  tape  playback  duration  and  listen- 
ing duration  must  be  maintained,  on  the  order  of  1 
min.  of  playback  followed  by  5  min.  of  listening, 
possibly  repeated,  at  each  station.  Training  and  data 
treatment  are  similar  to  Method  1. 


Method  2 — Tape  Playback  Response.  The 

tape  playback  technique  is  more  widely  used,  re- 
quires little  extra  effort,  and  increases  the  chance  of 
hearing  a  rail  that  is  present  during  the  station 
count.  Glahn  (1974)  located  71%  more  territories  of 
rails  in  Colorado  using  playback  techniques,  com- 
pared with  territory  mapping  without  playback.  Ac- 
curacy is  therefore  increased  but  may  still  be  low, 
depending  mainly  on  the  consistency  with  which 
rails  respond. 

Although  taped  response  techniques  are  most 
effective  during  the  breeding  season,  they  also  have 
been  used  with  appropriate  taped  calls  to  assess 
winter  populations  (Tomlinson  and  Todd  1973;  Mar- 
ion et  al.  1981 ).  Use  of  a  variety  of  calls,  preferably 


Method  3 — Flood  Tide  Counts.  In  tidal  areas, 
some  rails  can  be  located  visually  during  extremely 
high  tides,  when  almost  all  marsh  vegetation  is  inun- 
dated. In  San  Francisco  Bay,  clapper  rails  are  forced 
onto  small  patches  of  high  ground  or  floating  debris 
during  the  highest  tides  of  the  year,  which  occur 
during  daylight  only  in  winter.  They  can  be  located 
from  vantage  points  surrounding  the  marsh  or  in 
extensive  marsh  areas  from  a  canoe  or  an  airboat 
traveling  through  the  marsh.  Detection  of  birds  can 
be  increased  if  the  observer  works  with  a  competent 
hunting  dog  (T.  Harvey,  pers.  commun. ). 

With  sufficiently  high  tides  and  scarce,  accessi- 
ble refuges,  this  method  can  be  accurate  in  terms  of 
total  birds  sighted,  but  the  distance  over  which  birds 
have  traveled  to  a  refuge  is  generally  unknown.  For 


Gallinule  in  a  ryp/ca/  marsh  habitat. 


Marsh  and  Shorebirds 


365 


this  reason  it  is  listed  first  as  a  relative  density  tech- 
nique, but  this  technique  and  the  two  previous  tech- 
niques can  all  be  adapted  to  absolute  density 
estimation. 

Method  4 — Open  Water  Counts  (Coots). 

American  coots  are  the  only  species  in  this  group  to 
regularly  forage  in  open  water  away  from  emergent 
vegetation.  Direct  visual  counts,  made  from  shore  or 
in  boats,  can  provide  an  index  of  numbers  of  coots 
nesting  in  wetlands.  The  fraction  of  the  population 
missed  will  remain  unknown  and  may  vary  in  differ- 
ent types  of  wetlands.  All  counts  of  comparative 
purposes  should  be  made  .at  the  same  time  of  day 
and  over  as  short  a  calendar  interval  as  practical. 


Absolute  Density. 

Method  1 — Breeding  Vocalization.  Combin- 
ing the  breeding  vocalization  technique  with  a  terri- 
tory mapping  effort,  a  variable  width  line  transect 
method,  or  a  variable  circular  plot  method  permits 
estimation  of  absolute  densities.  Accuracy  is  proba- 
bly low  in  most  situations,  however,  because  an  un- 
known percentage  of  rails  present  will  not  call 
during  the  census  period.  All  three  approaches  re- 
quire the  observer  to  estimate  distances  to  unseen 
vocalizing  rails,  a  difficult  task. 


Method  2 — Tape  Playback  Response.  Simi- 
larly, the  tape  playback  method  is  the  combination 
of  the  technique  discussed  under  Relative  Density 
and  the  area  estimating  techniques  of  mapping,  varia- 
ble width  line  transect,  or  variable  circular  plot 
method.  It  suffers  the  same  limitations  of  accuracy 
and  precision  as  Method  1  in  this  section,  but  the 
tape  playback  elicits  vocalizations  at  a  higher  fre- 
quency than  the  natural  calling  rates,  so  fewer  birds 
are  missed. 


Method  3 — Flood  Tide  Counts.  Methods  for 
this  technique  are  described  under  Relative  Density. 
However,  when  discrete  patches  of  saltmarsh  or 
extensive  areas  of  continuous  saltmarsh  can  be 
searched  thoroughly,  absolute  density  measurements 
result.  Local  movements  of  birds  to  refuges  within 
their  population  area  do  not  distort  the  density  esti- 
mate if  the  entire  population  area  is  censused.  (Tran- 
sect sampling,  under  these  circumstances,  may 
produce  large  errors. )  If  all  potential  refuges  can  be 
carefully  searched,  accuracy  and  precision  of  these 
estimates  should  be  high.  Costs  may  also  be  high, 
however,  because  of  logistics,  especially  with  the  use 
of  airboats  at  high  tides.  This  method  is  only  useful 
in  winter,  when  high  tides  occur  during  daylight 
hours. 


Method  4 — Rope-Drag  Nest  Mapping.  As 

described  for  Tundra — Breeding  Shorebirds,  two  ob- 
servers walk  through  potential  nesting  habitat  drag- 
ging a  rope  (about  50  m  [165  ft]  of  1/4-in.  rope) 
between  them;  a  third  observer  walks  behind,  re- 
cording locations  of  nests  and  adults  when  birds 
flush.  The  technique  greatly  increases  the  numbers 
of  nests  located  and  is  especially  valuable  when  used 
in  conjunction  with  Methods  1  and  2,  because  it 
provides  an  independent  estimate  of  breeding  den- 
sity. Furthermore,  information  from  one  method  can 
sometimes  be  added  to  information  from  the  other 
when  clear  discrepancies  occur.  The  combined  cen- 
sus may  be  more  accurate  than  either  census  alone. 


Method  5 — Nest  Searches  (Coots).  Direct 
nest  searches  in  marshes  can  provide  reliable  esti- 
mates of  coot  nest  densities  (Sugden  1979).  This  can 
be  combined  with  observations  of  bird-nesting  be- 
havior and  territoriality  to  narrow  the  search  areas, 
but  will  usually  require  a  time-consuming  field  effort. 


Discussion 

Rails  in  marshes  share  one  characteristic  with 
shorebirds  in  migration.  Their  habitats  are  often 
clearly  distinguished  from  surrounding  areas  (marsh 
or  mudflat  versus  uplands,  for  example)  and  are  eas- 
ily recognized  and  mapped.  Habitat  evaluation  at  this 
very  general  level  is  straightforward.  At  the  finer 
scale,  which  requires  distinguishing  better  from 
poorer  marshlands,  knowledge  is  limited. 

The  two  bird  groups  contrast  markedly  in  ease 
and  accuracy  of  censusing  populations  for  absolute 
density  estimates.  For  all  the  reasons  that  shorebirds 
are  extremely  tractable,  rails  are  difficult  to  treat. 
Their  secretiveness  and  the  densely  vegetated,  rela- 
tively inaccessible  habitats  they  favor  impose  stum- 
bling blocks  in  any  rail  population  studies.  The 
methods  just  described  have  limited  accuracy  and 
precision,  but  provide  useful  estimates  of  population 
sizes  in  most  situations. 


CONCLUSIONS 

The  54  species  of  marsh  and  shorebirds  that 
occur  in  Alaska  and  the  western  States  cover  such  a 
wide  range  of  sizes,  habitats,  and  natural  histories 
that  the  difficulty  of  any  neat  generalizations  of  bird- 
habitat  relationships  or  of  preferred  census  tech- 
niques is  not  surprising.  Most  of  these  species  are 
relatively  easy  to  assign  to  large-scale  habitat  types 
(marsh  versus  forest,  for  example)  and  some  can  be 
partially  defined  by  minor  habitat  differences  in  habi- 
tat features,  such  as  vegetation  height  or  water 
depth.  In  almost  no  situations  beyond  a  local  study 


366 


Marsh  and  Shorebirds 


area,  however,  can  one  predict  densities  of  bird  use 
based  on  habitat  measurements  with  a  sufficient  de- 
gree of  accuracy  to  justify  such  an  approach.  For 
most  shorebirds  and  cranes,  this  is  not  a  serious  lack, 
because  these  species  are  usually  so  visible  in  their 
breeding  and  non-breeding  habitats  that  a  direct 
censusing  method  can  be  efficient  and  acceptably 
accurate.  Rails,  by  their  secretive  habits  and  the 
dense  vegetative  cover  in  their  habitats,  present  a 
much  more  difficult  censusing  challenge. 

For  these  reasons,  presence  and  absence  tech- 
niques are  usually  warranted  only  for  rails  and  gallin- 
ules.  Relative  density  census  techniques  have  much 
wider  application  for  many  shorebirds,  rails,  coots, 


and  cranes.  With  an  increase  in  effort  and  minor 
modifications,  many  of  these  techniques  provide  ab- 
solute density  measurements.  The  range  of  absolute 
density  census  techniques  is  especially  large  for 
shorebirds  nesting  on  tundra  or  migrating  along 
shorelines,  because  in  these  situations  a  large  mix- 
ture of  species  often  occurs  at  high  densities  under 
conditions  of  good  detectability.  Even  these  generali- 
zations, however,  gloss  over  individual  species  differ- 
ences that  may  be  of  major  importance  in  selecting 
research  and  management  techniques.  The  Literature 
Cited  section  in  this  chapter  should  be  consulted 
directly  for  additional  details  concerning  particular 
species,  areas,  or  habitats  before  field  work  is 
initiated. 


Marsh  and  Shorebirds 


567 


LITERATURE  CITED 


BENT,  A.C.  1926.  Life  histories  of  North  American  marsh 
birds.  U.S.  Natl.  Mus.  Bull.  135. 

.  1929.  Life  histories  of  North  American  shorebirds. 

U.S.  Natl.  Mus.  Bull.  146. 

BERGMAN,  R.D.,  R.  HOWARD,  K  ABRAHAM,  and  M. 
WELLER.  1977.  Water  birds  and  their  wetland  re- 
sources in  relation  to  oil  development  at  Storkersen 
Point,  Alaska.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv. 
Resour.  Publ.  129.  39pp. 

BRYANT,  D.M.  1979.  Effects  of  prey  density  and  site  char- 
acter on  estuary  usage  by  overwintering  waders 
(Charadrii).  Estuarine  and  Coastal  Marine  Science  9: 
369-384. 

CONNORS,  P.G.  1984.  Ecology  of  shorebirds  in  the  Alas- 
kan Beaufort  littoral  zone.  Pages  403-416  in  Barnes, 
P.,  E.  Reimnitz,  and  D.  Schell,  eds.  The  Alaskan  Beau- 
fort Sea:  Ecosystems  and  Environments.  Academic 
Press. 

,  J.P.  MYERS,  C.S.W.  CONNORS,  and  FA.  PITELKA. 

1981.  Interhabitat  movements  by  sanderlings  in  rela- 
tion to  foraging  profitability  and  the  tidal  cycle.  Auk 
98:49-64. 

, ,  and  FA.  PITELKA.  1979.  Seasonal  habitat 


use  by  arctic  Alaskan  shorebirds.  Studies  Avian  Biol. 
2:101-111. 

DREWIEN,  R.C.  1973.  Ecology  of  Rocky  Mountain  greater 
sandhill  cranes.  Ph.D.  dissertation,  Univ.  Idaho,  Mos- 
cow. 152pp. 

EMLEN,  J.T.  1971.  Population  densities  of  birds  derived 
from  transect  counts.  Auk  88:323-342. 

.  1977.  Estimating  breeding  season  bird  densities 

from  transect  counts.  Auk  94:455-468. 

GERSTENBERG,  R.H.  1979.  Habitat  utilization  by  winter- 
ing and  migrating  shorebirds  on  Humboldt  Bay,  Cali- 
fornia. Studies  Avian  Biol.  2:33-40. 

GIBSON,  F.  1971.  The  breeding  biology  of  the  American 
avocet  (Recurvirostra  americana)  in  central  Oregon. 
Condor  73:444-454. 

GILL,  R.E.,  Jr.  1979.  Status  and  distribution  of  the  Califor- 
nia clapper  rail  (Rallus  longirostris  obsoletiis).  Cali- 
fornia Fish  and  Game  65:36-49- 

and  CM.  HANDEL.  1981.  Shorebirds  of  the  eastern 

Bering  Sea,  Pages  719-738  in  Hood,  D.W.  and  J.A. 
Calder,  eds.  Eastern  Bering  Sea  Shelf:  Oceanography 
and  Resources,  Vol.  2.  Univ.  Washington  Press,  Seattle. 
-,  MR.  PETERSEN,  and  P.D.  JORGENSEN.  1981.  Birds 


of  the  north  Alaska  Peninsula,  1976-1980.  Arctic 

34:286-306. 
GLAHN,  J.F.  1974.  Study  of  breeding  rails  with  recorded 

calls  in  north-central  Colorado.  Wilson  Bull.  86:206- 

214. 
GOSS-CUSTARD,  J.D.  1970.  The  response  of  redshank 

(Tringa  totanus  L.)  to  spatial  variations  in  the  density 

of  their  prey.  J.  Animal  Ecol.  39:91-1 13. 
GRAUL,  WD.  1975.  Breeding  biology  of  the  mountain 

plover.  Wilson  Bull.  87:6-31. 
GRIESE,  H.J.,  R.A.  RYDER,  and  C.E.  BRAUN.  1980.  Spatial 

and  temporal  distribution  of  rails  in  Colorado.  Wilson 

Bull.  92:96-102. 
HAMILTON,  R.C.  1975.  Comparative  behavior  of  the 

American  avocet  and  the  black-necked  stilt  (Recurvi- 

rostridae).  A.O.U.  Monographs,  17. 


HOLMES,  R.T.  and  C.P.  BLACK.  1973.  Ecological  distribu- 
tion of  birds  in  the  Kolomak  River-Askinuk  Mountain 
Region,  Yukon-Kuskokwim  Delta,  Alaska.  Condor 
75:150-163. 

and  FA.  PITELKA.  1 968.  Food  overlap  among  coex- 
isting sandpipers  on  northern  Alaska  tundra.  System- 
atic Zoology  17:305-318. 

ISLEIB,  M.E.P.  1979.  Migratory  shorebird  populations  on 
the  Copper  River  Delta  and  eastern  Prince  William 
Sound,  Alaska.  Studies  Avian  Biol.  2:125-130. 
JARVINEN,  O.  and  R.A.  VAISANEN.  1975.  Estimating  rela- 
tive densities  of  breeding  birds  by  the  line  transect 
method.  Oikos  26:316-322. 
JENNI,  DA.,  R.L.  REDMOND,  and  IK.  BICAK.  1982.  Be- 
havioral ecology  and  habitat  relationships  of  long- 
billed  curlews  in  western  Idaho.  Unpubl.  Rep.  to  U.S. 
Dep.  Inter.,  Bur.  Land  Manage.,  Boise,  ID.  234pp. 
JOHNSON,  R.R.,  B.T.  BROWN,  L.T.  HAIGHT,  and  J.M. 
SIMPSON.  1981.  Playback  recordings  as  a  special 
avian  censusing  technique.  Studies  Avian  Biol.  6:68- 
75. 
JOHNSON,  S.R.  and  W.J.  RICHARDSON.  1981.  Beaufort 
Sea — Barrier  Island —  lagoon  ecological  processes 
studies:  Final  report,  Simpson  Lagoon:  birds.  Pages 
109-383  in  Environmental  Assessment  Alaskan  Conti- 
nental Shelf.  Vol.  7. 

KRAPU,  G.L.  1974.  Feeding  ecology  of  pintail  hens  during 
reproduction.  Auk  91:278-290. 

LATKA,  DC.  and  J.W.  YAHNKE.  1984.  Simulating  roosting 
habitat  of  sandhill  cranes  and  validation  of  suitability - 
of-use  indices.  Unpubl.  abstract  of  poster  presented 
at  Wildlife  2000  Conference,  Fallen  Leaf  Lake,  CA. 

LITTLEFIELD,  CD.  and  R.A.  RYDER.  1968.  Breeding  biol- 
ogy of  the  greater  sandhill  crane  on  Malheur  National 
Wildlife  Refuge,  Oregon.  Trans.  North  Am.  Wildl. 
Nat.  Resour.  Conf.  33:444-454. 

MARION,  W.R.,  T.E.  O'MEARA,  and  D.S.  MAEHR.  1981.  Use 
of  playback  recordings  in  sampling  elusive  or  secre- 
tive birds.  Studies  Avian  Biol.  6:81-85. 

MARTIN,  P.D.  and  C.S.  MOITERET.  1981.  Bird  populations 
and  habitat  use,  Canning  River  Delta,  Alaska.  Unpubl. 
Rep.  to  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.,  Fair- 
banks, AK  196pp. 

MICKELSON,  PS.,  J.S.  HAWKINS,  DR.  HERTER,  and  S.M. 
MURPHY.  1980.  Habitat  use  by  birds  and  other  wild- 
life on  the  Eastern  Copper  River  Delta,  AK.  Unpub- 
lished Report.  Alaska  Coop.  Wildl.  Res.  Unit,  Univ. 
Alaska,  Fairbanks.  189pp. 

MYERS,  J  P.,  P.G.  CONNORS,  and  FA.  PITELKA.  1979. 

Territory  size  in  wintering  sanderlings:  The  efifects  of 
prey  abundance  and  intruder  density.  Auk  96:551- 
561. 

and  FA.  PITELKA.  1980.  Effect  of  habitat  condi- 
tions on  spatial  parameters  of  shorebird  populations. 
Rep.  to  the  U.S.  Dep.  Energy.  82pp. 

NUDDS,  T.D.  1982.  Ecological  separation  of  grebes  and 

coots:  Interference  competition  or  microhabitat  selec- 
tion? Wilson  Bull.  94:505-514. 

ORING,  L.W.,  D.B.  LANK  and  S.J.  MAXSON.  1983.  Popula- 
tion studies  of  the  polyandrous  spotted  sandpiper. 
Auk  100:272-285. 

PAGE,  G.W.,  L.E.  STENZEL,  D.W.  WINKLER,  and  C.W. 

SWARTH.  1983.  Spacing  out  at  Mono  Lake:  Breeding 
success,  nest  density,  and  predation  in  the  snowy 
plover.  Auk  100:13-24. 

, ,  and  CM.  WOLFE.  1979.  Aspects  of  the 

occurrence  of  shorebirds  on  a  central  California  estu- 


368 


Marsh  and  Shorebirds 


ary.  Studies  Avian  Biol.  2:15-32. 

RALPH,  C.J.  and  J.M.  SCOTT,  eds.  1981.  Estimating  num- 
bers of  terrestrial  birds.  Studies  Avian  Biol.  630pp. 

RANDS,  M.R.W.  and  J.P.  BARKHAM.  1981.  Factors  control- 
ling within-flock  feeding  densities  in  three  species  of 
wading  bird.  Ornis  Scandinavica  12:28-36. 

RECHER,  H.F.  1966.  Some  aspects  of  the  ecology  of  mi- 
grant shorebirds.  Ecology  47:393-407. 

REDMOND,  R.L.  1984.  The  behavioral  ecology  of  long- 
billed  curlews  (Numenius  americanm)  breeding  in 
western  Idaho.  Ph.D.  dissertation,  Univ.  Montana, 
Missoula. 

,  T  K  BICAK,  and  DA.  JENNI.  1981.  An  evaluation 

of  breeding  season  census  techniques  for  long-billed 
curlews  (Numenius  atnericanus).  Studies  Avian  Biol. 
6:197-201. 

RENAUD,  WE.  1979.  The  piping  plover  in  Saskatchewan. 
Blue  Jay  37:90-103. 

REPKING,  C.F.  and  R.D.  OHMART.  1977.  Distribution  and 
density  of  black  rail  populations  along  the  lower 
Colorado  River.  Condor  79:486-489. 

RILEY,  J. L.  1982.  Habitats  of  sandhill  cranes  in  the  south- 
ern Hudson  Bay  lowland,  Ontario.  Canada  Field-Nat. 
96:51-55. 

RYAN,  MR.  1985.  Marbled  godwit  habitat  selection  in  the 
northern  prairie  region.  J.  Wildl.  Manage.,  in  press. 

SAYRE,  M.W.  and  WD.  RUNDLE.  1984.  Comparison  of 
habitat  use  by  migrant  soras  and  Virginia  rails.  J. 
Wildl.  Manage.  48:599-605. 

SENNER,  S.E.  1979.  An  evaluation  of  the  Copper  River 
Delta  as  a  critical  habitat  for  migrating  shorebirds. 
Studies  Avian  Biol.  2:1 31- 146. 

SHD2LDS,  G.F.  and  L.J.  PEYTON.  1979.  Avian  community 
ecology  of  the  Akulik-Inglutalik  River  Delta,  Norton 


Bay-Alaska.  Environmental  Assessment,  Alaskan  Conti- 
nental Shelf,  Ann.  Reps.  5:608-710.  NOAA,  Boulder, 
CO. 

SPLNDLER,  MA.  1978.  Bird  populations  and  habitat  use  in 
the  Okpilak  River  Delta  area,  Arctic  National  Wildlife 
Range,  Alaska,  1978.  Unpubl.  Rep.  to  U.S.  Dep.  Inter., 
Fish  and  Wildl.  Serv.,  Fairbanks,  AK  86pp. 

STEWART,  RE.  and  H.A.  KANTRUD.  1971.  Classification  of 
natural  ponds  and  lakes  in  the  glaciated  prairie  re- 
gion. U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  Resour. 
Publ.  92.  57pp. 

SUGDEN,  L.G.  1979.  Habitat  use  by  nesting  American 

coots  in  Saskatchewan  parklands.  Wilson  Bull.  91:599- 
607. 

TOMLLNSON,  RE.  and  R.L.  TODD.  1973-  Distribution  of 
two  western  clapper  rail  races  as  determined  by 
responses  to  taped  calls.  Condor  75:177-183- 

TROY,  DM.,  D.  HERTER,  and  R.  BURGESS.  1983.  Prudhoe 
Bay  Waterflood  Project  tundra  bird  monitoring  pro- 
gram. Unpubl.  Rep.  to  Dep.  Army,  Corps  of  Engineers, 
Anchorage,  AK.  98pp. 

WEBER,  M.J.,  PA.  VOHS,  Jr.,  and  L.D.  FLAKE.  1982.  Use  of 
prairie  wetlands  by  selected  bird  species  in  South 
Dakota.  Wilson  Bull.  94:  550-554. 

WESELOH,  D.V.  and  L.M.  WESELOH.  1983  Numbers  and 
nest  site  characteristics  of  the  piping  plover  in  central 
Alberta,  1974-1977.  Blue  Jay  41:155-161. 

WIENS,  J.A.  1973.  Pattern  and  process  in  grassland  bird 
communities.  Ecol.  Monogr.  43:237-270. 

WILLIAMSON,  F.S.L.,  M.C.  THOMPSON,  and  J.Q.  HINES. 
1966.  Avifaunal  investigations,  Pages  437-480  in  Wili- 
movsky,  N.  J.,  ed.  Environment  of  the  Cape  Thompson 
Region,  Alaska.  U.S.  Atomic  Energy  Commission,  Oak 
Ridge,  TN. 


Marsh  and  Shorebirds 


369 


18 


WATERFOWL 


Robert  L.  Eng 

Montana  State  University 
Bozeman,  MT  59717-0001 


Editor's  Note:  Waterfowl,  because  of  their  economic 
and  recreational  value  have  received  much  atten- 
tion for  many  years.  And  because  of  their  obligate 
association  with  wetlands,  habitat  has  long  been  a 
concern  with  these  species.  Thus,  inventory,  moni- 
toring and  management  of  waterfowl  habitat  has 
a  long  history  compared  to  species  groups  whose 
management  has  focused  on  population  manage- 
ment. However,  waterfowl  biology  has  become  a 
specialized  discipline  and  has  tended  to  focus  on 
major  wetland  areas.  Yet  large  numbers  of  water- 
fowl depend  on  smaller  wetlands  scattered  through- 
out the  West.  These  areas,  if  managed  at  all,  are 
probably  managed  by  biologists  with  no  special- 
ized expertise  in  waterfowl.  This  chapter  provides 
such  biologists  an  introduction  to  inventory  and 
monitoring  of  waterfowl  and  waterfowl  habitat. 


INTRODUCTION 

The  term  waterfowl  includes  ducks,  geese,  and 
swans.  In  the  most  recently  published  classification,  a 
single  family,  the  Anatidae,  is  separated  into  two 
subfamilies,  the  Anserinae  (whistling-ducks,  swans, 
and  geese)  and  the  Anatinae  (ducks;  Check- List  of 
North  American  Birds,  American  Ornithologists 
Union,  1983).  Both  subfamilies  are  further  divided 
into  species,  groups,  or  tribes — the  Anserinae,  into 
three  and  the  Anatinae,  into  five. 

Most  species  of  waterfowl  are  migratory  to 
some  extent,  with  many  participating  in  lengthy  sea- 
sonal movements.  Thus  habitat  features  for  different 
species  or  species  groups  will  vary  considerably 
throughout  the  year,  both  as  a  result  of  geographic 
availability  and  by  demands  placed  upon  the  species 
as  a  result  of  seasonal  behavior. 

As  with  upland  game  birds,  most  early  water- 
fowl investigations  emphasized  the  reproductive 
season  (Chabreck  1979).  This  is  somewhat  under- 
standable because  it  covers  the  period  when  off- 
spring are  produced.  The  nesting  season  also 
embraces  the  time  of  the  year  when  waterfowl  habi- 
tat is  most  important,  because  self-imposed  isolation 
of  pairs  results  in  a  density  far  below  that  which  may 
be  found  during  other  seasons.  Also,  nesting  habitat 
seemingly  was  the  most  threatened  by  land  uses 
by  man  and  consequently  was  thrust  into  the 
limelight. 

In  recent  years,  attention  has  increased  on  win- 
tering and  migration  habitat.  Even  though  the  gregar- 
ious nature  of  waterfowl  during  these  periods 
permits  rather  large  numbers  in  relatively  small 
areas,  winter  habitats  in  parts  of  the  Atlantic  Flyway 
have  deteriorated  to  the  extent  that  they  are  no 
longer  suitable  for  ducks  (Addy  1964).  As  was 
pointed  out  by  Chabreck  (1979),  dabbling  ducks 
may  spend  two-thirds  of  the  year  on  the  winter  habi- 


Waterfowl 


371 


tat,  and  the  quality  of  that  habitat  may  greatly  influ- 
ence their  reproductive  performance  in  the  spring. 

The  habitat  features  for  the  various  tribes  of 
waterfowl  appear  more  distinguishable  for  the  breed- 
ing season  than  for  other  times  of  the  year.  This 
may  in  part  be  a  result  of  the  greater  investigative 
effort  during  the  breeding  season.  It  may  also  be 
a  result  of  less  diverse  activities  by  all  species  during 
winter  (largely  feeding  and  loafing),  permitting  a 
more  intense  use  of  a  resource  by  many  members  of 
the  same  or  different  species.  Thus,  in  this  chapter, 
the  habitat  features  for  the  nesting  season  are  dis- 
cussed by  species  groups  for  the  more  widely  dis- 
tributed or  numerically  important  tribes  with 
references  to  individual  species  when  deemed  perti- 
nent. Winter  habitat  is  discussed  in  a  more  general 
fashion  for  the  entire  family. 


Trumpeter  swans,  adults,  and  cygnet. 


HABITAT  FEATURES  CORRELATED  WITH 
WATERFOWL 

Swans  (Cygnini) 

Two  species  of  swans  are  native  to  North  Amer- 
ica. The  tundra  swan  (Cygnns  columbianus),  until 
recently  called  the  whistling  swan,  is  the  most  abun- 
dant; about  150,000  were  reported  in  1971  (Bellrose 
1978).  The  trumpeter  swan  (C  buccinator),  once 
believed  to  number  less  than  100  birds  in  the  U.S. 
and  Canada,  now  constitutes  a  population  of  about 
1,500  plus  about  9,000  in  Alaska. 

Physical  Features.  As  the  name  implies,  tundra 
swans  nest  in  the  arctic  from  north  of  Hudson  Bay  to 
the  coast  of  Alaska  (Bellrose  1978).  Lensink  (in 
Bellrose  1978)  reported  nest  site  locations  as 
follows:  50%  on  the  shore  of  a  lake  or  pond  within 
18  m  (20  yd)  of  water,  30%  on  small  islands  or 
points,  and  the  remainder  in  a  variety  of  situations. 
In  most  instances  the  nest  was  elevated 
aboveground. 

Trumpeter  swans  in  Alaska  showed  a  nest  pref- 
erence for  small  beaver  impoundments  over  other 
larger  bodies  of  water;  one-half  of  the  40  nests  were 
located  in  beaver  ponds  of  less  than  5.6  ha  ( 14  a.). 
This  preference  appeared  to  be  related  to  a  stable 
water  level  and  little  wave  action,  which  promoted 
the  development  and  maintenance  of  extensive 
stands  of  emerging  and  floating  vegetation  in  which 
and  from  which  the  nests  were  constructed  (Hansen 
et  al.  1971).  Trumpeter  swans  in  the  Targhee  Forest 
in  Idaho  and  Wyoming  use  older,  more  eutrophic 
lakes,  which  are  relatively  shallow  (average  1.2  m  [4 
ft]),  have  a  large  part  of  the  surface  covered  with 
vegetation  (80%  ),  and  have  at  least  25%  of  their  to- 
tal area  less  than  1  m  deep.  Aquatic  vegetation  is 
used  for  the  nest  site  and  nest  material  (Maj  1983). 


Vegetation. 


Structure.  Whereas  the  trumpeter  swan  nests 
almost  exclusively  over  water  in  emergent  aquatic 
vegetation,  the  tundra  swan  seems  to  prefer  more 
upland  sites  not  associated  with  emergent  vegetation 
(Hansen  et  al.  1971),  but  close  enough  to  water 
that  water  was  often  exposed  when  vegetation  was 
removed  from  a  circle  around  the  nest  (Lensink 
in  Bellrose  1978).  Both  species  also  feed  largely  on 
aquatic  plants,  both  submergents  and  emergents, 
although  the  tundra  swan  will  feed  on  upland 
sources  when  marshes  freeze  (Nagel  1965).  Upland 
feeding  in  cornfields  by  tundra  swans  has  become 
relatively  common  along  Chesapeake  Bay  and  can  be 
readily  observed  during  spring  migration  through 
Montana. 

Species  Composition.  Tundra  swans,  as  inhab- 
itants of  the  open  tundra  and  somewhat  more  terres- 
trial than  trumpeter  swans,  construct  nests  from 
material  which  may  include  mosses,  grasses,  and 
sedges  (Carex  sp.).  In  Alaska,  nests  of  trumpeter 
swans  were  usually  constructed  with  the  stems  and 
rhizomes  of  Carex  sp.,  Equisetum  fluviatile,  or  Po- 
tentilla  sp.  (Hanson  et  al.  1971 ).  In  Idaho  and  Wyo- 
ming, nests  were  frequently  in  and  constructed  of 
Typha  sp.,  Scirpus  sp.,  or  Carex  sp.  Occasional  nests 
were  found  on  islands  or  beaver  (Castor  canaden- 
sis) lodges,  near  the  water's  edge,  and  constructed 
largely  of  aquatic  vegetation  (Maj  1983). 

Both  species  feed  largely  on  aquatic  plants,  in- 
cluding the  seeds,  stems,  and  tubers  of  several  spe- 
cies. Sago  pondweed  (Potamogeton  pectinatus)  is  a 
prominent  plant  in  the  diet  of  tundra  (Sherwood 
I960;  Nagel  1965)  and  trumpeter  swans  (Bellrose 
1978).  Other  plants  frequently  mentioned  as  food 
are  widgeon  grass  (Ruppia  sp. ),  duck  potato  (Sagit- 
taria  sp. ),  other  pondweeds  (Potamogeton  sp. ), 
and  water  milfoil  (Myriophyllum  sp. ). 


372 


Waterfowl 


Geese  (Anserini) 

Of  the  several  species  of  geese  that  winter  in 
the  U.S.,  only  one,  the  Canada  goose  (Branta  cana- 
densis), nests  within  the  U.S.  south  of  Canada.  This 
species  has  great  morphological  and  ecological 
breadth,  with  sizes  ranging  from  a  large  duck  to  a 
small  swan,  and  nesting  distribution  from  farm  ponds 
to  forested  lakes  to  arctic  tundra.  This  diversity  has 
led  to  considerable  discussion  as  to  the  number  of 
identifiable  races.  Bellrose  (1978)  illustrated  the 
ranges  of  1 2  populations  of  this  species,  established 
largely  on  the  basis  of  the  close  alliances  shown 
during  the  winter.  Two  races  nest  south  of  Canada, 
the  giant  Canada  goose  (B.  c.  maxima)  and  the 
western  Canada  goose  (B.  c.  moffltti),  whereas  at 
least  nine  other  races  nest  across  the  breadth  of  the 
arctic  or  northern  Canada  (Bellrose  1978). 


Physical  Features.  Nesting  habitat  for  Canada 
geese  is  extremely  diverse,  particularly  for  the  two 
races  nesting  south  of  Canada.  This  is  probably 
because  of  the  diverse  habitats  into  which  the  two 
races  have  been  introduced  through  management 
(transplant)  programs.  Canada  geese  have  been 
observed  to  nest  in  the  vicinity  of  lakes,  ponds, 
marsh,  and  river  habitat;  on  the  ground;  on  muskrat 
houses;  in  trees  in  unused  raptor,  heron,  and  magpie 
(Pica  pica)  nests;  and  on  cliffs.  However,  in  spite 
of  the  great  variety  of  reported  nest  sites,  the 
preference  for  islands  as  nest  sites  appears  to 
dominate  (Hammond  and  Mann  1956;  Ewaschuk  and 
Boag  1972;  Giroux  1981;  Giroux  et  al.  1983).  This 
preference  is  also  shown  by  some  of  the  smaller 
arctic  nesting  Canada  geese  (Maclnnes  1962),  but 
less  so  by  the  colonial  nesting  snow  geese  {Chen 
caerulescens;  Lemieux  1959)  and  Ross'  goose  (C 
rossii;  Ryder  1967). 


Terrestrial  nesting  Canada  geese. 


Geese  are  very  terrestrial  compared  with  most 
ducks  and  swans.  Consequently,  secure  brood-rear- 
ing areas  are  an  integral  part  of  breeding  habitat.  The 
habitat  usually  consists  of  a  meadow  or  flat  area  pro- 
viding a  grazing  area  with  good  visibility  and  access 
to  water.  Such  areas  are  frequently  near  the  nest 
sites  but  some  birds  nesting  on  western  rivers  may 
move  several  miles  to  congregate  on  desirable 
brood-rearing  areas. 

Another  habitat  requirement  of  geese  used  dur- 
ing the  breeding  season,  but  usually  distinct  from 
breeding  areas,  is  the  molting  area.  Nonproductive 
geese,  primarily  yearlings  (65%  ),  2-year-olds  (20%  ), 
and  older  birds  ( 1 5%  ),  leave  the  breeding  areas  in 
late  spring  for  traditional  molting  areas  (Krohn  and 
Bizeau  1979).  The  migration  is  usually  in  a  northerly 
direction.  Although  several  molting  areas  are  present 
in  the  western  U.S.,  many  of  the  Canada  geese  from 
this  area  migrate  to  the  Northwest  Territories  to 
molt  (Krohn  and  Bizeau  1979).  Key  physical  charac- 
teristics of  a  molting  area  are  a  large  body  of  water 
for  security,  a  good  food  supply,  and  little  human 
disturbance.  One  such  area  in  southwestern  Mon- 
tana, under  the  jurisdiction  of  the  U.S.  Bureau  of 
Land  Management,  annually  harbors  8,000  to  10,000 
geese.  This  site  has  received  special  consideration 
to  maintain  its  integrity  as  an  important  molting  area 
(Hildebrand  1979). 

Vegetation. 


Structure.  Protective  cover  near  goose  nests 
seems  optimal  when  it  at  least  breaks  the  outline  of 
the  incubating  bird,  but  is  not  so  high  as  to  prevent 
visibility  from  the  nest.  Thus,  in  areas  where 
vegetation  was  relatively  short  and  nesting  densities 
high,  geese  selected  islands  that  had  greater 
coverage  of  forbs  and  grass  (Giroux  1981). 
Conversely,  where  vegetation  was  relatively  high  and 
nesting  densities  low,  geese  selected  islands  with 
less  dense  vegetation  (Kaminski  and  Prince  1977).  In 
areas  where  high  densities  of  geese  are  nesting, 
visual  barriers  between  nest  sites  (vegetation  or 
topography)  permit  pairs  to  nest  closer  to  one 
another  with  less  aggressive  encounters  (Giroux 
1981). 

Structure  of  vegetation  in  brood-rearing  areas 
should  be  short  enough  to  permit  visibility,  at  least 
by  the  adults.  Grazing  geese  generally  avoid  dense  or 
high  vegetation  because  both  adults  and  goslings 
are  often  flightless. 

Species  Composition.  Where  nesting  cover  is 
concerned,  vegetation  structure  is  more  critical  than 
species  composition.  Migrating  and  wintering  geese 
make  extensive  use  of  agricultural  crops,  including 
cereal  grains  and  green  forage.  However,  on  brood- 
rearing  areas  where  flightless  birds  must  graze  adja- 


Waterfowl 


373 


cent  to  the  water,  a  wide  variety  of  green  forage 
is  utilized.  At  Canyon  Ferry  Wildlife  Management 
Area  near  Townsend,  Montana,  where  about  1,000 
goslings  are  produced  annually,  I  have  observed 
Canada  geese  grazing  heavily  on  wheatgrass  (Agropy- 
ron  sp. ),  spikerushes  (Eleocharis  sp. ),  new  shoots  of 
bulrush  (Scirpus  sp. ),  alfalfa  (Medicago  sativa),  and 
reed  canarygrass  (Phalaris  arundinacea),  while 
virtually  ignoring  abundant  plants  of  sweet  clover 
(Melilotus  officinalis). 


Dabbling  Ducks  (Anatini)  and  Wood  Ducks 
(Cairinini) 

Under  the  subfamily  Anatinae  (ducks),  more 
than  one  tribe  may  be  treated  as  a  single  species 
group  because  they  will  frequently  share  some  com- 
mon habitat  feature.  Although  many  ducks  have 
breeding  ranges  extending  far  into  Canada,  unlike 
most  geese,  many  also  nest  south  into  the  U.S.  Some 
have  very  extensive  breeding  ranges,  being  abundant 
on  both  the  prairies  and  the  tundra  (e.g.,  pintail 
[Anas  acuta]  whereas  the  black  duck  (A  rubripes) 
and  the  mottled  duck  (A  fulvigula)  are  consider- 
ably more  restricted  geographically.  Many  dabbling 
ducks  and  some  diving  ducks  reach  their  highest 
density  of  breeding  birds  in  the  north-central  U.S. 
and  south-central  Canada  in  an  area  called  the  Prairie 
Pothole  region.  Among  the  dabblers  in  this  group 
are  the  mallard  (A  platyrhynchos),  pintail,  gadwall 
(A  strepera),  blue-winged  teal  (A  discors),  and 
shoveler  (A  clypeata). 

Physical  Features.  Most  species  of  this  group  are 
birds  that  breed  primarily  in  the  prairies,  parklands, 
or  tundra,  all  of  which  provide  relatively  open 
shallow  marshes.  Notable  exceptions  include  the 
black  duck,  which  is  found  in  fresh  and  marine 
marshes,  swamps,  and  lakes  (and  prefers  wooded 
marshes  for  breeding),  and  the  wood  duck  (Aix 
sponsa),  which  prefers  small,  quiet  inland  streams 
and  ponds  near  woodland. 

Because  all  of  these  ducks  are  territorial  to 
some  degree,  breeding  habitat  is  enhanced  with  an 
abundance  of  water  areas  of  varying  sizes,  depths, 
and  configurations  that  provide  a  diversity  of  aquatic 
vegetation  structure.  Patterson  (1976)  found  a  sea- 
sonal shift  in  habitat  requirements  by  prairie-dwell- 
ing ducks,  where  the  breeding  pairs  disperse  to  all 
bodies  of  water,  irrespective  of  the  productivity  of 
the  waters.  As  the  breeding  chronology  progressed, 
the  broods  and  fledged  ducks  (which  had  greater 
demands  for  nutrition  but  less  demands  for  isolation) 
moved  selectively  to  the  more  productive  waters, 
thus  allowing  full  utilization  of  a  spatially  heteroge- 
neous aquatic  habitat. 

Stock  ponds  often  provide  valuable  habitats 
(Bue  et  al.  1964).  Although  in  comparatively  low 


densities  (seldom  exceeding  1/km    [1/247  a."]  on  arid 
grazing  lands  in  the  West),  these  ponds  are  often 
more  stable  (Brewster  et  al.  1976)  and  more  pro- 
ductive over  the  years  than  many  natural  wetlands. 
The  merits  of  different  types  of  stock  ponds  as  wa- 
terfowl habitat  were  discussed  by  Eng  et  al.  ( 1979). 


Stock  pond  in  Great  Basin  Desert  used  by  waterfowl. 

As  with  geese,  several  species  of  ducks  are  at- 
tracted to  island  habitats  for  nesting.  From  this  spe- 
cies group,  mallards  and  gadwalls  are  very  likely 
to  nest  on  islands.  In  the  absence  of  mammalian  pre- 
dation,  a  high  nest  success  combines  with  a  high 
degree  of  homing  in  ensuing  years  by  surviving 
young,  resulting  in  some  rather  high  nest  densities 
(Duebbert  1982;  Duebbert  et  al.  1983). 


Vegetation. 


Structure.  Most  dabbling  ducks  nest  on  the 
ground.  Black  ducks  nest  in  tree  cavities  and  stumps 
(Cowardin  et  al.  1967),  whereas  wood  ducks  are 
obligate  cavity  nesters  (Bellrose  et  al.  1964).  The 
mallard,  although  typically  a  ground-nesting  bird, 
will  nest  in  a  variety  of  elevated  sites,  natural  and 
artificial  (Cowardin  et  al.  1967;  Bishop  and  Barratt 
1970). 

Of  the  ground-nesting  ducks  in  this  group,  the 
pintail  appears  more  prone  to  nest  in  the  least  cover 
situation  than  any  of  the  others.  In  a  study  of  duck 
nesting  in  intensively  farmed  areas,  pintail  nests  were 
in  near-equal  densities  in  summer  fallow,  mulched 
stubble,  standing  stubble,  and  untitled  uplands, 
whereas  72%  or  more  of  other  duck  nests  were  in 
the  unfilled  uplands  (Higgins  1977). 

Duck-nesting  density  in  idle  grasslands  was  stud- 
ied in  North  Dakota  for  3  years;  the  3-year  average 
nesting  density  was  76.6  nests/km    (30  nests/mi.  ; 


374 


Waterfowl 


Duebbert  and  Lokemoen  1976).  Nesting  density  on 
the  idle  acres  for  one  of  the  years  was  four  times 
higher  than  that  found  on  nearby  agricultural  lands 
(farmed  and  grazed)  during  the  same  year.  Vegeta- 
tion measurements  on  the  idle  lands  in  May  showed 
an  average  of  90  dead  stems/m    (108/yd")  and 
heights  of  71,  47,  and  28  cm  (28,  19,  and  11  in.)  on 
dead  stems,  dead  leaves,  and  live  leaves,  respectively. 
The  preference  that  nesting  hens  had  for  this  vegeta- 
tion density  and  structure  was  shown  by  flights  up 
to  1.6  km  (1  mi.)  from  wetlands  to  nest  in  the  idle 
lands. 


Species  Composition.  As  with  geese,  when 
vegetation  is  used  as  cover,  species  composition 
appears  to  take  a  secondary  role  to  structure.  Several 
species  used  heavily  for  nesting  cover  by  dabbling 
ducks  in  Alberta  (Calamovilfa  longifolia,  Hordeum 
jubatum,  Spartina  gracilis)  had  one  common  de- 
nominator— they  were  unpalatable  to  cattle  and  con- 
sequently provided  the  structure  for  cover  (Keith 
1961 ).  The  same  could  probably  be  said  for  dense 
thickets  of  snowberry  (Symphoricarpos  sp. )  and  rose 
{Rosa  sp.),  which  are  frequently  found  on  the  prairie 
and  have  had  high  usage  by  nesting  ducks  (Hines 
and  Mitchell  1983;  Lokemoen  et  al.  1984). 

Successful  brood-rearing  involves  a  proper  com- 
bination of  cover  and  food.  Aquatic  cover  can  be 
provided  by  a  variety  of  emergent  plants,  some  of 
which  provide  cover  only  (Cattails  [Typha  sp.]) 
whereas  others  (bulrush,  spikerush,  arrowhead  [Sag- 
ittaria  sp.],  smartweeds  [Polygonum  sp.])  provide 
food  from  tubers,  foliage,  or  fruiting  bodies,  as  well 
as  cover.  Cattail,  although  often  used  as  cover  for 
nesting  ducks,  can  detract  from  the  attractiveness  of 
a  pond,  at  least  to  certain  species.  Keith  (1961)  re- 
ported an  increase  in  waterfowl  usage  of  ponds 
when  shoreline  cattail  was  reduced. 


Submerged  aquatics,  such  as  water  milfoil  and 
pondweeds  combined  with  seeds  of  spikerush,  were 
used  heavily  by  adult  dabblers  in  the  spring  (Keith 
1961).  Ducklings  showed  a  heavy  use  of  spikerush 
seeds  suggesting  a  feeding  pattern  along  the  shore 
where  this  plant  was  prevalent.  Flying  juveniles  fed 
more  heavily  on  submergents,  indicating  more  in- 
tense feeding  in  the  open  water. 

Major  foods  found  in  stock  pond  habitat  in  Mon- 
tana include  pondweeds,  water  milfoil,  common 
hornwort  (Ceratophyllum  demersum),  Canadian 
waterweed  (Elodea  canadensis),  watercrowfoot  but- 
tercup (Ranunculus  aquatilis),  and  longspike  spike- 
rush ( Eleocharis  macrostachya).  At  least  5  years 
must  pass  for  newly  constructed  stock  ponds  to  de- 
velop aquatic  vegetation  to  provide  adequate  food 
and  cover  (Hudson  1983). 


Bay  Ducks  (Aythyini)  and  Ruddy  Ducks 
(Oxyurini) 

Five  species  of  the  tribe  Aythyini,  all  of  a  single 
genus,  are  common  to  North  America.  These  five  are 
the  canvasback  (Aythya  valisineria),  redhead  (A 
americana),  ring-necked  duck  (A  collaris),  greater 
scaup  (A  marila),  and  lesser  scaup  (A  affinis).  The 
ruddy  duck  (Oxyura  jamaicensis),  although  a  mem- 
ber of  a  different  tribe  (stiff-tailed  ducks),  is  included 
here  because  it  has  some  similar  habitat  require- 
ments during  the  breeding  season.  These  two  tribes 
and  a  third,  Mergini,  are  often  included  under  the 
descriptive  term  diving  ducks  as  opposed  to  the 
dabbling  ducks  discussed  earlier. 

Although  many  diving  and  dabbling  ducks  may 
be  observed  on  the  same  bodies  of  water,  particu- 
larly during  the  breeding  season,  the  morphological 
differences  between  the  two  groups  are  reflected 
in  many  of  their  activities  and  often  in  the  segments 
of  a  pond  that  each  uses.  Divers  have  lobed  hind 
toes  and  generally  much  larger  feet  than  the  dab- 
bling ducks,  which  aid  in  their  underwater  swim- 
ming capabilities.  The  legs  on  divers  are  located 
farther  back  on  the  body  which  enhances  their  div- 
ing ability,  but  reduces  their  walking  capabilities — 
resulting  in  little  or  no  utilization  of  terrestrial 
habitats. 

Body  contour,  method  for  taking  flight,  and 
feeding  methods  also  separate  divers  from  dabblers. 
On  the  water,  divers  provide  a  much  lower  profile 
than  dabbling  ducks;  their  shorter  tail  feathers  pro- 
vide an  appearance  of  the  back  sloping  down  to  the 
water's  surface.  In  taking  flight,  divers  patter  along 
the  water  before  lift-off,  whereas  the  dabblers  leap 
immediately  into  the  air.  Divers  commonly  go  well 
below  the  surface  to  feed,  whereas  the  dabbling 
ducks  simply  tip  to  reach  below  the  surface  for 
aquatic  foods.  Primarily  because  of  the  divers'  inabil- 
ity to  maneuver  on  land,  ducks  observed  feeding  in 
grain  fields  will  be  dabblers. 

Physical  Features.  All  except  two  of  the  species 
within  this  group  normally  inhabit  the  open  prairie 
marshes.  One  of  these  exceptions,  the  ring-necked 
duck,  seems  to  show  preference  for  marshes 
bordered  by  woodland.  The  other,  the  greater  scaup, 
is  primarily  an  inhabitant  of  marshy  tundra.  The 
lesser  scaup,  although  commonly  found  associated 
with  prairie  marshes,  also  extends  its  breeding  range 
into  boreal  lakes  and  tundra  habitat. 

When  nesting,  canvasbacks,  redheads,  and  ruddy 
ducks  all  inhabit  a  variety  of  prairie  marshes,  sloughs, 
and  potholes  with  stands  of  emergent  vegetation.  Of 
the  three  species,  canvasbacks  seem  most  prone  to 
nest  in  the  small  ponds;  ruddy  ducks  and  especially 
redheads  seem  to  favor  the  larger,  deeper  bodies 
of  water.  This  preference,  in  conjunction  with  its 


Waterfowl 


375 


Ring-necked  duck. 


gregarious  nesting  habits,  makes  the  redhead  an  im- 
portant nesting  species  in  many  marshes  of  the  Inter- 
mountain  West — 20-25  times  more  abundant  than 
canvasbacks.  Although  not  evenly  distributed 
throughout  the  Intermountain  West — nor  the  marsh 
habitat — redheads  in  particular,  but  also  ruddy 
ducks,  lesser  scaup,  and  canvasbacks  reach  rather 
high  breeding  densities  in  western  marsh  habitats. 
Some  Nevada,  Utah,  and  Montana  marshes  attain  very 
high  breeding  densities  (919  redheads/km    [355/ 
mi.2]  of  wetlands  on  Salt  Lake  marshes;  Bellrose 
1978). 

All  of  the  species  in  this  group  are  inclined  to 
nest  closer  to  the  water  than  most  dabbling  ducks,  a 
reflection  of  their  close  tie  to  the  aquatic  environ- 
ment. Canvasbacks  and  ruddy  ducks  invariably  nest 
over  water.  Redheads  nest  over  water,  but  also  on 
land  (Keith  1961)  and  often  on  islands  (Vermeer 
1970;  Carlsen  1984).  Ring-necked  ducks  usually  nest 
within  marshes;  over  85%  nest  on  small  clumps  of 
floating  vegetation  or  within  clumps  of  marsh  vegeta- 
tion (Mendall  1958).  Although  both  the  greater  and 
the  lesser  scaup  will  nest  on  islands,  the  latter  seems 
to  show  a  greater  preference  for  such  sites  (Vermeer 
et  al.  1972;  Vermeer  1970). 


Vegetation. 


Structure.  The  canvasback,  redhead,  and  ruddy 
duck,  all  of  which  commonly  nest  over  the  water, 
are  highly  selective  to  nesting  ponds  that  provide 
emergent  vegetation  and  often  band  around  the 
shoreline  of  small  ponds.  Redheads  tend  to  avoid 
large  stands  of  emergents  without  breaks  of  open 
water  (Low  1945).  Ring-necked  ducks  and  scaup  all 
prefer  to  nest  close  to  water  and  consequently  use 
a  variety  of  structural  cover  over  water  (ring-necks) 
or  near  the  water  (scaup). 

Canvasback  broods  mostly  used  permanent  type 
wetlands,  larger  and  deeper  wetlands  (0.4  ha  [1  a.]), 


and  wetlands  with  less  than  one-third  of  the  water 
area  covered  by  emergent  vegetation  (Stoudt  1982). 
Broods  of  redheads,  ruddy  ducks  ( Evans  and  Black 
1956),  and  lesser  scaup  (Sugden  1973)  showed  a 
propensity  for  feeding  in  the  deeper,  more  open 
water  segments  of  the  pond.  Although  amphipods,  a 
favored  food  of  young  scaup,  appeared  to  be  more 
numerous  in  larger  and  deeper  ponds,  Sugden 
(1973)  believed  that  the  movement  to  larger  ponds 
by  older  scaup  was  related  to  security  and  not  food. 

Species  Composition.  A  variety  of  species  of 
emergent  vegetation  have  been  noted  as  nesting 
cover  for  canvasbacks,  redheads,  and  ruddy  ducks. 
Cattails,  bulrush,  whitetop  (Fluminia  festucacea), 
willows  (Salix  sp),  sedges,  and  reeds  (Phragmites 
communis)  were  listed  by  Stoudt  (1982)  in  Mani- 
toba. Bellrose  ( 1978)  reported  that  for  nesting 
cover,  redheads  appeared  to  have  a  preference  for 
bulrushes,  cattails,  and  sedges,  in  that  order.  He  also 
stated  that  ruddy  ducks  seemed  to  select  the  vegeta- 
tion characteristic  of  the  nesting  pond,  availability 
playing  a  key  role.  I  have  observed  this  on  stock 
dams  in  eastern  Montana,  when  in  good  moisture 
years,  ruddy  ducks  will  readily  nest  in  a  relatively 
short  emergent,  longspike  spikerush,  the  only  emer- 
gent cover  present. 

Most  food  habits  studies  of  divers  point  out  the 
heavy  intake  of  animal  matter  by  young  ducklings 
and  an  increase  in  plant  food  as  the  birds  get  older 
(Bartonek  and  Hickey  1969;  Sugden  1973)-  Although 
specific  foods  will  vary  considerably  with  location 
and  availability  within  this  species  group,  ruddy 
ducks  appear  to  feed  most  heavily  on  plant  materials. 
Both  species  of  scaup  are  most  inclined  to  an  animal 
diet. 

Plants  frequently  used  are  pondweeds,  musk- 
grass  (Chara  sp. ),  bulrush,  wild  celery  (Valisneria 
americana),  duckweeds  (Lemna  sp. ),  water  milfoil, 
and  widgeon  grass.  Depending  on  the  species  of 
plant  and  season  of  year,  leafy  parts,  seeds,  and  tub- 
ers of  the  various  plants  may  be  fed  upon. 

Eiders,  Scoters,  Mergansers,  and  Allies 
(Mergini) 

This  relatively  diverse  tribe  is  represented  in 
North  America  by  seven  genera,  all  different  in  ap- 
pearance as  adults,  but  grouped  together  at  least 
partially  as  a  result  of  similarity  in  voice  and  display 
(Delacour  and  Mayr  1945).  Members  of  this  entire 
group  are  referred  to  as  "sea  ducks,"  primarily  be- 
cause most  of  them  winter  along  the  sea  coasts. 
However,  several  members  breed  inland  on  fresh 
water,  and  a  few  winter  in  the  interior. 

Included  in  the  genus  Somateria  are  three  ei- 
ders, the  common  eider  (S.  mollissima),  king  eider 
(S.  spectabilis),  and  spectacled  eider  (S.  fischeri),  all 


376 


Waterfowl 


of  which  breed  along  the  arctic  coasts.  The  common 
and  king  eiders  winter  along  the  northern  coasts; 
the  spectacled  eider's  wintering  areas  remain  some- 
what obscure.  Steller's  eider  (Polysticta  stelleri),  the 
sole  member  of  this  genus,  nests  along  the  northern 
coasts  of  Alaska  and  Siberia  and  winters  along  the 
Alaska  peninsula  and  the  Aleutians  (Bellrose  1978). 

The  harlequin  duck  (Histrionicus  histrionicus) 
has  a  distinct  eastern  and  western  range.  It  breeds 
from  the  northern  coasts  of  Canada  and  Alaska  south- 
ward into  the  Cascades  and  Sierra  Nevadas  and  the 
Rocky  Mountains  (Bellrose  1978),  and  winters  along 
both  coasts  as  far  south  as  California  and  Long  Island. 


Harlequin  duck. 


Oldsquaws  (Clangula  hyemalis)  are  found 
throughout  arctic  Canada  and  Alaska  during  the 
breeding  season;  they  winter  along  both  coasts,  com- 
monly south  to  northern  California  and  northern 
North  Carolina  and  interior  on  the  Great  Lakes. 

Three  scoters  are  included  under  the  genus 
Melanitta:  the  black  scoter  (M.  nigra  americana), 
surf  scoter  (M.  perspicillata),  and  white-winged 
scoter  (M.  fusca  deglandi).  The  black  scoter  breeds 
in  Alaska  and  probably  in  Canada  (Bellrose  1978). 
Breeding  surf  and  white-winged  scoters  are  both 
well  distributed  across  Canada  in  the  open  and 
closed  boreal  forest.  All  three  species  winter  along 
most  of  the  east  and  west  coasts  of  the  U.S. 

The  genus  Bucephala  includes  the  bufflehead 
(B.  albeola)  and  two  goldeneyes,  the  Barrow's  (B. 
islandica)  and  common  (B.  clangula  americana). 
The  bufflehead  has  a  rather  extensive  breeding  range 
throughout  Canada;  the  largest  numbers  are  in  north- 
western North  America  (Bellrose  1978).  A  few  iso- 
lated breeding  populations  are  found  in  mountainous 
areas  of  the  U.S.  and  in  some  timbered  areas  of  the 
northern  prairies.  The  primary  breeding  populations 
of  Barrow's  goldeneyes  are  in  northwestern  North 
America  extending  from  the  mountains  of  Wyoming 


and  northern  California,  north  into  Alaska.  The  com- 
mon goldeneye  is  more  widespread,  breeding  in 
forest  areas  across  the  breadth  of  Canada  and  into 
Alaska.  All  three  species  of  this  genus  concentrate 
along  both  the  Atlantic  and  Pacific  coasts  during  the 
winter,  but  can  also  be  found  in  small  concentrations 
along  rivers  and  lakes  in  the  interior. 


Bufflehead. 


Male  and  female  common  goldeneye. 


Three  species  of  mergansers  (Mergus)  are  com- 
mon to  North  America:  the  hooded  (M.  cucullatus), 
red-breasted  (M.  serrator),  and  common  (M.  mer- 
ganser americanus).  The  hooded  merganser  nests 
throughout  most  of  the  eastern  half  of  the  U.S.,  ex- 
tending into  adjoining  provinces  in  Canada.  It  also 
nests  in  the  northwest  U.S.  from  northern  California 
to  southeast  Alaska  (Bellrose  1978). 

The  red-breasted  merganser  breeds  throughout 
eastern  Canada,  across  the  southern  tundra  and 
northern  boreal  forest  into  Alaska.  The  common 
merganser  has  a  somewhat  similar  distribution  more 


Waterfowl 


377 


to  the  south  although  with  a  broad  overlap  zone 
where  both  may  be  found.  The  common  merganser 
breeds  in  the  northeast  and  lake  States,  and  extends 
southward  along  mountain  ranges  in  the  West. 

The  hooded  merganser  winters  along  the  east, 
west,  and  gulf  coasts  and  several  southeastern  States; 
the  red-breasted  winters  largely  on  the  Great  Lakes 
and  along  both  coasts.  The  common  merganser,  by 
contrast,  winters  largely  (  58%  )  in  the  interior 
(Bellrose  1978). 

Physical  Features.  Of  the  three  Somateria  eiders, 
the  common  is  the  most  marine,  occurring  primarily 
along  rocky  shores  and  islands  and  only  rarely  on 
open  fresh  water.  The  king  and  spectacled  eiders, 
although  they  both  winter  along  seacoasts,  breed 
near  fresh-water  ponds,  lakes,  deltas,  and  tidal  inlets. 
The  Stellar's  eider  nests  along  the  coast  or  inland 
near  grassy  ponds  or  lakes  but  winters  along  coasts 
in  shallow  marine  habitat. 

The  harlequin  duck  usually  nests  along  fast- 
flowing  coastal  or  mountain  streams  and  winters 
along  rocky  sea  coasts. 

Oldsquaws,  unlike  the  eiders,  nest  along  the 
coast  and  inland  on  lakes  and  ponds  throughout  the 
tundra.  Like  most  of  the  other  sea  ducks,  it  winters 
along  coastal  areas  but  also  on  large  inland  lakes 
(e.g.,  Great  Lakes). 

All  three  scoters  nest  in  fresh-water  situations, 
the  black  and  the  surf  showing  a  preference  for 
more  brushy,  taller  cover  near  the  water,  whereas 
the  white -winged  scoter  is  more  prone  to  nest  in 
open  prairie  or  tundra  and  frequently  at  considerable 
distances  from  the  water.  Dense  ground  cover  for 
nesting  is  a  prominent  characteristic  (Bellrose 
1978). 

Goldeneyes  and  buffleheads  are  cavity  nesters, 
so  breeding  habitat  usually  consists  of  trees  sur- 
rounding fresh-water  lakes,  ponds,  and  streams.  All 
three  species  winter  along  the  coasts  in  bays  or  estu- 
aries as  well  as  lakes  and  rivers  inland. 

Hooded  and  common  mergansers  are  also  cavity 
nesters.  The  hooded  merganser  seems  more  re- 
stricted to  tree  cavities  or  wood  duck  houses  than 
the  common  merganser,  which  frequently  uses  cavi- 
ties in  rock  cliffs  along  major  western  rivers.  The 
red-breasted  merganser  nests  on  the  ground,  often 
under  shrubs  (Weller  et  al.  1969). 

Wintering  habitat  for  the  three  mergansers  var- 
ies: the  common  winters  largely  on  open  lakes  and 
rivers  and  only  rarely  in  coastal  habitat;  the  hooded 
winters  mostly  in  fresh  water,  but  it  is  also  com- 
monly found  in  estuaries  and  bays;  the  red-breasted 


Hooded  merganser. 


merganser  winters  primarily  in  coastal  areas  and 
rarely  in  fresh-water. 

Vegetation.  Less  can  be  concluded  regarding  the 
vegetational  aspect  of  sea  duck  habitat  than  with 
most  other  tribes  of  waterfowl.  Several  reasons  may 
be  put  forth  as  a  partial  explanation. 

Nesting  by  many  of  the  species  is  carried  out  in 
the  tundra  or  in  open  or  closed  boreal  forest.  Resid- 
ual vegetation  from  previous  years'  growth  is  critical 
because  the  short  season  requires  that  nesting  be 
initiated  before  the  appearance  of  much  of  the  cur- 
rent season's  growth. 

The  importance  of  structural  cover  for  nests 
rather  than  species  composition  is  reflected  in  the 
selection  of  shrubs,  rocks,  burrows,  and  overhanging 
banks  by  many  species.  Several  species  (eiders,  olds- 
quaw,  some  scoters)  nest  on  islands  in  rather  large 
numbers  (colonial)  and  as  such  rely  less  on  cover 
and  more  on  the  mammal-free  island  habitat. 

Also,  several  species  (goldeneyes,  bufflehead, 
common  and  hooded  mergansers)  are  cavity  nesters, 
and  as  such  the  dominating  feature  is  the  proper- 
sized  cavity  and  not  the  plant  species  providing  the 
cavity.  However,  certain  trees  appear  more  prone  to 
natural  cavities  and  species  composition  may  thus 
influence  selection  of  a  particular  habitat.  Erskine 
(1972)  in  British  Columbia  noted  that  of  1 1  trees 
that  provided  cavities  for  nesting  buffleheads,  52% 
were  aspen  (Populus  tremuloides). 

The  relative  minor  importance  of  vegetation 
structure  and  composition  to  sea  ducks  is  also  indi- 
cated by  the  little    se  made  of  plants  in  their  food 
habits.  Results  from  several  studies  showed  that  ani- 
mal matter  in  the  diet  ranged  from  about  70  to  95% 
for  eiders,  harlequins,  oldsquaws,  scoters,  buffle- 
heads, and  goldeneyes,  and  100%  (largely  fishes)  for 
mergansers  (Bellrose  1978).  Many  of  the  insect  lar- 


378 


Waterfowl 


vae  taken  as  food  by  young  ducks  of  several  tribes  or 
adult  dabbling  ducks  are  in  close  association  with 
aquatic  vegetation.  However,  many  of  the  animal 
foods  eaten  by  sea  ducks  are  crustaceans  and  mol- 
lusks  taken  during  bottom  feeding  and  have  rela- 
tively less  relationship  to  aquatic  macrophytes. 


CHARACTERISTICS  OF  WINTER  HABITAT 

Winter  habitat  for  waterfowl  has  only  recently 
received  the  attention  it  deserves.  For  6  to  8  months 
of  the  year,  this  habitat  must  provide  security  and  a 
food  source  capable  of  sustaining  the  bird.  For  many 
species  it  must  also  provide  an  energy  base  for  pair- 
ing and  initiation  of  migration.  With  the  wide  range 
of  species  requirements  present  within  the  water- 
fowl for  breeding  habitat,  differences  would  also  be 
expected  in  winter  habitat  requirements. 

Probably  the  least  variability  in  winter  habitat  is 
found  among  the  tribes  that  are  most  aquatic. 
Groups  within  the  tribe  Mergini,  which  feed  largely 
on  aquatic  invertebrates,  are  confined  to  coastal 
waters,  or  if  wintering  inland,  are  usually  found  on 
larger  lakes  or  rivers.  Heitmeyer  and  Vohs  (1984),  in 
looking  at  wintering  waterfowl  habitat  in  Oklahoma, 
found  common  mergansers  and  common  goldeneyes 
more  abundant  on  the  large  reservoirs. 

The  diving  ducks  (Aythyini),  another  highly 
aquatic  group,  tend  to  concentrate  along  certain 
coastal  areas.  In  contrast  with  many  of  the  Mergini 
which  winter  in  marine  habitat,  the  diving  ducks 
seek  out  estuarine  habitat.  In  1955,  slightly  less  than 
one-half  of  the  canvasbacks  recorded  in  North  Amer- 
ica were  found  in  Chesapeake  Bay  (Perry  et  al. 
1981 ),  whereas  about  78%  of  the  redheads  in  exis- 
tence normally  winter  on  Laguna  Madre  along  the 
Texas  coast  (Weller  1964). 

Both  species  of  scaup  make  heavy  use  of  coastal 
areas,  the  greater,  more  abundant  in  the  Atlantic 
Flyway  and  the  lesser,  more  prominent  in  coastal 
marshes  of  the  Mississippi  Flyway.  Both  scaup  also 
winter  inland  as  far  north  as  the  Great  Lakes 
(Bellrose  1978). 

Dabbling  ducks  show  considerable  variability  in 
selecting  winter  habitat,  both  between  and  within 
species.  Mallards  and  pintails  show  different  feeding 
preferences  depending  on  which  portion  of  their 
respective  ranges  they  may  be  occupying  (Chabreck 
1979).  Fredrickson  and  Drobney  (1979)  pointed 
out  the  importance  of  a  food  resource  on  wintering 
areas  and  referred  to  specific  feeding  habits  that 
permit  several  species  of  ducks  to  fully  and  at  times 
simultaneously  exploit  an  aquatic  food  resource. 

Thus,  with  high  densities  of  ducks  so  often 
found  in  wintering  areas,  a  highly  productive  aquatic 


food  source  is  a  key  component  of  winter  habitat. 
Gadwalls  and  widgeon  in  Oklahoma  wintered  pri- 
marily on  wetlands  where  submergent  and  emergent 
vegetation  were  abundant  (Heitmeyer  and  Vohs 
1984).  All  species  of  dabbling  ducks  preferred  natu- 
ral wetlands  over  farm  ponds,  probably  a  reflection 
of  the  absence  of  biologically  productive  littoral 
zones  on  the  steep-sided  farm  ponds.  In  determining 
values  for  13  physical,  limnological,  and  vegetative 
features  on  a  series  of  flood-prevention  lakes  in 
Texas,  Hobaugh  and  Teer  (  1981 )  concluded  that  the 
most  important  characteristics  influencing  winter 
waterfowl  use  were  amounts  of  aquatic  vegetation 
and  lake  surface  area. 

Some  species  of  waterfowl  have  been  able  to 
capitalize  on  land-use  changes  by  modifying  their 
food  habits  to  include  certain  upland  agricultural 
crops.  Historically,  upland  feeding  in  cultivated  fields 
in  Manitoba  was  first  recorded  for  geese,  which 
could  be  expected  in  light  of  the  terrestrial  habits  of 
this  group.  Use  of  cultivated  fields  by  ducks  seem- 
ingly developed  later  and  initially  appeared  with 
mallards  and  pintails  (Bossenmaier  and  Marshall 
1958).  Field  feeding,  although  initially  considered  a 
fall  activity,  has  to  be  a  major  contributor  in  the 
northward  extension  of  wintering  ducks  (primarily 
mallards)  into  Montana  and  the  two  Dakotas.  Many 
of  these  birds  winter  on  large  bodies  of  water  that 
are  totally  devoid  of  available  aquatic  vegetation, 
thus  forcing  a  total  subsistence  on  cultivated  grains. 
On  wintering  areas  farther  south,  American  widgeons 
(Anas  americana),  and  blue-winged  and  green- 
winged  teal  (A  crecca)  join  the  mallard  and  north- 
ern pintail  as  regular  field  feeders  (Baldassarre  and 
Bolen  1984). 

POPULATION  MEASUREMENT  TECHNIQUES 

Virtually  all  techniques  employed  for  population 
measurements  of  waterfowl  are  based  on  direct  ob- 
servations of  the  birds.  This  contrasts  with  tech- 
niques, such  as  auditory  counts,  employed  in 
monitoring  upland  bird  populations.  The  obvious 
reason  for  this  difference  is  that  waterfowl  are 
closely  associated  with  water  for  much  of  the  time 
and  although  they  are  no  more  equally  visible  be- 
tween different  bodies  of  water  and  between  sea- 
sons, they  are  easier  to  observe  and  count  than  many 
other  species  of  birds.  Because  the  methods  are 
based  on  direct  observation,  they  all  detect  presence 
with  varying  degrees  of  precision  and  accuracy. 

On  a  continental  scale,  breeding  and  winter 
populations  are  surveyed  annually  by  the  U.S.  Fish 
and  Wildlife  Service,  Canadian  Wildlife  Service,  and 
State  and  Provincial  conservation  agencies.  Most 
of  such  surveys  are  conducted  from  the  air  and  are 
used  for  determining  long-term  trends  (Bellrose 
1978:17-19). 


Waterfowl 


379 


Breeding  Population 

Description.  Annual  trend  data  on  North  American 
populations  of  breeding  waterfowl  are  gathered  in 
May  and  June  along  aerial  east-west  transects 
(Bellrose  1978:18).  The  transects,  which  are 
censused  cooperatively  by  State,  Federal,  and 
Canadian  biologists,  are  distributed  throughout 
prime  breeding  areas  from  the  prairies  to  the  boreal 
forest  and  into  the  tundra.  Because  visibility  varies 
between  species  and  years,  segments  of  transects  are 
covered  intensively  by  a  ground  crew  within  24 
hours  of  the  air  count  and  a  visibility  index  applied 
to  the  aerial  count.  This  annual  survey,  which  also 
provides  data  on  habitat  (water)  conditions,  is 
conducted  over  about  80,500  km  ( 50,000  linear  mi. ) 
representing  about  2%  of  the  breeding  habitat 
(Bellrose  1978). 

This  basic  technique,  involving  ground  counts 
only,  has  been  modified  and  refined  making  it  more 
useable  on  a  smaller  scale.  The  count,  called  a  breed- 
ing pair  count,  may  be  conducted  on  specific  bodies 
of  water,  along  roadside  transects,  or  on  all  water 
areas  within  a  predescribed  area.  The  counts  are 
based  on  the  knowledge  that  breeding  pairs  of  dab- 
blers, once  established  on  a  breeding  home  range, 
will  localize  their  activities  to  one  or  two  ponds. 
Thus,  from  the  time  of  pair  bond  formation  (often 
on  arrival  because  many  species  form  pair  bonds  on 
the  wintering  ground  or  enroute  north)  until  the 
female  is  well  into  incubation,  the  locations  of  pairs 
or  lone  drakes  are  plotted  during  four  or  five  cen- 
suses. Pairs  or  lone  drakes  of  any  dabbler  species 
occurring  on  a  given  unit  of  habitat  three  or  four 
times  are  recorded  as  part  of  the  assigned  breeding 
population  (Dzubin  1969).  The  counts  should  be 
made  when  the  highest  percentage  of  pairs  are  in 
the  pre-nesting,  laying,  or  incubation  stage.  This  may 
include  a  period  of  20  to  30  days  to  include  early- 
nesting  pintails  and  mallards,  as  well  as  late-nesting, 
blue-winged  teal  and  gadwalls.  Although  this  tech- 
nique is  often  described  to  include  all  ducks,  Dzubin 
(1969)  believed  the  large  home  ranges  of  canvas- 
backs  and  redheads  and  the  tendency  for  pairs  of 
these  two  species  and  lesser  scaup  to  congregate  on 
deep  ponds  termed  "primary  waiting  areas,"  pre- 
cluded enumerating  divers  in  smaller  census  units. 

For  additional  precision,  Dzubin  (1969)  pro- 
posed an  "Indicated  Breeding  Population"  where,  un- 
til a  certain  date,  grouped  males  and  males  in  aerial 
flights  temporarily  on  censused  ponds  were  also 
considered  pairs.  After  that  date,  grouped  males  of 
five  or  more  were  not  counted  as  pairs  (six  or  more 
were  considered  post-breeding  males).  Different 
cutoff  dates  were  established  for  early  nesters  (pin- 
tails and  mallards),  intermediates  (widgeon  and 
shovelers),  and  late  nesters  (blue-winged  teal  and 
gadwalls);  dates  may  vary  between  areas  and  years 


but  are  based  on  nesting  chronology  of  the  respec- 
tive groups. 

Lastly,  a  sex-ratio  corrected  population  figure 
was  suggested.  This  consisted  of  taking  the  mean-in- 
dicated breeding  population  from  four  or  five  census 
efforts  and  applying  a  sex-ratio  correction  factor 
based  upon  sex-ratio  data  obtained  from  each  spe- 
cies before  egg  laying. 

The  following  are  some  of  the  recommendations 
by  Dzubin  (1969)  for  breeding-pair  census  on  a 
grassland  type.  This  type,  which  has  little  visual  ob- 
struction from  emergent  vegetation,  is  appropriate 
for  much  of  the  waterfowl  habitat  on  public  lands  in 
the  West. 

( 1 )  Census  during  that  portion  of  the  breeding 
season  when  site  attachment  by  pairs  and 
drakes  is  greatest  (pre-nesting,  laying,  and  early 
incubation). 

(2)  Census  between  0800  and  1200  hours;  this  is 
the  period  of  least  mobility,  and  most  pairs  and 
lone  drakes  will  be  on  waiting  stations. 

(3)  Census  only  on  bright  days  with  temperatures 
above  4°  C  (40°  F)  and  winds  less  than  24  km/ 
h  (15  mph).  Avoid  inclement  weather  (rain, 
heavy  overcast,  and  low  temperatures)  which 
affects  mobility  and  visibility  of  ducks  (Diem 
and  Lu  I960). 

(4)  Conduct  at  least  two  censuses  when  sampling 
a  multiple  species  population;  four  to  six 
counts  would  be  preferable,  permitting  a  calcu- 
lated average  number  of  pairs  for  each  species. 

(  5 )  Take  counts  from  a  vehicle,  positioned  at  a 
vantage  point  at  an  adequate  distance  to  pre- 
vent flushing  birds. 

(6)  Tally  all  lone  pairs  and  lone  drakes  greater 
than  5  m  ( 1 5  ft )  apart  as  pairs;  this  type  enu- 
meration, if  conducted  four  to  six  times  on  the 
census  area  with  proper  timing,  will  provide 
relative  abundance  data  between  areas  or  be- 
tween years  on  a  given  area.  Greater  precision 
can  be  obtained,  if  the  objectives  warrant,  by 
a  more  refined  timetable  as  recommended 
by  Dzubin  (1969:220). 

Breeding  populations  of  Canada  geese  are  com- 
monly censused  along  rivers  either  from  the  ground 
(boat)  or  air  at  which  time  pairs  and  lone  geese 
(ganders)  are  recorded  as  pairs  and  groups  as  non- 
breeders  (Hanson  and  Eberhardt  1971;  Allen  et  al. 
1978).  Unlike  most  ducks,  geese  remain  paired 
throughout  the  nesting  season  so  lone  ganders  will 
be  available  for  tally  (as  pairs)  throughout  the  incu- 
bation period. 


380 


Waterfowl 


Prime  breeding  areas  on  the  prairies  are  censused  in  May 
and  June  along  aerial  transects. 


Accuracy  and  Precision.  Sauder  et  al.  (1971),  in 
evaluating  the  roadside  census  for  breeding 
waterfowl  in  South  Dakota,  found  that  to  be  within 
20%  of  the  mean  at  the  90%  confidence  level  on  an 
86.4  km  (54  mi.)  route,  at  least  four  counts  of  lone 
drakes  and  pairs  were  needed  for  blue-winged  teal 
and  three  for  gadwalls  and  mallards.  They  also  stated 
that  additional  counts  were  necessary  to  maintain 
the  same  degree  of  accuracy  when  routes  were 
shorter  or  when  water  areas  were  less  numerous. 
Although  Dzubin  (1969)  did  not  recommend  lone 
drake  and  pair  counts  for  divers,  Sugden  and  Butler 
(1980),  after  an  evaluation,  concluded  that  lone 
male  and  pair  counts  for  canvasbacks  and  total 
female  counts  for  redheads  provided  the  best  index 
to  breeding  populations  of  these  species. 

In  spite  of  the  obvious  shortcomings  with  the 
breeding  pair  counts,  indications  are  good  that  it  will 
provide  useable  data  for  certain  management  prob- 
lems. Pair  and  lone  drake  counts  were  conducted  on 
33  stock  ponds  during  a  5-year  study  to  measure 
waterfowl  responses  to  a  changing  land  use  (Gjers- 
ing  1975;  Mundinger  1976).  Although  the  changes  in 
waterfowl  numbers  were  fairly  pronounced  over 
the  period,  production  data  obtained  from  intensive 
brood  counts  for  the  same  period  followed  a  similar 
trend.  For  many  evaluations  of  waterfowl  habitat 
on  western  ranges,  carefully  timed  and  conducted 
breeding  pair  counts  (lone  pairs  and  drakes)  should 
fulfill  the  objectives. 

Discussion.  Without  knowing  what  percentage  of 
the  actual  breeding  population  is  being  counted, 
counts  of  breeding  pairs  at  best  provide  yearly 
trends.  Timing  of  the  count  is  critical.  The 
chronology  of  the  nesting  effort  of  a  single  species 
may  extend  over  a  70-day  period  (Humburg  et  al. 
1978)  and  longer  when  several  species  are 
considered.  Thus  for  intensive  work,  several  counts 
are  necessary,  well-dispersed  over  the  entire 
breeding  season. 


Nest  Counts 

Nest  counts  are  often  incorporated  in  (or  substi- 
tuted for)  breeding  pair  counts.  Dzubin  (1969:2217) 
suggested  this  as  a  better  method  on  block-type 
studies  for  divers  and  ruddy  ducks,  although  he 
pointed  out  potential  problems  in  uneven  nest  distri- 
bution, failure  to  locate  all  nests,  and  difficulty  in 
separating  first  and  second  nest  attempts. 

Sugden  and  Butler  (  1980)  considered  nest 
searching  as  an  impractical  technique  for  recording 
canvasback  and  redhead  breeding  densities,  primarily 
because  the  variability  in  nest  distribution  made  it 
necessary  to  have  a  large  sample.  Nest  counts  are 
effective  for  enumerating  breeding  populations  of 
Canada  geese,  particularly  when  nesting  is  largely  ac- 
complished on  islands  in  rivers  (Hansen  and  Eber- 
hardt  1971)  or  ponds  (Childress  and  Eng  1979). 

A  measure  of  various  upland  habitats  may  be 
made  by  comparing  nest  densities  or  nest  success. 
Nests  are  located  by  systematically  covering  the  area 
in  question  at  specified  time  intervals  to  locate  early 
and  late  nesting  birds.  This  effort  is  frequently  facili- 
tated by  the  cable-chain  drag  method  (Higgins  et 
al.  1969). 

With  early,  intermediate,  and  late  nesting  spe- 
cies plus  renesting  by  most  species,  a  continued 
initiation  of  nests  will  be  occurring  throughout  the 
nesting  season.  It  is  not  practical  (nor  advisable) 
to  search  an  area  daily;  thus  many  nests  may  be  initi- 
ated and  destroyed  between  search  periods  and  con- 
sequently not  found.  Also,  nests  located  early 
(during  laying  or  early  incubation)  suffer  a  greater 
chance  of  destruction  before  hatching  than  a  nest 
first  located  shortly  before  hatching.  If  all  nests  are 
considered  equal  regardless  of  their  stage  when  first 
found,  the  nest  density  would  be  biased  downward 
(e.g.,  nests  destroyed  early  are  not  found)  and  the 
hatch  rate  (nest  success)  biased  upward.  The  recom- 
mended method  for  standardizing  waterfowl  nest 
density  and  success  data  is  the  Mayfield  method 
(Mayfield  1975;  Miller  and  Johnson  1978)  in  which  a 
hatch  rate  is  determined  and  applied  to  each  nest 
depending  on  its  stage  of  development  toward 
hatching. 


Brood  Counts 

Production  surveys  are  conducted  by  U.S.  Fish 
and  Wildlife  Service  biologists  to  assess  annual  wa- 
terfowl breeding  success.  Brood  counts  are  con- 
ducted from  the  air  along  many  of  the  same 
transects  as  the  breeding  bird  counts  and  are  subject 
to  many  of  the  same  problems.  However,  Pospahala 
et  al.  (1974)  supported  the  general  contention  that 
the  number  of  Class  II  and  III  broods  observed  along 
aerial  transects  in  July,  when  adjusted  for  visibility, 


Waterfowl 


381 


Nesting  coot. 


is  related  to  the  density  of  breeding  birds  observed 
in  the  May  survey. 


Although  broods  are  less  mobile  than  flying 
adults,  they  are  usually  more  secretive.  By  the  time 
most  broods  are  on  the  water,  the  emergent  vegeta- 
tion is  more  lush — reducing  visibility.  A  serious  need 
for  aerial-ground  indexes  as  adjustments  for  the  va- 
garies of  aerial  counts  alone  was  stressed  by  Diem 
and  Lu  (I960). 


Brood  counts  conducted  more  intensively  from 
the  ground  appear  more  acceptable.  Gollop  and  Mar- 
shall (1954)  presented  a  method  for  brood  counts 
that  basically  calls  for  several  counts  to  be  made  and 
broods  recorded  by  species  and  age,  which  tends 
to  eliminate  duplication  of  counts.  This  method  has 
been  employed  extensively  on  studies  where  more 
precise  production  data  are  important  (Berg  1956; 
Mundinger  1976;  Kirby  1980). 


Stock  ponds  in  Montana  generally  have  far  less 
emergent  vegetation  than  prairie  potholes.  Mundin- 
ger (1976),  in  making  brood  counts,  would  usually 
approach  the  pond  unobtrusively  in  an  attempt  to 
count  undisturbed  broods.  Even  so,  brood  move- 
ment ("transient  broods")  made  it  necessary  to  re- 
cord broods  by  age  class  to  prevent  duplication. 


Kirby  ( 1980)  analyzed  intensive  observations  of 
breeding  pairs  and  broods  of  ducks  in  forested  habi- 
tats of  northern  Minnesota.  Aerial  censuses  in  such 
areas  appeared  inadequate  because  of  the  difficulty 
in  observing  birds  at  low  densities.  Kirby  (1980) 
found  that  indexes  to  breeding  birds  and  brood 
numbers  were  highly  correlated,  and  that  censusing 
was  a  distinct  possibility  as  a  source  for  population 
trend  data  in  this  type  habitat. 


Winter  Surveys  (Counts) 

The  annual  January  inventory  was  initiated  in 
the  U.S.  in  1935  and  has  increased  in  scope  until  it 
now  encompasses  most  wintering  areas  in  Alaska, 
Canada,  the  conterminous  U.S.,  and  Mexico.  The 
survey  is  coordinated  by  the  U.S.  Fish  and  Wildlife 
Service,  and  although  the  purpose  is  to  determine 
the  size  and  distribution  of  the  midwinter  waterfowl 
population,  the  potential  for  error  (at  least  in  deter- 
mining total  population)  is  obvious  and  great.  Black 
ducks  and  wood  ducks  winter  in  small  groups  more 
easily  counted,  but  winter  in  habitat  where  they 
would  be  more  easily  overlooked.  Canvasbacks  and 
redheads  often  concentrate  in  open  areas  where  they 
can  be  easily  seen,  but  in  such  large  flocks  that  esti- 
mates of  numbers  are  often  subject  to  error.  Popula- 
tion status  of  wintering  canvasbacks  on  Chesapeake 
Bay  and  coastal  North  Carolina  were  efficiently  and 
accurately  obtained  with  low-level  aerial  photogra- 
phy (Haramis  et  al.  1985). 

A  detailed  description  of  the  winter  survey  was 
presented  by  Stewart  et  al.  (1958:366-367).  Bellrose 
(1978:20)  acknowledged  that  winter  counts,  for 
several  reasons,  were  undoubtedly  underestimated. 
However,  he  defended  their  use  because  they  pro- 
vided the  distribution  of  ducks  at  that  time  of  the 
year  and  provided  the  major  annual  population  data 
for  geese.  Most  geese  nest  in  such  widely  distributed 
parts  of  the  arctic  and  subarctic  that  annual  counts 
are  impractical. 


DISCUSSION 

Tiner  (1984)  provided  an  excellent  summary  of 
the  current  status  and  recent  trends  of  wetlands  in 
the  U.S.  Although  some  recent  modest  gains  were 
recorded  as  a  result  of  man-made  lakes  and  reser- 
voirs, a  net  loss  of  3-6  million  ha  (9  million  a.)  oc- 
curred in  a  20-year  period  starting  in  the  mid-1950s. 
Frayer  et  al.  (1983)  attributed  87%  of  these  losses 
to  agricultural  activities,  8%  to  urban  development, 
and  5%  to  other  development.  The  geographical 
distribution  is  no  more  evenly  distributed,  but 
clearly  related  as  to  the  causes.  In  the  fertile  farm- 
lands of  Iowa,  99%  of  the  natural  marshes  have  been 
drained  compared  with  32%  of  Wisconsin's  wetlands 
(Tiner  1984,  p. 34). 

The  major  waterfowl  habitats  within  the  U.S.  are 
important  primarily  as  breeding  habitat,  wintering 
habitat,  or  both  (Figure  1 ).  The  prairie  pothole  re- 
gion, for  example,  encompasses  the  most  valuable 
waterfowl  production  marshes  in  North  America. 
Representing  no  more  than  10%  of  the  continent's 
waterfowl  breeding  area,  in  a  wet  year  it  produces 
over  50%  of  the  ducks.  Included  are  many  of  the 
popular  species  including  the  mallard,  pintail,  teal, 
widgeon,  canvasback,  and  redhead.  Of  more  recent 


382 


Waterfowl 


AREA  NAME 


1 — Prairie  potholes  and  parklands 

2 — Central  Valley  of  California 

3 — Yukon  Kiiskokivin  Delta 

4 — Middle  Upper  Atlantic  Coast 

5 — Lower  Mississippi  River  Delta  and  Red  River  Basin 

6 — Izembek  Lagoon 

7 — Upper  Mississippi  River  and  Northern  Lakes 

8 — Northern  Great  Plains 

9 — Yukon  Flats 
10 — Intermountain  West  (Great  Basin) 
11 — Teshelpuk  Lake 
12— Middle  Upper  Pacific  Coast 
13 — Klamath  Basin 
14 — Upper  Alaska  Peninsula 
15 — Copper  River  Delta 
1(>—West  Central  Gulf  Coast 
17 — Upper  Cook  Inlet 
18 — San  Francisco  Bay 
19— NE  United  States 
20 — Sandhills  and  Rainwater  Basin 
21 — Playa  Lakes 


Figure  1.     Waterfowl  habitats  of  major  concern  (after  Tiner  1984). 


Waterfowl 


383 


and  justified  concern  has  been  the  importance  of 
wintering  habitat  (Fredrickson  and  Drobney  1979) 
and  its  rate  of  loss  reported  along  coastal  marshes 
(Tiner  1984:35).  Although  wetland  losses  are  not 
confined  to  the  areas  illustrated  in  Figure  1,  they 
represent  areas  of  high  use,  and  consequently  any 
effort  toward  their  preservation  or  enhancement 
would  pay  great  dividends  toward  waterfowl  welfare. 

One  of  the  annual  measures  of  potential  water- 
fowl productivity  has  been  an  index  to  the  number 
of  water  areas  along  the  aerial  transects  used  for 
breeding  birds.  It  has  been  assumed  that  good  water 
conditions  in  a  given  year  would  speak  well  for  the 
productivity  of  waterfowl  in  the  area.  However,  with 
ever-intensifying  agricultural  efforts  in  the  surround- 
ing uplands,  at  least  the  dabbling  ducks,  which  rely 
on  secure  upland  nesting  cover,  may  not  be  able 
to  respond  favorably  with  good  water  conditions 
only.  Higgins  (1977),  in  measuring  waterfowl  pro- 
duction in  an  area  of  intense  farming,  believed  that 
upland  nesting  ducks  could  not  reproduce  them- 
selves during  most  years  in  areas  that  were  85%  or 
more  tilled.  He  recorded  a  21%  nest  success  on 
his  study  area,  with  76%  of  the  failures  attributed  to 
predators  and  16%  to  farming  operations. 

In  the  northern  Great  Plains,  grazing  is  a  more 
prevalent  land  use,  and  where  grazing  is  not  exces- 


sive, waterfowl  nest  success  is  often  high  (77%, 
Smith  1953;  67%,  Rundquist  1973).  Of  the  two  ma- 
jor ingredients  to  successful  dabbling  duck  produc- 
tion, quality  water  areas  and  secure  upland  nesting 
habitat,  the  former  is  currently  at  a  premium  in  the 
northern  Great  Plains — the  latter  in  the  prairie  pot- 
hole area.  Thus  the  two  areas,  although  adjacent 
to  one  another,  offer  different  options  in  terms  of 
type  of  investment  and  return  for  effort  expended 
toward  enhancing  habitat. 

Intensive  agriculture,  although  detrimental  on 
the  breeding  habitat,  may  be  providing  at  least  an 
acceptable  substitute  for  some  food  sources  on  win- 
tering areas.  Fredrickson  and  Drobney  (1979)  sug- 
gested that  abundant  food  supplies  (domestic  grains) 
may  be  holding  some  populations  farther  north  than 
previously,  thereby  reducing  competition  on  tradi- 
tional (but  negatively  altered)  winter  habitat.  This  is 
particularly  true  for  the  mallard,  which  in  the  Co- 
lumbia Basin  in  Washington,  increased  from  a  win- 
tering population  of  86,000  in  the  entire  State  in 
1947  to  730,000  in  I960,  largely  as  a  result  of  addi- 
tional food  sources  from  irrigated  grain  farms  (Bell- 
rose  1978).  Baldassarre  and  Bolen  (1984)  suggested 
a  possible  increasing  importance  of  field-feeding  by 
wintering  waterfowl,  because  of  wetland  distribution, 
and  presented  management  recommendations  to- 
ward enhancement  of  this  activity. 


384 


Waterfowl 


LITERATURE  CITED 

ADDY,  C.E.  1964.  Atlantic  Flyway.  Pages  167-184  in  Lin 
duska,  J. P.  ed.  Waterfowl  Tomorrow.  U.S.  Govt.  Print. 
Off.,  Washington,  DC. 

ALLEN,  G.T.,  S.E.  FAST,  B.J.  LANGSTAFF,  D.W.  TOMRDLE, 
and  B.L.  TROUTMAN.  1 978.  Census  of  Canada  geese 
on  the  Palouse  River,  Washington,  during  the  spring 
of  1977.  The  Murrelet  59:96-100. 

AMERICAN  ORNITHOLOGISTS  UNION.  1983  Checklist 
of  North  American  birds,  Sixth  Edition.  Allen  Press, 
Inc.,  Lawrence,  KS. 

BALDASSARRE,  GA.  and  E.G.  BOLEN.  1984.  Field-feeding 
ecology  of  waterfowl  wintering  on  the  southern  high 
plains  of  Texas.  J.  Wildl.  Manage.  48:63-71. 

BARTONEK,  J.C.  and  J.J.  HICKEY.  1969.  Food  habits  of 
canvasbacks,  redheads  and  lesser  scaup  in  Manitoba. 
Condor  71:280-290. 

BELLROSE,  F.C.  1978.  Ducks,  geese  and  swans  of  North 
America.  The  Stackpole  Co.,  Harrisburg,  PA,  and 
Wildl.  Manage.  Inst.,  Washington,  DC.  540pp. 

,  KL.  JOHNSON,  and  T.U.  MEYERS.  1964.  Relative 

value  of  natural  cavities  and  nesting  houses  for  wood 
ducks.  J.  Wildl.  Manage.  28:661-676. 

BERG,  P.F.  1956.  A  study  of  waterfowl  broods  in  eastern 
Montana  with  special  reference  to  movements  and 
the  relationship  of  reservoir  fencing  to  production.  J. 
Wildl.  Manage.  20:253-262. 

BISHOP,  R.A.  and  R.  BARRATT.  1970.  Use  of  artificial  nest 
baskets  by  mallards.  J.  Wildl.  Manage.  34:734-738. 

BOSSENMAIER,  E.F.  and  W.H.  MARSHALL.  1958.  Field- 
feeding  by  waterfowl  in  southeast  Manitoba.  Wildl. 
Monogr.  1.  32pp. 

BREWSTER,  W.G.,  J.M.  GATES,  and  L.D.  FLAKE.  1976. 

Breeding  waterfowl  populations  and  their  distribution 
in  South  Dakota.  J.  Wildl.  Manage.  40:50-59. 

BUE,  I.G.,  H.G.  UHLIG,  and  D.J.  SMITH  1964.  Stock  ponds 
and  dugouts.  Pages  391-398  in  Linduska,  J.P.  ed., 
Waterfowl  Tomorrow,  U.S.  Govt.  Print.  Off.,  Washing- 
ton, DC.  770pp. 

CARLSEN,  T.L.  1984.  Waterfowl  nesting  in  two  ponds  of 
the  Canyon  Ferry  Wildlife  Management  Area,  Mon- 
tana. M.S.  Thesis.  Montana  State  Univ.,  Bozeman. 
91pp. 

CHABRECK,  R.H.  1979.  Winter  habitat  of  dabbling 

ducks — physical,  chemical,  and  biological  aspects. 
Pages  133-142  in  Bookhout,  T.A.  ed.,  Waterfowl  and 
Wetlands — An  Integrated  Review.  North-central  Sect., 
The  Wildl.  Soc,  Madison,  WI 

CHBLDRESS,  DA.  and  R.L.  ENG.  1979.  Dust  abatement 
project  with  wildlife  enhancement  on  Canyon  Ferry 
Reservoir,  Montana.  Pages  282-288  in  The  Mitigation 
Symposium:  a  National  Workshop  On  Mitigating 
Losses  of  Fish  and  Wildlife  Habitats.  Colorado  State 
Univ.,  Ft.  Collins. 

COWARDIN,  L.M.,  G.E.  CUMMINGS,  and  P.B.  REED,  Jr. 
1967.  Stump  and  tree  nesting  by  mallards  and  black 
ducks.  J.  Wildl.  Manage.  31:229-235. 

DELACOUR,  J.  and  E.  MAYR.  1945.  The  family  Anatidae. 
Wilson  Bull.  57:3-55. 

DIEM,  KL.  and  K.H.  LU.  I960.  Factors  influencing  water- 
fowl censuses  in  the  parklands.  Alberta,  Canada.  J. 
Wildl.  Manage.  24:113-133- 

DUEBBERT,  H.F.  1982.  Nesting  waterfowl  on  islands  in 

Lake  Audubon,  North  Dakota.  Wildl.  Soc.  Bull.  10:232- 
237. 


—  and  J.T.  LOKEMOEN.  1976.  Duck  nesting  in  fields 
of  undisturbed  grass-legume  cover.  J.  Wildl.  Manage. 
40:39-49. 
J.T.  LOKEMOEN,  and  D.E.  SHARP.  1983-  Concen- 


trated nesting  of  mallards  and  gadwalls  on  Miller  Lake 
Island,  North  Dakota.  J.  Wildl.  Manage.  47:729-740. 

DZUBIN,  A.  1969.  Assessing  breeding  populations  of  ducks 
by  ground  counts.  Pages  178-230  in  Saskatoon  Wet- 
lands Seminar.  Can.  Wildl.  Serv.  Rep.  Series  6. 

ENG,  R.L.,  J.D.  JONES,  and  F.M.  GJERSING  1979.  Con 

struction  and  management  of  stock  ponds  for  water- 
fowl. U.S.  Dep.  Inter.,  Bur.  Land  Manage.  Tech.  Note 
327.  Denver,  CO.  39pp. 

ERSKINE,  A.J.  1972.  Buffleheads.  Can.  Wildl.  Serv.  Monogr. 
Series  4.  240pp. 

EVANS,  CD.  and  KE.  BLACK  1956.  Duck  production 
studies  on  the  prairie  potholes  of  South  Dakota.  U.S. 
Dep.  Inter.,  Fish  and  Wildl.  Serv.  Spec.  Sci.  Rep.  Wildl. 
32.  59pp. 

EWASCHUK,  E.  and  DA.  BOAG.  1972.  Factors  affecting 
hatching  success  of  densely  nesting  Canada  geese.  J. 
Wildl.  Manage.  36:1097-1106. 

FRAYER,  W.E.,  T.J.  MONAHAN,  DC.  BOWDEN,  and  FA. 
GRAYBILL.  1983-  Status  and  trends  of  wetlands  and 
deep  water  habitats  in  the  conterminous  United 
States,  1950s  to  1970s.  Dep.  of  Forest  and  Wood  Sci- 
ences. Colorado  State  Univ.,  Ft.  Collins.  32pp. 

FREDRICKSON,  L.H.  and  R.D.  DROBNEY.  1979.  Habitat 
utilization  by  post-breeding  waterfowl.  Pages  119-131 
in  Bookhout,  T.A.  ed.,  Waterfowl  and  Wetlands — An 
Integrated  Review.  North-central  Sect.,  The  Wildl. 
Soc,  Madison,  WI. 

GIROUX,  J.F.  1981.  Use  of  artificial  islands  by  nesting 
waterfowl  in  southeastern  Alberta.  J.  Wildl.  Manage. 
45:669-679. 

,  D.E.  JELINSKI,  and  R.W.  BOYCHUK  1983-  Use  of 

rock  lands  and  round,  straw  bales  by  nesting  Canada 
geese.  Wildl.  Soc.  Bull.  11:172-177. 

GJERSING,  F.M.  1975.  Waterfowl  production  in  relation  to 
rest-rotation  grazing.  J.  Range  Manage.  28:37-42. 

GOLLOP,  J.G.  and  W.H.  MARSHALL.  1954.  A  guide  for 
aging  duck  broods  in  the  field.  Mississippi  Flyway 
Council  Tech.  Sect.  14pp.  (mimeo.). 

HAMMOND,  M.C.  and  G.E.  MANN.  1956.  Waterfowl  nest- 
ing islands.  J.  Wildl.  Manage.  20:345-352. 

HANSEN,  H.A.,  P.E.K  SHEPHERD,  J.G  KING,  and  W.A. 

TROYER.  1971.  The  trumpeter  swan  in  Alaska.  Wildl. 
Monogr.  26.  83pp 

HANSON,  W.C.  and  L.L.  EBERHARDT.  1971.  A  Columbia 
River  Canada  goose  population,  1950-1970.  Wildl. 
Monogr.  28.  6  lpp. 

HARAMIS,  G.M., JR.  GOLDSBERRY,  D.G.  MCAULEY,  and 
E.L.  DERLETH.  1985.  An  aerial  photographic  census  of 
Chesapeake  Bay  and  North  Carolina  canvasbacks.  J. 
Wildl.  Manage.  491:449-454. 

HEITMEYER,  ME.  and  PA.  VOHS,  Jr.  1984.  Distribution 
and  habitat  use  of  waterfowl  wintering  in  Oklahoma. 
J.  Wildl.  Manage.  48:51-62. 

HIGGINS,  KF.,  L.M.  KIRSCH,  and  I.J.  BALL.  1969.  A  cable- 
chain  device  for  locating  duck  nests.  J.  Wildl.  Manage. 
33(4):1009-1011. 

.  1977.  Duck  nesting  in  intensively  farmed  areas  of 

North  Dakota.  J.  Wildl.  Manage.  41:232-242. 

HILDEBRAND,  B.D.  1979.  Habitat  requirements  of  molt- 
ing Canada  geese  at  Lima  Reservoir,  Montana.  M.S. 
Thesis.  Montana  State  Univ.,  Bozeman.  79pp. 


Waterfowl 


385 


HINES,J.E.  and  G.J.  MITCHELL.  1983.  Gadwall  nest-site 
selection  and  nesting  success.  J.  Wildl.  Manage. 
47:1063-1071. 

HOBAUGH,  W.C  and  J.G.  TEER.  1981.  Waterfowl  use 

characteristics  of  flood-prevention  lakes  in  north-cen- 
tral Texas.  J.  Wildl.  Manage.  45:16-26. 

HUDSON,  M.S.  1983.  Waterfowl  production  in  three  age 
classes  of  stock  ponds  in  Montana.  J.  Wildl.  Manage. 
47.112-117. 

HUMBURG,  D.D.,  N.H.  PRINCE,  and  R.A.  BISHOP.  1978. 
The  social  organization  of  a  mallard  population  in 
northern  Iowa.  J.  Wildl.  Manage.  42:72-80. 

KAMINSKI,  R.M.  and  H.H.  PRINCE.  1977.  Nesting  habitat 
of  Canada  geese  in  southeastern  Michigan.  Wilson 
Bull.  89:523-531. 

KEITH,  L.B.  1 96 1 .  A  study  of  waterfowl  ecology  on  small 
impoundments  in  southeastern  Alberta.  Wildl.  Mon- 
ogr.  6.  88pp. 

KIRBY,  RE.  1980.  Waterfowl  production  estimates  on 
forested  wetlands  from  pair  and  brood  counts.  Wildl. 
Soc.  Bull.  8:273-278. 

KROHN,  W.B.  and  E.G.  BIZEAU.  1979.  Molt  migration  of 
the  Rocky  Mountain  population  of  the  western  Can- 
ada goose.  Pages  1 30- 1 40  in  Management  and  Biology 
of  Pacific  Flyway  Geese  Symp.  N.W.  Sect.  Wildl.  Soc. 
Portland,  OR. 

LEMIEUX,  L.  1959.  The  breeding  biology  of  the  greater 
snow  goose  on  Bylot  Island,  Northwest  Territories. 
Can.  Field-Nat.  73:117-128. 

LOKEMOEN,  J.T.,  H.F.  DUEBBERT,  and  D.E.  SHARP.  1984. 
Nest  spacing,  habitat  selection,  and  behavior  of  water- 
fowl on  Miller  Lake  Island,  North  Dakota.  J.  Wildl. 
Manage.  48:309-321. 

LOW,  J.B.  1945.  Ecology  and  management  of  the  redhead 
(Nyroca  americana)  in  Iowa.  Ecol.  Monogr.  15:35-69. 

MACINNES,  CD.  1962.  Nesting  of  small  Canada  geese 
near  Eskimo  Point,  Northwest  Territories.  J.  Wildl. 
Manage.  26:247-256. 

MAJ,  ME.  1983-  Analysis  of  trumpeter  swan  habitat  on  the 
Targhee  National  Forest  of  Idaho  and  Wyoming.  M.S. 
Thesis.  Montana  State  Univ.,  Bozeman.  102pp. 

MAYFIELD,  H.  1975.  Suggestions  for  calculating  nest  suc- 
cess. Wilson  Bull.  87:456-466. 

MENDALL,  H.L.  1958.  The  ring-necked  duck  in  the  North- 
east. Univ.  of  Maine  Bull.  60(16).  317pp. 

MILLER,  H.W.  and  D.H.  JOHNSON.  1978.  Interpreting  the 
results  of  nesting  studies.  J.  Wildl.  Manage.  42:471- 
476. 

MUNDINGER,  J.G   1976.  Waterfowl  response  to  rest- 
rotation  grazing.  J.  Wildl.  Manage.  40:60-68. 

NAGEL,  J.  1965.  Field-feeding  of  whistling  swans  in  north- 
ern Utah.  Condor  67:446-447. 

PATTERSON,  J.N.  1976.  The  role  of  environmental  hetero- 
geneity in  the  regulation  of  duck  populations.  J.  Wildl. 
Manage.  40:22-32. 


PERRY,  M.C.,  RE.  MUNRO,  and  G.M.  HARAMIS.  1981. 
Twenty-five-year  trends  in  Chesapeake  Bay  diving 
duck  populations.  Trans.  North  Am.  Wildl.  Nat.  Re- 
sour.  Conf.  46:299-310. 

POSPAHALA,  R.S.,  DR.  ANDERSON,  and  C.J.  HENNY. 

1974.  Population  ecology  of  the  mallard.  II.  Breeding 
habitat  conditions,  size  of  the  breeding  populations 
and  production  indices.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  Resour.  Publ.  115.  73pp. 

RUNDQUIST,  V.M.  1973.  Avian  ecology  on  stock  ponds  in 
two  vegetational  types  in  north-central  Montana. 
PH.D.  Thesis,  Montana  State  Univ.,  Bozeman.  116pp. 

RYDER,  J. P.  1967.  The  breeding  biology  of  Ross'  goose  in 
the  Perry  River  region,  Northwest  Territories.  Can. 
Wildl.  Serv.  Rep.  Series  3-  56pp. 

SAUDER,  D.W.,  T.L  LINDER,  R.B.  DAHLGREN,  and  W.L. 
TUCKER.  1 97 1 .  An  evaluation  of  the  roadside  tech- 
nique for  censusing  breeding  waterfowl.  J.  Wildl. 
Manage.  35:538-543. 

SHERWOOD,  G.A.  I960.  The  whistling  swan  in  the  West, 
with  particular  reference  to  Great  Salt  Lake  Valley, 
Utah.  Condor  62:370-377. 

SMITH,  R.H.  1953-  A  study  of  waterfowl  production  in 
artificial  reservoirs  in  eastern  Montana.  J.  Wildl.  Man- 
age. 17:276-291. 

STEWART,  RE.,  A.E.  GEIS,  and  CD.  EVANS.  1958.  Distri- 
bution of  populations  and  hunting  kill  of  the  canvas- 
back  J.  Wildl.  Manage.  22:333-370. 

STOUDT,  J.H.  1982.  Habitat  use  and  productivity  of  can- 
vasbacks  in  southwestern  Manitoba,  1961-72.  U.S. 
Dep.  Inter.,  Fish  and  Wildl.  Serv.  Spec.  Sci.  Rep.  Wildl. 
248.  31pp. 

SUGDEN,  L.G.  1973.  Feeding  ecology  of  pintail,  gadwall, 
American  widgeon,  and  lesser  scaup  ducklings.  Can. 
Wildl.  Serv.  Rep.  Series  24.  45pp. 

and  G  BUTLER.  1980.  Estimating  densities  of 

breeding  canvasbacks  and  redheads.  J.  Wildl.  Manage. 
44:814-821. 

TINER,  R.W.,  Jr.  1984.  Wetlands  of  the  United  States:  Cur- 
rent status  and  recent  trends.  U.S.  Dep.  Inter.,  Fish 
and  Wildl.  Serv.  Habitat  Resources.  Newton  Corner, 
MA.  59pp. 

VERMEER,  K  1970.  Some  aspects  of  the  nesting  of  ducks 
on  islands  in  Lake  Newell,  Alberta.  J.  Wildl.  Manage. 
34:126-129- 

,  D.R.M.  HATCH,  and  J. A.  WINDSOR.  1972.  Greater 

scaup  as  common  breeder  on  northern  Lake  Winni- 
peg. Can.  Field-Nat.  86:168. 

WELLER,  M.W.  1964.  Distribution  and  migration  of  the 
redhead.  J.  Wildl.  Manage.  28:64-103- 

,  D.L.  TRAUGER,  and  G.L  KRAPU.  1969.  Breeding 

birds  of  the  West  Mirage  Islands,  Great  Slave  Lake, 
Northwest  Territories,  Can.  Field  Nat.  83:344-360. 


386 


Waterfowl 


19 

COLONIAL 
WATERBIRDS 

Steven  M.  Speich 


Cascadia  Research  Collective 
Waterstreet  Building,  Suite  201 
218  1/2  W.  4th  Avenue 
Olympia,  WA  98501 


Editor's  Note:  Like  marsh  and  shorebirds,  colonial 
waterbirds  are  a  taxonomically  diverse  group,  yet 
they  require  similar  management  and  survey  tech- 
niques. Because  of  their  colonial  nesting  behavior, 
they  can  be  very  sensitive  to  impacts  and  distur- 
bances to  nesting  areas,  including  those  caused  by 
biologists  conducting  inventories  and  monitoring 
studies.  This  chapter  provides  guidelines  on  survey 
techniques  and  methods  to  minimize  impacts  to 
the  birds  from  such  efforts. 


INTRODUCTION 

Any  bird  that  predominantly  feeds  in  aquatic 
systems,  marine,  fresh  water,  or  both,  and  tends  to 
nest  in  groups  of  closely  associated  nests,  is  consid- 
ered a  colonial  waterbird  (Table  1 ).  Some  species 
nest  both  in  groups  and  singly,  and  all  colonial  nest- 
ing birds  nest  as  solitary  pairs  at  times.  Discussions 
of  the  biology  of  colonial  waterbirds  are  found  in 
Emlen  and  DeMong  (1975),  Evans  (1982),  Krebs 
(1974,  1978),  Kushlan  (1978),  and  Pratt  (1980). 
Waterfowl  (Anseriformes)  are  generally  not  thought 
of  as  colonial  nesting  waterbirds  and  therefore,  are 
not  discussed  in  this  chapter. 

Species  of  colonial  waterbirds  range  from  small 
storm  petrels  to  large  pelicans.  Some  species,  such  as 
storm  petrels,  lead  a  truly  pelagic  existence,  except 
when  they  come  to  offshore  rocks  and  islands  to 
nest,  often  in  dense  colonies.  Other  colonial  nesting 
species  (alcids  and  cormorants,  brown  pelican  [Pele- 
canus  occidentalism  and  some  gulls  and  terns)  are 
less  pelagic,  with  coastal  and  nearshore  affinities 
throughout  the  year,  or  at  least  portions  of  the  year. 
A  few  species  bridge  the  gap  between  marine  and 
fresh-water  environments,  with  at  least  segments  of 
the  species'  population  feeding  in  marine  or  fresh 
water  systems.  Indeed,  some  individuals  (some  gulls 
and  terns,  great  blue  heron  [Ardea  herodias],  and 
double-crested  cormorant  [Phalacrocorax  auritus]) 
feed  in  both  systems  while  nesting.  And  still  other 
species  are  truly  specific  to  fresh  water  while  nest- 
ing, i.e.,  grebes;  white  pelican  (P.  erythrorhynchos); 
and  some  egrets,  gulls,  and  terns. 

There  are  several  reasons  why  it  is  important  to 
survey  and  census  colonial  nesting  waterbirds.  Most 
species  are  near  or  at  the  top  of  food  chains  and 
are  thus  sensitive  to  the  health  of  fresh-water  and 
marine  ecosystems.  Changes  or  reductions  in  food 
chain  biomass  and  species  composition  often  cause 
stress  in  colonial  waterbirds,  expressed  as  failure 
to  breed,  abandonment  of  eggs  and  young,  late  nest- 
ing, depressed  growth  rates,  or  reduced  fledging 
success.  Over  long  periods,  populations  may  decline; 
this  can  be  monitored  through  a  censusing  program. 


Colonial  Waterbirds 


387 


Table  1.     Habitat  and  nesting  characteristics  of  colonial  waterbirds  found  in  western  North  America. 


Family  and  Species 

Habitat 

Nest  Type  and  Substrate 

Nest  Location 

Grebes  (Podicipedidae) 

Red-necked  (Podiceps 

coastal  waters,  lakes 

floating,  reeds 

surface 

grisegena) 

Eared  (P.  nigricollis) 

lakes,  bays,  ocean 

floating,  reeds 

surface 

Western  (Aechmophorus 

lakes,  sloughs,  bays 

floating,  reeds 

surface 

occidentalis) 

Fulmars  (Procellariidae) 

Northern  (Fulmarus  glacialis) 

ocean 

herbage,  dried  grass 

seal  cliffs 

Storm  Petrels  (Hydrobatidae) 

Fork-tailed  (Oceanodroma 

ocean 

burrow 

sea  islands 

furcata) 

Leach's  (0.  leucorhoa) 

ocean 

burrow 

sea  islands 

Ashy  (0.  homochroa) 

ocean 

under  rocks 

sea  islands 

Black  (0.  melania) 

ocean 

burrow 

sea  islands 

Pelicans  (Pelecanidae) 

American  White 

lakes,  marshes, 

on  ground  or 

surfaces,  bare 

(Pelecanus  erythrorhynchos) 

salt  bays,  beaches 

in  bulrushes 

ground 

Brown  (P.  occidentalis) 

ocean,  salt  bays 

on  ground 

islands 

Cormorants  (Phalacrocoracidae) 

Double-crested  (Phalacrocorax 

coast,  bays, 

sticks,  brush, 

trees,  marshes, 

auritus) 

lakes,  rivers 

seaweed 

islands,  or  cliffs 

Brandt's  (P.  penicillatus) 

ocean,  coast,  littoral 

seaweed,  sticks 

rocky  island,  cliffs 

Pelagic  (P.  pelagicus) 

coast,  bays,  sounds 

seaweed,  grass 

ledges,  cliffs 

Red-faced  (P.  urile) 

ocean,  coast 

seaweed,  grass 

rocky  islands 

Herons  and  Egrets  (Ardeidae) 

Great  Blue  Heron 

marshes,  swamps, 

sticks 

trees,  rocky  islands 

(Ardea  herodias) 

streams,  shores, 
tideflats,  ditches 

Great  Egret  (Casmerodius 

marshes,  ponds, 

sticks,  brush 

trees,  brush  over 

albus) 

shores 

water 

Snowy  Egret  (Egretta  thula) 

marshes,  ponds, 

sticks,  tules, 

willow  thickets  in 

tideflats 

bushes  in  marsh 

marshes 

Cattle  Egret  (Bubuleus  ibis) 

cow  pastures,  fields 
and  marshes 

sticks 

in  bush  or  trees 

Black-crowned  Night  Heron 

marshes,  lake  margins, 

sticks,  stalks 

trees,  willow 

(Nycticorax  nycticorax) 

shores 

thickets  in 
marshes 

Ibises  (Threskiornithidae) 

Glossy  Ibis  (Plegadis  falcinellus) 

marshes,  lake  margins 

sticks,  stalks 

bushes  in  marshy 
areas 

Gulls  and  Terns  (Laridae) 

Laughing  Gull  (Larus  atricilla) 

salt  water  marshes, 
bays 

grass  platforms 

marsh  islands 

Franklin's  Gull  (L.  pipixcan) 

prairies,  marshes,  lakes 

among  reeds 

prairie  marshes 

Bonaparte's  Gull 

ocean,  bays,  rivers, 

in  conifers 

wooded  muskeg 

(L  Philadelphia) 

lakes 

Mew  Gull  (L.  canus) 

coasts,  rivers,  lakes 

rocks  or  trees 

marsh  or  beach 

Ring-billed  Gull 

coast,  estuaries,  rivers, 

dried  grass, 

islands  in  lakes 

(L  delawarensis) 

lakes,  fields,  refuse 
dumps 

weeds,  sticks 

California  Gull  (L.  californicus) 

coast,  beach,  lake, 
river,  farmland,  cities 

on  ground 

islands  in  lakes 

Herring  Gull  (L.  argentatus) 

ocean  coast,  bays, 

grass  or  seaweeds 

ground  or  cliff, 

beaches,  lakes, 

islands 

farmlands,  dumps 

388 


Colonial  Waterbirds 


Table  1.     Habitat  and  nesting  characteristics  of  colonial  waterbirds  found  in  western  North  America 
(concluded). 


Family  and  Species 

Habitat 

Nest  Type  and  Substrate 

Nest  Location 

Gulls  and  Terns  (Laridae)  (continued) 

Thayer's  Gull  (L  thayeri) 

ocean  coast,  bays, 
beaches 

vegetation  on  rocks 

on  ground 

Western  Gull  (L.  Occidentalis) 

coastal  waters, 

grass 

offshore  islets, 

estuaries,  beaches 

seacliffs 

Glaucous-winged  Gull 

ocean,  coast,  bays, 

seaweed  or  grass 

coastal  islands, 

(L.  glaucescens) 

beaches,  piers, 
dumps,  waterfronts 

headlands 

Black-legged  Kittiwake 

ocean 

seaweed 

ledges  of  seacliffs 

(Rissa  tridactyla) 

Red-legged  Kittiwake 

ocean 

seaweed  or  grass 

ledges  of  seacliffs 

(R.  brevirostris) 

Sabine's  Gull  (Xema  sabini) 

ocean,  tundra 

grass 

depressions  in  low 

(summer) 

tundra 

Gull-billed  Tern  (Sterna  nilotica) 

lakes,  marshes,  fields 

surface  depression 
lined  with  shells 
and  straw 

sandy  islands 

Caspian  Tern  (S.  caspia) 

lakes,  bays,  coast 

large,  deep  hollows 

sandy  islands  or 

lined  with  sticks 

dikes  of  lakes 

and  shells 

Elegant  Tern  (S.  elegans) 

coast 

scrape  on  sand 

islands 

Common  Tern  (S.  hirundo) 

ocean,  bays,  beaches, 
lakes  (summer) 

surface  depression 

lake  islands 

Arctic  Tern  (S.  paradisaea) 

ocean,  coast, 

surface  depression 

island,  beach, 

tundra  lakes  (sum- 

tundra 

mer) 

Aleutian  Tern  (S.  aleutica) 

ocean,  coast 

depression  in  moss 

islands 

Forster's  Tern  (S.  forsteri) 

ocean,  lakes,  bays, 

marsh  vegetation 

marshes  or  on 

marshes,  beaches 

muskrat  houses 

Least  Tern  (S.  antillarum) 

ocean,  bays,  beaches, 

scrape 

sandy  beaches  or 

estuaries 

gravel  bars 

Black  Tern  (Chlidonias  niger) 

coasts,  lakes,  fresh- 

marsh vegetation 

floating  or 

water  marshes 

on  muskrat 
houses 

Murres,  Guillemots,  Auks,  etc.  (Alci- 

dae) 

Common  Murre  (Uria  aalge) 

ocean,  large  bays 

bare  rock  or 

bare  ledges,  cliffs, 

loose  soil 

or  rocky  islands 

Thick-billed  Murre  (U.  lomvia) 

ocean 

bare  ledges 

bare  rock,  sea  cliffs 

Black  Guillemot 

ocean,  rocky  coast 

crevices,  burrows 

sea  cliffs 

(Cepphus  grylle) 

Pigeon  Guillemot 

ocean,  rocky  coast 

crevices,  burrows 

sea  cliffs 

(C.  columba) 

Xantu's  Murrelet 

ocean 

rock  crevices 

sea  islands 

(Synthliboramphus  hypolecus) 

Ancient  Murrelet  (S.  antiquus) 

ocean,  sounds 

burrows  or  bare  rock 

sea  islands 

Cassin's  Auklet 

ocean 

crevices  or  burrows 

sea  islands 

(Ptychoramphus  aleuticus) 

Parakeet  Auklet 

ocean 

crevices 

sea  cliffs 

(Cyclorrhynchus  psittacula) 

Least  Auklet  (Aethia  pusilla) 

ocean 

under  rocks 

sea  islands 

Wiskered  Auklet  (A.  pygmaea) 

ocean 

crevices 

sea  cliffs 

Crested  Auklet  (A.  cristatella) 

ocean,  tide-rip 

crevices 

sea  cliffs 

Rhinoceros  Auklet 

ocean,  tide-rip 

burrows 

sea  islands 

(Cerorhinca  monocerata) 

Tufted  Puffin 

ocean 

burrows 

sea  islands  or 

(Fratercula  cirrhata) 

headlands 

Horned  Puffin  (F.  corniculata) 

ocean 

burrows  or  crevices 

sea  cliffs 

Colonial  Waterbirds 


389 


Colonial  nesting  waterbirds  have  served,  coinci- 
dently,  as  indicators  of  man-made  contaminants  in 
the  environment.  The  effects  have  included  repro- 
ductive failure,  which  can  be  detected  through  care- 
ful censusing  and  surveying  programs.  Knowledge 
of  the  numbers  and  locations  of  colonies  provides  an 
immediate  source  of  biological  materials  for  the  geo- 
graphical monitoring  of  the  levels  and  types  of  pollu- 
tants in  the  environment  and  their  subtle  effects  on 
body  and  egg  tissues. 

Information  that  can  be  obtained  from  censusing 
and  surveying  programs,  includes  quantification  of 
the  numbers  and  geographic  locations  of  colonies, 
numbers  and  species  present  in  each  colony,  and 
habitats  occupied  by  species  and  colonies.  Collecting 
quantified  information  about  colonies  for  long  pe- 
riods, decades  and  preferably  longer,  allows  a  mean- 
ingful monitoring  of  populations.  There  is  often  a 
large  variation  in  reproductive  success  from  year  to 
year.  Long-term  population  trends  cannot  be  reliably 
derived  from  short-term  studies  (those  lasting  5 
years  or  less). 

Determining  the  numbers  of  waterbirds  is  more 
efficient  when  they  are  concentrated  at  breeding 
sites.  During  the  nonbreeding  portion  of  the  year, 
the  individuals  of  most  species  are  dispersed  over 
large  geographic  areas,  such  as  all  or  part  of  North 
America,  Central  and  South  Americas,  the  continental 
margins,  or  the  Pacific  Ocean.  Even  when  numbers 
are  determined  away  from  breeding  colonies  in  re- 
mote areas,  the  origin  of  observed  birds  is  almost 
always  unknown.  The  efficiency  and  accuracy  obtain- 
able from  censusing  breeding  colonies  is  usually 
superior  to  those  from  censusing  wintering  colonies. 

The  material  presented  here  is  intended  to  aid 
investigators  in  the  design  and  execution  of  census- 
ing, surveying,  and  monitoring  programs.  Primarily, 
the  biology  of  colonial  nesting  waterbirds  relevant  to 
such  programs,  measurement  techniques,  examples 
and  comparisons  of  techniques  actually  applied,  and 
limitations  on  observers  of  colonial  nesting  water- 
birds  relevant  to  disturbance  and  its  effects  are  dis- 
cussed. These  materials  are  meant  to  guide  potential 
users  in  formulating  specific  programs,  according 
to  species,  nesting  sites,  and  program  goals. 


COLONIAL  WATERBIRD  BIOLOGY 

Several  aspects  of  the  biology  of  colonial  nesting 
waterbirds  are  both  unique  and  important.  Since 
most  of  these  aspects  vary  greatly,  it  is  helpful  to  be 
aware  of  them.  Knowledge  of  these  aspects  of  biol- 
ogy will  improve  the  design  and  conduct  of  any 
inventory  or  monitoring  study.  The  following  aspects 
of  the  biology  of  colonial  nesting  waterbirds  are 
relevant  to  study  design  and  execution. 


Colony  Size 

In  most  cases,  it  is  not  possible  to  make  a  defini- 
tive statement  of  the  exact  size  of  a  given  colony. 
The  failure  of  nesting  birds,  and  the  occurrence  of 
renesters  and  late  nesters,  makes  any  determination 
imprecise.  Thus,  any  statement  of  colony  size  must 
be  accompanied  by  qualifying  statements,  such  as 
survey  date,  stage  of  nesting,  and  timing  of  the  sur- 
vey in  relation  to  the  colony  nesting  cycle.  If  precise 
information  on  any  of  these  nesting  aspects  is 
known,  it  should  accompany  the  colony  census  data. 

For  each  species,  there  is  a  range  in  the  number 
of  active  nests  found  in  its  colonies.  Solitary  nesting 
pairs  occur  in  most  species.  The  upper  colony  size 
limit  varies  with  each  species.  If  a  histogram  of  col- 
ony size  were  constructed  for  each  species,  all 
would  be  different.  The  frequency  distribution  of 
colony  size  is,  more  or  less,  species  specific  (Black- 
lock  et  al.  1978;  Buckley  and  Buckley  1972;  Des- 
Granges  and  Laporte  1979,  1981;  DesGranges  et  al. 
1979;  Sowls  et  al.  1978,  1980;  Speich  and  Wahl 
1985;  Thompson  1982;  Werschkul  et  al.  1976, 
1977). 

Colony  size  varies  geographically  in  some  spe- 
cies. In  marine  species,  a  good  example  is  the  com- 
mon murre  {Uria  aalge).  In  the  Pacific  Ocean  basin, 
it  breeds  along  the  continental  margin  from  central 
California  to  Alaska.  Colonies  in  California,  Oregon, 
and  Washington  usually  number  hundreds,  thou- 
sands, or  low  tens-of-thousands.  Colonies  of  just  a 
few  pairs  or  tens-of-pairs  are  present.  However,  in 
Alaska,  colonies  are  known  to  contain  up  to 
hundreds-of-thousands  of  birds.  A  similar  pattern  is 
present  in  Leach's  storm  petrel  (Oceanodroma  leu- 
corhoa),  and  to  a  lesser  extent,  tufted  puffin  (Frater- 
cula  cirrhata).  However,  other  species  do  not  show 
geographic  variation  in  colony  size  along  the  same 
coast  line:  double-crested  cormorant,  pelagic  cormo- 
rant (Phalacrocorax pelagicus),  glaucous-winged 
gull  (Larus  glaucescens),  and  pigeon  guillemot  (Cep- 
phus  columba)  (Sowls  et  al.  1978,  1980;  Speich 
and  Wahl  1985). 

Inland  fresh-water  species  also  show  a  range  in 
colony  size,  but  the  upper  limit  is  dramatically  lower 
than  the  large  colonies  of  marine  birds  (English 
1978). 

Colony  Size  Stability 

The  number  of  nesting  individuals  of  a  species 
at  a  site  varies  over  time.  To  document  this,  it  is 
necessary  to  determine  the  number  of  nests  or  nest- 
ing adults  at  a  site  with  the  same  methods  at  the 
same  time  in  the  nesting  cycle  over  several  years, 
preferably  at  least  a  decade  (Erwin  et  al.  1981 ).  In 
general,  survey  data  are  lacking  for  most  species  and 
most  sites  over  long  periods  of  time,  seldom  allow- 


390 


Colonial  Waterbirds 


ing  analysis  of  colony  site  trends.  Data  are  usually 
fragmentary  and  imprecise.  There  are  examples  of 
species  colonizing  new  sites,  ceasing  to  nest  at  sites, 
and  switching  sites  (Ainley  and  Lewis  1974;  Buckley 
and  Buckley  1980;  Conover  and  Conover  1981; 
DesGranges  1980;  Drent  and  Guiguet  1961;  Drent  et 
al.  1964;  Nettleship  1975;  Ogden  1978;  Penland 
1982;  Pratt  1972,  1974;  Speich  and  Wahl  1985; 
Stowe  1982;  Thompson  et  al.  1979;  Tuck  1961). 

Nesting 

Of  critical  importance  to  the  censusing  and 
monitoring  of  colonial  nesting  species  is  the  knowl- 
edge of  the  time  of  nesting  and  the  breeding  syn- 
chrony of  the  subject  species,  as  a  whole,  and  its 
colonies.  The  time  of  breeding  and  synchrony  are 
distorted  by  renesting  and  late  nesting  pairs.  Al- 
though the  time  of  nesting  at  any  site  tends  to  take 
place  at  the  same  time  each  year,  there  is,  nonethe- 
less, variation  in  the  time  of  nesting  (Bayer  and 
McMahon  1981,  Drent  1965;  Ogden  1978;  Pratt 
1970,  1972,  1974;  Thompson  and  Tabor  1981;  Tuck 
1961 ).  At  times  this  can  vary  as  much  as  a  month 
(Drent  et  al.  1964;  Speich  and  Wahl  1985).  A  few 
studies  have  documented  geographical  variation  in 
the  time  of  nesting.  Even  colonies  close  together  can 
be  out  of  phase  (Bailey  and  Terman  1983;  Bayer 
and  McMahon  1981;  Speich  and  Wahl  1985;  Thomp- 
son and  Tabor  1981;  Tuck  1961;  Werschkul  et  al. 
1977;  pers.  obs.). 

Renesting.  Generally  egg  laying  is  clumped  and  can 
be  described  statistically  by  calculating  mean, 
variance,  range,  etc.  However,  some  nesters  fail 
through  loss  of  nest,  eggs,  young,  or  even  mates.  The 
number  of  failed  nests  can  range  from  a  small 
percentage  to,  at  times,  the  entire  colony. 


Often  the  individuals  that  failed  in  their  initial 
attempt  will  renest  later  in  the  same  season.  After 
nest  failure,  there  is  a  recycle  time  before  they  at- 
tempt to  renest.  The  re-entering  of  these  individuals 
into  the  colony  prolongs  the  active  nesting  phase 
of  the  colony  and  negates  the  usual  pronounced 
breeding  synchrony  observed  at  the  onset  of  nesting 
in  the  colony. 

Non-nesters.  In  most  populations  there  is  a 
segment  that  does  not  breed.  It  may  be  quite  small, 
or  sizable  and  measurable,  depending  on  the  species. 
These  non-nesting  individuals  may  not  be  present 
in  the  nesting  colony  at  all,  or  may  be  present  for 
only  a  short  time.  Generally,  non-nesters  are  thought 
to  be  young  inexperienced  birds,  without  previous 
nesting  experience  (Rodgers  1978).  However, 
occasionally  yearling  birds  do  attempt  to  nest  (Lowe 
1954;  Pratt  1973;  Rodgers  1978;  Speich  and 
Manuwal  1974). 

Late  Nesters.  Although  nesting  colonies  often 
initially  appear  to  be  synchronous  in  nesting,  this 
frequently  is  dispelled  by  the  arrival  of  late  nesting 
individuals.  The  numbers  can  be  just  a  few  pairs 
or  so  numerous  that  they  double  the  colony  size,  or 
more  (Rodgers  1978).  The  appearance  of  late 
nesting  birds  is  often  gradual,  or  in  mass  (Massey 
and  Atwood  1981). 

Mixed  Species  Colonies.  In  mixed  species 
colonies,  nesting  species  are  often  out  of  phase  with 
one  another.  This  is  usually  preceded  by  different 
arrival  times  for  each  species  (Burger  1978;  Drent  et 
al.  1964;  McCrimmon  1978a;  Thompson  1981; 
Thompson  and  Tabor  1981;  pers.  obs.). 


A  colony  of  adult  nesting  pelicans. 


A  mixed-species  colony  consisting  of  great  blue  herons, 
double-crested  cormorants,  and  black-crowned  night  her- 
ons. 


Colonial  Waterbirds 


391 


Attendance  Patterns 

When  censusing  colonial  nesting  species  and 
counting  adults,  it  is  critical  to  determine  daily  col- 
ony attendance  patterns  of  the  nesting  adults  be- 
cause the  proportion  of  nesting  adults  in  the  colony 
varies  with  the  time  of  day,  the  stage  of  nesting, 
and  other  environmental  patterns  (Birkhead  1978; 
Brun  1972;  Burger  1976;  Conover  and  Miller  1980; 
Custer  and  Osborn  1978;  Drent  1965,  1967;  Galusha 
and  Amlaner  1978;  Gibson  1950;  Lloyd  1972;  Manu- 
wal  1974b;  Patterson  1965;  Slater  1980;  Tuck  1961). 


Distribution  and  Habitat 

When  working  with  colonial  nesting  waterbirds, 
two  aspects  of  distribution  should  be  considered: 
( 1 )  the  distribution  of  colony  sites  over  often  broad 
geographic  regions  and  (2)  the  distribution  of  the 
nests  in  a  colony. 

The  geographic  distribution  of  a  species'  colo- 
nies may  cover  a  considerable  area,  such  as  large 
portions  of  western  North  America  (Godfrey  1966; 
Palmer  1962;  Ryder  1978;  Sowls  et  al.  1978,  1980; 
Speich  and  Wahl  1985;  general  bird  guides).  The 
distribution  of  colony  sites  within  overall  ranges  may 
evidence  one  or  more  distribution  patterns.  Colonies 
may  be  distributed  non-randomly,  with  clusters  near 
lakes,  streams,  or  coastlines  (Drent  and  Guiguet 
1961;  English  1978;  Sowls  et  al.  1978,  1980;  Speich 
and  Wahl  1985).  Within  an  area,  such  as  a  corridor 
along  a  river  or  around  a  lake,  the  nests  there  may 
be  evenly  or  randomly  spaced.  Knowledge  of  the 
study  species'  distribution  pattern  is  important  to 
any  plan  to  locate  colony  sites  for  all  programs. 

The  geographic  distribution  of  a  species  is  par- 
tially determined  by  the  interactions  of  its  ability 


to  reach  new  areas,  the  extent  it  is  a  specialist  or 
generalist  in  feeding  and  nesting  requirements,  and 
the  extent  of  its  preferred  habitat(s)  (Udvardy 
1969). 

Colony  Site  Selection 

The  factors  involved  in  colony  site  selection  are 
only  partially  understood  and  undoubtedly  vary  with 
each  species.  They  may  include  ( 1 )  suitable  nest 
site  substrate  or  support  for  nest  placement,  (2) 
shelter  from  adverse  environmental  conditions,  (3) 
apparent  freedom  from  predators,  and  (4)  proximity 
to  food  sources  (English  1978;  Buckley  and  Buckley 
1972).  Individuals  of  colonial  nesting  species  usually 
nest  in  colonies  with  conspeciflcs  and  often  with 
other  species  as  well,  while  suitable  nearby  sites  are 
often  left  unused  with  all  individuals  in  the  area  nest- 
ing at  one  or  a  few  sites  (Weise  1978).  Behavioral 
factors  may  be  as  important  as  those  of  selecting 
nest  sites  (Buckley  and  Buckley  1980). 

Nest  Site  Selection 

Nest  density  varies  with  each  species.  Inter-nest 
(how  far  apart  nests  are  constructed)  distance  is 
the  result  of  nest  site  selection  and  appears  to  be 
affected  by  nest  site  structure  as  it  relates  to  habitat 
suitability  and  social  interactions  (Buckley  and  Buck- 
ley 1977;  Burger  1978;  McCrimmon  1978a). 

Habitat  Suitability.  The  placement  of  a  nest 
ultimately  depends  on  whether  the  substrate  is 
appropriate  for  a  particular  species'  nests.  The  nests 
of  some  species  are  found  in  several  habitat  types; 
the  habitat  structure  for  nest  placement  is  apparently 
important  (Burger  1978).  Great  blue  heron  nests 
are  found  on  the  ground,  cliff  ledges,  in  bushes  and 
shrubs,  deciduous  trees,  coniferous  trees,  dead  trees, 
and  man-made  structures  (DesGranges  1979;  Henny 
and  Kurtz  1978;  Werschkul  et  al.  1976).  In  this 


SM 


Typical  white  pelican  colony. 


Great  blue  heron  nests  in  cottonwood  grove. 


392 


Colonial  Waterbirds 


species,  if  the  substrate  will  support  the  nest,  it  can 
be  used  for  nesting.  Finding  suitable  places  that 
support  a  nest  in  a  colony  limit  in  part  nest 
placement,  nest  density,  and  colony  size.  For  most 
species  that  nest  in  vegetation,  the  species  of  plants 
used  are  not  necessarily  important,  as  long  as  the 
structures  of  the  plants  meet  the  support  and  shelter 
requirements  of  the  species'  nest  (Bailey  and  Terman 
1983;  Bayer  and  McMahon  1981;  Werschkul  et  al. 
1976). 

Burrow  nesting  species  require  a  substrate  that 
allows  the  digging  of  a  burrow,  yet  will  not  collapse 
once  the  burrow  is  complete.  In  rocky  soil,  density 
may  be  limited  by  a  scarcity  of  "clean"  soil  that  al- 
lows burrowing.  Talus  slopes  contain  only  a  limited 
number  of  openings  that  lead  deep  enough  into  the 
slope  to  afford  nesting  sites  for  alcids  and  storm- 
petrels  (pers.  obs.). 


Social  Interactions.  The  presence  of  a  species 
nesting  in  a  colony  may  affect  the  choice  of  nest 
sites  of  another  species  (Burger  1978).  If  all  suitable 
nest  sites  are  taken,  other  individuals  (conspecincs 
or  otherwise  )  have  four  choices:  ( 1 )  displace  the 
occupants  of  a  nest  site,  (  2 )  wait  until  a  site  is  open, 
(3)  nest  in  a  less  favorable  site,  or  (4)  go  elsewhere 
(Thompson  1981). 

Inter-nest  distance  is  determined  in  part  by  so- 
cial interaction  (territory  defense)  with  conspecincs 
(Buckley  and  Buckley  1977;  Meyerriecks  I960)  or 
other  species  (Burger  1978).  The  space  between 
nests  is  open  if  a  new  nest,  territory,  can  "fit"  be- 
tween existing  nests. 

The  density  of  burrowing  birds,  such  as  Cassin's 
auklet,  is  limited  by  aggressive  interactions  on  the 
colony  surface  (Manuwal  1974a,  b;  Thoresen  1964). 
Burrow  occupants  vigorously  defend  the  burrow 
opening  and  small  area  around  the  burrow  entrance. 
Glaucous-winged  gulls  and  western  gulls  (L.  occiden- 
talis)  nest  in  higher  densities  where  the  substrate 
surface  is  irregular  with  many  large  objects  on  the 
ground.  These  are  visual  barriers  and  allow  closer 
packing  of  the  nests.  Herons  and  egrets  nest  in  lower 
densities,  allowing  for  large  body  size  and  the  reach 
of  long  necks  of  birds  at  their  nests  (Burger  1978). 
Pelecaniformes  defend  nest  sites  and  a  small  area 
around  nests  (Van  Tets  1965),  probably  resulting 
from  the  observed  nest  densities  in 
colonies. 


Nest  Types,  Placement,  and  Habitat 

Colonial  nesting  waterbirds  in  western  North 
America  construct  a  variety  of  nests  and  place  them 
in  characteristic  locations  and  habitats  depending  on 
the  needs  of  the  particular  species. 


Floating  Platform.  Floating  nests  are  generally 
constructed  in  marshes  or  areas  of  slow  moving 
water.  The  nests  are  attached  to  the  edge  of  patches 
of  anchored,  emergent  vegetation;  areas  of  variable- 
plant  density;  and  often  in  thick  stands  of  emergent 
vegetation.  The  visibility  of  nests  varies,  with  edge 
nests  easily  observable,  while  nests  in  areas  of  dense 
vegetation  are  often  visible  only  when  the  observer 
is  next  to  the  nest.  Nests  in  dense  stands  are  often 
obscured  from  sight  from  the  air;  the  nests  are  built 
with  vegetation  forming  a  natural  overhead  shelter. 
(For  species  nest  descriptions  see  Bent  1922; 
Godfrey  1966;  Palmer  1962.) 

Subsurface  Chamber.  Subsurface  nests  are  used  or 
constructed  by  a  variety  of  marine  birds.  (For 
species'  nest  descriptions  see  Bent  1919;  Palmer 
1962;  Sowls  et  al.  1978.)  In  some  cases,  nesting 
chambers  are  naturally  occurring.  The  sites  range 
from  talus  slopes,  rubble  piles,  boulder  piles,  sea 
caves,  driftwood,  various  holes  and  cracks  in  cliffs, 
and  burrows  dug  by  other  species  (Manuwal  1974b). 
Existing  holes  are  often  extended  and  enlarged,  but 
at  least  usually  cleaned  out.  Birds  nesting  in  natural 
cavities  are  often  particularly  difficult  to  census  and 
the  numbers  present  in  a  colony  cannot  always  be 
determined  from  surface  features  (Byrd  et  al.  1983; 
Sowls  et  al.  1978). 

In  other  cases,  nesting  chambers  are  con- 
structed. For  instance,  many  species  of  marine  birds 
excavate  burrows  for  nesting.  Chambers  are  often 
excavated  new  each  season  or  used  for  several  nest- 
ing seasons.  Burrow  sites  vary  with  each  species 
and  show  considerable  range  in  some  species.  Sites 
must  allow  burrowing  and  be  stable  enough  to  per- 
mit the  structure  to  survive  the  nesting  season.  Bur- 
rows vary  in  length,  shape,  and  depth,  depending 
on  the  species  and  the  nesting  substrate.  Surface 
cover  is  not  critical  providing  the  vegetation  can  be 
penetrated. 


Glaucous  gull. 


Colonial  Waterbirds 


393 


Surface  Nests.  Nests  placed  on  the  surface  of  the 
ground  range  from  simple,  virtually  no  nest  at  all,  to 
large  complex  structures.  (For  species'  nest 
descriptions  see  Bent  1919,  1921,  1922,  1927; 
Godfrey  1966;  Palmer  1962;  others.) 

The  simplest  nests  are  in  reality  no  nest  at  all, 
with  the  egg  placed  directly  on  the  substrate.  This  is 
seen  in  the  ledge  nesting  common  murre  and  thick- 
billed  murre  (U.  lotnvia).  The  murres  bring  a  few 
small  stones  and  other  objects  to  the  nest  site  (Tuck 
1961).  Birds  nesting  on  ledges  are  usually  highly 
visible  at  the  proper  angle  of  observation. 

The  simplest  "nest"  of  ground  nesting  birds  is 
the  scrape,  a  small  hollow  made  in  the  soil,  often 
sand.  This  is  common  in  terns,  such  as  the  Caspian 
tern  {Sterna  caspia;  Penland  1981).  These  nests  are 
precarious,  and  are  often  filled  with  drifting  sand. 
Nests  of  this  type  are  difficult  to  see  from  a  distance, 
especially  from  aircraft.  The  birds  themselves  are 
counted,  unless  the  colony  is  entered. 

The  next  order  of  nest  complexity  is  seen  in  the 
small  rings  of  material  that  are  constructed  of  gath- 
ered vegetation  and  other  debris.  These  nests  have  a 
bottom  and  a  rim  that  hold  the  eggs  in  place  and 
off  the  substrate.  These  are  commonly  seen  in  gulls 
and  cormorants.  Nests  of  sticks,  such  as  those  built 
by  cormorants,  often  survive  the  winter  and  are  used 
again,  with  new  material  added.  Other  rim  nests, 
made  of  more  substantial  non-woody  plant  material 
often  survive  to  be  used  again.  Many  species,  such  as 
the  kittiwakes,  compact  the  nest  with  mud  giving 
permanency  to  the  nest.  Nests  of  these  types  are 
often  clearly  visible  and  are  often  easily  counted 
when  the  observer  is  remote  from  the  colony. 

Finally,  in  some  species  the  nest  can  become  a 
sizeable  structure.  Such  nests  usually  are  the  result 
of  use  over  several  years,  with  each  nesting  pair 


Nesting  double-crested  cormorants. 


adding  more  material  to  the  top  of  the  nest  (e.g., 
double-crested  cormorant,  Brandt's  cormorant  [P. 
auritus],  and  brown  pelican).  These  nests  are  easily 
observed  from  a  distance,  are  easily  counted,  and  are 
usually  discernible  in  photographs  (Sidle  and  Fergu- 
son 1982). 

Arboreal  Nests.  Arboreal  nests  are  built  at  various 
heights  from  just  a  few  centimeters  to  tens  of 
meters.  The  nest  support  can  vary  from  bulrushes, 
willows,  and  sagebrush  to  large  alders,  cottonwoods, 
or  conifers.  Nest  construction  material  varies 
considerably  with  the  large  nests  of  some  herons  and 
cormorants,  often  containing  sticks  and  small 
branches.  Nests  are  often  in  full  view  from  the 
ground  or  air;  however,  arboreal  nests  of  some 
species  may  be  placed  in  concealed  locations,  and 
locating  them  can  be  difficult.  Nests  placed  in  trees 
may  be  fully  exposed  at  the  onset  of  nesting,  but  are 
soon  concealed  by  emerging  leaves. 


POPULATION  MEASUREMENT  CRITERIA 

This  section  deals  with  the  mechanisms  and 
necessary  requirements  to  design  and  conduct  a  sur- 
vey, monitor,  or  census  program  (McCrimmon 
1976). 

Biologists  should  remember,  two  important 
points  when  working  with  colonial  nesting  water 
birds:  ( 1 )  that  actions  can  severely  affect  the  study 
species,  and  (2)  that  any  program  undertaken  must 
be  done  well.  This  includes  taking  steps  to  ensure 
the  program  results  are  available  to  present  day  re- 
searchers and  to  those  in  the  distant  future  (see 
McCrimmon  1978b;  Herman  1980).  Following  are 
the  requirements  needed  before  a  survey  or  moni- 
toring and  census  program  can  be  designed  or 
conducted. 

What  to  Measure:  Populations  or  Habitats 

There  are  two  different  types  of  nesting  distribu- 
tions that  must  be  considered.  One  is  the  distribu- 
tion of  the  study  species'  colonies  over  often  broad 
geographic  areas  of  one  or  more  habitat  types.  The 
other  is  the  distribution  of  nests  within  colonies.  The 
approaches  for  sampling  colonies  versus  nests  are 
usually  different;  the  sampling  program  used  depends 
on  the  program  goals.  If  the  goal  is  general  distribu- 
tion and  total  numbers  of  colonies,  with  imprecise 
knowledge  of  species  and  their  numbers,  then  sur- 
veys over  large  areas  are  employed.  However,  if  ac- 
curate numbers  of  species  at  each  colony  are 
needed,  then  censuses  of  colonies  are  also  required. 

Waterbirds  (those  that  commonly  nest  in  colo- 
nies) usually  concentrate  their  nests  in  a  few  colo- 
nies, leaving  apparently  suitable  nesting  habitat 
unused.  The  colony  sizes  and  species  present  usually 


394 


Colonial  Waterbirds 


vary.  Sampling  a  portion  of  apparently  available  and 
suitable  habitat  to  derive  values  for  the  number  of 
colonies  and  individuals  per  unit  area  and  then  ex- 
trapolating to  the  total  area  are  done  only  with  cau- 
tion. The  reliability  of  derived  values  is  often 
improved  with  larger  samples,  when  colony  size 
variation  is  small,  and  species  composition  is  con- 
stant. To  determine  the  total  number  of  colonies 
present  for  any  species  it  is  desirable  to  locate  each 
colony.  However,  this  is  not  always  practical  and 
extrapolations  from  samples  must  be  considered. 

When  colonies  are  censused,  the  goal  is  usually 
to  determine  the  total  number  of  nesting  individuals 
of  each  species  present.  The  possibilities  of  accom- 
plishing this  depend  on  the  species  present,  their 
numbers,  the  visibility  of  nests,  the  resources  avail- 
able, and  the  limitations  imposed  by  the  conse- 
quences of  disturbance.  Preferably,  a  direct  count  of 
each  nest  in  the  colony  by  personnel  visiting  each 
nest  is  most  desirable.  However,  this  is  often  not 
possible  or  desirable.  In  such  cases,  samples  of  the 
colony  are  obtained  and  values  extrapolated  to  the 
total  area  occupied  by  the  colony.  A  variety  of  tech- 
niques are  available  to  obtain  sample  counts  (Buck- 
ley and  Buckley  1978). 

Where  to  Look 

The  process  of  locating  colonies  can  take  con- 
siderable time  and  resources;  therefore,  learning 
as  much  as  possible  about  the  study  species'  colonies 
in  the  study  area  is  needed  before  field  work  begins. 
The  time  spent  reading  literature,  field  notes,  and 
reports,  and  talking  with  local  observers  is  usually  a 
good  investment.  When  surveys  do  begin,  the  known 
sites  can  serve  as  starting  points.  Later  as  more  expe- 
rience is  gained  through  study,  other  habitats  can 
be  searched. 

If  the  habitat  occurs  as  widely  scattered  patches, 
as  continuous  belts,  or  large  expanses,  the  use  of  an 
airplane  for  observations  may  be  necessary.  Depend- 
ing on  the  species,  colonies  may  be  visible  from 
the  air  and  easily  plotted  on  maps.  The  use  of  a  boat 
may  be  required  to  locate  colonies  along  rivers, 
waterways,  marshes,  and  coastlines.  For  secretive  or 
hard  to  observe  species  it  may  be  necessary  to 
closely  inspect  likely  colony  sites,  on  foot  or  from  a 
boat. 

The  colonies  of  some  species  are  not  specific  to 
a  particular  habitat  or  a  few  habitats;  other  colonies 
are  concealed.  In  both  cases,  these  colonies  would 
be  difficult  to  locate  by  direct  searches.  The  loca- 
tions of  colonies  can  be  narrowed  down  to  smaller 
areas  by  observing  where  foraging  birds  are  located. 
Observers  can  return  to  these  areas  and  attempt  to 
follow  birds  to  their  colonies.  It  is  also  possible  to 
determine  the  locations  of  colonies  by  observing 
where  birds  emerge  from  vegetation  when  leaving 


their  colonies  to  forage.  Nocturnal  species  may  re- 
quire systematically  checking  colony  sites  at  night,  a 
potentially  long  process.  Burrowing  species  may 
require  the  excavation  of  burrows  to  determine  spe- 
cies composition  of  colonies. 

Once  the  survey  of  an  area  or  habitats  is  com- 
pleted, observers  can  return  to  the  located  sites  and 
make  systematic  censuses  of  the  species,  their  num- 
bers and  status.  The  methods  used  to  locate  colonies 
in  the  beginning  can  then  be  used  for  monitoring 
the  colonies.  Unfortunately  there  is  no  one  way  to 
survey  for  the  presence  of  colonies.  The  methods 
must  be  adjusted  for  the  study  species,  the  habitats 
present,  the  area  covered,  the  resources  available, 
and  the  program  goals.  The  researcher  must  decide 
the  precision  needed  to  cover  the  target  area.  Should 
all  the  colonies  be  located,  or  just  a  few  for  later 
projects?  In  recent  years,  there  were  several  pro- 
grams to  locate  and  count  all  breeding  colonies  and 
species  of  the  North  American  continental  margin 
(for  examples  see  Blacklock  et  al.  1978;  Clapp  et  al. 
1982a,  b;  Erwin  1979;  Erwin  and  Korschgen  1979; 
Korschgen  1979;  Nesbitt  et  al.  1982;  Portnoy  1977; 
Portnoy  et  al.  1981;  Sowls  et  al.  1978,  1980;  and 
Speich  and  Wahl  1985). 

How  to  Observe 

Counting  the  number  of  active  nests  is  an  un- 
likely prospect  for  many  species  at  many  sites,  al- 
though it  can  be  done  under  ideal  situations.  Often, 
nests  are  hard  to  find,  and  may  be  concealed  in  vege- 
tation, or  both.  Even  when  nests  can  be  counted,  it 
is  unlikely  that  the  status  of  every  nest  can  be  deter- 
mined. Impacts  from  disturbance  on  the  study  spe- 
cies also  need  to  be  considered.  In  this  case, 
observers  must  often  obtain  sample  counts  and  sta- 
tus determinations  from  portions  of  the  colony  and 
apply  correction  factors  to  samples. 


High  density  nesting  species,  such  as  white  pelicans,  suc- 
cessfully reproduce  with  no  social  conflict. 


Colonial  Waterbirds 


395 


If  nests  cannot  be  counted,  less  precise  methods 
must  be  employed.  These  include  determining  the 
net  movements  of  adults  to  and  from  feeding  areas, 
counting  adults  at  feeding  areas  and,  for  nocturnal 
burrowing  species,  trying  to  determine  numbers 
from  the  sounds  of  adults.  Although  imprecise,  these 
methods  are  often  the  best  that  can  be  used. 

When  to  Observe 

Equally  important  to  knowing  how  and  what  to 
observe  is  to  know  when  to  observe.  Counts  of  nests 
or  burrows,  and  adults  on  nests,  are  usually  made 
during  daylight  hours.  The  time  of  day  observations 
are  made  can  be  important.  If  adults  are  counted, 
it  is  ideal  to  count  when  the  adult-to-nest  ratio  ap- 
proaches 2:1.  When  adults  might  be  frightened  from 
their  nests,  especially  if  the  nests  contain  eggs  and 
nestlings  that  would  be  exposed  to  overheating, 
chilling,  or  predation,  a  time  of  day  for  observations 
should  be  chosen  when  stress  is  minimal.  Ideally, 
numbers  are  best  determined  after  the  peak  of  egg 
laying  when  most  nests  at  a  colony  have  incubating 
adults. 


MEASUREMENT  TECHNIQUES 
Counts  of  All  Nests  or  Adults 


to  colony  conditions  and  program  goals.  Whether 
the  samples  are  obtained  by  personnel  in  the  colony, 
at  remote  vantage  points  or  in  aircraft,  or  through 
stereo  photography,  is  decided  on  a  colony-by-col- 
ony basis.  The  sample  plots  are  adjusted  as  necessary 
to  match  the  habitat  differences  within  the  colony. 

If  sample  counts  of  adults  are  required,  the  pro- 
cedures are  similar  to  those  used  to  obtain  sample 
counts  of  nests. 

Adult  Counts  Away  from  the  Colony 

When  estimates  of  the  numbers  of  birds  nesting 
in  a  colony  cannot  be  obtained,  counting  adults 
away  from  the  colony  is  an  alternative.  One  method, 
flight  line  counts,  counts  adult  birds  as  they  fly  to 
and  from  foraging  areas.  This  is  accomplished  by 
observing  birds  from  vantage  points  that  offer  a  clear 
view  of  the  area  surrounding  the  colony.  Numbers 
derived  from  net  movement  counts  are  affected  in 
several  ways.  In  part  these  are  taken  into  considera- 
tion in  calculations  to  derive  total  numbers.  The 
total  number  derived  can  be  affected  by  ( 1 )  the  por- 
tion of  the  colony's  birds  that  come  by  the  point  of 
observation,  (2)  the  portion  of  the  adults  observed 
that  are  actually  nesting,  and  (  3 )  the  portion  of  the 
birds  observed  that  are  from  the  subject  colony  and 
not  from  other  colonies. 


Counting  every  nest  is  usually  practical  only  in 
small  colonies;  larger  colonies  can  be  counted  when 
there  are  sufficient  personnel.  When  making  direct 
counts  of  burrows,  ground  nests,  or  arboreal  nests,  it 
is  desirable  to  mark  each  nest  visited  and  then  re- 
count all  nests,  marked  and  unmarked,  to  derive 
correction  factors.  It  is  useful  to  partition  the  colony 
to  ensure  a  thorough  census.  Counts  from  aircraft 
are  necessarily  less  precise,  but  often  necessary. 
However,  counting  nests  from  aerial  photographs, 
especially  through  stereo  photography,  is  often  a 
practical  way  to  determine  numbers  of  nesting  birds. 
The  counting  of  all  nests,  especially  by  observers  in 
the  colony,  is  limited  by  the  effects  of  disturbance. 

As  stated,  counts  of  adults  are  desirable,  but 
often  impractical,  especially  when  nests  are  difficult 
to  recognize.  Counts  of  adults  are  usually  made  from 
remote  vantage  points,  from  aircraft,  or  through 
stereo  photography.  Helicopters  are  often  superior 
to  fixed  wing  aircraft,  as  they  allow  prolonged  view- 
ing of  the  colony  from  chosen  vantage  positions 
above  the  colony. 

Samples  of  Nests  or  Adults 

Estimates  of  the  total  numbers  of  nesting  birds 
are  derived  from  counts  of  sample  sections  in  the 
colony.  Sample  plots  are  established  through  random 
strip  transects,  evenly  spaced  belts,  point-centered- 
quarter,  random  plots,  or  other  methods  suitable 


Another  method  is  to  obtain  counts  of  birds 
when  they  are  on  foraging  areas  away  from  the  col- 
ony. The  considerations  are  the  same  as  for  flight 
line  counts.  Additionally,  for  diving  birds,  at  any 
moment  of  observation,  a  portion  of  the  birds  may 
be  below  the  water  surface,  out  of  view.  Counts 
of  the  birds  are  made  directly  from  vantage  points, 
aircraft,  or  through  stereo  photography.  The  time 
taken  to  understand  the  foraging  schedule  of  the 
birds  is  well  spent. 

Correction  Factors 

When  definitive  numbers  of  birds  of  any  species 
nesting  at  any  given  colony  site  cannot  be  obtained, 
numbers  can  be  approximated  through  a  census 
program.  The  number  derived  is  qualified  by  the 
date  of  the  census  in  relation  to  the  nesting  stage  of 
the  colony.  Usually  nests,  burrows,  or  adults  are 
counted,  but  only  a  proportion  of  the  adults  nesting 
at  a  colony  may  be  present  at  the  time  of  the  census. 
Likewise,  a  count  of  nests  or  burrows  will  contain 
sites  that  are  not  active.  And,  of  the  active  nests 
found  on  any  given  census,  only  a  portion  will  be 
successful,  producing  fledging  young. 

Correction  factors  are  applied  to  counts  to  bring 
them  more  in  line  with  the  actual  number  of  adults 
probably  present  and  nesting  at  the  colony.  The 
usual  approach  is  to  correct  for  the  nests  or  burrows 
that  are  not  active,  and  for  those  missed  during  a 


396 


Colonial  Waterbirds 


census.  The  only  difference  is  that  generally  the  cor- 
rection factors  for  burrowing  species  are  much 
harder  to  obtain,  especially  for  nocturnal  species. 

If  adults  are  counted,  the  most  important  ratio 
to  determine  is  between  adults  present  and  the  num- 
ber of  active  nests.  Thus,  counts  from  the  air  or  from 
remote  observation  points  are  verified  by  site  checks 
of  sample  sections  in  the  colony  where  the  ratio  is 
determined  (Nisbet  1973;  Kadlec  and  Drury  1968). 

Checking  Accuracy 

Determining  the  numbers  of  a  species  present 
and  nesting  in  a  colony  is  often  elusive  because  of 
the  known,  uncontrollable  and  usually  unquantifi- 
able,  variables  involved.  However,  the  numbers  gen- 
erated are  often  good.  The  accuracy  or  inaccuracy  of 
numbers  obtained  may  preclude  any  detailed  analysis 
(Ralph  1981 ).  This  and  large  values  may  allow  state- 
ments only  when  numbers  are  different  by  factors 
of  2,  4,  or  even  orders  of  magnitude.  The  application 
of  statistical  tests  may  indicate  significant  differences 
or  trends  in  the  data,  but  significant  statistical  differ- 
ences may  not  translate  to  significant  biological  dif- 
ferences or  be  real. 

Nevertheless,  the  accuracy  of  a  survey  or  census 
should  be  checked.  There  are  two  main  approaches. 
First,  the  survey  or  census  can  be  repeated  by  the 
same  observers  or  different  observers.  Hopefully,  all 
methods  are  constant  and  the  environmental  condi- 
tions remain  the  same.  Obtaining  the  same  colony 
attendance  pattern  during  the  repeat  of  the  survey 
or  census  is  desirable.  Repeating  a  survey  or  census 
with  different  methods  could  be  employed  when 
there  is  reason  to  suspect  the  results  of  the  original 
survey  or  census. 

Sample  Methods  and  Comparisons 

The  following  examples  of  methods  used  to 
observe  and  obtain  numbers  of  colonial  nesting  wa- 
terbirds  are  only  a  sample;  many  other  methods 
could  be  included  (Buckley  and  Buckley  1978).  The 
examples  show  the  results  of  actual  studies  from  a 
variety  of  methods,  on  several  species  in  different 
habitats.  These  examples  should  help  the  researcher 
select  the  methods  needed  to  satisfy  program  goals. 
Studies  where  different  methods  are  compared  are 
presented  in  more  detail  to  demonstrate  the  variabil- 
ity of  results  and  efficiency  of  methods  in  specific 
situations.  These  examples  also  show  that  considera- 
ble effort  must  be  extended  to  obtain  data  on  spe- 
cies and  their  numbers,  nesting  in  waterbird 
colonies,  and  that  the  inherent  difficulties  should  not 
be  underestimated! 

Aerial  Photography — Ground  Nest  Count.  On 

June  2  and  3,  1980,  ground  counts  were  conducted 
and  aerial  photographs  taken  of  nesting  white 


pelicans  on  Small  Island,  Chase  Lake  National 
Wildlife  Refuge,  North  Dakota  (Sidle  and  Ferguson 
1982).  The  aerial  photography  revealed  1,368 
pelican  nests,  and  the  ground  census  1,355  pelican 
nests,  a  difference  of  only  1  % . 

Ground  Nest  Counts — Duplications.  On  May  24, 

1980,  herring  gulls  (L.  argentatus)  and  lesser  black- 
backed  gulls  (L.  fuscus)  nests  were  counted  by  10 
observers  on  10  separate  plots  in  Flat  Holm,  United 
Kingdom  (Ferns  and  Mudge  1981 ).  The  goal  of  the 
observers  was  to  systematically  search  their 
respective  plots  for  all  nests  and  place  a  spray  paint 
mark  by  each  nest  found.  The  observers  then 
switched  plots,  and  counted  all  marked  and 
unmarked  plots,  establishing  a  ratio  of  the  two  nest 
counts.  For  nine  plots  the  average  percentage  of 
nests  missed  during  the  first  count  was  16.9  ±  33% 
(SE).  The  range  for  nine  plots  was  5.0%  to  27.3%. 
In  total  2,758  nests  were  marked  and  counted,  and 
the  corrected  total  number  of  nests  was  3,651. 

Similar  results  were  obtained  from  duplicating 
the  count  and  marking  the  nests  in  gull  colonies 
in  Massachusetts.  The  error  ranged  from  4%  to  22% 
on  eight  sets  of  counts.  Plot  totals  varied  from  27 
to  442  nests,  on  plots  of  0.06  to  0.73  ha  (0.15  to 
1.8  a.)  with  varied  vegetation  density  (Erwin 
1980b).  Even  in  sparsely  vegetated  small  plots,  there 
is  considerable  error  in  finding  ground  nests. 

Ground  Censuses  of  Ground  Nests.  A  few  of  the 

methods  used  to  determine  the  numbers  of  ground 
nesting  species  by  counting  on  the  ground  are 
presented  here.  Numbers  of  ring-billed  gulls  (L. 
delaivarensis)  nesting  in  a  colony  were  determined 
by  dividing  the  colony  into  sections  and  then 
counting  all  the  nests  in  each  section  (Southern  and 
Southern  1981 ).  Total  counts  of  nesting  gulls  and 
terns  in  colonies  were  obtained  by  counting  and 
marking  all  the  nests  in  distinct  areas  of  the  colony. 


Ring-billed  gull. 


Colonial  Waterbirds 


397 


The  areas  were  then  counted  again  and  the  numbers 
of  marked  and  unmarked  nests  were  recorded.  A 
Lincoln  Index  was  applied  to  get  the  "true"  total 
(Erwin  1980b). 

Strip  transects  were  randomly  placed  through 
large  colonies,  and  all  nests  were  counted  in  the 
transects.  These  transect  counts  were  continued  un- 
til 10%  to  20%  of  the  colony  area  was  surveyed. 
The  densities  obtained  were  used  to  calculate  the 
total  numbers  nesting  in  the  whole  colony  ( Erwin 
1980b). 

In  large  wading  bird  colonies,  belt  transect 
counts  are  a  practical  way  to  obtain  an  idea  of  the 
total  numbers  nesting.  Transects  are  made  through 
the  colony  and  all  nests  are  counted  in  the  belt. 
These  are  repeated  until  10%  of  the  total  colony 
area  is  censused.  This  technique  at  least  obtains  or- 
der of  magnitude  counts  for  a  colony  (Portnoy 
1980).  Although  this  technique  may  be  practical  for 
arboreal  or  surface  ground  nesting  species,  the  pros- 
pect of  obtaining  a  10%  area  sample  in  a  large  bur- 
rowing species  colony  is  discouraging. 

In  a  burrowing  species  colony,  the  nests,  in 
many  cases,  must  be  individually  examined  to  iden- 
tify species  and  to  determine  whether  the  nests  are 
active  before  correction  factors  can  be  calculated 
(pers.  obs.).  The  point-centered-quarter  technique  of 
obtaining  samples  has  been  used.  Between  8  and  1 5 
samples  are  taken  along  an  arbitrarily  drawn  line 
through  a  colony.  However,  the  strip  transect 
method  is  recommended  (Erwin  1980b).  Total 
counts  of  nests  are  practical  only  in  small  colonies 
(Portnoy  1980)  and  the  accuracy  can  be  quite  good 
(pers.  obs.). 

Censusing — Helicopter,  Airplane,  and  Ground 
Nest  Counts.  The  census  results  obtained  from 
a  helicopter,  a  fixed  wing  aircraft,  and  on  the 
ground,  of  a  mixed  species  colony  of  mangrove 
(Rhizosphora  mangle)  were  compared  (Kushlan 
1979).  For  large  white  and  dark  birds,  nesting  in  the 
tree  tops,  the  helicopter  results  had  an  error  10% 
to  16%  of  the  ground  nest  count.  The  error  for  fixed 
wing  aircraft  compared  to  ground  censuses  ranged 
from  32%  to  100%  (also  see  Erwin  1980a). 


142%.  Five  samples  were  below  11%,  and  one  each 
at  26%  and  54% .  Point-centered-quarter  error  rates 
ranged  from  3%  to  400% .  Most  were  larger  than  the 
corresponding  strip  transect  sample  errors.  The  strip 
transect  method  is  apparently  equally  reliable 
regardless  of  the  spatial  distribution  of  the  sample 
population.  Laboratory  studies  with  artificial 
populations — random,  uniform,  and  patchy — yielded 
similar  results  (Erwin  1980b). 

Flight  Line  Count — Ground  Count.  Erwin  (1981) 
compared  flight  line  counts  and  ground  nest  count 
results  at  1 3  mixed  species  wading  bird  colonies.  For 
three  hours  in  the  morning,  all  birds  entering  and 
leaving  colonies  were  counted  and  identified.  Then 
all  nests  were  counted  or  belt  transects  were  used  to 
generate  the  total  nests  present.  Species  and  colony 
differences  in  flight  rate  were  observed.  Individual 
colony  size  error  reached  a  factor  of  2,  and  the 
actual  populations  were  predicted  with  10%  error. 
Species  correction  factors  to  flight  line  data  may 
be  necessary.  Hour  to  hour  variation  in  flight  rates 
can  be  substantial,  and  the  calculation  of  a  mean  rate 
is  suggested.  For  the  use  of  flight  line  counts, 
comparisons  with  other  methods  of  obtaining  the 
number  of  nesting  birds  present,  and  for  diurnal 
burrowing  species,  see  the  discussion  in  Byrd  et  al. 
(1983). 

Aerial  Visual  Estimates,  Aerial  Photograph 
Counts,  and  Transect  Sampling.  Portnoy  (1980) 
compared  aerial  visual  estimates,  aerial  photograph 
counts,  and  transect  counts  in  19  colonies  of  nesting 
wading  birds.  Aerial  visual  estimates  of  great  egrets 
(Casmerodius  albus)  at  these  colonies  were  greater 
4  times  and  less  15  times  than  the  actual  numbers 
present.  Of  33  colonies,  mean  colony  size  was  511 
nests;  the  average  error  of  the  visual  estimates  was 
—  4.2%  (SD  12.2).  Of  six  snowy  egret  colonies, 
mean  size  was  1 ,992  nests;  the  visual  error  was 
-8.7%  (SD  92.0).  For  Louisiana  heron  {Hydranassa 
tricolor)  colonies,  n  =  8;  mean  size,  3,192  nests;  and 
the  visual  error  was  79.1%  (SD  25.1).  And  in  black- 
crowned  night  heron  colonies,  n  =  7;  mean  size,  573 
nests;  and  the  visual  error  was  —84.4%  (SD  14.7). 
Clearly,  aerial- visual  samples  of  colonial  nesting 
wading  birds  greatly  underestimate  numbers  of 
smaller  cryptic  and  concealed  nesting  species  in 
trees  or  marsh  vegetation  (see  also  Erwin  1980a). 


Census  Accuracy — Total  Count,  Strip  Count,  and 
Point-Centered-Quarter.  Three  census  methods 
were  compared  in  six  gull  and  three  heron  colonies 
(Erwin  1980b).  Strip  transect  and  point-centered- 
quarter  methods  were  compared  to  known  total 
counts  in  the  colonies.  Sample  areas  were  within 
10%  to  20%  of  plots  for  the  strip  transect  method. 
Between  8  and  1 5  points  were  sampled  for  the  fixed 
point-centered-quarter  method.  Only  the  strip 
transect  method  showed  reasonable  accuracy  in  the 
field,  and  the  observer  error  ranged  from  3%  to 


Gull  Colony  Ground  Nest  Counts,  Aerial 
Photograph  Counts,  and  Aerial  Visual  Estimates. 

Three  methods  of  determining  nesting  numbers  of 
gulls  at  colonies  were  compared  (Kadlec  and  Drury 
1968).  Photographs  of  colonies  and  visual  estimates 
were  made  of  colonies  from  helicopters.  Double 
ground  counts  of  gull  nests  were  made  in  the 
colonies.  On  the  first  ground  count,  95%  of  all  nests 
were  found.  At  some  colonies,  two  additional  ground 


39S 


Colonial  Waterbirds 


counts  were  made.  The  ratios  of  adults,  counted  on 
photographs  and  from  aerial  estimates,  to  ground 
counts  were  established  for  all  sites.  Ratios  tended  to 
be  similar  within  each  census,  although  the  ratios 
varied  between  censuses.  Confidence  intervals  are 
given  for  predicting  the  number  of  nests  from 
photograph  counts  of  adults  and  visual  estimates  of 
adults.  For  regional  determinations  of  gull  numbers, 
visual  estimations  are  a  valid  approach.  More 
detailed  studies  within  a  region  require  photography. 
In  predicting  numbers  from  visual  estimates  and 
photograph  counts,  errors  of  19%  to  31%  and  26% 
to  35% ,  respectively,  can  be  expected.  Annual 
changes  of  25%  or  less  may  not  be  detectable  with 
either  method. 

Belt  Transect  Sampling  and  Confidence  Limits. 

In  1976,  gulf  coast  heronries  were  sampled  with 
randomly  placed  ground  belt  transects,  in  which  all 
nests  were  sampled  (Portnoy  1980).  Transects  were 
run  until  10%  of  the  colony  area  was  sampled.  The 
95%  confidence  limits  generated  for  nest  densities 
per  transect  were  large:  for  estimated  numbers  of 
nests,  n=  16,880;  the  confidence  interval  was  2,672 
nests;  n=  14,938,  3,090;  n=  14,279,  1,809; 
n=  12,600,  1,337;  n  =  3,240,  810.  When  dealing  with 
largely  inaccessible  colonies,  i.e.,  concealed  nests, 
colonies  over  water,  and  cryptic  species,  this  is 
about  the  best  one  can  expect.  When  surveying  large 
areas,  and  a  large  number  of  difficult  large  colonies, 
economic  and  personnel  limits  may  not  allow  better 
work.  In  large  colonies,  order-of-magnitude 
measurements  may  be  all  that  can  be  expected. 

Estimating  Numbers  from  Aerial  Photographs — 
An  Experiment.  Nine  observers  were  asked  to 
estimate  the  numbers  of  birds  on  10  different 
photographs  each  of  5  consecutive  days.  The 
observers  were  in  three  groups:  inexperienced,  past 
experience,  and  recent  experience.  The  effects  of 
observer  differences,  prior  experience,  training,  and 
numerical  magnitude  were  examined.  With  no 
training,  experienced  observers  were  more  accurate 
than  inexperienced  observers.  Trends  in 
underestimation  were  observed.  With  training, 
estimates  by  eight  observers  were  within  10%  of  the 
actual  number.  This  study  by  Erwin  (1982)  is 
interesting  and  suggests  that  training  prior  to 
censuses  or  surveys  is  important. 

Aerial  Photography  of  Large  Colonies  of  Cliff 
and  Island  Top  Nesting  Species.  Aerial 
photography  was  used  to  determine  colony  size  of 
nesting  common  murres  (Nettleship  1980).  This 
species  can  nest  in  numbers  in  the  100,000s.  It  nests 
on  cliff  ledges,  the  flat  tops  of  islands,  in  boulder 
fields,  and  in  caves.  There  is  no  easy  way  to  census 
these  large  colonies.  Correction  factors  for  counts 
from  photographs  are  generated  from  field  counting 
and  counting  from  photographs,  groups  of  about  200 


birds.  If  photographs  cannot  be  obtained  for  all 
areas,  then  numbers  may  be  estimated  for  these 
areas  from  densities  in  the  known  areas.  The  total 
"number"  may  be  a  composite  from  photograph 
counts,  area  extrapolations,  and  direct  counts  or 
estimates  from  land  or  aircraft.  Correction  factors  for 
flat,  accessible  island  top  areas  are  generated  by 
counting  birds  on  photographs,  field  counting  adults, 
or  both,  and  then  flushing  the  adults  and  counting 
their  eggs  in  distinct  areas.  For  large  colonies,  this 
procedure  is  very  time-consuming,  taking  one 
observer  up  to  6  weeks  to  complete  a  census.  To 
obtain  population  samples  and  detect  population 
trends,  numbers  are  determined  in  sample  plots  and 
extrapolated  to  known  areas  of  similar  habitat.  In  a 
large  colony,  this  can  take  an  observer  up  to  10  days. 
Several  photographs  appear  in  this  publication 
showing  birds,  habitat,  and  sampling  areas  (see  also 
Nettleship  1976). 

Aerial  Photography  of  Cliff  and  Island  Top 
Nesting  Species.  Aerial  photographs  of  nesting 
gannets  on  ledges,  on  cliff  faces,  and  on  flat  ground 
at  the  tops  of  cliffs  were  used  to  determine  nesting 
numbers  (Nettleship  1975).  Overlapping  aerial 
photographs  of  incubating  birds  were  taken  from  a 
fixed-wing  aircraft  with  a  35mm  camera,  using  a 
50mm  lens  and  Kodak  Plus-X,  black-and-white  film. 
The  photographs  were  taken  about  600  m  ( 1 ,980  ft ) 
from  the  colony.  Glossy  prints,  7  x   10  in.  or  9  x 
1 3  in.  were  made.  It  was  possible  to  count  the 
number  of  birds  associated  with  each  nest  from  the 
photographs  by  using  hand  lenses  and  1   x   1  cm 
(0.4  in.)  plastic  overlays.  This  gave  a  count  of  nest 
site  holders,  not  the  actual  number  of  birds  that 
build  nests  and  actually  lay  eggs.  The  error  was 
thought  to  be  less  than  20% .  The  error  largely 
resulted  from  the  imprecise  boundaries  on 


Common  murre  and  nest  in  typical  cliff  ledge  site. 


Colonial  Waterbirds 


399 


photographs  and  difficulty  in  counting  the  nests  at 
the  back  of  broad,  flat  areas  at  cliff  tops. 

Aerial  Photograph  Counts  and  Estimates  of 
Ground  Nests.  Estimates  of  the  total  numbers  of 
ground  nests  were  made  from  airplanes,  and  these 
were  corrected  with  ground  counts  of  nests  (Erwin 
1980a).  Counts  of  ground-nesting  terns  and  black 
skimmers  {Runchops  niger)  were  determined  from 
aerial  photographs  and  then  corrected  with  ground 
counts  of  nests  (Portnoy  1980). 

Aerial  Estimates  of  Arboreal  Nests.  Total 
estimates  of  all  visible  arboreal  nests  were  made 
from  an  airplane,  and  the  counts  were  then 
corrected  with  "good"  data  (nest  counts,  etc.;  Erwin 
1980a).  Large,  conspicuous  wading  birds  can  be 
counted  in  this  manner,  but  the  technique  is  not 
suitable  for  determining  the  numbers  of  cryptic,  and/ 
or  concealed,  nesting  species  (Portnoy  1980). 


DISTURBANCE 

Disturbance  includes  any  activity  by  an  observer 
that  results  in  any  perturbation  of  the  "normal"  con- 
dition on  the  colony  under  observation.  The  distur- 
bance created  ranges  from  that  undetectable  by 
observers  to  major  disruptions  of  the  colony  (Table 
2).  Unless  observations  are  conducted  from  a  posi- 
tion remote  and  distant  to  the  colony,  they  will  re- 
sult in  disturbance  to  the  colony. 

Sources  of  Disturbance 

The  nature  and  magnitude  of  disturbance  often 
depend  on  the  experience  and  skill  of  the  observers. 
The  level  of  disturbance  is  usually  a  function  of  the 
platform  of  observation,  and  the  techniques  used 
to  make  observations.  Table  2  lists  types  and  effects 
of  disturbance  on  several  colonial-nesting  species, 
with  references  to  pertinent  studies. 


Table  2.     Types  and  Effects  of  Disturbance. 


Types  of  Disturbance 

Authors 

Species  Groups 

AIRCRAFT: 

HELICOPTERS: 
OBSERVERS  IN  OR  NEAR 
COLONIES: 

Dunnet  1977 
Burger  1980 
Kushlan  1979 
Burger  1981 
Robert  and  Ralph  1975 

seabird  colonies 
herring  gulls 
wading  bird  colonies 
gulls 
gulls 

Effects  of  Disturbance 

AGGRESSIVE  BEHAVIOR: 

Burger  1980,  1981 
Robert  and  Ralph  1975 

gulls 
gulls 

THERMAL  STRESS: 
(when  adults  leave  eggs  or 
nestlings  unattended) 

Bartholomew  et  al.  1953 

Bartholomew  and  Dawson  1954 

Vermeer  1963 

Drent  et  al.  1964 

Harris  1964 

Drent  1967 

Hunt  1972 

Anderson  and  Keith  1980 

Dawson  and  Bennett  1981 

Bennett  et  al.  1981 

white  pelicans 
pelicans,  gulls,  herons 
glaucous-winged  gulls 
seabird  colonies 
gulls 

herring  gulls 
herring  gulls 
brown  pelicans 
Western  gulls 
Western  gulls 

PREDATION: 

Drent  et  al.  1964 
Verbeek  1982 
Manuwal  1974b 
Penland  1981 

by  crows  at  colonies 
by  crows  at  colonies 
by  gulls  at  colonies 
of  Caspian  tern  eggs 

PREDATION  FROM  OBSERVER- 
INDUCED  DISTURBANCES: 

Anderson  and  Keith  1980 
Anderson  and  Keith  1980 
Ellison  and  Cleary  1978 
DesGranges  and  Reed  1981 
Buckley  and  Buckley  1972 
Kury  and  Gochfeld  1975 
Johnson  1938 

pelicans  by  gulls  and  ravens 
tern  eggs/young  by  gulls 
cormorant  eggs  by  gulls 
eggs/young  by  gulls 
tern  eggs  by  gulls 
cormorant  eggs  by  gulls 
murre  eggs/young  by  gulls 

CHICK  MOBILITY: 
(flushing  of  young  from  ground 
nests) 

Robert  and  Ralph  1975 
Penland  1981 
Burger  1981 

gulls 

Caspian  terns 
gulls 

400 


Colonial  Waterbirds 


Table  2.     Types  and  Effects  of  Disturbance  (concluded). 


Effects  of  Disturbance 

Authors 

Species  Groups 

NEST  AND  COLONY 
ABANDONMENT: 
(disturbance  during  nesting) 

Werschkul  et  al.  1976 
Manuwal  1978 
Anderson  and  Keith  1980 
Jehl  1973 

Tremblay  and  Ellison  1979 
Ellison  and  Cleary  1978 
Burger  1981 
Penland  1981 
Manuwal  1978 
Manuwal  1978 

great  blue  herons 

storm  petrels 

brown  pelicans 

brown  pelicans 

black-crowned  night  herons 

double-crested  cormorants 

gulls 

Caspian  terns 

tufted  puffins 

rhinoceros  auklets 

DESTRUCTION  OF  NESTS: 

Giles  and  Marshall  1954 

marsh-nesters 

CANNIBALISM: 
(killing  of  eggs  and  young  by 
conspecific  adults) 

Anderson  and  Keith  1980 

Hand  1980 

Harris  1964 

Hunt  and  Hunt  1975,  1976 

Parson  1971 

Patterson  1965 

Penland  1981 

Robert  and  Ralph  1975 

Vermeer  1963 

seabirds 
Western  gulls 
gulls 
gulls 

herring  gulls 
black-headed  gulls 
Caspian  terns 
Western  gulls 
glaucous-winged  gulls 

REPRODUCTIVE  SUCCESS: 
(additional  studies) 

Anderson  et  al.  1976 
Ashmole  1963 
Brun  1972 
Gillett  et  al.  1975 
Kadlec  and  Drury  1968b 
Schreiber  1979 
Schreiber  1979 

seabirds 

tropical  oceanic  birds 

gannet 

glaucous-winged  gulls 

herring  gulls 

brown  pelicans 

brown  pelicans 

DISCUSSION  AND  CONCLUSIONS 

A  large  variety  of  species  of  birds  are  considered 
colonial  nesting  waterbirds.  They  are  found  in  the 
interior  of  the  continent  in  association  with  aquatic 
systems  and  in  marine  environments  along  the  conti- 
nental margin.  The  habitat  preferences  for  nest  sites 
vary  between  species  and  often  within  species.  Nests 
range  from  burrows  dug  in  the  soil,  ground  nests  of 
various  sizes  and  types,  floating  nests,  and  arboreal 
nests.  Nests  can  be  completely  concealed  or  easy  to 
see.  The  feeding  methods,  positions  on  food  chains, 
foraging  habitats,  and  feeding  schedules  are  varied, 
and  the  sizes  of  the  species  range  from  tens  of  grams 
to  the  low  kilograms.  These  factors  are  usually  im- 
portant in  the  design  and  execution  of  any  program. 

The  large  variance  in  the  breeding  biology  of 
colonial  nesting  waterbirds  requires  that  many  differ- 
ent measurement  techniques  be  employed  in  survey, 
census,  and  monitoring  programs.  The  specific  meas- 
urement techniques  depend  on  the  program  goals, 
the  resources  available  to  perform  the  program,  the 
specific  characteristics  of  the  placement  and  visibility 
of  nests  and  nesting  adults,  the  size  of  the  nesting 
colonies,  the  geographic  distribution  of  the  colonies, 
and  the  effects  of  disturbance  from  performing  the 


program.  Unfortunately,  no  one  method  or  few 
methods  can  be  considered  for  obtaining  quantified 
information.  Each  program  must  be  tailor-made  and 
adapted  to  each  species  and  program  goals.  It  is  ad- 
visable to  be  well-prepared  before  a  program  begins, 
but  flexibility  must  be  allowed  for  inevitable  adjust- 
ments. However,  despite  all  the  inherent  problems  in 
working  with  colonial  nesting  waterbirds,  reasonably 
accurate  results  can  be  obtained  from  measurement 
programs. 

A  variety  of  methods  can  be  employed  to  meas- 
ure the  numbers  of  individual  species  in  colonial 
nesting  waterbird  colonies.  Observations  are  made  in 
the  colony,  from  the  edge  of  the  colony  or  other 
remote  positions  on  land,  from  boats  or  aircraft,  and 
by  stereo  photography.  Aircraft  are  efficient  when 
large  areas  need  to  be  covered  quickly.  Observations 
and  photographs  of  colonies  are  made  from  aircraft. 
Helicopters  provide  stable  platforms  and  can  be  ac- 
curately positioned  above  colonies  for  prolonged 
viewing  and  photography.  Population  estimates  can 
be  derived  from  counts  of  nests  or  adults.  Counts  of 
adults  or  nests  by  stereo  photography  are  often  effi- 
cient. In  some  situations,  every  nest  or  adult  in  a 
colony  can  be  counted,  but  most  often,  sample 
counts  from  plots  in  the  colony  are  required. 


Colonial  Waterbirds 


401 


A  variety  of  methods  can  be  useful  for  partition- 
ing a  colony  for  sample  counts.  Correction  factors 
should  be  applied  to  census  numbers  to  adjust  for 
inactive  nests,  adults  not  nesting,  adults  not  present 
in  the  colony,  when  counts  of  adults  are  made,  etc. 
Duplicate  counting  of  nests  or  adults  in  colony  plots 
generates  accuracy  correction  factors  for  the  per- 
centage of  nests  or  adults  missed  on  the  first  count- 
ing. Additional  counts  or  estimates  can  be  obtained 
by  using  a  completely  different  method  to  check 
accuracy. 

Before  colonies  can  be  censused,  they  first  have 
to  be  located.  This  is  accomplished  first  by  searching 
literature,  reports,  and  field  notes,  and  interviewing 
local  observers.  This  first  step  can  save  time  and  give 
good  indications  of  the  locations  of  colonies,  the 
species  present,  and  the  habitats  they  occupy.  If  the 
goal  of  the  program  is  to  find  all  colonies,  then  wide 
area  searches  are  necessary,  with  the  first  effort  con- 
centrated in  likely  habitats  or  locations.  Surveys  are 
conducted  on  foot  and  from  cars,  boats,  and  aircraft. 
Once  the  species,  habitats,  and  geography  become 
familiar,  the  survey  can  be  refined.  If  the  program 
goal  is  the  numbers  of  each  species  in  one  or  more 
colonies,  then  a  census  is  conducted  at  each  colony. 

The  main  factor  limiting  the  methods  and  ulti- 
mately the  accuracy  of  the  data  obtained  from  pro- 
grams, especially  censuses  of  colonies,  is  the  effects 
of  observer-induced  disturbance  to  the  nesting  birds. 
Disturbance  to  colonies  in  the  past  has  resulted  in 
significant  mortality  to  eggs  and  nestlings.  Every 
precaution  must  be  taken  to  ensure  that  disturbance 
by  observers  is  kept  to  a  minimum.  When  working 
in  and  near  colonies,  constant  monitoring  of  the 
colony  is  required.  The  effects  of  disturbance  are  not 
always  obvious  and  immediately  evident. 


Colonial  waterbirds  are  censused  from  the  ground,  at 
water  level,  and  from  the  air. 


402 


Colonial  Waterbirds 


LITERATURE  CITED 


AINLEY,  D.G.  and  T.J.  LEWIS.  1974.  The  history  of  the 
Farallon  Island  marine  bird  populations,  1854-1972. 
Condor  76:432-446. 

ANDERSON,  D.W.  and  JO.  KEITH.  1980.  The  human 
influence  on  seabird  nesting  success:  Conservation 
implications.  Biol.  Conserv.  18:65-80. 

,J.E.  MENDOZA,  and  JO.  KEITH.  1976.  Seabirds  in 

the  Gulf  of  California:  A  valuable,  international  re- 
source. Nat.  Resour.  J.  16:483-505. 

ASHMOLE,  N.P.  1963.  The  regulation  of  numbers  of  tropi- 
cal oceanic  birds.  Ibis  103b:458-473 

BAILEY,  V.  and  MR.  TERMAN.  1983.  A  comparative  study 
of  a  Great  Blue  Heron  colony  in  Chase  County,  Kan- 
sas. Trans.  Kansas  Acad.  Sci.  86:81-88. 

BARTHOLOMEW,  G.A.  and  W.R.  DAWSON.  1954.  Tern 
perature  regulation  in  young  pelicans,  herons,  and 
gulls.  Ecol.  35:466-472. 

, ,  and  E.J.  O'NEILL.  1953.  A  field  study  of 

temperature  regulation  in  young  white  pelicans,  Pele- 
canus  erythrotyncbos.  Ecol.  34:554-560. 

BAYER,  R.D.  and  E.  MCMAHON.  1981.  Colony  size  and 
hatching  synchrony  of  Great  Blue  Herons  in  coastal 
Oregon.  Murrelet  62:73-79. 

BENNETT,  A.F.,  W.R.  DAWSON,  and  R.W.  PUTMAN.  1981. 
Thermal  environment  and  tolerance  of  embryonic 
western  gulls.  Physiological  Zoology  54:146-154. 

BENT,  AC.  1919.  Life  histories  of  North  American  diving 
birds.  U.S.  Natl.  Mus.  Bull.  107. 

.  1921.  Life  histories  of  North  American  gulls  and 

terns.  U.S.  Natl.  Mus.  Bull.  1 1 3. 

.  1922.  Life  histories  of  North  American  petrels  and 


tion  24:177-184. 
— .  1981.  Effects  of  human  disturbance  on  colonial 


pelicans  and  their  allies.  U.S.  Natl.  Mus.  Bull.  176. 
— .  1927.  Life  histories  of  North  American  marsh 


birds.  U.S.  Natl.  Mus.  Bull.  135. 
BIRKHEAD,  T.R.  1978.  Attendance  patterns  of  Guillemots 

(Uria  aalge)  at  breeding  colonies  on  Skomer  Island. 

Ibis  120:219-229- 
BLACKLOCK,  G.W.,  DR.  BLANKINSHIP,  S.D.  KENNEDY, 

KA.  KING,  R.T.  PAUL,  R.D.  SLACK,  J.C.  SMITH,  and 

R.C.  TELFAIR  II,  (compilers).  1978.  Texas  Colonial 

Waterbird  Census,  1973-1976.  Texas  Park  Wildl. 

Dept.,  FA  Rep.  1 5. 
BRUN,  E.  1972.  Establishment  and  population  increase  of 

the  Gannet  (Sulci  bassana)  in  Norway.  Ornis  Scand. 

3:27-38. 
BUCKLEY,  F.G.  and  PA.  BUCKLEY.  1972.  The  breeding 

ecology  of  royal  terns,  Sterna  ( Thalasseus )  maxima 

maxima.  Ibis  114:344-359. 
and .  1980.  Habitat  selection  and  marine 

birds.  Pages  69-112  in  Burger,  J.,  B.L.  Olla,  and  HE. 

Winn,  eds.  Behavior  of  Marine  Animals,  Vol.  4. 

Plenum  Publishing  Co. 
BUCKLEY,  PA.  1978.  Guidelines  for  the  protection  and 

management  of  colonially  nesting  waterbirds.  North 

Atlantic  Regional  Office,  Natl.  Park  Serv.,  Boston,  MA. 
and  F.G.  BUCKLEY.  1977.  Hexagonal  packing  of 

royal  tern  nests.  Auk  94:36-43- 
BURGER,  J.  1976.  Daily  and  seasonal  activity  patterns  in 

breeding  laughing  gulls.  Auk  93:308-32.3. 
.  1978.  The  pattern  and  mechanism  of  nesting  in 

mixed-species  heronries.  Pages  45-58  in  Sprunt,  A.S., 

IV,  J.C.  Ogden,  and  S.  Winckler,  eds.  Wading  birds. 

Natl.  Audubon  Soc,  Res.  Rep.  7. 
.  1980.  Behavioral  responses  of  herring  gulls  (Larus 


species,  particularly  gulls.  Colonial  Waterbirds 
4:28-36. 

BYRD,  G.V.,  R.H.  DAY,  and  E.P.  KNUDTSON.  1983.  Pat- 
terns of  colony  attendance  and  censusing  of  auklets  at 
Buldir  Island,  Alaska.  Condor  85:274-280. 

CLAPP,  R.B.,  R.C.  BANKS,  D.  MORGAN-JACOBS,  and  W.A. 
HOFFMAN.  1982a.  Marine  birds  of  the  southeastern 
United  States  and  Gulf  of  Mexico.  Part  I.  Gaviiformes 
through  Pelecaniformes.  U.S.  Dept.  Inter.,  Fish  Wildl. 
Serv.,  FWS/OBS-82/01. 

, , ,  and .  1982b.  Marine  birds 

of  the  southeastern  United  States  and  Gulf  of  Mexico. 
Part  II.  Anseriformes.  U.S.  Dept.  Inter.,  Fish  Wildl. 
Serv.,  Biol.  Serv.  Prog,  FWS/OBS-82/02. 

CONOVER,  MR.  and  DO.  CONOVER.  1981.  A  docu- 
mented history  of  ring-billed  gull  and  California  gull 
colonies  in  the  western  United  States.  Colonial  Water- 
birds  4:37-43. 

and  D.E.  MILLER.  1980.  Daily  activity  patterns  of 

breeding  Ring-billed  and  California  Gulls.  J.  Field 
Ornith.  51:329-339. 

CUSTER,  T.W.  and  R.G.  OSBORN.  1978.  Feeding  habitat 
use  by  colonially  breeding  herons,  egrets,  and  ibises 
in  North  Carolina.  Auk  95:733-743. 

DAWSON,  W.R.  and  A.F.  BENNETT.  1981.  Field  and  labo- 
ratory studies  of  the  thermal  relations  of  hatching 
western  gulls.  Physiological  Zoology  54:155-164. 

DESGRANGES,  J.L  1979.  Abandoned  windmill  used  as  a 
nesting  site  by  great  blue  herons.  Can.  Field-Naturalist 
93:439-440. 

.  1980.  Competition  entre  le  cormoran  a  aigrettes  et 

Le  Grand  Heron  au  moment  de  la  nidification.  Natu- 
raliste  Can.  107:199-200. 

and  P.  LA  PORTE.  1979.  Second  tour  of  inspection 

of  Quebec  heronries,  1978.  Can.  Wildl.  Serv.,  Prog. 
Notes  105. 

and .  1981.  Third  tour  of  inspection  of  Que- 


bec heronries,  1979.  Can.  Wildl.  Serv.,  Prog.  Notes 

123. 
— .  and  A.  REED.  1981.  Disturbance  and  control  of 

selected  colonies  of  double-crested  cormorants  in 

Quebec,  Colonial  Waterbirds  4:12-19. 
— , ,  and  G.  CHAPELAINE.  1979.  First  tour  of 


inspection  of  Quebec  heronries,  1977.  Can.  Wildl. 
Serv.,  Prog.  Notes  93. 

DRENT,  R.H.  1965.  Breeding  biology  of  the  pigeon  guille- 
mot (Cepphus  columba).  Ardea  53:99-160. 

.  1967.  Functional  aspects  of  incubation  in  the  her- 
ring gull  (Larus  argentatiis  Pont. ).  Behaviour  Supple- 
ment 17:1-32. 

and  C.J.  GUIGUET.  1961.  A  catalog  of  British  Co- 
lumbia seabird  colonies.  Occ.  Papers  British  Columbia 
Provincial  Mus.  12. 

,  G.F.  VAN  TETS,  F.  TOMPA,  and  K  VERMEER. 


argentatiis)  to  aircraft  noises.  Environmental  Pollu- 


1964.  The  breeding  birds  of  Mandarte  Island,  British 
Columbia.  Can.  Field-Naturalist  78:208-263. 

DUNNET,  G.M.  1977.  Observations  on  the  effects  of  low 
flying  aircraft  at  seabird  colonies  on  the  coast  of 
Aberdeenshire,  Scotland,  Biol.  Conserv.  12:55-63- 

ELLISON,  L.N.  and  L.  CLEARY.  1978.  Effects  of  human  dis- 
turbance on  breeding  double-crested  cormorants. 
Auk  95:510-517. 

EMLEN,  S.T.  and  N.J.  DEMONG  1975.  Adaptive  signifi- 
cance of  synchronized  breeding  in  a  colonial  bird:  A 
new  hypothesis.  Science  188:1029-1031. 

ENGLISH,  S.M.  1978.  Distribution  and  ecology-  of  Great 


Colonial  Waterbirds 


403 


Blue  Heron  colonies  on  the  Willamette  River,  Ore- 
gon. Pages  235-244  in  Sprunt,  AS.,  IV,  J.C.  Ogden, 
and  S.  Winckler,  eds.  Wading  birds.  Natl.  Audubon 
Soc,  Res.  Rep.  7. 

ERWIN,  R.M.  1979.  Coastal  waterbird  colonies:  Cape  Eliza- 
beth, Maine,  to  Virginia.  U.S.  Dept.  Inter.  Fish  Wildl. 
Serv.,  Biol.  Serv.  Prog.,  FWS/OBS-79/10. 

.  1980a.  Censusing  colonial  waterbirds:  Problems 

and  progress.  Atlantic  Naturalist  33:19-22. 

.  1980b.  Censusing  waterbird  colonies:  Some  sam- 
pling experiments.  Trans.  Linnaean  Soc.  New  York 
9:77-86. 

.  1981.  Censusing  wading  bird  colonies:  An  update 

on  "flight-line"  count  method.  Colonial  Waterbirds 
4:91-95. 

.  1982.  Observer  variability  in  estimating  numbers: 

An  experiment.  J.  Field  Ornithol.  53:159-167.1 

,  J.  GALLI,  and  J.  BURGER.  1981.  Colony  site  dy- 
namics and  habitat  use  in  Atlantic  coast  seabirds.  Auk 
98:550-561. 

and  C.E.  KORSCHGEN.  1979.  Coastal  waterbird 


colonies:  Maine  to  Virginia,  1977.  U.S.  Dept.  Inter. 
Fish  Wildl.  Serv.,  Biol.  Serv.  Prog.,  FWS/OBS-79/08. 

EVANS,  R.M.  1982.  Flock  size  and  formation  in  black-billed 
gulls.  Can.  J.  Zoology  60:1806-1811. 

FERNS,  P.N.  and  G.P.  MUDGE.  1981.  Accuracy  of  nest 

counts  at  a  mixed  colony  of  herring  and  lesser  black- 
backed  gulls.  Bird  Study  28:244-246. 

GALUSHA,  J.G.,  Jr.  and  CJ.  AMLANER,  Jr.  1978.  The  ef- 
fects of  diurnal  and  tidal  periodicities  in  the  numbers 
and  activities  of  herring  gulls  (Larus  argentatus)  in 
a  colony.  Ibis  120:322-328. 

GIBSON,  J.A.  1950.  Methods  of  determining  breeding  cliff 
populations  of  guillemots  and  razorbills.  British  Birds 
43:329-331. 

GILES,  L.W.  and  D.  B.  MARSHALL.  1954.  A  large  heron  and 
egret  colony  on  the  Stillwater  Wildlife  Management 
Area,  Nevada,  Auk  71:322-325. 

GILLET,  W.H.,  J.L.  HAYWARD,  Jr.,  AND  J.F.  STOUT.  1975. 
Effects  of  human  activity  on  egg  and  chick  mortality 
in  glaucous-winged  gull  colony.  Condor  77:492-495. 

GODFREY,  WE.  1966.  The  birds  of  Canada.  Nat.  Mus. 
Can.,  Bull.  203,  Biol.  Serv.  73. 

HAND,  J.L.  1980.  Human  disturbance  in  western  gull  (La- 
nds occidentalis  livens)  colonies  and  possible  amplifi- 
cation by  intraspecific  predation.  Biol.  Conserv. 
18:59-63. 

HARRIS,  MP.  1964.  Aspects  of  the  breeding  biology  of  the 
gulls  (Larus  argentatus,  L.  fiiscns,  and  L.  marinits). 
Ibis  106:432-456. 

HENNY,  C.J.  and  J.E.  KURTZ.  1978.  Great  blue  herons 
respond  to  nesting  habitat  loss.  Wildl.  Soc.  Bull. 
6:35-37. 

HERMAN,  S.G.  1980.  The  naturalist's  field  journal:  A  man- 
ual of  instruction  based  on  a  system  established  by 
Joseph  Grinnell.  Published  privately,  S.G.  Herman, 
Biol.  Dep.,  The  Evergreen  State  College,  Olympia,  WA. 
200pp. 

HUNT,  G.L.,  Jr.  1972.  Influence  of  food  distribution  and 
human  disturbance  on  the  reproductive  success  of 
herring  gulls  Ecol.  53:1051-1061. 

and  M.W.  HUNT.  1975.  Reproductive  ecology  of 

the  western  gull:  The  importance  of  nest  spacing.  Auk 
92:270-279. 

and .  1976.  Gull  chick  survival.  The  signifi- 
cance of  growth  rates,  timing  of  breeding  and  terri- 
tory size.  Ecol.  57:62-75. 

JEHL,  JR.,  Jr.  1973  Studies  of  a  declining  population  of 


brown  pelicans  in  northwest  Baja,  California.  Condor 

75:69-79. 
JOHNSON,  R.A.  1938.  Predation  of  gulls  in  murre  colonies. 

Wilson  Bull.  50:161-170. 
KADLEC,  J.A  and  W.H.  DRURY.  1968a.  Structure  of  the 

New  England  herring  gull  population.  Ecol. 

49:644-676. 
and .  1968b.  Aerial  estimation  of  the  size  of 

gull  breeding  colonies.  J.  Wildl.  Manage.  32:287-293. 
KORSCHGEN,  C.E.  1979.  Coastal  waterbird  colonies: 

Maine.  U.S.  Dept.  Inter.,  Fish  Wildl.  Serv.,  Biol.  Serv. 

Prog.,  FWS/OBS-70/07. 
KREBS,  JR.  1974.  Colonial  nesting  and  social  feeding  as 

strategies  for  exploiting  food  resources  in  the  great 

blue  heron  (Ardea  herodias).  Behaviour  51:99-134. 
.  1978.  Colonial  nesting  in  birds,  with  special  refer- 
ence to  the  Ciconiformes.  Pages  299-311  in  Sprunt, 

AS.,  IV,  J.C.  Ogden,  and  S.  Winckler,  eds.  Wading 

Birds.  Natl.  Audubon  Soc,  Res.  Rep.  7. 
KURY,  C.R.  and  M.  GOCHFELD.  1975.  Human  interference 

and  gull  predation  in  cormorant  colonies.  Biol.  Con- 
serv. 8:23-24. 
KUSHLAN,  J.A.  1978.  Feeding  ecology  of  wading  birds. 

Pages  249-297  in  Sprunt,  A.S.,  IV,  J.C.  Ogden,  and  S. 

Winckler,  eds.  Wading  Birds.  Natl.  Audubon  Soc,  Res. 

Rep.  7. 
-.  1979.  Effects  of  helicopter  censuses  on  wading 

bird  colonies.  J.  Wildl.  Manage.  43:756-760. 
LLOYD,  C.S.  1972.  Attendance  at  auk  colonies  during  the 

breeding  season.  Skokholm  Bird  Observation  Rep. 

1972:15-23. 
LOWE,  FA.  1954.  The  heron.  Collins,  St.  James  Place, 

London. 
MANUWAL,  DA.  1974a.  Effects  of  territoriality  on  breed- 
ing in  a  population  of  Cassin's  Auklet.  Condor 

55:1399-1406. 
.  1974b.  The  natural  history  of  Cassin's  auklet  (Pty- 

choramphns  aleuticus).  Condor  76:421-431- 
.  1978.  Effect  of  man  on  marine  birds:  a  review. 


Proc  J.S.  Wright  Forestry  Conf.  4:140-160. 

MASSEY,  B.W.  and  J.L  ATWOOD.  1981.  Second-wave 

nesting  of  the  California  least  tern:  Age  composition 
and  reproductive  success.  Auk  98:596-605. 

MCCRIMMON,  DA.  1976.  A  review  of  some  methods  and 
considerations  for  the  assessment  of  breeding  popula- 
tions of  colonial  waterbirds.  Proc.  4th  Ann.  Meeting, 
Texas  Fish-Eating  Bird  Conf.  1976:36-49. 

MCCRIMMON,  DA,  Jr.  1978a.  Nest-site  characteristics 
among  five  species  of  herons  on  the  North  Carolina 
coast.  Auk  95:267-280. 

.  1978b.  The  collection,  management,  and  exchange 

of  information  on  colonially  nesting  birds.  Pages  187- 
196  in  Sprunt,  A.S.,  IV.  J.C.  Ogden,  and  S.  Winckler, 
eds.  Wading  Birds,  Natl.  Audubon  Soc,  Res.  Rep.  7. 

MEYERRIECKS,  A.J.  I960.  Comparative  breeding  behavior 
of  four  species  of  North  American  herons.  Publ.  Nut- 
tal  Ornith.  Club  2. 

NESBITT,  S.A.,  J.C.  OGDEN,  H.W.  KALE,  II,  B.W.  PATTY, 
and  LA.  ROWSE.  1982.  Florida  atlas  of  breeding  sites 
for  herons  and  their  allies:  1976-78.  U.S.  Dept.  Inter., 
Fish  Wildl.  Serv.,  Biol.  Serv.  Prog.,  FWS/OBS-81/49. 

NETTLESHIP,  D.N.  1975.  A  recent  decline  of  gannets, 
Morns  bassanus,  on  Bonaventure  Island,  Quebec. 
Can.  Field-Naturalist  89:125-133- 

.  1976.  Census  techniques  for  seabirds  of  arctic  and 

eastern  Canada.  Can.  Wildl.  Serv.,  Occ  Pap.  25. 

.  1980.  Census  methods  for  murres,  Uria  species:  A 


unified  approach.  Can.  Wild.  Serv.,  Occ.  Pap.  43- 


404 


Colonial  Waterbirds 


NISBET,  I.C.T.  1973-  Terns  in  Massachusetts:  Present  num- 
bers and  historical  changes.  Bird-Banding  44:27-55. 
OGDEN,  J.C.  1978.  Recent  population  trends  of  colonial 

wading  birds  on  the  Atlantic  and  Gulf  coastal  plains. 

Pages  137-153  in  Sprunt,  AS.,  IV,  J.C.  Ogden,  and 

S.  Winckler,  eds.  Wading  Birds.  Natl.  Audubon  Soc, 

Res.  Rep.  7. 
PALMER,  R.S.  (ed.).  1962.  Handbook  of  North  American 

birds.  Vol.  1.  Loons  through  flamingos.  Yale  Univ. 

Press.,  New  Haven,  CT.  567pp. 
PARSON,  J.  1971.  Cannibalism  in  herring  gulls.  British 

Birds  64:528-537. 
PATTERSON,  I.J.  1965.  Timing  and  spacing  of  broods  in 

the  black-headed  gull,  Lanis  ridibundus.  Ibis 

107:433-459. 
PENLAND,  S.  1981.  Natural  history  of  the  Caspian  tern  in 

Grays  Harbor,  Washington.  Murrelet  62:66-72. 
.  1982.  Distribution  and  status  of  the  Caspian  tern 

in  Washington  State.  Murrelet  63:73-79. 
PORTNOY,  J.W.  1977.  Nesting  colonies  of  seabirds  and 

wading  birds —  coastal  Louisiana,  Mississippi,  and 

Alabama.  U.S.  Dept.  Inter.,  Fish  Wild!.  Serv.,  Biol.  Serv. 

Prog.,  FWS/OBS-77/07. 
.  1980.  Censusing  methods  for  gulf  coast  waterbirds. 

Trans.  Linnean  Soc.  New  York  9:127-134. 
-,  R.M.  ERWIN,  and  T.W.  CUSTER.  1981.  Atlas  of  gull 


and  tern  colonies.  North  Carolina  to  Key  West,  Flor- 
ida (including  pelicans,  cormorants,  and  skimmers). 

U.S.  Dept.  Inter.,  Fish  Wildl.  Serv.,  Biol.  Serv.  Prog., 

FWS/OBS-80/05. 
PRATT,  H.M.  1970.  Breeding  biology  of  great  blue  herons 

and  common  egrets  in  central  California.  Condor 

72:407-416. 
.  1972.  Nesting  success  of  common  egrets  and  great 

blue  herons  in  the  San  Francisco  Bay  region.  Condor 

74:447-453. 
.  1973.  Breeding  attempts  by  juvenile  great  blue 

herons.  Auk  90:897-899. 
.  1974.  Breeding  of  great  blue  herons  and  great 

egrets  at  Audubon  Canyon  Ranch,  California.  Western 

Birds  5:127-136. 

1980.  Directions  and  timing  of  great  blue  heron 


foraging  flights  from  a  California  colony:  Implications 
for  social  facilitation  of  food  finding.  Wilson  Bull. 
92:489-496. 

RALPH,  C.J.  1981.  Terminology  used  in  estimating  num- 
bers of  birds.  Studies  Avian  Biol.  6:577-578. 

ROBERT,  H.C.  and  C.J.  RALPH.  1975.  Effects  of  human 

disturbance  on  the  breeding  success  of  gulls.  Condor 
77:495-499. 

RODGERS,  J.A.,  Jr.  1978.  Display  characteristics  and  fre- 
quency of  breeding  by  subadult  Little  Blue  Herons. 
Pages  35-39  in  Sprunt,  A.S.,  IV,  J.C.  Ogden,  and  S. 
Winckler,  eds.  Wading  Birds.  Natl.  Audubon  Soc,  Res. 
Rep.  7. 

RYDER,  R.A.  1978.  Breeding  distribution,  movements,  and 
mortality  of  snowy  egrets  in  North  America.  Pages 
197-205  in  Sprunt,  AS.,  IV,  J.C.  Ogden,  and  S.  Winck- 
ler, eds.  Wading  Birds.  Natl.  Audubon  Soc,  Res. 
Rep.  7. 

SCHREIBER,  R.W.  1979.  Reproductive  performance  of  the 
eastern  brown  pelican.  Contr.  Sci.,  Natl.  Hist.  Mus. 
Los  Angeles  Co.  CA.317T-43. 

and  E.A.  SCHREIBER.  1972.  Studies  of  the  brown 

pelican.  Wilson  Bull.  84:119-135. 

SEALY,  S.G.  and  J.  BEDARD.  1973.  Breeding  biology  of  the 
parakeet  auklet  (Cyclorrhynchus psittacula)  on  St. 
Lawrence  Island,  Alaska.  Astarte  6:59-68. 


SIDLE,  J.G.  and  EL.  FERGUSON.  1982.  White  pelicans 
populations  at  Chase  Lake,  North  Dakota,  evaluated 
by  aerial  photography.  Prairie  Naturalist  14:13-26. 

SLATER,  J. B.  1980.  Factors  affecting  the  numbers  of  guille- 
mots ilria  aalge  present  on  cliffs.  Ornis  Scand. 
11:155-163. 

SOUTHERN,  W.F.  and  L.K  SOUTHERN.  1981.  Colony 

census  results  as  indicators  of  pre-hatching  perturba- 
tions. Colonial  Waterbirds  4:143-149- 

SOWLS,  A.L.,  A.R.  DE  GANGE,  J.W.  NELSON,  and  GS.  LES- 
TER. 1980.  Catalog  of  California  seabird  colonies.  U.S. 
Dept.  Inter.,  Fish  Wildl.  Serv.,  Biol.  Serv.  Prog.,  FWS/ 
OBS-80/37. 

,  S.M.  HATCH,  and  C.J.  LENSINK  1978.  Catalog  of 

Alaskan  seabird  colonies.  U.S.  Dept.  Inter.,  Fish  Wildl. 
Serv.,  Biol.  Serv.  Prog.,  FWS/OBS-78/78. 

SPEICH,  S.M.  and  DA.  MANUWAL.  1974.  Gular  pouch 
development  and  population  structure  of  Cassin's 
auklet.  Auk  91:291-306. 

and  T.R.  WAHL.  1985.  Catalog  of  Washington  sea- 
bird  colonies.  U.S.  Dept.  Inter.,  Fish  Wildl.  Serv.,  Biol. 
Serv.  Prog.,  FWS/OBS.  In  press. 

STOWE,  J.J.  1982.  Recent  population  trends  in  cliff-breed- 
ing seabirds  in  Britain  and  Ireland.  Ibis  124:502-510. 

THOMPSON,  B.C.  and  J.E.  TABOR.  1981.  Nesting  popula- 
tions and  breeding  chronologies  of  gulls,  terns,  and 
herons  on  the  upper  Columbia  River,  Oregon  and 
Washington.  Northwest  Sci.  55:209-218. 

THOMPSON,  L.S.  1981.  Nest-tree  sharing  by  herons  and 
cormorants  in  Montana.  Can.  Field-Naturalist 
95:257-260. 

.  1982.  A  1979  census  of  Great  Blue  Heron  colonies 

in  Montana.  Proc  Montana  Acad.  Sci.  41:23-27. 

THOMPSON,  S.P.,  CD.  LITTLEFIELD,  and  R.A.  RYDER. 

1979.  Historical  review  and  status  of  colonial  nesting 
birds  on  Malheur  National  Wildlife  Refuge,  Oregon. 
Proc.  Colonial  Waterbird  Group  3:156-164. 

THORESEN,  AC.  1964.  The  breeding  behavior  of  the 
Cassin's  auklet.  Condor  66:456-476. 

TREMBLAY,  J.  and  L.N.  ELLISON.  1979.  Effects  of  human 
disturbance  on  breeding  of  black-crowned  night  her- 
ons. Auk  96:364-369. 

TUCK,  L.K  1961.  The  murres.  Can.  Wildl.  Serv. 
Monogr.  1. 

UDVARDY,  M.D.F.  1969.  Dynamic  zoogeography:  with 
special  reference  to  land  animals.  Van  Nostrand 
Rheinhold  Co.,  New  York.  445pp 

VAN  TETS,  G.F.  1965.  A  comparative  study  of  some  social 
communication  patterns  in  the  Pelecaniformes.  Am. 
Ornith.  Union,  Ornith.  Monogr.  2. 

VERBEEK,  N.A.M.  1982.  Egg  predation  by  northwest 
crows:  Its  association  with  human  and  bald  eagle 
activity.  Auk  99:347-352. 

VERMEER,  K  1963.  The  breeding  ecology  of  the  glau- 
cous-winged gull  (Larus  occidentalism  on  Mandarte 
Island,  B.C.  British  Columbia  Province  Mus.,  Occ  Pap. 
13. 

WEISE,  J.H.  1978.  Heron  nest-site  selection  and  its  ecolog- 
ical effects.  Pages  27-34  in  Sprunt,  AS.,  IV,  J.C.  Og- 
den, and  S.  Winckler,  eds.  Wading  Birds.  Natl. 
Audubon  Soc,  Res.  Rep.  7. 

WERSCHKUL,  D.F.,  E.  MC  MAHON,  and  M.  LEITSCHUH. 
1976.  Some  effects  of  human  activities  on  the  great 
blue  heron  in  Oregon.  Wilson  Bull.  88:660-662. 

, , ,  S.  ENGLISH,  C.  SKIBINSKI,  and  G. 

WILLIAMSON.  1977.  Observations  on  the  reproduc- 
tive ecology  of  the  great  blue  heron  {Ardea  herodias) 
in  western  Oregon.  Murrelet  58:7- 12. 


Colonial  Waterbirds 


405 


20 


Upland  Game 
Birds 


Robert  L.  Eng 

Biology  Department 
Montana  State  University 
Bozeman,  MT  59717-0001 


Editor's  Note:    Upland  game  birds,  like  other  eco- 
nomically important  species,  have  received  much 
attention.  And  even  though  most  of  these  species 
are  avidly  hunted,  in  most  cases  habitat  quantity 
and  quality  is  the  factor  controlling  abundance 
and  distribution.  Although  much  research  has  been 
conducted  on  upland  game  bird  habitat,  much  of 
this  has  focused  on  species  associated  with  agricul- 
ture such  as  pheasants,  rather  than  on  wildland 
species.  For  example,  the  effects  of  logging  grazing 
and  recreation  on  species  of  grouse  are  poorly 
understood  Therefore  enlightened  management  of 
habitat  for  these  species  will  require  better  identifi- 
cation of  limiting  habitat  factors,  together  with 
careful  inventory  and  monitoring  of  habitats  and 
populations. 


INTRODUCTION 

Usually  a  reference  to  upland  game  birds  is  an 
exclusive  reference  to  members  of  the  family  Phas- 
ianidae:  non-migratory,  chicken-like  birds  including 
the  partridge,  grouse,  turkey,  and  quail.  Recently, 
increasing  mention  may  be  found  to  migratory  up- 
land game  birds  in  reference  to  members  of  the  fam- 
ily Columbidae,  the  pigeons  and  doves. 

Emphasis  here  will  be  placed  on  members  of 
the  family  Phasianidae,  primarily  because  their  non- 
migratory  status  requires  a  more  labor-intensive  in- 
ventory and  assessment  of  year-long  habitat  needs. 
Also,  within  this  group,  emphasis  will  be  placed  on 
those  members  in  the  western  U.S.  that  inhabit  non- 
cultivated  lands,  at  least  for  a  major  portion  of  the 
year. 

Although  life  histories  or  the  biology  of  each 
species  will  not  be  dealt  with  here,  some  basic 
knowledge  of  a  species'  life  history  is  necessary  be- 
fore sound  management  decisions  can  be  made. 

This  does  not  imply  the  need  for  an  in-depth 
study  of  each  species.  Rather,  many  species  have 
certain  habits  similar  enough  to  permit  the  use  of 
common  denominators  in  monitoring  the  birds  or 
their  habitat.  For  example,  most  quail  and  partridge 
feed  heavily  on  seeds  from  a  variety  of  annuals  that 
may  show  wide  year-to-year  fluctuations  and  are 
often  products  of  an  early  serai  stage.  These  birds 
have  evolved  with  a  very  high  reproductive  poten- 
tial, including  a  high  rate  of  first-year  breeding  and 
large  clutch  sizes,  and  thus  can  quickly  respond  to  a 
bountiful  food  supply.  Conversely,  blue  grouse  (Den- 
dragapus  obscurus)  and  sage  grouse  (Centrocercus 
urophasianus)  feed  during  the  winter  on  the  vegeta- 
tive parts  of  a  few  species  of  dominant  plants,  a  food 
supply  that  is  produced  in  a  climax  or  near  climax 
stage,  and  shows  comparatively  little  annual  variation 
in  abundance.  These  birds  have  a  relatively  low  re- 
productive potential  with  respect  to  age  of  consist- 


Upland  Game  Birds 


407 


ent  first  breeding  and  clutch  size;  this  appears  to 
be  in  keeping  with  their  more  stable  food  supply. 
The  contrasts  in  the  two  facets  of  life  history  of 
these  two  groups  would  result  in  equally  contrasting 
habitats  and  habitat  management  concerns. 

That  there  is  a  close  relationship  between  the 
quality  of  a  bird's  habitat  and  its  welfare  (and  result- 
ing population  status)  requires  little  explanation.  The 
difficulty  arises  in  quantifying  the  quality  of  a  habitat 
which  for  a  given  species  may  be  found  in  a  contin- 
uum from  marginal  to  excellent.  Available  in  the 
literature  are  qualitative  descriptions  of  game  bird 
habitats  listing  food  and  cover  components  which 
may  be  preferred  or  required  by  the  species  in  ques- 
tion. When  measurements  are  taken  of  the  amount 
and  distribution  of  these  components  within  a  unit 
area,  the  description  assumes  a  quantitative  air.  Fi- 
nally, with  the  assistance  of  a  series  of  assumptions, 
numbers  are  assigned  to  the  descriptions  and  meas- 
urements, and  a  numerical  rating  is  assigned  placing 
that  particular  piece  of  habitat  at  a  fixed  point  along 
the  continuum.  However,  the  quality  of  the  overall 
habitat,  which  may  rate  high  as  a  composite,  may  in 
reality  rate  rather  low  as  a  result  of  poor  quality  of 
one  facet.  For  example,  two  areas,  one  a  homogene- 
ous stand  of  smooth  brome  and  the  other  85% 
smooth  brome  and  15%  forbs,  may  both  be  re- 
corded as  ungrazed,  unmowed  pastures  of  smooth 
brome.  However,  the  field  with  a  sprinkling  of  forbs 
would  provide  better  cover  and  a  greater  diversity  of 
insects,  making  it  a  superior  habitat  for  brood 
rearing. 

The  above  is  not  intended  to  entirely  discredit 
efforts  at  measuring  the  quality  of  game  bird  habitats. 
Certainly,  the  available  descriptions  provide  land 
managers  with  the  capability  to  recognize  the  gen- 
eral habitat  type  in  which  they  could  expect  to  find 
the  various  species.  They  could  also,  with  some  ex- 
perience, place  an  assessment  of  quality  within  a 
broad  scale.  However,  a  more  realistic  measure  of 
the  quality  can  be  attained  through  a  measure  of  the 
bird's  response  to  its  habitat,  the  home  range  size, 
or  the  population  density. 

The  average  size  of  a  home  range,  whether  sea- 
sonal or  annual,  when  compared  between  habitats, 
will  reflect  the  quality  of  a  given  habitat.  Birds  in 
high  quality,  secure  habitat  tend  to  have  smaller 
home  ranges  than  those  in  habitats  of  lesser  quality. 
Unfortunately,  home  range  determination  is  time  and 
budget  consuming,  usually  involving  trapping,  mark- 
ing, and  relocating  individual  birds. 

Census  techniques,  whereby  relative  densities 
can  be  determined,  present  a  far  more  workable 
solution.  Most  techniques  permit  only  an  index  to 
the  population;  but  if  conducted  by  standardized 


methods,  data  between  areas  or  time  periods  provide 
workable  comparisons.  For  the  land  managers  inter- 
ested in  habitat  quality,  population  density  compari- 
sons between  areas  provide  a  measure  of  the  various 
areas  to  support  birds  of  that  species.  Similarly,  com- 
parable data  gathered  from  a  single  area  over  a  pe- 
riod of  several  years  may  provide  a  measure  of  the 
population  trend  which  in  turn  could  reflect  changes 
in  habitat  quality. 


HABITAT  FEATURES  CORRELATED  WITH 
SPECIES  GROUPS 

Major  habitat  characteristics  of  the  various  spe- 
cies are  presented  in  summary  form  with  selected 
references  in  Table  1.  Supplemental  statements  to  as- 
sist in  habitat  identification  are  presented  in  the  text. 

Gray  Partridge  (Perdix  perdix) 

The  European  gray  or  Hungarian  partridge  is 
one  of  three  exotic  game  birds  successfully  intro- 
duced into  the  U.S.  The  others  are  the  ring-necked 
pheasant  (Phasianus  colcNcus)  and  the  chukar 
(Alectoris  chukar).  The  "nun"  is  native  to  Europe 
and  Asia  although  North  American  stock  appears  to 
have  originated  largely  in  central  Europe. 


Although  present  in  a  rather  wide  variety  of 
topographic  conditions,  the  more  dense  populations 
of  gray  partridges  are  generally  associated  with  flat 
or  moderately  rolling  land.  Stiehl  (1984)  referred  to 
three  geographic  populations  in  North  America: 
western,  central,  and  the  Great  Lakes. 


Gray  partridge  are  frequently  associated  with  a 
combination  of  cultivated  and  non-cultivated  lands. 
Weigand  (1980)  reported  land  uses  from  his  study 
area,  in  decreasing  order  of  abundance,  as  small 
grain  and  fallow  land,  rangeland,  hayfields,  and  agri- 
culturally idle  areas.  Small  grain  fields  and  fallow 
lands  were  used  heavily  for  feeding  (both  seeds  and 
green  vegetation)  during  a  large  part  of  the  year 
whereas  the  other  types  provided  an  herbaceous 
cover  primarily  for  loafing,  roosting,  and  nesting. 
Their  affinity  for  edges  probably  accounts  for  their 
great  ability  to  successfully  cope  with  modern  agri- 
culture (Allen  1984). 


Weigand  (1980)  found  wheat  and  barley  seeds 
and  leaves  a  major  food  component  throughout  the 
year;  seeds  and  leaves  of  such  forbs  as  dandelion 
(Taraxacum  officinale),  knotweeds  (Polygonum  sp.), 
mustard  ( Brassica  kaber),  and  white  clover  (  Trifol- 
ium  repens)  were  prominent  at  various  seasons.  He 
also  reported  that  of  10  nests  located,  5  were  con- 
cealed by  grass,  the  others  by  forbs. 


408 


Upland  Game  Birds 


Table  1.     Synopsis  of  habitat  features  for  species  or  species  groups. 


Species  or 
Species  Group 

Habitat  Characteristics 

(Physical  features,  vegetation,  and  species  composition) 

References 

Gray  partridge 

(Perdix  perdix) 

Agricultural  lands,  primarily  small  grains,  interspersed  with  idle 
areas  Croplands  provide  major  food  source  and  idle  areas  pro- 
vide loafing,  roosting,  and  nesting  cover.  Nesting  cover  often 
in  grassland/shrub  areas — successful  nesting  efforts  dependent 
on  residual  cover  from  the  previous  year.  Seeds  from  cultivated 
grains  and  a  variety  of  forbs  and  other  green  succulent  vegeta- 
tion provide  primary  food 

Johnsgard  1973,  Wei- 
gand  1980;  Smith  et  al. 
1982;  Mendel  and  Peter- 
son 1983;  Potts  1984; 
Stiehl  1984 

Chukar 

(Alectoris  chukar) 

Rugged,  rocky  terrain  in  arid  to  semi-arid  climates,  about  equally 
divided  between  rocky  outcrops  and  low  shrub/grassland.  Low 
winter  temperatures  are  less  devastating  than  persistent  snow 
cover.  Water  availability  may  significantly  affect  summer  distribu- 
tion. Rocky  outcrops  may  provide  loafing  and  roosting  cover. 
Nesting  cover  similar  to  shrub/grassland  types  used  for  foraging, 
often  on  a  south  facing  slope  up  from  creek  bottoms  Sagebrush 
(Artemisia  sp),  cheatgrass  (Bromus  tectorum),  and  wheatgrass 
(Agropyron  sp  )  are  common  species  with  cheatgrass  seeds  and 
leaves  providing  a  major  food  source. 

Galbreath  and  Moreland 
1953;  Bohl  1957;  Harper 
et  al.  1958;  Christensen 
1970;  Molini  1976. 

Ring-necked  pheasant 
(Phasianus  colchicus) 

Agricultural  land  highly  interspersed  with  non-cultivated  land  (i.e., 
dense  clumps  of  grasses,  forbs,  and  shrubs).  Non-cultivated 
areas  (preferably  idle  areas  such  as  roadsides,  railroad  rights-of- 
way,  and  temporary  wetlands)  are  necessary  for  nesting,  roosting, 
and  loafing  cover,  particularly  during  winter  and  early  spring 
when  crops  are  removed.  Farm  crops,  primarily  waste  cereal 
grains,  may  account  for  over  75%  of  the  food  with  weed  seeds 
associated  with  cultivation  making  up  another  10-15%. 

Hiatt  1947;  Yeager  et  al. 
1951;  Snyder  1974; 
Weigand  and  Janson 
1976. 

Spruce  grouse 

(Dendragapus 
canadensis) 

Coniferous  forests  of  Canada  and  northern  U.S.  A  trend  toward 
increased  use  of  upland  over  lowland  areas  from  eastern  to 
western  distribution.  Preference  shown  for  coniferous  stands  that 
are  neither  mature  nor  extremely  dense,  thus  permitting  adequate 
understory  for  food  and  cover.  Spruce  (Picea  sp.)  and  jackpine 
(Pinus  banksiana)  in  Michigan  or  lodgepole  pine  (Pinus  contorta) 
in  the  mountain  west  are  key  to  the  species.  A  variety  of  berries 
(Vaccinium  sp.,  Symphoncarpus  sp.,  Rubus  sp.)  are  important 
summer  foods  whereas  coniferous  needles  (Picea,  Pinus,  Larix) 
dominate  the  winter  diet. 

Stoneberg  1967;  Pender- 
gast  and  Boag  1970; 
Robinson  1980;  Ratti  et 
al.  1984. 

Blue  grouse 

(Dendragapus  obscurus) 

Coniferous  forests  of  western  and  coastal  mountain  ranges  Win- 
ter at  high  elevations  in  stands  of  conifers;  in  spring,  found  at 
lower  elevations  at  forest-grassland  ecotone,  while  females  with 
broods  may  follow  brushy  draws  2  to  3  miles  (3  to  5  km)  below 
conifers.  Prefer  open,  dry  sites  but  near  clumps  of  conifers  for 
escape  cover.  Structural  height  of  vegetation  at  least  8  inches  (20 
cm)  preferred  in  brood  rearing  areas.  Douglas  fir  (Pseudotsuga 
menziesn)  a  key  cover  and  food  plant  (winter).  Berries  (Vaccinium 
sp.,  Symphoncarpos  sp.)  and  leafy  forage  (Trlfolium  sp.,  Fragana 
sp.)  are  heavily  used  at  other  times  of  the  year. 

Mussehl  1963;  Bendell 
and  Elliott  1966;  Martinka 
1972. 

Ptarmigan 
Willow 

(Lagopus  lagopus) 

Rock 

(Lagopus  mutus) 

White-tailed 

(Lagopus  leucurus) 

Willow  ptarmigan  inhabit  open  tundra  and  show  a  preference  for 
areas  with  level  to  moderate  terrain,  heavily  vegetated  with 
grasses,  forbs,  and  shrubs.  Rock  ptarmigan  frequent  the  open 
tundra  but  on  more  hilly  terrain  and  with  less  luxuriant  vegetation 
than  utilized  by  the  willow  ptarmigan.  White-tailed  ptarmigan 
occupy  alpine  tundra,  ridges,  and  steep  slopes  where  vegetation 
is  sparse  and  dwarfed;  rocks  provide  the  structural  cover.  Willow 
(Salix  sp.)  is  a  regular  shrub  for  all  three  species,  the  buds  and 
twigs  being  dominant  food  for  a  large  part  of  the  year.  Also  used 
are  a  variety  of  leaves,  flowers,  and  berries  as  they  become 
available  during  a  relatively  short  growing  season. 

Choate  1963;  Weeden 
1963,  1965,  1967;  Johns- 
gard 1973. 

Upland  Game  Birds 


409 


Table  1.     Synopsis  of  habitat  features  for  species  or  species  groups  (continued). 


Species  or 
Species  Group 

Habitat  Characteristics 

(Physical  features,  vegetation,  and  species  composition) 

References 

Ruffed  grouse 

(Bonasa  umbellus) 

Deciduous  forests  or  mixed  deciduous/coniferous  forest  edges.  In 
the  Mountain  West,  this  species  is  more  prevalent  along  stream 
courses  or  other  moist  areas.  Drumming  males  prefer  areas  of 
high  vegetative  density  (overhead  cover  important)  whereas 
broods  prefer  more  open  stands  but  with  high  degrees  of  herba- 
ceous cover.  Aspen  (Populus  tremuloides  or  P.  grandidentata) 
key  species  in  preferred  stands  A  preference  shown  for  early 
serai  stages  of  aspen  although  an  interspersion  of  different  age 
classes  desirable  for  yearlong  food  source  (i.e.,  aspen  buds  from 
mature  stands,  berries,  fruits,  and  leafy  forage  from  younger 
stands). 

Svoboda  and  Gullion 
1972;  Gullion  1977,  1984; 
Stauffer  and  Peterson 
1985,  1985b;  Kubisiak 
1985. 

Sage  grouse 

(Cenfrocercus 
urophasianus) 

Semi:arid  sagebrush  or  sagebrush/grassland  types  of  the  western 
plains  and  intermountain  basins  at  elevations  from  750  m  (2,500 
ft)  to  2,100  m  (7,000  ft)   Preference  for  gentle  topography  over 
steep-sided  canyons  or  slopes.  Sagebrush  is  essential  to  survival 
with  stands  of  greatest  canopy  coverage  (20-35%)  preferred  in 
winter  and  least  (5-10%)  in  summer.  Food  is  reflected  in  this 
seasonal  cover  preference:  sagebrush  in  winter,  a  variety  of  suc- 
culent forbs  in  spring  and  summer. 

Patterson  1952;  Eng  and 
Schladweiler  1972;  Wal- 
lestad  1975;  Wallestad  et 
al.  1975. 

Greater  prairie  chicken 

(Tympanuchus  cupido) 

Arid  native  grasslands  of  the  Southwest  intermixed  with  shrub  or 
half-shrub  component.  A  portion  (5-40%)  of  an  area  in  cultivation, 
if  well  interspersed,  may  enhance  the  habitat  through  increased 
food  source.  Shrubs,  less  than  1  m  (3  ft)  tall  in  savannah  aspect, 
are  desirable  although  shrub  species  may  determine  preferred 
shrub  density;  high  densities  of  sand  sagebrush  (Artemisia  fili- 
folia)  appear  more  desirable  than  high  densities  of  shinnery  oak 
(Quercus  havardii). 

Jones  1963;  Crawford 
and  Bolen  1976;  Davis  et 
al.  1980;  Cannon  and 
Knopf  1981a;  Doerr  and 
Guthery  1983. 

Sharp-tailed  grouse 

(Tympanuchus 
phasianellus) 

Grasslands  that  have  a  prominent  deciduous  shrub  or  woodland 
component.  Although  both  western  races  (T.  p.  columbianus 
and  jamesi)  inhabit  a  grassland/shrub/tree  complex,  the  Colum- 
bian sharptail  is  usually  found  in  habitat  with  a  higher  overall 
shrub/tree  cover.  Columbian  sharptails  are  associated  with  a  vari- 
ety of  shrubs  and  trees  (sagebrush,  hawthorne  [Crataegus  sp.], 
serviceberry  [Amelanchier  sp],  Gambel  oak  [Quercus  gambellii], 
and  aspen)  which  may  be  used  for  cover  and  produce  food  in 
the  form  of  berries  or  winter  buds.  A  similar  group  of  plants  exists 
for  the  prairie  race  (juniper  [Juniperus  sp.],  buffaloberry  [Shep- 
herdia  sp.]),  cottonwood,  and  aspen. 

Miller  and  Graul  1980; 
Nielson  and  Yde  1981. 

Wild  turkey 

{Meleagns  gallopavo) 

Open,  mature,  or  nearly  mature  forests,  deciduous  but  often 
deciduous-coniferous,  particularly  in  mountainous  areas.  Mature 
forests  provide  a  more  dependable  mast  crop  and  permit  greater 
visibility.  Merriam's  turkey  shows  a  preference  during  spring, 
summer,  and  fall  for  ponderosa  (Pinus  ponderosa)  forests  with  an 
open  grassland  understory;  in  winter  it  often  frequents  a  lower 
elevation  in  a  mast  producing  habitat  or  in  the  north,  grain  fields. 
The  Rio  Grande  race,  found  in  more  arid  areas  to  the  south,  often 
occupies  river  bottom  habitat,  but  also  oak-grassland  intersper- 
sions.  Scattered  openings  in  forest  habitat  (10-40%  of  the  total) 
are  highly  desirable.  Mast  producers  such  as  oak  (Quercus  sp.), 
pinyon  pine  (Pinus  edulis),  and  juniper  and  seeds  and  leaves 
from  a  variety  of  grasses  are  major  food  sources. 

Ligon  1946;  Spicer  1959; 
Hoffman  1962;  Jonas 
1966;  Beasom  1970. 

Mearn's  quail 
(Cyrtonyx  montezuma 
mearnsi) 

Oak-grasslands  and  pine-oak  woodlands  ranging  from  1,100  to 
3,600  m  (4,000  to  12,000  ft)  in  the  highlands  of  the  Southwest. 
Preference  is  shown  for  open  woodlands  with  a  grass-forb  un- 
derstory. Grass  is  used  commonly  for  cover;  the  bulbs  and  seeds 
of  woodsorrel  (Oxalis)  and  flat  sedges  (Cyperus)  are  prominent 
foods.  During  years  of  scarcity  of  forbs  and  sedges,  acorns  may 
be  a  common  food. 

Leopold  and  McCabe 
1957;  Ligon  1961;  Bishop 
and  Hungerford  1965; 
Brown  1979;  Brown  1982 

410 


Upland  Game  Birds 


Table  1.     Synopsis  of  habitat  features  for  species  or  species  groups  (concluded). 


Species  or 
Species  Group 

Habitat  Characteristics 

(Physical  features,  vegetation,  and  species  composition) 

References 

Masked  bobwhite 

(Colinus  virginianus 
ndgwayi) 

Open  desert  grasslands  at  elevations  from  300  to  1 ,200  m  (1 ,000 
to  4,000  ft).  Although  mesquite  (Prosopsis  sp.)  and  catclaw  (Aca- 
cia sp.)  are  frequently  found  in  the  area,  this  bird  seemingly 
prefers  the  grassy  areas  adjacent  to  shrubs.  Coarse  grasses 
(Sporobolus  sp.)  provide  cover  whereas  species  of  grama  (Bou- 
teloua  sp.)  and  three-awns  (Anstida  sp.)  provide  both  food  and 
cover. 

Tomlinson  1972a,  b; 
Johnsgard  1973. 

Scaled  quail 

{Callipepla  squamata) 

Desert  grassland,  usually  below  2,100  m  (6,900  ft),  composed  of 
low  growing  grasses,  forbs,  and  shrubs  with  an  overall  ground 
cover  between  10  and  50%.  Shrubs  used  for  overhead  cover  with 
preference  shown  for  those  less  than  2  m  (7  ft).  As  little  as  10% 
of  an  area  left  in  clumps  of  mesquite  will  provide  cover  and  food 
(seeds).  Major  vegetative  foods  consist  of  conspicuous  forb 
and  shrub- seeds  such  as  croton  (Croton  sp.),  cycloloma  (Cyclo- 
loma  atnplicifolium),  snakeweed  (Gutierrezia  sarothrae)  and 
mesquite. 

Campbell  et  al.  1973, 
Davis  et  al.  1975;  Good- 
win and  Hungerford 
1977,  Ault  and  Stormer 
1983 

Gambel's  quail 

(Callipepla  gambelii) 

Desert  shrub  type  usually  in  habitats  below  1,800  m  (6,000  ft) 
Preference  is  shown  for  shrub  cover  of  sufficient  density  to  shade 
50  to  75%  of  the  ground  and  where  60  to  80%  of  the  shrubs  are 
taller  than  2  m.  Mesquite,  hackberry  (Celtis  reticulata),  wolfberry 
(Lycium  sp),  and  catclaw  provide  cover  for  most  of  the  bird's 
activities  Many  preferred  broods  are  derived  from  members  of 
the  Leguminosae,  i.e.,  mesquite,  catclaw,  mimosa  (Mimosa  sp.), 
deervetch  (Lofus  sp.),  paloverde  (Cercidium  sp.),  and  lupine 
(Lupinus  sp.) 

Gullion  1960;  Hungerford 
1962;  Goodwin  and 
Hungerford  1977 

California  quail 

(Callipepla 
calif  or  nica) 

May  occur  in  desert,  rangeland,  and  dry  and  irrigated  farmland  at 
elevations  from  sea  level  to  2,600  m  (8,500  ft).  Rangeland  with 
the  following  components  are  most  widely  used:  clumps  of  woody 
vegetation  for  roosting,  shrubs  or  herbaceous  growth  for  escape 
cover,  and  herbaceous  cover  for  nesting.  Food  habit  studies 
repeatedly  show  importance  of  legumes  (esp.  seeds).  Some  im- 
portant legumes  include  the  following:  bur  clover  (Medicago  sp), 
lupines,  deervetches,  clover  (Trifolium  sp),  and  vetches  (Vicia 
sp.).  Filaree  (Erodium  sp.)  leaves  and  seeds  occur  frequently  in 
the  diet. 

Emlen  and  Glading  1945; 
Edminster  1954;  Shields 
and  Duncan  1966;  Johns- 
gard 1973;  Gutierrez 
1980. 

Band-tailed  pigeon 
Coastal  race 

(Columba  fasciata 
monilis) 
Interior  race 
(C.  f.  fasciata) 

The  coastal  race,  although  primarily  an  inhabitant  of  mountain 
habitat,  can  be  found  from  sea  level  to  4,200  m  (13,850  ft).  The 
race  may  occupy  a  variety  of  habitats,  but  the  preferred  type 
consists  of  large  conifers  and  deciduous  trees  interspersed  with 
berry  and  mast  producing  trees  and  shrubs.  The  interior  race  also 
shows  preference  for  mountainous  areas  supporting  a  mixed 
conifer-deciduous  type  and  achieves  greatest  densities  in  this 
type  at  elevations  between  1 ,675  m  (5,500  ft)  and  2,575  m  (8,500 
ft).  Both  races  feed  heavily  on  mast  produced  by  species  of 
pine  and  oak  as  well  as  a  variety  of  berries  produced  by  such 
shrubs  as  elderberries  (Sambucus  sp.),  wild  cherries  (Prunus 
sp.),  huckleberries,  and  dogwood  (Cornus  sp).  During  certain 
years,  both  races  will  move  to  lower  elevations  and  feed  on  culti- 
vated grain. 

Braun  et  al.  1975;  Jeffrey 
1977;  Tomlinson  1983. 

Mourning  dove 

(Zenaida  macroura) 

The  wide  breeding  distribution  of  this  species  almost  precludes 
describing  habitat  features  with  precision.  It  is  primarily  an  inhab- 
itant of  woodland-grassland  edge  Thus,  both  clearing  of  large 
areas  of  forest  in  the  East  and  planting  of  shelterbelts  in  the 
Plains  States  enhanced  mourning  dove  habitat.  Although  tree 
nesting  is  most  common,  in  the  absence  of  trees  or  shrubs,  the 
nests  are  readily  placed  on  the  ground  Food  supplies  for  doves 
have  been  improved  by  species  that  produce  more  seeds  than 
native  grasses. 

Keeler  1977;  Dunks  et  al 
1982 

Upland  Game  Birds 


411 


Chukar  (Alectoris  chukar) 

The  chukar  has  been  widely  introduced  into  the 
U.S.,  but  was  successfully  established  only  in  arid 
areas  of  the  West.  Seven  species  of  this  genus  have 
been  recognized,  all  of  which  are  native  to  parts 
of  Europe  or  Asia.  It  is  believed  that  the  only  species 
that  have  been  successful  in  the  U.S.  were  those 
originating  in  India  (Christensen  1970). 

Chukars  characteristically  inhabit  very  rugged, 
rocky  terrain  in  arid  to  semiarid  climates.  They  have 
been  found  from  sea  level  to  4,800  m  (16,000  ft) 
in  their  native  Asian  habitat  and  to  3,600  m  ( 1 2,000 
ft)  in  North  America.  The  Great  Basin  habitat  of  the 
West  with  its  combination  of  mountain  range-valley 
(basin)  floor  provides  excellent  conditions  for  this 
exotic  species  (Christensen  1970). 

Chukars  frequently  use  the  rugged  topography 
for  cover,  although  they  are  also  found  in  a  shrub- 
grassland  type.  The  understory  may  consist  of  an 
abundant  grass-forb  combination  of  the  desert.  Sage- 
brush (Artemisia  tridentata)  is  the  dominant  shrub  in 
chukar  habitat  in  Nevada  (Christensen  1970)  and 
probably  over  most  of  its  range  in  the  U.S.  Other 
structurally  similar  shrubs  such  as  greasewood  (Sar- 
cobatus  sp.),  rabbitbrush  (Chrysothamnus  sp.),  and 
bitterbrush  (Purshia  tridentata)  may  be  found  in- 
stead of  sagebrush  in  different  parts  of  this  bird's  U.S. 
distribution.  A  food  source  is  largely  derived  from 
an  understory  of  grasses  and  forbs  although  grasses 
frequently  provide  the  major  portion.  Christensen 
(1970)  reported  heavy  use  of  seeds  and  leaves  of 
cheatgrass  (Bromus  tectorum)  and  believed  that  the 
introduction  of  this  exotic  grass  into  the  western 
U.S.  may  have  greatly  influenced  the  later  success  of 
the  chukar  introduction. 


within  and  adjacent  to  agricultural  lands,  the  preser- 
vation of  grass/forb/shrub  cover  is  frequently  the 
key  to  maintaining  a  healthy  pheasant  population. 

Although  the  pheasant  is  primarily  associated 
with  agricultural  lands,  a  few  populations  exist  in  the 
western  U.S.  completely  removed  from  cultivation. 
In  such  areas,  usually  a  drainage  supporting  some 
woody  cover,  the  birds  substitute  weed  seeds  and 
fruits  from  shrubs  such  as  buffaloberry  (Shepherdia 
sp.),  rose  (Rosa  sp.),  and  snowberry  (Symphoricar- 
pos  sp. )  for  the  cultivated  grains  that  usually  make 
up  a  large  part  of  the  diet.  In  such  areas,  pheasant 
densities  are  much  lower  than  in  comparable  cover 
adjacent  to  agriculture,  but  the  existence  of  even  the 
low  numbers  is  highly  dependent  upon  the  shrubby 
(protective)  cover. 


Ring-necked  pheasant. 


Ring-necked  Pheasant  (Phasianus 
colchicus) 

Ring-necked  pheasants  consist  of  a  group  of 
subspecies  from  Asia  that  were  successfully  intro- 
duced into  the  U.S.  in  the  1880s.  The  habitats  of  this 
group  of  birds  in  Asia  were  similar  to  those  found 
in  the  midwestern  and  western  U.S.,  and  the  birds 
became  established  in  many  of  these  areas  as  a  very 
popular  upland  game  bird. 

The  pheasant  is  closely  allied  with  agricultural 
lands.  In  the  West,  areas  placed  under  irrigation  ini- 
tially created  islands  of  excellent  pheasant  habitat, 
where  as  little  as  5%  of  the  area  in  permanent,  well 
interspersed  cover  provided  adequate  nesting,  brood 
rearing,  and  winter  cover  (Yeager  et  al.  1951).  As 
agriculture  became  more  intensive,  land  use  changes 
(largely  the  removal  of  permanent  cover)  reduced 
the  potential  for  producing  pheasants  (Baxter  and 
Wolfe  1972;  Snyder  1974;  Taylor  et  al.  1978).  Thus, 


Spruce  Grouse  (Dendragapus  canadansis) 

The  western  representative  of  this  small  timber 
grouse,  for  many  years  considered  a  separate  species, 
is  presently  regarded  as  a  race — the  Franklin's  spruce 
grouse.  Among  the  many  local  names  applied  to  this 
bird,  the  most  common  is  undoubtedly  the  fool  hen, 
a  direct  result  of  its  lack  of  fear  of  man. 

Spruce  grouse  are  distributed  across  the  whole 
of  Canada  and  portions  of  the  northern  U.S.,  closely 
coinciding  with  the  distribution  of  the  coniferous 
forest.  Robinson  (1980),  in  reviewing  literature  on 
habitat  for  this  bird,  suggested  a  trend  toward  in- 
creased use  of  upland  over  lowland  areas  in  its  distri- 
bution from  east  to  west.  The  most  dense  forests 
are  seemingly  avoided  because  they  do  not  provide 
an  adequate  ground  cover  and  openings  large 
enough  for  territorial  display  flights  (Stoneberg  1967; 
Robinson  1980). 


412 


Upland  Game  Birds 


Although  early  descriptions  of  spruce  grouse 
habitat  emphasize  the  mature  coniferous  forest  with 
species  of  spruce  being  the  primary  ingredient,  more 
recent  studies  have  shown  that  habitat  often  consists 
of  mixed  stands  of  spruces  (Picea  sp.)  and  jack  pine 
(Pinus  banksiana)  in  Michigan  (Robinson  1980)  and 
spruces  and  lodgepole  pine  (Pinus  contorta)  in  west- 
ern States  (Stoneberg  1967;  Ratti  et  al.  1984). 


Spruce  grouse. 


Blue  Grouse  (Dendragapus  obscurus) 

The  blue  grouse  is  the  largest  of  the  North 
American  forest  grouse,  the  males  commonly  weigh- 
ing 1,200-1,250  g  (2.6-2.7  lb)  or  about  twice  the 
weight  of  spruce  or  ruffed  grouse.  The  subspecies 
(D.  fuliquinosus)  inhabiting  the  coastal  mountain 
ranges  is  often  referred  to  as  the  sooty  grouse 
whereas  the  one  in  the  Rocky  Mountain  States  is 
often  called  the  dusky  grouse. 

Blue  grouse  often  participate  in  a  seasonal,  alti- 
tudinal  migration.  In  winter,  they  are  most  fre- 
quently found  at  high  elevations  in  dense  stands  of 
coniferous  forest.  In  spring,  a  downward  migration 
occurs  into  a  forest-grassland  ecotone  where  males 
establish  territories.  Females  may  brood  in  open 
meadows  at  this  elevation,  or  in  some  parts  of  their 
range,  will  follow  small  brushy  draws  out  into  prairie 
foothill  types.  In  Montana  and  Colorado,  sharp-tailed 
grouse  (  Tympanuchus  phasianellus),  sage  grouse, 
and  blue  grouse  have  been  observed  on  common 
brood-rearing  ranges  in  late  summer  and  early  fall 
(pers.  obs.). 

Although  an  inhabitant  of  forested,  mountain 
areas,  blue  grouse  appear  to  prefer  meadows  and 
grassland  adjacent  to  woody  cover  (Bendell  and 
Elliott  1966).  Breeding  habitat  is  frequently  an  open, 
dry  site  with  scattered  shrubs  but  near  dense  clumps 
of  trees  for  escape  cover  (Martinka  1972).  Brood 


rearing  often  occurs  in  meadows  or  an  open  stand  of 
timber  with  a  good  ground  cover  of  grass  and  forbs. 
Mussehl  (1963)  listed  the  following  basic  physical 
requirements  of  blue  grouse  brood  range:  a  relatively 
high  degree  of  canopy  coverage,  an  effective  height 
of  ground  cover  of  about  20  cm  (8  in.),  which  is 
an  important  consideration  in  areas  grazed  by  do- 
mestic livestock,  diversity  of  plant  life  forms,  and 
small  amounts  of  bare  ground.  Winter  habitat,  which 
usually  includes  a  more  dense  canopy  of  conifers 
for  a  food  source,  is  nonetheless  often  in  a  parkland 
type  or  along  open  ridges. 

In  overall  distribution,  blue  grouse  seem  closely 
associated  with  Douglas  fir  (Pseudotsuga  menziesii). 
Needles  from  this  species  are  an  important  winter 
food  source  although  others  (Larix  sp.,  Pinus  sp.,  and 
Abies  sp. )  have  been  reported  as  common  in  the 
winter  diet. 


Blue  grouse. 


In  breeding  and  brood  rearing  habitat,  life  form 
is  probably  as  critical  as  species  composition.  Mar- 
tinka ( 1972)  pointed  out  the  importance  of  Douglas 
fir  thickets  as  escape  cover  for  males  displaying  in 
open  logged-over  stands  of  ponderosa  pine  (Pinus 
pondcrosa).  Mussehl  (1963),  in  discussing  brood 
habitat,  described  the  bunchgrass-balsamroot  open- 
ings at  the  edge  of  coniferous  stands  in  one  study 
area  and  in  another,  a  similar  grass-forb  association 
in  an  open  stand  of  mature  ponderosa  pine.  In  years 
when  balsamroot  was  not  abundant,  broods  were 
found  in  association  with  snowberry  or  large  sedges 
(Carex  sp.).  In  late  summer,  before  moving  to  higher 
elevations  and  a  conifer  diet,  blue  grouse  in  the 
Rocky  Mountains  feed  heavily  on  berries  such  as 
currants  (Ribes  sp.),  juneberries  (Amelanchier  sp.), 
bearberry  (Arctostaphylos  sp.),  and  huckleberry 
(Vaccinium  sp.)  (Beer  1943). 


Upland  Game  Birds 


413 


Willow  Ptarmigan  (Lagopus  lagopus) 
Rock  Ptarmigan  (I.  mutus) 
White-tailed  Ptarmigan  (I.  leucurus) 

Although  considerable  overlap  exists  in  the 
North  American  geographical  distribution  of  these 
three  species,  particularly  so  with  willow  and  rock 
ptarmigan,  differences  in  seasonal  habitat  selection 
assist  in  maintaining  them  as  separate  species.  Be- 
cause willow  (Salix  sp.)  is  a  shrub  group  commonly 
used  by  all  three  species,  the  major  difference  in 
habitat  characteristics  is  structural.  From  willow  to 
rock  to  white-tailed  ptarmigan,  the  trend  is  toward 
shorter  and  less  dense  vegetation  and  increased  top- 
ographic relief.  Thus  during  the  spring  and  summer, 
the  willow  ptarmigan  shows  a  preference  for  level 
ground  or  gentle  to  moderate  slopes,  luxuriant  plant 
growth  with  shrub  height  usually  1-2.5  m  (3-8  ft), 
and  an  elevation  at  the  upper  edge  of  timberline 
among  widely  scattered  trees.  The  rock  ptarmigan 
shows  a  preference  for  moderate  slopes  in  hilly  boul- 
der-strewn country,  vegetative  cover  complete  but 
sparse  with  shrubs  (in  ravines)  0.3  to  1.4  m  (1-4  ft), 
and  most  are  found  at  an  elevation  of  30-300  m 
(100-1,000  ft)  above  timberline.  The  white-tailed 
ptarmigan  prefers  steep  slopes  with  rocky  outcrops, 
with  plant  cover  rarely  continuous,  shrubs  sparse 
and  in  a  dwarf  form,  and  are  usually  found  1 50-600 
m  (500-2,000  ft)  above  timberline  (Weeden  1965). 

All  three  species  rely  heavily  on  willow  for  food. 
Dominant  winter  foods  eaten  by  the  willow  ptarmi- 
gan are  willow  buds  and  twigs;  dwarf  birch  (Betula 
nana)  buds  and  catkins  are  second  in  importance. 
Rock  ptarmigan  eat  the  same  species  but  in  reverse 
order  of  prominence.  During  spring,  summer,  and 
fall,  both  species  feed  commonly  on  leaves  and  ber- 
ries of  cranberry  and  blueberry,  horsetail  (Equisetum 
sp.)  tips,  leaves  of  mountain  avens  (Dryas  octope- 
tala),  and  crowberries  (Empetrum  nigrum)  (Weeden 
1965).  White-tailed  ptarmigan  in  the  Rocky  Moun- 
tain States  (primarily  Colorado)  rely  heavily  on  al- 
pine willows  (Salix  nivalis  and  S.  arctica),  and  the 
distribution  and  abundance  of  these  plants  dictate 
the  distribution  of  this  ptarmigan  (Braun  1970). 

Ruffed  Grouse  (Bonasa  utnbellus) 

Ruffed  grouse,  the  most  widely  distributed 
grouse  in  North  America,  does  not  have  the  same 
popularity  as  a  game  bird  in  the  western  U.S.  that  it 
has  in  the  Lake  States  and  farther  east.  This  distribu- 
tion of  popularity  is  also  reflected  by  the  distribution 
of  ruffed  grouse  research  conducted  to  date. 

Ruffed  grouse  are  closely  associated  with  decid- 
uous forests  or  mixed  deciduous-coniferous  forest 
edges.  This  permits  a  more  continuous  distribution 
in  the  Lake  States  where  large  expanses  of  hardwood 
forest  are  present.  In  the  mountain  West,  ruffed 
grouse  are  more  prevalent  along  (but  not  entirely 


confined  to)  stream  courses  or  moist  areas  where 
deciduous  trees  and  shrubs  are  found.  During  the 
winter  in  mountain  habitat,  the  birds  prefer  a  south- 
ern exposure  and  elevations  above  the  stream  bot- 
tom (Stauffer  1983),  both  related  to  higher 
temperatures,  snow  melt,  and  resulting  food 
availability. 

The  plant  most  often  associated  with  thriving 
ruffed  grouse  populations  is  quaking  aspen  (Populus 
tremuloides).  Several  authors  (Svoboda  and  Gullion 
1972;  Doerr  et  al.  1974)  have  suggested  that  aspen 
buds  (especially  flower  buds  from  mature  male  as- 
pen) are  critical  winter  and  spring  foods.  Gullion 
(1977),  in  discussing  forest  management  for  ruffed 
grouse,  pointed  out  the  value  of  each  different  age 
aspen  stand  to  a  particular  seasonal  ruffed  grouse 
requirement.  In  the  West,  fire  suppression  in  many 
areas  has  probably  prevented  regeneration  of  com- 
plete stands  of  aspen,  and  early  serai  stages  are  few 
or  non-existent  (Stauffer  and  Peterson  1985b). 


Ruffed  grouse. 


In  many  areas  of  the  mountain  West,  aspen  is 
limited  in  distribution  and  seldom  in  the  large  con- 
tinuous stands  so  common  to  the  Lake  States.  Under 
these  conditions,  ruffed  grouse  rely  on  a  variety  of 
understory  shrubs  for  food  and  cover  such  as  willow, 
serviceberry  ( Amelanchier  sp.),  blueberry,  haw- 
thorne  (Crataegus  sp.),  chokecherry  (Prunus  sp.), 
and  snowberry. 

Sage  Grouse  (Centrocercus  urophasianus) 

The  sage  grouse  is  decidedly  the  largest  grouse 
in  North  America;  an  adult  male  may  weigh  3,175  g 
(7  lb ),  about  twice  the  size  of  the  next  largest,  the 
blue  grouse.  Unlike  many  other  North  American 
grouse,  sexual  dimorphism  is  pronounced  in  the  sage 
grouse  in  both  size  and  plumage. 


414 


Upland  Game  Birds 


Sage  grouse  are  characteristic  of  the  semi-arid 
sagebrush  or  sagebrush-grassland  types  found  in  the 
western  plains  and  intermountain  basins.  Generally 
following  the  distribution  of  the  plant  genus  Artemi- 
sia, sage  grouse  can  be  found  at  elevations  from 
about  750  m  (2,500  ft)  in  the  western  Dakotas  to 
over  2,100  m  (7,000  ft)  in  some  intermountain 
basins. 

Sagebrush  is  essential  to  sage  grouse  survival.  A 
proper  combination  of  height  and  canopy  of  this 
plant  must  be  available  to  meet  the  birds'  seasonal 
needs.  Dependency  on  the  plant  is  greatest  during 
the  late  fall,  winter,  and  early  spring.  A  sage  grouse 
food  habits  study  in  Montana  (Wallestad  et  al.  1975) 
showed  that  during  8  months  of  the  year. (October 
through  May ),  the  frequency  and  volume  of  big  sage 
leaves  found  in  crops  remained  above  60%  and  90% , 
respectively.  For  three  of  these  months  (December 
through  February)  it  was  the  only  food  present.  Thus 
for  a  large  part  of  the  year,  sage  grouse  are  found  in 
or  near  stands  of  sagebrush  with  a  canopy-coverage 
exceeding  20%  (Eng  and  Schladweiler  1972;  Walles- 
tad and  Schladweiler  1974).  In  addition  to  the  sage- 
brush canopy  coverage  requirement,  winter  ranges 
are  frequently  on  areas  with  little  or  no  slope  (Eng 
and  Schladweiler  1972;  Beck  1977),  although  Beck 
(1977)  reported  some  use  of  slopes  exceeding  10% 
and  also  use  of  windswept  ridges  for  feeding  sites. 

During  the  4  months  of  summer  and  early  fall, 
sage  grouse  seek  out  succulent  forbs,  and  frequently 
leave  the  dense  sage  and  move  to  scattered  sage 
distribution  with  a  1-10  and  10-25%  canopy  cover- 
age (Wallestad  1975).  Sage  grouse  at  this  time  are 
often  found  near  seeps,  streams,  or  irrigated  fields 
where  succulent  forbs  are  available.  By  mid  Septem- 
ber to  late  September,  frequently  coinciding  with 
regular  heavy  frosts,  the  birds  move  toward  the  win- 
tering areas.  This  movement  between  summer  and 


winter  ranges  may  be  as  little  as  3-5  km  (1.8-3  mi.) 
(Eng  and  Schladweiler  1972)  or  up  to  80  km 
(48  mi.)  (Dalke  et  al.  1963). 

Although  big  sage  is  the  species  most  frequently 
associated  with  sage  grouse,  several  others  (Artemi- 
sia cana,  A.  frigida,  A.  ludoviciana,  A.  nova)  have 
been  listed  as  used  for  both  food  and  cover  (Patter- 
son 1952;  Wallestad  et  al.  1975).  Species  of  forbs 
that  are  prominent  in  the  diet  of  sage  grouse  are 
dandelion,  prickly  lettuce  (Lactuca  sp.),  salsify  (Tra- 
gopogon  sp.),  gumweed  (Grindelia  sp.),  yarrow 
(Achillea  sp.),  and  sweet  clover  (Melilotus  sp.).  Al- 
falfa (Medicago  sp.)  is  frequently  heavily  used  in  late 
summer  and  early  fall. 

Prairie  Chicken  (Tytnpanuchus  cupido) 

This  species,  also  called  the  pinnated  grouse, 
was  once  well  represented  throughout  the  central 
U.S.  from  the  southern  prairie  provinces  of  Canada  to 
the  Gulf  Coast,  coinciding  closely  with  the  distribu- 
tion of  mid-  and  tallgrass  prairie.  The  1957  American 
Ornithological  Union  (A.O.U.)  checklist  recognized 
the  lesser  and  greater  prairie  chicken  as  separate 
species;  the  1983  version  lists  the  two  as  a  superspe- 
cies  although  many  consider  them  to  be  a  single 
species. 

The  lesser  prairie  chicken  is  characteristic  of  the 
more  arid  grasslands  of  the  Southwest,  whereas  the 
greater  is  associated  with  the  more  moist  tallgrass 
prairie  of  the  eastern  great  plains  (Johnsgard  1973). 
Both  races  appeared  to  have  benefited  from  the 
breaking  of  the  prairies  for  cultivation,  providing 
winter  food  in  the  form  of  grain,  and  permitted  a 
northward  range  extension  (Johnsgard  and  Wood 
1968).  This  extension  was  relatively  short-lived  for 
both  races  and  as  the  agricultural  effort  became 
more  intense,  distribution  receded  to  the  remnant 
status  of  today  (Johnsgard  1973). 


€§y^i^'i  '^%'%f>  ■^^vj%W"»*«**»<i 


m 


Sage  grouse  droppings. 


Sage  grouse  at  water  hole. 


Upland  Game  Birds 


415 


A  very  explicit  comparison  of  habitat  for  the 
lesser  and  greater  prairie  chickens  in  Oklahoma  was 
presented  by  Jones  (1963).  He  stated  "...  the  habi- 
tat of  the  lesser  prairie  chicken  consisted  of  small 
units  of  shortgrass  prairie  intermixed  with  larger 
units  of  shrub  or  half-shrub  vegetation;  that  of  the 
greater  prairie  chicken  consisted  of  small  units  of 
shortgrasses  or  midgrasses  intermixed  with  larger 
units  of  tall  grasses."  This  difference  was  also  shown 
in  day-resting  activities  with  the  lesser  using  primar- 
ily half-shrub  vegetation  whereas  the  greater  favored 
the  edges  of  tallgrass  and  midgrass  vegetation.  Both 
tended  to  rear  broods  in  areas  of  abundant  forb 
growth,  at  least  in  part  for  the  insect  fauna  in  associ- 
ation with  such  areas. 

The  emphasis  of  shrub  component  in  the  lesser 
prairie  chicken  habitat  is  more  pronounced  than 
with  the  greater  prairie  chicken.  The  half-shrub  (Ar- 
temisia filifolia)  is  common  in  Oklahoma  and  New 
Mexico  (Jones  1963).  In  New  Mexico,  the  shinnery 
oak  (  Quercus  havardii)  is  heavily  used;  the  habitat 
consists  of  a  shinnery  oak-tallgrass  community  which 
included  such  grasses  as  sand  bluestem  (Andropogon 
hallii),  little  bluestem  (Andropogon  scoparius), 
three-awn  (Aristida  sp.),  dropseed  (Sporobolus  sp. ), 
and  hairy  grama  (Bouteloua  hirsuta)  in  addition  to 
a  variety  of  forbs  (Davis  et  al.  1980).  However,  in 
Oklahoma,  densities  of  displaying  males  were  nega- 
tively correlated  with  percentage  coverage  of  brush 
and  positively  correlated  with  percentage  coverage 
of  grass  in  a  shinnery  oak  rangeland;  the  reverse  was 
found  on  a  sand  sagebrush  rangeland  (Cannon  and 
Knopf  1981a). 


Sharp-tailed  grouse  (Tympanuchus 
phasian  ellus) 

Six  races  of  sharp-tailed  grouse  are  recognized 
(Hamerstrom  and  Hamerstrom  1961),  two  of  which 
extend  their  ranges  south  of  Canada  in  the  western 
U.S.  The  Columbian  sharptail  (  T.  p.  columbianus) 
is  found  in  western  Montana  and  eastern  Washington 
south  to  Utah  and  western  Colorado.  The  plains  race 
(  T.  p.  jamesi)  is  the  most  abundant  and  widespread 
of  the  races  in  the  U.S.,  being  found  in  eastern  Mon- 
tana, Wyoming,  and  Colorado,  and  in  western  Ne- 
braska and  the  Dakotas. 

Sharp-tailed  grouse  prefer  habitat  with  structural 
features  intermediate  to  the  early  forest  stages  of 
ruffed  grouse  and  the  tallgrass  prairie  habitat  of  the 
greater  prairie  chicken.  Columbian  sharptails  cur- 
rently exist  in  small  isolated  populations,  with  the 
most  viable  segments  in  western  Colorado;  they 
associate  largely  with  Kuchler's  (1964)  sagebrush 
steppe  type  (Miller  and  Graul  1980).  The  plains  race 
is  the  most  prairie  dwelling  of  the  three,  and  is 
found  in  association  with  patches  of  shrubs  on  the 
prairie. 


Both  races  utilize  a  combination  of  trees-shrubs- 
grasslands  but  in  varying  degrees  of  composition  of 
these  structural  components.  The  Columbian  sharp- 
tail  favors  rolling  hills  or  benchland  with  a  bunch- 
grass,  forb,  and  shrub  combination.  In  valleys  of 
western  Colorado,  this  bird  overlaps  the  zones  of 
sagebrush  and  Gambel  oak  (  Quercus  gambellii) 
(Hoffman  and  Alexander  1980).  Frequently  agricul- 
tural fields  are  interspersed  on  the  ridgetops  or  val- 
ley floors  and  may  serve  as  feeding  sites  in  the  fall 
before  snow  cover. 

The  plains  sharp-tail  habitat  consists  of  short  to 
mid-grass  prairie,  interspersed  with  shrubby  draws 
and  agricultural  land.  Grazing  intensity  greatly  influ- 
ences the  distribution  of  grouse  in  this  habitat.  Niel- 
son  and  Yde  ( 1981 )  noted  a  uniform  use  of  pastures 
where  grass  height  averaged  22.8  cm  (91  in.)  com- 
pared with  those  with  grass  averaging  9-11  cm  (  3.6- 
4  in. )  where  use  was  more  confined  to  sections 
where  shrubs  provided  cover. 


Sharp-tailed  grouse. 


Columbia  sharptails  are  most  frequently  associ- 
ated with  aspen,  sagebrush,  hawthorne,  serviceberry, 
and  rose.  In  western  Colorado,  a  rather  close  winter 
association  is  frequently  seen  with  Gambel  oak 
which  may  provide  both  mast  and  buds  for  food. 
Plains  sharptails  are  linked  with  such  key  species  as 
aspen,  cottonwood  ( Populus  sp. ),  buffaloberry,  juni- 
per (Juniperus  sp. ),  snowberry,  and  rose.  Nielson 
and  Yde  (  1981 )  indicated  a  close  association  with 
buffaloberry  during  winters  of  heavy  snow.  Swenson 
( 1985)  quantified  a  shift  by  plains  sharptails  from 
feeding  in  grain  stubble  to  shrubs  as  snow  depth 
increased  beyond  140  mm  (5.5  in.). 

Wild  Turkey  (Meleagris  gallopavo) 

Two  subspecies  of  wild  turkey,  the  Merriam's 
(M.  g.  merriami)  and  the  Rio  Grande  (M.  g.  interme- 


416 


Upland  Game  Birds 


dia),  are  common  in  portions  of  the  western  U.S.  A 
third,  the  Mexican  or  Gould's  turkey  (M.  g.  mexi- 
cana),  has  limited  distribution  in  the  U.S.,  being  con- 
fined to  a  couple  of  small  areas  in  New  Mexico  and 
Arizona  (Lee  1959).  The  Merriam's  turkey  was  origi- 
nally associated  with  the  mountains  of  Colorado, 
New  Mexico,  and  Arizona.  However,  trapping  and 
transplanting  has  extended  the  range  to  most  States 
west  of  the  Mississippi  to  the  Canadian  border.  This 
range  expansion  is  not  continuous;  rather  the  distri- 
bution consists  of  discrete  populations,  many  iso- 
lated in  small  islands  of  mountain  habitat.  The  Rio 
Grande  turkey  is  found  largely  in  Texas  and  the 
northeast  corner  of  New  Mexico. 


The  Merriam's  turkey  in  its  original  range  is  a 
mountain-dwelling  bird  ranging  from  1,800  to  3,000 
m  (6,000  to  10,000  ft)  in  elevation  in  the  ponderosa 
pine-oak  forest  (Ligon  1946).  It  frequently  has  a 
distinct  summer  and  winter  range,  moving  to  lower 
elevations  in  the  winter  to  zones  producing  mast 
crops,  which  are  more  available  due  to  lighter  snow 
conditions.  Many  of  the  more  successful  transplant 
sites  north  of  its  original  distribution  are  character- 
ized by  relatively  low  rainfall  (and  snowfall),  ponde- 
rosa pine  and  grassland  areas,  and  foothill  grainfields 
adjacent  to  stringers  of  timber  or  shrubby  draws. 
Mast  producing  trees  and  shrubs  are  mostly  absent 
from  the  northern  areas  and  domestic  grain  is  fre- 
quently a  substitute. 

Ponderosa  pine  forests  with  open  grassland  un- 
derstory  characterize  Merriam's  turkey  spring,  sum- 
mer, and  fall  habitat.  In  a  study  in  southeastern 
Montana,  Jonas  (1966)  observed  6,271  turkeys  in 
468  groups,  recording  the  observations  by  plant 
communities.  Over  a  third  of  the  observations,  ex- 
cluding those  made  during  the  winter,  were  in  grass- 
lands, but  only  9%  were  more  than  100  m  (109 
yd)  from  tree  or  shrub  cover.  Twenty-seven  percent 
of  all  observations  were  in  snowberry,  but  mostly 
in  the  brush-grass  ecotone  and  not  in  pure  stands. 
Much  use  of  snowberry  was  for  cover  by  young 
poults — older  poults  used  the  forest  for  cover.  Over 
20%  of  all  observations  were  in  pine  stands  and  78% 
of  these  were  in  stands  of  pole  size  class.  Stands  of 
pole  size  pine,  in  contrast  to  saplings,  were  open 
enough  to  permit  growth  of  a  variety  of  low  growing 
food  plants  but  still  provided  overhead  protection 
from  avian  predators.  Only  in  winter  did  the  use  of 
pine  forests  decrease  to  less  than  20% .  During  this 
time,  73%  of  the  observations  were  in  deciduous 
tree-shrubby  draws  at  the  lower  elevations  where 
weather  was  less  severe.  Beasom  (1970)  reported  on 
Rio  Grande  turkeys  inhabiting  mesquite  (Prosopsis 
juliflora)  and  live  oak  (Quercus  virginiana)  vegeta- 
tive types  in  south  Texas.  Kothmann  and  Litton 
(1975)  reported  that  this  subspecies  extended  its 
range  into  14  million  ha  (35  million  a.)  of  shrub 
mesquite  prairie  in  west  Texas. 


Ponderosa  pine  is  undoubtedly  the  single  most 
important  plant  species  in  Merriam's  turkey  habitat. 
Sapling  doghair  stands  of  this  species  are  used  for 
cover  (Hoffman  1962),  the  older  more  open  stands 
are  used  for  cover  and  feeding  areas,  and  large  ma- 
ture trees  for  roost  trees  and  mast  producers  for 
winter  feed.  Other  mast  producers  in  the  more 
southern  ranges  are  oak,  primarily  Gambel  oak,  pin- 
yon  pine  (Pinus  edulis),  and  alligator  juniper  (Juni- 
perus  deppeana)  (Ligon  1946). 

Because  most  mast  producers  do  not  produce 
heavy  crops  annually,  other  food  sources  must  often 
be  relied  upon.  This  is  frequently  provided  by  the 
grass  family,  both  seed  heads  and  vegetative  parts, 
from  a  variety  of  species  (Ligon  1946;  Spicer  1959; 
Hoffman  1962;  Jonas  1966).  In  northern  extensions 
of  the  range,  mast  crops  are  even  less  dependable, 
and  grasses  play  a  greater  role  in  winter  foods. 
Often,  waste  domestic  grains  in  stubble  fields  pro- 
vide supplemental  winter  foods,  particularly  when 
grazing  by  domestic  livestock  is  excessive  and  native 
grass  seed  production  is  nil. 


Wild  turkeys. 


The  extension  of  the  Rio  Grande  turkey  beyond 
its  primary  range  into  the  relatively  treeless  area  of 
west  Texas  was  assisted  by  man-made  structures 
(primarily  power  lines  and  poles)  for  roosting,  and 
available  food  supplies  in  the  form  of  grasses,  grass 
seeds,  and  forbs  (Kothmann  and  Litton  1975).  The 
importance  of  roost  sites  to  Merriam's  turkeys  is 
illustrated  by  a  reported  64%  reduction  in  popula- 
tion following  a  pinyon-juniper  reduction  program 
which  isolated  roost  sites  over  300  m  (330  yd)  from 
cover  (Scott  and  Boeker  1977). 

Mearn's  Quail  (Cyrtonyx  montezuma 
mearnsi) 

The  largest  part  of  the  overall  distribution  of 
this  quail  is  in  Mexico,  although  one  race,  the 


Upland  Game  Birds 


417 


Mearn's  (  C.  m.  mearnsi),  is  found  in  westcentral 
Texas,  southwest  New  Mexico,  and  southeast  Ari- 
zona. The  male  has  a  striking  facial  plumage  of 
bluish-black  patches  mixed  with  white,  hence  the 
common  name  harlequin  quail. 

Throughout  its  range,  the  Mearn's  quail  dis- 
places the  Gambel's  (  Callipepla  gambellii)  and  scaled 
quails  (G  squamata)  of  the  desert  grasslands  at 
higher  elevations  (1,200-3,600  m  [4,000-12,000  ft]). 
In  Arizona,  Mearn's  quail  inhabit  the  oak-grasslands 
and  oak-woodlands  (Bishop  and  Hungerford  1965). 
In  New  Mexico,  Mearn's  quail  habitat  is  typified 
more  by  pinyon  pine  and  junipers  as  well  as  ponde- 
rosa  pine  (Ligon  1961). 

Although  this  species  of  quail  was  considered  to 
be  rare  in  the  U.S.,  Brown  (1982)  noted  that  major 
populations  are  present  in  mountain  ranges  of  south- 
eastern Arizona  in  oak  woodland  and  grassland  type, 
and  is  common  enough  to  be  an  important  game 
bird  in  Arizona  (Brown  1979).  In  the  Mexican  high- 
lands, Leopold  and  McCabe  (1957)  believed  that  this 
quail  was  so  closely  associated  with  the  pine-oak 
forest  that  it  could  well  be  considered  an  avian  indi- 
cator of  the  type. 

The  overstory  of  pine  and  oak  is  not  the  impor- 
tant habitat  component  of  this  species.  Rather  it 
signifies  a  climatic  zone  harboring  a  complex  of  per- 
ennial forbs  and  sedges  that  provide  the  food  and 
water  source  in  the  form  of  bulbs  (Leopold  and 
McCabe  1957).  Although  heavy  grazing  by  livestock 
was  believed  responsible  for  a  severe  reduction  in 
quail  food,  more  recent  studies  have  shown  that 
forage  removal  >  55%  will  eliminate  a  quail  popula- 
tion, not  because  of  a  short  food  supply  but  because 
of  cover  removal  and  subsequent  vulnerability  of 
the  birds  (Brown  1982). 

Most  food  habit  studies  show  that  acorns  are 
fairly  high  in  both  frequency  and  volume,  particu- 
larly in  years  when  some  of  the  other  foods  may  be 
scarce  (Bishop  and  Hungerford  1965).  Thus  the 
various  species  of  scrubby  oaks  so  characteristic  of 
the  U.S.  range  of  this  quail  probably  play  a  dual  role 
as  food  and  cover.  Woodsorrel  (  Oxalis  sp. )  and 
sedges  are  both  heavily  used  for  food  (bulbs  and 
seeds)  and  only  during  July  through  September  did 
these  two  genera  make  up  less  than  50%  of  the  vol- 
ume of  food  in  a  study  in  Arizona  (Bishop  and 
Hungerford  1965). 


Masked  Bobwhite  (Colinus  virginianus 
ridgwayf) 

The  masked  bobwhite  quail  formerly  occupied 
mesquite-grassland  areas  of  southern  Arizona  and 
northern  Sonora,  Mexico.  They  disappeared  from  Ari- 
zona by  1900,  presumably  because  of  cover  removal 


resulting  from  a  series  of  droughts  coupled  with 
excessive  grazing  (Tomlinson  1972b). 

Descriptions  of  masked  bobwhite  habitat  have 
been  obtained  from  studies  conducted  in  Mexico 
(Tomlinson  1972a,  b)  or  on  transplanted  birds  in  Ar- 
izona (Goodwin  and  Hungerford  1977).  Masked 
bobwhite  show  preference  for  flat,  mesquite-grass- 
land at  elevations  from  300  to  1,200  m  (1,000  to 
4,000  ft).  This  is  in  contrast  to  the  other  quails  of 
the  area  that  exhibit  a  preference  for  a  more 
wooded  and  broken  terrain  (Tomlinson  1972b).  Pen- 
reared  bobwhites  released  in  Arizona  also  showed  a 
preference  for  open,  grass-forb  sites  but  adjacent 
to  mesquite  lined  drainages  (Goodwin  and 
Hungerford  1977). 

Masked  bobwhites  prefer  heavy  forb  growth  in 
moderate  to  dense  stands  of  grasses.  Grasses  com- 
mon to  their  Mexican  habitat  are  several  species  of 
grama  (Bouteloua  sp.),  three-awn,  bristlegrass  (Se- 
taria  sp.),  and  panic  grasses  (Panicum  sp.;  Tomlinson 
1972a).  Quail  released  on  Arizona  sites  were  later 
found  in  areas  with  a  dense  understory  of  pigweed 
(Amaranthus  sp. )  and  an  overstory  of  mesquite,  pa- 
loverde  (Cercidium  sp.),  and  wolfberry  (Lycium  sp.). 
Like  the  birds  observed  in  Sonora,  a  preference  was 
shown  for  a  dense  grass-forb  complex  with  adjacent 
brush  or  trees  (Goodwin  and  Hungerford  1977).  The 
woody  cover  was  used  for  cover  by  the  quail  during 
the  winter  (Tomlinson  1972a). 


Scaled  Quail  {Callipepla  squamata) 

The  scaled  or  blue  quail,  as  it  is  often  called,  has 
a  geographic  distribution  which  conforms  closely  to 
the  Chihuahuan  desert  and  adjacent  grasslands 
(Johnsgard  1973).  In  New  Mexico,  it  is  found  over 
most  nonforested  sections  up  to  2,100  m  (6,990  ft) 
(Ligon  1961).  Hoffman  (1965)  described  the  scaled 


Scaled  quail. 


418 


Upland  Game  Birds 


quail  range  in  Colorado  as  extremely  variable  from 
the  flat  farmlands  at  1,020  m  (3,400  ft)  to  the  rocky 
slopes  and  canyons  adjacent  to  the  Rocky  Mountains 
at  2,100  m  (7,000  ft). 

In  Arizona  good  scaled  quail  habitat  consists  of 
low-growing  grasses,  forbs,  and  shrubs  with  an  over- 
all ground  cover  between  10  and  50%  (Goodwin 
and  Hungerford  1977).  In  contrast  to  Gambel's  quail 
(which  seemed  to  prefer  taller  shrubs — 60  to  80% 
over  2  m  [6.6  ft]),  scaled  quail  were  found  to  fre- 
quent areas  where  less  than  2%  of  the  shrubs  ex- 
ceeded 2  m  (6.6  ft)  in  height.  In  the  absence  of 
shrubby  cover  in  western  Oklahoma,  they  readily 
used  man-made  structures  (Schemnitz  1961).  In 
Colorado,  scaled  quail  are  most  commonly  found  in 
sand  sagebrush,  dense  cholla  cactus  (Opuntia  sp.),  or 
yucca  (soapweed  [Yucca  glauca])-grassland,  and  pin- 
yon- juniper  woodlands  (Hoffman  1965). 

Species  of  plants  used  for  cover  by  scaled  quail 
seemed  secondary  to  structure.  Schemnitz  (1961) 
identified  habitat  use  in  Oklahoma  based  on  over 
2,000  observations  in  three  vegetation  types  and 
found  29%  in  man-made  cover  (such  as  buildings 
and  machinery)  and  54%  in  a  shrub  life-form,  pri- 
marily skunkbush  (Rhus  aromatica),  small  soapweed  - 
sandsage,  and  small  soapweed.  In  New  Mexico, 
prominent  shrubs  in  scaled  quail  habitat  were  mes- 
quite,  catclaw  (Acacia  greggii),  whitehorn  (Acacia 
constricta),  and  snakeweed  (Guticrrezia  sp.) 
(Campbell  et  ai.  1973). 

Based  on  the  analysis  of  221  crops  collected  in 
the  fall  and  winter  in  New  Mexico,  seven  foods 
made  up  over  75%  of  the  total  volume:  seeds,  leaves, 
and  stems  of  whitehorn,  snakeweed,  doveweed  ( Cro- 
ton  sp.),  whitemargin  euphorbia  (Euphorbia  albo- 
marginata),  insects,  mesquite,  and  amaranth  (Ama- 
ranthus  sp.)  (Campbell  et  al.  1973). 

Gambel's  Quail  (Callipepla  gatnbeUii) 

Gullion  (I960)  reported  the  Gambel's  quail  as 
the  most  promising  upland  game  bird  in  parts  of  the 
Southwest  because  in  many  desert  areas  it  is  the 
only  resident  upland  game  species  in  abundance.  Al- 
though records  exist  of  scattered  populations  of 
Gambel's  quail  occurring  at  higher  elevations,  maxi- 
mum elevations  of  normal  occurrence  are  more  in 
the  1,220-  to  1,830-m  (4,000-  to  6,000-ft)  range. 
Only  marginal  populations  exist  where  annual  snow- 
fall is  greater  than  50  cm  (20  in.)  or  where  2.5  cm 
(1  in. )  or  more  snow  is  on  the  ground  for  more  than 
40  days  each  year  (Gullion  I960). 

A  preference  for  dense  stands  of  desert  shrub 
was  clearly  evident  from  Goodwin  and  Hungerford's 
(1977)  Arizona  study.  They  found  this  species  of 
quail  favoring  shrub  cover  of  sufficient  density  to 
shade  from  50  to  75%  of  the  ground  and  where  60 


to  80%  of  the  shrubs  were  taller  than  2  m  (6.5  ft). 
They  suggested  that  a  dense  understory  was  unim- 
portant to  Gambel's  quail. 

Considerable  overlap  exists  in  the  distribution 
of  scaled  and  Gambel's  quail  and  although  they  may 
be  found  in  the  same  habitats,  their  habitat  require- 
ments are  different.  Generally,  Gambel's  quail  are 
found  in  the  river  bottoms  and  are  more  associated 
with  dry  washes  or  brushy  draws;  scaled  quail  will 
more  likely  occur  on  the  surrounding  mesas  and 
plains  in  association  with  grassy  open  lands  (R.L. 
Tomlinson  pers.  commun.). 

Mature  mesquite,  hackberry  (  Celtis  reticulata), 
wolfberry,  and  catclaw  provided  cover  for  feeding, 
loafing,  roosting,  nesting,  and  raising  broods  in  Ari- 
zona (Goodwin  and  Hungerford  1977).  Gullion 
(I960)  listed  many  of  the  same  plants  plus  other 
shrubs  providing  a  comparable  structure.  In  addition, 
he  listed  a  variety  of  noxious  weeds  such  as  redroot 
amaranth  (Amaranthus  rctroflexux),  goosefoot 
(Chenopodium  sp.),  sunflower  (Helianthus  annuus), 
and  white  sweetclover  (Melilotus  alba)  as  low  grow- 
ing forms  providing  a  good  food  source. 

Hungerford  (1962),  in  studying  food  habits  of 
this  quail  in  Arizona,  reported  members  of  the  Leg- 
uminosae  as  very  important  food  producers.  Six  of 
nine  genera  listed  as  important  food  items  were  of 
the  legume  family  including  mesquite,  catclaw,  mi- 
mosa (Mimosa  sp.),  deervetch  (Lotus  sp.),  paloverde, 
and  lupine. 


Gambel's  quail. 


California  Quail  (Callipepla  californica) 

This  close  relative  of  the  Gambel's  quail  is  dis- 
tributed along  the  West  Coast.  It  has  also  been  suc- 
cessfully introduced  into  the  interior  in  Washington, 


Upland  Game  Birds 


419 


Oregon,  Idaho,  Nevada,  and  Utah.  Although  hybridi- 
zation with  Gambel's  quail  occurs  where  the  two 
species  coexist,  ecological  differences  prevent  exten- 
sive overlap  in  distribution  (Johnsgard  1973). 

With  the  extensive  north-south  distribution  of 
this  quail,  at  least  eight  races  are  found  over  a  rather 
wide  range  of  climatic  and  moisture  conditions.  In 
California,  this  quail  can  be  found  from  near  sea 
level  to  2,550  m  (8,500  ft).  Although  several  races 
occur  in  California  and  a  wide  variety  of  habitat  con- 
ditions are  utilized,  brushy  vegetation  (stiff- twigged, 
dense-foliaged  trees  for  night  roosting)  adjacent  to 
open  grassy,  weedy  types  and  available  water  charac- 
terize this  species'  requirements  throughout  the 
state  (Grinnel  and  Miller  1944).  In  comparing  habi- 
tats of  the  mountain  quail  (  Oreortyx  pictus)  and 
California  quails,  Gutierrez  (1980)  listed  the  latter  as 
preferring  the  open  woodland,  grassland,  and  chapar- 
ral habitats  on  less  steep  hillsides. 

California  quail  habitat  reported  by  Emlen  and 
Glading  (1945)  includes  the  following  general  types: 
desert,  rangeland,  dry  farming  land,  and  irrigated 
land.  Rangeland  is  the  most  extensive  and  most 
widely  used  of  the  four  types.  Each  type  provides  for 
the  basic  requirements  of  the  quail  in  various  de- 
grees. Water,  which  is  apparently  a  more  critical 
factor  than  with  Gambel's  quail,  is  usually  at  a  pre- 
mium on  dry-land  farms  and  portions  of  the  desert 
habitat.  Cover  (roosting,  loafing,  and  nesting)  is  often 
lacking  in  both  irrigated  and  dry-land  farm  areas. 
Deserts  and  rangelands  (particularly  the  latter )  will 
most  often  provide  the  variety  of  conditions  needed 
for  food,  water,  and  various  cover  components  to 
house  the  more  dense  populations  (Emlen  and 
Glading  1945). 

Like  many  of  the  quail  that  feed  on  a  wide  vari- 
ety of  foods  (including  seeds  and  leafy  parts  of 
plants),  species  composition  of  cover  is  probably  far 
less  important  than  structural  or  life -form  character- 
istics. This  is  particularly  true  of  the  California  quail 
in  light  of  its  very  wide  distribution  through  diverse 
vegetational  zones.  A  variety  of  food  habit  studies 
repeatedly  show  the  importance  of  legumes,  espe- 
cially the  seeds  (Edminster  1954;  Shields  and  Dun- 
can 1966;  Gutierrez  1980).  Important  legumes 
appearing  in  food  include  bur  clover  (Medicago 
hispida)  lupines,  deervetches  (Lotus  sp.),  clover  (Tri- 
folium  sp. ),  and  vetches  (  Vicia  sp. ).  Another  herb 
that  appears  frequently  as  food,  both  as  leafage  and 
seeds,  is  filaree  (  Erodium  sp. ). 

Band- tailed  Pigeon  (Columba  fasciata) 

The  band-tailed  pigeon  is  a  large  migratory  dove 
similar  in  size  and  color  to  the  non-migratory  blue- 
phase  domestic  pigeon  (Columba  livia).  Two  races 
of  this  bird,  the  interior  race  (C.  f.  fasciata)  and  the 
coastal  race  (C.  f.  monilis),  have  been  identified 


north  of  Mexico.  The  interior  race  migrates  into 
Mexico  for  6  months  during  the  winter;  the  coastal 
race  winters  primarily  in  southern  California  (Braun 
et  al.  1975;  Tomlinson  1983). 


Band- tailed  pigeon. 


Both  races  of  band-tailed  pigeons  are  closely 
associated  with  forests  that  provide  their  major  food 
source  in  the  form  of  mast,  berries,  and  small  fruits. 
In  some  areas,  when  natural  forest  foods  are  scarce, 
usually  during  migration,  grain  fields  and  orchard 
crops  are  invaded  to  a  point  of  human  conflict.  The 
coastal  race  occupies  a  diverse  array  of  mountain 
forests  from  moist  coast  forests  in  Washington  and 
Oregon  to  drier  Sierran  montane  forests  further 
south  and  east  (Jeffrey  1977).  The  interior  race  oc- 
cupies a  wide  range  of  habitats  ranging  from  agricul- 
tural types  near  forests  to  berry-producing  forest 
sites  at  elevations  to  3,300  m  ( 10,900  ft).  Habitat  is 
almost  always  mountain  related  although  occasion- 
ally feeding  birds  will  be  observed  in  grassland  or 
desert  shrub. 

Habitat  structure  for  the  coastal  race  in  its 
northern  range  consists  of  forested  land,  well  inter- 
spersed with  forest  openings  in  early  successional 
stages  favoring  berry-producing  shrubs.  California 
habitat  includes  mountain  forests,  woodlands,  and 
chaparral  if  accompanied  by  abundant  oak.  Ranges  of 
the  interior  race  are  primarily  diverse  mountain  con- 
iferous forests  but  most  frequently  accompanied  by 
a  common  denominator — a  pine-oak  combination 
(Jeffrey  1977).  In  southern  Arizona  and  New  Mex- 
ico, breeding  habitat  may  be  found  in  oak  communi- 
ties (Braun  et  al.  1975). 

In  the  northern  habitat  of  the  coastal  race,  a 
variety  of  overstory  trees  such  as  western  hemlock 
(Tsuga  heterophylla),  red  cedar  (Thuja  plicata),  and 
Douglas  fir  are  dominants,  whereas  in  California  the 
redwood  (Sequoia  sempervirens)  predominates  in 
the  coastal  forest.  However,  because  food  supplies 


420 


Upland  Game  Birds 


appear  to  control  the  distribution  and  abundance  of 
bandtails  more  than  any  other  single  factor,  species 
of  understory  trees  and  shrubs  are  probably  most 
important.  Various  species  of  oak  are  key  deciduous 
trees.  Several  species  of  shrubs  that  are  good  food 
producers,  and  most  vigorous  and  productive  in 
serai  stages  following  logging  or  fire,  are  elderberries 
(Sambucus  sp.),  wild  cherries,  huckleberries,  and 
dogwood  (Cornus  sp. )  (Jeffrey  1977). 

The  primary  winter  range  of  the  coastal  race  is 
in  California.  During  winter  they  primarily  inhabit 
the  pine-oak  woodland  and  coastal  chaparral  plant 
associations,  although  wintering  populations  are  mo- 
bile and  may  shift  areas  of  concentration  from  year 
to  year  depending  on  available  food  sources 
(Tomlinson  1983). 

Although  the  interior  race  can  be  found  at  ele- 
vations supporting  Engelmann  spruce  (Picea  engel- 
manni)  and  alpine  fir  (Abies  lasiocarpa),  it  is  most 
abundant  at  the  lower  ponderosa  pine/Gambel  oak 
zone.  The  close  association  with  various  species 
of  oak  and  pine  is  related  to  the  staple  fall  and  win- 
ter foods  of  acorns  and  pine  nuts. 

Mourning  Dove  (Zenaida  macroura) 

Two  races  of  mourning  doves  breed  in  the  U.S., 
Z.  m.  carolinensis  in  the  eastern  one-third  and  Z. 
m.  marginella  in  the  western  two-thirds.  These  doves 
breed  in  all  of  the  lower  48  States,  extending  into 
the  Canadian  provinces  to  the  north  and  Mexico  to 
the  south.  Most  mourning  doves  migrate,  spending 
the  winter  in  the  southern  U.S.,  Mexico,  Central 
America,  or  the  West  Indies  (Keeler  1977). 

The  mourning  dove  has  probably  benefited  from 
or  has  been  able  to  adapt  to  man's  activities  more 
than  most  other  native  bird  species.  It  is  primarily  a 


Mourning  dove. 


bird  of  open  woodland  or  the  edge  between  forest 
and  prairie.  Thus,  the  clearing  of  large  areas  of  decid- 
uous forest  in  the  East  and  the  planting  of  trees  on 
the  prairie  (i.e.,  shelterbelts  for  farms  and  fields,  ur- 
ban tree  planting)  has  enhanced  the  dove  population 
over  large  areas.  The  dove  is  basically  a  tree  nester 
but  will  commonly  nest  on  the  ground  in  the  ab- 
sence of  trees  or  shrubs.  Early  in  the  nesting  season 
before  deciduous  trees  leaf  out,  doves  prefer  conifers 
for  nest  sites.  Thus  the  inclusion  of  conifers  in  shel- 
terbelts and  urban  plantings  has  also  favored  this 
bird.  The  conversion  of  large  tracts  of  treeless  prairie 
to  domestic  grains  and  farmsteads  (trees)  has  cre- 
ated an  excellent  combination  of  food  (waste  grain) 
and  nesting  cover  for  doves.  Intensive  grazing  on 
many  ranges  has  encouraged  invader  plant  species 
that  often  produce  more  seeds  than  native  grasses. 
Also  intensive  grazing  management  includes  stock- 
dams  that  provide  water  in  areas  where  it  is  at  a 
premium,  and  often  trees  and  shrubs  for  nesting  and 
loafing  (Dunks  et  al.  1982). 

Population  Measurement  Techniques 

Most  techniques  for  inventorying  upland  game 
birds  provide  direct  or  indirect  data  on  the  presence 
or  absence  of  the  target  species  within  the  area 
being  sampled.  Such  techniques  may  consist  of  enu- 
merations of  actual  sightings  or  audible  responses  by 
the  bird  to  a  stimulus  (direct)  or  the  observation  of 
droppings,  feathers,  or  tracks  (indirect),  either  of 
which  can  provide  the  manager  with  at  least  a  basis 
for  initial  habitat  management  decision. 

Many  techniques,  primarily  those  based  on  a 
seasonal  activity  of  the  bird  (usually  breeding  activ- 
ity), show  sufficient  precision  to  provide  relative 
abundance  data.  Thus  comparisons  can  be  made  be- 
tween areas  or  between  years  on  a  single  area  per- 
mitting an  evaluation  of  the  bird's  response  to 
different  or  changing  habitat  conditions. 

Few  if  any  techniques  measure  the  actual  num- 
ber of  birds  present.  Even  on  limited  areas  where 
a  large  percentage  of  the  male  population  may  be 
engaged  in  a  common  activity  (breeding  display),  an 
unknown  percentage  of  the  males  are  not  displaying 
(Dorney  et  al.  1958;  Rippin  and  Boag  1974). 

Although  a  few  upland  game  census  techniques 
involve  complete  coverage  of  a  unit,  most  are  based 
on  variations  of  transect  sampling.  Likewise,  a  basic 
walking  transect  census,  such  as  the  King  Census  for 
ruffed  grouse  originally  described  by  Leopold 
(1933),  has  been  modified  for  a  variety  of  species 
and  conditions  as  pointed  out  by  Hayne  (1949)  in 
his  examination  of  the  technique.  Thus,  many  meth- 
ods for  determining  indexes  to  upland  game  bird 
numbers  are  derived  from  a  basic  technique  but 
modified  slightly  to  accommodate  a  particular  spe- 
cies, habitat  type,  or  season.  A  very  comprehensive 


Upland  Game  Birds 


421 


and  exhaustive  coverage  of  transect  sampling  was 
presented  by  Eberhardt  (  1978). 

All  of  the  techniques  listed  below  will  at  least 
indicate  presence  or  absence.  The  degree  of  preci- 
sion and  accuracy  beyond  that  will  be  indicated  with 
each  method. 

Dropping  Counts. 

Description.  Although  this  technique  (pellet- 
group  counts)  has  been  used  extensively  with  large 
ungulates,  it  has  been  little  used  with  birds.  Pyrah 
(1972)  used  this  method  successfully  in  determining 
relative  use  by  sage  grouse  on  several  study  areas 
subjected  to  various  degrees  of  herbicide  treatment. 
In  areas  previously  gridded,  points  were  randomly 
selected  on  which  to  establish  30-m  (100-ft) 
transects.  At  7.5-m  (25-ft)  intervals  along  the 
transect,  9-m    (100-ft  )  circular  plots  (radius  [1.7 
m]5.64  ft)  were  established  and  searched  for 
droppings.  Distinction  was  made  between  single 
droppings  and  clusters  resulting  from  a  night  roost. 

Accuracy  and  Precision.  No  attempt  was 
made  to  correlate  the  data  with  the  actual  number  of 
birds  using  the  area.  However,  based  on  field  obser- 
vations during  this  rather  intensive  study,  the  data 
reflected  relative  use  by  sage  grouse  among  the  dif- 
ferent areas  during  the  previous  winter  (D.  Pyrah, 
pers.  commun.). 

Discussion.  This  indirect  census  method  has 
some  distinct  advantages  over  many  for  game  birds. 
Sampling  can  be  conducted  throughout  the  day  in 
contrast  to  short  periods  of  bird  activity.  By  clearing 
permanent  plots  after  each  count,  seasonal  use  can 
be  measured  in  a  relatively  short  period  of  time 
compared  with  measuring  directly  through  actual 
bird  observation.  Gates  (1983)  used  the  same  tran- 
sects and  in  part  the  sample  plots  to  determine  rela- 
tive use  by  sage  grouse,  pygmy  rabbits  (Sylvilagus 
idahoensis),  black-tailed  jack  rabbits  (Lepus  californi- 
cus),  and  pronghorn  (Antilocapra  americana). 

The  method  also  has  some  disadvantages:  its  use 
is  largely  restricted  to  habitats  with  moderate  to 
light  ground  cover  and  to  the  period  after  snow  melt 
and  before  excessive  green-up;  it  is  probably  re- 
stricted to  species  (several  of  the  grouse)  on  a  diet 
including  adequate  roughage  to  prevent  rapid  deteri- 
oration of  the  droppings;  and  until  more  information 
on  defecation  and  dropping  deterioration  rates  is 
available,  the  method  will  be  restricted  to  a  relative 
use  index. 

King  Strip  Census. 

Description.  The  King  Strip  census  has  been 
modified  and  discussed  many  times  since  first 


employed  on  ruffed  grouse  by  King  ( 1973).  In  its 
simplest  form,  a  series  of  parallel  transects  of  known 
length  are  walked,  and  birds  observed  are  recorded 
including  the  flushing  distance.  The  average  flushing 
distance  is  calculated,  which  when  multiplied  by 
the  total  length  of  the  transects  provides  the  area 
covered.  The  total  area  divided  by  the  number  of 
birds  observed  provides  the  density  index.  A  more 
detailed  description  of  this  method  is  presented  by 
Overton  (1971,  pp.  420-424). 

Accuracy  and  Precision.  Under  most  condi- 
tions this  method  would  denote  presence  and  with 
adequate  samples  may  indicate  distinct  trends  in  the 
population.  One  of  the  longest  continuous  uses  of 
this  method  on  a  single  area  (Marshall  1954)  sug- 
gested similar  magnitudes  in  population  when  com- 
pared to  grouse  numbers  with  a  second  index  from 
the  same  area. 

Discussion.  This  method  has  some  merit  in 
areas  or  seasons  where  the  use  of  the  other  methods 
are  not  feasible;  it  can  indicate  presence  and  to  a 
degree,  relative  abundance  of  the  target  species. 
Because  so  many  variables  are  potentially  present  in 
using  such  a  method,  it  is  imperative  that  conditions 
be  standardized  as  much  as  possible,  i.e.,  time  of  year 
(phenologically),  time  of  day,  and  comparable 
weather  conditions.  One  advantage  of  this  method  is 
that  it  forces  the  biologist  onto  the  ground  to  sys- 
tematically walk  through  the  habitat. 

This  method  does  not  provide  usable  indexes 
when  used  on  low  density  populations — the  amount 
of  sampling  necessary  would  be  prohibitive. 

Roadside  Counts. 


Description.  As  the  name  implies,  this  method 
consists  of  driving  along  a  predetermined  route 
under  a  prescribed  set  of  conditions  and  counting 
the  number  of  birds  observed  (Kosicky  et  al.  1952). 
One  should  attempt  to  standardize  driving  speed, 
number  of  observers,  season,  time  of  day,  and 
weather  conditions  (wind  speed,  temperature,  and 
cloud  cover).  Thus  many  of  the  variables 
encountered  in  walking  strip  counts  (King  Strip 
census)  would  be  the  same  as  in  roadside  counts; 
the  latter  permits  more  extensive  coverage  in  less 
time. 

Accuracy  and  Precision.  Basically  the  same  as 
for  the  preceding  method. 

Discussion.  Roadside  counts  have  been  used 
extensively  in  the  Midwest  to  obtain  indexes  on 
pheasants,  quail,  doves,  and  cottontails.  In  the  West 
where  game  bird  densities  are  frequently  lower  than 
in  the  Midwest,  the  roadside  count  is  superior  to  a 
strip  count  conducted  on  foot.  In  addition  to  popula- 
tion indexes,  roadside  counts  have  also  been  used 


422 


Upland  Game  Birds 


to  obtain  cover  or  habitat  preferences  by  recording 
habitat  use  in  relation  to  availability  along  the  route 
(Hoffman  1965). 

Like  the  strip  census,  the  greatest  drawback  to 
this  method  is  the  number  of  variables  that  can  influ- 
ence the  results.  The  availability  of  cottontails  for 
enumeration  can  vary  considerably  from  day  to  day 
as  discussed  by  Newman  (1959).  The  basic  problems 
would  be  similar  for  game  birds. 

Complete  Census. 

Description.  The  complete  census  attempts  to 
tally  all  birds  in  a  unit  of  habitat  of  known  size.  The 
procedure  can  vary  from  one  individual  carefully 
cruising  back  and  forth  to  one  or  two  persons  with 
well  trained  dogs,  to  many  individuals  essentially  in  a 
drive.  In  all  instances  care  must  be  exercised  to 
prevent  duplication  in  tally.  The  distance  between 
lines,  whether  covered  singly  or  by  many,  is  largely 
dictated  by  the  type  of  cover. 

Accuracy  and  Precision.  This  technique, 
when  properly  executed  in  cover  that  is  not  too 
dense,  has  been  considered  to  be  more  accurate  and 
reliable  than  many  others. 

Discussion.  For  most  species  of  game  birds, 
the  term  complete  count  would  be  a  misnomer. 
However,  on  tracts  without  heavy  cover,  with  spe- 
cies more  prone  to  flush  than  run,  and  with  adequate 
personnel  to  keep  intervals  between  participants  at 
a  minimum,  a  drive  count  can  be  effective  in  making 
a  large  percentage  of  the  birds  available  for  tallying. 
The  results  are  influenced  by  many  of  the  same  vari- 
ables mentioned  for  strip  counts,  and  maximum 
standardization  of  the  variables  is  very  critical  to 
comparing  the  results  between  areas  or  years. 

This  method  is  more  manpower-expensive  than 
strip  sampling  or  roadside  counts  and  does  not  sam- 
ple as  large  an  area  for  the  effort.  One  modification 
that  has  been  employed  on  a  number  of  game  birds, 
particularly  covey  species,  is  to  incorporate  well- 
trained  pointing  dogs  into  the  effort.  This  can  be 
very  effective  but  all  too  often  simply  adds  another 
variable.  This  method  is  obviously  impractical  for 
use  over  wide  areas  (county  or  state);  its  application 
is  best  confined  to  areas  of  intensive  study. 

Auditory  Census. 

Description.  As  used  here,  this  method 
includes  all  sounds  produced  by  a  bird.  Most  of  the 
sounds  are  vocalizations  although  some,  such  as 
those  produced  by  a  drumming  ruffed  grouse,  are 
mechanically  produced.  The  sounds  used  in 
obtaining  population  indexes  are  associated  with  the 
breeding  season. 


The  most  basic  auditory  method  is  to  travel 
over  a  predetermined  route,  stopping  and  listening 
at  established  intervals  (listening  stops)  for  a  given 
length  of  time,  during  which  period  the  observer 
records  the  number  of  sounds  produced  or  the  num- 
ber of  birds  that  can  be  heard  per  stop.  Applications 
or  evaluations  of  this  method  have  been  reported  for 
a  variety  of  game  bird  species:  crowing  count  for 
pheasants  (Kimball  1949),  a  drumming  count  for 
ruffed  grouse  (Petraborg  et  al.  1953),  mourning 
dove  call-count  survey  (Armbruster  et  al.  1978),  and 
woodcock  (Scolopax  Minor)  singing  ground  count 
(Tautin  1982). 

A  modification  of  this  method  has  been  used  for 
lekking  species;  an  observer  listens  for  displaying 
males  along  a  roadside  route,  plots  the  locations  of 
arenas,  and  counts  the  number  of  males  per  arena  on 
subsequent  days.  The  effective  width  of  such  a  strip 
is  determined  by  twice  the  average  distance  the 
target  species  can  be  heard.  A  similar  effort  is  often 
expended  in  locating  all  arenas  on  a  block  of  habitat. 
Intensive  coverage  of  a  block  of  land  with  its  redun- 
dancy through  overlap  perhaps  gives  greater  assur- 
ance of  locating  all  the  arenas  per  unit  by  using 
listening  stops  along  a  belt. 

Still  another  variation  of  an  auditory  census  is  to 
employ  a  recorded  sound  of  the  species  in  question 
to  stimulate  a  response.  This  may  be  a  male  call  to 
elicit  a  response  by  a  territorial  male  as  suggested  by 
Kimball  (1949)  for  pheasants;  or  a  female  call  which 
may  also  cause  the  male  to  respond  as  reported  by 
Stirling  and  Bendell  (1966)  for  blue  grouse;  and 
Levy  et  al.  (1966)  for  Mearn's,  Gambel's,  and  scaled 
quail.  This  variation  introduces  another  potential 
variable — the  proportion  of  the  population  that  re- 
sponds to  the  stimulus  may  vary  from  year  to  year 
depending  on  changes  in  cover  conditions  and  popu- 
lation density. 

Accuracy  and  Precision.  The  auditory  census 
techniques  must  be  used  for  indexes  only,  at  least 
under  the  present  methods  employed.  The  data  gath- 
ered do  not  permit  conversion  to  actual  numbers  of 
birds.  Even  with  the  total  coverage  of  a  unit  for  are- 
nas and  high  counts  of  males  attending,  data  are 
available  to  show  the  presence  in  the  population  of 
an  unknown  number  of  non-participants  (Dorney 
et  al.  1958;  Rippin  and  Boag  1974). 

Certain  of  the  methods,  if  conducted  properly, 
will  provide  trend  data  of  value  to  the  manager 
(Dorney  et  al.  1958 — ruffed  grouse;  Smith  and  Galli- 
zioli  1965 — Gambel's  quail;  Jenni  and  Hartzler 
1978 — sage  grouse). 

Discussion.  Auditory  counts  hold  the  most 
promise  as  a  usable  technique  for  the  habitat  man- 
ager. Most  calls  can  be  more  easily  heard  than  birds 
can  be  seen.  Consequently,  at  least  presence  can 


Upland  Game  Birds 


423 


be  determined.  Because  most  calls  are  related  to  a 
breeding  cycle,  and  this  is  often  an  all  out  seasonal 
effort,  certain  variations  are  avoided.  Because  a  high 
proportion  of  the  male  population  usually  partici- 
pates, auditory  counts  can  be  used  on  less  dense 
populations  than  methods  relying  on  sight 
observations. 

The  major  disadvantage  to  auditory  counts  is 
that  most  are  restricted  to  the  breeding  season  and 
consequently  provide  only  an  index  to  the  adult 
population.  Indexes  to  the  adult  population  are  valu- 
able in  comparing  populations  between  years  or 
between  habitats,  but  frequently  are  inadequate  for 
use  in  predicting  a  harvestable  surplus  for  the  fol- 
lowing fall.  However,  Smith  and  Gallizioli  (1965) 
suggested  that  the  index  they  obtained  from  calling 
quail  reflected  both  numbers  of  birds  and  breeding 
success,  thus  providing  some  predictability  of  fall 
numbers  and  hunter  success.  Calling  counts  for 
mourning  and  white-winged  doves  (Zenaida  asiatica) 
were  believed  to  be  a  valid  technique  for  determin- 
ing spring  population  levels,  but  only  for  mourning 
doves  did  the  counts  provide  reasonable  predictabil- 
ity for  fall  hunting  success  (Brown  and  Smith  1976). 


Lek  (Arena)  Surveys. 


Description.  Lekking  species  (prairie  chickens, 
sharp-tailed  grouse,  and  sage  grouse  in  North 
America)  are  those  which  conduct  group  breeding 
display  on  traditional  areas  (arenas).  The  arenas  may 
be  located  through  an  auditory  census 
(triangulation),  auditory-visual  (systematic  ground 
coverage),  or  visual  (aerial  coverage).  Persistent 
attendance  by  some  of  the  males  early  in  the  season 
permits  locating  the  arenas  2  to  3  weeks  before  the 
seasonal  peak  attendance.  Peak  male  attendance  on 
sage  grouse  arenas  occurs  in  1  to  3  weeks  following 
peak  female  attendance  (Eng  1963;  Jenni  and 
Hartzler  1978).  Three  to  four  counts  of  male  sage 
grouse  on  each  arena  during  the  same  period  each 
year  will  produce  the  most  reliable  trend  data  (Jenru 
and  Hartzler  1978;  Emmons  and  Braun  1984). 
Counting  only  the  number  of  arenas  (leks)  for  lesser 
and  greater  prairie  chickens  provides  an  index  to 
density  of  displaying  males  (Cannon  and  Knopf 
1981b). 

Accuracy  and  Precision.  Over  90%  of  the  33 
radio-marked  male  sage  grouse  attended  the  display 
grounds  during  the  period  of  peak  male  attendance, 
indicating  that  most  of  the  males  of  an  area  can  be 
found  on  the  arenas  (Emmons  and  Braun  1984). 
However,  in  a  sharp-tail  population,  substantial  num- 
ber of  males  (about  one-half  during  one  year)  were 
non-territorial  and  non-participants  in  display  (Rip- 
pin  and  Boag  1974). 


Discussion.  Determining  trends  in  populations 
of  lekking  grouse  through  counts  of  displaying  males 
is  considered  an  accepted  practice.  For  sage  grouse, 
available  data  suggest  that  most  of  the  males  will 
be  present  if  counts  are  scheduled  properly.  Sage 
grouse  arenas  are  very  traditional;  arenas  are  less 
likely  to  be  abandoned  or  new  ones  started  than 
with  sharp-tailed  grouse.  Thus  for  sage  grouse, 
counts  of  males  on  arenas  is  necessary  to  reflect 
trend  data.  Counts  of  arenas  only  on  a  unit  area,  as 
recommended  for  obtaining  trend  data  for  prairie 
chickens  (Cannon  and  Knopf  1981b),  may  also  suf- 
fice for  sharptails. 


DISCUSSION 

Wildlife  biologists  responsible  for  habitat  man- 
agement on  public  lands  frequently  find  themselves 
in  the  uncomfortable  position  of  having  to  make 
recommendations  based  on  scanty  data.  Further- 
more, the  size  or  diversified  nature  of  many  areas 
make  it  difficult  to  maintain  the  intimacy  necessary 
for  sound  management.  Thus  an  ongoing  search  is 
conducted  for  yardsticks  to  be  used  in  determining 
suitability  of  cover  for  or  the  status  of  a  particular 
population. 

Probably  as  a  result  of  the  popular  use  of  key 
plant  species  in  evaluating  big  game  habitat,  game 
bird  habitat  has  been  evaluated  in  much  the  same 
manner.  Yet  a  species  of  game  bird  may  have  a  dis- 
tribution of  sufficient  breadth  that  the  plant  species 
composition  may  be  very  different  from  one  end 
of  the  range  to  the  other.  It  appears,  then,  that  at 
least  with  respect  to  cover,  structure  of  the  vegeta- 
tion is  probably  a  better  common  denominator 
throughout  the  birds'  range  than  plant  species 
composition. 

On  the  other  hand,  when  food  requirements  are 
considered,  plant  species  are  important.  Many  of 
the  same  plants,  at  least  members  of  the  same  gen- 
era, are  listed  as  food  sources  throughout  much  of 
the  range  of  Gambel's  quail.  The  entire  distribution 
of  the  sage  grouse  is  governed  by  the  plant  genus 
Artemisia.  Although  the  sage  grouse  uses  Artemisia 
for  both  food  and  cover,  their  dependence  on  this 
plant  for  winter  food  dictates  this  inseparable  rela- 
tionship. This  plant  structure  (cover)  and  plant  spe- 
cies (food)  relationship  has  been  clearly  depicted  for 
marsh  birds  by  Weller  (1978). 

Studies  of  upland  game  populations  and  of  habi- 
tat change  are  often  conducted  simultaneously.  Un- 
fortunately many  habitat  studies  are  not  initiated 
until  a  decided  decline  in  bird  numbers  has  been 
noted.  This  situation  often  results  in  attempts  to 
evaluate  deteriorated  habitat  from  which  key  ingredi- 
ents may  already  be  missing. 


424 


Upland  Game  Birds 


This  sequence  should  not  be  totally  unexpected 
because  land  use  (habitat)  changes  can  be  very  eas- 
ily overlooked  when  they  occur  gradually.  Even  the 
occasional  conversion  of  rangeland  to  cropland  may 
seem  somewhat  innocuous  to  a  manager  of  grassland 
fauna  until  it  is  conducted  on  a  scale  of  some  of  the 
recent  sod-busting  ventures.  Thus  it  is  often  the  con- 
sistent, annual,  downward  trend  in  numbers  of  a 
species  of  game  birds  that  calls  one's  attention  to  the 
habitat  problems.  A  study  is  then  initiated,  and  as 
Robel  (1980)  pointed  out,  the  population  and  habi- 
tat studies  become  inseparable. 

Knowledge  of  a  game  bird's  annual  home  range 
is  of  prime  importance  both  in  assessing  its  habitat 
or  determining  its  presence  or  relative  abundance. 
With  a  species  like  the  gray  partridge  (and  many  of 
the  quail),  the  home  range  is  sufficiently  small  so 
that  most  transects  established  for  sampling  habitat 
or  birds  would  be  of  sufficient  length  to  include 
areas  occupied  throughout  the  year.  As  such,  sea- 
sonal timing  of  the  sampling  effort  would  be  less 
critical.  Conversely,  more  mobile  species  like  sage 
grouse  (Berry  and  Eng  1984),  blue  grouse  (Mussehl 
I960),  as  well  as  migratory  doves,  may  seasonably 
occupy  habitats  somewhat  different  in  type  and 
measurably  different  in  location.  Therefore,  with 
such  species,  timing  of  the  effort  within  a  particular 
habitat  is  most  critical. 

Overton  (1971)  defined  a  census  index  as  "a 
count  or  ratio  which  is  relative  in  some  sense  to  the 
total  number  of  animals  in  a  specified  population." 


The  problem  here,  of  course,  is  determining  how  the 
count  or  ratio  relates  to  the  total  number,  since  the 
latter  figure  favors  playing  the  role  of  the  unknown. 

Most  efforts  directed  toward  measuring  the  sta- 
tus of  a  game  bird  population  result  in  data  being 
compared  to  similar  data  collected  from  the  same 
area  in  previous  years  (population  trend  data),  or  to 
similar  data  gathered  from  different  areas  (relative 
abundance  data).  Trend  data  are  most  frequently 
used  by  agencies  concerned  with  annual  population 
status  as  it  applies  to  harvest  recommendations,  al- 
though long-term  trends  can  also  provide  informa- 
tion on  the  status  of  the  habitat.  Relative  abundance 
data  from  different  areas  are  probably  of  greater 
use  to  those  agencies  primarily  concerned  with  habi- 
tat management.  Such  data  enable  personnel  to  as- 
sess the  role  of  a  given  species  within  a  unit  of 
habitat  in  relation  to  other  potential  uses. 

In  most  instances,  when  target  species  and  sea- 
son are  appropriate,  a  version  of  the  auditory  census 
will  probably  provide  the  best  return  for  the  effort. 
This  would  be  true  even  for  a  simple  presence  or 
absence  determination  but  especially  so  if  relative 
abundance  data  are  of  concern.  As  mentioned  earlier, 
the  fact  that  the  birds  (at  least  the  males)  are  en- 
gaged in  a  seasonal  activity  which  will  have  a  meas- 
urable seasonal  and  daily  peak  permits  the  observer 
to  capitalize  on  this  predictable  sequence  and 
thereby  reduce  considerably  the  variables  encoun- 
tered. A  reduction  in  variables  will  provide  a  more 
usable  measure  of  the  population  level. 


Upland  Game  Birds 


425 


LITERATURE  CITED 


ALLEN,  A.W.  1984.  Habitat  suitability  index  models:  gray 
partridge.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  Biol. 
Rep.  82(1073).  23pp. 

ARMBRUSTER,  M.H.,  T.S.  BASKETT,  W.R.  GOFORTH,  and 
K.C.  SADLER.  1978.  Evaluating  call-count  procedures 
for  measuring  local  mourning  dove  populations. 
Trans.  Missouri  Acad.  Sci.  12:75-90. 

AULT,  S.C.  and  FA.  STORMER.  1983.  Seasonal  food  selec- 
tion by  scaled  quail  in  northwest  Texas.  J.  Wildl. 
Manage.  47:222-228. 

BAXTER,  W.L.  and  C.W.  WOLFE,  Jr.  1973.  Life  history  and 
ecology  of  the  ring-necked  pheasant  in  Nebraska. 
Nebraska  Game  and  Parks  Comm.,  Lincoln.  58pp. 

BEASOM,  S.L  1970.  Turkey  productivity  in  two  vegetative 
communities  in  south  Texas.  J.  Wildl.  Manage.  34:166- 
175. 

BECK,  T.D.I.  1977.  Sage  grouse  flock  characteristics  and 
habitat  selection  in  winter.  J.  Wildl.  Manage.  41:18-26. 

BEER,  J.  1943.  Food  habits  of  the  blue  grouse.  J.  Wildl. 
Manage.  7:32-44. 

BENDELLJ.  and  P.W.  ELLIOTT.  1966.  Habitat  selection  in 
blue  grouse.  Condor.  68:431-446. 

BERRY,  J.D.  and  RL.  ENG.  1984.  Interseasonal  movements 
and  fidelity  to  seasonal  use  areas  by  female  sage 
grouse.  J.  Wildl.  Manage.  49:237-240. 

BISHOP,  R.A.  and  C.R.  HUNGERFORD.  1965.  Seasonal 

food  selection  of  Arizona  Mearn's  quail.  J.  Wildl.  Man- 
age. 29:813-819. 

BOHL,  W.H.  1957.  Chukars  in  New  Mexico,  1931-1957. 
Bull.  6,  New  Mexico  Dep.  Game  and  Fish.  Santa  Fe. 
69pp. 

BRAUN,  C.E.  1970.  Distribution  and  habitat  of  white-tailed 
ptarmigan  in  Colorado  and  New  Mexico.  Abstract  of 
paper  presented  at  46th  Annual  Meeting,  Southwest- 
ern and  Rocky  Mountain  Division,  A.A.A.S.,  Las  Vegas, 
New  Mexico. 

,  D.E.  BROWN,  J.C.  PETERSON,  and  T.P.  ZAPATKA 

1975.  Results  of  the  Four-Corners  Cooperative  band- 
tailed  pigeon  investigation.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  Resour.  Publ.  126.  20pp. 

BROWN,  D.E.  1979.  Factors  influencing  reproductive  suc- 
cess and  population  densities  in  Montezuma  quail. 
J.  Wildl.  Manage.  43:522-526. 

BROWN,  D.E.  and  R.H.  SMITH.  1976.  Predicting  hunter 
success  from  call  counts  of  mourning  and  white- 
winged  doves.  J.  Wildl.  Manage.  40:743-749- 

BROWN,  R.L.  1982.  Effects  of  livestock  grazing  on  Mearn's 
quail  in  southeastern  Arizona.  J.  Range  Manage. 
35:727-732. 

CAMPBELL,  H.,  D.K  MARTIN,  P.E.  FERKOVICH,  and  B.K 
HARRIS.  1973-  Effects  of  hunting  and  some  other 
environmental  factors  on  scaled  quail  in  New  Mexico. 
Wildl.  Monogr.  34.  49pp. 

CANNON,  R.W.  and  F.L  KNOPF.  1981a.  Lesser  prairie 

chicken  densities  on  shinnery  oak  and  sand  sagebrush 
rangelands  in  Oklahoma.  J.  Wildl.  Manage.  45:521- 
524. 

.  1981b.  Lek  numbers  as  a  trend  index  to  prairie 

grouse  populations.  J.  Wildl.  Manage.  45:776-778. 

CHOATE,  T.S.  1963-  Habitat  and  population  dynamics  of 
white-tailed  ptarmigan  in  Montana.  J.  Wildl.  Manage. 
27:684-699. 


CHRISTENSEN,  G.C.  1970.  The  chukar  partridge,  its  intro- 
duction, life  history,  and  management.  Biol.  Bull.  4. 
Nevada  Dep.  Fish  and  Game.  Reno.  82pp. 

CRAWFORD,  J.A.  and  E.G.  BOLEN.  1976.  Effects  of  land 
use  on  lesser  prairie  chickens  in  Texas.  J.  Wildl.  Man- 
age. 40:96-104. 

DALKE,  P.D.,  D.B.  PYRAH,  DC.  STANTON,  J.E.  CRAW- 
FORD, and  E.F.  SCHLATTERER.1963.  Ecology,  pro- 
ductivity and  management  of  sage  grouse  in  Idaho.  J. 
Wildl.  Manage.  27:811-841. 

DAVIS,  C.A.,  R.C.  BARKLEY,  and  W.C.  HAUSSAMEN.  1975. 
Scaled  quail  foods  in  southeastern  New  Mexico.  J. 
Wildl.  Manage.  39:496-502. 

,  T.Z.  RILEY,  RA.  SMITH,  and  M.J.  WISDOM.  1980. 

Spring-summer  foods  of  lesser  prairie  chickens  in 
New  Mexico.  Pages  75-80  in  Proc.  of  the  Prairie 
Grouse  Symposium.  Oklahoma  State  Univ.  Stillwater. 

DOERR,  P.D.,  LB.  KEITH,  D.H.  RAUSCH,  and  C.A.  FISHER. 
1974.  Characteristics  of  winter  feeding  aggregations 
of  ruffed  grouse  in  Alberta.  J.  Wildl.  Manage.  38:601- 
615. 

DOERR,  T.B.  and  F.S.  GUTHERY.  1983-  Effects  of  tebuthiu- 
ron  on  lesser  prairie-chicken  habitat  and  food.  J. 
Wildl.  Manage.  47:1138-1142. 

DORNEY,  R.S.,  DR.  THOMPSON,  J.B.  HALE,  and  R.F. 

WENDT.  1958.  An  evaluation  of  ruffed  grouse  drum- 
ming counts.  J.  Wildl.  Manage.  22:35-40. 

DUNKS,  J.H.,  RE.  TOMLINSON,  H.M.  DEEVES,  D.D.  DOL- 
TON,  C.E.  BRAUN,  and  T.P.  ZAPATKA.  1982.  Migra- 
tion, harvest,  and  population  dynamics  of  mourning 
doves  banded  in  the  Central  Management  Unit,  1967- 
77.  U.S.  Dep.  Inter.  Fish  and  Wildl.  Serv.  Spec.  Sci. 
Rep— Wildl.  249.  128pp. 

EBERHARDT,  L.L.  1978.  Transect  methods  for  population 
studies.  J.  Wildl.  Manage.  42:1-31. 

EDMINSTER,  F.C.  1954.  American  game  birds  of  field  and 
forest.  Charles  Scribner's  Sons.  New  York,  NY. 

EMLEN,JT,  JR.  and  B.  GLADING.  1945.  Increasing  valley 
quail  in  California.  Univ.  Calif.  Agric.  Exp.  Sta.  Bull. 
695. 

EMMONS,  S.R.  and  C.E.  BRAUN.  1984.  Lek  attendance  of 
male  sage  grouse.  J.  Wildl.  Manage.  48:1023-1028. 

ENG,  R.L.  1963-  Observations  on  the  breeding  biology  of 
male  sage  grouse.  J.  Wildl.  Manage.  27:841-846. 

and  P.  SCHIADWEILER  1972.  Sage  grouse  winter 

movements  and  habitat  use  in  central  Montana.  J. 
Wildl.  Manage.  36:141-146. 

GALBREATH,  D.S.  and  R.  MORELAND.  1953-  The  Chukar 
partridge  in  Washington.  Biol.  Bull.  1 1 ,  Washington 
State  Game  Dep.  55pp. 

GATES,  R.J.  1983.  Sage  grouse,  lagamorph  and  pronghorn 
use  of  a  sagebrush  grassland  burn  site  on  the  Idaho 
National  Engineering  Laboratory.  M.S.  Thesis,  Montana 
State  University,  Bozeman.  1 35pp. 

GOODWIN,  J.G.  and  C.R.  HUNGERFORD.  1977.  Habitat 
use  by  native  Gambel's  and  scaled  quail  and  released 
masked  bobwhite  quail  in  southern  Arizona.  U.S.  Dep. 
Agric.  For.  Ser.  Res.  Pap.  RM-197.  8pp. 

GRINNELL,  J.G.  and  AH.  MILLER.  1944.  The  distribution 
of  birds  of  California.  Cooper  Ornithological  Club. 
Pacific  Coast  Avifauna  27. 

GULLION,  GW.  I960.  The  ecology  of  Gambel's  quail  in 
Nevada  and  the  arid  southwest.  Ecology.  41:518-536. 

.  1977.  Forest  manipulation  for  ruffed  grouse.  Trans. 

North  Am.  Wildl.  Nat.  Resour.  Conf.  42:449-458. 

.  1984.  Ruffed  grouse  management — where  do  we 


stand  in  the  eighties?  Pages  169-181  in  Robinson, 


426 


Upland  Gome  Birds 


W.L.  ed.  Ruffed  Grouse  Management:  State  of  the  Art 
in  the  Early  1980s.  North  Cent.  Sect.  Wildl.  Soc. 

GUTIERREZ,  R.J.  1980.  Comparative  ecology  of  the  moun- 
tain and  California  quail  in  the  Carmel  Valley,  Califor- 
nia. The  Living  Bird.  71-93- 

HAMERSTROM,  FN, JR.,  and  F.  HAMERSTROM.  1961. 
Status  and  problems  of  North  American  grouse.  Wil- 
son Bull.  73:284-294. 

HARPER,  H.T.,  B.H.  HARRY,  and  WD.  BAILEY.  1958.  The 
chukar  partridge  in  California.  California  Fish  and 
Game.  44(l):5-50. 

HAYNE,  D.W.  1949.  An  examination  of  the  strip  census 
method  for  estimating  animal  populations.  J.  Wildl. 
Manage.  13:145-157. 

HIATT,  R.W.  1947.  The  relation  of  pheasants  to  agricul- 
ture in  the  Yellowstone  and  Big  Horn  River  valleys  of 
Montana.  Montana  Fish  and  Game  Dep.,  Wildl.  Rest. 
Div.,  Proj.  1-R.  72pp. 

HOFFMAN,  DM.  1962.  The  wild  turkey  in  eastern  Colo- 
rado. Colo.  Dep.  of  Game  and  Fish.  Tech.  Publ.  12. 
47pp. 

.  1965.  The  scaled  quail  in  Colorado.  Colo.  Dep.  of 

Game,  Fish  and  Parks  Tech.  Publ.  18.  47pp. 

HOFFMAN,  G.R.  and  R.R.  ALEXANDER.  1980.  Forest  vege- 
tation of  the  Routt  National  Forest  in  northwestern 
Colorado:  a  habitat  type  classification.  U.S.  Dep.  Agric, 
For.  Ser.  Res.  Pap.  RM-221.  4lpp. 

HUNGERFORD,  C.R.  1962.  Adaptations  shown  in  selec- 
tion of  food  by  Gambel's  quail.  Condor.  64:213-219. 

JACOBS,  K.F.  1959.  Restoration  of  the  greater  prairie 

chicken.  Oklahoma  Dep.  of  Wildl.  Conserv.  Oklahoma 
City. 

JEFFREY,  R.G.,  Chairman.  1977.  Band-tailed  Pigeon  (Col- 
umba  fasciata).  Pages  210-245  in  Sanderson,  G.C. 
ed.  Management  of  Migratory  Shore  and  Upland  Game 
Birds  in  North  America.  International  Assoc,  of  Fish 
and  Wildl.  Agencies.  Washington,  DC.  358pp. 

JENNI,  DA.  and  J.E.  HARTZLER.  1978.  Attendance  at  a 
sage  grouse  lek:  implications  for  spring  census.  J. 
Wildl.  Manage.  42:46-52. 

JOHNSGARD,  PA.  1973  Grouse  and  quails  of  North 
America.  Univ.  of  Nebraska  Press.  Lincoln.  553pp. 

and  R.W.  WOOD.  1968.  Distributional  changes  and 

interactions  between  prairie  chickens  and  sharp- 
tailed  grouse  in  the  Midwest.  Wilson  Bulletin.  80:173- 
188. 

JONAS,  R.  1966.  Merriam's  turkeys  in  southeastern  Mon- 
tana. Montana  Fish  and  Game  Dep.  Tech.  Bull.  3. 
36pp. 

JONES,  RE.  1963-  Identification  and  analysis  of  lesser  and 
greater  prairie  chicken  habitat.  J.  Wildl.  Manage. 
27:757-778. 

KEELER,  J.E.  1977.  Mourning  Dove  (Zenaida  macroura). 
Pages  274-298  in  Sanderson,  G.C,  ed.  Management  of 
Migratory  Shore  and  Upland  Game  Birds  in  North 
America.  International  Assoc,  of  Fish  and  Wildl.  Agen- 
cies. Washington,  DC.  358pp. 

KIMBALL,  J.W.  1949.  The  crowing  count  pheasant  census. 
J.  Wildl.  Manage.  13:101-120. 

KING,  R.T.  1937.  Ruffed  grouse  management.  J.  Forestry. 
35:523-532. 

KOSICKY,  E.L.,  GO.  HENDERSON,  P.G.  HOMEYER,  and 
E.B.  SPEAKER.  1952.  The  adequacy  of  the  fall  road- 
side pheasant  census  in  Iowa.  Trans.  North  Am.  Wildl. 
Nat.  Resour.  Conf.  17:293-305. 

KOTHMANN,  H.G.  and  G.W.  LITTON.  1975.  Utilization  of 
man-made  roosts  by  turkey  in  West  Texas.  Proc. 


Third  Natl.  Wild  Turkey  Symp.  159-163- 

KUBISIAK,  J.F.  1985.  Ruffed  grouse  habitat  relationships 
in  aspen  and  oak  forests  of  central  Wisconsin.  Tech. 
Bull.  151.  Wisconsin  Dep.  Nat.  Resour.  Madison. 
22pp. 

KUCHLER,  A.W.  1964.  Potential  natural  vegetation  of  the 
conterminous  United  States.  Am.  Geogr.  Soc.  Spec. 
Pub.  36.  39pp. 

LEE,  L.  1959.  The  present  status  of  the  wild  turkey  in  New 
Mexico.  Proc.  First.  Natl.  Wild  Turkey  Manage.  Symp.: 
11-18. 

LEOPOLD,  A.  1933-  Game  Management.  Charles  Scribner's 
Sons.  New  York,  NY.  281pp. 

LEOPOLD,  AS.  and  R.A.  MCCABE.  1957.  Natural  history  of 
the  Montezuma  quail  in  Mexico.  Condor.  59:3-26. 

LEVY,  S.H.,  J.J.  LEVY,  and  R.A.  BISHOP  1966.  Use  of  tape 
recorded  female  quail  calls  during  the  breeding  sea- 
son. J.  Wildl.  Manage.  30:426-428. 

LIGON,  J.S.  1946.  History  and  management  of  Merriam's 
wild  turkey.  New  Mexico  Game  and  Fish  Dep.  84pp. 

.  1961.  New  Mexico  birds  and  where  to  find  them. 

Univ.  of  New  Mexico  Press.  Albuquerque. 

MARSHALL,  W.H.  1954.  Ruffed  grouse  and  snowshoe  hare 
populations  on  the  Cloquet  Experimental  Forest, 
Minnesota.  J.  Wildl.  Manage.  18:109-112. 

MARTINKA,  R.R.  1972.  Structural  characteristics  of  blue 
grouse  territories  in  southwestern  Montana.  J.  Wildl. 
Manage.  36:498-510. 

MENDEL,  G.W.  and  S.R.  PETERSON.  1983.  Management 
implications  of  gray  partridge  habitat  use  on  the  Pal- 
ouse  Prairie,  Idaho.  Wildl.  Soc.  Bull.  11:348-356. 

MILLER,  G.C.  and  WD.  GRAUL  1980.  Status  of  sharp- 
tailed  grouse  in  North  America.  Pages  18-28  in  Proc. 
Prairie  Grouse  Symposium.  Oklahoma  State  Univer- 
sity. Stillwater. 

MOLINI,  W.A.  1976.  Chukar  partridge  species  manage- 
ment plan.  PR  Proj.  W-8-R,  W-43-R  and  W-48-12.  Ne- 
vada Dep.  Fish  and  Game.  Carson  City. 

MUSSEHL,  T.W.  I960.  Blue  grouse  production,  move- 
ments, and  populations  in  the  Bridger  Mountains, 
Montana.  J.  Wildl.  Manage.  24:60-68. 

.  1963-  Blue  grouse  brood  cover  and  land-use  impli- 
cations. J.  Wildl.  Manage.  27:547-555. 

NEWMAN,  D.E.  1959.  Factors  influencing  the  winter  road- 
side counts  of  cottontails.  J.  Wildl.  Manage.  23:290- 
294. 

NIELSON,  L.S.  and  C.A.  YDE.  1981.  The  effects  of  rest- 
rotation  grazing  on  the  distribution  of  sharp-tailed 
grouse.  Pages  147-165  in  Proc.  Wildl. — Livestock 
Relationships  Symposium.  Univ.  of  Idaho.  Moscow. 

OVERTON,  W.S.  1971.  Estimating  the  numbers  of  animals 
in  wildlife  populations.  Pages  402-455  in  Giles,  R.H. 
Jr.,  ed.  Wildlife  Management  Techniques,  Third  Edi- 
tion: Revised.  The  Wildl.  Soc,  Washington,  DC. 
633pp. 

PATTERSON,  R.L.  1952.  The  sage  grouse  in  Wyoming. 
Wyoming  Game  and  Fish  Comm.,  and  Sage  Books, 
Inc.  Denver,  CO.  34  lpp. 

PENDERGAST,  B.A.  and  DA.  BOAG.  1970.  Seasonal 
changes  in  the  diet  of  spruce  grouse  in  central  Al- 
berta. J.  Wildl.  Manage.  34:605-611. 

PETRABORG,  W.H.,  E.G.  WELLEIN,  and  V.E.  GUNVALD- 
SON.  1953-  Roadside  drumming  counts  a  spring  cen- 
sus method  for  ruffed  grouse.  J.  Wildl.  Manage. 
17:292-295. 

POTTS,  G.R.  1984.  Gray  partridge  population  dynamics: 
comparisons  between  Britain  and  North  America. 


Upland  Game  Birds 


427 


Pages  7-12  in  Dumke,  R.T.,  R.B.  Stiehl,  and  R.  Kohl, 
eds.  Proc.  of  PERDIX  III:  Gray  Partridge/Ring-necked 
Pheasant  Workshop.  Wisconsin  Dep.  Nat.  Resour. 
Madison. 

PYRAH,  D.B.  1972.  Effects  of  chemical  and  mechanical 
sagebrush  control  on  sage  grouse.  Pages  16-25  in 
Ecological  Effects  of  Chemical  and  Mechanical  Sage- 
brush Control.  Montana  Fish  and  Game  Dep.,  Job 
Prog.  Rep.  W-105-R6.  79pp. 

RATTI,  J.T.,  D.L.  MACKAY,  and  JR.  ALDREDGE.  1984. 
Analysis  of  spruce  grouse  habitat  in  north-central 
Washington.  J.  Wildl.  Manage.  48:1188-1196. 

RIPPIN,  A.B.  and  DA.  BOAG.  1974.  Recruitment  to  popu- 
lations of  male  sharp-tailed  grouse.  J.  Wildl.  Manage. 
38:616-621. 

ROBEL,  R.J.  1980.  Current  and  future  research  needs  for 
prairie  grouse.  Pages  34-41  in  Proc.  of  the  Prairie 
Grouse  Symp.  Oklahoma  State  Univ.  Stillwater. 

ROBINSON,  W.L.  1980.  Fool  hen— the  spruce  grouse  on 
the  Yellow  Dog  Plains.  Univ.  of  Wise.  Press.  Madison. 
221pp. 

SCHEMNITZ,  S.D.  1961.  Ecology  of  the  scaled  quail  in  the 
Oklahoma  Panhandle.  Wildl.  Monogr.  8.  47pp. 

SCOTT,  V.E.  and  EL.  BOEKER.  1977.  Responses  of  Mer- 
riam's  turkey  to  pinyon-juniper  control.  J.  Range 
Manage.  30:220-223- 

SHIELDS,  P.W.  and  DA.  Duncan.  1966.  Fall  and  winter 
food  of  California  quail  in  dry  years.  California  Fish 
and  Game.  52:275-282. 

SMITH,  L.M.,  J.W.  HUPP,  andJ.T.  RATTI.  1982.  Habitat  use 
and  home  range  of  gray  partridge  in  eastern  North 
Dakota.  J.  Wildl.  Manage.  46:580-587. 

SMITH,  R.H.  and  S.  GALLIZIOLI.  1965.  Predicting  hunter 
success  by  means  of  a  spring  call  count  of  Gambel's 
quail.  J.  Wildl.  Manage.  29:806-812. 

SNYDER,  W.D.  1974.  Pheasant  use  of  roadsides  for  nesting 
in  northeast  Colorado.  Colo.  Div.  Wildl.  Spec.  Rep. 
26.  24pp. 

SPICER,  R.L.  1959.  Wild  turkey  in  New  Mexico— an  evalu- 
ation of  habitat  development.  New  Mexico  Dep. 
Game  and  Fish.  Bull.  10.  64pp. 

STAUFFER,  D.F.  1983.  Seasonal  habitat  relationships  of 
ruffed  and  blue  grouse  in  southeastern  Idaho.  Ph.D. 
Dissertation.  Univ.  of  Idaho.  Moscow.  108pp. 

and  S.R.  PETERSON.  1985a.  Ruffed  grouse  and  blue 

grouse  habitat  use  in  southeastern  Idaho.  J.  Wildl. 
Manage.  49:461-468. 

and  S.R.  PETERSON.  1985b.  Seasonal  micro-habitat 


relationships  of  ruffed  grouse  in  southeastern  Idaho.  J. 
Wildl.  Manage.  49:605-610. 

STIEHL,  R.B.  1984.  Critical  habitat  components  for  gray 
partridge.  Pages  129-134  in  Dumke,  R.T.,  R.B.  Stiehl, 
and  T.  Kahl,  eds.  Proc.  of  PERDIX  III:  Gray  Partridge/ 
Ring-Necked  Pheasant  Workshop.  Wisconsin  Dep.  Nat. 
Resour.  Madison. 

STIRLING,  I.  andJ.F.  BENDELL.  1966.  Census  of  blue 

grouse  with  recorded  calls  of  a  female.  J.  Wildl.  Man- 
age. 30:184-187. 


STONEBERG,  R.P.  1967.  A  preliminary  study  of  the  breed- 
ing biology  of  the  spruce  grouse  in  northwestern 
Montana.  M.S.  Thesis.  Univ.  of  Montana.  Missoula. 
82pp. 

SVOBODA,  F.J.  and  G.W.  GULLION.  1972.  Preferential  use 
of  aspen  by  ruffed  grouse  in  northern  Minnesota.  J. 
Wildl.  Manage.  36:1166-1180. 

SWENSON,  J.E.  1985.  Seasonal  habitat  use  by  sharp-tailed 
grouse  (  Tympanuchus  phasianellus)  in  mixed  grass 
prairie  in  Montana.  Can.  Field  Nat.  99:40-46. 

TAUTIN,  J.  1980.  Assessment  of  some  important  factors 
affecting  the  singing-ground  survey.  Proc.  Woodcock 
Symp.  7:6-11. 

TAYLOR,  M.W.,  C.W.  WOLFE,  and  W.L.  BAXTER.  1978. 
Land-use  change  and  ring-necked  pheasants  in  Ne- 
braska. Wildl.  Soc.  Bull.  6:226-230. 

TOMLINSON,  RE.  1972a.  Current  status  of  the  endan- 
gered masked  bobwhite  quail.  Trans.  North  Am. 
Wildl.  Nat.  Resour.  Conf.  37:294-311. 

.  1972b.  Review  of  literature  on  the  endangered 

masked  bobwhite.  U.S.  Dep.  Inter.  Fish  and  Wildl. 
Serv.  Res.  Pub.  108.  Washington,  DC. 

Chairman.  1983.  Pacific  flyway  management  plan 


for  the  Pacific  coast  band-tailed  pigeon.  Pacific  Flyway 
Council.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  Port- 
land, OR.  21pp. 

TRIPPENSEE,  RE.  1948.  Wildlife  management.  McGraw 
Hill,  New  York.  499pp. 

WALLESTAD,  R.  1975.  Life  history  and  habitat  require- 
ments of  sage  grouse  in  central  Montana.  Montana 
Dep.  Fish  and  Game.  66pp. 

and  P.  SCHLADWEILER.  1974.  Breeding  season 

movements  and  habitat  use  of  male  sage  grouse  in 
central  Montana.  J.  Wildl.  Manage.  38:634-637. 
-,  J.G.  PETERSON,  and  R.L.  ENG.  1975.  Foods  of 


adult  sage  grouse  in  central  Montana.  J.  Wildl.  Man- 
age. 39:629-630. 

WEEDEN,  R.B.  1963.  Management  of  ptarmigan  in  North 
America.  J.  Wildl.  Manage.  27:673-683. 

.  1965.  Grouse  and  ptarmigan  in  Alaska:  their  ecol- 
ogy and  management.  Mimeo.  Alaska  Dep.  Fish  and 
Game.  110pp. 
-.  1967.  Seasonal  and  geographic  variation  in  foods 


of  adult  white-tailed  ptarmigan.  Condor.  69:303-309. 

WEIGAND,  J.P.  1980  Ecology  of  the  Hungarian  partridge 
in  north-central  Montana.  Wildl.  Monogr.  74.  106pp. 

and  R.G.  JANSON.  1976.  Montana's  ring-necked 

pheasants — history,  ecology  and  management.  Game 
Manage.  Div.,  Montana  Dep.  of  Fish  and  Game.  Hel- 
ena. 178pp. 

WELLER,  M.W.  1978.  Management  of  freshwater  wetlands 
for  wildlife.  Pages  267-284  in  Good,  R.,  D.  Whigham, 
and  R.  Simpson,  eds.  Freshwater  Wetlands.  Academic 
Press.  NY.  378pp. 

YEAGER,  L.E.,  WW.  SANFORT,  and  L.J.  LYON.  1951.  Some 
problems  of  pheasant  management  on  irrigated  land. 
Trans.  North  Am.  Wildl.  Nat.  Resour.  Conf.  16:351- 
367. 


428 


Upland  Game  Birds 


21 

RODENTS  AND 
INSECTIVORES 

Mayo  W.  Call 


P.O.  Box  893 
Afton,  WY  83110 


Editor's  Note:  Although  mamma logists  have  been 
studying  small  mammals  for  years,  only  recently 
have  management  agencies  begun  to  include  ro- 
dent and  insectivore  data  in  their  wildlife  surveys. 
These  recent  attempts  have  resulted  in  many  ineffi- 
cient and  unproductive  efforts.  Surveys  of  rodents 
and  insectivores  may  be  important  in  land  man- 
agement decisions  especially  in  areas  where  (I) 
threatened  or  endangered  populations  are  or  may 
be  present,  (2)  rodents  and  insectivores  indicate 
habitat  changes,  (3)  rodents  and  insectivores  are 
important  as  a  prey  base  for  animals  higher  on  the 
food  chain  such  as  raptors,  (4)  the  larger  more 
economically  important  species  such  as  beaver  and 
muskrat  are  involved,  and  (5)  rodents  and  insecti- 
vores are  locally  important  as  pest  species.  These 
situations  usually  require  careful  planning  and 
designing  efforts.  This  chapter  provides  guidance  on 
planning  and  designing  studies  for  these  diverse 
groups  of  species. 


INTRODUCTION 

Rodents  and  insectivores  (shrews  and  voles) 
live  in  almost  every  kind  of  habitat  in  North  Amer- 
ica. Some  are  common  and  widespread  in  many 
kinds  of  habitats,  such  as  deer  mice  (Peromyscus 
maniculatus),  whereas  others  live  only  in  certain 
parts  of  the  country  or  in  specific,  limited  habitats. 
For  example,  beaver  (Castor  canadensis)  and  mus- 
krats  (Ondatra  zibethicus)  are  restricted  to  riparian 
or  other  aquatic  zones  and  their  associated 
vegetation. 

Most  rodents  are  a  major  source  of  food  for 
predators,  such  as  birds  of  prey,  coyotes  (Canis  la- 
trans),  badgers  (Taxidea  taxus),  and  others.  Thus, 
they  need  to  be  highly  prolific  and  have  some  type 
of  defensive,  elusive,  or  other  protective  behavior  to 
survive.  Some  of  these  protective  mechanisms  com- 
plicate efforts  to  catch  them  or  estimate  their 
numbers. 

Before  commencing  to  inventory  or  monitor 
any  species  or  population  of  rodents  or  insectivores, 
the  biologist  should  first  define  his  or  her  objectives 
for  the  study.  The  method  used  and  intensity  of  the 
work  will  be  determined  by  the  information  that 
is  needed  for  making  the  resource  management  deci- 
sions at  hand. 

Biologists  should  recognize  the  kinds  of  prob- 
lems they  will  be  faced  with  as  they  attempt  to  esti- 
mate numbers  and  to  analyze  differences  in  species 
or  numbers  from  different  habitats  ( or  the  same 
habitat  and  area)  in  different  seasons  or  years.  Many 
variables  can  affect  results;  even  the  most  intensive 
studies,  using  the  most  sophisticated  methods,  may 


Rodents  and  Insectivores 


429 


not  provide  as  accurate  estimates  as  might  be  ex- 
pected because  one  or  more  assumptions  are  not 
met  (Schemnitz  1980). 

Some  of  the  models  used  for  estimating  rodent 
numbers  require  an  assumption  that  the  population 
is  "closed,"  i.e.,  there  is  no  recruitment  or  loss  of 
animals  to  the  population  during  the  sampling  period 
(White  et  al.  1982).  This  rarely  happens  in  the  real 
world  and  the  complicated  procedures  required 
to  adjust  for  probable  changes  in  the  population  may 
be  difficult  to  use  for  most  field  biologists.  Proce- 
dures in  this  chapter  will  be  geared  toward  simpler 
methods;  references  are  provided  for  those  who  may 
need  to  justify  more  intensive  studies. 

Populations  of  many  species  of  rodents  and  in- 
sectivores  fluctuate  greatly  in  local  areas  at  various 
times  of  the  year.  High  birth  rates,  high  mortality, 
invasion  from  surrounding  areas,  and  emigration  to 
other  areas  must  be  considered  in  inventory  and 
monitoring  studies.  Some  small  mammal  populations 
are  cyclic;  they  alternately  irrupt  and  subside  in  a 
more  or  less  uniform  manner  between  high  and  low 
levels  of  density  (Anderson  et  al.  1977).  These  popu- 
lation fluctuations  sometimes  follow  a  general  pat- 
tern with  respect  to  time  (monthly,  seasonal,  annual) 
but  other  microtine  rodents  also  experience  cycles 
of  3-  to  4-,  4-  to  7-,  and  9-  to  10-year  intervals,  as 
well  as  annual  fluctuations  (Speirs  1939;  Elton  1942; 
Dymond  1947;  Keith  1963). 

The  causal  mechanisms  of  cycles  include  biotic 
as  well  as  abiotic  factors  (Anderson  et  al.  1977). 
Biotic  factors  are  inherent  in  the  populations  them- 
selves and  in  the  interrelationships  of  different  spe- 
cies. These  include  disease,  predation,  food,  and 
physiological  mechanisms.  Abiotic  factors  are  physi- 
cal and  chemical  elements  of  the  environment,  in- 
cluding moisture,  winds,  and  solar  radiation.  An 
example  of  the  effects  of  abiotic  factors  on  nocturnal 
rodents  in  the  Lower  Colorado  River  Valley  in  Ari- 
zona was  provided  by  Anderson  et  al.  (1977).  The 
authors  stated  that,  although  there  was  a  general 
decrease  in  rodent  populations  during  the  entire 
study  period  (1974-77),  there  was  significant  intra- 
specific  asynchrony  among  the  populations  in  differ- 
ent vegetation  types.  There  was  also  a  significant 
degree  of  interspecific  asynchrony  in  population  fluc- 
tuations "which  renders  the  task  of  evaluating  habitat 
difficult  and  subject  to  error  unless  carried  out  for 
several  years  in  various  vegetation  types"  (Anderson 
et  al.  1977). 

Numerous  studies  were  reviewed  by  Jorgensen 
and  Smith  ( 1974)  relative  to  estimating  the  numbers 
and  densities  of  small  mammals  in  unrestricted  popu- 
lations. They  noted  that  these  studies  quickly  ex- 
posed some  major  difficulties  that  must  be 
accounted  for  to  produce  a  single  estimate  for  one 
restricted  location  and  that  "the  effort  required  often 


challenges  the  necessity  of  the  data,  particularly 
when  one's  interest  is  frequently  focused  on  popula- 
tions in  much  larger  areas." 

Biologists  must  consider  these  variables  and 
recognize  limitations  that  may  be  placed  on  the  data. 
For  example,  what  might  appear  to  be  a  simple  den- 
sity calculation,  based  on  a  series  of  parallel  trap 
lines  within  a  designated  area  of  habitat,  turns  out  to 
be  not  so  simple  because  one  does  not  know  the 
extent  of  movement  into  or  out  of  the  trapping  area 
by  the  various  species,  factors  affecting  trappability 
of  individuals  or  species,  times  of  night  of  greatest 
activity  of  the  different  species,  influences  of  subtle 
changes  in  weather  during  the  trapping  period,  sizes 
of  home  ranges  of  the  different  species  for  that  par- 
ticular habitat,  and  many  other  factors. 

Monitoring  studies  should  be  conducted — 

( 1 )  at  the  same  time  of  year  as  previous  studies; 

(2)  in  the  same  kind  of  weather  (as  nearly  as 
possible)  and  with  approximately  the  same 
amount  of  moonlight; 

(3)  in  the  same  habitat;  and 

(4)  with  the  same  kind  of  data  collected  each 
time. 

There  are  few  situations  that  require  informa- 
tion other  than  species  occurrence  and  relative  den- 
sity of  rodents  and  insectivores  for  Federal  land 
management  programs.  These  include — 

( 1 )  determination  of  presence  and  estimated 
numbers  of  rare,  endangered,  or  sensitive 
species  (either  Federal  or  State); 

(2)  determination  of  species,  estimated  numbers 
(biomass),  and  trends  of  small  mammals  that 
have  been  serving  as  a  prey  base  for  raptors 
or  carnivores  whose  populations  are  critical 
or  declining; 

(3)  determination  of  estimated  numbers  or 
trends  of  a  species  of  economic  or  scientific 
importance  (such  as  beaver  and  muskrat); 
and 

(4)  determination  of  effects  of  environmental 
contaminants  on  either  the  animals  them- 
selves or  on  dependent  predators. 

These  are  only  guidelines,  however,  and  the  biologist 
must  ultimately  decide  and  justify  what  information 
should  be  obtained  to  help  land  managers  maintain 
the  ecosystem  desired  in  the  area  of  interest. 

In  this  chapter,  I  will  emphasize  a  few  species 
that  are  important  in  land  management  programs 


430 


Rodents  and  Insectivores 


because  of  their  economic,  aesthetic,  or  scientific 
values.  I  will  give  beaver  special  attention  because 
they  have  the  ability  to  modify  their  habitat  more 
than  any  other  rodent  and  create  changes  that  affect 
many  other  species  of  wildlife. 

HABITAT  FEATURES  CORRELATED  WITH 
SPECIES  GROUPS 

Because  the  species  included  in  this  chapter 
occupy  virtually  every  kind  of  habitat  on  the  conti- 
nent, it  will  not  be  possible  to  discuss  individual 
species  habitats.  I  will,  however,  discuss  general  hab- 
itat types  occupied  by  the  different  families  of  ro- 
dents and  insectivores. 

Beaver  (Family  Castoridae) 

Beaver  are  found  in  most  parts  of  the  U.S.  In  the 
West  their  preferred  foods  and,  therefore,  habitat 
types  include  aspen  (Populus  sp. ),  willow  (Salix 
sp.),  cottonwood  {Populus  sp.),  alder  (Alnus  sp.), 
and  other  deciduous  plants  (Call  1970).  For  short 
periods  of  time,  beaver  may  subsist  on  conifers  or 
some  of  the  less  preferred  deciduous  species,  such  as 
hackberry  (Celtis  sp.)  and  hickory  (Carya  sp.),  but 
they  will  not  thrive  on  them. 

These  animals  are  most  often  found  along  low 
gradient  streams  where  preferred  foods  are  available 
within  about  80  m  (80-90  yd).  This  permits  more 
surface  area  to  be  impounded  with  a  given  amount 


of  beaver  dam  construction.  Valley  grades  of  1  to  6% 
are  considered  excellent  for  beaver;  those  from  7  to 
12%  are  good,  and  those  beyond  12%  are  question- 
able for  occupation  (Rutherford  1964).  Beaver  are 
sometimes  found  occupying  small  springs,  streams, 
or  seeps  on  timbered  mountain  slopes  (Call  1970). 


Muskrat  (Family  Cricetidae) 

Muskrats  are  found  in  almost  any  kind  of  flowing 
water  or  in  any  kind  of  impoundment,  natural  lake, 
pond,  or  marsh  from  sea  level  to  high  mountain 
areas.  Their  nests  may  be  in  holes  in  stream  banks  or 
in  piles  of  vegetation  and  debris  constructed  in 
ponds.  Their  presence  is  tied  largely  to  the  presence 
of  suitable  aquatic  vegetation  which  they  use  for 
food  and  nest  construction. 


Moles  (Family  Talpidae) 

Moles  {Neurotrichus  sp.,  Scapanus  sp.,  Parascal- 
ops  sp.,  Scalopus  sp.  and  Condylura  sp.)  live  most 
of  their  lives  beneath  the  ground's  surface,  usually  in 
the  upper  dirt  layers  where  digging  is  easy.  Their 
presence  may  be  detected  by  the  low  ridges  they 
push  up  as  they  move  along  just  beneath  the  surface; 
also,  by  the  earth  mounds,  each  consisting  of  0.5  to 
2  ft   of  earth,  which  they  push  up  from  beneath. 
There  is  no  indication  of  an  entrance  to  the  burrow 
(Burt  and  Grossenheider  1964). 


Beaver  dam. 


Beaver  lodge. 


Rodents  and  Insectivores 


431 


Shrews  (Family  Soricidae) 

Shrews  usually  inhabit  moist  locations,  but  some 
species  are  found  in  sagebrush  areas  of  arid  western 
States.  They  are  common  along  most  streams  and  are 
usually  found  in  moist  areas  along  streams,  under 
logs,  and  in  thick  vegetation  of  seepages. 

Squirrels  (Family  Sciuridae) 

Marmots  (Marmota  sp.),  ground  squirrels  (Sper- 
mophilus  sp.),  prairie  dogs  (Cynomys  sp. ),  and  chip- 
munks (Tamias  sp.)  nest  in  burrows  in  the  ground 
or  beneath  rocks  or  logs.  Marmots  are  most  often 
found  living  in  boulder  fields  on  talus  slopes  but  may 
also  live  in  small  rock  outcroppings,  under  isolated 
boulders  on  hillsides,  in  piles  of  logging  debris  on 
hillsides  or  in  gullies,  or  under  abandoned  houses  or 
within  bridge  abutments. 

Tree  squirrels  (Sciurus  sp. )  and  flying  squirrels 
(Glaucomys  sp. )  nest  in  tree  cavities  or  leaf  and 
grass  nests  in  tree  canopies.  The  red  squirrel  (Tam- 
iasciurus  hudsonicus)  and  northern  flying  squirrel 
(G  sabrinus)  are  the  most  common  tree  squirrels  in 
the  West.  Red  squirrels  are  found  mostly  in  pine 
(Pinus  sp.)  and  spruce  (Picea  sp.)  or  mixed  hard- 
wood forests  and  in  swamps.  The  presence  of  red 
squirrels  in  an  area  is  usually  noted  quickly  by  an 
observer  who  walks  through  a  forested  area  by 
either  the  presence  of  piles  of  shucked  cones  or  by 
the  sharp  "barking"  of  the  animals  when  intruders 
come  near. 

Chickarees  (T.  douglasii)  have  habits  similar  to 
red  squirrels  but  live  almost  exclusively  in  conifer- 
ous forests.  The  Arizona  gray  squirrel  (S.  arizonen- 
sis)  is  found  chiefly  in  oak  (Quercus  sp. )  and  pine 
forests.  Tassel-eared  squirrels  (S.  aberti)  are  found 
mostly  in  yellow  pine  (P.  ponderosa)  forests  of 


northern  Arizona  and  New  Mexico,  southeastern 
Utah,  and  southwestern  Colorado. 


'^^'auA  /&<!/     ^jLoOt^y-L^- 


432 


Rodents  and  Insectivores 


Pocket  Gophers  (Family  Geomyidae) 

Pocket  gophers  (Thomomys  sp.,  Geotnys  sp.) 
are  true  earth  burrowers,  seldom  seen  aboveground. 
Their  presence  is  easily  detected  by  the  mounds  of 
earth  they  push  out  as  they  excavate  subterranean 
tunnels.  The  mounds  are  characteristically  fan-shaped 
with  an  indication  of  the  position  of  the  opening,  a 
round  earth  plug.  The  burrow  entrance  is  seldom 
left  open  for  long.  They  prefer  soil  that  is  slightly 
moist  and  easy  to  work  (Burt  and  Grossenheider 
1964). 

Pocket  Mice,  Kangaroo  Mice,  Kangaroo  Rats 
(Family  Heteromyidae) 

Pocket  mice  (Perognathus  sp. ),  kangaroo  mice 
(Microdipodops  sp. ),  and  kangaroo  rats  (Dipodomys 
sp.)  are  all  adapted  for  living  in  arid  or  semiarid 
conditions  and  all  burrow  into  the  ground  for  nest 
sites.  They  prefer  the  more  pliable,  sandy  soils. 

Mice,  Rats,  Voles,  and  Lemmings  (Family 
Cricetidae) 

Mice  (Reitbrodontomys  sp.,  Peromyscus  sp., 
Onychomys  sp.,  Sigmodon  sp.),  New  World  rats 
(Sigmodon  sp.,  Neotoma  sp. ),  and  voles  (Cletbriono- 
mys  sp.,  Phenacomys  sp.,  Microtus  sp.,  and  Lagurus 
sp. )  live  mostly  on  the  ground,  some  in  rocks,  and 
a  few  in  trees.  Grasshopper  mice  (Onchomys  sp. ) 
live  chiefly  on  prairies  and  southwest  desert  areas. 
Harvest  mice  (Reitbrodontomys  sp.)  seem  to  prefer 
rather  dense,  low  vegetation.  White-footed  (P.  leuco- 
pus)  and  deer  mice  are  nocturnal,  living  in  woods, 
prairies,  and  around  rocks  and  buildings.  Most  are 
ground  dwellers,  but  some  nest  in  trees  and  are 
largely  arboreal. 


Woodrats  (Neotoma  sp. )  normally  occupy  the 
pinyon-juniper  (Pinus  edulus-Juniperus  sp. )  plateau 
region,  frequenting  rocky  places  and  even  cliffs 
(Hoffmeister  and  Luis  de  la  Torre  I960),  but  some 
forms  will  be  found  in  Southwest  deserts  and  others 
in  mountain  brush  of  the  Northwest. 

Woodrats  are  active  mostly  at  night.  They  pile 
sticks  and  rubbish  together  for  nests  in  cliff  holes  or 
on  ledges.  In  the  plains  and  deserts,  they  build  stick 
and  cactus  nests,  often  1  to  2  m  (4  ft  or  more)  in 
diameter,  in  clumps  of  cactus,  yucca,  or  brush.  On 
the  West  Coast  their  nests  may  be  found  in  live  oak 
trees. 

Various  species  of  lemmings  (Synaptomys  sp. ) 
may  be  found  from  bogs  in  the  eastern  U.S.  to  arctic 
and  tundra  habitats  of  the  far  north.  The  bog  lem- 
mings (S.  borealis)  prefer  wet  bogs  and  meadows 
where  there  is  a  thick  mat  of  ground  vegetation 
(Burt  and  Grossenheider  1964). 


Deer  mouse. 


Redback  voles  (Clethrionomys  sp.)  are  small 
rodents  of  forested  areas  in  Alaska,  Canada,  and  scat- 
tered forested  areas  of  the  U.S.  Some  of  the  meadow 
voles  and  other  species  may  be  found  in  most  re- 
gions where  there  is  a  good  grass  cover,  especially 
where  it  is  moist.  At  least  one  species,  the  Townsend 
vole  (Microtus  townsendii),  inhabits  moist  fields 
and  meadows  from  tidewater  up  to  alpine  meadows. 
The  presence  of  narrow  runways,  2.5  to  5.0  cm  ( 1 
to  2  in. )  wide,  through  the  matted  grass  usually  indi- 
cate their  presence.  In  a  few  places,  voles  are  found 
among  rocks  or  on  forest  floors  where  there  is  no 
grass. 

Phenacomys  sp.  may  inhabit  grassy  areas  within 
the  cold  forested  regions  of  Canada  and  Alaska,  the 
high  mountain  tops  of  western  U.S.,  or  dense  for- 
ested regions  of  the  Northwest.  Most  are  ground-liv- 


ing, but  the  tree  Phenacomys  feed  almost  entirely  on 
leaves  of  the  tree  in  which  they  are  living — spruce, 
hemlock  (Tsuga  sp),  or  fir  (Abies  sp). 

Jumping  Mice  (Family  Zapodidae) 

Jumping  mice  (Zapus  sp.  and  Napaeozgapus 
sp. )  normally  inhabit  damp  meadows  and  forests  and 
hibernate  during  the  winter. 

Porcupines  (Family  Erethizontidae) 

Porcupines  (Erethizon  dorsatum)  may  be  seen 
lumbering  through  the  forest,  along  canyon  bottoms, 
or  hunched  into  large  dark  balls  high  in  trees.  They 
may  be  found  in  deciduous  or  coniferous  forests  and 
strangely,  sometimes  in  shrublands  or  prairies,  miles 
from  the  nearest  trees. 


Porcupines  sometimes  hide  in  trees  by  curling  into  a  large 
dark  ball. 


Old  World  Rats  and  Mice  (Family  Muridae) 

This  group  includes  the  Norway  rat  (Rattus 
norvegicus),  black  rat  (R.  rattus),  and  house  mouse 
(Mus  musculus).  None  of  these  animals  normally 
occur  far  from  man-made  structures  and  will  not  be 
considered  in  this  chapter. 


POPULATION  MEASUREMENT  TECHNIQUES 

With  the  variety  of  rodents  and  insectivores 
present  in  most  habitats,  more  than  one  method  is 
usually  required  to  determine  species  present,  rela- 
tive density,  or  estimated  numbers.  Space  is  not 
available  here  to  present  the  many  variations  in  ro- 
dent/insectivore  behavior  as  may  apply  to  sampling 
techniques.  In  fact,  there  remains  much  to  be 
learned  about  small  mammal  home  ranges,  behavior 
in  different  seasons,  cycles,  responses  to  traps,  and 
other  aspects  of  their  ecology  that  affect  surveys. 


Rodents  and  Insectivores 


433 


Volumes  of  material  have  been  written  about 
statistical  procedures  that  may  be  used  in  sampling 
wild  populations.  Depending  on  their  needs  for  ani- 
mal population  data,  biologists  must  decide  if  statisti- 
cal procedures  are  needed  and  refer  to  the  Literature 
Cited  section  for  detailed  accounts  of  the  necessary 
applications.  Not  every  sampling  of  rodent  or  insecti- 
vore  populations  needs  to  be  analyzed  by  elaborate 
statistical  procedures  or  complicated  sampling  pro- 
cedures. Much  useful  information  can  be  derived  on 
species,  relative  density,  and  trends  from  the  raw 
data  itself,  together  with  simple  statistics  such  as 
means  and  variances. 


General  Considerations 

Before  starting  any  project,  the  biologist 
should — 

( 1 )  determine  the  objective  (what  kind  and  how 
much  information  is  necessary); 

(2)  determine  what  kinds  of  inventory  methods 
or  sampling  will  be  required;  and 

(3)  design  the  study. 

For  most  land  management  programs,  only  infor- 
mation on  relative  density  or  species  trends  will  be 
required  for  rodents  and  insectivores.  At  times,  how- 
ever, the  biologist  may  decide  that  additional  data 
are  needed  on  either  a  specific  species  or  total  popu- 
lations of  several  species.  Computer  programs  are 
available  to  assist  with  the  more  difficult  computa- 
tions (Schultz  1961;  Snedecor  and  Cochran  1967; 
Sokal  and  Rohlf  1969;  Petrusewicz  and  MacFayden 
1970;  Seber  1973;  Anderson  et  al.  1976;  Burnham  et 
al.  1980;  White  et  al.  1982). 

For  most  rodents  and  insectivores,  counts  of 
individual  animals  are  a  practical  impossibility. 
Therefore,  one  normally  takes  a  sample  from  a  popu- 
lation and  attempts  to  derive  a  reasonable  estimate 
of  the  total  population  from  that  sample.  The  esti- 
mate for  the  total  population  is  then  based  on  the  as- 
sumption that  all  animals  in  the  population  occur 
throughout  the  defined  study  area  in  the  same  ratio 
as  they  did  within  the  sampled  area. 

An  index  may  be  employed  for  monitoring  a 
population.  An  index  is  an  object  (such  as  a  beaver 
lodge)  that  is  related  to  the  number  of  animals  in 
the  area.  The  object  is  usually  easier  to  count  than 
the  animal  itself.  For  example,  the  number  of  mus- 
krat  houses  or  beaver  lodges  is  an  index  to  the  num- 
ber of  animals  living  in  the  area.  Through  intensive 
trapping,  or  by  using  values  derived  from  other  re- 
searchers in  the  general  area,  one  can  estimate  the 
number  of  animals  in  the  area  by  multiplying  the 
number  of  houses  or  lodges  by  the  normal  number 


of  occupants  per  house  or  lodge.  One  of  the  prob- 
lems in  using  an  index  is  the  necessary  assumption 
that  the  ratio  of  the  index  to  the  population  is  the 
same  in  the  several  areas  or  populations  being  com- 
pared (Schemnitz  1980). 

The  basic  premise  used  in  rodent  and  insecti- 
vore  monitoring  is  that  a  similar  proportion  of  the 
total  population  is  tabulated  each  time  the  same 
technique  is  used  under  similar  field  conditions  with 
the  same  species.  No  statistical  procedure  will  make 
data  collected  under  different  conditions  compara- 
ble. With  all  techniques  for  rodents  and  insectivores 
that  employ  direct  counts,  the  major  problem  is  to 
conduct  counts  under  sufficiently  similar  conditions, 
so  that  comparison  of  counts  is  valid  (Schemnitz 
1980). 

Most  populations  of  rodents  and  insectivores  are 
not  distributed  randomly.  Each  species  will  tend  to 
be  more  abundant  within  small  microhabitats  that 
commonly  occur  within  a  general  habitat  type,  even 
when  the  general  habitat  appears  to  be  homogene- 
ous. Cover  and  food  are  not  usually  distributed  ran- 
domly, so  neither  are  the  animals.  Some  burrowing 
rodents  prefer  certain  soil  textures  and  types  for 
digging  their  tunnels,  and  these  soils  will  vary  some 
in  what  appears  to  be  a  fairly  homogeneous  vegeta- 
tion type.  Thus,  using  the  same  transect  site  for  mon- 
itoring in  subsequent  years,  or  selecting  sites  for 
comparative  purposes,  which  are  as  nearly  alike  as 
possible,  may  reduce  variances. 


Sampling  for  Large  to  Medium-Sized 
Rodents 

Because  of  the  diversity  of  rodents  and  insecti- 
vores, a  number  of  different  survey  techniques  are 
required  in  any  given  area  to  determine  all  species 
present,  their  relative  density  (abundance),  or  esti- 
mated numbers.  I  will  discuss  some  of  the  general 
methods  suitable  for  most  species  and  provide  addi- 
tional references  for  more  sophisticated  techniques. 
In  some  instances  the  biologist  may  need  to  devise 
innovative  methods  to  determine  presence  or  esti- 
mated numbers  of  endangered  or  rare  species  or 
animals  that  are  trap-shy  or  elusive  in  other  ways. 


Beaver. 


Presence.  The  presence  of  beaver  along  a 
drainage  is  determined  by  looking  for  beaver  dams, 
lodges,  or  beaver  cuttings  (Figure  1).  Along  desert 
streams  and  larger  rivers,  beaver  live  in  bank  dens 
and  do  not  build  dams,  but  their  tree-  or  shrub- 
cutting  activities  will  be  obvious.  Most  stream 
drainages  in  western  mountains  that  contain  aspen, 
willow,  or  cottonwood  will  contain  beaver. 


454 


Rodents  and  Insectivores 


Population  Estimate.  The  average  number  of 
beaver  per  colony  for  the  entire  U.S.  and  Canada  is 
about  five  (Denney  1952).  Where  beaver  trapping 
occurs  over  widespread  areas,  the  normal  colony 
composition  is  disrupted.  Although  this  makes  it  less 
precise  to  use  a  normal  average  number  for  popula- 
tion estimates,  the  method  described  will  suffice 
for  most  management  purposes. 

The  suggested  population  estimate  method  is  to 
cruise  streams  in  October  or  November,  counting 
the  active  colonies  (as  evidenced  by  food  caches  and 
repairs  on  dams  and  visible  lodges)  and  multiplying 
the  number  found  by  five.  Food  caches  are  easily 
discernible  and  are  the  most  obvious  indication  of  a 
wintering  beaver  colony.  The  size  of  the  food  cache 
is  some  indication  of  the  number  of  beaver  in  the 
colony  (Figure  2).  However,  there  is  no  direct  rela- 
tionship between  the  size  of  the  food  cache  and 
the  number  of  beaver  in  the  colony  because  ( 1 )  dif- 
ferent colonies  may  use  different  kinds  of  food  mate- 
rials in  their  caches  and  (  2 )  some  colonies  will  have 
much  more  readily  available  food  (aspen,  willow, 
etc. )  adjacent  to  their  primary  pond.  In  a  few  places, 
such  as  western  Oregon,  beaver  may  not  use  food 
caches  but  will  cut  the  needed  food  all  winter  long. 
In  such  areas,  the  beaver  can  be  surveyed  by  count- 
ing the  number  of  individual  cutting  areas  along 
the  river  which  should  be  concentrated  near  their 
bank  dens  (C.  Maser,  pers.  commun.). 


Even  where  colonies  do  not  have  a  normal 
makeup  because  of  trapping  each  year,  this  method 
will  be  satisfactory  for  most  management  purposes.  A 
few  colonies  are  usually  overlooked  and  a  few  bank- 
dwelling  beaver  are  usually  present  but  not  dis- 
cerned, which  will  largely  make  up  for  a  reduction 
in  the  average  number  of  beaver  per  colony  caused 
by  trapping. 

If  the  task  of  determining  the  number  of  beaver 
involves  many  miles  of  streams  or  relatively  inacces- 
sible drainages,  the  best  method  to  use  is  an  aerial 
survey,  either  by  fixed-wing  or  helicopter.  The  ter- 
rain will  dictate  the  type  of  aircraft  to  use.  It  is  best 
to  fly  within  90  to  151  m  (300  to  500  ft)  of  the 
ground  so  that  all  food  caches  may  be  seen  and 
counted.  Such  food  caches  can  be  seen  on  aerial 
surveys,  even  when  ice  has  formed  on  the  ponds 
(Figure  3).  This  is  not  difficult  in  a  fixed-wing  air- 
craft along  wide  valley  bottoms  with  low  stream 
gradients,  but  becomes  more  challenging  in  rough 
terrain. 

Muskrat. 


Presence.  Muskrats  are  noted  throughout  most 
of  their  geographical  distribution  for  their  conical 
or  dome-shaped  houses  of  vegetation  set  in  marshes 
or  in  shallow  pond  waters.  However,  in  many  areas, 


Figure  1.     A  two-lodge  beaver  pond  in  southeastern 
Wyoming.  This  pond  is  about  1  acre  and  was 
about  20  years  old  at  the  time  this  picture  was 
taken. 


Figure  2.     The  same  beaver  pond  as  shown  in  Fig- 
ure 1,  as  viewed  from  the  air.  Food  caches,  not 
present  in  this  photograph,  are  usually  built 
adjacent  to  dens  or  lodges,  and  are  easy  to  dis- 
cern from  this  altitude. 


Rodents  and  Insectivores 


435 


Figure  3.  A  relatively  large  food  cache  in  a  beaver 
primary  pond.  Eleven  beaver  were  trapped  from 
this  colony. 


especially  along  ditches  and  rivers,  they  dig  bank 
burrows.  Some  marshes  contain  dozens  of  the  dome- 
shaped  houses  that  leave  no  doubt  as  to  the 
occupants,  but  detecting  the  presence  of  muskrats 
that  live  along  streams  is  more  difficult.  Burrow 
entrances  are  usually  underwater.  Their  presence  in 
such  areas  can  best  be  detected  by  watching  for 
cuttings  of  cattails,  rushes,  sedges,  small  willows, 
pondweeds,  water  lilies,  and  other  aquatic  and 
terrestrial  plants  on  which  they  feed.  They  are  active 
during  the  day  and  may  occasionally  be  seen 
swimming  slowly  up  or  down  the  stream  or  sitting 
quietly  on  a  sandbar,  partly  submerged  tree  trunk,  or 
other  convenient  object  where  they  may  sun 
themselves  and  still  be  close  to  water  for  a  quick 
escape  from  danger. 

Population  Estimate.  Estimating  numbers  of 
muskrats  in  an  area  is  very  difficult.  The  number 
of  houses  in  a  marsh  is  only  an  index  to  the  number 
of  muskrats. 

One  method  for  monitoring  trends  of  these 
animals  is  to  question  local  trappers  concerning  their 
harvest  from  local  habitats  (C.  Maser,  pers.  com- 
mun.).  For  example,  if  trappers  believe  they  have 
harvested  about  60%  of  the  existing  population  of  a 
marsh  or  stream  and  have  taken  70  muskrats,  then 
the  population  at  the  beginning  of  the  trapping  pe- 
riod would  have  been  about  1 17  animals.  Regardless 
of  trappers'  estimated  percentage  of  harvest,  trends 
in  the  annual  harvest  would  give  an  indication  of 
trends  in  the  population  level.  The  basic  assumption 


is  that  prices  and,  therefore,  interest  in  trapping, 
make  it  worthwhile  for  trappers  to  conduct  about 
the  same  intensity  of  trapping  over  the  years. 

A  mark  and  recapture  method,  using  the  Lincoln 
Index,  could  be  used  to  estimate  the  population  but 
would  be  time-consuming.  High  mortality  from  pre- 
dators could  distort  the  results  if  too  much  time 
were  consumed  in  live  trapping  and  marking  a  num- 
ber of  animals. 

Prairie  Dog. 


Presence.  The  black-tailed  prairie  dog 
(Cynomys  ludovicianus)  is  especially  gregarious, 
i.e.,  it  lives  in  "towns."  Its  presence  is  usually 
revealed  by  a  group  of  bare  mounds  from  8  to  23  m 
(25  to  75  ft)  apart  and  each  mound  from  0.03  to 
1  m(l  to  3ft)  high  (Burt  and  Grossenheider  1964). 
Colonies  of  several  hundred  animals  are  not 
uncommon  on  some  of  the  dry  upland  prairies  on 
which  they  live.  Populations  may  vary  from  12  to  85 
per  ha  (5  to  34  per  a.)  or  more. 

Some  white-tailed  prairie  dogs  (C.  leucurus)  are 
widely  scattered  across  sagebrush  flats  and  rolling 
grasslands.  For  family  groups  that  occur  in  habitats 
where  they  are  largely  concealed,  their  presence 
is  often  revealed  by  their  danger  signal — a  two-sylla- 
ble bark,  issued  at  about  40  per  minute. 

Utah  prairie  dogs  (C.  panndens)  have  been  ob- 
served in  both  dense  and  scattered  colonies.  I  have 
counted  more  than  500  prairie  dogs  in  a  pasture 
of  about  4  ha  ( 10  a.)  near  Cedar  City,  Utah. 

Population  Estimate.  Prairie  dogs  live  in  fam- 
ily groups  and  exhibit  some  territoriality  around 
their  dens  but  may  use  communal  feeding  areas 


436 


Rodents  and  Insectivores 


away  from  the  dens.  An  estimate  of  numbers,  prefer- 
ably when  their  population  is  somewhat  stabilized 
in  the  fall,  can  be  obtained  by  the  mark  and  recap- 
ture method  using  Lincoln  Index  calculations  (see 
small  mammal  sampling).  However,  this  procedure  is 
very  time-consuming. 

I  recommend  that  inventorying  and  monitoring 
of  these  animals  be  based  on  the  number  of  active 
dens,  rather  than  on  number  of  animals  observed  at 
any  given  time.  Areas  containing  prairie  dogs  should 
be  mapped  out  and  then  a  grid  established  over  the 
colony  area  to  facilitate  the  counting  of  burrows. 
Burrows  can  be  plotted  on  the  map  for  comparison 
of  numbers  in  future  years.  Stakes  or  prominent 
shrubs  can  be  used  as  corners  of  the  grid  and 
brightly  colored  flagging  used  for  easy  reference 
while  walking  and  counting  burrows.  Additional  flag- 
ging should  be  placed  along  the  edges  of  the  colony 
at  intervals  of  about  10  m  (30  yd)  along  each  side  to 
effectively  cut  the  colony  area  into  squares  so  that 
burrows  can  be  counted  and  plotted  on  a  map.  Grid 
square  numbering  can  be  done  the  same  as  that  for 
the  International  Bird  Plot  Census  Method,  or  in  any 
other  convenient  manner  for  future  comparisons. 

Marmots. 


Presence.  Presence  of  these  animals  is  usually 
revealed  by  their  sharp  chirps  when  intruders 
approach.  The  den  area  is  rank  with  the  odor  of 
their  feces.  By  sitting  near  a  rockslide  in  early 
morning  and  watching  for  activity,  the  observer  can 
usually  determine  the  presence  of  these  animals 
without  doing  any  trapping.  Watching  an  area  of 
interest  for  a  few  hours  will  generally  give  the 
observer  a  rough  idea  of  how  many  of  the  animals 
are  present.  Both  young  and  old  animals  like  to  sun 
themselves  on  rocks  or  logs  near  their  dens. 

Population  Estimate.  Most  colonies  of  these 
animals  are  small.  A  mark  and  recapture  method  can 
be  used  for  estimating  numbers  by  marking  some 
of  the  animals  with  easily  visible  tags  and  then 
spending  a  period  of  3  or  4  days  observing  the  ani- 
mals and  tallying  all  marked  and  unmarked  animals. 
Lincoln  Index  procedures  can  then  be  used  to  esti- 
mate total  numbers  present. 

Woodrats. 


Presence.  Presence  is  revealed  by  large  stick 
or  cactus  houses  in  clumps  of  brush,  cactus,  yucca, 
or  in  crevices  in  cliffs.  They  may  also  take  over 
abandoned  cabins  or  other  old  buildings  in  foothills 
and  canyons  where  they  build  large  nests  containing 
rocks,  bones,  sticks,  leaves,  and  other  debris.  Nests 
built  on  ledges  of  cliffs  or  rocky  outcroppings  are 
sometimes  confused  with  nests  of  ravens  or  birds  of 


prey.  But  the  inclusion  of  bones  or  fragments  of 
human  refuse,  including  pieces  of  glass  or  other 
shiny  objects,  usually  identify  it  as  a  woodrat  nest. 
There  may  be  from  one  to  six,  or  more,  rats  per 
nest.  The  smelly  urine  stains  on  rafters  or  narrow 
cliff  ledges  near  their  nests  also  reveal  the  identity  of 
the  occupants. 


Rocky  slopes  are  home  for  the  yellow-bellied  marmot. 


Population  Estimate.  Colonies  of  woodrats 
are  usually  widely  scattered.  Rough  estimates  of  their 
abundance  may  be  obtained  by  walking  transects 
along  the  desert  floor  or  canyon  hillsides  while 
watching  for  their  large  stick  nests  which  are  often 
1.5  to  2  m  (4  to  6  ft)  in  diameter  and  1  m  (3  ft)  tall. 
When  located,  they  can  be  checked  for  tracks  and 
fresh  material  to  determine  if  they  are  active.  Rela- 
tive density  of  this  animal  can  be  determined  along 
with  other  small  mammals  through  the  periodic  use 
of  trapping  transects  (see  section  on  small  mammal 
sampling). 

Tree  Squirrels.  The  tree  squirrels,  including  the 
red  squirrel,  western  gray  squirrel  (5.  griseus),  fox 
squirrel  (S.  niger),  tassel-eared  squirrels,  and 
northern  flying  squirrels,  can  usually  be  identified  by 
a  combination  of  region  of  country  where  found 
and  the  signs  which  they  leave. 

Presence.  They  usually  have  favorite  stumps  or 
racks  where  shucks  from  pinecones  or  nuts  may 
accumulate  in  piles  of  a  bushel  or  more.  Their  nests 
are  either  in  cavities  in  trees,  on  outside  branches 
built  of  leaves,  twigs,  and  shredded  bark,  or  some- 
times in  cavities  at  the  bases  of  trees.  Conifer  cones 
and  nuts  are  stored  in  caches  under  logs  or  in  cavi- 
ties in  the  ground  or  among  tree  roots.  They  are 
active  mostly  in  daylight  hours. 


Rodents  and  Insectivores 


437 


There  are  seldom  more  than  two  chickarees  or 
red  squirrels  per  ha  (per  5.0  a.)  of  forest,  but  there 
may  be  as  many  as  10  per  0.4  ha  (1  a.;  Burt  and 
Grossenheider  1964).  Chickarees  are  fairly  quiet  dur- 
ing spring  and  summer  while  raising  their  young 
but  become  vociferous  in  fall.  Thus,  fall  is  the  best 
time  to  check  for  their  presence  (Maser  et  al.  1981). 
Nests  may  be  bulky  structures  built  of  twigs,  mosses, 
or  lichens,  well  within  the  crown  of  the  tree  or  in 
hollows  of  trees. 

Northern  flying  squirrels  are  strictly  nocturnal. 
They  usually  nest  in  tree  cavities  but  may  also  build 
outside  nests  of  leaves  and  twigs.  Normally,  nests  are 
round  balls  of  material  situated  in  a  fork  of  the  trunk 
of  a  tree  or  on  a  whorl  of  limbs  against  the  trunk 
(Maser  et  al.  1981).  In  most  areas,  few  people  realize 
how  abundant  these  night-gliders  really  are.  Maser 
et  al.  (1981)  stated  that  trappers  who  trap  marten 
sometimes  catch  hundreds  of  these  squirrels  each 
winter  in  traps.  These  animals  are  fairly  quiet  and 
generally  must  be  detected  by  observations  at  night 
in  suspected  habitats  or  caught  with  traps  baited 
with  fresh,  raw,  or  putrid  meat,  which  they  seem  to 
relish  (Jackson  1961).  Much  of  their  diet  consists 
of  green  vegetation,  insects,  nuts,  seeds,  and  fruits 
and  they  leave  few  signs  of  their  feeding. 


The  Arizona  gray  squirrel  (5.  arizonensis)  nests 
either  in  tree  cavities  or  in  tree  branches.  It  stores 
nuts  and  seeds  singly  in  small  holes  in  the  ground 
rather  than  in  large  food  caches.  The  squirrels  will 
cling  to  the  opposite  sides  of  trees  or  lie  quietly 
along  a  branch  and  never  reveal  themselves  as  long 
as  an  intruder  is  moving  near  them.  Visual  observa- 
tions are  the  best  means  for  detecting  presence  of 
these  animals  and  for  estimating  numbers.  Two  per- 
sons working  together,  but  a  short  distance  apart,  are 


effective  in  locating  the  animals.  Depending  on  the 
habitat,  there  may  be  from  2  to  20  squirrels  per  0.4 
ha  (1  a.;  Burt  and  Grossenheider  1964). 

Tassel-eared  squirrels  build  bulky  nests  high  in 
the  pines.  They  are  usually  found  from  2,121  to 
2,424  m  (7,000  to  8,000  ft)  in  elevation.  They  may 
be  heard  "barking"  when  excited  but  are  usually 
quiet.  Their  presence  is  best  detected  by  listening 
for  their  calls  and  observing  them  in  appropriate 
habitats. 


Porcupine. 


Presence.  The  presence  of  porcupines  is  most 
often  determined  by  watching  for  areas  of  bark 
removal  in  coniferous  trees  and  bushes  or  in  winter 
by  finding  their  waddling  trails  in  the  snow.  They  are 
often  observed  at  great  distances  in  winter  as  dark 
balls  in  the  tops  of  bushes  or  deciduous  trees  as  they 
feed  on  the  bark.  Their  brownish  droppings, 
consisting  of  woody  materials,  are  very 
characteristic.  They  often  have  a  home  den  where 
they  may  "hole  up"  during  bad  weather  or  before 
the  birth  of  the  single  young.  The  den  may  be  a 
small  cave,  crevice  in  a  cliff  or  among  boulders  and 
talus,  or  in  a  hollow  log.  Their  feces  accumulate  in 
such  places  and  reveal  their  presence. 

Population  Estimate.  Porcupines  seem  to  be 
scattered  sparsely  over  most  areas.  Some  estimate  of 
numbers  can  be  obtained  by  running  a  1.6-  to  32- 
km  (1-  to  2-mi.)  transect  through  open  forest  or 
mixed  deciduous-conifer  forest  types  while  watching 
for  signs.  Being  fairly  sparse,  they  would  be  easy  to 
miss  entirely.  Driving  unimproved  roads  through 
forested  areas  at  night  would  probably  reveal  more 
porcupines  than  daytime  observations.  Late  one 
night  in  the  fall  of  1950,  I  counted  21  porcupines 
along  32  km  (20  mi.)  of  Forest  Service  road  in  west- 
ern Wyoming.  I  doubt  I  could  have  found  half  as 
many  during  daylight  hours. 

Caution  should  be  used  in  trying  to  estimate 
numbers  of  porcupines  per  unit  of  land  area  because 
the  animals  will  sometimes  concentrate  along  can- 
yon bottoms  or  other  desirable  habitats. 


Sampling  for  Small  Rodents  and 
Insectivores 

Inventorying  of  most  small  rodent  and  insecti- 
vore  species  can  best  be  accomplished  with  trapping 
transects.  Subsequent  monitoring  of  species  and  pop- 
ulations should  be  done  at  the  same  sites  and  with 
the  same  methods  as  the  original  inventory.  All  trap 
lines  will  provide  information  on  species  present  and 
relative  density;  if  certain  designs  are  used,  one  can 


438 


Rodents  and  Insectivores 


also  calculate  an  estimate  of  the  populations  in- 
volved. However,  density  is  a  more  complicated 
problem  and  if  such  data  are  needed,  the  biologist 
will  need  to  research  literature  that  describes  the 
required  field  designs  and  statistical  analyses  (Smith 
et  al.  1971a;  Smith  et  al.  1972;  Smith  et  al.  1975; 
O'Farrell  et  al.  1977;  Scott  et  al.  1978;  Burnham  et 
al.  1980;  White  et  al.  1982).  The  degree  of  accuracy 
of  either  population  or  density  estimates  will  depend 
largely  on  meeting  certain  assumptions.  The  follow- 
ing procedures  can  be  used  for  trapping  small  mam- 
mals to  obtain  species,  relative  density,  and 
estimated  numbers — but  not  density. 

Species  Occurrence  and  Relative  Density.  For 

this  information,  I  suggest  following  the  procedures 
used  by  Anderson  et  al.  (1977)  and  Kepner  (1978): 


( 1 )    Be  sure  the  vegetation  communities  are  de- 
scribed so  that  relationships  between  plant 
and  animal  communities  can  be  suggested. 
Soil  types  should  also  be  described. 


(5)    Tabulate  and  analyze  your  results.  Catches 
will  normally  be  expressed  as  numbers  of 
each  species  per  270  trap  nights  (if  the  above 
system  is  used).  If  the  same  type  of  trapping 
grid  is  used  in  different  communities,  com- 
parison between  species  and  relative  density 
can  be  made  for  different  communities.  It 
will  soon  become  obvious  which  communi- 
ties are  most  important  in  terms  of  total  spe- 
cies and  numbers.  These  communities, 
however,  may  not  be  the  ones  where  rare  or 
indicator  species  are  found  that  will  require 
special  management,  and  you  will  need  to 
sample  the  preferred  habitats  used  by  these 
species. 

Live  traps  suitable  for  taking  medium-sized  ro- 
dents, such  as  squirrels  and  woodrats,  may  be  substi- 
tuted for  the  Victor  rat  trap,  if  desired  (Figure  5). 
However,  if  this  substitution  is  made,  the  animals 
captured  should  be  removed  from  the  area  or  kept 
in  captivity  until  all  trapping  is  completed. 


(2)    Establish  sampling  grids  within  each  vegeta- 
tion community  where  data  are  needed.  Each 
grid  will  consist  of  snap  trap  stations  placed 
along  two  parallel  lines  of  traps  15  m  (50  ft) 
apart  (Figure  4).  Each  line  will  consist  of 
15  stations  that  are  also  15  m  apart.  At  each 
station,  set  two  museum  special  traps  and 
one  Victor  rat  trap.  Bait  them  with  rolled  oats 
and  peanut  butter  with  a  chemical  mixed  in, 
such  as  dimethyl  pthalate,  to  repel  insects 
(Anderson  and  Ohmart  1977). 


(3)    Set  the  traps  in  early  evening  and  check  them 
every  hour  until  midnight,  if  possible.  Set 
them  again  at  midnight  and  leave  until  morn- 
ing, then  check  them,  remove  the  animals, 
and  leave  the  traps  sprung  until  evening. 
Checking  the  traps  every  hour  before  mid- 
night will  remove  most  of  the  easily  caught 
animals,  such  as  deer  mice,  and  leaving  the 
traps  set  will  increase  chances  to  catch  less 
abundant  or  more  trap-shy  animals  for  the 
remainder  of  the  night. 


(4)    Run  the  trap  lines  for  three  consecutive 
nights.  Tabulate  all  catches  by  species,  sex, 
and  age  group  (sub-adult  and  adult).  In  moni- 
toring a  small  mammal  population,  recognize 
that  animals  are  caught  differentially  by  spe- 
cies, sex,  and  age,  and  the  catch  will  depend 
largely  on  season  of  the  year.  In  any  future 
monitoring  of  the  same  community,  be  sure 
to  trap  at  the  same  season  to  obtain  compara- 
tive results. 


25 

20 

|         rock 

15 

shrub         |  V3 

^r 

1p  f°  in 

10 

I            trap 

5 

Enlargement  of  Station   10 

1 

Figure  4.     Diagram  of  a  simple  linear  trap  line  with 
an  enlarged  view  of  one  station  (No.  10)  to 
show  an  example  of  trap  placements. 


Rodents  and  Insectivores 


439 


L  *       W  "P   *»," 


wmmmwzn 


-J*. Ji 


ncnrwarrzrm 


Figure  5.     Live  traps  are  efficient  for  a  wide  variety  of  small  mammals.  This  bushy-tailed  woodrat  (Neotoma 
cinerea)  was  captured  in  western  Wyoming  using  peanut  butter  as  bait. 


Trap  placement  at  each  station  may  be  biased 
by  placing  them  in  obvious  runs,  near  other  rodent 
signs,  or  in  preferred  microhabitats.  Small  rodents 
and  insectivores  have  specific  microhabitat  prefer- 
ences (e.g.,  washes,  certain  kinds  of  bushes  or  trees, 
rocky  areas,  rotten  logs,  wet  areas).  Although  all 
animals  move  around  an  area  within  certain  inherent 
limits  or  home  ranges,  it  is  best  to  place  traps  within 
or  adjacent  to  these  preferred  sites.  The  objective  is 
usually  to  catch  as  many  as  possible  of  each  species 
present  in  each  vegetative  community  during  the 
predetermined  trapping  period.  Setting  traps  2  to  4 
m  (6  to  13  ft),  or  more,  from  the  centerline  of  the 
transect  to  improve  the  chance  of  taking  additional 
specimens  is  suggested. 


Some  species  of  small  mammals,  such  as  squir- 
rels and  some  mice  and  voles,  are  largely  arboreal. 
Consequently,  they  may  not  be  captured  easily  in 
the  usual  ground  surface  trap  lines.  To  collect  these 
species,  place  live  or  small  traps  on  platforms  in 
trees  or  fasten  them  to  limbs.  Traps  may  also  be 
placed  on  the  ground  around  the  base  of  the  trees 
where  evidence  of  the  animals  is  found.  Red  tree 
voles  (Arborimus  longicaudus)  in  Oregon  are 
strictly  nocturnal  and  cannot  be  trapped.  They  must 
be  evicted  from  their  arboreal  nests,  often  as  high 
asl5m(50ft)  above  the  ground,  and  captured  by 
hand  (Maser  et  al.  1981).  Small  mammals  observed 
during  the  daytime  may  be  shot  if  the  specimens  are 
needed.  Special  Conibear  traps,  which  catch  the 
animals  around  the  neck  or  chest,  killing  them  al- 
most instantly,  are  useful  for  taking  animals  ranging 
in  size  from  squirrels  to  beaver  (Collier  1957). 


Gophers  and  moles  are  almost  entirely  fossorial. 
Gophers  may  be  captured  with  Macabee  gopher 
traps  which  are  available  at  many  hardware  stores. 
The  traps  should  be  set  in  all  gopher  tunnels  leading 
from  one  or  more  mounds  and  should  be  staked 
down  to  prevent  gophers  from  pulling  them  into 
their  burrow  complexes.  Occasionally,  gophers  are 
captured  in  snap  traps  on  the  ground  surface. 

Moles  may  be  captured  with  gopher  or  special 
mole  traps.  A  gopher  trap  set  upside  down  in  the 
mole's  burrow  is  often  effective  (Ingles  1965).  Spe- 
cial commercial  mole  traps  consisting  of  a  box-like 
housing,  which  opens  at  the  ends  and  has  an  impal- 
ing device  actuated  by  a  trigger  mechanism,  are  also 
useful  in  obtaining  specimens  for  identification 
(Giles  1969). 

Shrews  and  other  insectivores  are  rarely  cap- 
tured with  traditional  trapping  methods  for  small 
mammals.  Because  they  are  insectivorous,  they  are 
often  not  readily  attracted  to  bait  composed  of  pea- 
nut butter  or  oatmeal.  One  of  the  most  effective 
methods  for  collecting  them  is  to  use  pitfall  or  can 
traps,  which  are  also  effective  in  catching  amphibi- 
ans, reptiles,  and  other  small  mammals.  For  best  re- 
sults, the  cans  or  buckets  should  hold  more  than 
1  gallon  and  have  a  round  aperture  in  the  center, 
which  helps  reduce  the  possibility  of  escape.  How- 
ever, regular  ice  cream  or  pickle  buckets  or  cans 
will  also  serve  the  purpose  well. 

Cans  or  buckets  should  be  buried  in  the  ground 
so  that  the  tops  are  flush  with  the  substrate  (Figure 
6).  For  best  results,  the  pitfall  traps  should  be  in- 


440 


Rodents  and  Insectivores 


stalled  and  then  covered  with  a  lid  for  about  10  days 
to  allow  for  accommodation  by  the  resident  animals. 
Maximum  trapping  success  for  shrews  does  not  oc- 
cur until  the  traps  have  been  in  place  for  about  10 
days  (S.  Cross,  pers.  commun.).  Setting  up  a  trapping 
array  with  either  three  or  four  arms  will  capture 
shrews  in  moist  areas,  as  well  as  small  mammals  and 
reptiles  in  arid  areas. 

The  trapping  arrays  shown  in  Figure  6  are  set  so 
that  each  can  is  connected  by  a  drift  fence  of  wire 
mesh  or  tin,  about  0.3  m  (10  in.)  high.  The  drift 
fence  should  be  set  about  3  cm  (1  to  2  in. )  into  the 
substrate.  If  the  animals  are  believed  to  occur  in 
very  low  densities,  the  size  or  number  of  the  sample 
plots  should  be  expanded  and  the  area  of  survey 
increased. 

Because  shrews  move  about  very  rapidly  or 
perhaps  because  they  smell  meat  (or  smell  and  hear 
other  small  animals  on  which  they  feed)  in  the  bot- 
tom of  the  pit  trap,  they  frequently  drop  into  the  pit 


and  cannot  climb  back  out.  If  the  pitfalls  are  baited 
with  sardines,  earthworms,  rodent  brains  or  liver,  or 
other  meat,  more  shrews  may  be  obtained.  If  you 
cannot  examine  the  traps  on  a  daily  basis,  10  cm  (4 
in.)  of  formalin  or  AFA  (Giles  1969),  covered  by  a 
thin  layer  of  mineral  oil,  should  be  placed  in  each 
can  or  bucket  to  retard  decomposition  of  the  speci- 
mens. This  will  also  help  to  prevent  the  shrews  from 
destroying  or  eating  other  trapped  animals  in  the 
pit.  Some  trappers  fill  each  bucket  about  one-third 
full  of  water  when  they  are  activated  to  accomplish  a 
similar  purpose  and  to  help  keep  the  shrews  from 
jumping  out  (S.  Cross,  pers.  commun.). 


Shrews  are  frequently  found  near  permanent 
water  sources.  Therefore,  the  biologist  should  seek 
out  such  areas  to  check  for  their  presence.  Shrew 
sign  is  not  easy  to  detect,  but  with  some  experience, 
the  worker  can  learn  to  recognize  their  runs,  bur- 
rows, and  nests.  Leaf  litter,  matted  grass,  and  moist 
ground  are  good  habitat  signs  to  look  for. 


^ Z3  it. 

^— '     drift  fenc 


O 


drift  fenc 


o 


6 


6 


Figure  6.     Configuration  of  pitfalls  with  associated  drift  fences  to  capture  small  mammals,  lizards,  and  snakes. 
Arrays  with  either  three  or  four  drift  fences  are  effective.  Buckets  (insert)  are  placed  flush  with  the 
ground  surface  with  a  cover  supported  by  rocks  to  provide  cover  and  protection  for  the  trapped  animals. 


Rodents  and  Insectivores 


441 


Another  method  for  capturing  shrews  is  to  dig 
small  trenches  between  the  pitfalls,  instead  of  using 
drift  fences  (Figure  7).  Good  success  is  often 
achieved  by  laying  out  a  grid  of  pitfalls  (coffee  cans 
or  other  suitable  containers)  adjacent  to  small 
streams  or  other  moist  areas  (Figure  8).  Bait  is 
placed  in  the  bottom  of  the  cans  and  along  the 
trenches;  the  trenches  should  be  moistened  so  they 
are  damp.  The  trap  grid  should  be  checked  once 
during  the  night  and  again  before  sunrise.  The  traps 
should  be  closed  with  plastic  lids  during  the  day 
unless  they  are  going  to  be  visited  frequently. 

Care  should  be  taken  in  handling  live  shrews  as 
some  species  secrete  a  toxin  in  the  saliva  that  can 
make  a  bite  quite  painful.  I  recommend  using  gloves 
when  handling  them.  Unless  you  intend  to  retain 
them  as  specimens  for  identification,  they  should  not 
be  kept  in  captivity  longer  than  about  2  hours. 

Most  researchers  seem  to  prefer  a  combination 
of  snap  traps  and  live  traps  to  obtain  a  representative 
sample  of  all  species  present.  Varying  results  have 
been  reported  when  only  one  type  of  trap  (snap  vs. 
live)  has  been  used  for  sampling.  Live  traps  come 
in  different  sizes.  Common  sizes  for  small  mammals 
are  3  x  3-5  x  9  in.  and  2  x  2.5  x  6  in.  Sherman  live 
traps  are  available  from  H.B.  Sherman  Traps,  Aenon 
Creek  Industrial  Park,  Route  4,  Box  529  X-2,  Talla- 
hassee, Florida  32304;  telephone  (904)  575-0424. 

Permanently  molded  live  traps  seem  to  be  pre- 
ferred to  collapsible  ones.  However,  the  number 
to  be  set  and  the  distance  they  need  to  be  carried 
will  have  a  bearing  on  the  type  used.  Also,  rodents 
are  more  apt  to  chew  through  aluminum  live  traps 
and  escape  than  they  are  with  tin  traps.  Triggers  on 
collapsible  traps  need  to  be  sensitized  each  time 
they  are  set  to  be  sure  they  will  operate  properly. 


In-hand  examination  of  small  mammals  is  often 
required  for  their  identification.  Certain  members 
of  genera  Dipodomys,  Peromyscus,  Perognathus,  and 
Neotoma  are  extremely  difficult  to  identify  in  the 
field  and  may  require  several  individual  animals  to 
confirm  the  identifications.  Juveniles  are  particularly 
difficult  to  identify  and  may  require  use  of  compara- 
tive skulls  and  skins.  Because  of  the  lack  of  expertise 
of  most  biologists  in  identifying  subspecies  (some 
of  which  may  be  endangered  or  threatened  forms),  I 
recommend  that  arrangements  be  made  to  send 
specimens  to  qualified  museums  or  universities  for 
identification.  If  this  is  done,  be  sure  proper  meas- 
ures are  taken  to  preserve  the  specimens  during 
shipment  so  they  will  be  in  good  condition  upon 
arrival  at  the  institution.  Obviously,  this  will  not  be 
necessary  if  qualified  taxonomists  are  available  in- 
house  or  with  cooperating  agencies. 

Traps  should  be  checked  as  early  as  possible  the 
morning  following  a  night's  trapping.  Many  desert 
areas  heat  up  rapidly  and  dead  animals  will  spoil 
quickly  in  the  heat.  When  using  live  traps,  consider 
placing  a  shingle  (or  board)  on  top  of  the  trap  with 
a  rock  to  hold  the  shingle  on.  This  seems  to  improve 
efficiency  and  provides  some  shade  in  early  morning 
heat.  Also,  I  recommend  using  bedding  material  such 
as  wool  (fleece)  in  live  traps.  When  sampling  for 
diurnal  rodents,  set  the  traps  as  early  in  the  day  as 
possible.  Check  them  before  noon  and  spring  them. 
Reset  them  in  the  afternoon  and  rerun  the  line  be- 
fore dark.  Remove  animals  and  spring  the  traps. 

The  terminal  ends  of  each  trap  line  should  be 
marked  with  stakes,  flagging,  or  other  visible  mate- 
rial. Permanent  stakes  are  desirable  for  future  refer- 
ence; the  locations  of  these  markers  should  be 
marked  on  maps  so  they  may  be  relocated  for  future 
monitoring  studies.  In  forested  areas  or  in  heavy 


ground  level 

X 


K7 

trench 

1  -  2'  deep 


:X    coffee  can 


Figure  7.     Diagram  of  pitfall  and  associated  trenches  for  catching  shrews.  The  trenches  should  be  slightly 
moistened. 


442 


Rodents  and  Insectivores 


brush,  you  may  wish  to  run  a  cord  through  the  habi- 
tat and  attach  tags  at  appropriate  places  to  facilitate 
finding  the  traps.  Strips  of  flagging  material  may  be 
tied  to  branches  for  identifying  trapping  stations. 

Provide  a  thorough  description  of  the  site 
where  the  trap  line  has  been  placed.  Include  the 
following  items: 

•  abundance  of  trees,  shrubs,  and  grasses 

•  canopy  coverage  and  height  of  the  plants 

•  serai  stage  of  the  plant  community 

•  litter  depth  and  quantity;  disturbance  by  live- 
stock, fire,  or  man 

•  presence  of  logs,  debris,  rocky  cover 

•  size  of  trees  and  species  in  forest  situations 

Include  a  brief  description  of  topography  (e.g.,  slope, 
elevation,  relation  to  drainage  system,  nature  of  the 
soil,  aspect).  For  future  monitoring  studies,  photo- 
graphs of  habitat  sites  are  helpful.  Describe  special 


microhabitats  where  traps  might  have  been  placed 
within  the  general  habitat  type  and  note  differences 
in  species  caught  there.  Make  special  note  of  the 
weather  conditions  at  the  time  the  traps  were  in  op- 
eration and  just  preceding  the  trapping  effort. 


Variability  is  considerable  in  the  daily  and  sea- 
sonal behavior  of  rodents  and  insectivores,  including 
the  time  species  are  aboveground  and  available  for 
trapping.  Some  hibernate  or  estivate,  whereas  others 
spend  parts  of  the  year  underground  where  they 
are  active.  Some  species  spend  most  or  all  of  their 
time  in  tree  canopies  during  spring  and  summer, 
then  spend  time  on  the  forest  floor  in  late  fall  and 
winter. 


Therefore,  determine  from  the  literature  or  local 
experts  when  various  species  are  expected  to  be 
most  active  aboveground  and  concentrate  on  trap- 
ping during  those  periods.  Trapping  for  species  oc- 
currence and  relative  density  during  all  four  seasons 
is  desirable  to  obtain  a  more  complete  picture  of 
total  species  using  an  area. 


cans 


-=?fe 


Figure  8.     Grid  pattern  of  pitfalls  and  trenches  for  collecting  shrews  along  a  stream.  Trenches  are  intercon- 
nected with  pitfalls  to  guide  shrews  into  the  pits. 


Rodents  and  Insectivores 


443 


Obviously,  it  will  be  difficult  to  obtain  a  true 
relative  density  of  all  species  present,  from  gophers 
to  tree  squirrels.  They  will  not  have  the  same  trappa- 
bility  nor  be  taken  by  the  same  kind  of  traps.  For 
example,  if  you  do  not  set  gopher  or  mole  traps 
where  signs  of  these  species  are  present  along  the 
trapping  transect,  you  may  not  get  any  in  your  sam- 
ple, even  though  they  may  be  fairly  numerous.  Spe- 
cial traps  or  techniques,  then,  will  be  necessary  if 
you  want  a  true  representation  of  all  species  present 
in  a  given  habitat  type. 

Population  Estimates.  Three  methods  for 
estimating  numbers  of  animals  will  be  discussed:  ( 1 ) 
the  strip  census,  (2)  Lincoln  Index  or  mark  and 
recapture,  and  (  3  )  probability  of  capture  method. 
Any  of  these  methods  will  give  satisfactory  results 
when  used  for  monitoring  the  same  populations 
during  some  future  point  in  time,  providing  all 
assumptions  are  met  to  the  same  degree  as  in  the 
original  sampling. 

(1)  Strip  Census.  One  method  for  estimating 
the  number  of  medium-sized  rodents,  such  as  tree 
squirrels  or  ground  squirrels,  within  a  habitat  type  is 
to  count  the  number  of  animals  observed  within  a 
strip,  transect,  and  then  apply  the  results  to  the  gen- 
eral area.  The  area  censused  should  be  representa- 
tive of  the  entire  habitat  type.  Counts  of  several 
similar  areas  or  several  counts  of  the  same  area  pro- 
vide numbers  needed  for  calculation  of  an  estimate 
of  total  numbers  for  the  species  being  surveyed. 
Detailed  versions  of  the  strip  census  method  can  be 
found  in  Anderson  et  al.  1976  and  Burnham  et  al. 
1980. 


The  strip  census  method  entails  following  a 
straight  transect  line  or  series  of  straight  line  seg- 
ments (Anderson  et  al.  1976),  counting  the  animals 
observed,  and  recording  the  right  angle  distances 
from  the  transect  line  to  the  animals  observed.  The 
requirement  to  follow  a  straight  line  is  usually  com- 
promised in  the  field  because  of  physical  obstacles 
and  the  near-impossibility  of  traveling  a  straight  line 
and  searching  for  animals  at  the  same  time.  Marking 
the  transect  line  in  some  manner  will  improve  the 
ability  to  maintain  a  fairly  straight  line.  This  would 
be  done  a  few  days  before  making  a  count. 

Four  assumptions  are  critical  to  achieve  good 
results  (Anderson  et  al.  1976): 


(  1 )    Animals  directly  on  the  line  are  seen  with 
probability  of  one  (all  seen). 


(2)    Animals  are  pinpointed  at  the  initial  sighting 
position  (they  do  not  move  before  being 
sighted )  and  none  are  counted  twice. 


(3)    There  are  no  measurement  errors  (distances 
are  measured  exactly). 


(4)   Animal  sightings  are  independent  events. 

Obviously,  one  or  more  of  these  assumptions 
may  not  be  entirely  met.  Ground  squirrels  or  other 
burrowing  animals  may  duck  into  a  hole  on  or  near 
the  transect  line  and  not  be  seen  at  all,  or  may  run 
some  distance  before  being  observed.  Many  animals 
respond  to  intruders  by  shying  away  or  running 
some  distance  under  cover  before  flushing  into  the 
open.  This  tends  to  exaggerate  the  right  angle  dis- 
tance from  the  transect  line  to  the  animal  when  first 
observed.  Some  animals  run  ahead  of  the  observer 
and,  after  being  counted  once,  may  be  counted  again 
if  they  continue  to  run  ahead.  Also,  the  flushing  of 
one  animal  is  not  supposed  to  cause  another  animal 
to  flush,  but  it  sometimes  does.  All  these  situations 
tend  to  violate  the  assumptions.  Nevertheless,  if  the 
same  procedures  are  used  each  time  the  population 
is  sampled,  comparison  of  results  will  provide  satis- 
factory information  on  trends.  As  in  all  methods,  a 
replication  along  the  same  transect  a  few  days  later 
or  in  a  similar  area  will  add  credibility  to  the  results 
and  provide  a  means  to  calculate  standard  deviation 
and  variance. 

See  Figure  9  for  a  diagrammatic  display  of  a  line 
transect  and  the  various  measurements  required. 

(2)  Lincoln  Index  or  Mark-Recapture 
Method.  The  first  mark-recapture  method  for  terres- 
trial animals  was  known  as  the  Lincoln  Index  and  it 
is  still  widely  used  (Seber  1973).  In  using  the  Lin- 
coln Index,  part  of  the  animals  in  a  population  are 
live-trapped,  marked  in  some  manner,  and  then  re- 
leased back  into  the  population.  Samples  are  next 
captured  to  observe  the  proportion  marked  in  the 
entire  population,  which  now  consists  of  a  known 
number  of  marked  animals  along  with  an  unknown 
number  of  unmarked.  An  estimate  of  the  total  num- 
ber is  computed  by  dividing  the  total  number 
marked  in  the  population  by  the  proportion  marked 
in  the  samples,  under  the  assumption  that  the  sam- 
ples will  estimate  closely  the  proportion  marked 
throughout  the  entire  population. 

There  are  three  general  methods  for  marking 
captured  animals:  mutilation,  tagging,  and  coloring 
(Mosby  1963). 

Mutilation.     Forms  of  mutilation  include  toe- 
clipping,  ear-cropping,  hole-punching,  fur-clipping, 
tail-notching  (beaver),  and  branding.  Any  of  these 
forms  of  marking  may  receive  public  criticism  in  that 
they  maim  the  animal  to  some  extent  and  possibly 
affect  its  behavior  or  survival.  The  main  advantages 
of  mutilation  are  that  marks  may  be  readily  applied 
with  a  minimum  of  equipment  and  the  marks,  in 


444 


Rodents  and  Insectivores 


some  instances,  may  be  identified  at  a  distance.  Sys- 
tems for  marking  small  mammals  by  toe-clipping  and 
ear-punching  have  been  described  by  several  work- 
ers (Blair  1941;  Reynolds  1945;  Sanderson  1961). 
Two  methods  for  numbering  systems  used  in  toe- 
clipping  are  shown  in  Figure  10.  Toe-clipping  is  a 
widely  used  technique  for  both  small  mammals  and 
reptiles.  A  small  pair  of  sharp  scissors  or  toenail  clip- 
pers works  very  efficiently.  Only  the  distal  portion 
of  the  animal's  toe  is  removed.  This  is  done  system- 
atically and  recorded  as  a  unique  mark  to  identify 
a  specific  animal. 

Shearing,  clipping,  or  burning  off  the  tips  of  the 
hairs  can  be  used  when  the  hair-base,  or  underfur, 
is  of  a  contrasting  color,  as  it  often  is,  such  as  in 
newborn  mice  (Svihla  1934)  and  chipmunks  (Euta- 
mias  sp.;  Yerger  1953)-  This  method  is  useful  only 
until  the  next  molt  but  has  three  advantages:  it  is 
easy  to  apply,  painless  to  the  animal,  and  it  consti- 
tutes a  good  mark  for  field  identification.  On  Norway 
rats,  Chitty  and  Shorten  (1946)  used  a  number 
drawn  on  the  pelage  with  a  depilatory.  Such  marks 
last  2  to  3  weeks,  normally  long  enough  for  sampling 
and  making  population  estimates. 

In  the  beaver,  the  flat,  hairless  tail  has  been 
marked  by  branding  (Bradt  1938)  and  by  cutting 


notches.  Bradt  found  straight-line  brands  on  the  up- 
per surface  of  the  tail  recognizable  for  at  least  4 
years.  Cuts  into  the  edge  of  the  tail  presumably  last 
for  life  (Mosby  1963). 

Tagging.     Tagging  involves  the  attachment  to 
the  animal  of  a  piece  of  metal  or  plastic  on  which  is 
stamped  an  individual  number.  Advantages  are  that 
they  are  easy  to  affix  and  to  see.  The  tag  should  be 
as  light  as  possible  and  should  be  placed  so  that  it 
does  not  pinch  the  animal.  It  is  desirable  to  place 
tags  low  on  the  ear,  where  the  cartilage  is  heavy,  and 
on  the  inner  edge  where  there  is  greater  protection 
of  the  tag  (Mosby  1963).  In  some  mammals  where 
the  ears  are  very  small  or  there  may  be  high  losses 
of  tags  through  freezing  or  tearing  out,  tags  may 
be  placed  elsewhere. 

Muskrats  have  been  tagged  by  slipping  a  5  by  24 
mm  aluminum  tag  through  two  slits  in  the  skin  of 
the  back  (Errington  and  Errington  1937)  and  is  most 
successful  with  muskrats  not  over  2  months  old, 
due  to  the  superior  regrowth  of  blood  vessels  across 
the  strip  of  skin  between  the  two  incisions  in  young 
animals  (Errington  1944).  This  method  might  cause 
infection  if  the  water  they  live  in  is  polluted  (Takos 
1943). 


Figure  9.     The  unit  to  be  censused  has  area  A.  The  transect  length  is  L  and  the  transect  width  is  2W.  Z  is  the 
position  of  an  observer  when  an  animal  or  object  is  detected  at  X.  P  is  the  point  on  the  line  which  is 
perpendicular  to  the  animal.  The  sighting  distance  is  r„  the  sighting  angle  is  Oj,  and  the  perpendicular 
distance  from  the  object  to  the  center  line  of  the  transect  is  y;.  W  can  be  left  unspecified  or  can  be  a 
fixed,  finite  constant  (adapted  from  Anderson  et  al.  1976). 


Rodents  and  Insectivores 


445 


Figure  10.     Illustration  of  toe-clipping  procedures.  In  the  left  diagram  only  the  first  four  toes  are  considered 
for  clipping,  counting  from  the  outer  toe.  A  numbering  system  used  in  toe-clipping  squirrels  is  shown 
(from  Mosby  1963)  in  the  diagram  on  the  right. 


Small  bird  bands  have  been  used  as  loose-fitting 
bracelets  around  the  hind  legs  of  small  mammals 
(Chitty  1937).  Use  of  such  bands  is  more  time-con- 
suming than  toe-clipping,  however,  and  on  some 
species  caused  irritation,  swelling,  skin-puncture,  and 
infection  (Takos  1943).  Some  of  these  difficulties 
may  be  overcome  by  encircling  the  band  around  the 
Achilles  tendon  rather  than  the  entire  leg.  Smaller 
bands  have  been  clamped  loosely  around  a  hind  toe 
on  squirrels  and  are  kept  in  place  by  the  toe  pad 
(Cooley  1948). 

A  different  type  of  marker,  used  on  muskrats 
and  other  medium-sized  rodents,  consists  of  a  water- 
proof plastic  button,  threaded  through  the  loose 
skin  of  the  back  with  a  flexible  steel  needle  and  se- 
cured with  another  button.  This  plastic  marker  has 
the  advantage  of  being  visible  from  a  distance  on 
a  swimming  muskrat  (Hensley  and  Twining  1946; 
Mosby  1963). 

Coloring.     Pelage  coloring  with  commercial  fur 
dyes  is  another  successful  method  of  marking  ani- 
mals and  lasts  until  the  next  molt.  Most  of  these  dyes 
are  poisonous  and  must  be  used  with  care.  Two 
common  dyes  are  Nyanzol  4R  (reddish-brown)  and 
Nyanzol  D  (black).  Both  may  be  fixed  after  applica- 
tion with  hydrogen  peroxide  ( Evans  and  Holdenried 


1943)  or  mixed  with  hydrogen  peroxide  before 
application  (Fitzwater  1943).  Dyes  are  useful  for 
mammals  of  light  pelage  but  the  range  of  colors  is 
small.  The  series  can  be  increased  by  applying  spots 
to  different  parts  of  the  body  (Mosby  1963).  Table  1 
shows  a  few  applications  of  these  dyes  to  small 
mammals. 

(3)  Probability  of  Capture  Method  (Re- 
moval Trapping).  Where  animals  are  removed 
from  a  population  as  captured,  it  is  necessary  to  use 
a  different  approach  from  the  mark-recapture 
method.  However,  data  from  the  mark  and  recapture 
samples  may  be  examined  by  either  the  Lincoln  In- 
dex method  or  by  the  removal  trapping  method, 
by  using  only  the  records  for  animals  trapped  the 
first  time. 

The  number  of  animals  captured  during  any 
trapping  period  may  be  viewed  as  the  product  of 
two  quantities,  the  first  being  the  probability  of  cap- 
ture and  the  second  the  number  of  animals  present 
at  the  beginning  of  the  period  (Hayne  1949a).  The 
probability  of  capture  is  assumed  to  be  constant, 
describing  the  hazard  in  which  any  animal  stands  in 
relation  to  capture  in  the  set  of  traps  during  one 
period.  The  number  of  animals  present  at  the  begin- 
ning of  each  trapping  period  is  assumed  to  be  the 


446 


Rodents  and  Insectivores 


original  population  minus  the  number  of  animals 
previously  captured.  Any  great  departure  from  these 
assumptions  seems  to  invalidate  this  method  for 
estimating  the  population. 


Monitoring  Small  Mammal  Populations 

Monitoring  small  mammals  is  simply  the  check- 
ing of  the  status  of  species,  relative  density,  or  num- 
bers at  some  point  before  an  initial  survey  or 
inventory  of  the  population.  Baseline  information  is 
required  to  determine  how  species  and  numbers  are 
changing  over  time.  The  objectives  for  monitoring 
small  mammals  must  be  determined  by  the  biologist, 
e.g.,  to  determine  trends  in  populations  of  endan- 
gered, threatened,  or  sensitive  species;  to  determine 
trends  in  species  serving  as  important  prey  bases; 
or  to  determine  trends  in  numbers  of  other  species 
for  economic,  scientific,  or  other  reasons.  This  might 
involve  monitoring  only  one  species,  such  as  beaver, 
to  maintain  their  numbers  within  prescribed  limits 
to  prevent  damage  to  vegetation  or  scenic  resources, 
or  it  might  be  to  monitor  several  small  rodent  spe- 
cies that  are  essential  to  the  continued  production  of 
an  endangered  or  sensitive  species  or  raptor.  Rea- 
sons will  vary  with  locality.  Time  intervals  between 
checks  of  species  and  populations  will  vary  accord- 
ing to  productivity,  initial  status,  importance,  and 
possible  decimating  factors  for  the  species  involved. 
Susceptibility  to  cycles  must  also  be  considered. 


DISCUSSION  AND  SUMMARY 

Larger  rodents  may  often  be  surveyed  visually 
or  by  their  signs,  but  small  mammals,  such  as  mice 
and  shrews,  are  difficult  to  observe  directly.  Noctur- 
nality,  protection  by  overhanging  vegetation,  and 
retirement  to  nest  or  burrow  upon  detection  of  an 
approaching  human  contribute  to  this  difficulty  of 


observation.  For  this  reason  biologists  use  a  variety 
of  sampling  devices  and  techniques.  In  previous  sec- 
tions I  have  discussed  general  kinds  of  habitats  occu- 
pied by  different  groups  of  small  mammals  and  some 
of  the  methods  for  determining  their  presence,  rela- 
tive density,  and  estimated  numbers.  Here  I  will 
discuss  some  of  the  problems  involved  in  rodent  and 
insectivore  studies  that  should  be  kept  in  mind  when 
designing  an  inventory  or  monitoring  study  and 
some  of  the  limitations  of  the  data  derived.  Anyone 
can  set  a  trap  line,  but  only  a  person  with  some  un- 
derstanding of  rodent  or  insectivore  ecology  can  put 
the  data  derived  in  their  proper  perspective. 

Sampling  Assumptions 

Species  Occurrence  and  Relative  Density.  In 

setting  out  single  trap  lines  or  a  series  of  parallel 
lines,  it  might  be  assumed  that  most  of  the  individual 
animals  and  species  within  easy  travel  range  of  the 
traps  will  be  equally  susceptible  to  being  caught. 
This  is  not  true.  First,  using  bait  in  the  traps 
immediately  biases  the  catch.  Some  species  (and 
probably  some  individuals)  will  be  attracted  to  this 
bait  much  more  quickly  than  others.  Animals,  such 
as  deer  mice,  may  be  drawn  from  considerable 
distance  by  the  smell  of  the  bait,  whereas  others  are 
only  mildly  attracted.  Therefore,  in  many  localities, 
a  preponderance  of  deer  mice  are  caught  during 
early  evening.  In  fact,  most  traps  may  become 
occupied  by  deer  mice,  excluding  other  species  for 
the  remainder  of  the  night  unless  the  traps  are 
emptied  and  reset.  This  is  the  reason  for  checking 
and  resetting  traps  each  hour  until  midnight.  By  so 
doing,  most  of  the  deer  mice  or  other  easily  caught 
species  will  be  caught  early  and  leave  the  traps 
ready  to  capture  more  secretive,  less  numerous,  or 
more  trap-shy  animals.  By  following  this  procedure,  a 
more  accurate  relative  density  among  species  can 
be  obtained.  Otherwise,  the  most  easily  trapped 
species  will  appear  the  most  numerous  in  the 


Table  1.     Techniques  for  coloring  the  pelage  of  mammals  (adapted  from  Mosby  1963). 


Species 

Coloring  Agent 

Special  Techniques 

Duration 
of  Color 

Authority 

Pocket 
gopher 

Human  black  hair  dye  with  oil 
base:  apply  equal  parts 
3%  hydrogen  peroxide  plus 
granulated  soap  until  liquid  is 
thick. 

Ammonium  hydroxide  (4%) — 
1  part  to  2  parts  3%  hydrogen 
peroxide  plus  soap  as  above. 

Recommended  for  light- 
pelaged  mammals 

Recommended  for  dark- 
pelaged  mammals 

Until 
molt 

Until 
molt 

Morejohn  and 
Howard  1956 

Morejohn  and 
Howard  1956 

Squirrel 

Picric  acid  in  5%  formalin. 

Nyanzol  A:  20  gm  per  L  of 
water-hydrogen  peroxide 
mixture  in  ratio  of  2:1. 

Use  on  light  pelage 

Ring  animal's  body  in  broad 
bands 

Until 
molt 

Fitzwater  1943 
Fitzwater  1943 

Rodents  and  Insectivores 


447 


composition,  whether  it  is  in  actuality  or  not. 
Regardless  of  the  procedure  you  use  in  checking 
traps,  do  it  the  same  way  each  time  in  order  to  have 
comparative  information. 

When  monitoring  populations,  be  sure  all  envi- 
ronmental conditions  and  biotic  factors  are  about 
the  same  as  for  prior  sampling.  For  example,  species 
will  vary  as  to  time  of  reproductive  peaks,  times  of 
greatest  seasonal  behavior,  and  months  of  greatest 
dispersal.  Results  may  be  erroneous  if  sampling  is 
done  before  the  reproductive  peak  one  year  and 
after  the  peak  when  sampling  two  years  later. 

In  spring  and  early  summer,  a  preponderance  of 
males  of  some  species  may  be  caught  when  they 
are  especially  active  in  seeking  mates.  Times  vary 
with  different  species.  There  is  always  a  reason  for  a 
preponderance  of  one  sex  or  age  group  in  a  catch. 
Some  knowledge  about  rodent  or  insectivore  ecol- 
ogy will  usually  reveal  the  reason.  In  making  com- 
parisons of  numbers  or  species  among  years,  be 
aware  that  many  factors  may  be  influencing  the  re- 
sults. The  reasons  may  be  more  important  than  the 
numbers  themselves. 


Of  all  factors  affecting  capture  probability,  time 
is  the  most  easily  controlled.  The  biologist  can  select 
the  season  of  the  year  when  studies  are  to  be  con- 
ducted, the  length  of  the  trapping  period,  and  the 
time  of  day  when  trapping  is  to  be  done.  The  objec- 
tive is  to  reduce  variation  in  capture  probabilities 
over  time.  For  usable  comparisons  equal  effort 
should  be  expended  on  each  sampling  occasion.  The 
number  of  traps  should  be  the  same  each  time,  trap- 
ping should  be  done  at  the  same  time  of  day  and,  if 
bait  is  used,  the  type  and  amount  should  be  the 
same  on  all  occasions. 


Weather  is  an  extremely  important  factor  in 
small  mammal  trapping  (or  any  kind  of  sampling). 
Animals  vary  greatly  in  their  trappability,  behavior, 
and  activity  depending  on  such  factors  as  moonlight, 
cloudiness,  precipitation,  high  winds,  and  tempera- 
ture. Where  seasonal  weather  conditions  have  pro- 
duced a  poor  seed  crop,  animals  will  likely  be  more 
attracted  to  baits.  It  is  also  important  to  be  aware 
of  possible  damage  to  habitats  caused  by  storms  or 
flooding  before  the  sampling  period.  Many  animals  in 
the  study  area  might  have  been  lost  by  drowning  or 
emigrated  to  safer  habitat  elsewhere.  Drought  can 
similarly  reduce  populations  by  lowering  productiv- 
ity. On  the  other  hand,  unusually  good  survival  con- 
ditions may  produce  much  higher  than  usual 
numbers  during  the  sampling  period.  Severe  storms 
and  low  temperatures  or  dry  conditions  that  inhibit 
seed  production  are  among  the  most  important  fac- 
tors to  consider  when  evaluating  species  changes, 
relative  densities,  or  population  fluctuations. 


Relative  rather  than  absolute  density  has  been 
assumed  to  suffice  for  solving  many  biological  prob- 
lems, but  the  area  factor  in  density  has  been  ignored. 
If  all  procedures  are  maintained  constant,  including 
the  time  over  which  samples  are  taken,  it  is  assumed 
that  the  relative  densities  obtained  will  suffice  for 
comparing  different  habitats  or  the  same  habitat  at 
different  times.  However,  these  relative  densities  can 
be  misleading.  For  example,  on  the  trap  lines  run 
by  the  North  American  Census  of  Small  Mammals 
from  1948  to  1951,  1,901  male  and  1,521  female 
deer  mice  were  trapped.  The  conclusion  might  be 
that  males  in  these  areas  were  25%  more  abundant 
than  females,  but  this  is  probably  not  true.  It  is  more 
likely  that  males  have  a  larger  home  range  than  do 
females  and,  therefore,  more  males  than  females  are 
exposed  to  the  trap  lines  (Stickel  1948). 

Similar  inaccuracies  of  assumptions  that  relative 
densities  are  proportionate  to  true  densities  also 
apply  to  comparisons  of  the  density  of  different  gen- 
era or  of  the  same  genus  in  different  habitats.  Until 
one  knows  more  about  the  influence  exerted  upon 
home  ranges  by  species,  sex,  and  habitat,  it  is  well  to 
use  caution  in  drawing  conclusions  from  relative 
densities  other  than  for  those  densities  that  concern 
a  single  species  and  sex  from  different  times  in  the 
same  habitat. 

Many  small  mammal  species  have  home  ranges, 
whereas  other  species  seem  to  wander  continually 
with  no  defined  center  of  activity  (Calhoun  and 
Casby  1958).  When  using  the  Lincoln  Index  method, 
one  assumes  that  the  marked  animals  when  released 
into  the  population  will  become  distributed  through- 
out and  that  further  samples  will  take  these  marked 
animals  with  no  greater  or  lesser  probability  of  cap- 
ture than  the  unmarked  animals  (Hayne  1949a).  If 
the  marked  animals  return  to  the  use  of  their  origi- 
nal home  ranges,  they  may  or  may  not  present  a 
random  distribution  within  the  population  for  subse- 
quent sampling.  In  fact,  White  et  al.  (1982)  stated 
frankly  that  "there  is  simply  no  basis  for  thinking 
that  samples  are  drawn  randomly  in  capture  studies 
of  animal  populations." 

Another  important  assumption  is  that  no  signifi- 
cant replacement  of  the  population  by  unmarked 
animals  will  occur  between  the  marking  and  the 
sampling  periods.  Mortality  does  not  present  a  signif- 
icant problem,  so  long  as  it  happens  in  equal  propor- 
tion among  marked  and  unmarked  animals  and  does 
not  result  in  replacement  of  the  dead  animals  by 
unmarked  individuals  from  other  sources  (Hayne 
1949a).  Movement  out  of  a  population  is  treated  the 
same  as  mortality;  movement  into  it,  as  part  of  re- 
cruitment (Schemnitz  1980).  If  either  mortality  or 
recruitment  is  significant  during  the  initial  inventory 
or  monitoring  period,  subsequent  sampling  will  be 
difficult  to  analyze  correctly  without  conducting 
time-intensive  studies  (White  et  al.  1982). 


448 


Rodents  and  Insectivores 


Turnover  is  rapid  in  many  small  mammal  popu- 
lations. Individuals  are  replaced  constantly,  either 
by  dispersal  with  replacement  by  individuals  from 
adjacent  areas  or  by  natural  mortality  with  replace- 
ment by  younger  individuals  growing  to  adult  status. 
If  too  much  time  lapses  between  the  preliminary 
marking  period  and  the  subsequent  sampling  period, 
an  overestimation  of  the  size  of  the  population  that 
was  present  during  the  marking  period  inevitably 
results.  In  either  this  method  or  methods  for  estimat- 
ing density  (Scott  et  al.  1978;  O'Farrell  et  al.  1977; 
White  et  al.  1982),  the  importance  of  being  able 
to  assume  no  recruitment  or  mortality  is  empha- 
sized. It  is  important  not  to  wait  more  than  a  few 
days  between  the  marking  period  and  the  subse- 
quent recapture  period. 


situation  will  tend  to  produce  underestimates  of  the 
population  (Hayne  1949a).  In  other  situations  ani- 
mals learn  to  avoid  traps.  This  may  happen  to  ani- 
mals that  have  a  toe  clipped  or  have  some  other 
unpleasant  experience  during  or  after  capture  in  a 
live  trap  or  after  a  bare  escape  from  a  snap  trap. 
After  first  capture,  an  animal's  capture  probability  on 
subsequent  capture  occasions  changes,  often  greatly 
(Getz  1961;  Bailey  1969;  White  et  al.  1982).  There- 
fore, marked  animals  probably  do  not  have  the  same 
catchability  as  previously  uncaught  animals.  Some 
factors  affecting  small  mammal  responses  to  traps  in- 
clude weather,  season,  individual  inquisitiveness, 
population  density,  food  availability,  type  of  traps 
used,  social  dominance,  sex,  and  activity  patterns 
(Sheppe  1972;  Sarrazin  and  Bider  1973;  Summerlin 
and  Wolfe  1973). 


General  Considerations  for  all  Small  Mammal 
Studies.  Sufficient  time  and  money  for  meaningful 
studies  will  continue  to  be  problems  facing  most 
biologists.  Good  judgment  must  be  exercised  to 
obtain  the  best  information  possible  with  limited 
funding.  Most  small  mammal  studies  in  the  past  have 
been  classified  by  Hayne  ( 1976)  as  "descriptive 
studies"  and  discuss  animals  found  at  the  time  and 
place  for  each  particular  study  but  have  little 
application  to  other  times  and  areas  because  they  are 
seldom  replicated.  Hayne  (1976)  emphasized  that  a 
sample  size  of  even  just  two  is  very  much  better 
than  a  sample  size  of  one.  He  did  not  imply  that  a 
sample  size  of  two  is  adequate  for  precision. 
Obtaining  precision  in  nature  can  be  very  costly,  but 
the  value  of  even  one  replication  of  a  study  can  shed 
great  light  on  the  results  of  the  first  study.  Usually, 
the  results  of  two  samples  can  be  lumped  or 
averaged  to  give  a  better  picture  of  the  real 
situation.  Unreplicated  studies  can  lead  to 
generalizations  and  unrestrained  speculations;  even 
one  replication  of  a  sample  in  a  comparable  habitat 
type  should  put  some  limitations  on  how  the  results 
are  interpreted. 

One  should  not  be  intimidated  by  criticism  that 
two  studies  in  the  same  kind  of  habitat  in  different 
areas  cannot  be  compared  because  they  are  not 
identical  habitats.  There  is  no  such  thing  as  any  two 
areas  being  identical;  there  will  always  be  some  dif- 
ferences, but  if  two  similar  areas  are  selected  care- 
fully, either  for  replication  of  an  initial  study  or  for 
comparison  with  other  years,  the  results  should  be 
valid  for  comparisons. 

The  assumption  of  equal  trappability  is  probably 
false  in  most  situations,  even  though  it  is  a  necessary 
assumption  in  most  sophisticated  statistical  proce- 
dures (White  et  al.  1982).  In  some  studies,  a  dispro- 
portionate number  of  animals  are  captured,  which 
had  a  high  probability  of  capture;  these  may  be 
termed  as  "repeaters"  or  described  as  having  a  "trap 
habit."  When  using  the  Lincoln  Index  method,  this 


Considerable  differences  occur  in  the  numbers 
of  animals  trapped  in  different  types  of  traps;  also, 
different  sizes  of  the  same  type  of  trap  may  catch  dif- 
ferent numbers  of  animals.  It  would  be  advisable, 
therefore,  to  use  the  same  kind  of  traps  and  in  the 
same  sequence  (live  vs.  snap  traps)  for  obtaining 
samples  that  are  to  be  compared  between  years,  sea- 
sons, or  habitat  types  (Schemnitz  1980). 


There  have  been  few  studies  to  check  popula- 
tion estimates  against  a  known  population.  Most 
present-day  techniques  of  estimation  have  such  low 
precision  that  only  large  changes  of  the  population 
or  a  large  influence  of  a  factor  can  be  detected 
(Schemnitz  1980).  Fortunately,  low  precision  esti- 
mates may  be  adequate  for  management  purposes  on 
small  mammals,  because  even  a  method  of  low  preci- 
sion can  detect  a  large  effect  in  a  population.  How- 
ever, if  better  precision  is  required  to  meet  a  specific 
objective,  procedures  may  be  found  in  the  literature 
(Hayne  1949a;  DeLury  1951;  Leslie  1952;  Fredin 
1954;  Davis  1956;  Nixon  et  al.  1967;  Eberhardt 
1969;  Burnham  1972;  Hansson  1974;  Jensen  1975; 
O'Farrell  et  al.  1977;  Burnham  and  Overton  1979; 
Cormack  1981;  Seber  1982;  White  et  al.  1982). 

In  the  typical  mark-recapture  study,  a  main  ob- 
jective is  to  estimate  the  population  size  (N).  Nei- 
ther the  true  value  of  (N)  nor  the  correct 
assumptions  to  make  about  capture  probabilities  are 
known  (White  et  al.  1982).  Delving  into  the  litera- 
ture one  will  find  dozens  of  published  estimators. 
The  biologists  should  select  one  and  proceed  with 
the  calculations.  However,  they  rarely  can  test  the 
assumptions  they  have  made  in  collecting  samples, 
nor  do  they  take  time  to  estimate  the  sampling  vari- 
ance of  the  population  estimate.  In  some  instances 
assumptions  hold  for  one  species  but  not  for  another 
living  in  the  same  habitat.  Unfortunately,  the  devel- 
opment of  the  mathematics  and  statistics  has  pro- 
ceeded far  more  rapidly  than  has  the  testing  of  the 


Rodents  and  Insectivores 


449 


V 


assumptions  usually  used  for  statistical  tests  (Schem- 
nitz  1980).  The  task  of  testing  assumptions  about 
small  mammal  populations  is  tremendous  and  be- 
yond the  scope  of  most  investigations.  Because  there 
are  numerous  published  estimators,  biologists  using 
different  estimators  can  get  different  estimates  with 
the  same  data.  It  is  difficult  to  decide  which  esti- 
mator will  give  the  most  accurate  results  for  any 
given  situation.  Another  difficulty  in  analyzing  differ- 
ences in  species,  relative  densities,  or  estimated 
numbers  during  monitoring  studies  over  a  sequence 
of  years  is  that  of  determining  whether  apparent 
differences  are  real  or  only  variations  that  might  be 
expected  among  any  set  of  samples. 

At  times,  the  biologist  may  wish  to  determine  an 
estimate  of  density  of  small  mammals  in  a  particular 
habitat  type.  Density  is  defined  as  the  number  of 
animals  per  unit  area.  Density  estimation  extends 
population  size  estimation  to  include  an  estimate  of 
the  area  to  which  the  population  estimate  relates. 
Density  estimates  are  usually  derived  with  the  use  of 
a  trapping  grid.  However,  density  estimation  is  not 
as  simple  as  dividing  the  estimated  population  of 
animals  by  the  area  of  the  trapping  grid,  because  of 
the  phenomenon  known  as  edge  effect  (White  et  al. 
1982).  For  example,  animals  at  the  edge  of  a  grid 
will  not  spend  all  of  their  time  on  the  grid,  because 
the  grid  area  contains  only  a  part  of  their  home 
ranges.  Thus,  the  effective  area  trapped  is  somewhat 
larger  than  the  grid  area,  because  the  area  to  which 
the  population  estimate  applies  includes  the  entire 
home  ranges  of  such  animals. 

A  boundary  strip  should  be  included  as  part  of 
the  area  occupied  by  the  population.  Dice  (1941) 
suggested  that  the  width  of  the  boundary  strip  be 
taken  as  one-half  the  average  diameter  of  the  home 
range  of  the  species  in  question.  Other  authors  have 
made  additional  recommendations  (Hayne  1949b; 
Stickel  1954;  Tanaka  1972;  Otis  et  al.  1978)  but  all 
approaches  are  subject  to  difficulties.  For  example, 
the  estimates  depend  on  trap  spacing  and  numbers 
of  recaptures  (White  et  al.  1982).  A  second  ap- 
proach involves  the  use  of  assessment  lines  (M.H. 
Smith  et  al.  1971;  R.  Smith  et  al.  1975;  H.D.  Smith  et 
al.  1972;  Scott  et  al.  1978;  O'Farrell  et  al.  1977). 
Although  this  approach  has  produced  good  results, 
the  method  can  become  complex  and  is  heavily 
dependent  on  the  design  of  the  trap  layout.  How- 
ever, it  can  be  demonstrated  that  the  importance  of 
the  boundary  strip  decreases  as  the  size  of  the  grid 


increases.  Placing  traps  in  a  rectangular  configuration 
or  in  a  series  of  parallel  lines  reduces  this  edge  effect 
even  when  using  the  simpler  methods,  such  as  trap 
lines  set  out  to  determine  relative  density  or  species 
occurrence. 

Biologists  desiring  to  calculate  densities  of  one 
or  more  species  of  small  mammals  should  refer  to 
Literature  Cited  for  needed  statistical  procedures. 
White  et  al.  (1982)  stated  that: 

"computer  programs  have  become  essential  in 
the  analysis  of  capture-recapture  and  removal  stud- 
ies. The  iterative  nature  of  many  estimators  of  popu- 
lation size  under  closure  makes  them  nearly 
impossible  to  compute  with  a  hand  calculator,  and 
the  testing  and  model  selection  procedures  are  tedi- 
ous. The  notation  and  algebra  for  the  estimators  of 
open-population  parameters  are  difficult,  and  re- 
cently developed  models  do  not  have  closed-form 
estimators.  In  all  cases,  rounding  errors,  especially 
for  the  estimates  of  sampling  variances,  can  be  seri- 
ous on  a  calculator.  Now  and  in  the  future,  a  com- 
prehensive analysis  of  any  set  of  multiple  capture 
data  will  require  the  use  of  sophisticated  computer 
programs.  A  good  computer  system  and  its  accompa- 
nying software  are  now  mandatory,  if  biologists  are 
to  benefit  from  the  statistical  and  theoretical  ad- 
vances made  in  the  past  decade.  A  general,  flexible, 
easy-to-use  system  will  be  a  great  help  in  future  re- 
search on  biological  populations." 

The  foregoing  thoughts  have  been  presented  to 
help  the  biologist  comprehend  the  magnitude  of 
the  factors  that  influence  results  of  small  mammal 
studies.  They  are  not  meant  to  create  doubt  but, 
rather,  to  create  caution  in  making  too  broad  gener- 
alizations based  on  small  samples  and  to  emphasize 
the  need  to  design  all  monitoring  studies  to  be  com- 
parable with  previous  studies.  Only  when  the  effects 
of  variables  discussed  in  this  section  are  about  equal 
can  results  of  sampling  be  compared  with  any  confi- 
dence from  year  to  year. 

Most  small  mammal  surveys  for  land  manage- 
ment purposes  are  to  determine  trends  in  species 
and  numbers,  not  necessarily  to  determine  densities 
(Kepner  1978).  Even  though  the  Lincoln  Index 
method  or  Probability  of  Capture  method  may  pro- 
vide somewhat  crude  results,  data  derived  can  be 
effectively  used  for  trends  so  long  as  conditions  are 
comparable  at  each  sampling  period. 


450 


Rodents  and  Insectivores 


/ 


LITERATURE  CITED 


ANDERSON,  B.W.J.  DRAKE,  and  R.D.  OHMART.  1977. 
Population  fluctuations  in  nocturnal  rodents  in  the 
Lower  Colorado  River  Valley.  Pages  183-192  in  Proc. 
Symp.  on  the  Importance,  Preservation,  and  Manage- 
ment of  the  Riparian  Habitat.  U.S.  Dep.  Agric,  For. 
Serv.  Gen.  Tech.  Report  RM-43. 

ANDERSON,  DR., J.L.  LAAKE,  BR.  CRAIN,  and  K.P.  BURN- 
HAM.  1976.  Guidelines  for  line  transect  sampling  of 
biological  populations.  Utah  Coop.  Wildl.  Res.  Unit. 
Logan.  27pp. 

and  R.D.  OHMART.  1977.  Rodent  bait  additive 

which  repels  insects.  J.  Mammal.  58:242. 

BAILEY,  J.A.  1969.  Trap  response  of  wild  cottontails.  J. 
Wildl.  Manage.  33(l>48-58. 

BLAIR,  W.F.  1941.  Techniques  for  the  study  of  mammal 
populations.  J.  Mammal.  22(2):  148- 157. 

BRADT,  G.W.  1938.  A  study  of  beaver  colonies  in  Michi- 
gan. J.  Mammal.  19:139-162. 

BURNHAM,  K.P.  1972.  Estimation  of  population  size  in 
multiple  capture-recapture  studies  when  capture 
probabilities  vary  among  animals.  Ph.D.  dissertation. 
Oregon  State  University,  Corvallis. 

DR.  ANDERSON,  and  J.L.  LAAKE.  1980.  Estimation 

of  density  from  line  transect  sampling  of  biological 
populations.  Wildl.  Monogr.  72:1-202. 

and  W.S.  OVERTON.  1979-  Robust  estimation  of 


population  size  when  capture  probabilities  vary 
among  animals.  Ecology  60(5):927-936. 

BURT,  W.H.  and  R.P.  GROSSENHEIDER.  1964.  A  field 
guide  to  the  mammals.  The  Riverside  Press.  Cam- 
bridge, MA.  284pp. 

CALHOUN,  J.B.  and  J.V.  CASBY.  1958.  Calculation  of  home 
range  and  density  of  small  mammals.  Public  Health 
Service,  Public  Health  Mongr.  55.  24pp. 

CALL,  M.W.  1970.  Beaver  pond  ecology  and  beaver-trout 
relationships  in  southeastern  Wyoming.  Ph.D.  disserta- 
tion. Univ.  of  Wyoming,  Laramie.  319pp. 

CHITTY,  D.  1937.  A  ringing  technique  for  small  mammals. 
J.  Animal  Ecol.  6:36-53- 

and  M.  SHORTEN.  1946.  Techniques  for  the  study 

of  the  Norway  rat  (Rattus  norvegicus).  J.  Mammal. 
27:63-78. 

COLLIER,  E.  1957.  Revolutionary  new  trap.  Outdoor  Life. 
Sept.  1957,  p.  38-41,  80,  and  Oct.  1957,  p.  70-73, 
80,  82. 

COOLEY,  ME.  1948.  Improved  toe-tag  for  marking  fox 
squirrels.  J.  Wildl.  Manage.  12:213- 

CORMACK,  R.M.  1981.  Log-linear  models  for  capture- 
recapture  experiments  on  animal  populations.  Univ. 
of  Otago.  Dunedin,  New  Zealand. 

DAVIS,  D.E.  1956.  Manual  for  analysis  of  rodent  popula- 
tions. Edwards  Brothers,  Inc.  Ann  Arbor,  MI.  82pp. 

DELURY,  D.B.  1951.  On  the  planning  of  experiments  for 
the  estimation  of  fish  populations.  J.  Fish.  Res.  Board 
of  Canada.  8:281-307. 

DENNEY,  R.N.  1952.  A  summary  of  North  American  bea- 
ver management.  Colo.  Game  and  Fish  Dep.  Current 
Report  28.  Denver,  CO.  58pp. 

DICE,  L.R.  1941.  Methods  for  estimating  populations  of 
animals.  J.  Wildl.  Manage.  5(4>398-407. 

DYMOND,  JR.  1947.  Fluctuations  in  animal  populations 
with  species  reference  to  those  of  Canada.  Trans. 
Royal  Soc.  Canada  4l(5):l-34. 


EBERHARDT,  L.L.  1969.  Population  estimates  from  recap- 
ture frequencies.  J.  Wildl.  Manage.  33(l>28-39. 

ELTON,  C.  1942.  Voles,  mice  and  lemmings.  Clarendon 
Press.  Oxford.  496pp. 

ERRINGTON,  P.L.  1944.  Additional  studies  on  tagged 
young  muskrats.  J.  Wildl.  Manage.  8:300-306. 

and  C.S.  ERRINGTON.  1937.  Experimental  tagging 

of  young  muskrats  for  purposes  of  study.  J.  Wildl. 
Manage.  24:231-260. 

EVANS,  EC.  and  R.  HOLDENRIED.  1943-  A  population 
study  of  the  Beechey  ground  squirrel  in  central  Cali- 
fornia. J.  Mammal.  24:231-260. 

FITZWATER,  W.D.,  Jr.  1943.  Color  marking  of  mammals, 
with  special  reference  to  squirrels.  J.  Wildl.  Manage. 
7:190-192. 

FREDIN,  R.A.  1954.  Causes  of  fluctuations  in  abundance  of 
Connecticut  River  shad.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  Fishery  Bull.  88.  Washington,  DC. 

GETZ,  L.L  1961.  Response  of  small  mammals  to  live-traps 
and  weather  conditions.  Am.  Midi.  Nat.  66(  1  ):160- 
169. 

GILES,  R.H.,  ed.  1969-  Wildlife  management  techniques. 
3rd  ed.  The  Wildl.  Soc.  Washington,  DC.  623pp. 

HANSSON,  L.  1974.  Influence  area  of  trap  stations  as  a 
function  of  number  of  small  mammals  exposed  per 
trap.  Acta  Theriol.  19:19-25. 

HAYNE,  D. W.  1 949a.  Two  methods  for  estimating  popula- 
tion from  trapping  records.  J.  Mammal.  30:399-41 1. 

1949b.  Calculation  of  size  of  home  range.  J.  Mam- 
mal. 30(1):1-18. 

1976.  Experimental  designs  and  statistical  analyses 


in  small  mammal  population  studies  in  Populations  of 
Small  Mammals  Under  Natural  Conditions.  The  Pyma- 
tuning  Symposia  in  Ecology,  Vol.  5.  Special  Publ. 
Series,  Pymatuning  Lab.  of  Ecology,  Univ.  of  Pitts- 
burgh. Symposia  held  at  the  Pymatuning  Lab.  of  Ecol- 
ogy, May  14-16,  1976. 

HENSLEY,  A.L.  and  H.  TWINING.  1946.  Some  early  sum 
mer  observations  on  muskrats  in  a  northeastern  Cali- 
fornia marsh.  Calif.  Fish  and  Game  32:171-181. 

HOFFMEISTER,  D.F.  and  L.  DE  LA  TORRE.  I960.  A  revi- 
sion of  the  woodrat  Neotoma  stephensi.  J.  Mammal. 
4l(4):482. 

INGLES,  L.G.  1965.  Mammals  of  the  Pacific  States.  Stanford 
Univ.  Press.  Stanford,  CA.  506pp. 

JACKSON,  H.H.T.  1961.  Mammals  of  Wisconsin.  Univ. 
Wisconsin  Press.  Madison.  504pp. 

JENSEN,  T.S.  1975.  Trappability  of  various  functional 

groups  of  the  forest  rodents  Clethrionomys  glareolus 
and  Apodemus  Jlavicollis,  and  its  application  in  den- 
sity estimations.  Oikos  26(2):  196-204. 

JORGENSEN,  CD.  and  H.D.  SMITH.  1974.  Mini-grids  and 
small  mammal  estimates  in  Proceedings  of  the  Utah 
Academy  of  Sciences,  Arts,  and  Letters.  Vol.  51,  Part  I. 

KEITH,  LB.  1963.  Wildlife's  ten-year  cycle.  Univ.  Wiscon 
sin  Press.  Oxford.  201pp. 

KEPNER,  W.G.  1978.  Small  mammals  of  the  Black  Canyon 
and  Skull  Valley  planning  units,  Maricopa  and  Yavapai 
Counties,  Arizona.  U.S.  Dep.  Inter.,  Bur.  Land  Manage. 
Tech  Note  350.  37pp. 

KREFTING,  L.W.,  J.H.  STOECKLER,  B.J.  BRADLE,  and  WD. 
FITZWATER.  1962.  Porcupine-timber  relationships  in 
the  Lake  States.  J.  For.  60:325-330. 

LESLIE,  PH.  1952.  The  estimation  of  population  parame- 
ters from  data  obtained  by  means  of  the  capture- 
recapture  method:  II.  The  estimation  of  total  num- 
bers. Biometrika  38(3/4):368-388. 


Rodents  and  Insectivores 


451 


and  D.H.S.  DAVIS.  1939.  An  attempt  to  determine 

the  absolute  number  of  rats  on  a  given  area.  J.  Anim. 
Ecol.  8:94-113. 

MASER,  BR.  MATE,  J.F.  FRANKLIN,  and  C.T.  DYRNESS. 
1981.  Natural  history  of  Oregon  coast  mammals.  U.S. 
Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep.  PNW-133. 
496pp. 

MOREJOHN,  G.V.  and  WE.  HOWARD.  1956.  Moult  in  the 
pocket  gopher,  Thomomys  bottae.  J.  Mammal. 
37:201-212. 

MOSBY,  H.S.  ed.  1963.  Wildlife  investigational  techniques, 
2nd  Edition.  The  Wildl.  Soc.  Washington,  DC.  4 19pp. 

NIXON,  C.N.,  W.R.  EDWARDS,  and  L.L.  EBERHARDT. 

1967.  Estimating  squirrel  abundance  from  live  trap- 
ping data.  J.  Wildl.  Manage.  31(1):96-101. 

OTARRELL,  M.J.,  D.W.  KAUFMAN,  and  D.W.  LUNDOHL. 
1977.  Use  of  live-trapping  with  the  assessment  line 
method  for  density  estimation.  J.  Mammal.  58(4):575- 
582. 

OTIS,  D.L.,  K.P.  BURNHAM,  G.C.  WHITE,  and  DR.  ANDER 
SON.  1978.  Statistical  inference  from  capture  data 
on  closed  animal  populations.  Wildl.  Monogr.  62:1- 
135. 

PETRUSEWICZ,  K.  and  A.  MACFAYDEN.  1970.  Productiv- 
ity of  terrestrial  animals  — principles  and  methods. 
Inter.  Biol.  Prog.  Handbook  1 3.  FA.  Davis  Co.  Phila- 
delphia, PA.  190pp. 

REYNOLDS,  H.C.  1945.  Some  aspects  of  the  life  history 
and  ecology  of  the  opossum  in  central  Missouri.  J. 
Mammal.  26:361-379. 

RUTHERFORD,  W.H.  1964.  The  beaver  in  Colorado,  its 
biology,  ecology,  management  and  economics.  Colo. 
Game,  Fish  and  Parks  Dep.  Fed.  Aid  in  Wildl.  Rest. 
Proj.  W-83-R  Denver,  CO.  49pp. 

SANDERSON,  G.C.  1961.  Estimating  opossum  population 
by  marking  young.  J.Wildl.  Manage.  25:20-27. 

SARRAZLN,  J.P.R.  and  JR.  BIDER.  1973-  Activity,  a  ne- 
glected parameter  in  population  estimates — the  devel- 
opment of  a  new  technique.  J.  Mammal.  54(2  ):369- 
382. 

SCHEMNITZ,  S.D.  ed.  1980.  Wildlife  management  tech- 
niques. 4th  ed.  revised.  The  Wildl.  Soc.  Washington, 
DC.  686pp. 

SCHULTZ,  V.  1961.  An  annotated  bibliography  on  the  uses 
of  statistics  in  ecology — a  search  of  3 1  periodicals. 
Publication  TID-3908.  U.S.  Atomic  Energy  Comm. 
Office  of  Tech.  Info,  Environmental  Science  Branch, 
Div.  of  Bio.  and  Medicine.  Washington,  DC.  31 4pp. 

SCOTT,  D.T.,  CD.  JORGENSEN,  and  H.D.  SMITH.  1978. 
Comparison  of  live  and  removal  methods  to  estimate 
small  mammal  densities.  Acta  Therio.  23(8):  173- 193- 


SEBER,  G.A.F.  1973-  The  estimation  of  animal  abundance. 
Hafner  Press,  New  York,  NY.  506pp. 

.  1982.  Estimation  of  animal  abundance  and  related 

parameters.  2nd.  ed.  Hafner  Press.  New  York,  NY. 
654pp. 

SHEPPE,  W.  1972.  The  annual  cycle  of  small  mammal 
populations  on  a  Zambian  floodplain.  J.  Mammal. 
53(3>445-460. 

SMITH,  H.D.,  C  D.  JORGENSEN,  and  H.D.  TOLLEY.  1972. 
Estimation  of  small  mammals  using  recapture  meth- 
ods: partitioning  of  estimator  variables.  Acta  Therio. 
17:57-66. 

SMITH,  M.H.,  R.  BLESSING,  J.G  CHELTON,  J.B.  GENTRY, 
G.  GOLLEY,  andJ.T.  MCGINNIS.  1971.  Determining 
density  for  small  mammal  populations  using  a  grid 
and  assessment  lines.  Acta  Theriol.  16:105-125. 

SMITH,  R.,  H.  GARDNER,  J.B.  GENTRY,  D.W.  DAUFMAN, 
and  M.J.  OTARRELL.  1975.  Density  estimation  of 
small  animal  populations,  Pages  25-53  in  F.G.  Golley, 
K.  Petrusewicz,  and  L.  Ruszkowski,  eds.  Small  Mam- 
mals: Their  Production  and  Population  Dynamics, 
Intern.  Biol.  Prog.  5.  Cambridge  Univ.  Press,  London. 

SNEDECOR,  G.W.  and  W.G.  COCHRAN.  1967.  Statistical 
methods,  6th  ed.  Iowa  State  Univ.  Press,  Ames.  593pp. 

SOKAL,  R.R.  and  F.J.  ROHLF.  1969.  Biometry:  the  princi- 
ples and  practice  of  statistics  in  biological  research. 
W.H.  Freeman  and  Co.  San  Francisco,  CA.  776pp. 

SPEIRS,  J.M.  1939.  Fluctuations  in  numbers  of  birds  in  the 
Toronto  region.  Auk  56:411-419. 

STICKEL,  L.F.  1948.  The  trap  line  as  a  measure  of  small 
mammal  populations.  J.  Wildl.  Manage.  12:153-161. 

1954.  A  comparison  of  certain  methods  of  measur- 
ing ranges  of  small  mammals.  J.  Mammal.  35(1):1-15. 

SUMMERLIN,  C.T  andJ.L.  WOLFE.  1973.  Social  influences 
on  trap  response  of  the  cotton  rat,  (Sigmodon  hispi- 
dus).  Ecology  54(5):  11 56- 11 59. 

SVIHLA,  A.  1934.  Development  and  growth  of  deer  mice 
(Peromyscus  maniculatus  artemisiae),  J.  Mammal. 
15:99-104. 

TAKOS,  M.J.  1943-  Trapping  and  banding  muskrats.  J. 
Wildl.  Manage.7:400-407. 

TANAKA,  R  1972.  Investigation  into  the  edge  effect  by 
use  of  capture-recapture  method  in  a  vole  population. 
Res.  Pop.  Ecol.  12(1  ):1 11-125. 

WHITE,  G.C,  DR.  ANDERSON,  K.P.  BURNHAM,  and  D.L. 
OTIS.  1982.  Capture-recapture  and  removal  methods 
for  sampling  closed  populations.  Los  Alamos  National 
Laboratory.  Los  Alamos,  NM.  235pp. 

YERGER,  R.W.  1953-  Home  range,  territoriality, 

and  populations  of  the  chipmunk  in  central  New 
York.  J.  Mammal.  34:448-458. 


452 


Rodents  and  Insectivores 


22 


LAGOMORPHS 


Joseph  A.  Chapman 

Utah  State  University 
Logan,  UT  84322 


Editor's  Note:  Lagomorphs,  among  other  midsize 
mammals,  have  often  been  neglected  in  wildlife 
surveys.  Although  often  locally  important  as  a 
game  or  pest  species,  they  have  not  received  the 
same  amount  of  attention  as  other  more  conspicu- 
ous or  economically  important  species.  Further- 
more, typical  inventorying  techniques  and 
equipment  used  for  other  species  are  not  useful  in 
studying  lagomorphs.  This  chapter  identifies  some 
possible  techniques  that  can  be  used  for  inventory- 
ing and  monitoring  this  often  overlooked  group 
of  species. 


Gale  R.  Willner 


The  Ecology  Center 
Utah  State  University 
Logan,  UT  84322 


INTRODUCTION 

Lagomorphs  are  important  components  of  the 
world's  ecosystems,  particularly  in  the  western  U.S. 
and  Alaska  where  they  make  the  base  of  many  carni- 
vore food  chains.  In  addition,  they  are  important 
indicators  of  habitat  quality,  sport  animals,  depreda- 
tors, and  models  for  ecological  research. 

The  order  Lagomorpha  occurs  worldwide  and 
contains  two  families,  Leporidae  and  Ochotonidae, 
both  found  in  the  U.S.  The  North  American  Lepori- 
dae includes  hares  and  jackrabbits  (Lepus  sp.),  cot- 
tontails (Sylvilagus  sp. ),  and  volcano  rabbit 
(Romerolagus  sp.),  found  only  in  Central  Mexico. 
Five  species  of  Sylvilagus  and  four  species  of  jackrab- 
bits occur  in  the  western  half  of  the  U.S.  Snowshoe 
hares  (L.  americanus)  are  widely  distributed  in 
Alaska  and  the  western  U.S.,  but  the  Alaskan  hare  (L. 
othus)  occurs  only  in  Alaska.  The  pikas  (Ochotona 
princeps)  are  the  only  members  of  the  family  Ocho- 
tonidae and  they  live  on  talus  slopes  throughout 
most  of  the  higher  mountains  in  the  western  U.S.  and 
southeastern  Alaska. 

All  Leporidae  are  adapted  for  quick  movement 
and  flight  from  danger.  Their  hind  legs  are  long  and 
adapted  to  cursorial  locomotion.  Their  ears  are  large 
and  moveable  to  permit  detection  of  approaching 
enemies.  Unlike  pikas,  their  eyes  are  large  and  suited 
to  their  crepuscular  and  nocturnal  habits.  These 
features  are  most  developed  in  the  hares.  Many  cot- 
tontails rely  on  thick  brush  or  even  burrows  to  avoid 
predators.  Pikas  are  rodent-like  in  appearance  and 
are  well-suited  to  living  among  the  rocks  and  boul- 
ders of  their  talus  slope  environment. 

In  general,  Lagomorphs  are  highly  fecund  and 
short-lived.  Population  levels  are  regulated  by  mor- 
tality (survival)  and  dispersal.  Indirectly,  the  compo- 
sition and  structure  of  the  plant  community  within 
their  habitat  plays  an  important  part  in  the  survival 
of  these  mammals.  They  are  usually  associated  with 
disturbed  or  subclimax  plant  communities. 


Lagomorphs 


453 


The  purpose  of  this  chapter  is  to  summarize 
techniques  that  are  most  useful  in  determining  Lago- 
morph  population  levels  as  well  as  examine  habitat 
features  that  are  important  to  this  order.  This  should 
enable  the  land  manager  to  evaluate  the  relationship 
of  density  and  habitat  and  thus  determine  the  status 
of  Lagomorphs  in  a  given  area. 

In  this  chapter,  the  Lagomorphs  are  divided  into 
four  species  groups  on  the  basis  of  taxonomy,  repro- 
ductive characteristics,  and  their  most  common  habi- 
tat requirements. 


HABITAT  FEATURES  CORRELATED  WITH 
SPECIES  GROUPS 


Jackrabbits 

Jackrabbits  are  true  hares,  primarily  associated 
with  arid  regions  of  the  1 1  western  States.  They  are 
found  at  elevations  from  below  sea  level  in  Death 
Valley  to  nearly  4,000  m  (13,000  ft).  Jackrabbits  will 
drink  water,  but  researchers  do  not  consider  it  a 
requirement  of  their  habitat. 


60c 


^^ 


WjSS*. 


o 

vo         r 


Miles 

400  800 
1 


400         800 


Kilometers 

120; 


Figure  1.     Distribution  of  the  black-tailed  jackrabbit  in  the  western  United  States  (adapted  from  Dunn  et  al. 
1982). 


454 


Lagomorphs 


White-tailed  jackrabbits  are  often  found  in  prai- 
ries and  in  open  flats  and  alpine  zones  above  tree- 
line.  Black-tailed  jackrabbits  are  the  most  abundant 
hare  in  the  western  U.S.;  white-tailed  jackrabbits  are 
locally  abundant.  The  antelope  and  white-sided  jack- 
rabbits are  uncommon  to  rare  throughout  their 
rather  restricted  ranges.    (See  Figures  1,2,  and  3  for 
distribution. ) 


White-tailed  jackrabbits  usually  occur  in  low- 
growing  scrub,  grasslands,  or  prairie  areas.  In  the 
western  States  they  are  often  found  at  higher  eleva- 
tions, even  above  treeline  where  they  inhabit  tundra- 
like vegetation  (Braun  and  Streeter  1968).  In  moun- 
tainous regions  they  also  occur  in  areas  vegetated 
with  disjunct  clumps  of  sagebrush.  See  Tables  1  and 
2  for  cover  and  food  plants  found  in  areas  occupied 
by  jackrabbits. 


White-tailed  jackrabbit. 


Black-tailed  jackrabbits  are  often  associated  with 
shortgrass-sagebrush  areas  in  arid  regions.  They  also 
inhabit  agricultural  areas.  White-tailed  jackrabbits  are 
also  found  in  sagebrush  areas  in  higher  mountain 
regions,  but  are  most  often  found  in  prairie  grass- 
lands. Black-tailed  jackrabbits  appear  to  displace  the 
white-tailed  jackrabbit  in  disturbed  or  cultivated 
areas  where  the  two  occur  sympatrically. 


Although  black-tailed  jackrabbits  seem  to  prefer 
diverse  plant  communities  made  up  of  disjunct 
cover,  there  are  little  quantitative  data  available. 
Black-tailed  jackrabbits  are  found  in  a  wide  array  of 
habitat  types,  all  of  which  are  characterized  or  domi- 
nated by  low-growing  trees  or  shrubs  and  well- 
drained  soils.  High  populations  occur  in  sagebrush 
{Artemisia  sp.)  interspersed  with  sparse  grass.  In 
general,  only  open  or  semi-open  areas  are  occupied 
by  this  species.  No  area  appears  to  be  too  open  as 
long  as  it  provides  adequate  hiding,  resting,  and 
birthing  sites.  Black-tailed  jackrabbits  are  usually 
associated  with  vegetation  that  is  no  more  than 
about  1  m  (3  ft)  high. 


The  antelope  jackrabbit  occurs  exclusively  in 
the  lower  Sonoran  life  zone  in  Arizona  where  its 
preferred  habitat  is  grassy  slopes.  However,  it  is  also 
found  in  the  cactus  belt  in  areas  dominated  by  mes- 
quite  (Prosopis  sp. ),  catclaw  (Acacia  greggi),  and 
various  grasses.  Smaller  populations  occupy  creosote 
bush  (Larrea  sp. )  deserts  and  valley  bottoms  (Figure 
2).  It  occurs  at  elevations  from  sea  level  to  1,200 
m  (3,960  ft)  (Vorheis  and  Taylor  1933).  Similarly, 
the  white-sided  jackrabbit  is  found  exclusively  in  the 
shortgrass  habitats  of  New  Mexico  (Figure  2).  The 
preferred  habitat  is  composed  of  65%  or  more 
grasses,  25%  or  less  forbs,  and  1%  or  less  shrubs 
(Bednarz  1977). 


Black-tailed  jackrabbit. 


Antelope  jackrabbit. 


Lagomorphs 


455 


Snow  shoe  and  Alaskan  Hares 

Alaskan  hares  are  among  the  least  known  Lago- 
morphs  in  North  America.  Recent  work  indicates 
that  L.  othus,  L.  arcticus,  and  L.  timidus  are  perhaps 
the  same  species  with  a  Holarctic  distribution 
(Dixon  et  al.  1983).  There  are  two  distinct  ecotypes 
of  Alaskan  hares:  L.  o.  othus,  associated  with  tundras 
and  alluvial  plains,  and  L.  o.  poudromus,  associated 
with  coastal  lowland  areas  of  the  Aleutian  Island 
chain  (Bittner  and  Rongstand  1982). 

Snowshoe  hares  are  found  in  most  higher  moun- 
tains of  the  western  States,  except  Arizona,  as  well 
as  throughout  Alaska  (Figure  3). 


Snowshoe  hares  inhabit  subclimax  forests  and 
transition  zones  adjacent  to  heavy  cover.  High  densi- 
ties of  snowshoe  hares  will  not  be  found  at  distances 
greater  than  200  to  400  m  (660  to  1,320  ft)  from 
conifer  stands.  High  densities  are  unlikely  in  areas 
with  a  solid  canopy  (Conroy  et  al.  1979).  Through- 
out their  range,  these  hares  appear  partial  to  dense 
thickets  and  windfalls  where  they  seek  cover.  In 
Alaska,  they  are  found  in  a  mosaic  of  plant  communi- 
ties that  include  upland  black  spruce/ledum/moss, 
willow-alder  thickets,  and  white  spruce/birch/aspen 
habitat  types  (Wolff  1978). 

Snowshoe  hares  are  also  found  in  dense,  second- 
growth  forests,  often  associated  with  aspen  (Populus 


60 


□   White-tailed 
|   Antelope 


^QtfcC?^        □   White-sided 


Miles 

400 
1 — 


Kilometers 


Figure  2.     Distribution  of  the  white-tailed,  antelope,  and  white-sided  jackrabbits  in  the  western  United  States 
(adapted  from  Dunn  et  al.  1982.) 


456 


Lagomorphs 


tremuloides)  thickets  and  coniferous  swamps.  In 
the  central  Rocky  Mountains,  snowshoe  hares  use 
areas  of  spruce  (Picea  sp.),  fir  {Abies  sp.  and  Pseu- 
dotsuga  sp.),  and  lodgepole  pine  (Pinus  contorta).  In 
those  habitats  they  prefer  conifer  stands  with  dense 
brushy  understories  (Dolbeer  and  Clark  1975).  In 
Utah,  prime  snowshoe  hare  habitat  occurs  in  areas 
with  high  sapling  numbers,  high  basal  area,  and 
dense  canopy  cover.  Optical  cover  densities  of  less 
than  40%  above  the  snowline  limit  winter  use  of 
a  habitat  by  snowshoe  hares  (Wolff  et  al.  1982). 
Snowshoe  hare  numbers  are  greatest  in  areas  sub- 
jected to  periodic  burning  (Grange  1965). 


Little  is  known  about  the  habitat  of  the  Alaskan 
hare  except  that  one  subspecies  prefers  tundra  or 
alluvial  plain,  and  the  other  prefers  coastal  lowlands 
(Bittner  and  Rongstand  1982). 

In  the  western  States  snowshoe  hares  are  invari- 
ably linked  with  conifers  at  higher  elevations.  How- 
ever they  may  range  away  from  coniferous  timber  at 
high  elevations  where  streams  are  thickly  bordered 
with  vegetation  (Dice  1926;  Orr  1940).  These  hares 
also  utilize  logged-over  areas  adjacent  to  standing 
conifers.  See  Tables  1  and  3  for  detailed  list  of  cover 
and  food  plants  found  on  areas  occupied  by 
snowshoe  hares. 


60 


|  Snowshoe 


Miles 

400  800 


Kilometers 

120 


Figure  3.     Distribution  of  the  snowshoe  and  Alaskan  hares  in  the  western  United  States  and  Alaska  (adapted 
from  Bittner  and  Rongstand  1982). 


Lagomorphs 


457 


Table  1.     Some  plants  occurring  in  habitats  occupied  by  various  Lagomorph  species. 


Lagomorph  and  Plant  Species 

Area 

Black-tailed  jackrabbit 

Rabbitbrush  (Chrysothamnus  sp.) 
Sagebrush  (Artemisia  tridentata) 
Saltbush  {Atriplex  sp.) 
Horsebrush  {Tetradymia  sp.) 

California 

Creosotebush  (Larrea  tridentata) 

Southern  Nevada  and 
southeastern  California 

Joshua  trees  (Yucca  brevifolia) 

Creosotebush 

Brig  ham  tea  (Ephedra  sp.) 

Southern  Utah 

Shadscale  (Atriplex  confertifolia) 

Greasewood  (Scarcobatus  vermiculatus) 

Rabbitbrush 

Sagebrush 

Juniper  (Juniperus  sp.) 

Pinyon  pines  (Pinus  edulis) 

Salt  deserts 

of 
Great  Basin 

Sagebrush 

Rabbitbrush 

Mesquite  (Prosopis  sp.) 

Snakeweed  (Gutierrezia  sarothrae) 

Soaptree  yucca  (Yucca  sp.) 

Agricultural  crops — alfalfa  hay 

Central  and 
southern  Utah 
New  Mexico 

White-tailed  jackrabbit 

Juniper  (J.  occidentalis) 
Lodgepole  pine  (Pinus  contorta) 
Dwarf  juniper  (J.  communis) 
Creambush  (Holodiscus  discolor) 
Granite/gilia  (Leptodactylon  purges) 

California 
(Lake  Tahoe) 

Sagebrush 

Bitterbrush  (Purshia  tridentata) 

Giant  wild  rye  (Elymus  condensatus) 

Chokecherry  (Prunus  melanocarpa) 

Snowberry  (Symphoricarpos  rotundifolius) 

Mountain  myrtle  (Pachystima  myrsinites) 

Northern  Utah 

Deerbrush  (Ceanothus  velotinus) 
Aspen  (Populus  tremuloides) 

Utah  and  Idaho 

Sagebrush 

Greasewood 

Shadscale 

Saltbush 

Dogwood  (Cornus  canadensis) 

Wild  rhubarb  (Polygonum  alaskanum) 

Horsetail  (Equisetum  sp.) 

Uintah  Basin 

Audubon's  cottontail 

Willow  (Salix  sp.) 

Buttonwillow  (Cephalanthus  occidentalis) 

Wild  grape  (Vitis  californica) 

Arroyo  willow  (S.  lasiolepis) 

Brushland  (Adenostoma  fasilutatuium) 

(Eriodictyeon  sp.) 

(Eriogonum  sp.) 

California 

Arrow-weed  (Pluchea  sericea) 
Screw-bean  mesquite  (Prosopis  pubesans) 
Catclaw  (Acacia  greggii) 

Southern  Nevada 

Joshua  trees 
Cactus 

Southern  Utah 

458 


Lagomorphs 


Table  1.     Some  plants  occurring  in  habitats  occupied  by  various  Lagomorph  species  (continued). 


Lagomorph  and  Plant  Species  (continued) 

Area 

Audubon's  cottontail  (continued) 

Brigham  tea 

Creosotebush 

Junipers 

Pinyon 

Evergreen  oaks  (Quercus  turbinella) 

Gumine  bush  (Garrya  flaresceni) 

Saltbush 

Rabbitbrush  (Chrysothamnus  nauseous) 

Snakeweed 

Winterfat  (Eurotia  sp.) 

Western  Utah 

Nuttall's  cottontail 

Sagebrush 

Rabbitbrush 

Wild  rye  (Elymus  condensatus) 

Bitterbrush 

Chokecherry  (Prunus  virginiana) 

Northern  Utah 

Brush  rabbit 

Bramble  {Rubus  sp.) 

Oregon 

Buckbrush  (Ceanothus  sp.) 

Willows 

California  rosebay  {Rhododendron  californicum) 

Chaparral   broom  {Baccharis  pilularis) 

Wild  rose  (Rosa  californica) 

Poison  oak  {Rhus  diversiloba) 

Scrub  oak  (0.  dumosa) 

Snowberry  (Symphoricarpus  albus) 

Yellow  pine  (Pinus  ponderosa) 

Manzanita  (Arctostaphylos  sp.) 

California 

Antelope  jackrabbit 

Catsclaw  {Acacia  greggi) 

Mesquite 

Grasses  (Graminea) 

Creosotebush  (Larrea  sp.) 

Arizona 

White-sided  jackrabbit 

Grasses 

New  Mexico 

Grama  (Bouteloua  gracilis) 
Black  grama  (B.  eriopoda) 
Buffalo  grass  (Buchloe  dacyloides) 
Wolftail  (Lycurus  phleoides) 

Snowshoe  hares  (Continental  U.S.) 

Willows 

Douglas  fir  (Pseudotsuga  menziesii) 

Snowberry 

Snowbrush 

Mountain  myrtle 

Mountain  ash  (Sorbus  scopulina) 

Aspen 

Northern  Utah 

Lagotnorphs 


459 


Table  1.     Some  plants  occurring  in  habitats  occupied  by  various  Lagomorph  species  (concluded). 


Lagomorph  and  Plant  Species  (Concluded) 

Area 

Snowshoe  hares  (Continental  U.S.)  (continued) 

Alpine  fir  (Abies) 

Engleman  spruce  (P.  englemanii) 

Lodgepole  pine 

Alder  (Alnus  sp.) 

Willow 

Conifers 

Chaparral  {Ceanothus  sp.) 

Manzanita 

Buckbrush  (C.  velutinus) 

Douglas  fir 

Red  cedar  (Thuja  plicata) 

Uintah  Mountains 
of  Utah 
California 

Southern  California 

Snowshow  hares  (Alaska) 

White  spruce  (Picea  glauca) 

Black  spruce  (P.  mariana) 

Paper  birch  (Betula  papyrifera) 

Aspen 

Balsam  poplar  (Populus  balsamifera) 

Alder  (Alnus  crispa) 

Dwarf  birch  (Betula  gladulosa) 

Willows 

Labrador  tea  (Ledum  groenlandlcum) 

Blueberry  (Vaccinium  uliginosum) 

Lowbush  cranberry  (Vaccinium  vitis-idaea) 

Rose  (Rosa  acicularis) 

Grasses  (Calamagrostis  sp.) 

Fireweed  (Epilobium  angustifolium) 

Pikas 

Rabbitbrush 

Sagebrush 

Bitterbrush 

Columbine  (Aquilegia  pubescens) 

California 

Sources:  (Orr  1940,  Janson  1946;  Davis  et  al.  1975;  Bednarz  1977;  Dice  1926;  Wolff  1978;  Chapman  1971;  Smith  1974a). 


Cottontails 

Cottontails  are  widely  scattered  throughout  1 1 
western  States,  but  are  not  present  in  Alaska.  The 
Nuttall's  cottontail  inhabits  the  intermountain  region, 
whereas  the  Audubon's  cottontail  is  found  in  the 
arid  Southwest  and  high  deserts  north  to  Montana. 
Ranges  of  these  two  species  broadly  overlap  in  Colo- 
rado, Montana,  Utah,  and  Wyoming  (Figure  4).  The 
brush  rabbit  is  confined  to  the  coasts  of  California 
and  Oregon,  whereas  the  pygmy  rabbit  is  found  in 
the  Great  Basin  with  a  small  disjunct  population 
in  eastern  Washington  (Figure  5).  The  eastern  cot- 
tontail occurs  primarily  from  the  Midwest  east,  with 


the  exception  of  Arizona  and  New  Mexico,  which 
have  large  eastern  cottontail  populations.  This  spe- 
cies occurs  peripherally  in  eastern  Colorado,  Wyo- 
ming, and  Montana  and  has  been  introduced  into 
Oregon  and  Washington  (Figure  5). 

The  Audubon's  cottontail  occurs  in  the  lower 
Sonoran  life  zone  and  often  prefers  stream  bottom 
vegetation.  Shrubs  interspersed  with  pinyon  (Pinus 
edulis)  and  juniper  (Juniperus  sp.)  provide  excellent 
cover  for  this  species.  Downed  trees  or  shrubs  piled 
170  to  220  per  ha  (425  to  1,210  per  a.)  create  ex- 
cellent habitat  for  this  species  (Kundaeli  and  Rey- 
nolds 1972). 


460 


Lagomorphs 


The  Nuttall's  cottontail  is  found  in  rocky, 
wooded,  or  brushy  areas.  It  is  often  observed  in 
rocky  ravines  next  to  sagebrush-covered  hills.  When 
Audubon's  and  Nuttall's  cottontails  occupy  the  same 
region,  the  former  is  found  on  desert  valleys  and 
the  latter  on  higher  rocky,  sagebrush-covered  slopes. 
Where  cover  is  scarce,  Nuttall's  cottontails  will  use 
burrows  dug  by  badgers  (  Taxidea  taxus)  or  crevices 
under  rocks  (Dice  1926;  Orr  1940). 

Pygmy  rabbits  are  always  found  in  sagebrush. 
They  prefer  tall  sagebrush  (1  to  1.5  in  [3  to  5  ft]) 
often  in  ravines.  They  dig  their  own  burrows  which 
are  found  at  the  base  of  sagebrush  clumps  (Chapman 
et  al.  1982). 


The  brush  rabbit  gets  its  name  from  its  prefer- 
ence for  dense,  brushy  cover.  Clumps  of  brush 
smaller  than  460  m    (14,900  ft  )  are  probably  not 
permanently  occupied  (Chapman  1971).  Brush  rab- 
bits have  been  reported  to  use  burrows,  but  they  do 
not  dig  their  own. 

In  Arizona  and  New  Mexico,  eastern  cottontails 
are  found  in  open  areas  and  along  stream  bottoms  or 
near  farm  buildings.  Where  they  have  been  intro- 
duced in  Washington  and  Oregon,  they  have  become 
established  on  old,  overgrown  farmsteads,  in  fence 
rows,  and  around  buildings.  Eastern  cottontails  in 
Minnesota  have  a  distinct  preference  for  shrubby 
vegetation  (Swihart  and  Yahner  1984).  They  are 


60 


□    Nuttall's 
|   Audubon's 


Miles 

400 

1 — 


Ki  lometers 

120 


Figure  4.     Distribution  of  the  Nuttall's  and  Audubon's  cottontails  in  the  western  United  States  (adapted  from 
Chapman  et  al.  1982). 


Lagomorphs 


461 


most  abundant  in  disturbed  successional  areas 
(Chapman  et  al.  1982).  See  Tables  1  and  4  for  de- 
tailed list  of  cover  and  food  plants  found  in  areas 
occupied  by  cottontails. 


Pikas  are  found  throughout  the  higher  mountain 
ranges  of  the  western  States  except  Arizona  and  New 
Mexico.  There  is  also  a  population  of  collared  pikas 
in  the  mountains  of  southern  Alaska  (Figure  6). 


Pikas 

Pikas  are  guinea-pig-shaped  Lagomorphs,  having 
short,  rounded  ears  and  no  visible  tail.  Pikas  move 
with  a  running  gait  rather  than  the  typical  hopping 
or  leaping  of  other  Lagomorphs.  They  are  the  only 
North  American  Lagomorphs  that  store  or  cache 
food.  They  are  active  during  the  day  and  are  ex- 
tremely vocal. 


Pikas  inhabit  talus  slopes  where  they  feed  al- 
most exclusively  on  grasses,  forbs,  and  low  bushes 
adjacent  to  these  talus  slopes.  Pikas  require  a  boul- 
der-strewn habitat  on  high  mountain  slopes  with 
a  nearby  source  of  food.  Pikas  require  an  abundance 
of  plant  species  which  they  can  collect  and  store 
in  "hay"  piles  for  later  use  in  winter.  Most  often  they 
prefer  alpine  meadows  with  abundant  grass  and  low 
shrubs  adjacent  to  talus  slopes. 


60c 


□    Brush 

^s^M}       D  Py9my 

*4^0«fcC3  □  Eastern 


Miles 

400 


1 
400 


Kilometers 


Figure  5.     Distribution  of  the  brush  rabbit,  pygmy  rabbit,  and  eastern  cottontail  in  the  western  United  States 
(adapted  from  Chapman  et  al.  1982). 


462 


Lagomorphs 


Pikas  are  noticeably  absent  from  the  Olympic 
Mountains  in  Washington  even  though  suitable  habi- 
tat appears  to  be  available.  Researchers  attribute 
this  to  the  lack  of  talus  occurring  between  the  Olym- 
pics and  the  Cascade  Range  where  pikas  are  abun- 
dant (Dalquist  1948).  Barriers  to  dispersal,  such 
as  climatic  conditions,  prevent  juveniles  from  occu- 
pying suitable  habitats  and,  therefore,  limit  popula- 
tion densities  (Smith  1974b).  Table  1  lists  plant 
species  found  on  areas  occupied  by  pikas. 

POPULATION  MEASUREMENT  TECHNIQUES 

Standard  census  techniques  may  be  applied  to 
Lagomorph  populations.  Those  techniques  are  sum- 


marized in  the  Wildlife  Management  Techniques 
Manual  (Davis  and  Winstead  1980)  and  in  The 
Handbook  of  Census  Methods  for  Terrestrial 
Vertebrates  (Davis  1982).  These  census  techniques 
may  be  divided  into  three  broad  categories:  ( 1 )  pres- 
ence or  absence,  (  2  )  relative  abundance,  and  (  3 ) 
density  estimates. 

Absolute  density  and  relative  abundance  are 
somewhat  misleading  terms.  In  reality  it  is  difficult  to 
meet  the  requirements  for  estimates  of  absolute  den- 
sity because  of  problems  associated  with  delineating 
boundary  areas.  Drive  counts  and  mark-recapture 
techniques  are  often  thought  of  as  measures  of  abso- 
lute density,  but  it  is  probably  better  to  consider 


60 


SfrKSSP. 


Miles 

0  400 

I 


400         800 


K  ilometers 


Figure  6.     Distribution  of  the  pika  in  the  western  United  States  and  Alaska  (adapted  from  Chapman  1979). 

Lagotnorphs  463 


Table  2.     Food  of  jackrabbits  as  reported  in  the  literature. 


Black-tailed  Jackrabbit 

White-tailed  Jackrabbit 

SHRUBS 

SHRUBS 

Shadscale 

Winterfat 

Fringed  sage  {Artemisia  frigida) 

Greasewood 

Sagebrush 

Hopsage  {Grayia  spinosa) 

Four-winged  saltbush  (Atriplex 

canescens) 

Horsebrush 

Snakeweed 

Creosote 

FORBS 

FORBS 

Globemallow  {Sphaeralcea 

Aster  (Casfer  sp.) 

cocci  ni  a) 

Vetch  (Astragalus  sp.) 

Winterfat  (Eurotia  lanata) 

Saltbush 

Larkspur  (Delphinium  sp.) 

Indian  paintbrush  (Castilleja  integra) 

Spiderwort  (Tradescantia  sp.) 

Golden  aster  (Chrysopsis  villosa) 

Alfalfa  (Medicago  sativa) 

Goosefoot  (Chenopodium  sp.) 

Mallow  plants  (Malvaceae  sp.) 

Corydalis  (Corydalis  aurea) 

Curly  dock  (Rumex  crispu) 

Daisy  (Erigeron  sp.) 

Yucca  (Yucca  glauca) 

Gilia  (Gilia  aggregate) 

Prickly  pear  (Opuntia  sp.) 

Gaura  (Gaura  coccinea) 

Sand  dropseed  (Sporobolus 

Sunflower  (Helianthus  annuus) 

cyptandrus) 

Summer  cypress  (Kochia  sp.) 

Pigweed  (Amaranthus  retroflexus) 

Blazing  star  (Liatris  punctata) 

Russian  thistle  (Salsola  kali) 

Lupine  (Lupinus  sp.) 

Buffalo  burr  (Solanum  rostratum) 

Alfalfa 

Lily  (Lilium  sp.) 

Bluebells  (Mertensia  lanceolata) 

Sedge  (Carex  sp.) 

Four-o'clocks  (Mirabilis  linearis) 

Cactus  (Opuntia  sp.) 

Musineon  (Musineon  divaricatum) 

Prickly  pear  cactus 

Locoweed  (Oxytropis  sp.) 

Beardtongue  (Penstemon  sp.) 

Plaintain  (Plantago  purshii) 

Cinquefoil  (Potentilla  sp.) 

Psoralea  (Psoralea  tenluiflora) 

Rabbitbrush 

Russian  thistle 

Groundsel  (Senecio  sp.) 

Scarlet  globe  mallow  (Sphaeralcea 

cocci  ni  a) 

Dandelion  (Taraxacum  officinale) 

Spreading  thermopsis  (Thermopsis 

divaricarpa) 

Verbena  (Verbena  bracteata) 

Yucca 

GRASSES 

GRASSES 

Needle-and-thread  (Stipa  comata) 

Crested  wheatgrass  (Agropyron 

Downy  brome  (Bromus  tectorum) 

cristatus) 

Western  wheatgrass  (Agropyron 

Western  wheatgrass 

smithii) 

Wheatgrass  (A.  trachycaulum) 

Barley  (Hordeum  sp.) 

Red  three  awn  (Aristida  longiseta) 

Buffalo  grass  (Buchloe 

Oats  (Avena  sativa) 

dactyloides) 

Blue  grama 

Winter  wheat  (Triticum  aestivum) 

Brome  (Bromus  anomalus) 

Blue  grama 

Smooth  brome  (B.  inermis) 

Indian  ricegrass  (Oryzopsis 

Downy  brome 

hymenoicJes) 

Sedges 

Lovegrass  (Eragrostis  sp.) 

Hairgrass  (Deschampsia  caespitosa) 

464 


Lagomorphs 


Table  2.     Food  of  jackrabbits  as  reported  in  the  literature  (concluded). 


Black-tailed  Jackrabbit 

White-tailed  Jackrabbit 

GRASSES 

Barnyard  grass  (Echinochloa  sp.) 
Panicum  {Panicum  sp.) 
Dropseed  (Sporobolus  sp.) 
Trichloris  (Trichloris  sp.) 

GRASSES 

Love  grass 

Arizona  fescue  (Festuca  arizonica) 

Foxtail  barley 

Junegrass  (Koeleria  cristada) 

Ringgrass 

Bluegrass  (Poa  sp.) 

Squirreltail  (Sitanion  hystrix) 

Sand  dropseed 

Needle-and-thread 

Letterman  needlegrass  (S.  lettermanii) 

Wheatgrass 

White-sided  Jackrabbit 

Antelope  Jackrabbit 

GRASSES 

Buffalo  grass 

Tabossa  grass 

Fiddleneck  (Amsinickia  sp.) 

Wolftail 

Blue  grama 

Vine  Mesquite  (Pancium 

obstosum) 
Ring  muhly  (Muhlenbergia 

torreyi) 
Wooly  Indian  wheat  (Plantago 

purshii) 
Wright  buckwheat  (Eriogonum 

wrightii) 

FORBS 

Cactus  (Echinocactus  sp.) 

GRASSES 
Three  awn 
Grama 
Love  grass 
Spike  grass 

Sandbar  (Cenchrus  Sp.) 
Drop  Seed 

Red  top  (Agrostis  sp.) 
Barnyard  grass 
Finger  grass  (Chloris  sp.) 

Sources:  (Hanson  and  Flinders  1969;  Bear  and  Hansen  1966,  Flinders  and  Hansen  1972,  Bednarz  1977;  Vorheis  and  Taylor  1933). 


these  techniques  as  being  more  accurate  estimates 
rather  than  measures  of  absolute  density. 

Presence 

The  presence  of  Lagomorphs  can  be  determined 
by  direct  observation  of  the  animal  or  by  "activity 
signs."  Activity  signs  include  fecal  pellets,  active 
runways  and  trails,  tracks,  skeletal  remains,  and  feed- 
ing sites  (Figure  7).  For  example,  jackrabbits  can 
usually  be  observed  in  the  field  because  of  their 
rather  large  size.  Snowshoe  hares  use  runways,  espe- 
cially in  the  snow.  Hay  piles  along  talus  slopes  are 
good  indicators  of  pikas.  The  presence  of  these  signs 
does  not  necessarily  provide  an  estimate  of  numbers 
of  Lagomorphs  in  a  given  area,  but  will  help  deter- 
mine sites  where  they  occur  or  avoid. 

Relative  Abundance 

The  most  commonly  used  technique  for  census- 
ing  several  species  of  Leporids  is  the  roadside  count. 
Roadside  census  of  Leporids  is  best  in  the  evenings 


or  early  morning  when  Lagomorphs'  eyes  can  be 
reflected  from  a  spotlight.  Jackrabbits  and  desert  cot- 
tontails have  been  censused  in  the  evenings  along 
unpaved  roads  in  shortgrass  prairie  habitat  (Flinders 
and  Hansen  1973).  The  "eye  shine"  provided  by 
the  spotlight  and  subsequent  "freeze"  of  the  Lago- 
morphs permitted  easy  identification  of  species  and 
counting.  Roadside  counts  may  be  used  as  long  as 
factors  such  as  time  of  day,  time  of  year,  and 
weather  conditions  remain  constant  (equal).  This 
type  of  transect  count  provides  an  index  to  relative 
abundance  but  does  not  give  a  density  estimate 
(Chapman  et  al.  1982). 


Density  Estimates 

Density  estimates  can  be  obtained  by  drive 
counts.  Using  this  method,  several  people  spread 
equally  across  an  area,  walk  abreast  along  parallel 
transects  and  count  Leporids  as  they  are  flushed. 
Flushing  angles  and  flushing  distances  should  be  re- 
corded. The  flushing  angle  is  obtained  by  pinpointing 


Lagomorphs 


465 


Table  3.     Food  of  the  snowshoe  hare  in  western 
North  America  as  reported  in  the  literature. 


Alaska 


Birch 

Spruce 

Alder 

Willow 

Viburnum  (Viburnum  edule) 

Labrador  tea  (Ledum  groenlandicum) 

Aspen 

Rose 

Bramble  (Rubus  ideaus) 

Grass 

Dogwood  (Cornus  canadensis) 

Fireweed 

Horsetail 

Wild  rhubarb 


Ontario 


Pines  (Pinus  strobus  and  P.  resinosa) 
Aspen 
Alder 

Hazel  (Corylus  cornuta) 
(Ostrya  virginiana) 
Juneberry  (Amelanchier  sp.) 
Willow 

Poplar  (P.  balsamifera) 
White  birch 
Grasses 


Washington 


Spotted  catsear  (Hypochoeris  radicata) 


Sources:  (Radwan  and  Campbell  1968,  Wolff  1978;  de  Vos 
1964). 


Snowshoe  hare  with  winter  coat. 


Figure  7.     Cottontail  fecal  pellet  group. 

the  location  of  the  animal  and  noting  its  compass 
direction  by  using  the  transect  line  as  a  point  of 
reference.  The  flushing  distance  is  the  distance  be- 
tween the  center  of  the  transect  and  point  where 
the  Leporid  is  first  observed.  Flushing  distance  and 
angle  will  vary  according  to  the  density  and  height 
of  the  cover  in  which  the  Leporid  is  living.  In  gen- 
eral, the  greater  the  density  or  thickness  of  the 
cover,  the  shorter  the  flushing  distance.  The  flushing 
distance  and  angle  will  be  determining  factors  in 
the  width  of  the  transect  used.  The  greater  the  ob- 
servable flushing  distance  and  angle,  the  wider  the 
transect  will  be  and,  thus,  the  greater  the  area  cen- 
sused  (Webb  1942). 

Behavioral  patterns  and  density  of  vegetation 
tend  to  preclude  drive  counts  as  a  suitable  technique 
for  cottontails  and  snowshoe  hares.  Cottontails  tend 
to  freeze  rather  than  flee,  making  a  head  count 
difficult. 

Snowshoe  hares  inhabit  conifer  stands  with 
thick  understories,  making  them  difficult  to  see  and 
count.  However,  drive  counts  have  been  successfully 
used  in  jackrabbit  studies  where  sagebrush  was  the 
dominant  vegetation  type  (Gross  et  al.  1974).  Biases 
in  census  data  may  result  from  a  change  in  flushing 
behavior,  observer  behavior,  and  duplicate  counting 
of  hares.  Gates  (1969)  used  an  equation  to  adjust  for 


466 


Lagomorphs 


flushing  behavior  changes  in  jackrabbits  due  to  sea- 
son and  density.  Jackrabbits  in  sagebrush  habitat  are 
easily  visible  and  thus  can  be  censused  by  using 
the  drive  count  technique  (F.H.  Wagner,  pers.  com- 
mun.  )•  Broadbooks  (1965)  successfully  used  a  30-m 
(100-ft )  wide  transect  to  census  pikas  on  a  rock 
formation. 

Lagomorph  density  can  also  be  estimated  by 
mark-recapture  techniques  and  fecal  pellet  group 
counts.  Taylor  ( 1930)  was  one  of  the  first  investiga- 
tors to  use  fecal  pellets  as  an  indicator  of  Leporid 
abundance  in  the  West.  Pellet  counts  were  also  used 
to  evaluate  Audubon's  cottontail  population  density 
in  a  pinyon- juniper  forest  in  New  Mexico  (Kundaeli 
and  Reynolds  1972).  However,  pellet  counts  are 
more  useful  for  determining  the  habitat  preference 
of  Lagomorphs  than  for  determining  density.  The 
presence  of  pellets  was  used  to  determine  the  use  of 
montane  forest  types  by  snowshoe  hares  in  Utah 
(Wolff  et  al.  1982).  These  methods  are  useful  for 
comparing  Lagomorph  populations  in  different  years, 
different  habitat  types,  and  different  seasons.  When 
using  pellet  counts  to  estimate  populations,  adjust- 
ments for  varying  defecation  rates  may  be  necessary 
(See  Lord  1963). 

Mark-recapture  techniques  involve  sampling  a 
segment  of  a  population,  marking  and  releasing 
them,  and  at  a  later  date  recapturing  a  percentage  of 
the  marked  individuals.  Mark  and  recapture  tech- 
niques are  frequently  used  because  they  not  only 
give  census  data  but  also  specific  information  on 
factors  such  as  sex  and  age,  survival,  and  movements 
of  individuals.  These  techniques  have  been  exten- 
sively used  to  assess  Lagomorph  population  densities. 

The  Wildlife  Management  Techniques  Man- 
ual describes  several  models  such  as  the  Lincoln 
Index  (Lincoln  1930)  or  Jolly  Index  (Jolly  1963)  to 
estimate  the  size  of  a  population  by  the  mark-recap- 
ture method  (Davis  and  Winstead  1980).  Population 
levels  based  on  mark-recapture  techniques  often 
underestimate  the  actual  number  of  animals  present 
in  an  area  (Eberhardt  1969).  Otis  et  al.  (1978)  dis- 
cuss the  appropriateness  of  different  models  available 
for  estimating  the  population  size  from  capture  and 
recapture  data. 

Some  problems  associated  with  mark-recapture 
procedures  include  the  loss  of  marks  or  the  tend- 
ency of  some  individuals  to  become  "trap-happy"  or 
"trap-shy,"  all  resulting  in  inaccurate  population  esti- 
mates. Loss  of  ear  tags  can  result  in  inflated  popula- 
tion estimates.  Some  problems  associated  with  this 
technique  include  variation  in  recapture  rates  and 
the  behavioral  responses  to  traps  exhibited  by  indi- 
vidual animals  (Otis  et  al.  1978). 

Mark-recapture  techniques  have  been  used  to 
census  brush  rabbits  in  bush  lupine  ( Lupinus  arbus- 


tus)  in  Humboldt  Bay,  California  (Shields  1958), 
Audubon's  cottontails  in  mesquite-grassland  habitat 
associations  (Asserson  1967),  Nuttall's  cottontail 
in  shrub- juniper  scrubland  in  Oregon  (McKay  and 
Verts  1978),  and  jackrabbits  in  Utah  (Clark  1972). 
Factors  affecting  the  trap  responses  of  eastern  cot- 
tontails in  Oregon  included  barometric  pressure  and 
temperature,  but  not  precipitation  (Chapman  and 
Tretheway  1972).  Methods  to  capture,  handle,  and 
mark  Lagomorphs  are  outlined  in  The  Handbook  of 
Census  Methods  for  Terrestrial  Vertebrates  (Feist 
1982;  Smith  1982;  and  Wolff  1982)  and  Wild  Mam- 
mals of  North  America  (Chapman  and  Feldhamer 
1982). 

Jackrabbits  can  be  captured  in  winter  by  using 
large  Havahart  traps  baited  with  alfalfa  or  apples  (G. 
Smith,  pers.  commun.).  Cottontails  can  be  caught 
in  either  baited  or  unbaited  wooden  box  traps  (Fig- 
ure 8;  Edwards  1975).  Leporids  can  also  be  captured 
from  vehicles  by  using  a  spotlight  and  a  hand-held 
net  (Labisky  1959,  1968). 

Lagomorphs  can  be  marked  with  size  3  monel 
ear  tags,  hair  dye,  or  toe-clipping  to  identify  individ- 
uals (Wolff  1982).  Depending  on  the  size  of  the 
Lagomorph,  other  types  of  markers  include  commer- 
cially available  colored  discs,  colored  discs  cut  from 
surveyors  flagging,  an  alcohol-picric-acid  solution, 
and  notching  or  punching  holes  in  the  ears  (Figure 
9).  Approaching  and  handling  Lagomorphs  requires 
care  to  avoid  the  flight  response  and  subsequent 
injury  of  the  rabbit  or  hare  while  the  animal  is  still 
in  the  trap.  Placing  the  Lagomorph  in  a  cloth  bag 
tends  to  reduce  stress  on  the  animal  while  handling 
it  for  marking  or  weighing. 


Figure  8.     Typical  wooden  box  trap  used  for 
capturing  cottontails. 


Lagomorphs 


467 


Figure  9.     Black-tailed  jackrabbit  with  ear  tag 
in  place. 


POPULATION  LEVELS  AND  DENSITY 

Although  there  are  little  specific  data  on  the 
absolute  densities  of  Lagomorphs  associated  with 
specific  habitat  types,  some  general  guidelines  can  be 
obtained  from  the  literature.  These  guidelines  will 
be  useful  to  the  manager  because  they  can  compare 
densities  on  their  study  areas  with  existing  informa- 
tion in  the  literature.  Thus,  the  manager  can  gain 
insight  into  the  status  of  Lagomorph  populations  in  a 
particular  area. 

In  general,  Leporid  populations  do  not  reach 
extremely  high  densities.  There  are  of  course  excep- 
tions as  noted  in  localized  situations  with  black- 
tailed  jackrabbits  and  snowshoe  hares.  Uniform  den- 
sities of  more  than  3  or  4  Leporids  per  ha  ( 1  to  3 
per  a. )  would  be  considered  high  for  most  North 
American  species. 

Population  estimates  for  jackrabbits  in  the 
Southwest  vary  between  0.2  and  1.2  per  ha  (0.1  to 
0.5  per  a.);  in  California  they  were  reported  at  3  per 
ha  ( 1  per  a.),  and  at  35  per  ha  ( 14  per  a. )  in  culti- 
vated areas  of  Kansas  (See  Dunn  et  al.  1982  for 
references). 


Audubon's  cottontail  in  shortgrass-prairie  re- 
gions of  Colorado  reached  densities  of  0.02  per  ha 
(0.08  per  a.)  during  the  winter.  Densities  of  Nuttall's 
cottontails  in  shrub-juniper  scrubland  in  Oregon 
varied  between  0.06  to  2.5  per  ha  (0.25  to  6.25  per 
a.).  Island  populations  of  eastern  cottontails  in  Mary- 
land reached  maximum  density  levels  of  10.2  per 
ha  (25.5  per  a.)  in  December  (See  Chapman  et  al. 
1982  for  references). 

Snowshoe  hare  populations  exhibit  a  marked 
variation  in  density  levels.  During  low  population 
cycles,  densities  may  vary  from  0.13  to  0.26  per  ha 
(0.32  to  0.65  per  a.)  to  5.9  to  11.8  per  ha  (14.7 
to  19. 5  per  a.)  in  a  high  cycle  in  Alberta  (Keith  and 
Windberg  1978).  Continent-wide,  peaks  in  hare  pop- 
ulations have  been  reported  as  high  as  38.6  per  ha 
(96.5  per  a.)  to  lows  of  0.12  per  ha  (0.3  per  a.)  (See 
Bittner  and  Rongstand  1982  for  references). 

Pika  densities  fluctuate  greatly,  ranging  from  2 
to  3  per  ha  ( 5  to  7  per  a. )  to  as  high  as  70  to  80  per 
ha  (175  to  200  per  a.)  (Kawamichi  1984). 


Pika. 


DISCUSSION 

In  this  chapter  we  have  defined  the  major  plant 
communities  that  are  associated  with  the  various 
genera  and  species  of  Lagomorphs  found  in  the  con- 
tinental U.S.  and  Alaska.  We  have  also  discussed 
some  of  the  methods  used  to  estimate  Lagomorph 
population  sizes.  However,  to  understand  the  status 
of  a  Lagomorph  population  in  a  particular  location, 
season,  or  year,  one  must  know  a  considerable 
amount  about  the  relationship  between  population 
size  and  the  habitat. 

The  relationship  between  Lagomorphs  and  their 
habitats  is  complex.  The  factors  involved  include 
those  inherent  in  the  physiology  of  the  individual 
species  and  those  edaphic  factors  external  to  the 


468 


Lagomorphs 


species  such  as  weather,  topography,  and  vegetation. 
The  factors  which  we  will  consider  in  this  discussion 
include  weather,  vegetation,  reproduction,  food  hab- 
its, predation,  and  their  interrelations  with  Lago- 
morph  population  dynamics  and  habitat  use. 


The  population  dynamics  of  Lagomorphs  has 
been  the  subject  of  considerable  research  and  nu- 
merous publications,  particularly  for  the  snowshoe 
hare  (Keith  1979)  and  eastern  cottontail  (Edwards 
et  al.  1979).  The  population  dynamics  of  a  species 


has  been  defined  as  "numbers,  structural  and  distri- 
butional changes  in  population  with  respect  to 
changes  in  the  rate  of  reproduction,  survival,  move- 
ments, and  environmental  interactions,  which  in  part 
determine  these  rates"  (Keith  1979).  Much  has  been 
written  on  the  cyclic  nature  of  snowshoe  hare  popu- 
lations and  the  random  fluctuations  in  cottontail 
and  jackrabbit  populations.  We  define  cyclic  as  a  reg- 
ular predictable  phenomenon,  such  as  the  10-year 
cycle  reported  in  snowshoe  hares  at  northern  lati- 
tudes (Keith  1974).  Population  fluctuations  are  ran- 
dom sporadic  increases  and  decreases  occurring 


Table  4.     Food  of  cottontail  rabbits  (Sylvilagus)  in  Utah. 


Audubon's  Cottontail 

Nuttall's  Cottontail 

Eastern  Utah 

Saltgrass  {Distichlis  stricta) 

Galleta  grass  (Hilaria  sp.) 

Indian  ricegrass 

Sand  dropseed 

Alkali  sacaton  (S.  airoides) 

Blue  grama 

Downy  brome 

Western  wheatgrass 

Spike  rush  {Eleocharis  sp.) 

Sedge 

Wire  grass  (Juncus  balticus) 

Pigweed  (Chenopodium  sp.) 

Saltbush  (A.  nuttallii, 

A.  corrugata,  A.  wolfii, 

A.  powelli) 
Greasewood 
Seepweed  (Dondia  sp.) 
Russian  thistle 

Tansymustard  (Descuraiana  sp.) 
Yellow  cleome  (Cleome  lutea) 
Wild  rose 

Sand  bur  (Glycyrrhiza  lopidota) 
Globe  mallow 
Prickly  pear 

Dogbane  (Apocynum  cannabinum) 
Rabbitbrush 
Sagebrush 
Horsebrush 

Northeastern  Utah 

Wheat  grass 

Willows 

Knotweed  (Polygonum  sp.) 

Spiny  hopsage 

Oregon  grape  (Berberis  repens) 

Bitterbrush 

Wild  rose 

Chokecherry 

Snakeweed 

Rabbitbrush 

Sagebrush 

Pygmy  Rabbit 

Northeastern  Utah 

Rabbitbrush 

Little  rabbitbrush  (C. 

viscidiflorus) 
Horsebrush 
Russian  thistle 
Pigweed 
Smooth  wheatgrass  (Agropyron 

inerme) 
Spiked  wheatgrass  (A.  spicatum) 
Giant  wild  rye  (Elymus  condensatus) 
Small  scale  (Atriplex  pusilla) 
Four-winged  saltbush 
Western  wheatgrass 
Sagebrush 
Globe  mallow 
Snakeweed 

Western  Utah 

Galleta  grass 

Alkali  sacaton 

Saltgrass 

Four-winged  saltbush 

Mound  saltbush 

Winterfat 

Finchook  bassia  (Bassia 

hyssopi :  folia) 
Pigweed  (Amaranthus  sp.) 
Purple  cleome 
Snakeweed 
Sagebrush 

Source:  (Janson  1946) 


Lagomorphs 


469 


on  an  irregular  basis,  caused  by  man-induced 
changes  in  habitat  or  uncontrolled  changes  in  envi- 
ronmental factors  such  as  weather  (Edwards  et  al. 
1979).  Fluctuations  in  Lagomorph  population  levels 
may  occur  annually  or  seasonally.  Several  models 
have  been  developed  to  explain  the  cause  of  cycles 
in  Lagomorph  populations,  but  most  are  based  on 
Keith's  (1974)  model  of  habitat-hare-predation  inter- 
action (Bryant  1979,  1981;  Wolff  1980a).  Keith's 
(1974)  model  is  based  on  regular  increases  and  de- 
creases in  population  numbers  of  snowshoe  hares 
with  respect  to  changes  in  survival,  reproduction, 
and  movement  of  this  species  (Keith  1979).  Simi- 
larly, regular  cycles  have  been  reported  in  Utah  jack- 
rabbits  (Gross  et  al.  1974). 

The  effects  of  environmental  factors  on  Lago- 
morph populations  are  difficult  to  assess  and  quan- 
tify. However,  environmental  factors  such  as  weather 
are  believed  to  have  an  impact  on  the  regularity  of 
Lagomorph  population  fluctuations.  Weather  varia- 
bles such  as  snowfall  and  rainfall  may  have  a  major 
impact  on  the  early  survival  of  young  cottontails 
(Applegate  and  Trout  1976;  Havera  1973).  In 
snowshoe  hares,  the  depth  of  the  snowfall  can  deter- 
mine the  suitability  of  winter  habitat.  Optimal  habitat 
for  snowshoe  hares  in  the  montane  forests  of  north- 
ern Utah  had  an  understory  cover  between  1.0  and 
2.5  m  (3  and  8.25  ft)  high  (Wolff  et  al.  1982). 

Temperature  and  snow  depth  are  correlated 
with  the  size  of  snowshoe  hare  litters  (Meslow  and 
Keith  1971).  Similarly,  cold,  inclement  weather  af- 
fects the  onset  of  reproduction  and  fetal  survival  in 
Leporids  (Chapman  et  al.  1977). 

The  onset  and  duration  of  the  breeding  season 
may  be  regulated  by  certain  weather  factors.  The 
length  of  photoperiod  is  associated  with  earlier  onset 
of  the  breeding  season  in  snowshoe  hares  ( Meslow 
and  Keith  1971 )  and  in  cottontails  (Bissonnette  and 
Csech  1939).  The  length  of  the  breeding  season 
for  Nuttall's  cottontails  was  shortened  when  precipi- 
tation was  reduced.  Consequently,  population  densi- 
ties were  lower  in  years  of  low  precipitation. 
Densities  in  late  summer  reached  a  high  of  2.54  cot- 
tontails per  ha  (6.35  per  a.)  when  precipitation  was 
11.54  cm  (4.6  in.);  in  contrast,  population  levels 
dropped  to  0.66  per  ha  (1.65  per  a.)  when  precipita- 
tion was  5.21  cm  (208  in.)  the  following  year  (Verts 
et  al.  1984). 

Reproduction  is  an  important  factor  affecting 
population  cycles.  Keith's  (1974)  model  attributed 
decreases  in  fecundity  from  small  litter  sizes,  shorter 
breeding  seasons,  and  lower  pregnancy  rates,  due 
at  least  in  part  to  the  regulation  of  hare  populations. 
The  onset  of  the  breeding  season  has  been  corre- 
lated with  midwinter  photoperiod  (Meslow  and 


Keith  1971).  Litter  sizes  are  larger  in  areas  where 
days  are  longer;  hence  latitude  is  a  factor  in  repro- 
ductive output  (Keith  1979).  Litter  sizes  were  also 
larger  in  hares  when  over-winter  weight  loss  did  not 
occur  (Keith  and  Windberg  1978),  pointing  to  the 
importance  of  adequate  high  quality  browse.  There 
is  also  considerable  variation  in  the  size  of  first  and 
subsequent  litters  in  several  species  of  Leporids 
(Chapman  et  al.  1977;  Bittner  and  Rongstand  1982). 
Leporid  litter  sizes  also  vary  with  latitude  (Conaway 
et  al.  1974;  Dunn  et  al.  1982).  Similarly,  there  is  a 
correlation  with  latitude  and  length  of  gestation  pe- 
riod in  North  American  rabbits.  Those  species  occur- 
ring at  more  northern  latitudes  have  shorter 
gestation  periods  (Chapman  1984). 

Lagomorphs  are  important  in  many  food  chains, 
being  prey  items  for  many  avian  and  terrestrial 
predators  (Bittner  and  Rongstand  1982;  Chapman  et 
al.  1982;  Dunn  et  al.  1982).  The  selection  of  refugia 
or  escape  cover  from  predators  during  snowshoe 
hare  population  "lows"  was  considered  of  major 
importance  in  their  survival.  Refugia  for  snowshoe 
hares  in  Alaska  were  dense  thickets  of  alder  or  wil- 
low (Wolff  1980a).  Hare  vulnerability  to  predation 
increases  as  populations  rise,  forcing  individuals  to 
occupy  suboptimal  areas.  In  Utah,  these  areas  were 
generally  open,  where  the  vegetation  density  profile 
was  less  then  40%  above  the  snowline  (Wolff  et  al. 
1982).  Coyote  predation  on  jackrabbits  was  greatest 
when  jackrabbit  densities  were  at  low  or  moderate 
levels  (Wagner  and  Stoddart  1972).  Even  in  areas 
with  abundant  plant  food  resources,  cover  is  the 
most  important  habitat  component  (Wolff  1980a). 
Keith's  (1974)  model  includes  predation  as  a  factor 
in  the  population  regulation  of  snowshoe  hares,  par- 
ticularly in  northern  latitudes.  Thus  escape  cover  is 
an  essential  component  of  Lagomorph  habitat. 

Lagomorphs  eat  a  wide  variety  of  plants.  How- 
ever, plant  resins  have  been  identified  as  a  deterrent 
in  Lagomorph  herbivory  (Bryant  1979,  1981).  These 
resins  have  been  suggested  as  a  key  to  hare  popula- 
tion regulation  (Bryant  1979).  This  idea  is  new  and 
requires  further  verification  before  it  will  be  widely 
accepted. 

The  diameter  of  the  twigs  consumed  and  brows- 
ing intensity  has  been  used  as  an  indicator  of  "food 
stress"  in  habitat  utilized  in  excess  of  carrying  capac- 
ity (Wolff  1980a).  Hares  will  starve  in  winter  if 
forced  into  habitats  where  the  diameter  of  twigs 
used  for  browse  was  greater  than  3  mm  (0.12  in.) 
(Pease  et  al.  1979).  Clark  and  Wagner  (1984)  ob- 
served that  only  20%  of  a  browse  plant  "brown 
sage"  (Kochla  americana)  was  consumed  when  jack- 
rabbit  populations  were  at  their  highest  level.  Heavy 
grazing  by  sheep  in  addition  to  the  limited  foraging 
by  hares  caused  the  plants  to  decline.  Food  is  appar- 
ently an  important  factor  in  hare  population  regula- 
tion (Keith  1974).  As  hare  densities  increase,  the 


470 


Lagomorphs 


biomass  of  available  food  is  less  than  that  needed  by 
the  population,  which  ultimately  leaves  a  food  short- 
age and  thus  causes  mortality  in  hare  populations. 

The  relationship  of  Lagomorph  populations  to 
their  habitat  is  evident  from  studies  on  snowshoe 
hares,  jackrabbits,  cottontails,  and  pikas.  Periods  of 
drought  forced  jackrabbits  to  occupy  cultivated 
fields,  which  provided  an  abundant  food  source  and 
supported  higher  jackrabbit  population  levels  (Bron- 
son  and  Tiemeier  1959).  Similarly,  overgrazing  by 
cattle  in  conjunction  with  periods  of  drought  cre- 
ated more  open,  weedy  areas,  a  habitat  which  also 
supported  higher  jackrabbit  populations  (Taylor 
et  al.  1935).  In  the  Mojave  desert,  jackrabbit  popula- 
tion densities  were  higher  around  water  sources  in 
winter,  declining  substantially  from  spring  to  sum- 
mer (Hayden  1966).  Thus,  changes  in  jackrabbit 
populations  were  affected  by  food  availability  and 
changes  in  environmental  conditions  brought  on  by 
factors  such  as  drought.  Changes  in  cottontail  popu- 
lation levels  have  also  been  attributed  to  changing 
land-use  patterns  (Edwards  et  al.  1979).  In  pikas,  size 
of  territories  is  determined  at  least  in  part  by  vegeta- 
tion quality  (Smith  1979).  Thus,  the  higher  the  qual- 
ity of  vegetation,  the  more  pikas  an  area  will 
support. 


Habitat  patchiness  or  the  mosaic  pattern  of  veg- 
etation within  a  Lagomorph's  environment  plays  a 
key  role  in  the  survival  of  snowshoe  hares  during  a 
population  low  (Wolff  1980a).  All  habitat  types  from 
open  suboptimal  areas  to  optimal  habitat  types  of 
dense  cover  were  occupied  by  snowshoe  hares  dur- 
ing population  highs.  However,  during  lows,  the 
remaining  individuals  in  a  population  tended  to  seek 
refuge  in  dense  thickets  of  spruce,  willow,  or  alder. 
In  winters  of  high  densities,  hares  in  Alaska  utilized 
more  open  stands  and  consumed  all  of  the  available 
browse  (Wolff  1979,  1980b). 


It  is  apparent  that  the  inventory  of  Lagomorph 
habitat  must  take  into  account  many  factors.  Most  of 
these  factors  are  poorly  understood  or  known  only 
for  a  particular  genus  or  species.  This  is  especially 
true  for  the  pika  and  Alaskan  hare  which  remain 
among  the  least  known  Lagomorphs.  In  addition, 
there  are  little  quantitative  data  on  habitat  variables 
available  for  most  North  American  Lagomorphs.  This 
lack  of  quantitative  data  makes  it  difficult  to  assess 
changes  in  Lagomorph  abundance  in  relation  to  habi- 
tat quality. 


Table  5.     Lagomorph  species  groups,  reproductive  characteristics,  and  common  habitat  requirements. 


Reproductive 

Family 

Species 

Characteristics 

Habitat 

Leporidae 

Black-tailed  jackrabbit  (Lepus 

Precocial 

Desert  scrub  and 

californicus) 

dry  grasslands 

White-tailed  jackrabbit 

(L  townsendii) 

Antelope  jackrabbit 

(L  alleni) 

White-sided  jackrabbit 

(L.  callotis) 

Leporidae 

Snowshoe  hare  (L  americanus) 

Precocial 

Coniferous  forests, 

Alaskan  hare  (L.  othus) 

subalpine  and 

(Some  workers  feel  L.  othus  is  a 

climax  forest 

synonym  for  L.  timidus  [Dixon  et  al. 

1983]). 

Leporidae 

Audubon's  cottontail  {Sylvilagus 

Altricial 

Disturbed  and 

audubonii) 

successional  plant 

Nuttall's   cottontail  (S.  nuttallii) 

communities 

Brush  rabbit  (S.  bachmani) 

Eastern  cottontail  (S.  floridanus) 

Pygmy  rabbit  (S.  idahoenis) 

(Some  workers  consider  the  pigmy 

rabbit  to  be  in  the  genus  Brachylagus 

[Chapman  et  al.  1982]). 

Ochotonidae 

Southern  pika  {Ochotonidae  princeps) 

Altricial 

Rocky  talus  at  high 

(Some  workers  consider  the  collared 

altitudes 

pika  [0.  p.  collaris]  of  Alaska  a 

separate  species  [Chapman  1979;  Hall 

1981]) 

Lagomorphs 


471 


LITERATURE  CITED 


— ,  J.G.  HOCKMAN,  and  W.R.  EDWARDS.  1982.  Cot- 
tontails. Pages  83-123  in  Chapman,  J.A.  and  G.A.  Feld- 
hamer  eds.,  Wild  Mammals  of  North  America.  Johns 
Hopkins  Univ.  Press.  Baltimore,  MD. 

and  MM.  OJEDA.  1980.  Sylvilagus  flori- 


ADAMS,  L.  1959.  An  analysis  of  a  population  of  snowshoe 
hares  in  northwestern  Montana.  Ecol.  Monogr. 
29:141-170. 

APPLEGATE,  J.E.  and  JR.  TROUT.  1976.  Weather  and  the 
harvest  of  cottontails  in  New  Jersey.  J.  Wildl.  Manage. 
40:658-662. 

ASSERSON,  W.C.  1967.  Upland  Game  Investigations:  Cot- 
tontail rabbit  investigations.  California  PR  Rep.  W-47- 
R  15.  12pp. 

BEAR,  G.D.  and  R.M.  HANSEN.  1966.  Food  habits,  growth, 
and  reproduction  of  white-tailed  jackrabbits  in  south- 
ern Colorado.  Colorado  State  Univ.,  Agric.  Exp.  Sta. 
Tech.  Bull.  90.  39pp. 

BEDNARZ,  J.  1977.  The  white-sided  jackrabbit  in  New 
Mexico:  distribution,  numbers  and  biology  in  the 
grasslands  of  Hidalgo  County.  Res.  Rep.  New  Mexico 
Dep.  Game  and  Fish,  Endang.  Sp.  Program,  Santa  Fe. 
33pp. 

BISSONNETTE,  T.H.  and  AG.  CSECH.  1939.  Modified 
sexual  periodicity  in  cottontail  rabbits.  Biol.  Bull. 
17:364-367. 

BITTNER,  S.L.  and  O.J.  RONGSTAND.  1982.  Snowshoe 
hare  and  allies.  Pages  146-163  in  Chapman,  JA.  and 
GA.  Feldhamer  eds.,  Wild  Mammals  of  North  Amer- 
ica. Johns  Hopkins  Univ.  Press.  Baltimore,  MD. 
1147pp. 

BRAUN,  C.E.  and  R.G.  STREETER.  1968.  Observations  on 
the  occurrence  of  the  white-tailed  jackrabbit  in  the 
alpine  zone.  J.  Mammal.  49:160-161. 

BROADBOOKS,  HE.  1965.  Ecology  and  distribution  of 
the  pikas  of  Washington  and  Alaska.  Am.  Midi.  Nature. 
73:299-335. 

BRONSON,  F.H.  and  O.W.  TIEMEIER.  1959.  The  relation- 
ship of  precipitation  and  black-tailed  jackrabbit  popu- 
lations in  Kansas.  Ecology.  40:194-198. 

BRYANT,  J.P.  1979.  The  regulation  of  snowshoe  hare 

feeding  behavior  during  winter  by  plant  antiherbivore 
chemistry.  Pages  720-731  in  Proc.  World  Lagomorph 
Conf. 

.  1 98 1 .  Phytochemical  deterrence  of  snowshoe  hare 

browsing  by  adventitious  shoots  of  four  Alaskan  trees. 
Science.  213:889-890. 

CHAPMAN,  JA.  1971.  Orientation  and  homing  of  the 
brush  rabbit  (Sylvilagus  bachmani).  J.  Mammal. 
52:686-699. 

.  1 979.  Rabbits,  hares,  and  pikas.  Pages  8 1  -98  in 

Wild  Animals  of  North  America.  National  Geographic 
Soc,  Washington,  DC.  406pp. 

.  1984.  Latitude  and  gestation  period  in  New  World 

rabbits  (Leporidae:  Sylvilagus  and  Romerolagus). 
Am.  Nature.  124:442-445. 

and  G.A.  FELDHAMER  eds.  1982.  Wild  mammals  of 

North  America.  Johns  Hopkins  Univ.  Press.  Baltimore, 
MD.  1147pp. 

and  DEC.  TRETHEWAY.  1972.  Factors  affecting 

trap  responses  of  introduced  eastern  cottontail  rab- 
bits. J.  Wildl.  Manage.  36:1221-1226. 

,  A.L.  HARMAN,  and  D.E.  SAMUEL.  1977.  Reproduc- 
tive and  physiological  cycles  in  the  cottontail  com- 
plex in  western  Maryland  and  nearby  West  Virginia. 
Wildl.  Monogr.  56.  73pp. 


danns.  Mammalian  Species  136:1-8. 

CLARK,  F.W.  1972.  Influence  of  jackrabbit  density  on 
coyote  population  change.  J.  Wildl.  Manage.  36:343- 
356. 

CLARK,  W.R.  and  F.H.  WAGNER.  1984.  Role  of  livestock 
and  black-tailed  jackrabbits  in  changing  abundance  of 
Kochia  americana  Great  Basin  Naturalist.  44:635- 
645. 

CONAWAY,  C.H.,  KC.  SADLER,  and  D.H.  HAZELWOOD. 
1974.  Geographic  variation  in  litter  size  and  onset  of 
breeding  in  cottontails.  J.  Wildl.  Manage.  38:473-481. 

CONROY,  M.J.,  L.W.  GYSEL,  and  G.R.  DUDDERAR.  1979. 
Habitat  components  of  clear-cut  areas  for  snowshoe 
hares  in  Michigan.  J.  Wildl.  Manage.  43:680-690. 

DALQUIST,  W.W.  1948.  Mammals  of  Washington.  Univ.  of 
Kansas  Publ.  Mus.  Nat.  Hist.  2:1-444. 

DAVIS,  C.A.,  JA.  MEDLIN,  and  J.P.  GRJFFING.  1975.  Abun- 
dance of  black-tailed  jackrabbits,  desert  cottontail 
rabbits,  and  coyotes  in  southeastern  New  Mexico. 
New  Mexico  State  Univ.  Agric.  Exp.  Sta.  Res.  Rep. 
293. 

DAVIS,  D.E.  ed.,  1982.  Handbook  of  census  methods  for 
terrestrial  vertebrates.  CRC  Press,  Boca  Raton,  FL. 
397pp. 

and  R.L.  WINSTEAD.  1980.  Estimating  the  numbers 

of  wildlife  populations;  pages  221-245  in  Schemnitz, 
S.D.  ed.,  Wildl.  Manage.  Techniques  Manual.  Wildl. 
Soc,  Washington,  DC. 

DE  VOS,  A.  1964.  Food  utilization  of  snowshoe  hares  on 
Manitoulin  Island,  Ontario.  J.  For.  62:238-244. 

DICE,  L.R.  1926.  Notes  of  Pacific  Coast  rabbits  and  pikas. 
Occ.  Papers  Mus.  Zool.,  Univ.  of  Michigan,  Ann  Arbor. 
166:1-28. 

DIXON,  K.R,  JA.  CHAPMAN,  G.R.  WILLNER,  D.E.  WIL- 
SON, and  W.  LOPEZ-FORMENT.  1983.  The  New 
World  jackrabbits  and  hares  (genus  Lepus  L.  Part  2) — 
Numerical  taxonomic  analysis.  Acta.  Zool.  Fenica. 
174:53-56. 

DOLBEER,  R.A.  and  W.R.  CLARK  1975.  Population  ecol- 
ogy of  snowshoe  hares  in  the  central  Rocky  Moun- 
tains. J.  Wildl.  Manage.  39:535-549. 

DOUTT,  J.K,  C.A.  HEPPENSTALL,  and  J.E.  GUILDAY.  1967. 
Mammals  of  Pennsylvania.  Pennsylvania  Game  Comm. 
Harrisburg.  238pp. 

DUNN,  J.P,  JA.  CHAPMAN,  and  RE.  MARSH.  1982.  Jack- 
rabbits. Pages  124-145  in  Chapman,  JA.  and  GA. 
Feldhamer  eds.  Wild  Mammals  of  North  America. 
Johns  Hopkins  Univ.  Press,  Baltimore,  MD. 

EBERHARDT,  L.L.  1969.  Population  estimation  from  re- 
capture frequencies.  J.  Wildl.  Manage.  33:28-39. 

EDWARDS,  W.R.  1975.  Rabbit  Trap.  Illinois  Nat.  Hist. 
Survey,  Urbana,  IL.  1pp. 

,  SP.  HA  VERA,  R.F.  LABISKY,  JA.  EILLES,  and  RE. 

WARNER.  1979.  The  abundance  of  cottontails  in  rela- 
tion to  agriculture  land  use  in  Illinois  (U.S.A.)  1956- 
1978,  with  comments  on  mechanism  and  regulation. 
Pages  761-789  in  Proc.  World  Lagomorph  Conf. 

FEIST,  D.D.  1982.  Snowshoe  hare  (Alaska).  Page  139  in 
Davis,  D.E.  ed.  Handbook  of  Census  Methods  for 
Terrestrial  Vertebrates.  CRC  Press,  Boca  Raton,  FL. 

FLINDERS,  J.T  and  R.M.  HANSEN.  1972.  Diets  and  habi- 
tats of  jackrabbits  in  northeastern  Colorado.  Colorado 


472 


Lagomorphs 


State  Univ.,  Range  Sci.  Dep.  Sci.  Ser.,  Fort  Collins. 

12:1-29. 

—  and  —       — .  1973-  Abundance  and  dispersion  of 


leporids  within  a  shortgrass  ecosystem.  J.  Mammal. 
54:287-291. 

GATES,  C.E.  1969.  Simulation  study  of  estimates  for  the 
line  transect  sampling  method.  Biometrics.  25:317- 
328. 

GRANGE,  W.B.  1965.  Fire  and  tree  growth  relationships  to 
snowshoe  rabbits.  Proc.  Tall  Timbers  Fire  Ecol.  Conf. 
4:111-125. 

GROSS,  J.E.,  L.C.  STODDART,  and  F.H.  WAGNER.  1974. 
Demographic  analysis  of  a  northern  Utah  jackrabbit 
population.  Wildl.  Monogr.  40.  68pp. 

HALL,  E.R.  1981.  The  mammals  of  North  America,  Vol.  I. 
John  Wiley  and  Sons,  New  York,  NY.  600pp. 

HANSEN,  R.M.  and  JT.  FLINDERS.  1969.  Food  habits  of 
North  American  hares.  Colorado  State  Univ.,  Range 
Sci.  Dep.  Sci.  Serv.  Fort  Collins.  31:1- 18.. 

HAVERA,  S.P.  1973.  The  relationship  of  Illinois  weather 
and  agriculture  to  the  eastern  cottontail  rabbit.  Illi- 
nois State  Water  Surv.  Tech.  Rep.  4.  92pp.  (Mimeo.). 

HAYDEN,  P.  1 966.  Seasonal  occurrence  of  jackrabbits 

on  Jackass  Flat,  Nevada.  J.  Wildl.  Manage.  30:835-838. 

JANSON,  R.G.  1946.  A  survey  of  the  native  rabbits  of  Utah 
with  reference  to  their  classification,  distribution, 
life  histories  and  ecology.  Unpubl.  M.S.  Thesis,  Utah 
State  Agric.  College.  103pp. 

JOLLY,  G.M.  1963.  Estimates  of  population  parameters 

from  multiple  recapture  data  with  both  death  and  di- 
lution-deterministic model.  Biometrika.  50:113-128. 

KAWAMICHI,  T.  1984.  Pikas.  Pages  726-727  in  Macdon- 
ald,  D.  ed.  The  Encyclopedia  of  Mammals.  Facts  on 
File  Publ.  New  York,  NY. 

KEITH,  LB.  1974.  Some  features  of  population  dynamics 
in  mammals.  Proc.  Int.  Congr.  Game  Biol.  11:17-58. 

.  1979-  Population  dynamics  of  hares.  Pages  395- 

440  in  Proc.  World  Lagomorph  Conf. 

and  LA.  WINDBERG.  1978.  A  demographic  analysis 


of  the  snowshoe  hare  cycle.  Wildl.  Monogr.  58.  70pp. 

KUNDAELI,  J.N.  and  H.G.  REYNOLDS.  1972.  Desert  cot- 
tontail use  of  natural  and  modified  pinyon-juniper 
woodland.  J.  Range  Manage.  25:1 16-1 18. 

LABISKY,  R.F.  1959.  Nightlighting:  a  technique  for  captur- 
ing birds  and  mammals.  Illinois  Nat.  History  Survey 
Div.  Biol.  Notes.  11pp. 

.  1968.  Nightlighting:  Its  use  in  capturing  pheasants, 

prairie  chickens,  bobwhites,  and  cottontails.  Illinois 
Nat.  History  Survey  Div.  Biol.  Notes.  62.  12pp. 

LLNCOLN,  F.C  1930.  Calculating  waterfowl  abundance  on 
the  basis  of  banding  returns.  U.S.  Dep.  Agric.  Cir. 
118:1-4. 

LORD,  R.D.  1963-  The  cottontail  rabbit  in  Illinois.  Illinois 
Dep.  Conserv.  Tech.  Bull.  3:1-94. 

McKAY,  DO.  and  B.J.  VERTS.  1978.  Estimates  of  some 
attributes  of  a  population  of  Nuttall's  cottontails.  J. 
Wildl.  Manage.  42:159-168. 

MESLOW,  EC.  and  LB.  KEITH.  1971.  A  correlation  analy- 
sis of  weather  versus  snowshoe  hare  population  pa- 
rameters. J.  Wildl.  Manage.  35:1-14. 

ORR,  R.T.  1940.  The  rabbits  of  California.  Occ.  Papers 
California  Acad.  Sci.  19.  227pp. 

OTIS,  D.L.,  KP.  BURNHAM,  G.C  WHITE,  and  DR.  ANDER 


SON.  1978.  Statistical  inference  from  capture  data 
on  closed  animal  populations.  Wildl.  Monogr.  63:1- 
135. 

PEASE,  J.L.,  R.H.  VOWLES,  and  LB.  KEITH.  1979.  Interac 
tion  of  snowshoe  hares  and  woody  vegetation.  J. 
Wildl.  Manage.  43:43-60. 

RADWAN,  MA.  and  D.L.  CAMPBELL.  1968.  Snowshoe  hare 
preference  for  spotted  catsear  flowers  in  western 
Washington.  J.  Wildl.  Manage.  32:104-108. 

SHIELDS,  P.W.  1958.  Ecology  and  population  dynamics  of 
the  brush  rabbit  of  the  north  spit  of  Humboldt 
County,  California.  Unpublished  M.S.  Thesis.  Hum- 
boldt State  College.  98pp. 

SMITH,  A.T.  1974a.  The  distribution  and  dispersal  of  pikas: 
Influences  of  behavior  and  climate.  Ecology.  55:1368- 
1376. 

.  1974b.  The  distribution  and  dispersal  of  pikas: 

Consequences  of  insular  population  structure.  Ecol- 
ogy. 55:1112-1119. 

.  1979.  Population  dynamics  of  pikas.  Pages  572-586 


in  Proc.  World  Lagomorph  Conf. 

1982.  Pika(Ochotona).  Pages  131-133  in  Davis, 


D.E.  ed.  Handbook  of  Census  Methods  for  Terrestrial 
Vertebrates.  CRC  Press,  Boca  Raton,  FL 

SWIHART,  RK  and  R.H.  YAHNER.  1984.  Winter  use  of 
insular  habitat  patches  by  the  eastern  cottontail.  Acta 
Theriologica.  29:45-56. 

TAYLOR,  W.P.  1930.  Methods  of  determining  rodent  pres- 
sure on  the  range.  Ecology.  11:523-542. 

,  C.T.  VORHEIS,  and  P.B.  LISTER.  1935.  The  relation 

of  jackrabbits  to  grazing  in  southern  Arizona.  J.  For. 
33:490-498. 

VERTS,  B.J.,  S.D.  GEHMAN,  and  KJ.  HUNDERTMARK 

1984.  Sylvilagus  nuttallii:  a  semiarboreal  Lagomorph. 
J.  Mammal.  65(  1  ):1 31-135. 

VORHEIS,  C.J.  and  W.P.  TAYLOR.  1933.  The  life  histories 
and  ecology  of  the  jackrabbits  ( Lepus  alleni  and  L 
californicus )  in  relation  to  grazing  in  Arizona.  Univ.  of 
Arizona  Agric.  Exper.  Stat.  Tech.  Bull.  49:1-117. 

WAGNER,  F.H.,  and  L.C.  STODDART.  1972.  Influence  of 
coyote  predation  on  black-tailed  jackrabbit  popula- 
tions in  Utah.  J.  Wildl.  Manage.  25:242-329- 

WEBB,  W.L.  1942.  Notes  on  a  method  for  censusing 

snowshoe  hare  populations.  J.  Wildl.  Manage.  42:159- 
168. 

WOLFF,  ML,  N.V.  DeBYLE,  C.S.  WINCHELL,  and  T.R. 

McCABE.  1982.  Snowshoe  hare  cover  relationships  in 
northern  Utah.  J.  Wildl.  Manage.  46:552-670. 

WOLFF,  J.O.  1978.  Food  habits  of  snowshoe  hares  in  inte- 
rior Alaska.  J.  Wildl.  Manage.  42:148-153- 

.  1979.  Refugia,  dispersal,  predation,  and  geographic 

variation  in  snowshoe  hare  cycles.  Pages  441-449  in 
Proc.  World  Lagomorph  Conf. 

.  1980a.  The  role  of  habitat  patchiness  in  the  popu- 
lation dynamics  of  snowshoe  hares.  Ecol.  Monogr. 
50:111-130. 

.  1980b.  Moose-snowshoe  hare  competition  during 

peak  hare  densities.  Proc.  N.  Amer.  Moose  Conf.  & 
Workshop.  16:238-254. 

.  1982.  Snowshoe  hare.  Pages  141-149  in  Davis,  D.E. 

ed.  Handbook  of  Census  Methods  for  Terrestrial  Ver- 
tebrates. CRC  Press,  Boca  Raton,  FL. 


Lagomorphs 


473 


23 

CARNIVORES 


Richard  A.  Spowart   and  Fred  B.  Samson 

Colorado  Cooperative  Wildlife  Research  Unit 
Colorado  State  University 
Fort  Collins,  CO  80523 


Editor's  Note:  Carnivores  are  interesting,  elusive, 
and  controversial.  They  are  also  one  of  the  most 
difficult  groups  to  inventory  or  monitor.  Many 
studies  have  focused  on  prey  bases  for  carnivores, 
but  other  habitat  requirements  have  not  been  well 
defined. 

This  chapter  brings  together  much  of  the  exist- 
ing literature  on  carnivore  techniques,  and  should 
assist  in  study  design.  However,  the  biologist  will 
find  plenty  of  opportunity  for  innovation  when 
working  with  these  species. 


INTRODUCTION 

The  order  Carnivora  is  a  diverse  group  of  preda- 
tory mammals.  Trenchant  characteristics  of  the  order 
are  found  in  their  dentition,  which  indicates  diet. 
However,  many  carnivores  are  omnivorous  or  even 
herbivorous.  The  variability  in  diet  is  reflected  by 
the  large  array  of  carnivore  habitat  types  and  life- 
styles. Some  species  are  habitat  and  food  specialists, 
distributed  within  only  a  few  plant  communities; 
others  are  habitat  and  food  generalists,  widely  dis- 
tributed in  natural  and  disturbed  ecosystems. 

In  contrast,  the  very  adaptable  and  successful 
coyote  (Canis  latrans)  occupies  almost  every  habi- 
tat type  in  North  America,  whereas  the  endangered 
black-footed  ferret  (Mustela  nigripes)  depends  on 
undisturbed  prairie.  The  smallest  living  carnivore, 
the  least  weasel  (M.  nivalis),  weighs  about  45  g  ( 1.6 
oz);  the  largest  living  carnivore,  a  subspecies  of  the 
grizzly  or  brown  bear  ( Ursus  arctos ),  weighs  more 
than  700  kg  ( 1,550  lbs).  The  hyperactive  least  wea- 
sel feeds  almost  exclusively  on  mice  and  requires 
about  a  third  of  its  body  weight  in  food  per  day;  the 
brown  bear,  however,  is  omnivorous  and  hibernates 
for  about  half  the  year. 

Because  of  the  diverse  ways  carnivores  have 
evolved  to  exploit  their  environments,  we  chose  to 
treat  most  species  separately.  When  important  habi- 
tat features  are  similar  for  several  species,  we  com- 
bined their  descriptions.  Within  these  combined 
descriptions,  we  attempted  to  contrast  unique  habi- 
tat preferences  of  individual  species. 

We  have  not  attempted  to  describe  all  physical 
features  and  vegetation  types  that  occur  in  all  habi- 
tats of  species  that  are  generalists;  only  habitats  that 
support  the  densest  populations  or  are  considered 
optimum  are  discussed.  Carnivore  species  that  are 
rare  in  western  North  America,  either  because  they 
have  been  extirpated  from  much  of  their  original 
range  or  because  their  range  extends  only  into  the 

'Current  Address:  Colorado  Division  of  Wildlife.  Kremmling,  CO 
80459 
"Current  Address:  U.S.  Forest  Service,  Olympia,  WA  98195 


Carnivores 


475 


periphery  of  that  region,  are  also  not  discussed.  Cer- 
tain of  these  species  were  excluded  largely  because 
of  their  threatened  or  endangered  status.  Species  not 
included  are  arctic  fox  (Alopex  lagopus),  black- 
footed  ferret,  coati  (Nasua  nasua),  jaguar  {Felis 
onca),  jaguarundi  (F.  yagouaroundi),  margay  cat  (F. 
wiedi),  ocelot  (F.  pardalis),  polar  bear  (Ursus  mari- 
timiis),  and  red  wolf  (Cants  rufus).. 


> 

mm" 

fit 

■H  m  . 

>-     -- — ~~* m                             ■     \  ±-    ■'TriiniB 

Arctic  fox. 


Important  considerations  in  evaluating  habitat 
for  any  species  are  food  supply,  habitat  size,  and 
interspecific  competition.  Food  supply,  particularly 
the  availability  of  prey  species,  often  determines  the 
density  of  carnivore  populations.  Habitat  manipula- 
tions can  often  alter  prey  population  size.  Since  habi- 
tat manipulations  indirectly  influence  the  dynamics 
of  carnivore  populations,  we  included  a  brief  de- 
scription of  food  habits  of  each  species  or  species 
group.  Unfortunately,  there  is  little  information  on 
food  requirements  for  most  carnivores,  but  when 
available,  we  referenced  this  information.  Table  1 
summarizes  major  foods  and  habitats  of  carnivores. 

Small  populations  are  more  likely  to  become 
extinct  than  large  populations.  A  major  influence  in 
population  size  is  the  availability  of  habitat.  The  size 
or  area  of  habitat  must  be  large  enough  to  support 
self-sustaining  populations  of  carnivores.  The  number 
of  prey  species  in  all  likelihood  will  also  increase  as 
habitat  size  increases.  This  habitat  feature  is  likely 
most  critical  to  those  carnivores  that  are  territorial, 
have  large  home  ranges,  and  exploit  a  patchily  dis- 
tributed food  supply. 


boreal  and  mixed-hardwood  forests  and  tundra.  In  all 
habitats  occupied  by  wolves,  either  past  or  present, 
availability  of  their  principal  prey,  large  ungulates, 
likely  has  most  influenced  wolf  distribution  and 
density  where  wolf  control  programs  have  been 
absent.  The  primary  objective  of  many  studies  of 
wolf  ecology  in  Alaska,  Canada,  Michigan,  and 
Minnesota  has  been  to  determine  the 
interrelationships  of  wolves  and  large  ungulates 
(Murie  1944;  Mech  1966;  Pimlott  1967;  Kolenosky 
1972;  Kuyt  1972;  Van  Ballenberghe  et  al.  1975; 
Peterson  1977;  Fuller  and  Keith  1980;  Hollemann 
and  Stephenson  1981;  Gasaway  et  al.  1983). 
Difficulties  in  measuring  wolf  and  prey  densities  and 
predation  rates,  as  well  as  the  influence  of  other 
forms  of  wolf  and  prey  mortality,  have  left  many 
questions  unanswered.  Disparate  results  have  been 
reported  for  daily  food  consumption  rates  of  wolves 
and,  more  importantly,  the  rates  at  which  wolves  kill 
prey  (Mech  1966;  Pimlott  1967;  Kolenosky  1972). 
In  northeastern  North  America,  where  simple 
predator-prey  systems  exist,  wolves  are  a  major 
mortality  factor  for  prey  populations  (Mech  1966; 
Jordan  et  al.  1967).  In  addition,  declining  wolf 
populations  have  been  attributed  primarily  to 
decreasing  numbers  of  their  primary  prey  (Jordan  et 
al.  1967;  Mech  1977;  Mech  and  Karns  1977). 
However,  the  role  wolf  predation  plays  in  more 
complex  predator-prey  systems  in  Alaska  and  the 
Rocky  Mountain  national  parks  of  western  Canada  is 
unclear  (Murie  1944;  Cowan  1947;  Gasaway  et  al. 
1983). 

Inventorying  and  monitoring  habitat  in  areas  of 
the  continental  U.S.,  where  wolves  have  been  extir- 
pated, are  important  to  feasibility  and  evaluation 
studies  of  wolf  reintroductions.  Most  attempts  to  re- 
establish wolves  in  areas  of  their  original  range  have, 
at  best,  met  limited  success  (Mech  1966;  Henshaw 
and  Stephenson  1974;  Henshaw  et  al.  1975;  Weise  et 
al.  1975).  Many  believe  that  large  areas,  having  no 
human  habitation,  are  necessary  to  reestablish 
wolves  because  of  their  large  home  ranges  and  social 
structure  (reviewed  by  Henshaw  1975;  Henshaw  et 
al.  1975;  Mech  1975;  Brown  1983).  However,  in 
Europe  and  some  areas  of  Alaska,  wolves  live  near 
humans  (Peterson  1975;  Pulliainen  1975;  Zimen  and 
Boitani  1975;  Gasaway  et  al.  1983).  The  density  of 
prey  populations  over  a  large  area  may  be  the  most 
important  habitat  requisite  determining  feasibility 
of  reintroduction. 


HABITAT  FEATURES  CORRELATED  WITH 
SPECIES  GROUPS 

Canidae 

Wolves.  Today,  substantial  populations  of  gray  or 
timber  wolves  (Cants  lupus)  occur  in  wilderness 
areas  and  large  national  parks  in  Alaska,  Canada,  and 
northern  Minnesota  and  Michigan.  Here  they  inhabit 


Coyotes.  The  coyote  has  survived  relentless 
predator-control  programs  that  extirpated  the  wolf 
in  most  of  the  continental  U.S.  Moreover,  this 
ultimate  opportunist  has  expanded  its  range  (Gier 
1975;  Andrews  and  Boggess  1978;  Berg  and 
Chesness  1978;  Hilton  1978).  Coyotes  evolved  in  a 
plains  environment  and  were  once  most  numerous 
in  western  grasslands  where  large  ungulate 
populations  were  densest.  They  flourished  in  the 


476 


Carnivores 


Table  1.     Summary  of  major  foods  and  habitats  for  carnivores. 


1 

1 

Carnivore 
Species 

Food 

Habitat 

c 
o 

CO 

O 

<f> 
0) 
CO 
3 
O) 

c 
3 

en 
Q. 
O 

E 
o 

D) 
CO 

_l 

CO 

c 

CD 
"O 
O 

rr 

"O   en 
c    O) 
CO   CD 
co   *> 
"D  "p 

CO  £3 

E<2 

«    CO 

£5 

=  sz 
Q.Q- 
CD    E 
DC    CO 

CO 
o 

CD 

en 

C 

03 

*^    CO 

2  ■- 

-  CD 

CD  "O 
CD    C 
CO    CO 

CO 

Li- 

"O 
CO 

E 

CO 

LL 

C 
CO 
CO 
CO 
CO 

o 

SZ 
CO 

Z) 

CD 

O) 

CO 
C/) 

to 
CD 

o 

CD 

o 

CD 

tf> 

o 

§^ 

CO 
O 

is 

So 

CO 

c 

CO 

CO 
Q. 
CO 

sz 
O 

CO 

CD 
CO 
CD 

Q 

CO 

"O 

c 

O    co 

Q-  q3 
-  > 

co  -c 

CO    c 

_l    CO 

Wolf 

• 

• 

• 

• 

• 

Coyote 

• 

• 

• 

• 

• 

• 

• 

• 

• 

• 

• 

• 

Gray  fox 

• 

• 

• 

• 

• 

Red  fox 

• 

• 

• 

• 

• 

• 

• 

• 

• 

Kit  fox 

• 

• 

Swift  fox 

• 

• 

Mountain  lion 

• 

• 

• 

Bobcat 

• 

• 

• 

• 

• 

• 

• 

Lynx 

• 

• 

Black  bear 

• 

• 

• 

Grizzly  bear 

• 

• 

• 

• 

Fisher 

• 

• 

• 

• 

Marten 

• 

Wolverine 

• 

• 

• 

Mink 

• 

• 

• 

• 

River  otter 

• 

• 

• 

Badger 

• 

• 

• 

Raccoon 

• 

• 

• 

• 

Ringtail 

• 

Striped  skunk 

• 

• 

• 

• 

• 

• 

• 

• 

• 

Hooded  skunk 

• 

• 

• 

• 

• 

Hognosed  skunk 

• 

• 

• 

• 

• 

• 

• 

Spotted  skunk 

• 

• 

• 

• 

• 

• 

Least  weasel 

• 

• 

• 

Short-tailed 
weasel 

• 

• 

Long-tailed 
weasel 

• 

• 

shortgrass-steppe,  semiarid  sagebrush  (Artemisia  sp. ) 
grasslands,  and  deserts.  They  ranged  from  deserts 
and  plains  to  alpine  areas  of  adjacent  mountains. 
Today,  range  extensions  indicate  that  coyotes  can  be 
successful  in  broken  forests,  from  the  tropics  of 
Guatemala  to  the  eastern  U.S.,  up  to  northern  Alaska. 
Altitude,  latitude,  and  vegetation  type  do  not  seem 
to  restrict  their  survival.  Range  expansions  and 
population  increases  were  likely  caused  by 
elimination  of  gray  wolves,  clearing  of  forests,  and 
agricultural  practices  (Carbyn  1982). 


The  coyote's  ability  to  expand  its  range  and 
adapt  to  so  many  habitat  types  is  partly  due  to  its 
versatile  food  habits.  Coyotes  are  primarily  carnivo- 
rous, but  their  diets  depend  on  the  food  resources 
most  available.  They  easily  adapt  to  being  omnivo- 
rous and,  in  this  respect,  their  food  habits  resemble 
those  of  foxes  (Green  and  Flinders  1981).  An  exten- 
sive food  habits  study  (Sperry  1941 )  conducted  in 
17  western  states  showed  that  major  diet  items  were 
lagomorphs  (33%  ),  carrion  (25%  ),  rodents  ( 18%  ), 
and  domestic  livestock  (135%). 


Carnivores 


477 


Studies  conducted  locally  in  the  western  U.S. 
revealed  large  variances  in  coyote  diets  from  one 
area  to  another.  Murie  (1951)  studied  coyote  food 
habits  on  Arizona  cattle  ranges.  The  ranges  consisted 
primarily  of  open  grasslands,  oak  (Quercus  sp. ),  juni- 
per (Juniperus  sp. ),  prickly  pear  and  cholla  (Opun- 
tia  sp. ),  and  ponderosa  pine  (Pinus ponderosa). 
In  these  habitats,  coyote  diets  were  complex  but 
contained  high  percentages  of  plant  material.  Juniper 
berries  were  particularly  important,  followed  by 
prickly  pear  fruits.  Berries  of  manzanita  (Arctosta- 
phylos  pungens  and  A  pringlei)  and  pods  of  velvet 
mesquite  {Prosopsis  juliflora)  were  rare  but  pre- 
ferred foods.  Other  important  diet  items  were  car- 
rion, grasshoppers,  and  rodents.  Lagomorphs  were 
found  in  only  3%  of  coyote  scats  but  were  rare  in 
this  area. 

Short  ( 1979)  found  similar  results  in  another 
area  of  Arizona,  described  as  a  semidesert,  grass- 
shrub  habitat,  dominated  by  velvet  mesquite,  cholla, 
and  prickly  pear.  Although  coyote  food  habits  were 
complex,  the  incidence  of  some  foods  in  diets  fol- 
lowed seasonal  patterns.  Coyote  food  habits  also 
exhibited  seasonal  trends  in  a  ponderosa  pine  forest 
of  Arizona  (Turkowski  1980).  In  contrast,  coyote 
diets  in  other  western  states  have  been  less  varied 
and  contained  primarily  mammals.  Clark  ( 1972) 
reported  that  coyotes  in  sagebrush  habitat  of  north- 
eastern Utah  and  south-central  Idaho  ate  about  90% 
animal  matter;  black-tailed  jackrabbits  (Lepus  califor- 
nicus)  approached  75%  of  their  year-round  diet. 
This  dependence  on  a  single  prey  species  influences 
coyote  density  in  this  region  (Clark  1972;  Wagner 
and  Stoddart  1972;  Johnson  and  Hansen  1979). 

In  northeastern  California,  meadow  voles  (Mi- 
crotus  sp.)  occurred  in  about  half  of  all  coyote  scats 
analyzed  (Hawthorne  1972).  Other  important  diet 
items  were  mule  deer  and  cattle,  probably  eaten  as 


carrion.  Mule  deer  were  also  important  in  coyote 
diets  in  two  areas  of  southern  Utah  ( Pederson  and 
Tuckneld  1983).  In  central  Wyoming,  mule  deer, 
pronghorn  antelope  (Antilocapra  americana), 
white-tailed  jackrabbits  (L.  townsendii),  and  desert 
cottontail  (Sylvilagus  audubonii)  were  present  in 
63%  of  coyote  scats  (Springer  and  Smith  1981). 

Although  coyotes  are  extremely  adaptable  to 
whatever  foods  are  available,  like  most  species  of 
carnivores,  their  populations  are  regulated  largely  by 
food  abundance  (Clark  1972;  Nellis  and  Keith  1976; 
Todd  and  Keith  1983).  Population  densities  of  coy- 
otes and  their  principal  prey  are  strongly  correlated. 
Food  availability  appears  to  have  a  strong  influence 
on  litter  size  and  survival  but,  as  yet,  the  carrying 
capacity  of  coyote  habitat  in  terms  of  prey  abun- 
dance cannot  be  evaluated  on  an  absolute  basis.  Of 
the  few  studies  conducted  to  estimate  food  require- 
ments of  coyotes  (Clark  1972;  Wagner  and  Stoddart 
1972),  disparate  results  have  been  obtained  (re- 
viewed by  Gier  1975).  Furthermore,  the  decline  in 
prime  habitat  may  also  cause  decreases  in  coyote 
density  (Andelt  and  Andelt  1981 ). 

Foxes.  Like  the  coyote,  the  gray  fox  (Urocyon 
cinereoargenteus)  and  red  fox  (Vulpes  vulpes)  have 
adapted  to  a  wide  range  of  habitat  types  and  foods. 
Most  of  the  gray  fox  habitat  consists  of  shrublands 
and  brushy  woodlands  on  hilly,  rough,  rocky,  or 
broken  terrain.  In  the  western  U.S.,  the  gray  fox 


478 


Carnivores 


favors  chaparral,  woodlands  of  pinyon-juniper 
(Pinusjuniperus  sp. )  or  oak,  along  rocky  hillsides, 
mountainsides,  and  washes.  Chaparral  vegetation 
of  the  foothills  of  the  Pacific  states  includes 
ceanothus  {Ceanothus  sp. ),  chamise  (Adenostoma 
fasciculatum),  manzanita  (Arctostaphylos  sp. ), 
mountain  mahogony  (Cercocarpus  sp. ),  and  oak. 
Here  the  gray  fox  supplements  its  primary  diet  of 
rodents  with  manzanita  berries  (Ingles  1965). 

In  Zion  National  Park,  gray  foxes  are  abundant 
in  blackbrush  (Coleogyne  ramosissima),  brushy 
meadows,  open  meadows,  and  pinyon-juniper  and 
ponderosa  pine  at  lower  elevations  (Trapp  and  Hall- 
berg  1975).  Here  the  gray  fox  is  more  a  herbivore, 
insectivore,  or  scavenger  than  a  carnivore.  Juniper 
berries  are  an  important  food  during  winter,  as  is 
mule  deer  carrion  during  late  winter  and  prickly 
pear  fruits  during  autumn.  Similar  results  were  ob- 
tained from  a  food  habits  study  in  an  Arizona  ponde- 
rosa pine  forest  (Turkowski  1980). 

The  red  fox's  unspecialized  way  of  life  has  al- 
lowed it  to  adapt  and  thrive  in  many  habitat  types. 
While  it  ranges  from  deep  forests  to  the  most  ex- 
posed tundra,  it  prefers  a  mixture  of  forest  and 
meadows.  In  the  continental  U.S.,  the  red  fox  is 
widespread  except  in  the  Great  Plains  and  the  ex- 
treme Southeast  and  Southwest.  One  of  the  densest 
populations  of  red  foxes  in  North  America  inhabits 
southwestern  Wisconsin  (Richards  and  Hine  1953). 
This  country  is  a  patchwork  of  woodlots,  cropland, 
pasture,  and  stream  bottoms.  By  following  fox  trails 
in  the  snow,  Schofield  ( I960)  found  that  red  foxes 
in  Michigan,  at  least  during  winter,  preferred  low- 
land brush  and  oak  woods  but  avoided  swamps.  In- 
gles (1965)  believed  that  sheep  grazing  in  meadows 
adversely  affected  red  fox  populations  in  the  western 
mountains  by  reducing  their  rodent  food  supply. 
The  overall  implication  is  that  although  the  red  fox 
can  survive  in  many  habitat  types,  it  prefers  those 
with  a  mixture  of  plant  communities.  Major  factors 
determining  habitat  selection  are  probably  cover 
requirements  and  food  availability. 

Red  foxes  are  opportunistic  feeders,  eating  foods 
in  proportion  to  their  availability  (Errington  1935; 
Scott  1955).  Thus,  red  fox  diets  often  show  seasonal 
changes  in  composition  (Korschgen  1959).  Most 
red  fox  food  habit  studies  reviewed  by  Abies  ( 1 975 ) 
showed  major  foods  to  be  small  rodents,  rabbits 
(Sylvilagus  sp. ),  wild  fruits  and  berries,  and  insects. 

The  kit  fox  (Vulpes  macrotis)  and  swift  fox 
{V.  velox)  inhabit  the  deserts  and  Great  Plains  of 
North  America,  respectively — areas  not  occupied  by 
the  red  fox.  An  exception  is  in  western  Kansas 
where  the  red  fox  is  usually  found  near  towns,  and 
the  swift  fox,  in  open,  often  grassland  areas.  The 


Kit  fox. 


kit  fox  is  the  most  specialized  of  the  North  American 
canids  for  desert  existence.  It  prefers  open,  level, 
sandy  ground  with  low  desert  vegetation. 

Egoscue  ( 1962)  studied  kit  foxes  in  a  western 
Utah  desert  that  contained  three  plant  communities: 
vegetated  or  stabilized  dunes,  greasewood  (Sarcoba- 
tiis  vermiculatus)  flats,  and  shadscale  (Atriplex  con- 
fertifolia)  flats.  Vegetated  dunes  have  a  diversified 
plant  cover  dominated  by  desert  shrubs,  including 
fourwinged  saltbush  {Atriplex  canescens),  rabbit- 
brush  (Chrysothamnus  sp. ),  greasewood,  horsebrush 
(Tetradymia  sp. ),  and  shrubby  buckwheat  (Erio- 
gonum  dubium ).  With  a  single  exception,  no  kit  fox 
dens  were  discovered  in  this  plant  community. 
Greasewood  flats  are  dominated  by  widely  spaced 
greasewood  shrubs,  ltol.5m(3to5ft)  high. 
Ground  cover  between  shrubs  is  sparse  and  consists 
primarily  of  grey  molly  (Kockia  americana),  seep- 
weed  (Suaeda  torryand),  and  shadscale.  Numerous 
kit  fox  dens  were  located  in  this  community,  which 
also  was  a  favorite  foraging  area.  Shadscale  flats  con- 
sist of  sparsely  growing  vegetation  averaging  =g60cm 
(=524  in.)  high.  Dominant  species  were  shadscale, 
inkweed,  and  grey  molly.  This  community  was  pri- 
marily important  as  a  hunting  ground  for  black-tailed 
jackrabbits,  which  composed  90  to  95%  of  kit  fox 
diets  during  the  pup-rearing  season.  Although  jack- 
rabbit  populations  were  not  especially  dense  in 
shadscale  flats,  their  numbers  did  not  seem  to  fluc- 
tuate as  sharply  as  in  other  communities.  Thus,  this 
community  appeared  to  provide  a  stable  food 
source. 


The  swift  fox  is  considered  by  some  to  be  a 
subspecies  of  the  kit  fox  (Stains  1975).  The  swift 
fox,  once  common  and  widespread  on  the  high 
plains  of  the  central  U.S.  and  Canada,  now  has  a 
much  more  restricted  distribution  (Kilgore  1969; 
Moore  and  Martin  1980).  The  original  swift  fox  habi- 
tat, the  shortgrass  plains  of  the  Oklahoma  Panhandle, 


Carnivores 


479 


was  dominated  by  buffalograss  {Buchloe  dacty- 
toides)  and  blue  grama  (Bouteloua  gracilis)  (Kil- 
gore  1969).  Other  grass  species  were  bluestem 
(Andropogon  scoparius),  wiregrass  (Aristida  sp.), 
and  sideoats  grama  (B.  curtipendula).  Much  of  this 
region  is  now  extensively  cultivated;  as  a  result,  the 
original  vegetation  has  largely  been  replaced  by 
grain  crops  and  weeds.  In  Wyoming,  recent  records 
of  swift  foxes  have  come  from  areas  of  gently  rolling, 
shortgrass  prairie,  dominated  by  buffalograss  and 
blue  grama  (Floyd  and  Stromberg  1981 ).  In  this  area, 
the  native  shortgrass  prairie  is  interspersed  with 
winter  wheat,  alfalfa,  and  fallow  fields.  This  habitat  is 
very  similar  to  that  described  for  swift  foxes  in  the 
Oklahoma  Panhandle  by  Kilgore  ( 1969). 

Felidae 

Mountain  Lions.  In  western  North  America, 
mountain  lions  (Felis  concolor)  are  generally 
associated  with  mountainous  terrain,  canyons,  and 
rimrock.  Here  they  feed  primarily  on  mule  deer 
(Robinette  et  al.  1959;  Barnes  I960;  Spalding  and 
Lesowski  1971;  Toweill  and  Meslow  1977;  reviewed 
by  Anderson  1983  and  Ackerman  et  al.  1984). 

In  southern  Utah,  mountain  lion  habitat  is  vege- 
tatively  and  topographically  diverse  (Barnes  I960; 
Ackerman  1982;  Hemker  1982).  Desert  shrub  and 
sagebrush-grassland  communities  occur  at  lower 
elevations  (1,350  to  1,800  m  [4,445  to  5,940  ft]). 
Pinyon-juniper  woodlands,  oakbrush  (Quercus  gam- 
belii),  and  ponderosa  pine  forests  dominate  mid- 
elevations  (1,800  to  2,700  m  [5,940  to  8,910  ft]). 
Stands  of  quaking  aspen  (Populns  tremidoides), 
Engelmann  spruce  (Picea  engelmann  ii),  and  white 
fir  (Abies  concolor)  interspersed  with  subalpine 
meadows  occur  above  2,700  m  (8,910  ft).  Deep, 
rocky,  vertical-walled  river  canyons  within  these 
communities  contain  riparian  vegetation,  including 
Fremont  Cottonwood  (P.  fremontii)  and  willow 
(Salix  sp.). 

Vegetation  and  topography  of  the  Idaho  Primi- 
tive Area,  where  Hornocker  ( 1969,  1970)  and  Sei- 
densticker  et  al.  (1973)  conducted  extensive 
mountain  lion  research,  are  also  diverse.  Plant  com- 
munities are  interspersed,  depending  on  site  charac- 
teristics including  elevation,  slope,  and  aspect. 
Engelmann  spruce-subalpine  fir  (A  lasiocarpa)  and 
ponderosa  pine-Douglas-fir  (Pseudotsuga  menziesii) 
associations  grow  at  higher  elevations  and  in  pro- 
tected drainages.  At  lower  elevations  and  on  exposed 
slopes,  curlleaf  mountain  mahogany  (Cerococarpus 
ledifolius),  bitterbrush  (Purshia  tridentata),  and  big 
sagebrush  (A  tridenta  ta)-bunchgrass  associations 
occur. 

In  California,  mountain  lions  occur  primarily 
between  600  and  1,800  m  ( 1,980  and  5,940  ft)  ele- 
vations in  mixed  conifer  and  brush  habitats  (Koford 


1977).  At  higher  elevations,  in  pure  stands  of  coni- 
fers, and  at  lower  elevations,  in  pure  stands  of  cha- 
mise  brush  (Adenostoma  fasciculata),  mountain 
lions  are  rare. 

Within  these  western  habitats,  stream  courses 
and  ridgetops  are  frequently  used  as  travel  lanes  and 
hunting  routes  by  mountain  lions  (Barnes  I960). 
Riparian  vegetation  along  streams  provides  cover  for 
lions  traveling  in  open  areas,  and  ridgetops  with 
cover  allow  undetected  surveillance. 

From  these  studies,  mountain  lions  appear  to 
prefer  habitats  that  are  vegetatively  and  topographi- 
cally complex.  This  habitat  characteristic  and  prey 
availability  determine  the  amount  of  space  a  resident 
mountain  lion  requires.  Because  of  large  home 
ranges,  territoriality,  and  population  social  structure, 
large  areas  of  wilderness  or  remote  habitat  are  essen- 
tial to  sustain  viable  mountain  lion  populations  (Hor- 
nocker 1969,  1970;  Koford  1977;  Hemker  1982; 
reviewed  by  Anderson  1983).  Mountain  lions  wan- 
der close  to  civilization  (Barnes  I960),  and  some 
populations  live  near  humans  (Koford  1977).  This  is 
especially  true  in  California,  where  there  is  a  morato- 
rium on  mountain  lion  hunting.  It  is  questionable, 
however,  whether  mountain  lion  populations  can  be 
sustained  in  this  increasingly  modified  habitat 
(Figure  1A). 

Bobcats.  Bobcat  (Felis  rafus )  and  mountain  lion 
habitat  preferences  are  similar  in  that  both  use  a 
variety  of  habitats.  Bobcats  still  inhabit  much  of  their 
former  range  in  the  eastern  and  southern  U.S. 
(Figure  IB),  where  the  mountain  lion  has  been 
extirpated.  Western  bobcats  prefer  rocky  canyons  at 
elevations  between  1,400  and  2,100  m  (4,620  and 
6,930  ft )  with  ledges  and  areas  of  dense  vegetation 
(Young  1958;  Gashwiler  et  al.  1961;  McCord  and 
Cardoza  1982).  Common  tree  and  shrub  species  of 
western  habitats  are  manzanita,  mountain  mahogany, 
pinyon  pine,  sagebrush,  and  juniper.  In  the 
southwestern  and  western  U.S.,  bobcats  have  adapted 
to  living  in  even  the  driest  deserts  (Young  1958; 
Jones  and  Smith  1979). 

Prey  abundance,  topography,  and  vegetation 
structure  affect  bobcat  social  structure  and  home 
range  size  (Bailey  1974;  Beason  and  Moore  1977).  In 
turn,  social  organization  profoundly  affects  popula- 
tion density  and  the  ways  bobcats  use  their  environ- 
ment. Competitive  interaction  with  coyotes  also 
likely  affects  bobcat  population  density  and  behavior 
(Linhart  and  Robinson  1972). 

Lynx.  Lynx  (F.  lynx)  are  more  restrictive  in  habitat 
and  food  selection  than  bobcats,  making  them  more 
vulnerable  to  a  changing  environment.  Where  both 
these  species  are  sympatric,  competition,  should 
it  occur,  may  be  more  detrimental  to  the  lynx 
(Parker  et  al.  1983).  Dominant  tree  species  of  boreal 


480 


Carnivores 


PAST 


PRESENT 


Figure  1A.     Historical  distribution  of  the  mountain  lion. 


PAST 


PRESENT 


Figure  IB.     Historical  distribution  of  the  bobcat. 


Carnivores 


481 


forests  inhabited  by  lynx  are  balsam  fir 
(A  balsamea),  black  spruce  {P.  mariana),  white 
spruce  (P.  glauca),  and  paper  birch  (Betnal 
papyrifera)  (Brand  et  al.  1976;  Parker  et  al.  1983). 

In  Alberta,  lynx  inhabit  a  variety  of  vegetation 
types,  ranging  from  undisturbed  forest  to  90%  culti- 
vated farmland  (Nellis  and  Keith  1968;  Nellis  et  al. 
1972).  In  Montana,  lynx  preferred  dense  stands  of 
lodgepole  pine  (P.  contorta)  (Koehler  et  al.  1979). 
Snowshoe  hares  were  also  most  abundant  in  this 
habitat  type.  Throughout  their  range,  lynx  depend 
on  snowshoe  hares  for  most  of  their  diet  (Saunders 
1963;  van  Zyll  de  Jong  1966;  Nellis  and  Keith  1968; 
Nellis  et  al.  1972;  Brand  et  al.  1976;  More  1976; 
Koehler  et  al.  1979;  Parker  et  al.  1983).  This  de- 
pendence is  reflected  in  cyclical  fluctuations  of  lynx 
populations  with  changing  snowshoe  hare  densities 
(Figure  2)  (Keith  1963;  Brand  et  al.  1976;  Brand  and 
Keith  1979).  Because  of  this  phenomenon,  forest 
management  plans  that  incorporate  habitat  require- 
ments of  snowshoe  hares  are,  in  effect,  management 
plans  for  lynx  (Parker  et  al.  1983).  Clear-cutting 
and  burning  of  mature  conifer  forests  are  likely  to 
initially  decrease  hare  and  lynx  populations,  but 
in  the  long  term,  an  increase  in  food  and  cover  for 
hares  should  increase  densities  of  both  hares  and 
lynx  (Koehler  et  al.  1979;  Parker  et  al.  1983). 

Ursidae 

Black  Bear.  Bray  and  Barnes  ( 1967)  noted  that 
forested  environments  and  food  availability  were 
emphasized  in  almost  all  habitat  descriptions  of  black 
bear  (Ursus  americanus).  Forests  provide  escape 
cover,  thermal  cover,  and  den  sites.  However,  in 
western  North  America,  black  bear  habitat  ranges 
from  desert  chaparral  to  closed  coniferous  forest 


(Jonkel  and  Cowan  1971;  Amstrup  and  Beecham 
1976;  Lindzey  and  Meslow  1977;  Kelleyhouse  1980; 
LeCount  1980;  Graber  1982;  Novick  and  Stewart 
1982).  The  physical  features  and  vegetation  of  black 
bear  habitat  are  exceedingly  variable,  reflecting 
differences  in  latitude,  elevation,  slope,  aspect, 
precipitation,  and  land-use  patterns.  Because  black 
bears  have  adapted  to  so  many  habitat  types,  there  is 
no  clear  consensus  on  which  habitat  components 
are  most  important  (Treadwell  1979).  Determining 
local  habitat  use  may  be  needed  for  sound  manage- 
ment of  this  species. 


In  whatever  ecosystem  black  bears  occur,  food 
availability  greatly  influences  their  abundance  and 
distribution.  Jonkel  and  Cowan  (  1971 )  found  that 
seasonal  food  production  dictated  habitat  use  by 
black  bears  in  spruce-fir  {Picea-Abies)  forests  in 
northwestern  Montana.  The  spruce-fir/mountain 
lover  (Pachistima  myrsinites)  association  was  an 
important  component  of  black  bear  habitat  during  all 
seasons.  Other  habitat  components  were  important 
seasonally:  dry  meadows  in  early  spring,  snowslides 
and  stream  bottoms  in  early  and  midsummer,  and 
spruce-fir/rustyleaf  (Menziesia  fetrnginea)  and 
spruce-fir/bcargrass  (Xerophylum  tenax)  associations 
in  fall.  Recent  clearcuts  were  avoided.  The  home 
range  of  bears  included  the  seasonal  food  sources 
and,  when  a  seasonal  food  was  abundant,  bears  with 
adjacent  home  ranges  congregated  in  these  habitats. 

Lindzey  and  Meslow  ( 1977)  studied  home  range 
and  habitat  use  by  black  bears  in  a  western  hemlock- 
Sitka  spruce  (7:  heterophylla-P.  stichensis)  forest  in 
southwestern  Washington.  They  found  that  black 
bears  used  vegetation  types  disproportionately  to 
their  availability,  depending  on  availability  of  food 


250- 


200 


Figure  2.     Cyclical  fluctuations 
in  lynx  and  snowshoe  hare 
populations,  showing 
10-year  cycle  phenomenon, 
(after  Keith  1963). 


o 
o 
o 


*.  150-- 


uT   100  - 


50-- 


Hare 


Lynx 


1 1 1 1  n~i frTTTn  1 1 1 1  h  1 1 1 1 1 1 1  m  1 1 iii 


1850 


1860 


1870 


1880 


1890 


1900 


482 


Carnivores 


and  cover.  The  plant  communities  that  followed 
clear-cutting  provided  concentrations  of  foods,  espe- 
cially huckleberries  (Vaccinium  sp.). 

In  a  mixed  coniferous  forest  of  Douglas-fir,  pon- 
derosa  pine,  subalpine  fir,  and  Engelmann  spruce  in 
Idaho,  seven  forage  species  (wild  onions  [Allium 
sp.],  waterleaf  [Hydrophyllum  capitatwn],  biscuit 
root  [Lomatium  dissectum],  bitter  cherry  [Prunus 
emarginata],  chokecherry  [P.  virginiana],  mountain 
ash  [Sorbus  scopulina],  and  huckleberry  [Vaccinium 
globulare])  were  highly  correlated  with  locations 
of  radio-collared  black  bears  (Amstrup  and  Beecham 
1976).  Locations  of  black  bears  corresponded  to 
areas  of  maximum  food  availability,  which  depended 
on  phenological  stages  of  these  seven  key  food 
plants.  Researchers  in  California  also  correlated  habi- 
tat use  by  black  bears  with  seasonal  food  availability 
(Piekielek  and  Burton  1975;  Kelleyhouse  1980;  Nov- 
ick  et  al.  1981;  Novick  and  Stewart  1982;  Grenfell 
and  Brody  1983).  Habitat  types  ranged  from  chapar- 
ral, oak  woodlands,  and  coniferous  forest. 


Grizzly  Bears.  In  contrast  to  the  importance  of 
forested  habitats  to  black  bears,  the  importance  of 
nonforested  habitats  to  grizzly  bears  has  been 
emphasized  by  several  researchers.  Herrero  (1972, 
1978)  hypothesized  that  grizzly/brown  bears  evolved 
to  exploit  nonforested,  periglacial  environments. 
Martinka  (1976)  stressed  the  importance  of  fire  in 
reducing  or  eliminating  forest  canopies  to  improve 
habitat  for  grizzlies.  He  stated  that  without  natural 
fire,  successional  advance  towards  mature  forest 
creates  conditions  more  favorable  for  black  bears 
than  grizzlies.  Particularly  important  open  habitats 
are  wet  meadows  and  alluvial  plains  that  supply 
grizzlies  with  key  forbs  and  rhizomatous  grasses 
(Singer  1978). 

Craighead  et  al.  ( 1982)  included  extensive  tim- 
ber cover,  as  well  as  open  grasslands  and  meadows, 
in  their  descriptions  of  optimum  grizzly  habitat  in 
the  continental  U.S.  Grizzly/brown  bear  populations 
in  Alaska  and  Canada  apparently  have  no  intrinsic 
need  for  extensive  timber  cover,  as  they  thrive  in 
open  areas.  However,  grizzly  populations  in  the  con- 
tinental U.S.  may  require  forested  habitat  for  isola- 
tion from  human  activities.  Here  a  combination  of  a 
wide  range  of  vegetation  types  characterizes  prime 
grizzly  bear  habitat;  mountain  parks,  grasslands  inter- 
spersed with  timber,  alpine  meadows,  and  talus 
slopes  are  necessary  for  feeding  and  social  activities. 
Alder  {Alnus  tenuifolia)  thickets,  lodgepole  pine 
"downfalls,"  and  other  dense  vegetation  are  pre- 
ferred bedding  sites. 

Large  geographical  areas  are  needed  to  provide 
a  variety  of  alternative  foods.  Major  foods  are  car- 
rion, ungulates,  rodents,  insects,  berries,  pine  nuts, 
green  vegetation,  bulbs  and  tubers  and,  in  some 


areas,  fish  (Mealy  1980).  Habitat  use  by  grizzlies  is 
greatly  influenced  by  the  seasonal  availability  of 
these  foods.  In  spring,  the  location  of  carcasses  of 
ungulates  that  died  during  the  winter  or  individuals 
that  emerged  from  the  winter  in  poor  condition 
are  important  (Cole  1972;  Houston  1978).  In  spring 
and  summer,  areas  with  high  population  densities 
of  rodents  or  insects,  fish-spawning  sites,  or  succu- 
lent herbaceous  vegetation  or  bulbs  and  tubers  are 
important.  In  fall,  berries,  pine  nuts,  and  ungulates 
crippled  during  hunting  seasons  are  important  food 
sources.  Because  of  the  conflict  between  grizzlies 
and  humans,  along  with  human  land  uses,  wilderness 
areas  are  essential  to  separate  them.  Grizzly  habitat 
must  be  isolated  from  developed  areas  and  should 
receive  only  light  recreational,  logging,  and  livestock 
use.  Except  for  humans,  the  grizzly  has  no  enemies 
that  restrict  its  use  of  habitat. 


Arboreal,  Aquatic,  and  Terrestrial 
Furbearers 

Fisher,  Marten,  and  Wolverine. 


Fisher  (Martes  pennanti).  These  mustelids  are 
generally  associated  with  climax  coniferous  forests 
(de  Vos  1952;  Hornocker  and  Hash  1981;  Powell 
1982;  Allen  1982,  1983;  Douglass  et  al.  1983; 
Spencer  et  al.  1983;  Hargis  and  McCullough  1984) 
and  have  large  home  range  sizes  (Koehler  and 
Hornocker  1977;  Buck  et  al.  1978;  Powell  1982; 
Hornocker  and  Hash  1981).  Fishers  prefer 
continuous  coniferous  and  mixed  coniferous/ 
deciduous  forests  with  closed  canopies  ( reviewed  by 
Powell  1982;  Allen  1983).  Dense  overhead  cover  is 
the  one  characteristic  common  to  all  habitats 
preferred  by  fishers.  Canopy  closure  of  80  to  100% 
is  selected  and  canopy  closure  of  less  than  50%  is 
avoided.  Within  these  habitats,  fishers  eat  a  variety  of 
foods,  including  small  mammals,  carrion,  birds  and 
their  eggs,  insects,  reptiles,  amphibians,  and  various 
fruits  and  nuts  (reviewed  by  Strickland  et  al.  1982). 

Grinnell  et  al.  (1937)  and  Schempf  and  White 
( 1977)  reported  that  fishers  in  California  are  most 
commonly  found  in  Douglas -fir  and  mixed  conifer 
forests  between  610  and  2,440  m  (2,000  and  8,000 
ft)  elevations.  Here  an  unusual  but  important  diet 
item  is  false  truffles  (Rhizopogon  sp.;  Grenfell  and 
Fasenfest  1979). 

Marten  {Martes  americana ).  The  effects  of  fire 
and  timber  harvest  have  generally  been  considered 
harmful  to  fisher  and  marten  habitat  (de  Vos  1952; 
reviewed  by  Powell  1982).  However,  whether  this  is 
true  for  marten  habitat  is  subject  to  disagreement, 
because  marten  appear  to  be  less  dependent  than 
fisher  on  old  growth  forest.  Marten  habitat  use  varies 
with  the  season. 


Carnivores 


483 


During  winter,  marten  prefer  dense  overhead 
cover  of  mature  coniferous  forests  (Koehler  and 
Hornocker  1977).  In  Idaho,  marten  activity  was 
highest  in  stands  having  an  Engelmann  spruce-subal- 
pine  fir  overstory,  100  years  old  or  more,  and  a  can- 
opy cover  greater  than  30% ;  mesic  sites  also  had 
high  marten  activity.  Similarly,  marten  in  northern 
California  preferred  stands  of  lodgepole  pine  or  red 
fir  (A  magnified)  with  closed  canopies  (Spencer 
et  al.  1983).  Within  both  these  habitat  types,  marten 
selected  sites  with  Douglas  squirrel  (Tamiasciurus 
douglasii)  feeding  sign.  Old-growth  red  fir  with 
large  snags,  stumps,  and  logs  was  an  especially  pre- 
ferred habitat.  In  Yosemite  National  Park,  marten 
selected  dense  cover  less  than  3  m  (9  ft)  above 
snow  level  in  lodgepole  pine  and  western  juniper 
forests  (Hargis  and  McCullough  1984).  Tree  trunks 
allowed  access  to  the  subnivean  (below  the  snow 
surface)  zone,  and  logs  served  as  den  sites  and  hunt- 
ing areas.  Open  areas  were  avoided.  Preference  for 
closed  canopies  may  also  be  due  to  less  snow  accu- 
mulation, which  would  permit  marten  to  conserve 
energy  of  movement  (Raine  1982).  Another  means 
to  conserve  energy  is  use  of  red  squirrel  (Tamias- 
ciurns hudsoniens)  middens  as  roosting  sites 
(Buskirk  1984). 

Soutiere  (1979)  hypothesized  that  prey  availa- 
bility in  dense  forests  rather  than  density  of  over- 
head cover,  per  se,  explained  winter  habitat  use  by 
marten.  Small  mammals,  especially  the  red-backed 
vole  (Clethrionomys  gapperi),  make  up  the  major 
portion  of  marten  diets  during  winter  (Cowan  and 
MacKay  1959;  Weckwerth  and  Hawley  1962;  Sou- 
tiere 1979).  Red-backed  voles  are  associated  with 
downed  logs  and  mesic  conditions  of  uncut  conifer- 
ous forests  (Koehler  and  Hornocker  1977). 

In  summer,  effects  of  timber  harvest  and  fire 
may  be  beneficial  to  marten.  In  Idaho,  Koehler  and 
Hornocker  (1977)  reported  that  plant  communities 
created  by  fire  provide  marten  with  important  sum- 
mer foods.  Streeter  and  Braun  ( 1968),  working  in 
Colorado,  also  noted  that  marten  do  not  require  the 
cover  of  climax  forests  during  summer.  Contrary 
to  these  reports,  Spencer  et  al.  (1983)  found  that 
marten  in  the  northern  Sierra  Nevada  avoided  open 
areas  during  all  seasons.  Soutiere  ( 1979)  concluded 
that  marten  in  the  Northeast  were  not  limited  to 
climax  forests,  and  partial-cutting  methods  of  timber 
harvest  were  compatible  with  the  preservation  of 
marten  habitat.  In  California,  Hargis  and  McCullough 
( 1984)  found  that  mixed-aged  forests  were  impor- 
tant in  providing  protective  cover  and  subnivean 
access  over  a  wide  range  of  snow  conditions. 

Wolverine  (Gulo  gulo).  Like  marten  and  fisher, 
the  wolverine  is  associated  with  boreal  forests.  In 
Montana,  wolverines  selected  subalpine  fir  cover 
types  throughout  the  year;  this  selection  was  strong- 
est during  summer  (Hornocker  and  Hash  1981 ). 


Serai  lodgepole  pine  and  western  larch  (Larix  occi- 
dentalism sites  were  also  used  by  wolverines.  Large 
areas  of  mature  timber  were  preferred,  followed 
by  ecotonal  areas  and  rocky  areas  of  timbered 
benches.  Areas  of  dense,  young  timber;  recent  burns; 
and  wet  meadows  were  avoided. 

In  the  Pacific  states,  wolverines  are  associated 
with  coniferous  forests  containing  Shasta  red  fir  (A 
magnifica),  lodgepole  pine,  and  western  hemlock 
(T.  heterophylla);  at  higher  elevations,  subalpine  fir, 
Alaska  cedar  (P.  albicanlis),  and  western  juniper 
(J.  occidentalis)  (Ingles  1965).  In  Canada,  wolver- 
ines inhabit  tundra  between  the  northern  tree  line 
and  arctic  coasts  as  well  as  boreal  forests  (Banfield 
1974). 

Wilderness  is  important  in  protecting  wolverine 
populations  because  they  are  highly  susceptible  to 
human-caused  mortality  (van  Zyll  de  Jong  1975). 
Hornocker  and  Hash  ( 1981 )  found  no  difference  in 
wolverine  density  between  wilderness  and  nonwild- 
erness  areas,  but  concluded  that  wilderness  was  es- 
sential as  a  refuge  and  reservoir  for  wolverine 
populations.  Furthermore,  large  areas  of  habitat  are 
needed  to  support  wolverine  populations  because  of 
the  wolverine's  scavenging  life-style,  which  requires 
a  solitary  existence  and  a  large  home  range.  Carrion 
is  a  mainstay  in  wolverine  diets,  especially  during 
winter  (Rausch  and  Pearson  1972;  van  Zyll  de  Jong 
1975;  Hornocker  and  Hash  1981 ).  Van  Zyll  de  Jong 
( 1975)  hypothesized  that  wolverine  abundance  is 
correlated  with  biomass  and  turnover  of  large  ungu- 
late populations. 

Mink  and  River  Otter. 


Mink  (Mustela  vison ).  These  amphibious 
mustelids  are  widespread,  but  uncommon,  and  have 
irregular  distributions  in  western  North  America. 
With  the  exception  of  otters,  mink  lead  a  more 
aquatic  existence  than  any  other  mustelid.  They  are 
often  found  near  water,  where  a  diet  of  muskrats 
(Ondatra  zibethicus),  fish,  crayfish,  frogs,  and 
aquatic  invertebrates  are  available  (Haley  1975; 
Gilbert  and  Nancekivell  1982).  In  Canada  and  Alaska, 
habitat  types  with  some  of  the  densest  mink 
populations  are  tidal  flats  and  ocean  beaches  (Burns 
1964;  Banfield  1974).  Unlike  otters,  mink  are  also 
common  on  land  where  they  primarily  hunt  rodents 
and  lagomorphs.  In  inland  Canada,  snowshoe  hares 
are  often  their  principal  prey. 

River  Otter  (Lutra  canadensis).  The  more 
aquatic  river  otter  depends  on  fish  for  a  larger  per- 
centage of  its  diet  than  does  the  mink,  especially 
in  stream  habitats  (Melquist  et  al.  1981 ).  Mink  once 
concentrated  on  mammals  and  birds  in  stream  habi- 
tats but  switched  to  eating  mostly  fish  at  lakes  (Gil- 
bert and  Nancekivell  1982).  River  otters,  however, 


484 


Carnivores 


ate  primarily  fish  in  both  habitats.  Besides  fish,  im- 
portant river  otter  foods  are  crayfish,  frogs,  turtles, 
and  aquatic  insects.  Muskrat,  beaver  {Castor  cana- 
densis), and  water  birds  occur  infrequently  in  river 
otter  diets,  but  in  some  areas  they  may  be  seasonally 
important  (Greer  1955). 

Throughout  their  range,  river  otters  inhabit 
streams,  rivers,  lakes,  marshes,  and  ocean  bays  where 
fish  populations  are  dense.  In  Canada,  they  occur 
north  beyond  the  tree  line  in  tundra  lakes  and 
streams  (Banfield  1974).  In  west-central  Idaho,  prey 
availability  had  the  greatest  influence  on  habitat  use 
by  river  otters,  but  adequate  shelter  was  necessary 
for  extensive  habitat  use  (Melquist  and  Hornocker 
1983).  Rather  than  excavate  their  own  dens,  river 
otters  use  dens  constructed  by  other  animals,  often 
those  of  beavers  (Haley  1975;  Melquist  and  Hor- 
nocker 1983).  They  also  use  natural  or  man-made 
structures  as  dens.  Habitat  use  in  Idaho,  determined 
by  biotelemetry,  indicated  that  river  otters  preferred 
valley  to  mountain  habitats  and  stream-associated 
habitats  to  ponds,  lakes,  and  reservoirs  (Melquist  and 
Hornocker  1983).  This  was  believed  to  be  true  be- 
cause stream  habitats  provided  more  adequate  es- 
cape cover  and  shelter  and  less  human  disturbance. 
However,  if  food  and  shelter  were  adequate,  river 
otters  exhibited  high  tolerance  toward  human 
activity. 


During  winter,  habitat  use  by  mink  and  river 
otters  is  complicated  by  ice-covered  waterways 
(Banfield  1974;  Melquist  and  Hornocker  1983).  Hab- 
itat is  limited  to  open  water  during  early  winter. 
Outflows  from  lakes  are  particularly  favored  haunts 
at  this  time.  In  late  winter,  water  levels  usually  drop 
below  ice  levels  in  rivers  and  lakes,  leaving  a  layer  of 
air  that  allows  mink  and  river  otters  to  travel  and 
hunt  under  the  ice. 

Water  quality  and  quantity  are  habitat  variables 
that  directly  and  indirectly  affect  mink  and  river 
otters.  As  secondary  or  higher-order  consumers, 
these  carnivores  concentrate  water  pollutants  (Cum- 
bie  1975;  Henry  et  al.  1981).  Polluted  or  inadequate 
water  quantity  also  adversely  affects  prey  popula- 
tions, thereby  diminishing  the  carrying  capacity  of 
mink  and  river  otter  habitat. 

Badgers,  Raccoons,  Ringtails,  Skunks,  and 
Weasels.  The  geographic  distribution  of  most  of 
these  small-  to  medium-sized  carnivores  overlap 
broad  areas  in  western  North  America.  Many  of  these 
species  also  have  similar  food  habits.  Being 
opportunists,  they  eat  foods  locally  or  seasonally 
available  and,  except  for  the  weasels  (Mustela  sp.), 
all  are  omnivorous.  Because  of  these  similarities,  we 
have  grouped  these  species  so  important  habitat 
features  can  be  compared  and  constructed. 


River  otters. 


Carnivores 


485 


Badger  (Taxidea  taxus).  Badgers  are  wide- 
spread in  the  western  U.S.,  occurring  in  all  plant 
communities  except  the  rain  forests  of  Washington 
and  Oregon.  They  occur  from  sea  level  to  alpine 
meadows  and  from  deserts  to  dense  coniferous  for- 
ests (Long  1973;  Long  and  Killingley  1983).  Conse- 
quently, their  general  food  habits  are  not  too 
surprising.  Badgers  eat  almost  any  species  of  verte- 
brate or  invertebrate  they  can  capture,  but  especially 
rodents  captured  underground  (Messick  and  Hor- 
nocker  1981;  Lampe  1982).  They  are  most  numer- 
ous in  sagebrush-grasslands  where  pocket  gophers 
(Thomomys  sp.)  and  ground  squirrels  (Spermophi- 
his  sp.)  abound.  Their  home  range  size  varies  dra- 
matically with  the  season  (Sargent  and  Warner  1972; 
Lindzey  1978;  Lampe  and  Sovada  1981 ),  being  small- 
est during  winter  when  they  can  enter  a  state  of 
torpor  and  reduce  their  time  aboveground  by  93% 
(Harlow  1981). 


Badgers. 


Raccoon  (Procyon  lotor).  Raccoons  are  com- 
mon in  forests  of  the  western  U.S.  at  low  to  middle 
elevations,  where  they  greatly  depend  on  riparian 
areas  (Lotze  and  Anderson  1979;  Hart  1982).  They 
prefer  deciduous  stands  over  coniferous  forests,  but 
even  inhabit  basaltic  outcroppings  in  sagebrush  eco- 
systems. They  are  absent  from  deserts,  high  eleva- 
tions, and  the  boreal  forests  of  Canada  and  Alaska. 
Their  habitat  must  include  den  sites;  hollow 
branches  are  preferred,  but  hollow  trunks  and  logs, 
caves  or  crevices,  and  attics  of  abandoned  farm 
buildings  are  also  used  (Berner  and  Gysel  1967; 
Lynch  1974).  Raccoons  are  probably  the  most  omni- 
vorous native  carnivores  in  the  western  U.S.,  feeding 
on  whatever  foods  are  available.  They  forage  exten- 
sively in  farmyards  and  wetlands  (Ellis  1964;  Urban 
1970;  Fritzell  1978;  Greenwood  1982)  and  have 
adapted  to  living  in  suburban  residential  communi- 
ties (Hoffman  and  Gottschang  1977). 


Ringtail  (Bassariscus  astutas).  Depending  on 
habitat,  ringtails  may  also  be  omnivorous.  They  typi- 
cally inhabit  rocky  and  brushy  canyons  and  talus 
area.  They  are  also  common  in  shinnery  oak  (Q.  har- 
varclii)  woodlands  and  occur  in  live  oak  (Q.  virgi- 
niana)  savannas. 


In  Colorado,  ringtails  are  usually  associated  with 
arid  canyon  and  mesa  country  in  the  pinyon-juniper 
vegetation  type.  However,  they  have  been  reported 
in  a  wide  variety  of  habitats,  including  an  Engelmann 
spruce-lodgepole,  pine-aspen  community  (Rutherford 
1954).  Similarly,  in  southern  Nevada,  Bradley  and 
Hansen  (1965)  reported  ringtails  in  many  habitat 
types.  They  concluded  that  ringtails  were  present  in 
all  habitats  except  the  most  arid  lower  deserts. 


Skunk  (Conepatus  sp.,  Mephitis  sp.,  Spilogale 
sp.).  The  striped  skunk  (M  mephitis)  occurs  in  a 
wider  range  of  habitats  and  is  more  omnivorous  than 
the  ringtail  (Wood  1954;  Verts  1967;  Ewer  1973; 
Haley  1975;  Schowalter  and  Gunson  1982).  This,  the 
most  common  mustelid,  is  absent  from  colder 
climes,  likely  due  to  its  inability  to  remain  dormant 
for  such  long  winters  (Haley  1975).  Within  its  range, 
the  striped  skunk  flourishes  from  city  parks  to  wil- 
derness areas  because  of  its  generalized  habits.  Pre- 
ferred habitats  are  difficult  to  delineate  for  a  species 
so  adaptable  and  widely  distributed. 

Verts  (1967)  concluded  that  relative  abundance 
of  striped  skunks  determined  their  preferred  habitat. 
Using  this  criterion,  he  found  that  the  extent  and 
location  of  woodlots  in  farmlands  of  northeastern  Il- 
linois were  important  habitat  factors.  Presumably, 
woodlots  provided  needed  cover  and  den  sites.  Con- 
trary to  earlier  reports,  striped  skunks  were  more 
abundant  in  intensively  cultivated  areas  than  where 
forest,  brushland,  and  farmland  were  intermixed 
(Verts  1967;  Smith  and  Verts  1982).  Storm  (1972) 
found  that  striped  skunks  used  pasture  and  hay  crops 
more  frequently  than  expected  because  they  were 
conducive  to  night  foraging  and  retreats,  day  or 
night.  Other  crops  and  uncultivated  areas  were  used 
less  than  expected.  In  the  open,  dry  country  of 
Idaho,  the  importance  of  riparian  growth  along 
streams  was  deemed  important  for  the  same  reasons 
(Larrison  and  Johnson  1981).  Along  the  Pacific 
Coast,  sand  dunes  covered  with  sufficient  vegetation 
to  permit  digging  of  dens  are  excellent  habitat  (In- 
gles 1965).  Here  striped  skunks  scavenge  along 
beaches  at  night  for  washed-up,  dead  marine  animals. 
Similarly,  in  the  Mississippi  Lowland  Region  of  Mis- 
souri, available  denning  habitat  determines  the  distri- 
bution and  population  density  of  striped  skunks 
(Verts  1967).  The  high  water  table  makes  ground 
dens  unsuitable  in  many  areas  of  this  region. 


486 


Carnivores 


Life-styles  of  the  hooded  skunk  (M.  macroura) 
and  hognosed  skunk  are  similar  to  those  of  the 
striped  skunk.  The  hooded  skunk  prefers  rocky  and 
brushy  canyons  and  riparian  areas.  Two  or  more 
species  of  hognosed  skunks  have  been  recognized 
(Hall  and  Kelson  1959)  in  these  areas,  but  may  only 
be  geographical  races  of  the  same  species  (Haley 
1975).  This  skunk  inhabits  desert  valleys,  grasslands, 
and  partly  wooded  mountain  foothills. 

Spotted  skunks  (some  authors  recognize  a  west- 
ern species,  S.  gracilis,  and  an  eastern  species,  S. 
putorius)  are  less  common  than  the  striped  skunk 
and  less  widely  distributed.  They  generally  prefer 
more  open  country  than  striped  skunks,  but  still 
require  some  form  of  cover  such  as  a  fence  row  or 
gully  vegetation  between  their  den  and  foraging 
areas  (Crabb  1948).  In  agricultural  areas,  they  often 
den  under  old  farm  buildings  or  in  abandoned 
ground  squirrel  burrows.  In  the  southwestern  U.S., 
spotted  skunks  inhabit  rough,  broken  country  below 
2,400  m  (8,000  ft).  They  can  survive  in  deserts  and 
drier  regions  of  Nevada,  New  Mexico,  and  Arizona, 
where  the  striped  skunk  cannot  (Haley  1975).  How- 
ever, they  depend  on  riparian  areas  in  arid  country 
(Larrison  and  Johnson  1981).  Although  spotted 
skunks  are  highly  omnivorous,  as  are  striped  skunks, 
rodents  are  relatively  more  important  and  arthro- 
pods correspondingly  less  so  in  their  diets. 


Weasel  (M  us  tela  nivalis,  M.  frenata,  M.  ermi- 
nea).  Unlike  the  rest  of  the  species  in  this  group, 
weasels  are  primarily  carnivorous,  specializing  in 
killing  rodents,  lagomorphs,  and  birds.  Like  other 
species  in  this  group,  weasels  occupy  a  broad  range 
of  habitats  including  boreal  and  deciduous  forests, 
shrub-steppe  regions,  grasslands,  and  agricultural 
areas  (Hall  1951).  Their  distributions  are  not  closely 
linked  to  those  of  any  certain  prey  species,  but  may 
be  limited  by  competition  with  other  weasel  species 
(Simms  1979). 


Short-tailed  weasel. 


areas  around  farmsteads  and  the  edges  of  urban 
areas.  However,  Ingles  ( 1965)  reported  that  short- 
tailed  weasels  in  California,  Oregon,  and  Washington 
inhabit  high  mountain  red  fir,  lodgepole  pine,  and 
subalpine  forests.  Rodents,  especially  Microtus  sp., 
are  important  in  the  diet  of  ermine,  but  shrews  and 
birds  are  also  eaten  (Ewer  1973;  Simms  1979). 

The  long-tailed  weasel  is  more  of  a  generalist  in 
food  habits  and  habitat  use  than  the  ermine  or  least 
weasel  (Ewer  1973;  Haley  1975;  Gamble  1981 ).  It 
also  has  a  more  southerly  distribution  than  these 
smaller  species.  Simms  (1979)  hypothesized  that  the 
northward  distribution  of  the  long-tailed  weasel  is 
limited  by  snow  cover,  while  the  southward  distribu- 
tions of  the  two  smaller  species  are  limited  by  inter- 
actions with  the  long-tailed  weasel.  In  the 
southwestern  U.S.,  it  is  absent  from  deserts  and  de- 
sert grasslands. 


The  least  weasel  feeds  almost  entirely  on  small 
rodents,  primarily  Microtus  sp.  (Ewer  1973;  Haley 
1975),  and  prefers  meadows,  brushy  areas,  and  open 
woods.  Ermine  or  short-tailed  weasels  are  found  in 
habitats  similar  to  those  of  least  weasels,  but  their 
distribution  extends  farther  north  and  south. 


Working  in  southern  Ontario,  Simms  ( 1979)  and 
King  ( 1983)  noted  that  short-tailed  weasels  selected 
early  successional  communities  and  avoided  forested 
habitats,  whereas  long-tailed  weasels  showed  no 
habitat  preferences.  Similarly,  Larrison  and  Johnson 
(1981 )  reported  that  the  short-tailed  weasel  in  Idaho 
shuns  the  more  rugged  mountains  and  higher  foothill 
ridges  and  occupies  stream  bottoms,  rocky  slides, 
fence  rows,  and  shrub  growth  near  water,  as  well  as 


POPULATION  MEASUREMENT  TECHNIQUES 

Low  densities  and  nonrandom  distributions  of 
carnivore  populations  create  sampling  problems 
when  estimating  population  sizes.  Carnivores  are 
often  distributed  in  a  nonrandom  manner  because 
they  prefer  some  specific  type  of  habitat  or  because 
of  social  interactions  between  individuals  or  groups. 
Sampling  problems  are  often  compounded  by  the 
high  mobility  and  wariness  of  most  carnivore  spe- 
cies. No  single  or  best  technique  has  yet  been  devel- 
oped that  can  suitably  census  all  carnivore 
populations.  Basic  sampling  schemes  and  census 
methods  indicate  either  presence  or  absence  of  a 
species,  relative  abundance,  or  absolute  abundance 
(reviewed  by  Lewis  1970;  Seber  1973;  Sen  1982; 
Barrett  1983). 


Carnivores 


487 


How  and  when  a  biologist  should  measure  a 
carnivore  population  or  prey  base  depends  on  the 
type  of  information  needed.  If  information  is  re- 
quired to  assess  the  range  or  local  distribution  of  the 
species,  then  techniques  to  determine  presence  are 
satisfactory.  Techniques  to  obtain  relative  density 
estimates  should  be  implemented  after  strongly  de- 
clining harvests  by  trappers  and  hunters,  substantial 
animal  damage,  habitat  loss,  or  public  concern  for 
the  species.  Such  techniques  may  require  modifica- 
tions as  well  as  trial  and  error  to  achieve  the  desired 
results.  They  can  be  applied  to  most  species  within 
a  reasonable  time  frame.  An  estimate  of  absolute 
density  for  most  species  is  virtually  impossible.  The 
techniques  are  extremely  time-consuming  and  ex- 
pensive but  may  be  appropriate  when  a  carnivore  is 
faced  with  extinction.  These  techniques  are  normally 
outside  the  scope  of  most  management  agencies. 

The  need  to  assess  the  prey  base  is  most  appro- 
priate when  carnivores  are  being  censused  for  abso- 
lute density  estimates  and  where  knowledge  of  the 
prey  is  crucial  to  preserving  a  species  (e.g.,  the  prai- 
rie dog  and  black-footed  ferret).  Often  a  relative 
density  index  can  be  obtained  more  readily  for  prey 
species  than  for  carnivores.  Such  estimates  are  par- 
ticularly important  when  data  on  carnivores  reveal 
areas  of  low  and  high  abundance.  Importantly,  many 
other  ecological  factors  (e.g.,  denning  sites,  various 
aspects  of  cover,  access  to  subnivially  active  prey,  or 
interactions  with  other  species)  may  be  of  equal  or 
greater  importance.  Therefore,  sampling  the  prey 
base  may  only  be  appropriate  when  other  ecological 
factors  are  also  being  considered.  One  also  presup- 
poses a  knowledge  of  the  principal  prey,  and  rela- 
tively few  carnivores  depend  on  one  or  a  few 
species  of  prey  for  survival. 


Presence 

Sightings  or  sign  left  by  carnivores  indicate  their 
presence  in  a  particular  area.  Yocom  ( 1973,  1974) 
and  Yocom  and  McCollum  (  1973)  used  reports  of 
sightings  and  sign  to  indicate  the  presence  and  range 
of  wolverine,  marten,  and  fisher  in  California.  Tracks, 
scats,  and  landing  sites  of  river  otter  were  used  by 


Melquist  and  Hornocker  (1979)  to  identify  areas 
frequented  by  otters.  Harvest  records  indicate  geo- 
graphic areas  where  species  had  been  present.  Re- 
sponses to  man-made  stimuli,  such  as  signs  left  at 
scent  stations  or  howling  by  wolves  or  coyotes  to  si- 
rens, indicate  recent  or  immediate  presence  of 
carnivores. 


Some  states  provide  harvest  records  of  carnivores. 


Relative  Density 

Aerial  surveys,  bounty  and  harvest  records, 
questionnaires,  number  of  road  kills,  and  indexes 
developed  from  counts  of  sightings,  sign,  and  rate  of 
capture  or  catch  per  unit  effort  data  are  widely  used 
to  estimate  relative  abundance  of  carnivore  species. 
These  types  of  data  taken  over  time  reveal  popula- 
tion trends,  are  less  costly  than  censuses,  and  are 
usually  adequate  for  most  management  needs.  Nellis 
and  Keith  (1976)  and  Parker  (1973)  used  aerial 
surveys  to  estimate  coyote  and  wolf  densities,  re- 
spectively. However,  this  technique  is  not  always  re- 
liable for  wolves  (Miller  and  Russell  1977). 

Cahalane  (1964)  used  responses  from  mail  ques- 
tionnaires, based  mostly  on  bounty  and  harvest  rec- 
ords, to  estimate  population  trends  and  relative 
densities  of  mountain  lion,  grizzly  bears,  and  wolves 
throughout  North  America.  Fuller  and  Keith  (1980) 
obtained  trends  in  wolf  numbers  in  northeastern 
Alberta  from  records  of  trapping  and  poisoning.  Simi- 
larly, coyote  densities  were  estimated  from  fur  and 
hunter  harvest  estimates  obtained  from  annual  re- 
ports and  mail  and  telephone  contacts.  In  Oklahoma, 
Hatcher  and  Shaw  ( 1981 )  obtained  slightly  over  a 
50%  response  rate  on  3,000  questionnaires  sent  to 
farmers.  They  compared  the  results  of  this  mail  sur- 
vey with  those  of  two  scent-station  surveys  to  evalu- 
ate bobcat,  coyote,  and  gray  and  red  fox  populations. 


488 


Carnivores 


Their  results  suggested  that  mail  surveys  are  more 
effective  than  scent-station  surveys  in  estimating 
carnivore  populations.  Furthermore,  mail  surveys  are 
more  cost-effective.  Berg  et  al.  (1983)  mailed  ques- 
tionnaires to  trappers  and  mountain  lion  hunters, 
outfitters,  biologists  and  foresters,  and  ranchers  and 
farmers  to  obtain  data  on  distribution  and  relative 
densities  of  mountain  lions  in  Wyoming. 

Allen  and  Sargent  (  1975)  used  sightings  of  red 
fox  by  rural  mail  carriers  as  an  index  of  population 
trends  in  North  Dakota.  This  technique  correlated 
highly  (r  =  0.958)  with  population  estimates  de- 
rived by  aerial  searches  for  fox  families  in  the  spring. 
This  index  appears  best-suited  to  prairie  regions, 
where  foxes  are  relatively  conspicuous.  Its  accuracy, 
timing,  and  low  cost  make  this  technique  especially 
useful  in  management  programs. 

The  relative  abundance  of  coyotes  has  been 
estimated  by  scat  deposition  rates  (reviewed  by  An- 
delt  and  Andelt  1984).  Scat  deposition  rates  are  de- 
termined from  the  number  of  scats  found  along  a 
given  length  of  road  during  a  known  period.  Before 
coyote  density  indexes  and  trends  in  abundance  can 
be  reliably  estimated  from  scat  deposition  rates,  the 
effect  of  diet  on  deposition  rates  must  be  deter- 
mined. Numbers  of  interstate  road-kills  have  also 
been  used  to  determine  the  relative  density  of  wild- 
life, including  the  coyote  (Case  1978). 

Elicited  responses  to  man-made  stimuli  are  often 
used  to  estimate  relative  densities  of  carnivore  popu- 
lations, especially  those  of  coyotes.  Scent-station 
indexes  have  recently  gained  popularity  as  a  means 
of  estimating  seasonal  and  annual  trends  in  relative 
abundance  of  carnivores  (reviewed  by  Conner  et  al. 
1983).  Richards  and  Hine  (1953)  and  Wood  (1959) 
originally  developed  this  method  for  determining  the 
relative  abundance  of  red  and  gray  foxes.  Humphrey 
and  Zinn  (1982)  modified  it  for  river  otter  and  mink. 
Lindzey  and  Knowlton  ( 1975)  used  it  extensively 
for  coyotes  in  17  western  states.  Several  hundred 
scent-station  lines,  each  consisting  of  50  scent  sta- 
tions located  at  0.48-km  (0.3-mi. )  intervals  along 
a  23.5-km  (14.7-mi.)  route,  were  checked  for  five 
consecutive  days  in  September.  Each  station  con- 
tained a  perforated  plastic  capsule  containing  a  fer- 
mented-egg  attractant  placed  in  the  center  of  a  1  - 
m  (1-yd)  circle  of  sifted  dirt.  Animal  visits  (based  on 
tracks )  were  recorded  to  provide  an  index  by  which 
coyote  population  trends  could  be  compared  be- 
tween regions,  states,  and  years.  The  index  of  rela- 
tive abundance  was  calculated  as  follows: 


Index  =   1,000  x 


total  animal  visits 


total  operative  station  nights 


The  presence  of  many  other  carnivore  species 
was  recorded  at  scent  stations,  but  because  survey 
lines  were  located  primarily  in  coyote  habitat,  these 


records  are  not  likely  a  measure  of  relative  abun- 
dance of  these  species.  With  modifications,  similar 
scent  stations  could  be  used  to  estimate  relative 
densities  of  these  carnivore  species.  Lindzey  and 
Meslow  ( 1977)  modified  this  technique  to  make  it 
more  selective  for  black  bears,  and  Clark  and  Camp- 
bell (  1983)  modified  it  for  small  carnivores.  Conner 
et  al.  (1983)  ran  permanent  scent-station  transects  in 
Florida  and  evaluated  the  indexes  to  determine  sea- 
sonal and  annual  trends  in  abundance  of  bobcats, 
raccoons,  and  gray  foxes.  They  found  this  technique 
provided  a  reliable  index  of  species  abundance  when 
compared  to  estimates  based  on  trapping,  radioiso- 
tope tagging,  and  radiotelemetry.  Despite  the  utility 
of  this  technique,  Roughton  and  Sweeny  (1982)  and 
Conner  et  al.  ( 1983)  recognized  it  needed  to  be 
standardized  and  verified  before  between-area  com- 
parisons could  be  made.  Furthermore,  between-spe- 
cies  comparisons  of  population  densities  could  not 
be  made  directly  with  scent-station  indexes  because 
of  species-specific  visitation  rates. 

Siren-elicited  vocalization  has  also  gained  inter- 
est as  a  possible  method  for  estimating  relative  abun- 
dance of  coyotes  ( reviewed  by  Okoniewski  and 
Chambers  1984).  Wenger  and  Cringan  (1978)  evalu- 
ated the  technique  in  northeastern  Colorado  by  us- 
ing six  radio-instrumented  coyotes.  Three  of  the 
coyotes  readily  responded  to  a  siren,  but  the  other 
three  rarely  responded.  These  results  substantiated 
the  response  rate  assumed  by  Alcorn  ( 1971 )  for 
coyotes  in  Nevada.  However,  further  research  is 
needed  to  determine  if  an  average  response  rate  ex- 
ists for  all  areas.  With  yearly  sampling  on  a  given 
area,  this  technique  should  provide  good  indexes  of 
locally  specific,  annual  coyote  fluctuations.  Andrews 
and  Boggess  (1978)  reported  two  advantages  of 
using  this  technique  for  coyotes  over  the  more  pop- 
ular scent-station  survey:  ( 1 )  coyote  and  dog  vocali- 
zations are  more  easily  distinguished  than  coyote 
and  dog  tracks  and  (2)  the  elicited-vocalization  tech- 
nique is  not  complicated  by  regional  variation  in 
coyote  use  or  avoidance  of  roadways.  Okoniewski 
and  Chambers  (  1 984 )  considered  the  advantages  of 
the  elicited-vocalization  technique  to  be  especially 


Carnivores 


489 


important  in  areas  where  dog  densities  are  high  and 
habitat  types  diverse.  On  the  other  hand,  specialized 
equipment  is  needed,  and  a  substantial  amount  of 
field  effort  and  time  is  required. 

Catch  per  unit  effort  has  also  been  used  to  esti- 
mate the  relative  densities  of  coyotes  (Clark  1972; 
Knowlton  1972).  Clark  (  1972)  used  the  success  of  a 
summer  trapping  program,  expressed  as  coyotes 
captured  per  1,000  trap  days,  as  one  relative  index 
of  a  postbreeding  coyote  population.  Knowlton 
(1972)  used  humane  "coyote-getters,"  set  in  stand- 
ard lines,  as  a  similar  technique  to  quantify  relative 
coyote  abundance.  Live-trapping  techniques  can  also 
be  used  to  estimate  carnivore  relative  densities. 
Simms  (  1979)  live-trapped  species  of  weasels  to  de- 
termine their  habitat  preferences  and  population 
densities.  Trapping  success  was  measured  as  the 
number  of  captures  per  100  trap  days.  A  trap  day 
was  24  hours,  from  one  morning  check  to  the  next. 

Barrett  (1983)  used  smoked  aluminum  tracking 
stations  to  sample  carnivores  in  California.  The  ad- 
vantages of  the  technique  are  the  ease  in  transport- 
ing and  the  ability  to  identify  tracks.  The  major 
disadvantage  is  similar  to  that  of  other  bait-based 
techniques  ( variation  in  attractability  of  many 
species). 


den,  or  other  sign.  A  complete  count  of  all  individu- 
als in  a  population  is  a  census  in  the  true  meaning  of 
the  word.  This  is  very  difficult  for  most  carnivore 
populations.  An  exception  may  be  locating  wolf 
packs  by  radiotelemetry  and  then  counting  entire 
packs  from  aircraft  (Fuller  1982).  However,  counting 
packs  of  wolves  may  not  account  for  the  number  of 
transients,  which  in  wolves  make  up  8  to  20%  of  the 
population  (Mech  and  Frenzel  1971). 

The  Lincoln  or  Petersen  Index  method  is  a  cap- 
ture-recapture technique  commonly  used  to  estimate 
population  sizes  of  many  species  of  wildlife,  includ- 
ing carnivores  (Lewis  1970;  Seber  1973;  Davis  and 
Winstead  1980).  The  word  "index"  is  a  misnomer 
because  this  technique  provides  an  estimate  of  total 
population.  Accuracy  of  this  technique  depends  on 
the  proportion  of  the  population  sampled.  The 
Schnabel  Method  is  a  variation  of  the  Lincoln  Index. 
With  this  method,  samples  are  taken  over  time  so 
population  estimates  can  be  averaged.  With  the  Lin- 
coln or  Petersen  Index  method,  M  individuals  from  a 
population  are  caught,  marked,  and  released.  On  a 
second  occasion,  a  sample  of  n  individuals  is  cap- 
tured. If  m  is  the  number  of  marked  animals  in  this 
sample,  an  estimate  N2  of  the  population  size  N  and 
an  estimate  v(N?)  of  its  variance  are  calculated  as 
follows: 


Absolute  Density 

Direct  counts  and  capture-recapture  techniques 
are  most  commonly  used  to  determine  absolute  den- 
sity of  wildlife  species.  Ilie  term  "direct  count"  re- 
fers to  counting  each  animal  rather  than  its  tracks, 


490 


Carnivores 


$     =  n_(M) 

2       m 


v(N2) 


Nr(N2  -  Af)(N2  -  n) 
Mn  ( N_>  -  1  ) 


!■ 


Sign  left  by  carnivores  has  also  been  used  to 
determine  absolute  density.  Brand  et  al.  (1976)  be- 
lieved the  number  of  lynx  and  their  movements 
could  be  accurately  determined  on  a  study  area  by 
track  observations.  After  snowfalls,  they  drove  along 
all  roads  in  their  study  areas  each  day  and  recorded 
lynx  tracks  across  roads  until  conditions  prevented 
identifying  recent  tracks. 

Koford  (  1977)  derived  estimates  of  mountain 
lion  populations  for  several  study  areas  in  California, 
based  on  the  number  of  different  sets  of  tracks  de- 
tected per  160  km  (100  mi.)  of  road  and  trail  track- 
ing. By  comparing  these  results  with  the  total 
number  of  mountain  lions  detected  by  intensive 
searches  of  each  study  area,  track  counts  were  cali- 
brated to  indicate  numbers  of  mountain  lions  per 
260  km2  (100  mi.2).  Koford  ( 1977)  and  Shaw 
(1979)  believed  all  mountain  lions  on  a  study  area 
could  be  distinguished  by  the  shape  and  size  of  their 
footprints.  Both  authors  reported  that  the  width  of 
track  heel  pads  was  the  most  useful  measurement  to 
be  taken. 

Sightings  of  fox  dens  from  the  ground  ( Scott 
and  Seiko  1939)  and  from  the  air  (Sargent  et  al. 
1975)  have  been  used  to  census  red  fox  populations 
in  Iowa  and  North  Dakota,  respectively.  Sargent  et 
al.  (1975)  flew  systematic  aerial  searches  to  locate 
red  foxes  and  their  dens.  The  number  of  individual 
fox  families  was  used  as  a  census  of  the  population. 

DISCUSSION 

Carnivores  are  a  diverse  group  of  mammals. 
Their  respective  habitats  and  food  requirements  vary 


greatly  (Table  1 ).  However,  to  accurately  relate  habi- 
tat to  any  species  or  animal  community,  biologists 
must  be  able  to  identify  environmental  features  that 
accurately  reflect  a  species'  or  community's  require- 
ments. Moreover,  changes  in  these  environmental 
features  should  allow  biologists  to  make  accurate 
predictions  of  a  species'  response  to  habitat  altera- 
tion (Anderson  1982),  either  directly  or  indirectly, 
through  a  change  in  prey  base.  Usually  dominant 
soils,  plant  species,  and  landforms  have  been  listed 
and  qualitatively  described.  Although  these  descrip- 
tions convey  much  information  about  an  individual 
habitat  type,  they  do  not  allow  correlations  between 
specific  environmental  features  and  a  species'  re- 
quirements. Quantitatively  measuring  many  habitat 
features  allows  statistical  correlation  between  spe- 
cific habitat  components  and  presence  and  abun- 
dance of  a  wildlife  species. 

Habitat  requirements  or  preferences  of  carni- 
vore species  can  be  quantified  with  current  technol- 
ogy. Soil,  vegetation,  and  landforms  can  be 
quantitatively  evaluated  on  the  ground  by  techniques 
of  Nudds  ( 1977)  and  Hays  et  al.  ( 1981 )  and  those 
reviewed  by  Anderson  (1982).  Large-scale  aerial 
photography  (Aldrich  1979;  Ulliman  et  al.  1979)  and 
satellite  imagery  (Varney  1973;  Brabander  and  Bar- 
clay 1977;  Aldrich  1979;  Cannon  et  al.  1982;  Craig- 
head et  al.  1982)  can  be  used  to  classify,  inventory, 
and  monitor  general  vegetative  cover  types  (e.g., 
alpine  meadows,  forest,  grassland,  shrubland).  Devel- 
opment of  these  remote-sensing  techniques  to  evalu- 
ate vegetation  types  and  land  uses  has  greatly 
reduced  the  logistical  effort  and  the  cost  of  survey- 
ing and  monitoring  extensive  areas  by  conventional 
methods. 


Carnivores 


491 


LITERATURE  CITED 


ABLES,  E.D.  1975.  Ecology  of  the  red  fox  in  America. 

Pages  216-236  in  Fox,  M.W.  ed.  The  Wild  Canids.  Van 
Nostrand  Reinhold  Co.  New  York,  NY. 

ACKERMAN,  B.B.  1982.  Cougar  predation  and  ecological 
energetics  in  southern  Utah.  M.S.  Thesis.  Utah  State 
Univ.,  Logan.  103pp. 

,  F.G.  LINDZEY,  and  T.P.  HEMKER.  1984.  Cougar 

food  habits  in  southern  Utah.  J.  Wildl.  Manage. 
48:147-155. 

ALCORN,  JR.  1971.  A  discussion  of  coyote  censusing 

techniques.  U.S.  Bur.  Sport  Fish,  and  Wildl.,  Div.  Wildl. 
Serv.  8pp. 

ALDRICH,  R.C.  1979.  Remote  sensing  of  wildland  re- 
sources: A  state-of-the-art  review.  U.S.  Dep.  Agric,  For. 
Serv.  Gen  Tech.  Rep.  RM-71.  56pp. 

ALLEN,  AW.  1982.  Habitat  suitability  index  models:  Mar- 
ten. U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  FWS/OBS 
82/10.11.  9pp. 

.  1983.  Habitat  suitability  index  models:  Fisher.  U.S. 

Dep.  Inter.,  Fish  and  Wildl.  Serv.  FWS/OBS-82/0.45. 
20pp. 

ALLEN,  S.H.  and  A.B.  SARGENT.  1975.  A  rural  mail-carrier 
index  of  North  Dakota  red  foxes.  Wildl.  Soc.  Bull. 
3:74-77. 

AMSTRUP,  S.C.  and  J.  BEECHAM.  1976.  Activity  patterns 
of  radio-collared  black  bears  in  Idaho.  J.  Wildl.  Man- 
age. 40:340-348. 

ANDELT,  W.F.  and  S.H.  ANDELT.  1981.  Habitat  use  by 
coyote  in  southeastern  Nebraska.  J.  Wildl.  Manage. 
45(4):  1001-1005. 

and .  1984.  Diet  bias  in  scat  deposition-rate 

surveys  of  coyote  density.  Wildl.  Soc.  Bull.  12:74-77. 

ANDERSON,  A.E.  1983-  A  critical  review  of  literature 
on  puma  (Felis  concolor).  Colo.  Div.  Wildl.  Spec. 
Rep.  54.  91pp. 

ANDERSON,  S.H.  1982.  Comments  on  measurement  of 
habitat.  Pages  278-280  in  Davis,  D.E.  ed.  CRC  Hand- 
book of  Census  Methods  for  Terrestrial  Vertebrates. 
CRC  Press,  Inc.  Boca  Raton,  FL. 

ANDREWS,  R.D.  and  E.K.  BOGGESS.  1978.  Ecology  of 

coyotes  in  Iowa.  Pages  249-265  in  Bekoff,  M.  ed.  Coy- 
otes Biology,  Behavior  and  Management.  Academic 
Press.  New  York,  NY. 

BAILEY,  T.N.  1974.  Social  organization  in  a  bobcat  popula- 
tion. J.  Wildl.  Manage.  38:435-445. 

BANFIELD,  A.W.F.  1974.  The  Mammals  of  Canada.  Univ. 
Toronto  Press,  Toronto.  438pp. 

BARNES,  C.T.  I960.  The  cougar  or  mountain  lion.  The 
Ralton  Co.  Salt  Lake  City,  UT.  176pp. 

BARRETT,  R.H.  1983.  Smoked  aluminum  track  plots  for 
determining  furbearer  distribution  and  relative  abun- 
dance. California  Fish  and  Game  69:188-190. 

BEASON,  S.L.  and  R.A.  MOORE.  1977.  Bobcat  food  habit 
response  to  a  change  in  prey  abundance.  Southwest. 
Nat.  21:451-457. 

BERG,  R.L.,  L.L.  MCDONALD,  and  M.D.  STRICKLAND. 

1983  Distribution  of  mountain  lions  in  Wyoming  as 
determined  by  mail  questionnaire.  Wildl.  Soc.  Bull. 
11:265-268. 

BERG,  W.K  and  R.A.  CHESNESS.  1978.  Ecology  of  coyotes 
in  northern  Minnesota.  Pages  229-247  in  Bekoff,  M. 
ed.  Coyotes  Biology,  Behavior  and  Management.  Aca- 
demic Press.  New  York,  NY. 


BERNER,  A.  and  L.W.  GYSEL.  1967.  Raccoon  use  of  large 
tree  cavities  and  ground  burrows.  J.  Wildl.  Manage. 
31:706-714. 

BRABANDER,  J.J.  and  J.S.  BARCLAY.  1977.  A  practical 
application  of  satellite  imagery  to  wildlife  habitat 
evaluation.  Proc.  Ann.  Conf.  Southeast  Assoc.  Fish 
Wildl.  Agencies  31:300-306. 

BRADLEY,  W.G.  and  C.G.  HANSEN.  1965.  Observations  on 
the  distribution  of  the  ring-tailed  cat  in  southern 
Nevada.  Southwest.  Nat.  10:310-311. 

BRAND,  C.J.  and  LB.  KEITH.  1979-  Lynx  demography 
during  a  snowshoe  hare  decline  in  Alberta.  J.  Wildl. 
Manage.  40:416-428. 

,  LB.  KEITH,  and  C.A.  FISCHER.  1976.  Lynx  re- 
sponse to  changing  snowshoe  hare  densities  in  cen- 
tral Alberta.  J.  Wildl.  Manage.  40:416-428. 

BRAY,  O.E.  and  V.G.  BARNES.  1967.  A  literature  review  on 
black  bear  populations  and  activities.  U.S.  Dep.  Inter., 
Natl.  Park  Serv.,  and  Colo.  Coop.  Wildl.  Res.  Unit. 
34pp. 

BROWN,  D.E.  1983.  The  wolf  in  the  Southwest.  Univ. 
Arizona  Press,  Tucson.  195pp. 

BUCK,  S.,  C.  MULLIS,  and  A.  MOSSMAN.  1978.  A  radio 
telemetry  study  of  fishers  in  northwestern  California. 
Cal.-Nev.  Wildl.  Trans.  1979:166-172. 

BURNS,  J.J.  1964.  Comparisons  of  two  populations  of  mink 
from  Alaska.  Can.  J.  Zool.  42:1071-1079. 

BUSKIRK,  S.W.  1984.  Seasonal  use  of  resting  sites  by 
marten  in  south-central  Alaska.  J.  Wildl.  Manage. 
48:950-953. 

CAHALANE,  V.H.  1964.  A  preliminary  study  of  distribution 
and  numbers  of  cougar,  grizzly  and  wolf  in  North 
America.  NY.  Zool.  Soc.  12pp. 

CANNON,  R.W.,  FL.  KNOPF,  and  L.R  PETTINGER.  1982. 
Use  of  landsat  data  to  evaluate  lesser  prairie  chicken 
habitats  in  western  Oklahoma.  J.  Wildl.  Manage. 
46:915-922. 

CARBYN,  L.N.  1982.  Coyote  population  fluctuations  and 
spatial  distribution  in  relation  to  wolf  territories  in 
Riding  Mountain  National  Park.  Can.  Field-Nat. 
96:176-183- 

CASE,  R.M.  1978.  Interstate  highway  road-killed  animals,  a 
data  source  for  biologists.  Wildl.  Soc.  Bull.  6:8-13- 

CLARK,  F.W.  1972.  Influence  of  jackrabbit  density  on 
coyote  population  change.  J.  Wildl.  Manage.  36:343- 
356. 

CLARK,  T.W.  and  T.M.  CAMPBELL  1983.  A  small  carnivore 
survey  technique.  Great  Basin  Nat.  43:438-440. 

COLE,  G.F.  1972.  Grizzly  bear-elk  relationships  in  Yellow- 
stone National  Park.  J.  Wildl.  Manage.  36:556-561. 

CONNER,  M.C.,  R.F.  LABISKY,  and  DR.  PROGULSKE,  Jr. 
1983-  Scent-station  indices  as  measures  of  population 
abundance  for  bobcats,  raccoons,  gray  foxes,  and 
opossums.  Wildl.  Soc.  Bull.  11:146-152. 

COWAN,  I.  McT.  1947.  The  timber  wolf  in  the  Rocky 
Mountain  National  Parks  of  Canada.  Can.  J.  Res. 
25:139-174. 

and  R.H.  MACKAY.  1959.  Food  habits  of  marten 

(Mattes  americana)  in  the  Rocky  Mountain  region  of 
Canada.  Can.  Field-Nat.  64:100-104. 

CRABB,  W.G.  1948.  The  ecology  and  management  of  the 
prairie  spotted  skunk  in  Iowa.  Ecol.  Monogr.  18:201- 
233. 

CRAIGHEAD,  J.J,  J.S.  SUMMER,  and  G.B.  SCAGGS.  1982.  A 
definitive  system  for  analysis  of  grizzly  bear  habitat 
and  other  wilderness  resources  utilizing  IANDSAT 
multispectral  imagery  and  computer  technology. 


492 


Carnivores 


Wildlife-Wildlands  Inst.  Monogr.  1,  Univ.  Montana 

Foundation,  Missoula. 
CUMBIE,  P.M.  1975.  Mercury  levels  in  Georgia  otter,  mink 

and  freshwater  fish.  Environ.  Contam.  Toxicol. 

14:193-196. 
DAVIS,  D.E.  and  R.L.  WINSTEAD.  1980.  Estimating  the 

number  of  wildlife  populations.  Pages  221-246  in 

Giles,  R.H.  ed.  Wildlife  Management  Techniques  Man- 
ual. The  Wildl.  Soc.  Washington,  DC. 
DE  VOS,  A.  1952.  Ecology  and  management  of  fisher  and 

marten  in  Ontario.  Tech.  Bull.  Ontario  Dep.  Lands  and 

Forests. 
DOUGLASS,  R.J.,  L.G.  FISHER,  and  M.  MAIR.  1983.  Habitat 

selection  and  food  habits  of  marten,  Martes  atneri- 

cana,  in  the  Northwest  Territories.  Can.  Field-Nat. 

97:71-74. 
EGOSCUE,  H.J.  1962.  Ecology  and  life  history  of  the  kit 

fox  in  Tooele  County,  Utah.  Ecology  43:481-497. 
ELLIS,  R.J.  1964.  Tracking  raccoons  by  radio.  J.  Wildl. 

Manage.  28:363-368. 
ERRINGTON,  P.L.  1935.  Food  habits  of  mid-west  foxes.  J. 

Mammal.  16:192-200. 
EWER,  R.F.  1973.  The  Carnivores.  Cornell  Univ.  Press. 

Ithaca,  NY.  494pp. 
FLOYD,  B.L.  and  MR.  STROMBERG.  1981.  New  records  of 

the  swift  fox  (Vulpes  velox)  in  Wyoming.  J.  Mammal. 

62:650-651. 
FRITZELL,  E.K  1978.  Habitat  use  by  prairie  raccoons 

during  the  waterfowl  breeding  season.  J.  Wildl.  Man- 
age. 42:118-127. 
FULLER,  T.K  1982.  Wolves.  Pages  225-226  in  Davis,  D.E. 

ed.  CRC  Handbook  of  Census  Methods  for  Terrestrial 

Vertebrates.  CRC  Press,  Inc.  Boca  Raton,  FL. 
and  LB.  KEITH.  1980.  Wolf  population  dynamics 

and  prey  relationships  in  northeastern  Alberta.  J. 

Wildl.  Manage.  44:583-602. 
GAMBLE,  R.L.  1981.  Distribution  in  Manitoba  of  Mustela 

Frenata  Longicanda,  the  long-tailed  weasel,  and  the 

interrelation  of  distribution  and  habitat  selection 

in  Manitoba,  Saskatchewan  and  Alberta.  Can.  J.  Zool. 

59:1036-1039. 
GASAWAY,  W.C.,  R.O.  STEPHENSON,  J.L.  DAVIS,  P.E.K. 

SHEPHERD,  and  O.E.  BURRIS.  1983.  Interrelationships 

of  wolves,  prey,  and  man  in  interior  Alaska.  Wildl. 

Monogr.  84.  50pp. 
GASHWILER,  J.,  W.L.  ROBINETTE,  and  O.W.  MORRIS. 

1961.  Breeding  habits  of  bobcats  in  Utah.  J.  Mammal. 

42:76-84. 
GIER,  H.T.  1975.  Ecology  and  behavior  of  the  coyote 

(Canis  latrans).  Pages  246-262  in  Fox,  M.W.  ed.  The 

Wild  Canids.  Van  Nostrand  Reinhold  Co.  New  York, 

NY. 
GILBERT,  F.F.  and  E.G.  NANCEKJVELL.  1982.  Food  habits 

of  mink  (Mustelavison)  and  otter  (Lutra  canadensis) 

in  northeastern  Alberta  Canada.  Can.  J.  Zool.  60:1282- 

1288. 
GRABER,  DM.  1982.  Ecology  and  management  of  black 

bear  in  Yosemite  National  Park.  Cooperative  National 

Park  Resources  Study  Unit.  Univ.  of  California  at 

Davis.  Tech.  Rep.  5. 
GREEN,  J.S.  and  J.T.  FLINDERS.  1981.  Diets  of  sympatric 

red  foxes  and  coyotes  in  southeastern  Idaho.  Great 

Basin  Nat.  41:251-254. 
GREENWOOD,  R.J.  1982.  Nocturnal  activity  and  foraging 

of  prairie  raccoons  in  North  Dakota.  Am.  Midi.  Nat. 

107:238-243- 
GREER,  KR.  1955.  Yearly  food  habits  of  the  river  otter  in 


the  Thompson  I.akes  region,  northwestern  Montana, 
as  indicated  by  scat  analysis.  Am.  Midi.  Nat.  54:299- 
313. 

GRENFELL,  W.E.  and  A.J.  BRODY.  1983.  Seasonal  foods  of 
black  bears  in  Tahoe  National  Forest,  California.  Calif. 
Fish  and  Game  69:132-150. 

and  M.  FASENFEST.  1979.  Winter  food  habits  of 

fishers,  Martes  pennanti,  in  northwestern  California. 
Calif.  Fish  Game  65:186-189. 

GRINNELL,  J,  J.S.  DIXON,  andJ.M.  LINSDALE.  1937.  Fur- 
bearing  mammals  of  California,  Vol.  I.  Univ.  Calif. 
Press,  Berkeley.  375pp. 

HALEY,  D.  1975.  Sleek  and  savage.  North  America's  weasel 
family.  Pacific  Search.  Seattle,  WA.  298pp. 

HALL,  E.R.  1951.  Weasels.  Univ.  Kansas  Publ.  Mus.  Nat. 
Hist.  No.  4. 

and  KR.  KELSON.  1959.  The  mammals  of  North 

America,  Vol.  II.  The  Ronald  Press.  New  York,  NY. 

HARGIS,  CD.  and  DR.  McCULLOUGH.  1984.  Winter  diet 
and  habitat  selection  of  marten  in  Yosemite  National 
Park.  J.  Wildl.  Manage.  48:140-146. 

HARLOW,  H.J.  1981.  Torpor  and  other  physiological  adap- 
tations of  the  badger  (Taxidea  taxus)  to  cold  envi- 
ronments. Physiol.  Zool.  54:267-276. 

HART,  E.B.  1982.  The  raccoon  (Procyon  lotor)  in  Wyo- 
ming. Great  Basin  Nat.  42:599-600. 

HATCHER,  R.T.  and  J.H.  SHAW.  1981.  A  comparison  of 
three  indices  to  furbearer  populations.  Wildl.  Soc. 
Bull.  9:153-156. 

HAWTHORNE,  V.M.  1972.  Coyote  food  habits  in  Sagehen 
Creek  basin,  northeastern  California.  Calif.  Fish  and 
Game  58:4-12. 

HAYS,  R.L,  C.  SUMMERS,  and  W.  SEITZ.  1981.  Estimating 
wildlife  habitat  variables.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  FWS/OBS-81/47.  Washington,  DC.  111pp. 

HEMKER,  T.P.  1982.  Population  characteristics  and  move- 
ment patterns  of  cougars  in  southern  Utah.  M.S.  The- 
sis. Utah  State  Univ.,  Logan.  59pp. 

HENRY,  C.J.,  L.J.  BLUS,  S.V.  GREGORY,  and  C.J.  STAF- 
FORD. 1981.  PCBs  and  organochlorine  pesticides  in 
wild  mink  and  river  otters  from  Oregon.  Pages  1 763- 
1789  in  Chapman,  J.A.  and  Pursley,  D.  eds.  World- 
wide Furbearer  Conf.  Proc.  1980.  Frostburg,  MD. 

HENSHAW,  R.E.  1975.  Reintroduction  of  wolves  into  the 
wild.  Pages  420-444  in  Klinghammer,  E.  ed.  The 
Behavior  and  Ecology  of  Wolves.  Symp.  Proc.  Garland 
STPM  Press.  New  York,  NY. 

,  R.  LOCKWOOD,  R.  SHIDELER,  and  R.O.  STEPHEN- 
SON. 1975.  Experimental  release  of  captive  wolves. 
Pages  319-345  in  Klinghammer,  E.  ed.  The  Behavior 
and  Ecology  of  Wolves.  Symp.  Proc.  Garland  STPM 
Press,  New  York,  NY. 

and  R.O.  STEPHENSON.  1974.  Homing  in  the  gray 

wolf  (Canis  lupus).).  Mammal.  55:234-237. 

HERRERO,  S.  1972.  Aspects  of  evolution  and  adaptation  in 
American  black  bears  (Ursns  americanus  Pallus)  and 
brown  and  grizzly  bears  ( U.  asctos  Linne )  of  North 
America.  Pages  221-231  in  Herrero,  S.  ed.  Bears — 
Their  Biology  and  Management.  New  Series  23,  Inter- 
national Union  for  Conservation  of  Nature.  Morges, 
Switzerland. 

.  1978.  A  comparison  of  some  features  of  the  evolu- 
tion, ecology  and  behavior  of  black  and  grizzly/brown 
bears.  Carnivore  1:7-17. 

HILTON,  H.  1978.  Systematics  and  ecology  of  the  eastern 
coyote.  Pages  209-228  in  Bekoff,  M.  ed.  Coyotes 
Biology,  Behavior  and  Management.  Academic  Press. 


Carnivores 


493 


New  York,  NY. 

HOFFMAN,  CO.  and  J.L.  GOTTSCHANG.  1977.  Numbers, 
distribution,  and  movements  of  a  raccoon  population 
in  a  suburban  residential  community.  J.  Mammal. 
58:623-636. 

HOLLEMANN,  D.F.  and  R.O.  STEPHENSON.  1981.  Prey 
selection  and  consumption  by  Alaskan  wolves  in 
winter.  J.  Wildl.  Manage.  45:620-628. 

HORNOCKER,  M.G.  1969.  Winter  territoriality  in  moun- 
tain lions.  J.  Wildl.  Manage.  33:457-464. 

.  1970.  An  analysis  of  mountain  lion  predation  upon 

mule  deer  and  elk  in  the  Idaho  Primitive  Area.  Wildl. 
Monogr.  21.  39pp. 

and  H.S.  HASH.  1981.  Ecology  of  the  wolverine  in 

northwestern  Montana.  Can.  J.  Zool.  59:1286-1301. 

HOUSTON,  D.B.  1978.  Elk  as  winter-spring  food  for  carni- 
vores in  northern  Yellowstone  National  Park.  J.  Appl. 
Ecol.  15:653-662. 

HUMPHREY,  S.R.  and  T.L.  ZINN.  1982.  Seasonal  habitat 
use  by  river  otters  and  Everglades  mink  in  Florida.  J. 
Wildl.  Manage.  46:375-381. 

INGLES,  L.G.  1965.  Mammals  of  the  Pacific  States.  Stanford 
Univ.  Press,  Stanford,  CA.  506pp. 

JOHNSON,  M.K  and  R.M.  HANSEN.  1979.  Coyote  food 

habits  on  the  Idaho  Natural  Engineering  Laboratory.  J. 
Wildl.  Manage.  43:951-955. 

JONES,  J.H.  and  N.S.  SMITH.  1979.  Bobcat  density  and 
prey  selection  in  central  Arizona.  J.  Wildl.  Manage. 
43:666-672. 

JONKEL,  J.  and  I.  McT.  COWAN.  1971.  The  black  bear  in 
the  spruce-fir  forest.  Wildl.  Monogr.  27. 

JORDAN,  PA.,  PC.  SHELTON,  and  D.L.  ALLEN.  1967. 
Numbers,  turnover,  and  social  structure  of  the  Isle 
Royale  wolf  population.  Am.  Zool.  7:233-252. 

KEITH,  L.B.  1963.  Wildlife's  ten-year  cycle.  Univ.  Wiscon- 
sin Press,  Madison.  201pp. 

KELLEYHOUSE,  D.G.  1980.  Habitat  utilization  by  black 
bears  in  northern  California.  Pages  221-228  in  Mar- 
tinka,  C.J.  and  KL.  McArthur  eds.  Bears — Their  Biol- 
ogy and  Management.  Bear  Biol.  Assoc.  Ser.  4. 

KILGORE,  D.L.,  Jr.  1969.  An  ecological  study  of  the  swift 
fox  (Vulpes  velox)  in  the  Oklahoma  Panhandle.  Am. 
Midi.  Nat.  81:512-534. 

KING,  CM.  1983-  Mustela  erminea.  Mammalian  Species 
195.  8pp. 

KNOWLTON,  F.F.  1972.  Preliminary  interpretation  of 

coyote  population  mechanics  with  some  management 
implications.  J.  Wildl.  Manage.  36:369-382. 

KOEHLER,  G.M.  and  M.G.  HORNOCKER.  1977.  Fire  ef 
fects  on  marten  habitat  in  the  Selway-Bitterroot  Wil- 
derness. J.  Wildl.  Manage.  41:500-505. 

, ,  and  H.S.  HASH.  1979.  Lynx  movements 

and  habitat  use  in  Montana.  Can.  Field-Nat.  93:441- 
442. 

KOFORD,  C.B.  1977.  Status  and  welfare  of  the  puma  (Felis 
concolor)  in  California,  1973-1976.  Final  Rep.  De- 
fenders of  Wildlife  and  the  National  Audubon  Society. 
Mus.  Vert.  Zool.  Univ.  Calif.,  Berkeley.  57pp. 

KOLENOSKY,  G.B.  1972.  Wolf  predation  on  wintering 

deer  in  east-central  Ontario.  J.  Wildl.  Manage.  36:257- 
369. 

KORSCHGEN,  L.J.  1959.  Food  habits  of  the  red  fox  in 
Missouri.  J.  Wildl.  Manage.  23:168-176. 

KUYT,  E.  1972.  Food  habits  and  ecology  of  wolves  on 

barren-ground  caribou  range  in  the  Northwest  Terri- 
tories. Can.  Wildl.  Serv.  Rep.  Ser.  21.  36pp. 

LAMPE,  R.P.  1982.  Food  habits  of  badgers  in  east-central 


Minnesota.  J.  Wildl.  Manage.  46:790-795. 
—  and  M.  SOVADA.  1981.  Seasonal  variation  in  home 


range  of  a  female  badger  (Taxidea  taxiis).  Prairie  Nat. 
13:55-58. 

LARRISON,  E.J.  and  D.R.JOHNSON.  1981.  Mammals  of 
Idaho.  Univ.  Idaho  Press,  Moscow.  166pp. 

LECOUNT,  A.L.  1980.  Some  aspects  of  black  bear  ecology 
in  the  Arizona  chaparral.  Pages  175-180  in  Martinka, 
C.J.  and  KL.  McArthur  eds.  Bears — Their  Biology 
and  Management.  Bear  Biol.  Assoc.  Ser.  4. 

LEWIS,  J.C  1970.  Wildlife  census  methods:  A  resume.  J. 
Wildl.  Dis.  6:356-364. 

LINDZEY,  F.G.  1978.  Movement  patterns  of  badgers  in 
northwestern  Utah.  J.  Wildl.  Manage.  42:418-422. 

and  F.F.  KNOWLTON.  1975.  Determining  the  rela- 
tive abundance  of  coyotes  by  scent  station  lines. 
Wildl.  Soc.  Bull.  3:119-124. 

and  E.C  MESLOW.  1977.  Home  range  and  habitat 


use  by  black  bears  in  southwestern  Washington.  J. 
Wildl.  Manage.  41:408-415. 

LINHART,  SB.  and  W.B.  ROBINSON.  1972.  Some  relative 
carnivore  densities  in  areas  under  sustained  coyote 
control.  J.  Mammal.  53:880-884. 

LONG,  CA.  1973-  Taxidea  taxus.  Mammalian  Species  26. 
The  Amer.  Soc.  of  Mammal.  Lawrence,  KS.  4pp. 

and  CA.  KILLINGLEY.  1983.  The  badgers  of  the 

world.  Charles  C.  Thomas,  Publisher.  Springfield,  IL. 
404pp. 

LOTZE,  J.H.  and  S.  ANDERSON.  1979.  Procyon  lotor. 

Mammalian  Species  119.  The  Amer.  Soc.  of  Mammal. 
Lawrence,  KS.  8pp. 

LYNCH,  G.M.  1974.  Some  den  sites  of  Manitoba  raccoons. 
Can.  Field-Nat.  88:494-495. 

MARTINKA,  C.J.  1976.  Ecological  role  and  management  of 
grizzly  bears  in  Glacier  National  Park,  Montana.  Pages 
147-156  in  Pelton,  MR,  J.W.  Lentfer,  and  G.E.  Folk 
eds.  Bears — Their  Biology  and  Management.  IUCN 
Publ.  New  Ser.  40.  Morges,  Switzerland. 

McCORD,  CM.  and  J.E.  CARDOZA.  1982.  Bobcat  and  lynx. 
Pages  728-766  in  Chapman,  J.A.  and  G.A.  Feldhamer 
eds.  Wild  Mammals  of  North  America:  Biology,  Man- 
agement and  Economics.  John  Hopkins  Univ.  Press. 
Baltimore,  MD. 

MEALY,  S.P.  1980.  The  natural  food  habits  of  grizzly  bears 
in  Yellowstone  National  Park,  1973-74.  Pages  281-292 
in  Martinka,  C.J.  and  KL.  McArthur  eds.  Bears — Their 
Biology  and  Management.  Bear  Biol.  Assoc.  Ser.  4. 

MECH,  L.D.  1966.  The  wolves  of  Isle  Royale.  U.S.  Dep.  In- 
ter., Natl.  Park  Serv.  Fauna  Ser.  7.  Washington,  DC. 
210pp. 

.  1975.  Some  considerations  in  re-establishing 

wolves  in  the  wild.  Pages  445-457  in  Klinghammer,  E. 
ed.  The  Behavior  and  Ecology  of  Wolves.  Symp.  Proc. 
Garland  STPM  Press.  New  York,  NY. 

.  1977.  Productivity,  mortality,  and  population 

trends  of  wolves  in  northeastern  Minnesota.  J.  Mam- 
mal. 58:559-574. 

and  L.D.  FRENZEL.  1971.  Ecological  studies  of  the 

timber  wolf  in  northeastern  Minnesota.  U.S.  Dep. 
Agric,  For.  Serv.  Res.  Pap.  NC-52.  62pp. 

and  P.D.  KARNS.  1977.  Role  of  the  wolf  in  deer 

decline  in  the  Superior  National  Forest.  U.S.  Dep. 
Agric,  For.  Serv.  Res.  Pap.  NC-51.  62pp. 

MELQUIST,  WE.  and  M.G.  HORNOCKER.  1979.  Methods 
and  techniques  for  studying  and  censusing  river  otter 
populations.  Univ.  Idaho  For.  Wildl.  and  Range  Exp. 
Sta.  Tech.  Rep.  8.  1 7pp. 


494 


Carnivores 


and 


-.  1983-  Ecology  of  river  otters  in  west- 


central  Idaho.  Wildl.  Monogr.  83-  60pp. 
—  J.S.  WHITMAN,  and  M.G.  HORNOCKER.  1981. 


Resource  partitioning  and  coexistence  of  sympatric 
mink  and  river  otter  populations.  Pages  187-220 
in  Chapman,  J.A.  and  D.  Pursley  eds.  Worldwide  Fur- 
bearer  Conf.,  Vol.  1.  Frostburg,  MD. 

MESSICK,  J.P.  and  M.G.  HORNOCKER.  1981.  Ecology  of 
the  badger.  Taxidea  taxns  in  southeastern  Idaho. 
Wildl.  Monogr.  76.  53pp. 

MILLER,  F.L.  and  R.H.  RUSSELL.  1977.  Unreliability  of  strip 
aerial  surveys  for  estimating  numbers  of  wolves  on 
western  Queen  Elizabeth  Islands,  Northwest  Territo- 
ries. Can.  Field-Nat.  91:77-82. 

MOORE,  RE.  and  N.S.  MARTIN.  1980.  A  recent  record  of 
the  swift  fox  (Vulpes  velox)  in  Montana.  J.  Mammal. 
61:161. 

MORE,  G.  1 976.  Some  winter  food  habits  of  lynx  (Felix 
lynx)  in  the  southern  Mackenzie  district,  Northwest 
Territories.  Can.  Field-Nat.  90:499-500. 

MURIE,  A.  1944.  The  wolves  of  Mount  McKinley.  U.S.  Dep. 
Inter.,  Natl.  Park  Serv.  Fauna  Ser.  5.  238pp. 

.  1951.  Coyote  food  habits  on  a  southwestern  cattle 

range.  J.  Mammal.  32:291-295. 

NELLIS,  C.H.  and  LB.  KEITH.  1968.  Hunting  activities  and 
success  of  lynx  in  Alberta.  J.  Wildl.  Manage.  32:718- 
722. 

and .  1976.  Population  dynamics  of  coyotes 

in  central  Alberta,  1964-68.  J.  Wildl.  Manage.  40:389- 
399. 

,  S.P.  WETMORE,  and  LB.  KEITH.  1972.  Lynx-prey 

interactions  in  central  Alberta.  J.  Wildl.  Manage. 
36:320-329. 

NOVICK,  H.J.,  J.M.  SIPEREK  and  G.R.  STEWART.  1981. 
Denning  characteristics  of  black  bears  (Ursus  ameri- 
canus)  in  the  San  Bernardino  mountains  of  southern 
California.  Calif.  Fish  and  Game  67:52-61. 

and  G.R.  STEWART.  1982.  Home  range  and  habitat 

preferences  of  black  bears  (Ursiis  americantis)  in 
the  San  Bernardino  Mountains  of  southern  California, 
U.S.A.  Calif.  Fish  and  Game  68:21-35. 

NUDDS,  ID.  1977.  Quantifying  the  vegetative  structure  of 
wildlife  cover.  Wildl.  Soc.  Bull.  5:113-117. 

OKONIEWSKI,  J.C.  and  RE.  CHAMBERS.  1984.  Coyote 
vocal  response  to  an  electronic  siren  and  human 
howling.  J.  Wildl.  Manage.  48:217-221. 

PARKER,  G.R.  1973.  Distribution  and  densities  of  wolves 
within  barren-ground  caribou  range  in  northern  main- 
land Canada.  J.  Mammal.  54:341-346. 

,  J.W.  MAXWELL,  L.D.  MORTON,  and  E.J.  SMITH. 

1983-  The  ecology  of  the  lynx  (Lynx  canadensis)  on 
Cape  Breton  Island.  Can.  J.  Zool.  61:770-786. 

PEDERSON,  J.C.  and  R.C.  TUCKFIELD.  1983.  A  compara- 
tive study  of  coyote  food  habits  on  two  Utah  deer 
herds.  Great  Basin  Nat.  43:432-437. 

PETERSON,  R.O.  1975.  The  wolves  of  Isle  Royale:  New 
developments.  Pages  3-18  in  Klinghammer,  E.  ed.  The 
Behavior  and  Ecology  of  Wolves.  Symp.  Proc.  Garland 
STPM  Press.  New  York,  NY. 

1977.  Wolf  ecology  and  prey  relationships  on  Isle 

Royale.  U.S.  Dep.  Inter.,  Natl.  Park  Serv.  Sci.  Monogr. 
Ser.  11.  210pp. 

PIEKIELEK,  W.  and  T.  BURTON.  1975.  A  black  bear  pop- 
ulation study  in  northern  California.  Calif.  Fish  and 
Game  61:4-25. 

PIMLOTT,  D.H.  1967.  Wolf  predation  and  ungulate  popu- 
lations. Am.  Zool.  7:267-278. 


POWELL,  R.A.  1982.  The  fisher,  life  history,  ecology,  and 
behavior.  Univ.  Minnesota  Press,  Minneapolis.  217pp. 

PULLIAINEN,  E.  1975.  Ecology  of  the  wolf  in  the  settled 
areas  of  Finland.  Pages  84-92  in  Klinghammer,  E. 
ed.  The  Behavior  and  Ecology  of  Wolves.  Symp.  Proc. 
Garland  STPM  Press.  New  York,  NY. 

RAINE,  R.M.  1982.  Ranges  of  juvenile  fisher  (Martes  pen  - 
nanti)  and  marten,  (Martes  americana)  in  southeast- 
ern Manitoba.  Can.  Field-Nat.  96:431-438. 

RAUSCH,  R.A.  and  A.M.  PEARSON.  1972.  Notes  on  the 

wolverine  in  Alaska  and  the  Yukon  Territory.  J.  Wildl. 
Manage.  36:249-268. 

RICHARDS,  S.H.  and  R.L.  HINE.  1953-  Wisconsin  fox  pop- 
ulations. Tech.  Wildl.  Bull.  6.  Wisconsin  Conserv. 
Dep.,  Madison. 

ROBINETTE,  W.L.,  J.S.  GASHWILER,  and  O.W.  MORRIS. 
1959.  Food  habits  of  the  cougar  in  Utah  and  Nevada. 
J.  Wildl.  Manage.  23:261-273. 

ROUGHTON,  R.D.  and  M.W.  SWEENY.  1982.  Refinements 
in  scent-station  methodology  for  assessing  trends  in 
carnivore  populations.  J.  Wildl.  Manage.  46:217-229- 

RUTHERFORD,  W.H.  1954.  A  record  of  the  ringtail  (Bas- 
saricas  astutns)  in  Colorado.  J.  Mammal.  35:442-443 

SARGENT,  A.B.,  W.K  PFEIFER,  and  S.H.  ALLEN.  1975.  A 
spring  aerial  census  of  red  foxes  in  North  Dakota. 
J.  Wildl.  Manage.  39:30-39- 

and  D.W.  WARNER.  1972.  Movements  and  denning 

habits  of  a  badger.  J.  Mammal.  53:207-210. 

SAUNDERS,  J.K,  Jr.  1963-  Food  habits  of  the  lynx  in  New- 
foundland. J.  Wildl.  Manage.  27:384-390. 

SCHEMPF,  P.F.  and  M.  WHITE.  1977.  Status  of  six  furbear- 
ers  in  the  mountains  of  northern  California.  U.S.  Dep. 
Agric,  For.  Serv.  Calif.  Reg.  51pp. 

SCHOFIELD,  R.D.  I960.  A  thousand  miles  of  fox  trails  in 
Michigan's  ruffed  grouse  range.  J.  Wildl.  Manage. 
24:432-434. 

SCHOWALTER,  D.B.  and  JR.  GUNSON.  1982.  Parameters 
of  populations  and  seasonal  activity  of  striped  skunks. 
Mephitis  mephitis,  in  Alberta  and  Saskatchewan.  Can. 
Field-Nat.  96:409-420. 

SCOTT,  T.G.  1955.  Dietary  patterns  of  red  and  gray  foxes. 
Ecology  36:366-367. 

and  L.F.  SELKO.  1939.  A  census  of  red  foxes  and 

striped  skunks  in  Clay  and  Boone  counties,  Iowa. 
J.  Wildl.  Manage.  3:92-98. 

SEBER,  G.A.F.  1973-  The  estimation  of  animal  abundance. 
Griffin.  London.  506pp. 

SEIDENSTICKER,  J.C,  M.G.  HORNOCKER,  W.V.  WILES, 
and  J.P.  MESSICK  1973  Mountain  lion  social  organi- 
zation in  the  Idaho  Primitive  Area.  Wildl.  Monogr.  35. 
60pp. 

SEN,  A.R.  1982.  A  review  of  some  important  techniques  in 
sampling  wildlife.  Can.  Wildl.  Serv.,  Occ.  Paper  49- 
17pp. 

SHAW,  H.G.  1979.  A  mountain  lion  field  guide.  Arizona 
Game  and  Fish  Dep.  Spec.  Rep.  9,  Phoenix.  27pp. 

SHORT,  ILL.  1979.  Food  habits  of  coyotes  in  a  semidesert 
grass-shrub  habitat.  U.S.  Dep.  Agric,  For.  Serv.  Res. 
Note  RM-364.  4pp. 

SIMMS,  DA.  1979.  North  American  weasels  resource  utili- 
zation and  distribution.  Can.  J.  Zool.  57:504-520. 

SINGER,  F.J.  1978.  Seasonal  concentrations  of  grizzly 
bears.  North  Fork  of  the  Flathead  River,  Montana. 
Can.  Field-Nat.  92:283-286. 

SMITH,  J.W.  and  B.J.  VERTS.  1982.  Mephitis  mephitis. 

Mammalian  Species  173-  The  Amer.  Soc.  of  Mammal. 
Lawrence,  KS.  7pp. 


Carnivores 


495 


SOUTIERE,  EC.  1979.  Effects  of  timber  harvesting  on 
marten  in  Maine.  J.  Wildl.  Manage.  43:850-860. 

SPALDING,  DJ.  and  J.  LESOWSKJ.  1971.  Winter  food  of 
the  cougar  in  south-central  British  Columbia.  J.  Wildl. 
Manage.  35:378-381. 

SPENCER,  W.D.,  R.H.  BARRETT,  and  W.J.  ZIELINSKI.  1983. 
Marten  habitat  preferences  in  the  northern  Sierra 
Nevada.  J.  Wildl.  Manage.  47:1181-1186. 

SPERRY,  C.C.  1941.  Food  habits  of  the  coyote.  U.S.  Dep. 
Inter.,  Fish  and  Wildl.  Serv.  Res.  Bull.  4.  70pp. 

SPRINGER,  J. T.  and  J.S.  SMITH.  1981.  Summer  food  habits 
of  coyotes  in  central  Wyoming.  Great  Basin  Nat. 
41:449-456. 

STAINS,  HJ.  1975.  Distribution  and  taxonomy  of  the  cani- 
dae.  Pages  3-26  in  Vox,  M.W.  ed.  The  Wild  Canids. 
Van  Nostrand  Reinhold  Co.  New  York,  NY. 

STORM,  G.L.  1972.  Daytime  retreats  and  movements  of 
skunks  on  farmlands  of  Illinois.  J.  Wildl.  Manage. 
36:31-45. 

STREETER,  R.G.  and  C.E.  BRAUN.  1968.  Occurrence  of 
pine  marten  (Martes  americana)  in  Colorado  alpine 
areas.  Southwest.  Nat.  13:449-451. 

STRICKLAND,  M.A.,  C.W.  DOUGLAS,  M.  NOVAK,  and  N.P. 
HUNZIGER.  1982.  Martin.  Pages  599-612  in  Chap- 
man, J.  and  G.  Feldhamer  eds.  Wild  Mammals  of 
North  America:  Their  Biology,  Management  and  Eco- 
nomics. The  John  Hopkins  Univ.  Press.  Baltimore,  MD. 

TODD,  AW.  and  LB.  KEITH.  1983.  Coyote  demography 
during  a  snowshoe  hare  decline  in  Alberta.  J.  Wildl. 
Manage.  47:394-404. 

TOWEILL,  D.E.  and  EC.  MESLOW.  1977.  Food  habits  of 
cougars  in  Oregon.  J.  Wildl.  Manage.  41:576-578. 

TRAPP,  G.  and  D.L  HALLBERG.  1975.  Ecology  of  the  gray 
fox  (Urocyoti  cinereoargenteus):  a  review.  Pages 
164-178  in  Fox,  M.W.  ed.  The  Wild  Canids:  Their  Sys- 
tematics,  Behavioral  Ecology  and  Evolution.  Van  Nos- 
trand Reinhold  Co.  New  York,  NY. 

TREADWELL,  B.D.  1979.  A  provisional  framework  for 
denning  black  bear  habitat.  Pages  319-330  in 
LeCount,  A.  ed.  First  Western  Black  Bear  Workshop. 
Ariz.  Game  and  Fish  Dep.,  Phoenix. 

TURKOWSKI,  F.J.  1980.  Carnivora  food  habits  and  habitat 
use  in  ponderosa  pine  forests.  U.S.  Dep.  Agric,  For. 
Serv.  Res.  Pap.  RM-215.  9pp. 

ULLIMAN,  J.J.,  E.O  GARTON,  andJ.A.  KEAY.  1979.  Wild- 
life habitat  classification  using  large-scale  aerial  pho- 
tography. Pages  414-422  in  Frayer,  W.F.  ed.  Forest 
Resource  Inventories,  Vol.  1.  Colo.  State  Univ.,  Ft. 
Collins. 

URBAN,  D.  1970.  Raccoon  populations,  movement  pat- 
terns, and  predation  of  a  managed  waterfowl  marsh.  J. 
Wildl.  Manage.  34:372-382. 


VAN  BALLENBERGHE,  V.,  AW.  ER1CKSON,  and  D.  BY- 
MAN.  1975.  Ecology  of  the  timber  wolf  in  northeast- 
ern Minnesota.  Wildl.  Monogr.  43-  43pp. 

VAN  ZYLL  DE  JONG,  C.G.  1966.  Food  habits  of  the  lynx 
in  Alberta  and  the  Mackenzie  District,  N.W.T.  Can. 
Field-Nat.  80:18-23. 

.  1975.  The  distribution  and  abundance  of  the  wol- 
verine (Gulo  gulo)  in  Canada.  Can.  Field-Nat.  89:431- 
437. 

VARNEY,  JR.  1973.  An  evaluation  of  the  use  of  ERTS-1 
satellite  imagery  for  grizzly  bear  habitat  analysis. 
NASA  Prog.  Rep.,  Mont.  Coop.  Wildl.  Res.  Unit.  Univ. 
Montana,  Missoula. 

VERTS,  B.J.  1967.  The  biology  of  the  striped  skunk.  Univ. 
Illinois  Press,  Urbana.  218pp. 

WAGNER,  FH.  and  LC.  STODDART.  1972.  Influence  of 
coyote  predation  on  black-tailed  jackrabbit  popula- 
tions in  Utah.  J.  Wildl.  Manage.  36:329-342. 

WECKWERTH,  R.P.  and  V.D.  HAWLEY.  1962.  Marten  food 
habits  and  population  fluctuations  in  Montana.  J. 
Wildl.  Manage.  27(1):5574. 

WEISE,  T.F.,  W.L  ROBINSON,  R.A.  HOOK  and  L.D.  MECH. 
1975.  An  experimental  translocation  of  the  eastern 
timber  wolf.  Pages  19-42  in  Klinghammer,  E.  ed.  The 
Behavior  and  Ecology  of  Wolves.  Symp.  Proc.  Garland 
STPM  Press.  New  York,  NY. 

WENGER,  C.R.  and  AT  CRINGAN.  1978.  Siren-elicited 
coyote  vocalizations:  an  evaluation  of  a  census  tech- 
nique. Wildl.  Soc.  Bull.  6:73-76. 

WOOD,  J.E.  1954.  Food  habits  of  furbearers  of  the  upland 
post  oak  region  in  Texas.  J.  Mammal.  35:406-415. 

.  1959.  Relative  estimates  of  fox  population  levels.  J. 

Wildl.  Manage.  23:53-63. 

YOCOM,  C.F.  1973.  Wolverine  records  in  the  Pacific 

coastal  states  and  new  records  for  northern  California. 
Calif.  Fish  and  Game  59:207-209. 

.  1974.  Status  of  marten  in  northern  California, 

Oregon,  and  Washington.  Calif.  Fish  and  Game  60:54- 
57. 

and  M.T.  McCOLLUM.  1973.  Status  of  the  fisher  in 


northern  California,  Oregon,  and  Washington.  Calif. 
Fish  and  Game  59:305-309. 

YOUNG,  S.P.  1958.  The  bobcat  of  North  America.  The 
Stackpole  Co.  Harrisburg,  PA.  and  The  Wildlife  Man- 
agement Inst.  Washington,  DC. 

ZIMEN,  E.  and  L.  BOITANI.  1975.  Status  of  the  wolf  in 

Europe  and  the  possibilities  of  conservation  and  rein- 
troduction.  Pages  43-83  in  Klinghammer,  E.  ed.  The 
Behavior  and  Ecology  of  Wolves.  Symp.  Proc.  Garland 
STPM  Press.  New  York,  NY. 


496 


Carnivores 


24 

BATS 


Stephen  P.  Cross 

Department  of  Biology 
Southern  Oregon  State  College 
Ashland,  OR  97520 


Editor's  Note:  Most  biologists  have  little  experience 
with  bats,  and  therefore  tend  to  exclude  them  from 
inventories  or  monitoring  studies.  Bats,  however, 
are  an  important  resource  and  are  often  locally 
abundant. 

The  study  of  bats  is  a  specialized  discipline  that  re- 
quires the  use  of  distinctive  techniques.  This  chapter 
provides  a  summary  of  bat  study  methods  and  sug- 
gests appropriate  circumstances  for  their  use. 


INTRODUCTION 

Bats  are  often  a  neglected  segment  of  wildlife 
inventories;  they  are  difficult  to  sample,  they  are  mo- 
bile and  have  secretive  nocturnal  habits,  and  they 
are  considered  by  many  people  to  be  a  mysterious 
and  unlikable  group  of  wildlife.  However,  bats  play 
important  roles  in  many  ecosystems  and  they  cer- 
tainly deserve  attention  in  wildlife  habitat  and  popu- 
lation inventories. 

Roughly  40  species  of  bats  exist  in  North  Amer- 
ica north  of  Mexico  (Jones  et  al.  1982;  Hall  1981). 
In  some  habitats  the  number  of  bat  species  exceeds 
that  of  all  other  mammal  species  combined,  and 
some  species  can  occur  in  very  large  numbers  if 
conditions  are  favorable.  Bats  have  many  distinctive 
morphological,  physiological,  and  behavioral  adapta- 
tions that  are  associated  with  some  unique  uses  of 
habitat.  Consequently,  methods  of  inventory  and 
monitoring  of  habitat  and  populations  also  tend  to 
be  unique.  Only  methods  most  suitable  for  applica- 
tion to  the  bats  of  temperate  North  America  will  be 
considered  here. 

A  knowledge  of  basic  behavioral  characteristics, 
especially  daily  and  seasonal  activity  patterns,  is  es- 
sential when  contemplating  any  study  of  bats.  Be- 
cause bats  are  extremely  mobile,  some  habitat  inven- 
tory and  monitoring  techniques  are  similar  to  those 
for  birds.  Unlike  most  birds,  however,  bats  are  active 
at  night  and  are  not  easily  detected  or  distinguished 
by  direct  observation  or  unaided  audition.  Further, 
the  local  distribution  of  bats  is  usually  patchy  and 
uneven.  This  is  caused  by  variation  in  the  distribu- 
tion of  essential  habitat  resources  and  seasonally 
variable  tendencies  of  bats  to  congregate  at  specific 
and  often  unique  resting,  breeding,  feeding,  and 
watering  places.  Consequently,  although  population 
measurements  and  determination  of  species  presence 
may  be  accomplished  by  censusing  at  such  congrega- 
tion sites,  determination  of  absolute  density  is  vir- 
tually impossible. 

Daily  and  seasonal  activity  patterns  are  impor- 
tant aspects  of  behavior  that  are  directly  related  to 
inventory  of  populations  and  analysis  of  habitat  asso- 
ciations. During  the  summer  a  typical  daily  activity 
pattern  for  North  American  bats  begins  with  emer- 


Bats 


497 


Pallid  bat. 


gence  from  a  daytime  shelter  (roost)  during  evening 
twilight  or  after  dark.  A  roost  is  any  site  where  a  bat 
lands  when  it  voluntarily  ceases  flying.  By  conven- 
tion, roosts  usually  include  resting  sites  where  food 
may  be  devoured  and  digested  but  exclude  sites 
where  a  bat  lands  to  capture  prey.  Some  species  or- 
dinarily congregate  in  large  numbers  whereas  others 
roost  singly. 

Bats  may  begin  feeding  immediately  after  leav- 
ing the  day  roost  or  they  may  fly  several  miles  be- 
fore reaching  a  foraging  site.  Most  North  American 
species  feed  on  flying  insects.  After  feeding,  they  may 
fly  to  a  drinking  site  or,  alternatively,  a  night  roost.  A 
night  roost  is  generally  not  at  the  same  site  as  a  day 
roost.  Night  roosts  are  occupied  for  various  lengths 
of  time  depending  upon  the  species  and  purpose  for 
which  the  roost  is  being  used.  Some  night  roosts 
may  be  used  simultaneously  by  several  species  that 
do  not  normally  share  a  day  roost. 

The  active-inactive  pattern  may  be  repeated  one 
or  two  times  during  the  night  after  which  the  bats 
return  to  their  original  day  roost.  This  daily  pattern 


varies  depending  on  the  species,  geographic  location, 
season,  and  environmental  conditions.  Water  require- 
ments and  drinking  activity  vary  with  the  tempera- 
ture and  humidity.  During  late  spring  or  early  sum- 
mer the  females  of  many  species  congregate  to  form 
colonies  at  maternity  roost  sites  where  parturition 
and  maturation  of  the  young  take  place.  During  this 
period,  water  requirements  may  be  elevated  due  to 
lactation  and  high  environmental  temperatures,  and 
the  bats  may  begin  their  nocturnal  activity  with  a 
trip  to  a  drinking  site.  While  females  congregate  in 
maternity  colonies,  males  of  some  species  may  form 
separate  bachelor  colonies. 

Some  night  roosts  are  used  primarily  in  the  late 
summer  and  fall,  presumably  for  breeding.  Congrega- 
tion at  such  sites  is  often  referred  to  as  swarming. 
For  some  species  that  are  usually  found  solitary  or 
sexually  segregated,  late  summer  and  fall  is  a  time 
when  individuals  come  together,  thus  facilitating 
mating  and  reducing  the  probability  of  inbreeding. 

Roosts  used  during  the  winter  for  dormancy  are 
called  hibernacula.  Such  roosts  may  be  at  different 


498 


Bats 


elevations  or  in  different  geographic  regions  than 
those  of  summer  roosts.  Often  these  are  the  same 
sites  visited  during  swarming  earlier  in  the  year. 
Those  species  that  do  not  hibernate  may  undertake 
extensive  latitudinal  migration  to  find  suitable  condi- 
tions during  different  periods  of  the  year.  It  becomes 
evident  that  several  different  types  of  roosts,  each 
with  distinct  characteristics,  are  required  throughout 
the  year  for  many  species. 

Although  techniques  are  not  standardized,  there 
is  a  logical  sequence  of  procedures  that  may  be  fol- 
lowed when  investigating  bat  populations  and  their 
habitat  relationships  at  a  local  level.  The  first  step  is 
to  compile  a  list  of  species  potentially  in  the  area  of 
concern  by  reviewing  publications  on  mammal  dis- 
tribution in  North  America  (e.g.,  Hall  1981),  one 
that  deals  specifically  with  the  distribution  of  bats 
(e.g.,  Barbour  and  Davis  1969),  or  one  that  deals 
with  the  distribution  of  mammals  or  bats  in  a  more 
specific  region  (e.g.,  Maser  et  al.  1981).  The  known 
habitat  affinities  may  also  be  determined  by  consult- 
ing general  publications  on  bats,  especially  Barbour 
and  Davis  (1969),  or  on  regional  natural  history.  For 
many  regions  in  North  America,  one  may  consult  the 
very  specific  descriptions  of  wildlife-habitat  associa- 
tions and  distribution  that  have  been  compiled  in  re- 
cent years  (e.g.,  Thomas  1979). 

The  next  phase  of  study  depends  on  the  ques- 
tions) being  asked.  One  question  may  be  simply 
whether  a  given  species  or  group  of  different  species 
is  present  in  an  area  or  habitat.  Techniques  for  veri- 
fying the  presence  of  predicted  species  should  be  se- 
lected with  the  habits  of  the  species  in  mind.  For 
some  species  an  investigation  of  potential  roost  sites 
is  appropriate.  For  species  whose  roost  sites  are  inac- 
cessible or  little  known,  a  technique  such  as  mist- 
netting  at  water  holes  or  foraging  areas  may  be  best. 
Even  more  specialized  techniques,  such  as  the  use  of 
ultrasonic  detectors,  may  be  appropriate  for  some 
species.  These  techniques  may  complement  one  an- 
other depending  on  the  objectives  of  the  study  (e.g., 
see  Paige  et  al.  1985).  The  important  point  is  that 
many  species  or  groups  of  species  have  distinctive 
behaviors  that  must  be  taken  into  consideration 
when  attempting  to  verify  presence. 

Questions  related  to  the  use  of  specific  habitat 
features  are  sometimes  difficult  to  answer.  Determin- 
ing whether  bats  use  a  given  cave  as  a  roost  site  is 
relatively  simple,  but  the  investigator  must  be  aware 
that  just  because  bats  or  their  sign  are  not  found  at 
one  time  of  a  day  or  year  does  not  necessarily  mean 
that  they  do  not  use  the  site  at  another  time.  Be- 
cause some  roost  areas,  such  as  cliffs  and  trees,  are 
often  inaccessible,  it  is  very  difficult  to  establish  hab- 
itat relationships. 

Questions  regarding  population  size  also  require 
consideration  of  both  general  and  specific  behaviors. 


Although  measurement  of  absolute  density  is  not  a 
feasible  objective,  an  estimation  of  population  size 
utilizing  a  specific  area  or  habitat  feature  is  possible 
for  those  species  whose  members  congregate  at  spe- 
cific locations.  For  many  species  these  congregation 
sites  are  the  summer  day  roosts,  especially  maternity 
colonies,  or  winter  hibernacula.  Even  though  there 
may  be  a  great  distance  between  such  sites,  most 
bats  seem  to  exhibit  loyalty  to  them — at  least  on  a 
seasonal  basis.  This  fidelity  makes  it  possible  to  com- 
pare population  size  at  a  given  site  during  the  same 
time  from  year  to  year.  Other  species  may  use  a  par- 
ticular site  year  around.  In  any  event,  knowledge  of 
such  species-specific  behavior  is  imperative  when 
framing  questions  about  the  importance  of  a  habitat 
feature  or  change  that  might  influence  population 
size. 

For  species  that  do  not  congregate  at  accessible 
daytime  roosts  or  hibernacula,  it  may  be  possible  to 
obtain  rough  estimates  of  population  size  by  captur- 
ing them  in  other  areas  of  concentration.  Although 
drinking  sites  are  often  difficult  places  to  estimate 
species  population  sizes,  they  may  be  the  best  places 
to  assess  bat  community  structure.  Whereas  many 
species  segregate  for  day  roosting  and  dormancy, 
they  may  come  together  at  common  drinking  sites 
during  the  summer.  To  a  lesser  extent  this  may  also 
be  true  for  night  roosts  during  the  fall.  However, 
both  drinking  places  and  night  roosts  may  have  char- 
acteristics that  are  differentially  attractive  to  various 
species  so  one  must  not  assume  that  samples  taken 
at  those  sites  are  necessarily  representative  of  local 
community  structure. 

When  population  estimates  are  compared  spa- 
tially or  temporally,  the  techniques  need  to  be  iden- 
tical. In  contrast,  to  determine  bat  community  com- 
position in  a  particular  area,  as  many  sampling 
techniques  as  possible  should  be  employed.  Relative 
density  of  the  entire  community  may  not  be  easily 
measured,  but  it  is  possible  to  measure  the  portion 
of  the  community  using  a  common  local  resource, 
and  the  data  obtained  need  only  be  qualified  as  such 
to  be  useful  for  descriptive  or  comparative  purposes. 


HABITAT  FEATURES  CORRELATED 
WITH  BATS 

Many  bats  are  closely  tied  to  specific  habitat  fea- 
tures. The  primary  activities  associated  with  such 
features  are  drinking,  feeding,  and  roosting.  Drinking 
is  obviously  associated  with  physical  features,  partic- 
ularly the  presence  of  a  suitable  source  of  water, 
whereas  feeding  often  shows  some  association  with 
vegetative  features.  Because  roosting  behavior  may 
be  associated  with  both  physical  and  vegetative  fea- 
tures, a  brief  consideration  of  the  general  physical 
attributes  of  the  various  functional  categories  or 
roosts  follows. 


Bats 


499 


Day  roosts  are  generally  inaccessible  to  preda- 
tors and  have  relatively  stable  microclimatic  condi- 
tions thus  providing  safe  and  physiologically  efficient 
places  to  spend  most  of  the  inactive  time.  Much  less 
is  known  about  night  roosts  but  they  seem  essential 
for  bat  survival.  They  are  often  located  much  closer 
to  foraging  and  drinking  sites  than  day  roosts  and 
provide  a  relatively  safe  place  to  rest,  eat  and  digest 
food,  and  groom.  The  physical  structure  of  night 
roosts  is  often  distinctly  different  from  that  of  day 
roosts.  For  example,  many  bats  that  roost  in  crevices 
during  the  day  select  caves  or  other  relatively  open 
sites  as  night  roosts. 

The  largest  congregations  of  bats  occur  in  ma- 
ternity roosts.  Even  some  bats  that  are  solitary  dur- 
ing most  of  the  year  may  congregate  in  small  colo- 
nies for  birthing  purposes.  This  indicates  either  that 
sites  with  the  proper  conditions  for  birthing  are 
scarce  or  that  the  presence  of  large  numbers  of  bats 
is  necessary  to  create  the  proper  roost  conditions. 
Maternity  roosts  are  warmer  than  other  roosts — a 
condition  which  increases  the  growth  rate  of  young 
during  gestation  and  afterbirth  (McNab  1982). 


Maternity  colony  of  the  Yuma  Myotis  in  a  man-made  struc- 
ture. 


Conditions  in  bat  hibernacula  that  are  crucial 
for  survival  include  relatively  stable  temperatures 
that  are  low  enough  to  foster  the  reduced  metabolic 
activity  necessary  to  conserve  energy,  but  not  so  low 
as  to  be  lethal.  For  a  comprehensive  treatment  of 
current  knowledge  of  bat  roosting  ecology,  see  Kunz 
(1982). 

Physical  Features 

Compared  with  other  wildlife,  bats  exhibit  a  rel- 
atively strong  association  with  prominent  physical 
features  of  their  habitat.  Most  of  these  physical  fea- 
tures fall  into  the  category  that  Maser  et  al.  ( 1979) 
refer  to  as  "unique"  habitats.  These  sites  make  up  a 


very  small  percentage  of  the  total  land  base  but  are 
often  disproportionately  important  as  habitat.  Distri- 
bution and  abundance  of  bats  are  often  directly  re- 
lated to  the  availability  of  these  habitat  features, 
which  may  be  natural  or  man-made  structures. 

There  are  two  principal  activities  associated  di- 
rectly with  the  use  of  physical  habitat  features.  The 
first,  roosting,  occurs  in  a  variety  of  natural  and  man- 
made  structures.  The  second,  drinking,  usually 
requires  an  open  water  source.  Specific  structural 
features  are  described  under  these  two  functional 
headings. 

Roosting  sites.  Physical  features  that  serve  as  roost- 
ing sites  may  be  categorized  in  a  number  of  ways. 
The  following  list  is  not  intended  to  be  comprehen- 
sive but  does  include  the  structures  used  most  often 
for  roosts.  Bats  appear  to  choose  these  structures  on 
the  basis  of  their  specific  and  relatively  constant  mi- 
croclimatic conditions  and  lack  of  disturbance. 

Caves  and  Cave-like  Structures.  Caves  are  uti- 
lized by  bats  more  than  by  any  other  group  of  wild- 
life. Caves  provide  shelter  from  adverse  weather,  a 
relatively  stable  environment,  darkness,  and  protec- 
tion from  predators.  Bats  may  roost  in  caves  singly 
or  in  large  congregations.  As  many  as  25  million  Bra- 
zilian free-tailed  bats  (Tadarida  brasiliensis)  have 
been  found  in  single  caves  in  the  southwestern 
United  States  in  the  summer  (Cockrum  1969;  Davis 
et  al.  1962).  Some  species  use  caves  or  cave-like 
structures  for  all  roosting  activities  whereas  others 
use  them  only  for  one  type  of  roosting  behavior.  For 
example,  Townsend's  big-eared  bat  {Plecotus  town- 
sendii)  and  the  gray  myotis  {Myotis  grisescens) 
seem  restricted  to  cave-like  roosts,  whereas  the  long- 
eared  myotis  (M.  evotis)  and  long-legged  myotis  (M 
volans)  ordinarily  use  caves  only  as  night  roosts 
(Barbour  and  Davis  1969). 

Many  caves  are  unsuitable  as  bat  roosts  and 
some  are  used  regularly  for  only  a  portion  of  the 
year.  The  important  characteristics  of  a  cave  are  tem- 
perature, humidity,  light,  size,  and  interior  configura- 
tion and  surface  texture.  Caves  that  offer  the  most 
optimal  range  of  all  of  these  features  are  those  most 
likely  to  be  used  by  the  greatest  number  of  individu- 
als and  species.  Caves  with  large  spacious  chambers 
are  more  likely  to  house  large  aggregations.  Cold  and 
damp  caves  are  more  suitable  as  hibernacula  or  as 
summer  night  roosts.  Warm  caves  are  more  suitable 
for  maternity  roosts  or  for  species  that  are  active 
year-round.  Total  darkness  does  not  seem  to  be  an 
absolute  requirement  for  most  cave  dwelling  bats,  al- 
though they  may  select  the  darkest  portion  of  a  par- 
ticular cave.  Caves  used  as  day  roosts  are  often  more 
likely  to  be  occupied  simultaneously  by  fewer  spe- 
cies than  those  used  as  night  roosts. 


500 


Bats 


A  group  of  Townsend's  big-eared  bats  roosting  on  the  ceiling  of  a  lava  tube  at  Lava  Beds  National  Monument. 


Many  types  of  man-made  cave-like  structures 
have  access  to  the  surface.  Tunnels  and  mines  are 
the  most  obvious.  Although  not  usually  as  structur- 
ally diverse  as  natural  caves,  they  may  be  used  by 
bats  for  all  types  of  roosting.  Generally  they  are  eas- 
ier for  humans  to  find  and  access  than  natural  caves 
because  of  habitat  modifications  related  to  excava- 
tion. Not  only  are  the  locations  more  visible  but 
they  are  often  recorded  on  legal  or  topographic 
maps.  Old  mine  shafts  and  tunnels  are  often  very 
dangerous — great  care  should  be  exercised  when  in- 
vestigating such  abandoned  excavations. 


Cliffs,  Crevices,  and  Talus  Slopes.  Cliffs  pro- 
vide some  roosting  structures — especially  crevices — 
favored  by  many  bat  species.  Rock  crevices  in  cliffs 
may  have  more  suitable  temperature  and  moisture 
conditions  for  bats  than  rock  crevices  elsewhere  be- 
cause of  their  vertical  exposure  and  large  mass  asso- 
ciated with  large  heat-retaining  capacity.  Further, 
cliffs  provide  a  safer  refuge  from  predators  than 
roost  structures  more  accessible  to  flat  terrain. 


The  largest  and  smallest  species  of  bats  in  the 
U.S.  are  cliff-crevice  dwellers.  The  greater  mastiff  bat 
(Eumops perotis),  roosting  in  small  groups,  occupies 
large  crevices  with  downward  facing  openings  at 
least  twice  as  wide  as  its  body  (Vaughan  1959).  The 
western  pipestrelle  (Pipistrellus  hesperus)  usually 
roosts  singly  in  small,  up  to  2.5  cm  (1  in.)  wide,  ver- 
tical rock  crevices  (Cross  1965).  A  few  other  species 
have  been  found  in  intermediate  situations  but,  as 
Kunz  ( 1982)  pointed  out,  little  is  known  about  the 
ecology  of  crevice-dwelling  bats  because  of  the  diffi- 
culty of  finding  them  in  these  relatively  inaccessible 
places. 

In  general,  those  cliffs  with  greatest  fracturing 
are  more  likely  to  provide  a  diversity  of  rock  crev- 
ices from  which  bats  may  choose  roosting  sites.  Ori- 
entation of  cliff  faces  with  respect  to  the  sun  may  be 
important  for  determining  the  temperature  condi- 
tions of  the  associated  crevices  or  other  cavities.  Oc- 
cupied cliff  crevice  roosts  offer  some  thermal  buffer- 
ing and  a  variety  of  thermal  choices  (O'Shea  and 
Vaughan  1977;  Hayward  and  Cross  1979). 


Bats 


501 


Talus  is  the  accumulation  of  broken  rocks  at  the 
base  of  cliffs  or  other  steep  slopes.  Talus  creates  a 
variety  of  crevices,  cavities,  and  even  small  caves,  de- 
pending on  the  size  of  the  breakdown  debris.  Bats 
roosting  in  such  habitat  are  more  susceptible  to  nat- 
ural predation  than  in  caves  or  cliffs. 

Man-Made  Structures.  Many  species  of  North 
American  bats  readily  use  certain  man-made  struc- 
tures as  roost  sites.  These  structures  appear  to  sup- 
ply suitable  substitutes  for  "natural"  roost  sites.  Some 
species,  such  as  the  big  brown  bat  (Eptesicus  fus- 
cus),  the  little  brown  bat  (Myotis  lucifugus),  and 
the  Yuma  myotis  (M.  yumanensis),  show  a  great  de- 
pendency upon  these  roosts,  especially  as  maternity 
sites  (Barbour  and  Davis  1969).  There  is  evidence  to 
indicate  that  both  ranges  and  population  sizes  of 
some  species  have  increased  as  a  result  of  using 
man-made  structures  for  roosts  (see  Kunz  1982).  A 
great  variety  of  man-made  structures  are  used  by 
bats.  In  addition  to  the  cave-like  structures  men- 
tioned above,  the  other  structures  may  be  grouped 
into  three  categories:  buildings,  bridges,  and  miscel- 
laneous structures. 

Buildings  offer  an  endless  variety  of  roosting 
sites  that  environmentally  resemble  natural  crevices, 
caves,  or  cavities.  For  example,  spaces  under  roofing 
or  siding  provide  crevice-like  structures.  Attics,  base- 
ments, and  abandoned  buildings  provide  cave-like 
conditions.  Intermediate-sized  spaces  may  resemble 
natural  rock  or  tree  cavities.  Many  semi-open  struc- 
tures such  as  carports,  porches,  or  woodsheds  serve 
as  temporary  night  roosts.  Some  structures  have 
been  built  specifically  to  attract  roosting  bats  (see 
Greenhall  and  Paradiso  1968;  Greenhall  1982). 

Bridges  also  provide  a  variety  of  roosting  sites, 
sometimes  in  remote  areas  otherwise  having  rela- 
tively featureless  terrain  (Davis  and  Cockrum  1963). 
Some  concrete  bridges  have  cave-like  chambers  in 
the  ends.  Several  types  of  crevices  are  available,  es- 
pecially in  older  style  bridges.  Modern  bridges  tend 
to  have  fewer  protected  roosting  sites  but  some  are 
still  useful  as  night  roosts  and,  in  some  instances, 
could  be  modified  to  serve  as  day  roosts. 

Other  potentially  useful  man-made  structures  in- 
clude dams,  windmill  support  structures,  storm  sew- 
ers, and  any  structures  that  provide  cavities,  crev- 
ices, or  cave-like  areas  with  suitable  temperature, 
humidity,  light,  and  protection. 


Drinking  Sites.  The  availability  of  open  water  is 
crucial  to  the  survival  of  most  bats,  and  therefore 
greatly  influences  their  distribution  and  abundance. 
Most  bats  drink  while  flying,  which  means  that  the 
surface  area  of  the  open  water  must  be  sufficiently 
large  to  accommodate  this  behavior.  Rapid  flyers, 


Highway  bridge. 


Crevices  formed  between  wooden  planks  of  bridge. 


Bats  roosting  in  one  of  the  crevices. 


502 


Bats 


I 


such  as  members  of  the  family  Molossidae  (free- 
tailed  bats),  need  a  very  large  surface  area  from 
which  to  drink  whereas  some  of  the  highly  agile 
flyers,  such  as  several  small  Myotis,  can  drink  from  a 
pool  only  a  few  centimeters  in  diameter  while  still  in 
flight.  Because  free  water  is  an  essential  requirement 
for  most  species,  seasonal  fluctuations  in  the  amount 
and  distribution  of  open  water  greatly  influence  the 
local  distribution  of  bats,  especially  in  arid  regions 
where  water  is  scarce. 

Man  has  inadvertently  made  many  otherwise  un- 
inhabitable regions  useful  to  bats  by  providing  water 
in  the  form  of  seasonal  storage  reservoirs,  ponds  and 
watering  devices  for  livestock,  heliponds,  some  types 
of  wildlife  water  catchment  devices  (guzzlers),  and 
swimming  pools.  Any  habitat  modification  that  re- 
sults in  the  availability  of  open  water  during  dry  pe- 
riods may  result  in  greater  use  of  the  areas  by  bats  if 
other  requirements  are  available. 


teris  mexicana)  feeds  on  nectar.  The  Mexican  long- 
nosed  bat  (Leptonyc teris  nivalis)  and  the  other 
long-nosed  bat  (L.  yerbabuenae  or  sanborni)  feed 
on  nectar  but  also  utilize  pollen  and  fruit. 

Insectivorous  species  may  be  more  numerous 
near  areas  such  as  marshes  and  agricultural  lands 
where  the  vegetation  supports  high  insect  popula- 
tions. Often  in  such  places,  chemical  insecticides  are 
used  to  control  the  insects  and  there  is  evidence 
that  these  accumulate  secondarily  in  the  tissues  of 
some  bats  and  may  be  harmful  to  their  populations 
(Geluso  et  al.  1976;  Henny  et  al.  1982).  Some  forest 
dwelling  bats  respond  to  tree  density  on  the  basis  of 
foraging  style.  Those  that  are  strong  flyers  and  catch 
their  insect  prey  in  flight  need  open  forests  whereas 
those  that  are  very  agile  or  glean  their  food  from  the 
vegetation  can  utilize  more  densely  vegetated  areas. 

Discussion 


Vegetation  Features 

Little  information  is  available  regarding  bat-vege- 
tation associations.  It  appears  that  most  North  Ameri- 
can bats  respond  to  major  structural  differences  in 
vegetation  rather  than  to  vegetation  species  differ- 
ences. Roosting  and  feeding  are  two  prominent  activ- 
ities that  are  often  associated  with  vegetative  fea- 
tures. Because  bats  are  very  mobile,  these  two 
activities  may  occur  in  widely  separated  and  vegeta- 
tively  distinct  areas. 

Roosting  Sites.  Some  North  American  bats  roost  on 
vegetation  but  relatively  little  is  known  about  spe- 
cific requirements.  Most  known  vegetative  roosts  are 
daytime  retreats,  but  some  evidence  exists  that  bats 
may  also  use  trees  and  shrubs  as  night  roosts  (Kunz 
1982).  Vegetative  day  roosts  are  most  prominently 
associated  with  living  trees  or  snags  and  logs. 

A  few  North  American  species  regularly  roost  in 
the  foliage  of  trees.  These  include  the  hoary  bat 
(Lasiurus  cinereus),  red  bat  (L  borealis),  northern 
yellow  bat  (L.  intermedins),  seminole  bat  (L.  semi- 
nolus),  and  probably  the  silver-haired  bat  {Lasionyc- 
teris  noctivagans).  Some  species,  such  as  the  big 
brown  bat  and  the  pallid  bat  {Antrozous  pallidus), 
may  use  tree  cavities  or  crevices.  Species  such  as 
Keen's  myotis  (M.  keenii)  and  long-eared  myotis 
have  been  found  in  tree  crevices  or  under  exfoliat- 
ing bark,  primarily  in  standing  trees  (Barbour  and 
Davis  1969).  Many  other  species  are  also  thought  to 
use  these  sites  because  of  their  close  association 
with  forested  habitats. 


Because  bats  are  closely  associated  with  distinc- 
tive habitat  features,  their  occupancy  of  a  particular 
area  may  be  predicted  to  some  extent  on  the  basis 


Feeding  Relationships.  Three  species  with  limited 
distribution  in  the  southwestern  U.S.  use  plants  di- 
rectly for  food.  The  long-tongued  bat  (Cboeronyc- 


Conifer  snag  with  hole  and  exfoliating  bark — creating 
roosting  sites  for  bats. 


Bats 


503 


of  the  presence  or  absence  of  these  features.  The 
best  places  to  look  for  bats  or  signs  of  their  presence 
are  in  and  around  essential  habitat  features.  Any  geo- 
graphic area  of  interest  should  first  be  surveyed  for 
potential  roost  and  drinking  sites.  In  addition  to  the 
presence  or  absence  of  essential  habitat  features, 
their  proximity  to  other  features  is  also  important. 
For  example,  the  absence  of  open  water  within  a  few 
miles  of  potential  roosting  or  feeding  sites  will  limit 
some  species.  The  available  surface  of  open  water 
will  limit  others.  The  types  of  roost  sites  available 
will  also  limit  some  species.  Cave  and  crevice  dwell- 
ers are  often  restricted  to  just  those  specific  roost 
structures. 

Apparent  or  potential  roost  structures  should  be 
investigated  first.  If  bats  are  present,  it  is  possible  to 
obtain  samples  for  species  identification  and  esti- 
mates of  population  size.  However,  because  bats  are 
only  associated  with  some  roosts  seasonally,  absence 
from  a  potential  roost  site  may  not  mean  it  is  not 
used.  If  bat  guano  or  other  sign,  such  as  urine  stains 
or  remains  of  chewed  insects  are  found,  the  roost 
should  be  visited  at  other  times  for  sampling.  In 
some  instances,  bats  may  use  a  roost  structure  with- 
out leaving  obvious  evidence  of  their  visits.  This  is 
true  at  some  night  roosts  where  several  hundred 
bats  may  use  a  site  without  leaving  much  sign.  If 
such  a  situation  seems  possible  based  on  other  types 
of  evidence  of  bats  presence  in  the  area,  the  investi- 
gator must  visit  the  roost  at  different  times  to  ascer- 
tain usage. 

Potential  water  sources  should  be  investigated. 
If  water  is  limited,  bats  may  be  highly  concentrated 
at  available  sources.  The  presence  of  bats  in  an  area 
may  sometimes  be  determined  easily  by  simple  ob- 
servation at  an  isolated  water  source  during  the  eve- 
ning activity  period.  Observations  should  be  sched- 
uled on  partially  moon-lit  nights  or  may  be  aided  by 
a  low  intensity  light,  such  as  a  gas  lantern,  to  en- 
hance observation  after  dark. 

In  addition  to  drinking  places,  bats  may  also  be 
seen  in  their  foraging  areas,  especially  those  species 
that  begin  their  activities  during  the  twilight  period. 
Some  species  can  be  identified  by  their  characteristic 
flight  but  this  often  requires  considerable  experi- 
ence. Even  if  species  are  not  identified,  the  relative 
number  of  bats  associated  with  different  habitat  fea- 
tures may  be  estimated. 


POPULATION  MEASUREMENT 

Measuring  populations  of  nocturnal  flying  mam- 
mals presents  some  unusual  problems.  It  is  virtually 
impossible  or  impractical  to  achieve  accurate  mea- 
surements of  absolute  density.  However,  because 
many  North  American  bat  species  tend  to  congregate 
at  preferred  roosting  sites  or  at  isolated  water 


sources,  it  is  often  feasible  to  determine  presence 
and  broad  habitat  associations.  In  addition,  the  popu- 
lation features  of  size,  relative  density,  and  age  struc- 
ture of  bats  using  the  congregation  sites  may  often 
be  obtained. 

Presence 

Presence  may  often  be  determined  by  simple 
observation.  Bats  are  sometimes  readily  observed 
while  feeding,  drinking,  or  roosting.  Counts  may  be 
made  but  species  identification  usually  involves  cap- 
turing some  of  the  animals.  There  are  three  practical 
capture  methods:  hand  capture,  mist  netting,  and 
trapping.  These  capture  techniques  all  involve  simi- 
lar handling  methods  associated  with  species  identifi- 
cation, sexing,  aging,  and  marking.  Handling  tech- 
niques are  described  in  association  with  the  first 
sampling  method.  For  more  detailed  and  complete 
accounts,  see  "Bats  and  Bat  Banding"  by  Greenhall 
and  Paradiso  (1968)  or  "Behavioral  and  Ecological 
Methods  for  the  Study  of  Bats,"  T.H.  Kunz,  editor  (in 
press). 

Hand  Capture.  Opportunities  to  capture  bats  di- 
rectly with  the  hand  are  rare.  Such  instances  usually 
come  about  when  bats  are  in  a  torpid  state,  such  as 
when  hibernating  or  when  a  bat  is  sick  or  injured 
and  partially  immobilized.  Bats  should  be  handled 
only  if  absolutely  necessary  and  gloves  should  be 
worn.  Arousal  from  torpor  may  cause  serious  deple- 
tion of  stored  energy  that  might  lead  to  death  at  a 
later  time.  Handling  or  disturbance  of  some  species 
can  force  them  to  abandon  a  preferred  roost  for  one 
that  is  less  favorable  for  survival. 

Hand-capture  techniques  also  include  those  in- 
volving some  implement  manipulated  by  hand.  It  is 
common  to  encounter  bats  that  are  out  of  normal 
reach  or  grasp,  often  in  crevices  or  flying  in  a  con- 
fined space  such  as  a  cave  or  building.  There  are  two 
implements  that  are  particularly  valuable  aids  for 
capturing  such  bats.  Long-nose  scissor  forceps  (bot- 
tle forceps)  are  useful  for  reaching  into  cracks,  crev- 
ices, and  corners  where  a  hand  will  not  fit  or  cannot 
be  extended.  The  tips  of  the  forceps  should  be  pad- 
ded by  covering  them  with  rubber  tubing,  plastic 
rubber,  or  some  substance  that  will  reduce  chances 
of  injury.  Another  useful  capture  tool  is  a  wide- 
mouth  insect  net,  preferably  with  an  adjustable- 
length  handle.  The  net  can  be  used  to  capture  bats 
in  flight,  especially  in  confined  spaces,  or  may  be 
placed  over  a  small  roost  opening  or  over  bats  that 
are  roosting  in  the  open  but  are  out  of  reach.  For 
capturing  bats  in  flight,  sweeping  from  behind  may 
be  more  effective  and  have  less  chance  of  causing  in- 
jury than  a  head-on  sweep.  Both  of  these  equipment 
items  may  be  purchased  from  biological  supply  com- 
panies (see  Dowler  and  Genoways  1976)  and  are 
relatively  inexpensive.  No  special  training  is  required 
for  their  use.  Efficient  use  of  both  of  these  capture 


504 


Bats 


implements  will  usually  be  aided  by  a  hand-held 
flashlight  or,  preferably,  a  head  lamp. 

Handling  Techniques.  Whether  captured  by 
hand  or  by  some  other  method,  it  is  often  necessary 
to  handle  individual  bats.  Handling  may  be  necessary 
to  identify  species,  determine  sex  and  age,  or  mark 
the  bats  for  population  estimates.  Handling  bats 
should  be  done  with  care  to  avoid  injuring  the  ani- 
mals or  being  bitten.  Like  most  wild  mammals,  bats 
will  attempt  to  defend  themselves  by  biting.  They 
are  capable  of  carrying  rabies.  Although  the  inci- 
dence of  rabies  is  not  unusually  high  (estimated  at 
less  than  1%  in  natural  populations),  some  precau- 
tions should  be  taken  such  as  wearing  gloves  and, 
where  there  is  unusually  high  risk  of  bites,  acquiring 
protection  from  rabies  by  pre-exposure  immuniza- 
tion. Information  regarding  preexposure  immuniza- 
tion may  be  obtained  from  a  physician  or  local  pub- 
lic health  service  agency.  In  most  instances  a  pair  of 
pliable  leather  gloves  will  provide  adequate  protec- 
tion against  bites  and  allow  enough  feeling  for  sensi- 
tive manipulation  of  the  bat. 

Care  should  be  taken  not  to  apply  too  much 
pressure  to  a  bat  during  handling  because  their  wing 
(finger)  bones  are  easily  broken  and  some  species 
are  easily  suffocated  or  may  go  into  shock.  Placing 
the  thumb  or  forefinger  under  the  chin  of  the  bat 
while  cradling  the  body  in  the  palm  of  the  hand  al- 
lows immobilization  without  excessive  pressure  (Fig- 
ure 1 ).  Alternatively,  a  bat  may  be  held  with  its  two 


wings  open  by  grasping  each  wing  with  the  thumb 
and  forefinger  at  the  distal  end  of  the  bat's  forearm 
near  the  clawed  thumb.  Using  this  method  requires 
that  measurements  and  marking  be  done  by  a  second 
worker.  Although  the  bat's  head  (and  teeth)  are  far- 
ther from  the  handler's  fingers,  many  bats  will  strug- 
gle violently  in  this  position  and  there  is  often  a 
greater  chance  of  either  being  bitten  or  injuring  the 
bat.  Considering  these  problems  and  the  difficulty  in 
grabbing  a  bat  in  this  position  to  start  with,  it  seems 
more  sensible  to  use  the  finger-under-the-chin 


A  long-eared  Myotis  handled  by  grasping  wings  near  the 
wrist. 


Figure  1.  An  effective  and  relatively  safe  way  to  hold  a  bat  without  being  bitten 
or  injuring  the  bat.  The  worker's  forefinger  or  thumb  is  placed  under  the 
bat's  lower  jaw  while  the  remaining  fingers  restrain  the  body  and  wings.  This 
leaves  the  other  hand  free  for  manipulation,  marking,  or  measuring. 


Bats 


505 


A  big  brown  bat  held  with  the  wings  extended.  When  holding  by  the  wings,  take  care  to  avoid  grasping  the  delicate  finger 
(wing)  bones. 


Grasping  this  Townsend "s  big-eared  bat  with  thumb  and  forefinger  of  one  hand  allows  for  manipulation  with  the  other 
hand. 


506 


Bats 


method  of  immobilization.  The  most  difficult  part  of 
this  method  is  the  initial  seizure  of  the  animal,  which 
is  most  easily  accomplished  by  pinning  it  against 
something  solid. 

It  is  often  necessary  to  retain  some  individual 
bats  temporarily.  For  small  numbers  a  cloth  sack  is 
adequate  as  a  holding  container.  For  larger  numbers 
additional  sacks  or  some  type  of  cage-like  container 
may  be  necessary.  Greenhall  and  Paradiso  ( 1968)  de- 
scribed a  number  of  such  holding  cages.  If  a  cage  is 
deemed  necessary,  it  should  be  constructed  to  fit  the 
specific  conditions  under  which  it  is  likely  to  be 
used. 

The  first  task  of  inspection  is  species  identifica- 
tion. This  is  relatively  simple  for  some  species  but 
difficult  for  others.  A  good  place  to  start  is  with  a 
key  to  the  bat  species  of  North  America  such  as  that 
found  in  Barbour  and  Davis  (1969).  There  have  been 
some  taxonomic  revisions  since  that  work  was  pub- 
lished, so  it  would  be  wise  to  check  a  more  recent 
publication,  perhaps  one  that  is  focused  on  the  spe- 
cific region  of  work.  For  some  species  there  is  no 
foolproof  method  of  identification  in  the  field,  such 
as  distinguishing  the  Yuma  myotis  from  the  little 
brown  myotis  in  areas  where  they  occur  together. 
Some  other  species  of  Myotis  are  particularly  diffi- 
cult to  identify  without  experience.  Time  spent  with 
a  regional  key  and  collection,  such  as  one  might  find 
at  a  local  college  or  university,  may  be  helpful. 

It  is  often  necessary  to  determine  sex,  and 
sometimes  age,  of  individuals  so  that  the  type  of  ag- 
gregation (breeding,  maternity,  or  bachelor)  may  be 
ascertained.  The  sex  of  bats  is  generally  easy  to  de- 
termine. The  males  have  an  obvious  penis  and  the 
clitoris  of  the  females  is  not  enlarged.  Testes  of 
males  and  mammary  nipples  of  females  are  seasona- 
bly obvious.  Aging  is  difficult  after  the  juvenile  pe- 
riod. Juveniles  less  than  4-5  months  old  can  be  rec- 
ognized by  the  cartilaginous  areas  in  the  finger 
joints.  The  joints  are  lighter  in  color  and  relatively 
smooth  compared  with  the  darker,  knobby  appear- 
ance of  the  adult  joints.  Tooth  wear  can  only  give  a 
general  impression  of  age;  canines  and  premolars  be- 
come obviously  round  in  very  old  individuals. 

After  inspection,  bats  may  be  released  by  care- 
fully launching  them  into  flight  such  as  one  might 
pitch  a  horseshoe,  or  by  quickly  releasing  the  grip 
and  allowing  them  to  take  off  on  their  own.  The 
launching  method  is  preferred  for  fast  flying  species 
that  need  some  initial  speed  to  gain  flight  and  may- 
be used  for  all  species  unless  holding  time  is  pro- 
longed. Some  species  become  progressively  more  le- 
thargic the  longer  they  are  held  in  captivity,  espe- 
cially under  cold  conditions,  and  they  should  not  be 
launched  or  dropped.  The  best  release  method  in 
this  situation  is  to  simply  place  the  bat  as  high  as 
possible  above  the  ground  on  a  solid  substrate  such 


as  a  rock,  wall,  or  tree.  As  soon  as  the  bat  warms  up 
it  will  fly  off  on  its  own.  Bats  released  from  grasp  in 
this  fashion  will  often  attempt  to  bite  if  the  restrain- 
ing hold  is  gradually  reduced.  Those  released  by 
launching  or  rapid  relaxation  of  the  grip  usually 
make  no  attempt  to  bite. 

Marking  Techniques.  Bats  are  often  marked  to 
aid  in  determining  movements  or  to  identify  them  as 
previously  captured  individuals  for  population  esti- 
mates. A  variety  of  marking  techniques  are  available, 
but  two — banding  and  punch-marking — are  most 
useful  for  studying  populations  related  to  habitat  in- 
ventories. For  details  of  other  techniques,  see  Bar- 
clay (in  press). 

Banding  involves  placing  a  small  plastic  band 
around  the  distal  end  of  the  forearm  where  the  lead- 
ing edge  of  the  wing  membrane  is  very  narrow.  For 
the  North  American  species  being  considered  here, 
use  of  lipped  or  flanged  plastic  bands  applied  loosely 
so  they  slide  freely  along  the  forearm  will  reduce  in- 
juries to  the  bats  (Stebbings  1978). 

Banding  with  aluminum  bands  should  be 
avoided  because  identification  numbers  may  be  lost 
as  a  result  of  chewing  by  the  bats.  The  use  of  plastic 
bands  with  large  embossed  numbers  or  color  codes 
will  solve  this  problem  (Bonaccorso  et  al.  1976).  For 
local  studies  of  abundance  and  movements,  the  col- 
ored plastic  bands  are  easily  coded  to  represent  date 
and  location  of  banding.  Bands  may  be  purchased 
from  A.C.  Hughes,  1  High  St.,  Hampton  Hill,  Middle- 
sex, Eng.  TW121WA  (colored  plastic);  Gey  Band  and 
Tag  Co.,  Box  363,  Norristown,  PA  19404;  National 
Band  and  Tag  Co.,  721  York  St.,  Newport,  KY  41071; 
and  Ball  Chain  Mfg.  Co.,  Inc.,  741  South  Fulton  Ave., 
Mt.  Vernon,  NY  10550. 


Aluminum  band  on  the  forearm  of  a  long-legged  Myotis. 


Bats 


507 


A  second  method  of  marking  bats  is  punch- 
marking  (Bonaccorso  and  Smythe  1972).  The  tech- 
nique consists  of  punching  small  holes  in  the  form  of 
numbers  through  the  outstretched  wing  membrane. 
This  is  accomplished  with  a  tattoo  instrument  nor- 
mally used  to  mark  domestic  livestock.  These  instru- 
ments may  be  purchased  from  veterinary  supply 
companies  or  some  livestock  supply  stores,  and  they 
come  in  at  least  two  sizes,  for  large  and  small  ani- 
mals. With  care,  the  large  animal  tattoo  marker  can 
be  used  for  all  species  of  North  American  bats.  It  has 
the  advantage  of  larger  perforations  (both  size  of 
numbers  and  size  of  holes)  that  are  more  legible  and 
longer  lasting  than  those  made  by  the  small  animal 
instrument.  Although  initially  touted  as  a  suitable 
substitute  for  banding,  it  was  quickly  learned  that 
punch-marking  had  the  drawback  of  being  short- 
lived (Bonaccorso  et  al.  1976).  Punch  marks  remain 
legible  for  only  about  5  months.  Nevertheless,  the 
advantages  are  numerous  and  the  technique  is  suita- 
ble for  short-term  population  inventories  related  to 
habitat  monitoring. 


Several  publications  are  available  that  describe 
the  technique  in  considerable  detail.  See  Bleitz 
( 1984)  and  Keyes  and  Grue  ( 1982)  for  general  pro- 
cedure and  review  of  history  and  techniques.  See 
Greenhall  and  Paradiso  (1968)  and  Kunz  (in  press) 
for  application  of  the  techniques  to  bats.  A  general 
description  of  the  technique  will  be  given  here  with 
emphasis  on  those  areas  that  are  important  or 
unique  with  respect  to  bats.  This  information  should 
allow  an  investigator  to  determine  if  the  technique  is 
suitable  for  a  given  bat  sampling  situation.  The 
proper  use  of  mist  nets  requires  experience  and  is 


Punch-marking  a  long-eared  Myotis  with  a  large  animal 
tattoo  device. 


A  Yuma  Myotis  with  punch-marked  wing. 


greatly  facilitated  by  field  training.  If  instruction  is 
not  available  from  someone  working  with  bats,  the 
best  alternative  is  to  work  with  a  knowledgeable 
bird  netter.  Names  and  addresses  of  such  individuals 
may  be  obtained  from  a  regional  bird  banding  associ- 
ation (see  Bleitz  1984).  Attempts  to  "start  from 
scratch"  may  result  in  injury  or  stress  to  the  bats, 
damage  to  the  nets,  and  frustration  to  the  investiga- 
tor. 


Each  individual  bat  may  be  marked  with  a  dis- 
tinct number  but  this  is  usually  unnecessary.  A  sim- 
pler and  adequate  approach  for  population  census 
and  local  movement  studies  is  to  use  a  distinctive 
numerical  sequence  (letters  may  be  included)  for 
each  marking  episode  at  each  location.  If  necessary, 
additional  punch  marks  may  be  made  in  the  same 
wing  or  opposite  wing  on  subsequent  encounters. 

Mist-Netting.  The  mist  net  was  developed  by  the 
Japanese  to  capture  birds  and  is  still  used  extensively 
for  that  purpose.  Because  mist-netting  is  the  most 
commonly  used  capture  method  in  bird  research,  it 
has  received  a  great  deal  of  attention  from  ornitholo- 
gists (see  review  by  Keyes  and  Grue  1982).  The 
basic  technique  and  many  of  the  recent  innovations 
developed  by  bird  netters  are  readily  applicable  to 
bats. 


Basic  Technique.  The  basic  technique  consists 
of  placing  a  large  rectangular  net,  having  several  tiers 
made  of  fine  hairnet-like  material,  in  a  position 
where  it  will  be  in  the  pathway  of  flying  bats.  For 
best  results,  these  nets  are  placed  where  bats  are 
concentrated  such  as  near  their  unique  habitat  fea- 
tures or  drinking  sites.  Bats  attempting  to  enter  or 
leave  a  roost  site,  drink  water  from  an  open  water 
source,  or  forage  in  a  relatively  confined  area  are 
vulnerable  to  capture  with  mist  nets.  Bats  are  cap- 
tured when  they  strike  the  netting  and  fall  into  ham- 
mock-like pouches  at  the  lower  side  of  each  shelf  of 
the  net  (Figure  2).  Entangled  bats  are  removed  from 
the  net  by  hand. 

Equipment.  Mist  nets  are  available  in  a  variety 
of  colors,  sizes,  and  construction  material.  A  list  of 
suppliers  may  be  found  in  Dowler  and  Genoways 


508 


Bats 


The  wing  of  a  long-eared  Myotis  with  perforations  resulting  from  the  punch-marking  process.  The  tattoo  device  used  for 
making  the  marks  is  shown  below  the  wing. 


Figure  2.  Diagrammatic  view  of  a  mist  net  setup.  The  net  has  five  cross  lines  forming  four  tiers  and  associated 
pockets  where  the  bats  are  usually  entrapped. 


Bats 


509 


(1976).  The  capture  location,  situation,  and  size  of 
bats  to  be  captured  should  be  considered  when  de- 
termining the  most  suitable  net  characteristics.  Black 
nets  are  suitable  for  bats  because  most  capturing  is 
done  at  night  and  the  color  does  not  contrast  with 
the  background. 

Net  sizes  vary  in  length,  height,  number  of  tiers, 
and  mesh.  All  sizes  for  nets  are  given  in  the  fully- 
stretched  position.  The  most  common  sizes  used  for 
bat  netting  are  about  5m(18ft),  9m(30ft),  and 
13  m  (42  ft).  Nets  longer  than  13  m  (42  ft)  are  diffi- 
cult to  operate  without  some  sort  of  center  support 
or  heavy  guying.  Short-length  nets  are  useful  for 
roost  entrances,  small  ponds,  or  foraging  pathways  in 
heavy  vegetation.  The  9-m  (30-ft)  net  is  very  useful 
for  netting  over  small  ponds  or  in  foraging  areas 
with  larger  open  spaces.  Most  nets  used  for  captur- 
ing bats  are  about  2  m  (7  ft)  high  and  have  four  tiers 
(also  referred  to  as  shelves  or  pockets).  The  working 
height  becomes  less  than  2  m  (6.6  ft)  when  the 
cross  lines  (trammels)  are  compressed  to  form  the 
pockets.  Nets  with  two  or  five  shelves  are  also  com- 
mercially available.  Other  widths  (heights)  may  be 
obtained  by  removing  shelves  of  the  standard  nets  or 
by  stacking  nets  to  cover  a  greater  vertical  distance. 

Mesh  sizes  (designated  by  the  diagonal  of  each 
square)  useful  for  netting  North  American  bats  range 
from  3-8  cm  (1.5  in.)  for  small  species  to  6.4  cm 
(2.5  in.)  for  large  species.  Because  many  regions  sup- 
port a  combination  of  large  and  small  species,  it  is 
safest  to  use  a  relatively  small  mesh  size. 

Nylon  and  terylene  are  common  construction 
materials;  terylene  is  stronger  but  less  pliable. 
Strength  is  determined  by  the  "ply"  (number  of 
strands  per  thread)  and  "denier"  (a  weight  measure- 
ment). Nets  for  capturing  bats  are  usually  2  ply  and 
range  from  50  to  110  denier  (d),  are  black  color,  ny- 
lon, 70  d,  2  ply,  3-8-cm  (1.5-in.)  mesh,  4  tiers,  2  m 
(7  ft)  high,  and  9  m  (30  ft)  long. 

Many  types  of  poles  may  be  used  to  hold  the 
ends  of  a  mist  net  in  place.  Electrical  conduit,  which 
may  be  cut  into  portable  sections  that  can  be  cou- 
pled in  the  field,  is  very  commonly  used  for  this  pur- 
pose. Keyes  and  Grue  ( 1982)  described  many  of  the 
available  alternatives.  Because  deployment  condi- 
tions usually  vary,  sometimes  greatly  between  sites, 
it  is  advisable  to  have  versatile  equipment. 

Nylon  cord  is  an  essential  item  for  guying  the 
tops  of  any  pole  setup.  Other  useful  items  include  a 
headlamp,  so  both  hands  are  free  for  removing  bats 
from  the  net;  gloves;  cloth  holding  sacks  that  are  eas- 
ily carried  under  the  belt;  and  insect  repellent.  When 
netting  over  water,  hip  boots  or  chest  waders  are 
often  necessary  to  facilitate  removing  bats  from  the 
net.  Another  alternative  in  such  situations  is  to  use  a 
boat  or  raft.  Some  items  that  appear  to  be  luxuries, 


such  as  a  comfortable  chair  and  warm  clothing,  often 
become  necessities  as  a  night  of  netting  wears  on. 

Operation.  Nets  should  be  set  up  well  before 
dark  but  the  cross  lines  should  be  bunched  until  bats 
are  seen  flying  or  when  fully  dark.  This  prevents  cap- 
turing birds  that  are  active  during  the  daylight  or 
twilight  periods.  Position  of  the  nets  is  often  critical 
and  influences  capture  success.  If  possible,  the  nets 
should  be  set  so  they  are  perpendicular  to  the  direc- 
tion that  the  bats  are  most  likely  to  fly.  This  is  rela- 
tively easy  to  determine  at  roost  sites  but  often  in- 
volves experience  with  the  particular  situation  at 
drinking  and  foraging  sites  (Figure  3).  If  possible, 
several  nets  should  be  set  up  in  such  situations  so  as 
to  find  the  most  efficient  positions.  Generally,  when 
nets  are  set  over  somewhat  rectangular  bodies  of 
water  they  should  be  positioned  perpendicular  to 
the  long  axis  of  the  pond.  Other  features  such  as  ob- 
structions to  flight  should  also  be  considered  and  the 
nets  placed  where  there  is  the  least  interference 
with  the  bats'  flight  patterns.  In  open  areas  where 
there  are  no  criteria  for  orientation  of  a  single  net, 
several  nets  may  be  set  at  different  angles.  Common 
configurations  are  V,  L,  and  triangles.  Vertical  stack- 
ing of  nets  to  achieve  greater  height  is  a  variation 
particularly  useful  in  foraging  areas. 

Nets  should  be  checked  regularly — at  least 
every  15  minutes.  Bats  captured  in  a  net  should  be 
removed  as  quickly  as  possible  to  avoid  confounding 
entanglements  and  holes  in  the  net  resulting  from 
the  bats  chewing  or  the  netter  having  to  break  the 
strands  to  free  the  bats.  A  captured  bat  should  be  re- 
moved from  the  side  of  the  net  from  which  it  en- 
tered a  pocket.  A  good  sequence  for  untangling  is 
head,  one  wing,  other  wing,  and  body. 

When  removing  a  bat  from  a  net,  the  bat  may 
be  grasped  in  such  a  way  as  to  immobilize  the 
mouth,  as  described  previously,  or  by  using  a  cloth 
baffle  on  the  opposite  side  of  the  net  for  support  and 
to  provide  something  the  bat  can  get  its  teeth  into. 
The  cloth  baffle  is  also  useful  as  a  backing  to  get  an 
initial  hold  on  the  bat.  Inducing  a  bat  to  release  its 
bite  on  the  baffle  or  other  object  (such  as  a  finger) 
may  be  achieved  by  mild  pulling  pressure,  by  touch- 
ing the  back  of  the  bat's  head,  by  blowing  on  the  an- 
imal, or  by  a  combination  of  these  strategies.  After 
removal  from  the  net,  the  bats  may  be  held  in  a 
cloth  sack  or  holding  cage,  or  marked  and  released 
immediately.  Bats  flying  nearby  may  be  attracted  to 
the  calls  of  captives  held  near  a  net  or  trap  site. 

The  duration  and  timing  of  netting  depends 
upon  the  objectives  of  the  investigator.  It  is  rela- 
tively easy  to  determine  when  the  evening  emer- 
gence flight  of  bats  from  a  day  roost  is  finished.  In 
contrast,  the  period  of  greatest  bat  activity  at  night 
roosts  may  be  difficult  to  predict  and  may  occur  dur- 
ing the  middle  of  the  night.  Netting  at  foraging  sites 


510 


Bats 


***- y*-i 


wmrn 


Figure  3.  Some  possible  placements  for  mist  nets  to  capture  bats. 


Bats 


511 


1/7     // 
hi  hit 


y^i  ff/h 


\  :;J  f 


u    ■/  /. 


V  V 


\  \,  '  '  \ 


in  places  where  there  are  extremely  high  concentra- 
tions of  bats.  For  example,  at  the  entrance  of  some 
cave  roosting  sites,  the  captured  bats  cannot  be  un- 
tangled quickly  enough  to  avoid  bat  injury  or  dam- 
aged nets.  Windy  conditions  greatly  diminish  netting 
success  because  of  reduced  bat  activity  or  decreased 
net  effectiveness. 

Trapping.  An  alternative  to  mist-netting  is  trapping, 
first  developed  by  Constantine  (1958).  He  used  a  so- 
called  "harp  trap"  which  is  designed  to  capture  and 
hold  bats.  Like  a  mist  net,  a  trap  is  placed  where  it  is 
likely  to  intercept  passing  bats.  Unlike  the  mist  net, 
it  does  not  require  constant  attention  and  can  be 
used  where  the  density  of  flying  bats  is  very  high 
and  where  space  is  restricted. 

Basic  Technique.  The  basic  technique  consists 
of  placing  a  metal  frame,  strung  with  vertical  strands 
of  wire  or  nylon,  directly  in  the  path  of  flying  bats 
(Figure  4).  The  strands  of  the  trap  are  closely  spaced 
and  kept  taut  so  that  when  a  bat  strikes  them  it  will 
lose  control  of  its  flight  and  slide  to  the  bottom  of 
the  frame.  A  sack-like  container  is  suspended  from 
the  bottom  of  the  frame  to  serve  as  a  retaining  struc- 
ture or  as  a  funnel  to  direct  the  bats  into  a  holding 
container.  Bats  captured  in  such  a  device  may  be 
handled  immediately  or  left  until  a  more  convenient 
time  for  the  investigator. 


A  bat  entrapped  in  the  pocket  of  a  mist  net.  Captured  bats 
should  be  removed  as  soon  as  possible  to  avoid  confound- 
ing entanglements.  Bats  will  also  chew  the  netting  which 
results  in  holes  (pictured). 

or  at  drinking  sites  is  usually  most  productive  within 
2  to  3  hours  after  sunset.  However,  it  appears  that 
different  species  have  different  peak  periods  for  feed- 
ing and  drinking  so  it  is  desirable  to  sample  for  as 
long  as  possible  for  species  presence  and  to  sample 
for  a  constant  period  when  attempting  to  estimate 
population  size  for  comparative  purposes. 

Upon  completion  of  a  sampling  episode  the  nets 
should  be  dried,  if  necessary,  and  all  debris  removed. 
Each  net  should  be  folded  in  such  a  manner  that  it 
can  be  easily  unfolded  for  the  next  use  (e.g.,  see 
Greenhall  and  Paradiso  1968).  Proper  care  at  the 
end  of  one  session,  including  tying  lines  so  they  stay 
in  order  and  making  notations  regarding  the  condi- 
tion of  the  net,  will  greatly  facilitate  setup  at  the 
next  session. 

Although  mist-netting  is  a  versatile  and  fairly  ef- 
ficient method  of  capturing  bats,  it  has  some  limita- 
tions. Netting  is  most  efficient  under  very  specific 
circumstances,  e.g.,  in  areas  where  bats  concentrate 
to  forage  or  drink  and  where  there  is  sufficient  space 
to  deploy  the  net.  Mist  nets  are  often  inappropriate 


Equipment.  Bat  traps  are  not  readily  available 
commercially  but  are  fairly  easy  to  build.  Informa- 
tion presented  here  will  give  a  rough  idea  of  design; 
details  of  construction  are  given  in  the  references 
cited  below. 

The  original  design  (Constantine  1958;  see  also 
Greenhall  and  Paradiso  1968),  commonly  referred  to 
as  the  Constantine  trap,  consists  of  a  single  rectangu- 
lar aluminum  frame  with  vertically  arranged  taut 
steel  wires  spaced  1  in.  (2.5  cm)  apart.  Bats  stopped 
in  flight  slide  down  the  wire  and  are  directed 
through  a  large  funnel-shaped  plastic  catchment  into 
an  escape-proof  cage.  Constantine  later  modified  his 
original  design  by  varying  the  size  and  collapsibility 
of  the  trap  to  conform  to  specific  spatial  needs  and 
logistical  constraints. 

Tuttle  (1974)  further  modified  the  harp  trap 
and  provided  detailed  construction  plans.  Tuttle's 
trap  uses  two  adjacent  frames,  157.5  cm  (62  in.) 
square,  spaced  about  8  cm  (3  in.)  apart.  The  tautness 
of  the  vertical  strings  is  adjustable  as  is  the  distance 
between  frames.  These  adjustments  allow  for  differ- 
ences in  conditions,  and  size  and  flight  characteris- 
tics of  different  species  of  bats.  The  strings  may  con- 
sist of  0.20-mm  (0.008-in.)  stainless  spring-steel  wire 
or  6-  to  20-lb  (3-  to  9-kg)  test  monofilament  nylon 
line.  Ideally,  bats  hitting  the  strings  should  not 
bounce  off  but  slide  down  the  first  bank  of  strings  or 
pass  into  the  space  between  frames  and  then  slide 


512 


Bats 


TENSION  ADJUSTMENTS 


FRAMES 


NYLON  OR 
WIRE  STRANDS 


Figure  4.  Harp-type  bat  trap  with  major  part  labeled 
(after  Tuttle  1974). 


down.  The  canvas  holding  bag  is  lined  with  plastic 
that  is  attached  at  the  top  but  hangs  free  at  the  bot- 
tom. The  bats  slide  in  easily  but  instead  of  crawling 
up  the  outer  slick  plastic  surface,  they  crawl  up  the 
cloth  sides,  under  the  plastic  to  a  blind  ending.  Here, 
unless  large  numbers  are  present,  they  usually  re- 
main quiet  until  harvested  by  the  trapper.  The  width 
of  the  opening  of  the  holding  bag  is  adjustable  to  ac- 
commodate different  sized  species.  The  trap  also  has 
adjustable  legs  to  allow  for  changes  in  height. 

The  Tuttle  harp  trap  is  designed  to  be  portable 
but  field  assembly  requires  about  45  minutes.  If  sam- 
ple sites  are  relatively  accessible  and  the  frames  can 
be  transported  without  dismantling,  assembly  time 
may  be  reduced.  In  fact,  permanent  non-collapsible 
frames  may  be  practical  in  most  situations  where  the 
sample  sites  are  accessible  and  have  limited  physical 
variation. 

Tidemann  and  Woodside  ( 1978)  gave  design 
and  construction  details  for  a  light-weight  collapsible 
version  of  the  harp  trap.  By  using  telescoping  sec- 
tions of  aluminum  tubing,  the  trap  becomes  a  rela- 
tively portable  package  95  cm  (37.4  in.)  long  and  15 
cm  (5.9  in.)  in  diameter,  weighing  7  kg  ( 15  lb). 


They  maintain  that  the  trap  may  be  assembled  in 
about  5  minutes.  If  potential  sample  sites  are  inac- 
cessible or  if  transport  of  a  fully  assembled  trap  is 
not  possible,  this  construction  option  should  be 
considered. 

Operation.  Typically,  bat  traps  have  much  less 
capture  surface  area  than  mist  nets.  Consequently, 
they  are  particularly  appropriate  where  bats  are 
flying  in  rather  confined  spaces  such  as  cave  en- 
trances, small  heavily-canopied  streams,  small  iso- 
lated pools  of  water,  building  openings,  and  along 
narrow  pathways  in  thickly-vegetated  areas.  In  many 
situations,  the  open  area  around  the  trap  may  be 
blocked  off  to  direct  the  bats  into  the  trap.  When 
necessary,  several  traps  may  be  used  simultaneously 
because  the  bats  are  retained  without  harm  until  it  is 
convenient  for  the  investigator  to  deal  with  them.  In 
most  instances  it  is  necessary  to  check  a  trap  at  least 
every  hour  to  ensure  that  the  holding  bag  is  not 
overloaded  and  to  release  bats  as  soon  as  possible  so 
they  can  resume  their  normal  activity. 

Other  Techniques.  Some  other  techniques  are  oc- 
casionally useful  for  determining  presence  of  bats. 
Generally  these  techniques  are  used  under  more  spe- 
cialized circumstances  than  those  described  above 
and,  therefore,  will  not  be  treated  in  the  same  detail. 


Bat  trap  set  up  in  cave  opening.  Bats  slide  down  vertical 
strands  into  holding  bag. 


Bats 


513 


Bat  trap  set  up  over  open  area  of  small  stream. 


Detection  of  Echolocation  Calls.  Flying  bats 
produce  high  frequency  sounds  for  communication, 
orientation,  and  prey  capture  (see  Fenton  1982). 
Most  such  calls  are  above  the  range  of  unaided  hu- 
man detection  and  are  referred  to  as  "ultrasonic."  In 
recent  years,  equipment  has  become  available  that 
allows  detection  of  these  ultrasonic  calls  in  the  field. 
Echolocation  calls,  especially  those  used  for  pursuit 
and  capture  of  prey,  are  often  recognizable  as  spe- 
cies-specific (Fenton  1982). 

The  technique  has  several  features  that  are  use- 
ful for  wildlife  management  applications.  Species 
presence  in  a  general  area  or  specific  habitats  may 
often  be  determined  without  actually  handling  the 
animals  or  interfering  with  their  normal  activity.  Un- 
like mist-netting  and  trapping,  this  technique  is  not 
restricted  to  sites  with  very  specific  capture  condi- 
tions. It  may  be  especially  useful  in  areas  where  con- 
gregation sites  are  unknown,  inaccessible,  or  not 
conducive  to  sampling  by  conventional  methods. 
The  technique  may  also  be  useful  in  conjunction 
with  more  conventional  methods.  For  example,  an 
ultrasonic  detector  may  be  a  valuable  tool  in  deter- 
mining optimal  habitats  and  time  intervals  for  suc- 
cessful deployment  of  mist  nets  (Kunz  and  Brock 
1975).  In  other  instances,  a  detector  may  be  used  to 
monitor  degree  of  activity  or  to  aid  in  counting  the 
number  of  bats  at  a  specific  place  after  some  of  the 
bats  have  been  captured  to  determine  or  verify  spe- 
cies identity. 

The  technique  does  have  some  limitations.  Some 
bat  species  are  not  easily  detected  and  some  are  not 
easily  distinguished  from  others.  It  cannot  be  used  to 
determine  population  density  because  calls  of  spe- 
cific individuals  cannot  be  identified.  However,  it  can 
sometimes  be  used  to  count  the  number  of  individu- 
als passing  by  a  restricted  area  in  one  direction,  such 
as  a  roost  opening. 


Ultrasonic  detection  equipment  varies  in  com- 
plexity, cost,  and  availability  (Simmons  et  al.  1979; 
Fenton  in  press).  The  cost  of  some  sophisticated 
units  is  beyond  the  reach  of  local  management  units. 
In  addition,  equipment  operation  and  data  interpreta- 
tion often  require  extensive  training  and  experience. 
If  information  that  is  only  obtainable  with  this 
method  is  deemed  essential  for  a  land  management 
decision,  it  may  be  desirable  to  contract  specialists 
for  its  collection  rather  than  to  attempt  to  acquire 
the  equipment  and  training  at  the  local  level.  How- 
ever, some  inexpensive,  easy-to-build,  and  relatively 
easy-to-use  systems  are  becoming  available  (e.g.,  see 
Paige  et  al.  1985)  and  may  provide  some  valuable  in- 
formation useful  for  management  purposes. 

Observation.  The  presence  of  bats  in  general 
and  a  few  species  in  particular  can  be  determined  by 
direct  observation.  This  is  best  done  in  the  evening 
twilight  period.  Bats  may  also  be  seen  during  morn- 
ing twilight  although  they  are  not  usually  as  numer- 
ous as  in  the  evening.  It  is  often  possible  to  view 
bats  around  attractive  habitat  features  such  as  roost- 
ing, drinking,  and  foraging  sites,  with  the  aid  of  low- 
intensity  artificial  red  light.  Enhancement  of  vision  is 
also  possible  through  the  use  of  a  night-  vision 
scope,  sometimes  aided  by  an  infrared  light  source, 
but  the  value  of  this  instrument  is  somewhat  limited 
because  of  its  relatively  small  field  of  view.  Such  an 
instrument  is  best  used  where  areas  of  bat  flight  are 
restricted,  to  determine  population  size  of  known 
species  with  a  minimum  of  disturbance  (e.g.,  see 
Bagley  and  Jacobs  1985).  For  further  details  con- 
cerning observational  techniques,  see  Barclay  (in 
press). 

Advertising.  One  method  of  finding  concentra- 
tions of  bats,  especially  roosting  sites  in  areas  fre- 
quented by  humans,  is  to  advertise.  Most  local  news- 
papers will  run  stories  about  bats  that  include  a 
request  for  information  concerning  their  wherea- 
bouts. Some  radio  stations,  especially  public  service 
types,  will  do  the  same.  Finally,  posters  may  be 
printed  and  distributed  in  the  area  of  concern.  Most 
responses  obtained  from  these  methods  are  from 
people  wanting  to  know  how  to  get  rid  of  bats  in 
man-made  structures.  This  provides  the  opportunity 
for  some  conservation  education  in  addition  to  locat- 
ing local  populations  of  bats. 

Shooting.  Because  many  bats  emerge  from 
their  daytime  roosts  during  the  twilight  period,  it  is 
often  possible  to  collect  them  by  shooting.  This  may 
be  done  with  a  standard  shotgun  using  shells  loaded 
with  very  small  shot  (No.  9  or  higher).  A  small-gauge 
light-weight  shotgun  with  light  loads  is  usually  most 
efficient  for  hitting  erratically  flying  bats  in  dim  light. 
A  pistol  or  small-bore  rifle  loaded  with  shot  shells 
may  be  used  in  the  same  manner,  but  the  shot  pat- 
tern produced  is  smaller  and  less  uniform  than  that 
of  a  shotgun. 


514 


Bats 


Killing  an  animal  forecloses  the  option  of  subse- 
quent release,  and  shooting  should  only  be  used  if 
absolutely  necessary.  If  roost  sites  are  unknown  or 
inaccessible  and  other  techniques  are  not  applicable 
in  an  area  where  bats  are  known  to  occur,  this  op- 
tion might  be  justified  to  establish  presence.  If  possi- 
ble, all  specimens  collected  in  this  fashion  should  be 
prepared  as  study  skins  or  preserved  in  alcohol  and 
deposited  in  a  regional  museum  for  future  reference. 
Such  specimens  may  be  necessary  for  species  identi- 
fication or  may  serve  as  vouchers  to  verify  the  pres- 
ence of  species  in  an  area. 

The  drawbacks  of  the  technique  are  obvious. 
Whereas  only  one  or  two  specimens  may  be  neces- 
sary for  species  identification  or  to  serve  as  vouch- 
ers, many  more  may  be  killed  needlessly  while  trying 
to  collect  specific  species.  Shooting  at  twilight  is 
dangerous  and  illegal  in  some  areas  without  special 
permission.  The  technique  should  only  be  used  as  a 
last  resort  when  other  more  selective  and  less  de- 
structive methods  are  not  applicable. 


Relative  Density 

Relative  density  values  of  bats  may  be  computed 
from  data  obtained  by  some  of  the  techniques  de- 
scribed above.  It  is  important  to  realize  that  all  such 
density  values  are  biased  to  some  degree  because  of 
selectivity  for  certain  species.  The  degree  of  tech- 
nique selectivity  and  the  intended  use  of  the  relative 
density  values  should  be  considered  when  collecting 
and  analyzing  data.  There  are  two  primary,  but  not 
exclusive,  uses  of  relative  density  values.  First,  rela- 
tive densities  of  temporally  or  spatially  separated 
populations  may  be  compared  to  detect  possible  dif- 
ferences in  community  composition.  Second,  relative 
density  values  may  be  viewed  as  valid  indicators  of 
bat  community  structure.  Considering  the  built-in 
biases  and  possible  uses,  some  specific  sampling 
situations  can  be  described  to  illustrate  possible 
applications. 

Capture  at  day  roosts  is  often  highly  species-se- 
lective and  therefore  of  little  value  for  calculating 
relative  density  that  reflects  overall  bat  community 
structure.  However,  day  roosts  that  are  used  by 
more  than  one  species  may  yield  relative  density  val- 
ues that  are  useful  for  evaluating  changes  over  time 
at  a  particular  site.  Site-to-site  comparisons  are  usu- 
ally not  possible  because  different  physical  features 
make  it  impossible  to  sample  in  an  identical  manner. 
Some  night  roosts  such  as  caves  or  tunnels  may  be 
used  by  several  species  at  once  and  therefore  yield 
data  that  may  be  used  to  calculate  relative  densities. 
Again,  these  values  are  best  used  for  temporal  com- 
parisons related  to  habitat  or  seasonal  changes  at  a 
given  site.  They  are  not  good  indicators  of  total  bat 
community  structure  because  not  all  bats  in  an  area 
are  likely  to  use  the  same  type  of  night  roost.  They 


may,  however,  indicate  the  community  of  species 
that  is  utilizing  a  particular  habitat  feature. 

Sampling  at  foraging  sites  is  also  selective  and, 
therefore,  not  a  good  measure  of  total  community 
structure.  Sampling  at  communal  drinking  sites  offers 
perhaps  the  best  opportunity  to  collect  relative  den- 
sity data  that  reflect  community  structure.  This  is 
particularly  true  where  water  sources  are  scarce  and 
the  site  is  large  enough  to  accommodate  all  species 
potentially  in  an  area.  Data  obtained  at  such  sites 
may  also  be  used  to  calculate  species  diversity  values 
that  may  be  useful  for  comparative  and  descriptive 
purposes. 

To  determine  species  richness  for  a  habitat  or 
area,  one  should  use  as  many  forms  of  capture  and 
detection  as  are  available. 


Absolute  Density 

It  is  very  difficult,  if  not  impossible,  to  deter- 
mine the  absolute  density  of  bat  populations.  There 
are  two  basic  problems:  ( 1 )  samples  are  taken  at 
points  of  concentration  rather  that  at  random,  and 
(2)  it  is  extremely  difficult  to  determine  the  total 
area  used  by  individuals  in  the  sample.  A  few  at- 
tempts have  been  made  to  estimate  density  of  bats 
(see  Gaisler  1979)  but  there  are  no  standard  tech- 
niques similar  to  those  used  for  other  small  mam- 
mals or  birds. 

Whereas  measurement  of  absolute  density  is  not 
a  feasible  objective,  it  is  often  possible  to  measure 
population  size  at  the  places  where  bats  congregate. 
The  values  thus  derived  may  be  compared,  at  least 
temporally,  to  look  for  trends  (seasonal  or  annual) 
or  the  effects  of  environmental  alterations.  Several 
methods  have  been  used  to  estimate  the  size  of  se- 
lected bat  concentrations  (see  Laval  in  press).  Often, 
with  small  colonies  in  roosting  sites,  it  is  possible  to 
simply  count  the  bats  present.  This  is  especially  ap- 
plicable to  hibernacula  where  the  bats  are  extremely 
immobile  and  not  easily  disturbed  if  care  is  taken  by 
the  investigator.  In  large  aggregations,  where  it  is  im- 
possible to  count  every  individual,  the  number  of 
bats  in  several  sample  areas  of  known  size  may  be 
counted  and  the  estimate  made  by  extrapolation 
from  the  total  area  covered.  Photographic  analysis 
may  be  used  in  the  same  manner  and  may  also  be 
used  to  estimate  population  size  by  taking  pictures 
of  bats  emerging  from  day  roosts  (Altenbach  et  al. 
1979;  Warden  1980).  The  number  of  flying  bats  can 
be  counted  in  each  sample  picture  leading  to  the  es- 
timate of  bats  per  unit  of  time.  If  the  total  time  of 
the  emergence  flight  is  also  known  then  population 
size  may  be  estimated. 

Population  estimates  may  also  be  made  from 
capture-recapture  data.  However,  it  appears  that 


Bats 


515 


most  capture  methods  cause  some  alienation  to  a 
particular  site  and  chances  of  recapture  are  dimin- 
ished. Nevertheless,  if  such  estimates  are  obtained  in 
a  consistent  manner  from  place  to  place  and  time  to 
time,  they  can  provide  a  quantitative  value  that  is 
useful  for  comparison.  Such  values  are  usually  over- 
estimates of  the  real  population  because  animals  are 
more  likely  to  become  capture-shy  rather  than  cap- 
ture-prone. With  consideration  of  these  weaknesses, 
a  simple  Lincoln/Peterson  estimator,  or  some  modifi- 
cation, such  as  that  of  Gaisler  ( 1979),  will  yield  ade- 
quate estimates. 


DISCUSSION 

The  problems  of  monitoring  bat  populations  and 
their  habitats  are  challenging.  To  address  these  prob- 
lems, one  must  be  aware  of  the  peculiarities  of  bat 
behavior  and  the  limitations  of  the  specialized  tech- 
niques associated  with  their  study. 

Two  characteristics  of  many  bat  species  make 
their  populations  particularly  susceptible  to  sharp 
population  declines.  First,  their  tendency  to  congre- 


gate often  makes  them  especially  vulnerable  to  den- 
sity-independent mortality  factors.  These  factors  in- 
clude both  natural  events  and  human-caused  death 
and  disturbance.  A  few  thoughtless  or  misguided 
people  can  cause  severe  population  reductions 
either  by  direct  killing,  vandalizing  roosts,  or  disturb- 
ing the  bats  during  critical  periods.  Bats  disturbed 
during  hibernation  may  use  essential  energy  reserves 
during  the  arousal  process  and  those  disturbed  at 
maternity  sites  may  lose  contact  with  their  young. 

Second,  bats  have  a  very  low  reproductive  rate; 
the  females  of  most  species  only  produce  one  young 
per  year.  This  low  reproductive  potential  makes  re- 
covery from  sudden  increases  in  mortality  relatively 
slow.  Reduction  of  populations  of  North  American 
bats  during  recent  years  has  been  documented  for 
several  species  (Harvey  1976). 

Little  doubt  exists  that  bats  fill  important  niches 
and  the  impact  of  their  insectivorous  habits  in  tem- 
perate ecosystems  should  not  be  underestimated.  It 
is  important  to  acknowledge  this  vital  role  and  at- 
tempt to  inventory  and  monitor  bat  habitat  and  pop- 
ulations to  facilitate  their  effective  management  and 
conservation. 


516 


Bats 


LITERATURE  CITED 


ALTENBACH,  J.S.,  KN.  GELUSO,  and  D.E.  WILSON.  1979. 
Population  size  of  Tadarida  brasiliensis  at  Carlsbad 
Caverns  in  1973-  Pages  341-348  in  Genoways,  H.H. 
and  Baker,  R.J.  eds.  Biological  Investigation  in  the 
Guadalupe  Mountains  National  Park,  Texas.  Natl.  Park 
Serv.  Proc.  and  Trans.  Ser.  4. 

BAGLEY,  F.  and  J.  JACOBS.  1985.  Census  technique  for 
endangered  big-eared  bats  proving  successful.  Endang. 
Species  Tech.  Bull.  3:5-7. 

BARBOUR,  R.W.  and  W.H.  DAVIS.  1969.  Bats  of  America. 
University  Press  of  Kentucky,  Lexington.  286pp. 

BARCLAY,  R.M.R.  (In  press).  Marking  and  observational 
techniques,  in  Kunz,  T.H.  ed.  Behavioral  and  Ecologi- 
cal Methods  for  the  Study  of  Bats.  Smithsonian  Institu- 
tion Press.  Washington,  DC. 

BLEITZ,  D.  1984.  Mist  nets  and  their  use.  Bleitz  Wildlife 
Foundation,  5334  Hollywood  Blvd.,  Hollywood,  CA. 
18pp. 

BONACCORSO,  F.J.  and  N.  SMYTHE.  1972.  Punch-mark 
ing:  an  alternative  to  banding.  J.  Mammal.  53:389-390. 

, ,  and  S.R.  HUMPHREY.  1976.  Improved 

techniques  for  marking  bats.  J.  Mammal.  57:181-182. 

COCKRUM,  EL.  1969.  Migration  in  the  guano  bat,  Tadar- 
ida brasiliensis.  Pages  303-336  in  Jones,  J.K,  Jr.  ed. 
Contributions  in  Mammalogy.  Misc.  Publ.  51.  Univ. 
Kansas,  Lawrence 

CONSTANTINE,  D.G.  1958.  An  automatic  bat-collecting 
device.  J.  Wildl.  Manage.  22:17-22. 

CROSS,  S.P.  1965.  Roosting  habits  of  Pipistrellus  hesperus. 
J.  Mammal.  46:270-279. 

DAVIS,  R.B.,  C.F.  HERREID  II,  and  H.L.  SHORT.  1962.  Mex- 
ican free-tailed  bats  in  Texas.  Ecol.  Monogr.  32:311- 
346. 

DAVIS,  R.P.  and  EL.  COCKRUM.  1963.  Bridges  utilized  as 
day  roosts  by  bats.  J.  Mammal.  44:428-430. 

DOWLER,  R.C  and  H.H.  GENOWAYS.  1976.  Museology: 
supplies  and  suppliers  for  vertebrate  collections. 
Texas  Tech  Press.  Lubbock.  83pp 

FENTON,  MB.  1982.  Echolocation,  insect  hearing,  and 
feeding  ecology  of  insectivorous  bats.  Pages  261-28  in 
Kunz,  T.H.  ed.  Ecology  of  Bats.  Plenum  Publ.  Corp. 
New  York,  NY. 

.  (In  press).  Detecting,  recording,  and  analyzing  the 

vocalizations  of  bats,  in  Kunz,  T.H.  ed.  Behavioral  and 
Ecological  Methods  for  the  Study  of  Bats.  Smithsonian 
Institution  Press.  Washington,  DC. 

GAISLER,J.  1979.  Ecology  of  bats.  Pages  281-342  in  Stod- 
dard, D.M.  ed.  Ecology  of  Small  Mammals.  Chapman 
and  Hall,  London. 

GELUSO,  K.N,  J.S.  ALTENBACH,  and  D.E.  WILSON.  1976. 
Bat  mortality:  pesticide  poisoning  and  migratory 
stress.  Science  194:184-186. 

GREENHALL,  A.M.  1982.  House  bat  management.  U.S. 
Dep.  Inter.,  Fish  and  Wildl.  Serv.  Resour.  Publ.  143- 
53pp. 

and  J.L.  PARADISO.  1968.  Bats  and  bat  banding. 

U.S.  Dep.  Inter.,  Bur.  Sport  Fish,  and  Wildl.  72.  48pp. 

HALL,  E.R.  1981.  The  mammals  of  North  America.  2nd  Ed. 
John  Wiley  &  Sons.  New  York,  NY.  1 181pp. 

HARVEY,  M.J.  1976.  Endangered  Chiroptera  of  the  south- 
eastern United  States.  Proc.  Annu.  Conf.  Southeast.  As- 
soc. Game  Fish  Comm.  29:429-433- 


HAYWARD,  B.J.  and  S.P.  CROSS.  1979.  The  natural  history  of 
Pipistrellus  hesperus  ( Chiroptera:  Vespertilionidae  ).  Of- 
fice Res.  West.  New  Mexico  Univ.,  Silver  City.  3: 1-36. 

HENNY,  C.J.,  C.  MASER,  JO.  WHITAKER,  and  T.E.  KAISER. 
1982.  Organochlorine  residues  in  bats  after  forest 
spraying  with  DDT.  Northw.  Sci.  56:329-337. 

JONES,  J.K  Jr.,  DC.  CARTER,  and  H.H.  GENPWAUS.  1982. 
Revised  checklist  of  North  American  mammals  north 
of  Mexico.  Occ.  Pap.  62.  Mus.  Texas  Tech.  Univ.,  Lub- 
bock. 

KEYES,  BE.  and  C.E.  GRUE.  1982.  Capturing  birds  with 
mist  nets:  a  review.  N.  Am.  Bird  Bander.  6:1-14. 

KUNZ,  T.H.  1982.  Roosting  ecology.  Pages  1-46  in  Kunz, 
T.H.  ed.  Ecology  of  Bats.  Plenum  Publ.  Corp.  New 
York,  NY. 

,  ed.  (In  press).  Behavioral  and  ecological  methods 

for  the  study  of  bats.  Smithsonian  Institution  Press. 
Washington,  DC. 

and  C.E.  BROCK  1975.  A  comparison  of  mist  nets 


and  ultrasonic  detectors  for  monitoring  flight  activity 
of  bats.  J.  Mammal.  56:907-911. 

LAVAL,  R.K  (In  press).  Census  techniques,  in  Kunz,  T.H., 
ed.  Behavioral  and  Ecological  Methods  for  the  Study 
of  Bats.  Smithsonian  Institution  Press.  Washington,  DC. 

MASER,  C.J.E.  RODIEK  and  J.W.  THOMAS.  1979.  Cliffs, 
talus,  and  caves.  Pages  96-103  in  Thomas,  J.W.  tech. 
ed.  Wildlife  Habitats  in  Managed  Forests.  U.S.  Dep. 
Agric.  For.  Serv.  Agric.  Handbook  553- 

,  BR.  MATE,  J.F.  FRANKLIN,  and  C.T.  DYRNESS. 

1981.  Natural  history  of  Oregon  coast  mammals.  U.S. 
Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep.  PNW-133- 
496pp. 

MENAB,  B.K  1982.  Evolutionary  alternatives  in  the  physi- 
ological ecology  of  bats.  Pages  151-200  in  Kunz,  T.H. 
ed.  Ecology  of  Bats.  Plenum  Publ.  Corp.  New  York, 
NY. 

O'SHEA,  T.J.  and  T.A.  VAUGHAN.  1977.  Nocturnal  and  sea- 
sonal activities  of  the  pallid  bat,  Antrozous  pallidas.  J. 
Mammal.  58:269-284. 

PAIGE,  K.N.,  L.A.  MINK  and  V.R.  McDANIEL.  1985.  A 

broadband  ultrasonic  field  detector  for  monitoring  bat 
cries.  J.  Wildl.  Manage.  49:11-13. 

SIMMONS,  J.A.,  MB.  FENTON,  W.R.  FERGUSON,  M.  JUT- 
TING, and  J.  PALIN.  1979.  Apparatus  for  research  on 
animal  ultrasonic  signals.  Life  Sci.  Misc.  Publ.  Royal 
Ontario  Mus.  31pp. 

STEBBINGS,  R.E.  1978.  Marking  bats.  Pages  81-94  in 
Stonehouse,  R.  ed.  Animal  Marking.  Univ.  Park  Press, 
Baltimore,  MD. 

THOMAS,  J.W.  tech  ed.  1979.  Wildlife  habitats  in  managed 
forests:  the  Blue  Mountains  of  Oregon  and  Washing- 
ton. U.S.  Dep.  Agric,  For.  Serv.  Agric.  Handbook  553- 
512pp. 

TIDEMANN,  C.R.  and  DP.  WOODSIDE.  1978.  A  collapsi- 
ble bat-trap  and  comparison  of  results  obtained  with 
the  trap  and  with  mist  nets.  Australian  Wildl.  Res. 
5:355-362. 

TUTTLE,  M.D.  1974.  An  improved  trap  for  bats.  J.  Mam- 
mal. 55:475-477. 

VAUGHAN,  T.A.  1959.  Functional  morphology  of  three 
bats:  Eumops,  Myotis,  Macrotus.  Publ.  Mus.  Nat.  Hist. 
Univ.  Kansas,  Lawrence.  12:1-153 

WARDEN,  T.  1980.  Notes  on  the  determination  of  bat 
populations  using  photographic  measurements.  Nat. 
Speol.  Soc  Bulletin.  42:70-71. 


Bats 


517 


25 


Ungulates 


Raymond  J.  Boyd  and  Allen  Y.  Cooperrider 


U.S.  Bureau  of  Land  Management 
Service  Center 
Denver,  CO  80225 


Peter  C.  Lent 

U.S.  Bureau  of  Land  Management 
920  Valley  Road 
Reno,  NV  89512 


James  A.  Bailey 

Colorado  State  University 
Fort  Collins,  CO  80523 


Editor's  Note:  Ungulates  occupy  a  great  diversity  of 
habitats  in  North  America  and  have  life-history 
strategies  uniquely  adapted  to  these  habitats.  Yet, 
many  inventory  and  monitoring  techniques  are 
similar  for  each  of  the  species.  Ungulates  are  ex- 
tremely important  because  of  their  recreational 
hunting  and  viewing  values;  they  also  can  compete 
with  livestock  for  forage  and  cause  damage  to  var- 
ious agricultural  crops. 

Many  common  ungulates  in  North  America,  such 
as  elk  and  antelope,  were  once  quite  rare  and  extir- 
pated from  many  parts  of  their  historical  range. 
This  suggests  that  given  adequate  public  support, 
cooperation,  and  funding  ungulates  can  be  main- 
tained or  increased  in  natural  habitats  with  ade- 
quate food,  cover,  and  water.  Intensive  efforts  in 
studying  inventorying  and  monitoring  such  habi- 
tat factors  have  been  largely  responsible  for  the 
successful  increase  in  North  American  ungulate 
numbers.  These  efforts  have  been  detailed  in  much 
literature  over  the  past  50  years.  This  chapter,  there- 
fore, only  introduces  and  provides  an  overview  of 
this  knowledge  base.  A  biologist  working  with  un- 
gulates will  need  to  review  the  sources  listed  in  the 
chapter  for  more  detailed  information. 


"All  animals  are  created  equal,  but  some  are  created 
more  equal  than  others." 


INTRODUCTION 


-George  Orwell,  Animal  Farm 


Wild  ungulates  are  among  the  most  prized  wild- 
life groups,  and  management  of  their  habitat  is  an 
important  task  of  wildlife  biologists.  Most  areas  of 
North  America  contain  habitat  or  potential  habitat 
for  one  or  more  ungulate  species.  These  animals  are 
highly  valued  for  aesthetic  reasons,  and  public  hunt- 
ing of  wild  ungulates  is  an  important  cultural  heri- 
tage. Recreational  hunting  is  an  important  source  of 
revenue  for  local  economies  and  for  manufacturers 
of  hunting-related  equipment.  License  sales  for  these 
animals  provide  a  primary  source  of  funding  for  state 
wildlife  agencies. 

On  the  other  hand,  wild  ungulates  can  cause 
obvious,  short-term  habitat  alterations  through  graz- 
ing and  browsing.  They  can  damage  agricultural 
crops,  directly  compete  for  forage  with  livestock,  as 
well  as  degrade  their  own  habitat. 

Inventory  and  monitoring  of  habitat  for  wild 
ungulates  is  thus  one  of  the  most  important  tasks  of 
wildlife  biologists;  it  is  also  difficult,  challenging,  and 
interesting  work.  Although  biologists  have  studied 
and  attempted  to  manage  wild  ungulate  habitat  more 
intensively  than  most  other  species  habitats,  our 
understanding  and  ability  to  manage  it  is  still 
primitive. 

In  particular,  biologists  often  have  difficulty 
predicting  how  ungulate  populations  will  respond  to 
habitat  changes.  Because  ungulates  are  adaptable,  a 


Ungulates 


519 


Haystack  damage  by  elk. 


species  such  as  white-tailed  deer  (Odocoileus  virgi- 
nianus)  may  live  in  habitats  ranging  from  tropical 
Florida  swamps  to  hot  Arizona  deserts,  or  northern 
coniferous  forests  of  British  Columbia.  Furthermore, 
since  the  Pleistocene  epoch,  the  number  of  ungulate 
species  in  North  America  has  been  quite  limited, 
consisting  of  "generalists"  (as  opposed  to  some  of 
the  highly  specialized  forms  found  in  Africa).  In 
adapting  to  different  habitats,  ungulate  populations 
acquire  distinct  patterns  of  forage  selection,  habitat 
use,  and  population  dynamics.  Therefore,  generaliz- 
ing about  the  species  from  study  of  an  individual 
population  is  not  only  risky,  but  may  increase  or  per- 
petuate misinformation. 

Ungulates  are  highly  intelligent  animals.  For 
example,  choice  of  migration  routes  for  species  such 
as  bighorn  sheep  results  not  from  random  selection 
among  appropriate  alternative  routes,  but  rather 
selection  learned  from  travel  with  other  individuals 
during  previous  migrations.  Such  learned  behaviors 
greatly  affect  how  the  animals  use  habitat  and  re- 
spond to  habitat  changes.  Such  use  and  response  to 
management  cannot  always  be  determined  by  look- 
ing at  the  habitat. 

Because  ungulates  are  relatively  long-lived,  re- 
sponses to  habitat  changes  may  not  be  apparent  for 
many  years.  Such  time  lags  make  assessing  the  effects 
of  management  difficult  unless  careful,  well-planned 
records  are  kept  for  many  years. 

Finally,  habitat  suitability  often  is  not  the  sole 
factor  limiting  ungulate  populations.  Predation,  dis- 
ease, or  hunting  may  limit  a  population.  Each  of 
these  factors  can,  of  course,  be  related  to  habitat 
conditions  or  interact  with  habitat  factors.  However, 
there  are  well-documented  cases  where  changes  in 
predation,  disease,  or  hunting,  in  the  absence  of  any 
change  in  habitat,  have  caused  significant  changes 
in  ungulate  populations. 


Because  ungulates  have  been  studied  so  exten- 
sively, this  chapter  only  provides  a  broad  overview. 
It  is  intended  for  biologists  having  limited  experi- 
ence with  ungulates  and  for  those  wanting  guidance 
on  finding  additional  information  in  the  literature. 
In  the  following  sections  we  describe  general  habitat 
requirements  of  ungulates  and  methods  for  measur- 
ing population  variables.  We  then  describe  specific 
requirements  and  techniques  for  individual  North 
American  species.  These  accounts  are  necessarily 
brief,  but  should  assist  the  reader  in  finding  more  de- 
tailed information  in  the  literature. 


HABITAT  REQUIREMENTS 

Habitat  requirements  of  ungulates  are  analyzed 
in  terms  of  their  needs  for  food,  water,  cover,  and 
reproduction  as  well  as  the  way  they  must  be  ar- 
ranged in  space  and  time. 

Food 

Forage  supply  often  limits  ungulate  populations, 
either  by  itself  or  in  conjunction  with  some  other 
factor  such  as  cover.  Food  is  almost  always  a  qualita- 
tive rather  than  a  quantitative  problem  with  ungu- 
lates. By  this  we  mean  that  the  quality  of  available 
food  rather  than  the  total  amount  of  food  limits  a 
population.  Ungulates  have  died  of  malnutrition  with 
their  stomachs  full  of  food.  Thus,  the  habitat  biolo- 
gist must  consider  not  only  the  amount  of  food  avail- 
able, but  its  nutritional  content.  Furthermore,  forage 
availability  and  nutritional  quality  of  available  plants 
vary  seasonally  due  to  phenology  of  plants,  utiliza- 
tion by  other  wild  and  domestic  animals,  and 
weather  conditions  that  may  make  plants  unavailable 
(e.g.,  deep  snow)  or  alter  phenology  or  nutritional 
quality  (Figures  1  and  2).  Thus,  analyzing  the  ade- 
quacy of  the  forage  supply  for  an  ungulate  is  com- 
plex. Biologists  have  tried  to  simplify  the  problem  by 
identifying  measurable  limiting  factors  among  the 
vast  number  of  forage-related  factors.  This  approach 
has  been  quite  successful  in  local  areas,  but  unsuc- 
cessful when  knowledge  of  limiting  factors  from  one 
region  has  been  blindly  used  as  the  basis  for  evaluat- 
ing forage  supply  in  another.  As  background  for  un- 
derstanding methods  of  analyzing  forage  supply  for 
ungulates,  we  describe  the  annual  forage  cycle,  pat- 
terns of  forage  preference  and  food  habits,  potential 
limiting  factors,  and  mineral  licks. 

Annual  Forage  Cycle.  Since  ungulates  may  eat 
many  plant  species  in  different  regions  and  different 
seasons,  biologists  need  to  classify  forage  species 
according  to  similar  nutritional  and/or  phenological 
characteristics.  Wildlife  biologists  and  range  man- 
agers have  found  it  convenient  to  classify  forage  as 
grasses,  forbs,  and  browse  (Table  1 ).  This  classifica- 
tion is  based  primarily  on  structure  and  phenology 
of  the  plants,  but  also  correlates  well  with  seasonal 


520 


Ungulates 


840- 
5   800- 

a. 
O 
g  600- 

o 

z 

o 

z 

<    400- 

co 

CO 

< 

d   200- 
< 
> 
< 

/ 

TOTAL  ANNUAL  PRODUCTION 

NUTRITION  QUALITY  (K   cal   DE/g) 

O                                         O  (S3                                  O 

/                              N<                                               NUTRITIONAL  VALUES 

VSE 

1 

1   FORBS 

2  GRASS 
&  BRO\ 

/                                          N.                                         NECESSARY  FOR 
/                                                    X.                                 MAINTENANCE 

Mm 

Ji 

v         M         J          JASONDJ           F 

M         A 

\        M         J          J           ASODXJ           FMA 

Figure  1.     Typical  annual  cycle  of  available  standing 
crop  of  forage  (from  Cooperrider  and  Bailey 
1984). 

Table  1.     Classification  of  ungulate  forages. 


Figure  2.     Typical  annual  cycle  of  average  nutri- 
tional value  of  forage  on  temperate  ranges 
(modified  from  Cooperrider  and  Bailey  1984). 


Class 

Description 

Taxonomy 

Notes 

Grass  and  grass-like 
plants 

True  grass  and  grass- 
like sedges  and  rushes 

Family  Gramineae  and 
grass-like  Cyperaceae 
and  Juncaceae 

This  category  often  is 
abbreviated  to  grass, 
but  still  includes  sedges 
and  rushes. 

Forbs 

All  herbaceous  plants 
other  than  grass  and 
grass-like  plants 

Most  monocots  other 
than  three  families  listed 
above;  herbaceous 
dicots 

Since  classification  of 
forbs  and  browse  is  not 
based  on  taxonomy, 
classification  of  some 
semi-woody  species  is 
somewhat  arbitrary. 

Browse 

All  woody  plants 

Woody  dicots 

nutritional  qualities,  making  it  a  basis  for  describing 
such  an  annual  cycle.  The  nutritional  quality  of  a 
given  plant  species  varies  with  parts  of  a  plant,  indi- 
vidual plants,  ecotypes,  subspecies,  regions,  and 
physical  characteristics  of  the  site. 

A  general  picture  of  the  forage  cycle  can  be 
provided  by  considering  a  typical  pattern  of  produc- 
tion, availability,  and  forage  quality  of  grasses,  forbs, 
and  browse  in  temperate  and  arctic  areas.  Forbs 
are  mostly  available  for  a  short  time  in  the  spring 
and  summer  (Figure  1);  during  this  time  they  are 
generally  highly  nutritious  as  measured  by  digestible 
energy  content  or  similar  measures  (Figure  2). 
Grasses,  on  the  other  hand,  are  available  year-round 
but  are  most  nutritious  during  the  spring  and  early 
summer  when  they  are  green.  They  also  often  begin 
growing  earlier  than  any  other  forages  (Figures  1 
and  3).  Dry  dormant  grass,  on  the  other  hand,  has 
limited  nutritional  value.  Browse,  by  contrast,  is 
available  year-round  with  relatively  similar  nutri- 
tional value  among  seasons  (Figure  3)  Significant 
variations  may  occur  because  of  species  differences, 


annual  versus  perennial  grasses,  local  weather  pat- 
terns, deciduous  versus  evergreen  browse  species, 
etc.  However,  understanding  this  basic  model  along 
with  local  variations  should  help  in  understanding 
food  habits  and  forage  preferences  of  a  specific  ungu- 
late herd  and  thus  in  identifying  potential  limiting 
factors  related  to  forage. 

Forage  Preference  and  Food  Habits.  All  North 
American  ungulates  are  generalist  herbivores,  choos- 
ing opportunistically  among  hundreds  of  local  plant 
species.  A  typical  ungulate  will  eat  over  100  plant 
species  during  the  course  of  a  year.  However,  due  to 
availability  and  preference,  less  than  a  dozen  plants 
usually  constitute  over  75%  of  the  animal's  seasonal 
diet.  These  are  the  species  that  should  be  the  initial 
concern  of  management. 

Forage  cycles  can  be  used  to  help  interpret  food 
habits  of  the  various  ungulate  species.  Beginning  in 
spring,  ungulates  will  turn  to  plants  that  are  actively 
growing.  For  instance,  the  food  habits  of  pronghorn 


Ungulates 


521 


3.0-] 

90- 

vs 

___--■" 

_, 

^  FORBS 

\ 

S*~ 

O) 

80- 

LU 

/                 **  ""  " 

["  \ 

h- 

\ 

/ 

Q 

/              • 

V      N 

UJ 

o 

*:    2.0- 

> 
i- 

/             /      ,'- 

T         N  ~~  --  -. 

Q 

Z 
111 

o 
< 

DC 

70- 
60- 
50- 

\      BROWSE 

y 

/ 
/ 

/ 

/ 

/  -'; 

7'"  i 

1     i 

i 

X 

BROWSE 



< 

i 

GRASS 

o 

"■-- 

——■* 

O 

/ 

LL 

o 

/ 

LL 

40- 

_l 

/ 

o 

<    1.0- 

1 

H 

z 

/ 

Z 

30- 

o 

/ 

O 

V       FORBS 

cr 

// 

LU 

20- 

O 

// 

Q. 

/  GRASS 

z 

r  / 

10- 

0 

s       / 

■<„ 

>               i               i 

— i i 1^   i 

A         S        0         N 

D        J          F 

A 

A          M          J           J 

■           i           i 

ii            ii            ii 

A         M 

J          J          A 

S        O         N         D         J          F 

M 

A 

Figure  3.     Typical  annual  cycle  of  nutritional  qual- 
ity of  grass,  forbs,  and  browse. 


shown  in  Figure  4  can  be,  to  a  large  extent,  ex- 
plained by  the  pattern  of  forage  production  availabil- 
ity and  nutritional  value  shown  in  Figures  1  and  3. 
Typically,  the  actively  growing  plant  species  are  first 
grasses,  then  forbs.  As  the  season  progresses  and 
forbs  become  less  nutritious  or  available,  ungulates 
turn  more  to  browse  or  grass,  depending  on  the 
species  and  its  habitat  and  forage  preference.  During 
winter,  most  ungulates  depend  on  dry  grass  or 
browse  to  maintain  themselves.  Ungulates  reduce 
their  forage  intake  during  this  period  and  typically 
lose  weight  over  winter. 

Although  North  American  ungulates  are  oppor- 
tunistic, significant  differences  in  forage  preferences 
are  presumably  a  reflection  of  their  ability  to  gather 
and  digest  different  forage  classes.  Most  notable  of 
these  differences  is  the  distinction  between  grass 
eaters  and  browse  eaters.  All  North  American  ungu- 
lates eat  grass  or  browse  when  it  is  green,  succulent, 
and  nutritious.  However,  when  grass  or  browse  is 
dry  or  dormant,  some  ungulates  prefer  one  over  the 
other.  In  general,  grass  has  poorer  nutritional  quality 
at  this  time,  but  is  usually  available  in  greater  con- 
centrations, favoring  animals  that  can  gather  greater 
quantities  in  a  short  period  of  time.  Browse,  on  the 
other  hand,  is  usually  more  nutritious,  but  an  animal 
needs  to  be  more  selective  and  therefore  take  more 
time  to  gather  the  same  quantity.  In  general,  the 
smaller  ungulates  (pronghorn  antelope  [Antilocapra 
americana],  mule  [Odocoileus  hemionns]  and 
white-tailed  deer  [O.  virginianus\)  tend  toward 
browse  at  this  time,  while  the  larger  species  (bison 
[Bison  bison]  and  muskox  [Ovibos  moschatus]) 
prefer  grasses. 

Habitat  preferences  and  evolutionary  history 
also  play  a  role.  The  North  American  bovids  (big- 
horn [Ovis  canadensis]  and  Dall's  sheep  [O.  dalli], 


Figure  4.  Food  habits  of  pronghorn  antelope  on 
Trickle  Mountain,  Colorado  (from  Bailey  and 
Cooperrider  1982). 

muskox,  bison,  and  mountain  goat  [Oreamnos  amer- 
icana]) are  generally  animals  of  open  country  (grass- 
lands) and  are  capable  of  surviving  on  dry  grass  in 
winter.  On  the  other  hand,  North  American  cervids 
(mule  and  white-tailed  deer,  elk  [Cervus  elaphus], 
moose  [Alces  alces])  tend  to  be  forest  or  forest-edge 
animals  or  have  evolved  from  forest-dwelling  ances- 
tors; all  can  survive  on  dormant  browse.  Caribou 
(Rangifer  tarandus),  an  apparent  exception,  survive 
on  lichens  during  winter  over  much  of  their  range. 
Caribou  probably  evolved  as  forest  animals,  eating 
arboreal  lichens  but  some  learned  to  move  out  into 
the  tundra  to  eat  lichens  on  the  ground.  Some  North 
American  ungulates  seem  capable  of  surviving  either 
way,  most  notably  elk;  although  less  well-studied, 
mountain  goats  appear  to  have  the  same  capability. 

Food  habits  are  foods  that  an  animal  actually 
eats,  based  on  forage  availability  and  preference. 
Measurement  or  observation  of  food  habits  can  be  an 
important  tool  toward  understanding  limiting  factors 
related  to  forage  quality  and  quantity.  If  observed 
food  habits  differ  significantly  from  typical  or  ex- 
pected food  habits,  it  may  suggest  a  problem  or  lim- 
iting factor.  For  example,  if  deer  are  eating 
significant  quantities  of  dry  grass  during  winter  (an 
atypical  pattern),  it  could  suggest  that  their  winter 
range  lacks  suitable  browse  species.  This  could  be 
confirmed  by  observing  or  measuring  plant  species 
composition  on  that  range.  Similarly,  if  antelope  are 
eating  small  quantities  of  forbs  in  spring,  it  could 
suggest  that  the  range  is  lacking  good  quality  forbs. 

Food  habits  must  always  be  interpreted  in  terms 
of  local  conditions  and  the  ungulate  species  under 
consideration.  For  example,  the  food  habit  pattern 
for  mule  deer  shown  in  Figure  5  would  appear,  at 
first  glance,  to  be  atypical  of  the  species  as  a  whole. 
However,  it  can  easily  be  explained  by  recognizing 


522 


Ungulates 


that  these  deer  live  in  California  in  a  Mediterranean 
climate  where  winters  are  mild  and  all  precipitation 
comes  in  fall,  winter,  and  spring.  Therefore,  grass 
is  green  and  nutritious  during  winter  but  dry  and  un- 
palatable during  summer. 

Analysis  of  Limiting  Forage  Factors.  Several 
methods  have  been  developed  to  analyze  limiting 
factors  related  to  forage  supply.  In  order  of  increas- 
ing complexity,  these  are  the — 

( 1 )  Foraging  area  approach, 

(2)  Key  species  approach, 

(3)  Carrying-capacity  approach,  and 

(4)  Nutritional  approach. 


Items  2  and  3  have  been  used  in  one  form  or 
another  for  over  50  years  in  North  America.  The 
foraging  area  approach,  however,  has  only  been  for- 
malized in  recent  years  in  response  to  land  manage- 
ment needs.  Similarly,  the  nutritional  approach  has 
only  recently  been  developed  and  formalized,  al- 
though earlier  attempts  were  made  to  incorporate 
nutritional  considerations  into  analysis  methods. 


Shrubland  Diet 


GRASS   FORBS 


SHRUBS 


26% 

19% 

55% 

Figure  5.     Food  habits  of  mule  deer  in  California 
chaparral  (from  Taber  and  Dasmann  1958). 


Foraging  Area  Approach.  The  foraging  area 
approach  is  a  landscape-based  analysis  system.  With 
this  approach,  the  biologist  identifies  the  aspects 
of  topography,  vegetation  structure,  and  other  physi- 
cal features  required  to  provide  suitable  foraging 
areas  for  ungulate  species.  For  example,  suitable  win- 
ter foraging  areas  for  bighorn  sheep  in  northern  Col- 
orado might  be  considered  to  be  south-facing,  grass- 
covered  slopes  below  2,400  m  (  7,920  ft )  elevation, 
within  400  m  (1,320  ft)  of  suitable  rocky  escape 
cover.  The  biologist  then  analyzes  the  range  to  de- 
termine if  it  contains  an  adequate  acreage  or  distri- 
bution of  foraging  areas.  This  analysis  does  not  result 
in  a  quantitative  estimation  of  carrying  capacity. 
The  foraging  area  approach  was  developed  for  elk 
and  mule  deer  in  the  Blue  Mountains  of  eastern  Ore- 
gon by  Black  et  al.  (  1976)  and  Thomas  et  al.  (  1979). 
The  technique  has  the  advantage  of  requiring  only 
limited  data  collection,  with  data  frequently  obtained 
from  existing  sources  such  as  aerial  photographs, 
topographic  maps,  etc.  The  technique  is  limited  in 
that  forage  quality  is  not  considered  in  terms  of 
plant  species  composition,  range  condition,  total 
forage  weight,  browse  species  condition,  or  nutri- 
tional value.  However,  it  is  a  good  way  to  begin  an 
analysis  of  ungulate  range  and  may  serve  as  the  basis 
for  more  detailed  analyses  if  these  are  required  later. 

Key  Species  Approach.  The  key  species  ap- 
proach is  based  on  the  assumption  that  a  biologist 
can  identify  the  critical  (limiting)  range  of  the  ani- 
mal and  also  the  critical  or  key  forage  species  on 
that  range.  It  was  developed  primarily  for  mule  and 
white-tailed  deer  in  northern  areas  of  the  U.S.  and 
Canada.  In  winter,  these  deer  typically  concentrate 
on  traditional  winter  ranges  and  subsist  on  one  or 
more  species  of  palatable  browse.  The  browse  spe- 
cies on  these  ranges  are  assumed  to  be  the  factors 
limiting  these  herds.  By  concentrating  on  key 
browse  species  on  key  areas,  the  biologist  can  sim- 
plify a  complex  situation.  The  condition,  trend,  and 
production  of  the  key  species  can  be  measured  and 
monitored;  if  total  production  is  estimated,  it  can 
lead  to  a  carrying-capacity  determination. 

The  utility  of  this  technique  depends  entirely  on 
two  factors:  ( 1 )  the  degree  to  which  ungulates  de- 
pend on  key  species  and  range  and  ( 2 )  the  biolo- 
gists' ability  to  correctly  identify  these  key  species 
and  ranges.  Recent  research  has  demonstrated  the 
importance  of  all  seasonal  ranges  in  determining  the 
health  of  deer  and  other  ungulate  herds  (Mautz 
1978).  For  many  years,  biologists  concentrated  on 
winter  ranges  because  this  was  when  deer  and  other 
ungulates  lost  weight;  furthermore,  losses  from  star- 
vation typically  were  observed  during  late  winter 
and  early  spring  while  deer  were  still  on  winter 
ranges.  However,  recent  field  studies  and  nutritional 
research  have  demonstrated  that  the  condition  of 
ungulates  before  winter  may  be  as  important  or 
more  important  than  the  condition  of  the  winter 


Ungulates 


523 


food  supply  in  determining  overwinter  survival. 
Therefore,  biologists  have  begun  to  pay  more  atten- 
tion to  spring,  summer,  and  fall  range  conditions. 

The  key  species  approach  has  provided  and  will 
continue  to  provide  a  useful  way  of  analyzing  ungu- 
late forage  supplies.  However,  blind  adherence  to 
this  approach  has  caused  much  misinformation  and 
inefficient  use  of  limited  habitat  management  funds. 
In  extreme  cases,  dogmatic  belief  in  key  forages  as 
limiting  factors  has  caused  biologists  to  overlook 
limiting  factors  such  as  predation,  hunting,  or  dis- 
ease. Biologists  should  use  the  key  species  approach 
with  caution. 

Carrying-Capacity  Approach.  When  discuss- 
ing carrying  capacity,  biologists  tend  to  bog  down  in 
semantics;  for  the  purpose  of  this  discussion,  carry- 
ing capacity  is  the  number  of  animals  that  can  be 
supported  on  a  range  with  a  given  amount  of  forage 
production  without  adversely  affecting  the  long-term 
productivity  of  the  forage  supply.  For  a  more  de- 
tailed discussion  of  the  concept,  its  use  and  misuse, 
see  Bailey  (1984:280-288)  or  Caughley  ( 1 979 ). 

To  use  the  carrying-capacity  approach,  a  biolo- 
gist must  measure  or  estimate  the  total  amount  of 
forage  available  on  a  year-round  or  seasonal  range, 
determine  the  amount  of  forage  that  can  be  used 
without  damaging  future  productivity,  and  then  cal- 
culate the  number  of  animals  this  amount  of  usable 
forage  can  sustain.  Although  seemingly  simple,  the 
approach  is  difficult  at  each  stage.  Measuring  total 
forage  is  expensive  and  complex  in  itself.  Further- 
more, determining  the  amount  of  forage  that  can  be 
used  without  damage  is  more  of  an  art  than  a  sci- 
ence (Skiles  et  al.  1980).  Finally,  calculating  the 
number  of  animals  that  can  be  sustained  requires 
making  assumptions  or  estimates  on  the  quality  or 
condition  of  the  animals,  sex  and  age  ratios,  etc. 
Furthermore,  it  may  require  information  on  intake 
rates  or  food  habits  that  are  not  available  or  can  only 
be  obtained  at  great  expense. 

To  alleviate  these  problems,  biologists  tend  to 
make  simplifying  assumptions  and  use  "rules  of 
thumb"  in  making  the  calculations.  For  example,  the 
50%  or  "graze  half/leave  half"  rule  developed  by 
range  managers  is  commonly  used  to  determine  the 
percentage  of  total  forage  that  can  be  consumed. 
Similarly,  many  agencies  have  developed  tables  of 
forage  intake  rates  for  both  domestic  and  wild 
ungulates. 

Calculating  carrying  capacity  is  thus  a  very 
crude  measure  of  the  capability  of  the  habitat  to 
support  ungulates  on  a  sustained  basis.  Nonetheless, 
it  has  been  useful  and  will  continue  to  be  used.  It 
is  especially  useful  where  more  than  one  wild  or 
domestic  ungulate  is  found  on  the  same  range.  In 
particular,  when  livestock  are  grazing  areas  used  by 


Winter-killed  mule  deer  under  overbrowsed  juniper. 

wild  ungulates,  land  managers  must  determine  the 
amount  of  forage  to  be  allocated  for  livestock  and 
the  amount  to  be  reserved  for  wild  animals.  Forage 
allocation  requires  a  carrying-capacity  approach,  and 
several  systems  have  been  developed  for  formalizing 
these  calculations  ( MacPherson  et  al.  1982;  Cooper- 
rider  and  Bailey  1984;  Nelson  1984). 

Calculating  carrying  capacity,  therefore,  requires 
a  great  deal  of  information  which  may  be  costly  to 
obtain,  and  the  final  measure  may  only  be  a  very 
crude  determination  of  the  ability  of  the  habitat  to 
provide  forage  for  the  animals.  Furthermore,  most 
carrying-capacity  approaches  do  not  consider  infor- 
mation on  the  nutritional  requirements  of  animals 
nor  the  nutritional  content  of  plants. 

Nutritional  Approach.  The  nutritional  ap- 
proach attempts  to  consider  both  the  quality  and 
quantity  of  forage  in  a  carrying-capacity  model. 
When  considering  the  nutritional  value  of  forage, 
many  factors  can  be  measured.  For  many  years,  wild- 
life biologists  used  proximate  analyses  (Table  1 )  to 
evaluate  forages.  These  analyses  were  developed 
by  livestock  nutritionists  and  proved  to  have  some 
value  in  evaluating  wildlife  forages.  Crude  protein 
was  found  to  be  especially  useful  since  it  was  easy 
and  cheap  to  measure,  and  levels  of  crude  protein  in 
most  plants  seemed  to  correlate  with  their  palatabil- 
ity  to  wild  ungulates.  There  were  some  notable  ex- 
ceptions, however,  including  arid  land  browse 
species  plants  such  as  big  sagebrush  {Artemisia  tri- 
dentata)  and  conifers  such  as  Douglas  fir  (Pseudot- 
suga  menziesii).  These  types  of  plants  typically 
contained  high  levels  of  crude  protein  but  were  fre- 
quently far  less  palatable  or  digestible  than  other 
plants  with  similar  levels  of  crude  protein.  Under- 
standing wildlife  nutrition  was  greatly  enhanced  by 
the  discovery  of  essential  oils  in  plants,  which  inhib- 


524 


Ungulates 


ited  ruminant  digestion  and  made  the  plants  unpalat- 
able (Nagy  et  al.  1964;  Longhurst  et  al.  1968). 
Although  some  plants  can  have  relatively  high  levels 
of  certain  key  nutrients,  these  nutrients  are  unavail- 
able to  the  animal  because  of  the  presence  of  other 
non-nutritive  substances. 

This  discovery  and  other  developments  stimu- 
lated the  search  for  new  measures  of  nutritional 
quality.  Two  such  measures  are  commonly  used 
now:  digestible  protein  (DP)  and  digestible  energy 
(DE).  Both  effectively  measure  the  overall  quality  of 
a  forage  for  ungulates,  although  they  can  be  expen- 
sive to  obtain.  Furthermore,  these  measures  tend 
to  be  highly  correlated  with  each  other  and  with 
certain  other  measurements  such  as  concentration  of 
certain  vitamins.  Since  the  basic  requirements  for 
maintenance,  reproduction,  lactation,  etc.  have  been 
developed  for  animals  in  terms  of  digestible  energy, 
animal  requirements  can  be  related  directly  to 
intake. 

Nutritional  approaches  to  carrying  capacity  re- 
quire biologists  to  limit  analyses  to  a  very  few  possi- 
ble nutrients  or  nutritional  measurement.  Everything 
from  trace  minerals  to  vitamins  has  been  suggested 
as  factors  limiting  ungulate  populations,  and  such 
substances  may  be  important  in  local  situations. 
However,  the  biologist  is  well-advised  to  begin  a 
nutritional  investigation  or  analysis  by  measuring  the 
more  obvious  and  basic  parameters  such  as  crude 
protein,  digestible  protein,  or  digestible  energy. 

Systems  for  integrating  nutritional  measurements 
such  as  DP  and  DE  into  carrying-capacity  determina- 
tions have  been  developed  and  described  by  Mautz 
(1978),  Robbins  (1973),  Wallmo  et  al.  (1977), 
Hobbs  et  al.  (1982),  Cooperrider  and  Bailey  (1984), 
and  others. 

Nutritional  approaches  to  determining  carrying 
capacity  require  much  sophisticated  information 
on  nutritional  requirements  of  animals  and  nutrient 
content  of  forages.  In  addition,  much  data  are  re- 
quired for  traditional  carrying-capacity  models.  Se- 
lecting appropriate  measurement  s)  of  nutritional 
quality  is  critical  both  in  terms  of  costs  and  predic- 
tive power;  nutritional  approaches  should  only  be 
used  when  ungulate  populations  and  management  is- 
sues warrant  the  high  cost  in  money  and  personnel. 
Not  surprisingly,  we  do  not  know  of  any  situations 
where  a  nutritional  approach  is  being  used  in  an 
operational  mode  by  land-  or  wildlife-management 
agencies. 

Mineral  Licks.  On  some  ranges,  ungulates  are  at- 
tracted seasonally  to  mineral  licks.  While  these  areas 
may  certainly  affect  animal  distribution,  the  effect 
on  herd  productivity  has  not  yet  been  demonstrated. 
Mountain  sheep,  mountain  goats,  and  moose  may 
be  an  exception  as  discussed  later  in  this  chapter,  al- 
though the  effects  have  only  been  hypothesized. 


Biologists  once  provided  salt  on  ungulate  ranges  as 
livestock  operators  do  on  cattle  ranges.  This  prac- 
tice, however,  does  not  seem  to  affect  animal  pro- 
ductivity, survival,  or  abundance. 

Water 


Free  Water.  Although  some  desert  species  such  as 
bighorn  sheep  can  survive  for  extended  periods 
without  free  water  or  with  free  water  in  plants,  all 
North  American  ungulates  need  to  have  water 
sources  in  their  habitat  for  survival.  The  biologist 
must  determine  the  spacing  or  density  of  water 
sources:  what  is  minimal  for  a  given  species  and 
what  is  optimal.  In  most  northern  areas  of  the  conti- 
nent, water  is  abundant  and  relatively  well-distrib- 
uted. As  a  result,  lack  of  water  as  a  limiting  factor 
has  rarely  been  suggested  or  studied.  In  the  south- 
western desert  areas,  however,  lack  of  water  has 
been  identified  as  an  important  factor  limiting  bigh- 
orn sheep  (Turner  and  Weaver  1980),  pronghorn 
antelope  (Yoakum  1980),  and  mule  deer  (Wood  et 
al.  1970).  For  desert  bighorn,  minimal  spacing  of 
water  sources  has  been  estimated  at  10  km  (6  mi.) 
and  optimal  spacing  as  2  km  (1.2  mi.)  (Turner  and 
Weaver  1980).  Optimal  spacing  of  water  has  been 
estimated  at  4  to  5  km  (2.5  to  3  mi.)  for  mule  deer 
(Wood  et  al.  1970)  and  5  to  7  km  (3  to  4  mi.)  for 
pronghorn  (Yoakum  1978).  Without  further  study,  a 
biologist  may  conclude  that  distribution  of  free- 
water  sources  (streams,  lakes,  or  perennial  springs) 
every  2  to  5  km  (1  to  3  mi.)  is  an  optimal  free-water 
supply  for  ungulates  in  North  America. 

Water  Quality.  Water  quality  requirements  for  wild 
ungulates  have  seldom  been  investigated.  However, 
biologists  have  documented  that  waters  high  in  total 
dissolved  solids  (TDS)  or  with  a  high  pH  will  not 
be  used  by  some  species.  Sundstrom  ( 1 968 )  noted 
that  pronghorn  antelope  avoid  waters  with  TDS  in 
excess  of  5,000  mg/1  or  pH  higher  than  9.2.  In  gen- 
eral, biologists  need  not  worry  about  water  quality 
for  ungulates  unless  there  is  evidence  that  some 
waters  are  being  avoided.  In  these  cases,  water  qual- 
ity measurements  should  be  considered,  particularly 
if  water  quality  can  potentially  be  improved  through 
management  actions. 

Snow.  The  presence  or  absence  of  snow  packs  is 
such  an  obvious  determinant  of  habitat  suitability  for 
ungulates  in  the  northern  areas  of  this  continent 
that  it  is  frequently  omitted  from  elementary  discus- 
sions on  the  subject.  However,  more  than  one  nov- 
ice biologist  has  seriously  misjudged  habitat 
suitability  by  examining  a  range  during  summer 
without  understanding  conditions  of  the  range  in 
midwinter.  More  seriously,  some  agencies  have  for- 
malized procedures  that  calculate  earning  capacities 
based  on  forage  production  without  considering 
snow  depths  on  the  range.  By  using  such  systems, 


Ungulates 


525 


forage  may  be  considered  available  to  wild  ungulates 
even  though  it  is  under  10  feet  of  snow  during  the 
time  of  year  they  need  it. 

On  the  basis  of  much  observation  and  measure- 
ments, snow  depth  has  two  important  effects  on 
ungulates:  it  covers  up  and  eventually  makes  ground 
forage  unavailable  or  unlocatable,  and  it  reduces 
mobility.  However,  these  effects  are  quite  species- 
specific  and,  even  then,  conventional  wisdom  or 
"rules  of  thumb"  may  be  misleading. 

Most  ungulates,  even  those  that  are  not  particu- 
larly adapted  to  living  in  snowy  areas,  will  paw 
through  a  light  snow  cover  of  up  to  10  cm  (3  in. ) 
and  forage  without  difficulty.  Snow  covers  over  10 
cm  (3  in.)  may  inhibit  foraging  on  the  ground  for 
some  species,  although  others  such  as  caribou  can 
locate  by  smell  and  paw  for  food  through  depths  of 
approximately  15  to  18  cm  (6  to  7  in.)  (Bergerud 
1978).  Elk  have  been  reported  to  shift  from  herba- 
ceous forage  to  browse  when  depths  exceeded  61 
cm  (24  in.)  (Skovlin  1982). 

Effects  of  snow  on  mobility  are  partly  caused  by 
depth  but  also  by  snow  density  and  crusting.  As  a 
general  rule,  most  ungulates  can  travel  through  snow 
that  reaches  up  to  their  bellies  without  great  diffi- 
culty. Some  species,  such  as  caribou,  have  wide 
hooves  that  help  them  walk  on  snow  rather  than 
through  it.  Kelsall  (1969)  suggested  that  snow 
depths  from  36  to  43  cm  (14  to  17  in.)  would  re- 
strict white-tailed  deer  mobility.  Similarly,  snow 
depths  greater  than  76  cm  (30  in.)  seriously  curtail 
movement  of  elk  (Skovlin  1982).  Sweeney  and  Stein- 
hoff  (1976)  suggested  that  depths  over  71  cm  (28 
in.)  usually  prohibited  use  by  elk. 

Many  ungulates  move  to  areas  with  lighter  snow 
cover  during  winter.  This  may  require  long  migra- 
tions, movement  to  lower  altitudes,  or  more  limited 
shifts  in  use  of  particular  cover  types  or  areas  with 
special  topographic  features.  In  mountainous  areas, 
south-facing  slopes  generally  have  lower  snow 
depths,  and  many  ungulates  such  as  mule  deer  and 
elk  will  concentrate  there.  However,  some  ungulates 
such  as  mountain  goats  use  windswept  ridges  at 
high  elevations. 

Snow  cover  can  easily  be  measured.  However, 
obtaining  enough  samples  throughout  a  range  during 
winter  would  be  exceedingly  difficult  and  of  ques- 
tionable value  unless  conditions  were  uniform 
throughout  the  range  from  year  to  year.  Mountain- 
ous areas  that  are  likely  to  remain  relatively  snow- 
free  can  be  predicted  to  a  limited  extent  from 
knowledge  of  elevation,  slope,  and  aspect.  Such  areas 
can  be  delineated  simply  and  easily  through  direct 
observation,  or  through  direct  observation  combined 
with  use  of  low-level  aerial  photographs  taken  at 
appropriate  times  during  winter. 


Elk  tracks  and  pawing  craters  in  snow. 


Cover 

Cover,  for  purposes  of  this  discussion,  is  defined 
as  any  structural  feature  of  the  environment  that  is 
used  for  protection  from  the  environment  (e.g.,  ther- 
mal cover)  or  from  predators  (e.g.,  security  cover). 
As  Bailey  (1984:1 10)  pointed  out,  cover  has  con- 
noted vegetation  cover  in  the  past.  However,  for  an 
ungulate  that  relies  on  sight  to  detect  predators  and 
on  open  country  for  escape,  absence  of  vegetation 
may  serve  the  same  function  as  presence  of  vegeta- 
tion does  for  a  hiding  species.  For  example,  bighorn 
sheep  rely  on  sight  to  detect  predators  and  avoid 
areas  of  dense  vegetation;  escape  cover  for  bighorn 
sheep  consists  of  steep,  open,  rugged  areas  (Bailey 
1984:112). 

For  many  years,  cover  for  wildlife  species  was 
described  in  terms  of  the  habitat  they  favored,  with- 
out distinguishing  the  many  biological  functions  such 
cover  was  providing.  Biologists  have  begun  to  iden- 
tify the  purposes  that  a  given  type  of  cover  serves. 
Black  et  al.  (1976)  distinguished  between  thermal 
and  hiding  cover;  the  former  is  defined  as  overstories 
that  give  protection  from  the  weather  and  sun  and 
the  latter,  as  vegetation  used  for  escape  and  protec- 
tion from  predators  and  humans.  This  is  a  useful 
distinction,  but  will  be  used  here  to  include  all 
structural  features,  rather  than  just  vegetation,  and 
will  be  termed  shelter  or  security  cover.  In  some 
cases,  both  functions  may  be  served  by  the  same 
physical  structure,  but  frequently  this  is  not  the  case. 
Shelter  and  security,  however,  are  the  two  most 
commonly  recognized  year-round  cover  needs.  As 
with  many  other  wildlife  species,  cover  requirements 
for  reproduction  (birthing  and  early  rearing)  of  un- 
gulates are  quite  specific  and  will  be  discussed  sepa- 
rately. Some  species  also  have  special  cover 
requirements  and  these  are  mentioned. 


526 


Ungulates 


# •  w        1 

"    A 

(^    '  * 

?  ■#* 

^^^i 

ik^feF* 

i«    m 

'&&>*•& 

T%^ 

■' / . 

-- 

vs     \.     JH|4i 

JBP'^s 

'• 

■&>  -    w» '.  m       ^  h  v  ^  * 

^   ■kill 

t  w  ^?  win? 

MftiCi  (£>y     J 

Escape  cover  for  bighorn  sheep. 


Shelter.  Ungulates  use  shelter  to  escape  the  ex- 
treme conditions  of  weather,  particularly  extreme 
heat  or  cold  or  rain,  hail,  or  snow.  Most  forest  or 
forest-edge  species  use  tree  or  brush  overstories  for 
this  purpose.  During  winter,  conifer  or  other  ever- 
green species  are  obviously  most  effective.  Open 
country  animals  such  as  bighorn  sheep,  pronghorn 
antelope,  and  bison  may  appear  at  first  to  survive 
without  using  shelter  cover.  However,  many  field  bi- 
ologists have  noted  that  these  species  will  use  minor 
topographic  features  as  shelter  from  weather.  On 
cold,  windy,  winter  days,  pronghorn  antelope  will 
frequently  be  found  in  swales  where  wind  speed  and 
chill  factor  are  substantially  less.  Similarly,  bighorn 
sheep  will  often  take  advantage  of  overhanging  cliffs 
to  provide  thermal  cover  in  winter  and  shade  in 
summer;  desert  bighorn  have  even  been  observed 
using  caves  during  the  summer.  To  our  knowledge, 
no  system  for  quantitatively  identifying  such  topo- 
graphic cover  features  has  been  devised  or  pub- 
lished. There  are,  however,  many  guidelines  on  what 
constitutes  suitable  shelter  or  thermal  cover  in  terms 
of  vegetation  by  animal  species.  For  example,  Wit- 
mer  et  al.  ( 1985)  define  thermal  cover  for  elk  as 
a  "stand  of  evergreen  coniferous  trees  over  40  ft  tall 
with  crown  closure  over  70%  with  an  optimum 
size  of  30  to  50  acres." 

Security.  Security  cover  is  used  for  protection 
from  predators  or  humans.  North  American  ungu- 
lates have  several  distinct  strategies  for  avoiding 
predation.  Hiders  such  as  the  white-tailed  deer  rely 
on  suitable  vegetation  cover  to  conceal  them  from 
predators;  they  can  also  usually  run  effectively 
through  dense  vegetation  to  escape  predators.  Many 
quantitative  definitions  are  available  for  what  consti- 
tutes hiding  cover  for  an  individual  ungulate  species. 
For  example,  Thomas  et  al.  (1979)  has  defined  hid- 
ing cover  for  elk  as  "vegetation  capable  of  hiding  an 


eik  from  view  at  200  ft  or  less  with  an  optimum 
size  of  30  to  50  acres." 

Watchers  such  as  pronghorn  antelope  live  in 
open  country  and  rely  on  observation  to  protect 
them  from  predators.  Some  watchers  can  effectively 
outrun  their  predators  in  open  country,  which  may 
be  level  terrain  in  the  case  of  the  pronghorns  or 
rocky  slopes  in  the  case  of  bighorn  and  Dall's  sheep. 
Other  watchers  rely  on  group  defense  (e.g.,  musk- 
oxen)  or  individual  defense  (e.g.,  moose)  as  protec- 
tion. Watchers  tend  to  congregate  into  groups 
ranging  from  small  family  units  to  large  herds  (over 
100  animals).  For  these  ungulates,  security  cover 
consists  of  suitable  open  habitat.  Pronghorn  ante- 
lope, for  example,  prefer  habitat  with  a  mean  height 
vegetation  of  37.5  cm  (15  in.);  areas  with  vegetation 
over  60  cm  (24  in.)  are  less  preferred  and  those 
over  75  cm  (30  in.)  are  rarely  used  (Yoakum  1980). 
This  preference  can  be  attributed  to  security  cover 
since  suitable  forage  is  often  more  abundant  in  areas 
with  taller  vegetation. 

Many  ungulates  use  more  than  one  strategy.  For 
example,  mule  deer  tend  to  hide;  yet  on  many 
ranges  they  will  congregate  into  large  herds  in  open 
areas,  such  as  for  feeding  in  late  winter  and  early 
spring,  and  use  a  watcher  strategy.  Often  this  feeding 
takes  place  near  rocky  terrain  or  more  thickly  vege- 
tated areas  where  deer  can  more  effectively  outrun 
predators. 

Ungulates  may  also  use  different  strategies  in 
different  parts  of  their  zoogeographic  range  or  even 
from  one  range  to  another.  Elk,  for  example,  tend 
to  use  coniferous  forests  for  hiding  cover  in  much  of 
the  West;  however,  in  recent  years  they  have  natu- 
rally colonized  shrub-steppe  regions  of  Washington, 
which  are  devoid  of  trees,  and  also  sagebrush 


Ungulates 


527 


steppes  of  Wyoming,  where  they  use  tall  shrubs  for 
security  cover.  Similarly,  animals  may  use  different 
strategies  against  different  predators.  Bighorn  sheep, 
for  example,  will  run  from  wolves,  but  they  may 
turn  and  defend  themselves  from  coyotes. 

Because  of  the  different  strategies  and  the  diver- 
sity of  security  cover  that  ungulates  may  use,  a  biolo- 
gist must  be  cautious  in  assessing  what  is  suitable 
security  cover.  First-hand  knowledge  of  the  range 
being  analyzed  and  of  the  behavior  of  the  animals 
that  use  it  are  necessary. 

Special  Cover.  Some  ungulates  use  special 
areas  for  cover  from  other  forces.  For  example,  cari- 
bou may  seek  windy  areas  in  summer  where  they 
are  more  protected  from  biting  and  stinging  insects. 

Relationship  between  Cover  Preference  and 
Requirements.  Preference  for  a  cover  type  does 
not  necessarily  indicate  that  the  type  is  required  for 
survival  or  that  its  absence  will  limit  the  herd  (Allen 
1977;  Peek  et  al.  1982).  For  example,  bighorn  sheep 
have  evolved  a  predator-evasion  strategy  and  use 
rocky  terrain  as  escape  cover  from  wolves.  Bighorn 
sheep  continue  to  use  such  habitats  even  in  areas 
where  wolves  have  been  extirpated  for  over  50 
years.  Do  they  need  it  or  merely  prefer  it?  Such 
questions  are  difficult  to  answer  and  raise  questions 
of  value  (e.g.,  should  we  be  managing  for  "natural" 
animals;  when  does  a  wild  animal  become  domesti- 
cated?). These  questions  are,  for  the  most  part,  be- 
yond the  scope  of  this  publication.  However,  the 
subject  of  cover  deserves  some  cautionary  notes. 

If  an  animal  has  evolved  a  strong  preference  for 
using  a  certain  type  of  security  or  shelter  cover  and 
this  cover  is  available  within  its  range,  then  that 
preference  will  effectively  determine  which  portion 
of  the  range  the  animal  will  use.  For  example,  big- 
horn sheep  often  deplete  forage  on  small  areas  of 
suitable  open,  rugged  terrain  to  the  detriment  of  the 
herd,  even  though  vast  expanses  of  coniferous  forest 
with  suitable  forage  surround  such  areas.  Under 
these  conditions,  the  question  of  preference  or  re- 
quirement is  moot.  Biologists  should  be  cautious  in 
assuming  they  understand  better  than  the  animal 
what  it  needs.  If  an  animal  has  evolved  or  learned  a 
strong  preference  for  a  given  type  of  cover,  it  is 
probably  adaptive,  even  though  the  reasons  may  not 
be  obvious  to  the  biologist. 

Birthing  Areas.  Many  ungulate  species  use  special 
birthing  areas.  The  lambing  areas  of  wild  sheep 
(Geist  1971a)  and  calving  grounds  of  caribou  (Lent 
1974a)  are  well-known  examples.  Birthing  habitat 
requirements  of  species  such  as  pronghorn  ( Fichter 
1974)  have  been  of  management  concern.  Such  spe- 
cial areas  arc  frequently  identified  in  the  develop- 
ment of  land-use  plans,  habitat  management  plans, 
and  environmental  impact  assessments.  For  those 


species  in  which  clear  birthing  habitat  descriptions 
are  available,  the  wildlife  biologist  should  be  alert  to 
identify  areas  meeting  these  specifications.  Docu- 
menting early  infant  mortality  rates  in  ungulates  is 
extremely  difficult,  however,  and  data  to  describe 
the  demographic  consequences  of  loss  or  alteration 
of  birthing  areas  are  not  available.  Most  examples  are 
anecdotal  and  relate  to  direct  human  disturbance 
(Kurt  1968;  Lent  1974a). 

Special  habitat  requirements  for  cover  may  be 
anticipated,  especially  for  hider  species  (Lent  1974a) 
that  depend  on  predator  defense  to  isolate  the  partu- 
rient female  from  others  of  the  species  and  subse- 
quently isolate  the  newborn  for  long  periods  of  the 
day.  Such  antipredator  strategy  may  be  significant  for 
up  to  2  months  after  birth,  like  in  mule  deer  (Geist 
1971a,  1981). 

Migration  and  Movements 

Most  North  American  ungulate  populations 
show  predictable  patterns  of  movements  over  the 
course  of  a  year,  periodically  returning  to  some  orig- 
inal locality.  Nomadic  movements  are  those  in  which 
such  a  fixed,  recurrent  locality  seems  to  be  absent. 
The  more  general  term  "movements"  is  used  in  the 
literature,  particularly  in  reference  to  migrations 
of  relatively  short  distance  or  when  the  type  of 
movement  is  not  clear. 

Even  white-tailed  deer,  generally  considered  to 
be  relatively  sedentary,  may  make  seasonal  migra- 
tions of  20  to  100  km  (12  to  60  mi.).  They  occupy 
more  restricted  home  ranges  (X  =  44  ha  [110  a.]) 
in  winter,  larger  ones  in  summer  ( 1 00  to  500  ha 
[250  to  1,250  a.])  (Nelson  1979;  Nelson  and  Mech 
1981).  This  illustrates  the  phenomenon  of  seasonal 
home  range  prevalent  in  many  temperate  and  boreal 
populations.  That  is,  there  are  relatively  localized 
seasonal  use  areas  where  movements  are  over  short 
distances,  connected  by  corridors  of  longer  distances 
traversed  relatively  quickly.  Moose  in  Wyoming  rep- 
resent another  well-documented  example  (Houston 
1974). 

The  biologist  must  be  particularly  alert  for  situa- 
tions where  resident  (relatively  sedentary)  and  mi- 
gratory populations  of  the  same  species  may  overlap 
during  certain  seasons.  This  has  been  reported  for 
white-tailed  deer  (Nelson  and  Mech  1981 ),  caribou 
(NPR-A  Task  Force  1978,  1979;  Cameron  and  Whit- 
ten  1979),  and  moose  (LeResche  1974).  Such  over- 
laps, which  often  require  marking  or  telemetry 
programs  to  identify  and  understand,  have  obvious 
implications  for  management.  Whether  one  needs  to 
estimate  seasonal  carrying  capacities,  regulate  har- 
vest by  establishing  zones  or  seasons,  or  assess  im- 
pacts of  proposed  developments,  one  needs  to  be 
able  to  discriminate  between  such  resident  and  mi- 
gratory populations. 


528 


Ungulates 


POPULATION  MEASUREMENT  TECHNIQUES 

Measuring  ungulate  population  attributes  is  one 
of  the  most  important,  yet  most  difficult  and  expen- 
sive, aspects  of  ungulate  management.  Mule  deer 
have  been  studied,  measured,  and  censused  for  over 
50  years  in  the  West,  yet  Connolly  ( 1981:226) 
wrote — 

"No  reliable  estimate  of  mule  . . .  deer  numbers 
exists  for  any  state  or  province.  The  only  avail- 
able estimates  are  speculative  and  often  they 
are  no  more  than  guesses  by  the  best-qualified 
persons." 

Determining  numbers  or  density  of  ungulates 
and  trends  for  these  measurements  are  extremely 
important  not  only  as  a  measure  of  the  effectiveness 
of  habitat  management  actions,  but  also  as  a  basis  for 
setting  harvest  quotas  or  goals.  Biologists  have  stud- 
ied this  problem  extensively,  and  a  vast  body  of  liter- 
ature has  accumulated.  Some  state  agencies  are 
making  substantial  progress  in  improving  methodol- 
ogy (Gill  et  al.  1983).  There  are  many  disagreements 
on  the  subject,  however,  and  much  work  needs  to 
be  done. 

Because  so  much  work  has  been  done  on  this 
subject,  this  section  only  serves  as  a  brief  overview 
to  literature  on  the  subject.  Biologists  will  need  to 
consult  original  references  for  detailed  information 
on  methodology  and  extended  discussions  of  advan- 
tages and  limitations  of  specific  techniques. 


Neck-banded  cow  elk  in  summer  range  herd 


Because  state  wildlife  agencies  are  responsible 
for  managing  resident  ungulate  species,  they  are 
usually  more  active  in  developing  and  implementing 
techniques  for  gathering  population  data  on  ungu- 
lates. Biologists  in  land-managing  agencies,  who  re- 
quire population  data,  should  consult  with  state 
agencies  to  determine  whether  such  data  are  avail- 
able and  can  be  obtained  from  them  or  through  a 
cooperative  effort.  Ungulate  surveys  are  expensive; 
publicly  funded  wildlife  programs  cannot  afford 
duplications. 

Presence/ Absence 

Evidence  of  presence  or  absence  of  an  ungulate 
species  in  a  given  area  can  be  determined  by  di- 
rectly observing  animals  or  tracks,  droppings,  or 
other  signs.  Tracks  of  most  North  American  ungu- 
lates are  relatively  distinct  (Murie  1954).  Droppings 
may  be  more  difficult  to  distinguish;  droppings  of 
mule  and  white-tailed  deer,  pronghorn  antelope, 
bighorn  sheep,  and  mountain  goat  overlap  in  size, 
shape,  and  appearance.  However,  a  biologist  may  be 
able  to  rule  out  all  except  one  of  these  species 
through  knowledge  of  habitat  requirements  and  geo- 
graphic distribution. 

Although  data  on  presence/absence  are  rela- 
tively easy  to  collect,  they  are  of  limited  value  for 
ungulates.  Management  agencies  and  the  public  are 
more  concerned  with  numbers,  density,  or  relative 
density.  Presence  information  is  most  useful  in  deter- 
mining geographic  distribution  or  use  of  a  range. 
The  geographic  distribution  of  North  American  un- 
gulates is  well-known  on  a  continental  or  regional 
basis;  however,  local  patterns  of  distribution  are 
often  poorly  understood.  For  example,  only  in  the 
past  35  years  has  the  distribution  of  bighorn  sheep 
in  North  America  been  described  in  enough  detail  to 
indicate  whether  a  given  mountain  range  supported 
sheep  populations.  Furthermore,  much  of  this  infor- 
mation has  only  been  gathered  in  the  past  1 5  years; 
some  states  still  do  not  have  this  information.  Finally, 
detailed  information  on  habitat  use,  seasonal  ranges, 
lambing  areas,  etc.  is  lacking  for  most  of  these  herds, 
and  much  of  the  information  obtained  in  the  past 
needs  to  be  updated.  Some  ungulates  such  as  moose 
are  expanding  their  range;  other  species  are  being 
locally  extirpated  and  many  herds  are  abandoning  or 
no  longer  using  parts  of  their  range.  Knowledge  of 
seasonal  ranges  including  migration  corridors  and 
birthing  areas  is  essential  for  good  habitat  manage- 
ment. This  is  basic  and  very  important  information 
and  often  can  be  obtained  with  equipment  no  more 
sophisticated  than  a  field  notebook,  topographic 
maps,  field  glasses,  and  a  stout  pair  of  hiking  boots. 

Relative  Density  or  Relative  Abundance 

Relative  density  or  relative  abundance  are  com- 
monly used  to  measure  the  effectiveness  of  habitat 


Ungulates 


529 


management  actions.  Collecting  this  data  provides  an 
index  of  population  abundance,  such  as  deer  ob- 
served per  square  mile  or  antelope  counted  per  mile 
of  highway.  Many  such  indexes  have  been  devel- 
oped; some  require  directly  observing  animals, 
whereas  others  use  indirect  signs,  such  as  tracks  and 
fecal  pellets.  Indexes  can  then  be  compared  from 
year  to  year  or  from  area  to  area.  Population  indexes 
do  not  usually  require  the  biologist  to  make  the 
same  sort  of  assumptions  as  required  with  popula- 
tion estimation  methods.  Furthermore,  they  are  eas- 
ier and  cheaper  to  obtain  and,  for  measuring 
population  trends  or  comparing  areas,  they  may  be 
adequate. 

The  major  problem  with  indexes  is  ensuring 
observations  or  counts  are  replicated  as  nearly  as 
possible.  Factors  that  may  need  to  be  standardized 
include  routes,  flying  or  driving  conditions,  weather, 
experience  of  personnel,  training,  protocol  for  such 
things  as  counting  an  animal  in  or  out  of  an  area, 
time  of  day,  etc.  The  value  of  indexes  for  measuring 
year-to-year  changes  in  abundance  significantly  de- 
creases and  becomes  useless  if  the  techniques  or 
protocol  for  collecting  the  information  is  periodi- 
cally changed. 

One  technique,  the  pellet  group  count,  needs 
further  explanation.  Biologists  have  attempted  with 
varying  success  to  use  this  technique  for  quite  differ- 
ent purposes  for  over  50  years  (Neff  1968).  As  a 
method  for  determining  presence,  it  is  sound.  As  a 
method  for  population  density  estimation,  it  has 
significant  problems  (Neff  1968).  As  a  method  for 
determining  relative  density,  it  is  useful  for  compar- 
ing years  or  large  areas,  but  may  be  misleading  if 
used  to  compare  habitat  use  between  or  among 
small  adjacent  areas.  The  problem  is  that  defecation 
does  not  correlate  well  with  habitat  use.  Elk,  for 
example,  often  defecate  while  traveling  (40%  of  all 
defecations),  yet  this  activity  consumes  only  about 
5%  of  their  day  (Collins  and  Urness  1979).  Using 
such  a  technique  within  the  home  range  of  an  elk 
herd  would  erroneously  suggest  that  areas  used  for 
travel  were  heavily  used.  On  the  other  hand,  if  the 
technique  is  used  to  compare  average  relative  densi- 
ties on  areas  large  enough  to  contain  the  daily  home 
ranges  of  animals,  then  such  factors  are  of  negligible 
importance. 

Density- 
Estimating  density  or  total  numbers  in  a  herd  is 
difficult  and  expensive,  yet  if  done  carefully,  it  can 
provide  very  useful  information.  The  many  tech- 
niques available  fall  into  four  general  categories:  total 
counts  (census),  sample  counts,  ratio  methods,  and 
indirect  methods. 

Total  Counts.  Total  counts  are  most  successful 
with  diurnal  ungulates  on  open  country,  such  as 

530  Ungulates 


pronghorn  antelope.  Observers  on  foot  often  try  to 
totally  count  small  isolated  herds  that  are  concen- 
trated, such  as  on  a  winter  range  or  at  waterholes  in 
summer.  Such  counts  usually  require  good  planning 
and  coordination  to  ensure  that  all  animals  are  ob- 
served and  are  not  counted  more  than  once. 

Various  methods  have  been  devised  to  correct 
or  adjust  for  these  problems,  but  none  are  entirely 
satisfactory.  For  this  reason,  information  from  many 
such  efforts  are  used  as  indexes  rather  than  total 
counts. 

Total  counts  are  often  attempted  from  helicop- 
ters or  fixed-wing  aircraft,  although  more  often  some 
form  of  sampling  is  used.  The  practical  problems 
are  similar,  but  sampling  schemes  make  the  calcula- 
tions more  complex. 

Sample  Counts.  Sample  counts  are  of  two  general 
types,  strip  samples  and  plot  samples.  Strip  samples 
have  been  used  effectively  to  count  ungulates  from 
airplanes,  motor  vehicles,  or  on  foot  (Riney  1982:78- 
84).  Many  different  types  of  strip  samples  are  avail- 
able; Robinette  et  al.  (1974)  tested  many  of  these 
and  Burnham  et  al.  (1980)  summarized  assumptions 
and  statistical  analyses. 

Plot  counts  have  been  used  less  frequently,  es- 
pecially from  airplanes.  However,  a  helicopter  quad- 
rat system  (Kufeld  et  al.  1980)  is  being  used  in 
Colorado  and  initial  results  and  tests  appear  promis- 
ing (Gill  et  al.  1983). 

Ratio  Methods.  Ratio  methods,  including  mark- 
recapture  methods,  are  based  on  knowing  the  num- 
ber of  marked  animals  in  a  population  or  removed 
from  a  population  at  some  point.  These  methods 
generally  require  animals  to  be  marked  or  tagged — 
expensive  and  time-consuming  with  ungulates.  Many 
variations  on  the  basic  capture/recapture  models 
have  been  developed.  Refer  to  Riney  (1982:42-74) 
for  practical  aspects  of  capturing  and  marking  ani- 
mals, to  Bailey  (1984:311-315)  and  Gilbert 
( 1979:302-304)  for  a  general  discussion  of  ratio 
methods,  and  to  White  et  al.  (1982)  for  a  discussion 
of  detailed  statistical  treatment. 

Indirect  Methods.  Indirect  methods  correlate  ani- 
mal density  with  some  measurable  and  observable 
sign  left  by  the  animals.  With  ungulates,  the  indirect 
method  used  most  frequently  is  the  pellet  group 
density  measurement  (Freddy  and  Bowden  1983a,b; 
Neff  1968). 

General  Cautions.  Gill  et  al.  (1983)  suggested  that 
biologists  obtain  reliable  estimates  of  ungulate  popu- 
lation density  to  track  changes  over  time.  This  is 
certainly  laudable  and  biologists  with  agencies  re- 
sponsible for  population  management  would  be  well- 
advised  to  pursue  it.  However,  because  such  esti- 


mates  are  expensive  and  difficult  to  obtain  and  the 
results  often  debatable,  many  biologists  and  man- 
agers prefer  to  use  indexes.  Furthermore,  although 
many  techniques  have  serious  disadvantages  when 
used  to  count  and  estimate  populations,  most  of 
these  disadvantages  disappear  when  they  are  used  as 
indexes  (Riney  1982:85).  The  biologists  charged 
with  habitat  management  or  working  for  a  habitat 
management  agency  should  consider  whether  a  pop- 
ulation or  condition  index  will  be  adequate  for  their 
purposes. 

Animal  Condition 

Condition  indexes  have  been  developed  and 
used  experimentally  for  many  years  to  measure  un- 
gulate health,  but  few  have  been  used  operationally 
by  management  agencies.  Generally,  condition  meas- 
urements or  observations  are  collected  on  individual 
animals  and  then  averaged  to  derive  an  index  of  the 
condition  of  the  herd  or  population.  Animal  condi- 
tion indexes  can  be  considered  a  function  of  the 
habitat  and  the  number  of  animals.  Thus,  unlike  most 
other  population  measurements,  they  reflect  ecologi- 
cal density  or  the  number  of  animals  relative  to  the 
quantity  and  quality  of  habitat  available  (Bailey 
1984:318).  Indeed,  many  measures  of  population 
structure,  such  as  doe  to  fawn  ratios  (a  measure  of 
productivity)  or  behavioral  indexes,  can  be  consid- 
ered special  types  of  population  condition  indexes 
(Bailey  1984:318-324;  Hanks  1981). 

Population  Structure  and  Productivity 

With  most  North  American  ungulates,  sex  and  a 
limited  number  of  age  classes  (young  of  the  year, 
yearling,  adult)  can  be  distinguished  in  the  field.  In 
addition,  more  detailed  age  information  can  be  ob- 
tained from  dead  or  harvested  animals  using  tooth 
eruption  characteristics,  dental  cementum  annuli,  or 
horn  growth  rings.  Biologists  have  collected  much 
data  on  sex  and  age  structures  of  ungulate  popula- 
tions and  expended  much  effort  in  developing  and 
refining  methods  to  analyze  such  data  (Caughley 
1977;  Riney  1982;  Fowler  and  Smith  1981).  This 
type  of  information  is  now  commonly  used  in  popu- 
lation models  that  can  simulate  the  dynamics  of  herd 
numbers  and  structure  over  time  (Conley  1978). 
The  models  are  now  used  routinely  by  several  west- 
ern state  game  agencies  as  a  source  of  information  to 
be  used  with  other  sources  in  making  recommenda- 
tions about  harvest  goals  and  quotas  for  big  game 
species  (Pojar  and  Strickland  1979). 

Some  very  good  work  has  been  done  with  popu- 
lation structure  measurements.  However,  manage- 
ment biologists  must  keep  in  mind  the  original 
reasons  for  collecting  population  structure  data  and 
avoid  getting  bogged  down  in  the  complexities  of 
population  modeling.  For  management  purposes,  the 
simplest  and  directly  observable  measures,  such  as 


cow  to  calf  ratios,  have  often  proven  to  be  the  most 
useful. 

Behavioral  Indicators 

Interest  in  the  use  of  behavioral  attributes  as 
indexes  of  population  and  habitat  condition  is  in- 
creasing. Hanks  (1981:61)  reported  that  several  stud- 
ies of  population  condition  correlate  declines  in 
physiological  condition  or  demographic  vigor  with 
certain  behavioral  attributes  of  the  population.  He 
recommended  considering  these  attributes  as  a  pos- 
sible early  warning  system  of  adverse  population 
trends:  ( 1 )  rate  and  quality  of  social  interaction,  (2) 
population  density,  and  (  3  )  feeding  strategy.  As  an 
example,  Horejsi  (1976)  concluded  that  duration  of 
suckling  and  total  time  spent  suckling  could  be  re- 
lated to  population  and  habitat  quality  in  bighorn 
sheep. 

MAJOR  SPECIES 

The  following  descriptions  treat  each  major 
ungulate  species  or  species  group  separately  because 
of  the  importance  of  these  animals.  Because  many 
of  these  species  are  similar  in  many  ways,  the  sec- 
tions emphasize  aspects  of  the  biology  and  manage- 
ment that  are  unique  to  the  species. 

Many  books  and  monographs  are  available  on 
North  American  ungulates.  Some  of  the  more  recent, 
major  publications  about  population  or  habitat  meas- 


Jaw  board  used  as  aging  reference  at  big  game  check  sta- 
tions. 


Ungulates 


531 


Table  2.     Major  recent  references  on  North  American  ungulates  that  cover  habitat  and  population 


i 


measurements. 


Species 

Periodicals 

Books/Monographs 
Bibliographies 

General 
Individual  Species 

Schmidt  and  Gilbert 
(1978) 

Collared  peccary 

Sowls  (1984) 

Elk 

Proceedings  of  Western  States  Elk  Workshop 

Thomas  and  Toweill 
(1982) 

Mule  deer 

Wallmo  (1981) 
Kerr  (1979) 

White-tailed  deer 

Halls  (1984) 

Moose 

Alces  (Proceedings  of  North  American  Moose  Confer- 
ence and  Workshop 

Caribou 

Proceedings  of  International  Reindeer  and  Caribou  Sym- 
posiums 

Pronghorn 

Proceedings  of  Pronghorn  Antelope  Workshop 

Yoakum  (1980) 
Autenrieth  (1978) 

Bison2 

Mountain  goat 

Proceedings  of  Biennial  Symposium  of  the  Northern 
Sheep  and  Goat  Council 

Samuel  and  Mac- 
Gregor  (1977) 

Muskox 

Proceedings  of  International  Muskox  Symposiums 

Klein  et  al.  (1984) 

Bighorn  sheep 
and 

Proceedings  of  Desert  Bighorn  Council 

Monson  and  Sumner 

(1980) 
Trefethen  (1975) 

Dall's  Sheep 

Proceedings  of  Biennial  Symposiums  of  Northern  Sheep 
and  Goat  Council 

Krausman  et  al.  (1984) 

'Only  references  since  1975  are  cited. 
2No  recent  general  references  available. 


urements  are  listed  in  Table  2;  others  are  cited  in 
specific  species  subsections. 

Collared  Peccary 

Of  all  large  game  animals  in  North  America,  the 
collared  peccary  (Tayassu  tajacu)  has  been  the  last 
to  get  the  attention  of  wildlife  biologists.  In  many 
places  it  was  considered  vermin  or  a  pest  to  live- 
stock operators.  In  the  past  30  years,  however,  the 
peccary  has  become  a  protected  big  game  species  in 
Arizona,  New  Mexico,  and  Texas.  Careful  censuses 
and  regulated  harvests  now  ensure  its  survival  if 
habitats  can  be  preserved.  The  most  serious  limiting 
factor  to  the  future  survival  of  this  animal  is  habitat 
loss.  Areas  of  unspoiled  habitats  need  to  be  set  aside 
to  ensure  their  future  existence  (Sowls  1984).  Preda- 
tion  levels  are  thought  to  be  low,  although  hard 


i^*-"p. 


^^t?^^:.  * 


Collared  peccary. 


532 


Ungulates 


data  are  not  available  on  the  effects  of  predation  on 
population  levels  (Sowls  1984).  Although  the  col- 
lared peccary  could  have  both  ecto-  and  endoparas- 
ites,  they  do  not  seem  to  cause  significant  population 
losses  (Sowls  1984). 

The  species  is  not  long-lived  (mean  length  of 
life  is  about  4  years),  but  the  high  reproductive  rate 
and  its  ability  to  withstand  severe  food  and  water 
shortages  for  short  periods  enables  the  species  to 
maintain  good,  young  animal  populations. 

Habitat  Requirements  and  Analysis.  In  the  South- 
west, collared  peccaries  occupy  a  variety  of  habitats. 
Habitat  types  occupied  by  peccaries  in  Arizona  have 
been  described  by  Knipe  ( 1957),  Neal  (1959),  and 
Eddy  ( 1961 ).  Arizona  populations  are  found  primar- 
ily in  the  southeastern  quarter  of  the  state  and  are 
generally  contiguous  with  populations  in  New  Mex- 
ico (Bissonette  1982).  The  Arizona  populations  in- 
habit areas  of  low  desert-scrub  vegetation  where  the 
principal  plants  are  saguaro  (Carnegia  gigantea), 
mesquite  (Prosopis  sp. ),  paloverde  (Cercidium  sp. ), 
and  many  species  of  cactus.  The  oak-woodland,  chap- 
arral, and  lower  areas  of  ponderosa  pine  {Pinus  port- 
derosa)  forests  are  occupied  during  the  summer, 
but  can  be  too  cold  for  peccaries  in  the  winter.  Pec- 
caries can  inhabit  desert  environments  because  they 
make  full  use  of  microclimates,  and  during  very  hot 
summer  periods,  they  move  only  at  night. 

Peccaries  in  New  Mexico  occupy  two  types  of 
vegetation  cover.  One  is  the  oak-juniper  association 
between  elevations  of  1,500  and  1,940  m  (5,000  and 
6,400  ft)  just  below  the  ponderosa  pine  belt.  The 
other  cover  type  is  mesquite,  beginning  at  elevations 
about  1,303  m  (4,300  ft)  and  extending  to  about 
1,879  m  (6,200  ft)  on  some  exposures  (Donaldson 
1967).  Other  protective  cover  includes  caves  and 
mine  shafts,  natural  crevices,  rock  overhangs,  and  un- 
dercut arroyo  banks. 

Habitats  occupied  by  peccaries  in  Texas  vary 
from  dry  deserts  in  the  west  to  lush  chaparral  and 
deciduous  brushlands  in  the  east.  Peccaries  occupy 
about  260,000  km2  (100,000  mi.2)  in  this  state  in 
areas  having  rainfall  varying  from  40  cm  (16  in.)  per 
year  in  the  west  to  75  cm  (27  to  30  in.)  per  year 
in  the  east  (Jennings  and  Harris  1953;  Low  1970; 
Bissonette  1976;  Ellisor  and  Harwell  1979). 

Food  and  water  are  supplied  by  succulent  food 
(Sowls  1978).  The  prickly  pear  cactus  is  the  pre- 
ferred food  of  the  peccary  in  Arizona  (Knipe  1957), 
but  tubers,  bulbs,  and  rhizomes  are  also  important. 
In  Texas,  good  peccary  habitat  consists  of  heavy 
brush  with  an  abundance  of  prickly  pear  (Ellisor  and 
Harwell  1979).  In  New  Mexico,  major  food  items 
include  oak  (Quercus  sp.)  mast,  juniper  (Juniperus 
deppeana  and/  monosperma),  pinyon  (Pinus  edu- 
lis),  mesquite  beans,  leaves  and  fruits  of  Engelmann 


prickly  pear  (Opuntia  engelmanni),  and  mescal 
(Lophophora  williamsii)  or  Palmer  agave  (Agave 
palmer i)  (Donaldson  1967). 

Peccaries  do  not  depend  entirely  on  free  water, 
but  use  it  if  it  is  available.  Peccaries  can  usually  ob- 
tain ample  amounts  of  water  from  prickly  pear,  the 
basal  ends  of  mescal,  and  underground  bulbs  (Don- 
aldson 1967).  Thus  they  can  withstand  long  dry 
periods  without  ill  effects. 

Since  prickly  pear  cactus  seems  to  be  the  most 
common  of  all  peccary  foods,  habitat  measurements 
should  include  determination  of  abundance  of 
prickly  pear.  Because  the  greatest  threat  to  peccary 
populations  is  habitat  loss  (Sowls  1978),  the  vegeta- 
tion trend  of  peccary  habitat  should  be  monitored, 
particularly  that  of  prickly  pear. 

Population  Measurements.  Compared  with  more 
common  big  game  animals,  peccaries  present  new 
problems  to  researchers  trying  to  gather  and  inter- 
pret biological  data.  The  principal  problem  is  that 
peccaries  have,  a  year-round  breeding  season.  This, 
coupled  with  the  female's  ability  to  become  preg- 
nant within  a  few  days  after  giving  birth,  gives  the 
species  a  very  high  breeding  potential.  Because  of 
these  factors,  these  animals  do  not  fit  as  easily  into 
age  classes  or  cohorts  like  elk,  deer,  and  other  big 
game  animals  (Sowls  1984).  Life  tables  have  been 
constructed  from  ages  of  hunter-killed  animals,  but 
problems,  primarily  unequal  birth  and  survival  rates, 
make  use  of  these  analytical  techniques  suspect 
(Low  1970). 

In  Texas,  aerial  surveys,  road  censuses  by 
ground  vehicles,  and  track  counts  have  been  tried 
for  inventorying  these  animals.  These  census  meth- 
ods, however,  are  not  feasible  in  brushy  country. 
In  Arizona,  extensive  testing  of  peccary  census  meth- 
ods involved  track  counts,  counts  at  waterholes, 
helicopter  counts,  surveys  by  biologists  on  foot,  and 
questioning  hunters  about  numbers  of  animals  seen. 
Results  indicated  that  helicopter  counts  were  too 
expensive  and  variable,  and  track  counts  were  too 
variable.  The  Arizona  Game  and  Fish  Department 
found  that  deer  hunter  counts  of  peccaries,  obtained 
right  after  a  hunt,  were  more  accurate  than  sightings 
reported  later  in  questionnaires.  They  also  reported 
that  data  from  deer  hunters  compared  well  with  data 
obtained  by  biologists  surveying  on  foot  (Sowls 
1978).  Thus,  the  most  promising  method  seems  to 
be  questioning  hunters  in  peccary  range  immediately 
after  hunting  seasons  for  other  species,  particularly 
deer.  Combined  with  foot  or  horseback  observations 
by  biologists,  this  method  gives  good  indications  of 
population  number  and  trend.  Roadblocks  and  estab- 
lished hunter  checking  stations  also  provide  impor- 
tant data  on  peccary  populations,  and  data  on 
productivity  can  be  obtained  by  estimating  ages  of 
harvested  animals. 


Ungulates 


533 


Discussion.  The  most  important  management  con- 
sideration is  maintaining  as  much  suitable  peccary 
habitat  as  possible.  The  collared  peccary  has  high 
reproductive  potential  and  is  remarkably  adaptable. 
For  these  reasons,  healthy  and  productive  popula- 
tions of  this  unique  animal  can  be  effectively  man- 
aged throughout  its  range.  Although  habitat  factors 
limiting  peccary  populations  are  not  well-known, 
evidence  suggests  that  density  of  prickly  pear  is  very 
important  and  should  be  a  high  priority  for  inven- 
tory and  monitoring. 


Elk 


At  one  time,  the  American  elk  (Cervus  elaphus) 
was  the  most  widely  distributed  member  of  the  deer 
family  in  North  America.  It  was  found  from  the  At- 
lantic to  Pacific  coasts  and  from  Mexico  to  northern 
Alberta,  except  the  southern  coastal  plains  and  Great 
Basin  (Hall  and  Kelson  1959).  However,  as  people 
moved  westward,  elk  disappeared  from  settled  re- 
gions until  they  virtually  vanished  from  most  of  their 
historic  ranges.  Remnant  herds  were  scattered  in 
the  Rocky  Mountain  region  and  parts  of  the  Pacific 
Northwest  and  Canada.  In  the  past  50  years,  elk  have 
expanded  their  range  and  been  reintroduced  so  that 
they  are  fairly  widespread  throughout  the  mountain- 
ous areas  of  the  West. 

Habitat  Requir ements  and  Analysis.  Much  of  the 
work  in  recent  years  on  elk  habitat  has  concentrated 
on  ( 1 )  identifying  areas  used  for  foraging,  cover, 
and  other  needs,  and  the  necessary  spatial  relation- 
ships of  these  areas  using  the  Thomas  et  al.  ( 1979) 
landscape  approach,  and  (2)  determining  the  effects 
of  logging,  roads,  fire,  livestock  grazing,  and  other 
human  disturbance  on  habitat  suitability  for  elk. 

Thomas  et  al.  (1979)  defined  optimum  elk  habi- 
tat for  the  Blue  Mountains  of  Oregon  and  Washing- 
ton in  terms  of  a  proper  mixture  of  foraging  and 
cover  areas.  They  further  subdivided  cover  into  hid- 
ing cover  and  thermal  cover.  This  landscape  ap- 
proach has  been  modified  for  use  in  many  other 
areas  of  the  West.  For  example,  Witmer  et  al.  ( 1985) 
have  used  this  approach  for  the  elk  in  the  coastal 
forests  of  Oregon  and  Washington,  and  Wisdom  et  al. 
(1986)  have  formalized  this  into  a  habitat  evaluation 
model. 

The  greatest  limitation  of  such  a  landscape  ap- 
proach is  that  the  quantity,  quality,  or  availability  of 
forage  in  "foraging  areas"  is  not  considered.  Elk  con- 
sume more  forage  per  animal  than  most  of  the  other 
common  North  American  ungulates,  and  their  forage 
preferences  are  similar  to  those  of  domestic  cattle 
(Cooperrider  1982).  Therefore,  where  elk  are  reach- 
ing carrying  capacity  or  where  they  occupy  common 
range  with  cattle,  forage  quantity  and  quality  can  be 
a  problem. 


For  a  current  and  more  thorough  review  of  elk 
habitat  requirements  and  evaluations,  see  Skovlin 
(1982). 

Population  Measurements.  Enumerating  various 
components  (age  structure,  sex  and  age  ratios,  and 
total  population)  of  an  elk  population  is  difficult,  but 
important.  These  data  tell  the  biologist  or  manager 
much  about  the  dynamics  and  general  health  of  the 
herd,  and  with  adequate  sample  sizes,  they  provide 
the  information  needed  for  proper  elk  population 
management. 

Age  Structure.  Two  methods  are  used  to  de- 
termine age  structure.  The  first  involves  examining 
tooth  wear  and  replacement  in  elk  killed  by  hunters. 
Aging  by  this  method  is  based  on  the  sequence  of 
tooth  eruption  and  wear.  Several  good  aging  guide- 
lines are  available,  but  techniques  developed  by 
Quimby  and  Gaab  (1957)  are  used  by  most  wildlife 
researchers  and  managers.  One  problem  associated 
with  this  method,  however,  is  obtaining  a  large 
enough  sample  to  enable  the  manager  to  make  an 
accurate  estimate  of  the  herd's  age  structure. 

To  increase  samples,  hunters  are  invited  to  re- 
move the  lower  jaws  from  their  kills  and  bring  them 
to  check  stations.  However,  ages  obtained  by  this 
method  do  not  necessarily  represent  a  herd's  age 
composition.  If  a  wildlife  management  agency  is 
willing  to  spend  the  time  and  money,  rather  large 
sample  sizes  of  jaws  can  be  gathered  by  setting  up 
several  check  stations  around  one  herd  area.  This 
method  is  effective  for  special  research  areas  or  ex- 
tremely important  herds,  but  is  not  satisfactory  for 
statewide  age-structure  estimations. 

The  second  method  is  the  dental  cementum- 
aging  procedure,  which  requires  only  an  incisor 


Cross-section  of  an  incisor  from  an  unknown  age  elk  killed 
in  October;  age  estimated  at  six  years  by  cementum  annuli 
count. 


534 


Ungulates 


tooth  from  a  harvested  elk  (Keiss  1969).  By  using 
this  method,  age  samples  from  all  elk  herds  in  a  state 
can  be  greatly  increased.  The  tooth  is  sectioned  with 
a  special  sectioning  saw,  put  through  various  staining 
procedures,  and  then  examined  through  a  micro- 
scope for  cementum  annuli  which  can  be  counted  in 
a  manner  similar  to  counting  rings  in  a  tree  stump. 
By  sending  every  hunter,  or  a  statistically  reliable 
sample  of  hunters,  a  small  postage-paid  envelope  in 
which  to  collect  and  mail  an  incisor,  sample  sizes 
can  be  increased  substantially.  A  hunter  can  more 
easily  remove  and  carry  an  incisor  than  a  whole  jaw. 

Sex  and  Age  Ratios.  Ground  counts  and  aerial 
counts  are  two  general  methods  used  to  estimate  sex 
and  age  ratios.  Ground  counts  are  useful  in  areas 
where  elk  congregate  in  large  numbers,  such  as  feed- 
ing grounds  or  winter  ranges.  Ground  sex  and  age 
classifications  can  be  considerably  biased,  however, 
because  larger  bulls  tend  to  winter  away  from  the 
cow/calf/young  bull  groups.  Ground  sex  and  age 
classifications  are  generally  only  possible  from  De- 
cember through  February,  when  elk  are  congregated 
on  their  winter  ranges.  This  means  that  some  mortal- 
ity has  occurred  after  the  hunting  season  so  that 
ratios  are  not  exactly  the  posthunt  ratios  needed  for 
some  population  formulas.  In  addition,  large  samples 
of  prehunt  sex  and  age  ratios  can  rarely  be  obtained 
by  ground-count  methods.  While  the  total  number  of 


elk  on  winter  ranges  can  be  satisfactorily  estimated 
from  a  fixed-wing  aircraft,  sex  and  age  ratio  counts 
cannot.  Low  and  slow-flying  helicopters  are  needed 
to  accurately  determine  these  ratios. 

Prehunt  sex  and  age  ratio  counts  should  be 
made  in  mid-September  at  the  peak  of  the  rut  when 
the  largest  percentage  of  bulls  can  be  observed.  Elk 
can  be  classified  as  spike  bulls;  young  bulls  (typical 
antler  conformation,  but  with  light  beams  and  usu- 
ally four  or  five  points);  mature  bulls  (typical  antler 
conformation,  but  with  heavy  beams);  cows;  and 
calves.  Calves  can  be  distinguished  from  yearling 
cows  primarily  from  the  face  or  nose  conformation, 
but  color,  relative  thickness  of  the  neck,  and  com- 
parative size  are  other  criteria  (Boyd  1970,  1978). 

Biologists  can  facilitate  classifications  by  using 
tape  recorders  to  tally  various  categories  and  later 
extract  the  data.  This  method  allows  them  to  watch 
the  elk  continually  while  recording  the  sex  or  age  of 
each  animal. 

Sex  and  age  ratio  counts  can  be  made  in  early 
December,  before  the  larger  bulls  have  moved  to 
high  and  isolated  wintering  areas  away  from  the 
main  herd,  to  estimate  changes  caused  by  fall  hunt- 
ing. Techniques  and  criteria  of  posthunt  counts  are 
not  different  from  prehunt  counts. 


Calves  are  distinguished  from  yearling  cows  primarily  on  the  basis  of  nose  or  face  conformation  (a  calf  has  a  short,  blunt 
nose).  The  neck  on  a  calf  appears  to  be  shorter  and  heavier  than  that  of  a  mature  cow.  A  calf  has  a  fuzzy,  rounded  patch 
on  its  forehead  and  in  general  has  a  "puppylike"  appearance. 


Ungulates 


535 


Total  Population.  To  properly  manage  an  elk 
population,  the  wildlife  manager  must  estimate  the 
total  elk  population  on  a  given  range.  Several  census- 
ing  methods  can  be  used  to  make  these  estimates. 

Elk  on  winter  ranges  have  been  aerially  counted 
since  about  1946.  These  sample  counts,  for  the  most 
part,  are  accomplished  by  using  fixed-wing  aircraft 
but  do  not  encompass  all  elk  in  a  survey  area.  In 
Colorado,  the  percentages  of  deer  or  elk  observed 
from  the  air  were  compared  with  the  total  popula- 
tion present  for  several  years.  Only  about  37%  of  the 
elk  present  on  a  sagebrush-oakbrush-serviceberry 
(Artemesia-Quercus-Amelanchier  sp. )  wintering  area 
actually  were  seen  by  aerial  observers  (Boyd  1958). 
However,  winter  counts  are  used  by  many  wildlife 
management  agencies  to  evaluate  the  population 
trends  and  to  set  annual  hunting  seasons  and 
regulations. 

Indirect  census  methods  have  been  used  suc- 
cessfully by  several  management  agencies  by  con- 
trasting prehunt  and  posthunt  sex  ratio  counts  and 
harvest  estimates  (Boyd  1970).  Among  the  many 
references  concerning  these  methods,  one  excellent 
review  applicable  to  elk  is  by  Hanson  (1963),  which 
provides  a  discussion  on  calculating  abundance  of 
vertebrates  from  sex-age  ratios. 

While  population  figures  resulting  from  these 
methods  are  extremely  valuable,  they  do  not  tell  the 
entire  story  of  elk  herd  dynamics.  Missing  data  in- 
clude estimates  of  nonhunting  mortality  and  formu- 
las to  calculate  the  consequences  of  past  and  future 
management  on  elk  populations. 

A  wildlife  manager  also  needs  to  set  manage- 
ment goals  and  priorities  for  managing  wildlife  popu- 
lations. Obtaining  reliable  information  on  which  to 
base  these  objectives,  however,  is  difficult.  Without 


Mature  bull  elk. 


tools  or  methods  to  assist  in  gathering  and  analyzing 
data,  management  often  ignored  many  elk  population 
factors  when  setting  harvest  limits,  regulations,  and 
management  priorities.  Now,  with  the  development 
of  the  computer,  these  factors  can  be  considered 
(Gross  1970;  Pojar  1981).  All  factors  that  contribute 
to  the  life  cycle  of  an  elk  population  (i.e.,  sex-age 
ratios,  cow-calf  ratios,  harvest  by  sex  and  age, 
wounding  loss)  are  needed  to  conduct  a  simulation 
(Boyd  et  al.  1975). 

One  new  approach  to  censusing  elk  involves  the 
use  of  mark-recapture  techniques  to  estimate  num- 
bers of  elk  (Bear  and  Green  1980).  Individual  elk  in 
a  wintering  area  are  randomly  trapped  and  marked 
with  colored  ear  tags.  An  attempt  is  made  to  capture 
about  10%  of  the  estimated  herd.  The  recapture 
effort  is  accomplished  by  surveying  the  entire  winter 
range  with  a  helicopter  as  soon  after  trapping  as 
possible.  All  ear-tagged  individuals  are  counted  to 
determine  the  ratio  of  marked  to  unmarked  animals. 
Three  or  four  recapture  flights  are  made  each  year. 
This  system  is  quite  precise  (SE  ±  5%  ).  Accuracy  is 
unknown,  but  the  estimates  agree  very  well  with  life- 
table  analyses,  other  aerial  counts,  and  estimates 
from  biologists  familiar  with  the  area  (Gill  et  al. 
1983). 

Discussion.  The  major  concerns  of  managing 
elk  habitat  focus  on  the  effects  of  competing  land 
uses,  particularly  logging,  grazing,  and  other  human 
disturbance.  Logging  especially  is  a  concern  because 
it  physically  alters  the  habitat  through  cutting  trees, 
piling  slash,  and  other  forestry-related  activities.  In 
addition  to  the  immediate  disturbance  from  such 
activity,  habitat  is  also  subsequently  disturbed  from 
roads  built  as  part  of  the  logging  job.  Not  only  do 
elk  tend  to  avoid  well-used  roads,  such  roads  also 
provide  increased  hunter  access.  The  effects  of  log- 
ging, roads,  and  other  human  disturbances  are  de- 
scribed by  Lyon  et  al.  (1985),  Lyon  and  Ward 
(1982),  and  Heib(1976). 

Grazing  is  another  concern  because  elk  compete 
with  livestock  for  forage.  The  competition  is  most 
severe  with  elk  and  cattle  because  they  occupy 
many  common  ranges,  their  diets  are  similar,  and 
both  species  consume  large  amounts  of  forage 
(Cooperrider  1982). 


Mule  and  White-tailed  Deer 

The  mule  deer  {Odocoileus  hemionus)  ranges 
over  most  of  temperate  North  America  between  the 
Pacific  Coast  and  the  100th  Meridian.  The  white- 
tailed  deer  (O.  virginianus)  is  the  most  popular  and 
widespread  big  game  animal  in  North  America,  found 
in  almost  all  the  contiguous  48  states.  It  thrives  in  a 
wide  range  of  climatic  and  habitat  conditions  and 
adapts  quickly  to  changes. 


536 


Ungulates 


Mule  and  white-tailed  deer  are  often  victims  of 
automobile  accidents,  diseases,  parasites,  predation, 
starvation,  wounds,  fence  entanglements,  poaching 
and,  of  course,  legal  hunting.  A  primary  limiting  fac- 
tor for  these  species  would  be  habitat  degradation 
caused  by  urban  development,  deteriorating  forest 
habitat,  and  agricultural  development.  Of  all  these 
limiting  factors,  the  primary  management  concern  is 
habitat  loss. 

Habitat  Requirements  and  Analysis.  Mule  deer 
and  white-tailed  deer  have  been  intensively  studied 
for  many  years.  However,  because  they  occupy  such 
diverse  habitats,  generalizations  about  habitat  re- 
quirements are  extremely  difficult  to  make.  A  biolo- 
gist should,  therefore,  consult  the  literature  on  a 
particular  region  for  specific  habitat  requirements. 
Wallmo  (1981)  and  Halls  (1982)  provide  extensive 
reviews  of  habitat  evaluation  and  management  for 
mule  and  white-tailed  deer,  respectively. 

Because  mule  deer  are  adaptable,  they  may  be 
found  in  all  major  climatic  and  vegetational  zones  of 
the  western  U.S.  except  the  arctic,  tropics,  and  ex- 
treme deserts.  Generally,  mule  deer  inhabit  semiarid, 
open  forest,  brush,  and  shrublands  found  in  conjunc- 
tion with  steep,  broken,  or  otherwise  rough  terrain. 
However,  extensive  populations  are  also  found  in 
prairie  habitats  along  the  eastern  limits  of  their  dis- 
tribution and  in  semidesert  shrub  habitats  of  the 
Southwest  (Mackie  et  al.  1982).  Black-tailed  deer, 
which  consist  of  two  mule  deer  subspecies,  are 
found  in  temperate,  coniferous  forests  along  the  Pa- 
cific Coast  from  central  California,  north  to  south- 
eastern Alaska.  Populations  can  also  be  found  in 
adjacent  woodland-chaparral  areas  of  the  coastal 
ranges  of  California. 

In  general,  the  white-tailed  deer  does  best  in 
subclimax  habitat  (Hesselton  and  Hesselton  1982), 
particularly  cut  over  forests.  This  is  particularly  true 


in  the  eastern  U.S.  However,  west  of  the  100th  Me- 
ridian, white-tailed  deer  are  often  associated  with 
riparian  areas  along  major  river  drainages.  Many  of 
these  areas  consist  of  forested  stands  of  cottonwoods 
(Populus  sp. ),  willows  (Salix  sp. ),  and  other  tree 
species.  Although  white-tailed  deer  have  not  been 
studied  as  extensively  in  the  West,  the  evidence 
suggests  that  these  habitats  are  very  important  if  not 
essential  for  many  white-tailed  deer  herds. 

The  same  caution  applies  to  mule  deer.  Whereas 
early  research  and  writing  emphasized  the  impor- 
tance of  early  successional  habitat,  recent  research 
suggests  that,  in  some  areas,  climax  forest  is  the 
most  important  of  the  available  or  potential  habitat 
types  for  maintaining  high  year-round  carrying  ca- 
pacity (Wallmo  and  Schoen  1981 ).  This  is  particu- 
larly true  of  black-tailed  deer  at  the  northern  areas  of 
their  geographic  range. 

Both  mule  and  white-tailed  deer  require  a  diver- 
sity of  plant  species  as  food  at  any  one  time  and 
during  an  entire  year.  Diverse  vegetation  is  also  re- 
quired for  hiding,  escape,  and  thermal  cover.  Be- 
cause of  this,  interspersion  of  several  individual 
forage  and  habitat  types  may  be  more  important  than 
the  occurrence  of  individual  types. 


Population  Measurements. 


Direct  Counts.  Four  basic  methods  are  used  to 
count  mule  and  white-tailed  deer:  aerial  counts, 
drives  or  "complete"  surveys,  flushing  counts,  and 
automobile  counts.  For  direct  counts,  aerial  census  is 
probably  the  only  method  to  be  considered.  These 
counts  are  usually  conducted  during  the  winter 
when  there  is  a  good  counting  background, 
deciduous  trees  have  shed  their  leaves,  and  deer  are 
concentrated  on  limited  winter  ranges. 


Note  differences  in  antler  conformation  between  the  white-tailed  deer  buck  (left  photo)  with  the  points  coming  off  the 
main  beam  and  the  mule  deer  buck  (right  photo)  with  dichotomous  branching. 


Ungulates 


537 


Aerial  counts  from  fixed-wing  aircraft  are  gener- 
ally limited  to  total  counts  of  entire  herds  or  total 
counts  on  trend  areas,  flown  annually.  As  stated  for 
elk,  sex  and  age  ratio  data  cannot  usually  be  ob- 
tained with  the  use  of  fixed-wing  aircraft.  The  use  of 
aerial  censusing  has  been  revolutionized  by  the  heli- 
copter, however.  Sex,  age,  and  area  data  can  now 
be  obtained  that  were  not  possible  from  fixed-wing 
aircraft.  Helicopter  counts  are  more  expensive,  but 
better  quality  data  can  be  obtained. 

Present-day  aerial  census  has  moved  from  total 
counts  to  sample-based  estimates  of  population  size 
(Gill  et  al.  1983).  A  helicopter-quadrat  system  of 
censusing  deer  has  been  developed  in  Colorado 
(Bartmann  1974;  Kufeld  et  al.  1980)  that  consists  of 
1.6-km    (1-mi.  )  quadrats,  randomly  located  in  win- 
ter range  areas  and  permanently  marked  so  the  same 
areas  can  be  flown  each  year.  The  accuracy,  preci- 
sion, frequency,  and  resolution  of  the  quadrat  census 
method  are  currently  being  studied  (Gill  et  al. 
1983). 


Indirect  Counts.  The  primary  indirect  census 
method  is  the  pellet  counts.  Pellet  counts  are  best 
used  in  wintering  areas.  This  technique  is  based  on 
assumptions  of  the  number  of  pellet  groups  depos- 
ited every  24  hours  by  an  individual  deer.  A  winter 
range  is  censused  for  pellet  groups  after  the  deer 
leave  an  area.  All  pellet  groups  within  a  series  of  sta- 
tistically designed  plots  are  then  counted.  The  num- 
ber of  pellet  groups  per  hectare  (acre)  is  multiplied 
by  the  number  of  hectares  (acres)  per  habitat  type, 
resulting  in  an  estimate  of  the  total  number  of  pellet 
groups  per  habitat  type.  Another  factor  that  must 
be  closely  estimated  is  the  number  of  days  that  deer 
have  been  on  the  area  sampled.  Knowing  this,  the 
number  of  pellet  groups  per  hectare  (acre)  for  each 
habitat  type  and  the  daily  defecation  rate,  results  in 
an  estimate  of  the  number  of  deer  using  the  area 
(Neff  1968;  Ryel  1971;  Anderson  et  al.  1972). 

A  review  of  research  on  methods  to  estimate 
total  populations  of  mule  and  white-tailed  deer  over 
large  areas  revealed  that  the  helicopter  quadrat  and 
pellet  group  count  methods  mentioned  above  were 
conceptually  feasible  for  implementation  on  large 
areas  of  the  western  U.S.  No  single  system  can  apply 
to  all  areas  of  the  western  U.S.  Topography,  weather, 
and  vegetation  composition  are  primary  items  that 
dictate  the  system  to  be  used  at  different  localities. 

Helicopter  aerial  surveys  appear  to  be  the 
method  of  choice  wherever  possible,  because  they 
require  less  time,  are  more  accurate  than  fixed-wing 
aircraft  counts,  and  provide  direct  estimates  of  popu- 
lation size.  Aerial  censuses  cannot  be  used,  however, 
in  areas  of  turbulent  winds  or  where  combinations 
of  topography  and  mottled  snow  background  pre- 
clude accurate,  consistent  counts. 


Research  in  Colorado  revealed  that  pellet  group 
counts  on  permanently  marked  plots  and  on  tempo- 
rary plots  yielded  comparable  estimates  of  popula- 
tion size.  Seemingly,  temporary  plots  are  more 
practical  because  they  are  easier  to  implement  and 
are  less  costly  (Freddy  and  Bowden  1983a). 

Discussion.  Because  habitat  and  population  meas- 
urements are  so  similar  for  mule  and  white-tailed 
deer,  the  species  were  combined  in  this  section  to 
save  repeating  much  of  the  information.  Some  of  the 
problems  encountered  while  inventorying  and  moni- 
toring deer  include  a  lack  of  accurate  and  precise 
census  methods,  extreme  difficulty  in  determining 
carrying  capacity  on  seasonal  ranges,  and  the  fact 
that  some  populations  inhabit  areas  of  the  western 
U.S.  where  winters  are  so  mild  that  winter  range 
concentrations  are  unknown  and  year-round  ranges 
are  normal. 

Factors  limiting  deer  populations  in  the  short 
term  are  still  not  well  understood.  For  example, 
available  evidence  indicates  that  mule  deer  declined 
substantially  over  most  of  the  species'  range  from 
the  early  1960s  to  around  1976,  then  began  to  in- 
crease with  no  generally  accepted  explanation  for 
either  the  decline  or  recovery.  Connolly  (1981:243) 
wrote — 

"It  reveals  how  little  control  biologists  and 
managers  have  over  the  deer  they  purport  to 
manage.  Just  as  they  were  powerless  to  halt  the 
decline,  the  biologists  and  managers  now  are 
unable  to  show  in  any  scientifically  acceptable 
way  that  improved  management  put  the  herds 
on  the  road  to  recovery." 

In  spite  of  the  problems,  a  wealth  of  information 
on  techniques  can  be  used  as  is  or  modified  slightly 
to  fit  site-specific  situations.  The  brief  writeups  and 
literature  cited  in  this  section  should  give  guidance 
on  where  to  obtain  additional  information  for  either 
the  biologist  with  no  or  limited  experience  or  the 
experienced  biologist  who  has  need  for  more  de- 
tailed information. 

Moose 

The  moose  (Alces  alces)  is  a  valued  trophy  and 
game  animal  throughout  most  of  its  range.  In  many 
areas  (Isle  Royale,  Denali  National  Park,  Kenai  Moose 
Range),  it  is  also  important  as  a  large,  impressive, 
and  relatively  photogenic  focus  for  tourists  and 
photographers. 

Female  reproductive  performance  in  moose 
populations  varies,  apparently  reflecting  levels  of  nu- 
trition which,  in  turn,  can  be  related  to  habitat  char- 
acteristics. Yearling  pregnancy  rates  are  especially 
variable,  ranging  from  0  to  62%  according  to  various 
studies.  Adult  pregnancy  rates  are  somewhat  less 


538 


Ungulates 


variable,  but  the  twinning  rate  among  adult  females 
in  a  population  may  be  an  especially  good  indicator 
of  habitat  quality  (Franzmann  1978). 

Such  reproductive  performance  indicators 
should  relate  to  quite  specific  habitat  areas,  because 
moose  generally  have  small  home  ranges  (3  to  6  km' 
[1.8  to  3.6  mi.  ]).  In  some  cases,  however,  these  are 
seasonal  home  ranges,  separated  by  migration  corri- 
dors of  up  to  50  km  (30  mi.)  (LeResche  1974; 
Franzmann  1978). 

Moose  are  widely  distributed  in  North  America, 
extending  from  coastal  areas  bordering  the  Arctic 
Ocean  to  the  mixed  deciduous/coniferous  forests  of 
Minnesota  and  Maine  and  as  far  south  as  Utah  in 
the  western  montane  forests.  Dramatic  expansions  in 
the  Rocky  Mountains  have  occurred  in  this  century, 
as  have  apparent  expansions  and  increases  in  tundra 
areas  of  western  and  northern  Alaska.  Climatic 
changes  and  a  large  increase  in  the  proportion  of 
fire-seral  communities  are  two  reasons  most  fre- 
quently cited  for  these  distributional  dynamics  (Kel- 
sall  and  Telfer  1974). 

Average  adult  body  weights  for  females  are  4 14 
kg  (920  lb)  in  Alberta  and  445  kg  (980  lb)  in  Alaska; 
males  attain  weights  of  up  to  585  kg  (1,300  lb) 
(Franzmann  1978).  From  their  size,  moose  are 
clearly  an  ungulate  species  requiring  large  amounts 
of  forage.  Adult  energy  requirements  vary,  but  aver- 
age around  14,000  kcal  per  day  (Belovsky  1978). 
Although  moose  have  large  bite  sizes,  they  have  rela- 
tively low  bite  rates.  In  a  study  in  Alberta,  Renecker 
and  Hudson  (1986)  estimated  that  consumption 
rates  ranged  from  23  g/min  (0.8  oz/min)  in  July  to 
11  g/min  (0.4  oz/min)  in  January. 

In  Alaska,  calves  grow  from  birth  weights  of 
1 1.2-135  kg  (25-30  lb)  to  weights  of  180  kg  (400 


Mature  bull  moose. 


lb )  in  5  months.  Their  size  suggests  that  forage  abun- 
dance and  quality  are  critical  factors  for  moose.  In- 
deed, Belovsky  (1978)  presents  convincing  evidence 
that  moose  almost  always  forage  to  maintain  energy 
input.  They  have  little  room  for  error  if  they  are 
to  achieve  positive  energy  balances. 


Habitat  Requirements  and  Analysis.  Because  of 
the  high  forage  demand  of  moose,  habitat  evaluation 
for  the  species  has  concentrated  on  identifying  forag- 
ing areas  and  measuring  forage  quantity  and  quality. 

Browse  comprises  75  to  80%  of  the  moose  diets 
in  winter  but  becomes  less  important  by  varying 
degrees  in  summer.  Moose  can  break  off  large  stems 
and  branches  and  exploit  the  canopy  up  to  3  m  (10 
ft)  above  ground  level.  Throughout  the  range  of 
the  species,  willows  (Salix  sp. )  are  used  out  of  pro- 
portion to  their  abundance.  Other  highly  palatable 
and  apparently  preferred  plants  include  birch  (Be- 
tula  sp.),  aspen  (Populus  tremuloides),  and  cotton- 
wood  (Populus  sp. ).  When  snow  is  not  too  deep, 
low  shrubs  (such  as  Vaccinium  sp.)  or  sedges 
(Carex  sp.  and  Equisetum  sp. )  may  also  be  used. 

Brassard  et  al.  ( 1974)  examined  several  varia- 
bles potentially  useful  in  measuring  relative  habitat 
value  and  winter  carrying  capacity  in  Quebec.  They 
found  a  direct  relationship  between  degree  of  occu- 
pation and  quantity  of  available  deciduous  stems. 
Later  work  (Polequin  et  al.  1977)  has  shown  that  in 
late  winter,  cover  and  reduced  exposure  from  wind 
may  be  as  important  as  forage  availability  in  the  se- 
lection of  sites. 

In  most  parts  of  their  range,  moose  make  signifi- 
cant use  of  aquatic  plants  in  summer.  Much  of  this 
use  is  concentrated  at  a  time  when  twigs  and  leaves 
of  browse  species  have  peak  nutritional  value.  In 
terms  of  energy  needs,  there  is  no  obvious  advantage 
for  moose  switching  to  aquatics.  However,  Jordan  et 
al.  (1973)  and  Fraser  et  al.  ( 1984)  concluded  that 
sodium  content  is  being  selected  since  winter 
browse  may  have  been  insufficient. 

Similarly,  natural  mineral  licks  are  also  impor- 
tant sources  of  sodium  for  moose  (Tankersley  and 
Gasaway  1983).  Although  nothing  directly  links 
aquatic  forage  and  mineral  licks  to  moose  reproduc- 
tive dynamics,  such  links  can  be  hypothesized,  at 
least  physiologically.  Thus,  habitat  managers  should 
identify  and  maintain  such  features. 

For  many  populations,  riparian  and  floodplain 
communities  are  also  preferred  habitats,  apparently 
of  importance  far  beyond  their  relatively  restricted 
occurrence.  This  is  certainly  true  of  Shiras  moose  (A 
a  shirasi)  in  the  Rocky  Mountains  (Peek  1974)  and 
of  populations  that  extend  into  tundra  areas  (Mould 
1977,  1979). 


Ungulates 


539 


Browse  quality  has  been  determined  by  many 
measurement  techniques  (Oldemeyer  1977a).  On 
the  Kenai  Pensinsula,  Alaska,  deteriorating  range 
quality  has  been  circumstantially  linked  with  declin- 
ing moose  numbers  (Oldemeyer  et  al.  1977a).  Ani- 
mal physiological  factors,  such  as  hair  and  blood 
chemistry,  have  also  been  used  in  measuring  habitat 
quality  (LeResche  et  al.  1974;  Franzmann  et  al. 
1975).  LaPerriere  et  al.  ( 1980)  used  Landsat  images 
and  cluster  analysis  techniques  to  map  moose  habitat 
classes  and  features  in  interior  Alaska.  Through  these 
techniques,  large  areas  could  be  mapped  with  high 
accuracy  and  little  field  checking. 


A  model  has  been  developed  to  evaluate  moose 
habitat  suitability  and  quality  in  Alaska.  The  model 
allows  quality  to  be  estimated  quantitatively,  based 
primarily  on  data  from  studies  in  south-central  and 
interior  Alaska.  A  handbook  is  available  that  would 
help  a  biologist  apply  the  model  (Konkel  1980,  as 
cited  in  Mule  1982).  However,  Mule  (1982)  com- 
pared the  habitat  quality  ratings  generated  by  the 
model  with  those  derived  by  a  team  of  experts,  and 
he  concluded  that  the  model-based  ratings  were 
unacceptably  inaccurate.  A  revision  of  the  model  is 
described  but  not  fully  tested  (Mule  1982). 

Population  Measurements.  Techniques  for  accu- 
rately censusing  moose  need  to  be  perfected.  In 
spite  of  their  large  size,  their  habitats  and  frequent 
use  of  forested  cover  make  them  difficult  to  see  from 
aircraft,  even  helicopters.  Even  though  at  some  times 
of  the  year  some  populations  move  into  more  open 
terrain,  moose  are  not  sufficiently  consistent  in  such 
behavior.  Furthermore,  seasonal  aggregations,  when 
they  do  occur,  do  not  involve  large  segments  of 
the  population. 


Helicopter  surveys,  before  bulls  drop  their  ant- 
lers in  the  fall,  are  useful  in  obtaining  adult  sex  ra- 
tios. Yearling  bulls  may  also  be  reliably  identified 
from  the  air,  but  yearling  females  are  difficult  to 
reliably  identify  (LeResche  and  Rausch  1974).  After 
males  have  shed  their  antlers,  adult  females  may 
be  correctly  distinguished  from  the  ground  by  their 
white  vulvar  markings  (Lent  1974b).  More  detailed 
information  on  the  age  structure  of  populations  can 
only  be  obtained  by  sectioning  teeth  and  reading 
cementum  annuli  from  hunter  kills  or  other  dead  an- 
imals (Sergeant  and  Pimlott  1959). 


Timmermann  ( 1974)  reviewed  other  tried  tech- 
niques in  assessing  relative  population  levels  of 
moose,  ranging  from  track-and-fecal  dropping  counts 
to  infrared  scanning  devices.  All  these  techniques 
had  factors  that  severely  limited  their  usefulness. 


Discussion.  Because  moose  generally  prefer  early 
successional  vegetation,  carrying  capacity  can  be 
improved  by  manipulating  selected  habitats.  These 
manipulations  may  include  mechanical  rehabilitation 
(Oldemeyer  1977b),  prescribed  burns,  and  special 
logging  practices  (Teller  1974).  But  Peek  (1974) 
advised  caution  and  careful  planning  of  such  manipu- 
lations, particularly  in  areas  where  more  than  one 
large  herbivore  species  may  be  present. 

Perhaps  the  most  critical  management  need  for 
moose  is  the  development  of  truly  accurate  and  use- 
ful census  techniques.  Until  this  happens,  biologists 
will  have  difficulty  assessing  the  value  of  various 
management  practices,  including  habitat 
manipulations. 

Caribou 


Despite  these  problems,  most  census  techniques 
still  involve  aerial  surveys,  using  either  transects  or 
intensive  complete  counts  of  small  blocks  of  habitat. 
Transects  have  been  discontinued  in  some  areas 
because  of  their  extremely  low  accuracy  (Timmer- 
mann 1974);  in  other  areas,  they  are  used  only  as 
indexes  of  relative  abundance.  In  the  most  intensive 
tests  of  aerial  counts  to  date,  conducted  over  enclo- 
sures with  known  numbers  of  animals,  LeResche 
and  Rausch  (1974)  found  that  only  about  68%  of 
the  adults  were  observed  under  ideal  conditions, 
with  inexperienced  observers  seeing  only  43%.  Both 
precision  and  accuracy  were  low.  Recent  experience 
with  conditions  similar  to  the  test  area,  along  with 
planning  and  attention  to  details,  was  found  to  be 
essential  in  obtaining  valid  trend  indicators.  Similarly, 
Gasaway  et  al.  (  1985)  described  significant  and 
widely  variable  biases  in  summer  aerial  surveys  of 
moose  where  even  the  direction  of  population  trend, 
except  veiy  large  changes,  could  not  be  reliably 
detected. 


Populations  of  caribou  {Rangifer  tarandus) 
roam  a  variety  of  northern  boreal  habitats  in  North 
America,  from  the  barren  polar  deserts  of  the  Cana- 
dian Arctic  Archipelago  southward,  occupying  much 
of  Alaska  and  extending  into  woodland  and  alpine 
areas  of  southern  Canada.  A  small  remnant  popula- 
tion moves  into  the  contiguous  48  states  in  the  area 
of  the  Idaho  Panhandle  during  part  of  the  year.  The 
species  is  typically  divided  into  two  groups  of  sub- 
species: the  tundra  (cylindricornis)  group  and  the 
woodland  (compressicornis)  group  (Banneld  1962). 
Ecologically,  however,  at  least  four  categories  should 
be  considered: 


( 1 )  Small  groups  that  remain  in  tundra  habitats 
throughout  the  year; 

(2)  Large  migratory  populations,  some  portions 
of  which  typically  use  taiga  habitats  in  the 
winter  months; 


540 


I  'ngulates 


if 


Caribou. 


(3)  Mountain  caribou  that  make  marked  altitudi- 
nal  migrations  seasonally; 

(4)  Woodland  caribou  that  generally  make  only 
short  annual  movements. 

Generally  the  woodland  populations  are  small,  show- 
ing one  of  the  latter  two  strategies.  There  are  excep- 
tions however.  In  Labrador,  populations  of  woodland 
caribou  have  reached  nearly  200,000  animals  and 
migrate  over  200,000  km2  (80,000  mi.2)  of  tundra 
and  taiga  (Juniper  1980;  Parker  1981). 

Reindeer,  domesticated  members  of  the  same 
species,  have  been  introduced  into  North  America, 
primarily  from  Siberia.  Herds  exist  in  Alaska  and 
in  the  Mackenzie  River  Delta  of  Canada.  Management 
as  domesticated  stock  continues  in  some  areas  to 
varying  degrees.  In  other  areas,  reindeer  are  now 
feral,  and  some  have  mixed  and  interbred  with  na- 
tive subspecies. 

Habitat  Requirements  and  Analysis.  Surprisingly, 
few  biologists  agree  on  the  habitat  requirements  of 
caribou,  despite  an  extraordinary  amount  of  research 
and  management  concern  over  the  past  decades. 
Much  controversy  in  recent  years  has  focused  on  the 
importance  and  need  of  lichens  in  the  winter  diet 
of  caribou.  Populations  of  the  high  arctic  polar  de- 
sert and  some  introduced  populations  make  little  use 


of  lichens  and  yet  they  thrive.  Klein  (1982)  argued 
that  these  exceptions  in  no  way  excluded  or  refuted 
the  possibility  that  the  species'  evolution  may  be 
closely  linked  to  a  lichen-based  niche,  unexploited 
by  other  herbivores.  Lichens,  as  winter  forage,  are 
low  in  protein  but  high  in  digestible  carbohydrates. 

Late  winter-spring  fat  deposition  may  occur 
on  high-quality  lichen  ranges.  But  clearly,  caribou, 
like  other  northern  ungulates,  adapt  to  long  periods 
of  low-quality  forage  in  winter  and  indeed  do  not 
normally  maintain  body  weight  in  winter  even  on  a 
relatively  high-protein  diet  (McEwan  and  Whitehead 
1970,  1975).  However,  actual  examples  of  starvation 
or  extreme  malnutrition  among  adult  caribou  in 
winter  or  spring  are  rare,  although  it  has  occurred  in 
some  arctic  island  populations.  Bergerud  (1978:98) 
points  out  that — 

"It  has  not  been  demonstrated  that  an  absolute 
shortage  of  food  has  caused  caribou  declines 
....  I  do  not  know  of  any  studies  where  it 
could  be  shown  that  birth  and  death  rates  were 
altered  due  to  an  absolute  shortage  of  food." 

However,  the  massive  declines  and  die-offs  that  have 
occurred  in  insular  populations  demonstrate  that, 
under  certain  circumstances,  forage  can  be  abso- 
lutely limiting  (Klein  1968). 


Ungulates 


541 


Thing  and  Clausen  (1980)  attributed  a  high 
mortality  among  calves  in  Greenland  to  overuse  and 
depletion  of  lichens  on  winter  ranges.  Leader-Wil- 
liams (1980)  also  attributed  alterations  in  reproduc- 
tive performance  to  habitat  overutilization. 
Bergerud's  analysis  ( 1978),  emphasizing  hunting  and 
predation  as  the  key  mortality  factors  in  caribou 
population  dynamics,  tended  to  ignore  the  possible 
role  of  habitat  quality  and  nutrition  in  mediating 
mortality  factors. 

Most  specialists  have  stressed  the  importance  of 
identifying  and  protecting  the  calving  grounds  of 
migratory  caribou  populations.  Nevertheless,  there 
has  again  been  no  agreement  as  to  what  critical  habi- 
tat requirements  are  linked  with  such  calving 
grounds.  Isolation  from  human  centers  (Lent  1966) 
and  phenological  characteristics  ( Kuropat  and  Bryant 
1980;  Lent  1980)  have  been  suggested  as  important. 
Kelsall  (1968)  believed  that  most  Canadian  calving 
grounds  were  extremely  harsh  with  poor  forage,  and 
stressed  the  low  density  of  predators  as  a  factor. 
Reimers  et  al.  (1983)  showed  that  females  on  early 
greened  calving  grounds  had  less  weight  loss  during 
lactation  than  those  still  on  winter  forage  at  calving 
time. 

As  noted  above,  specific  habitat  requirements  of 
caribou  are  controversial  issues,  particularly  in  win- 
ter ranges  and  calving  areas.  In  addition,  many  now 
believe  that  the  quality  of  winter  forage  and,  within 
broad  limits,  quantity  of  winter  forage  are  of  rela- 
tively little  consequence  in  terms  of  caribou  demog- 
raphy. In  contrast,  managers  of  domestic  reindeer 
continue  to  be  concerned  about  winter  range  condi- 
tion techniques  for  evaluating  these  ranges  (Eriksson 
1980).  Workers  with  woodland  caribou  have  also 
developed  techniques  for  assessing  arboreal  lichen 
abundance  (Van  Daele  and  Johnson  1983).  The  rela- 
tive availability  of  forage,  as  influenced  by  patterns  of 
snow  cover  and  ice  accumulation,  have  also  been  of 
great  interest  in  terms  of  wild  and  domestic  popula- 
tions. These  techniques  are  described  in  Chapter 
8,  Tundra. 

A  broad  concensus  has  developed  that  Rangifer 
habitat  quality  is  presently  best  measured  by  examin- 
ing population  attributes.  Parker  ( 1981 )  measured 
both  physical  and  reproductive  characteristics  in 
females,  consistent  with  an  expanding  population. 
Reimers  et  al.  ( 1983)  noted  several  attributes  they 
believe  highly  correlate  with  spring-summer  range 
quality. 


cover  is  essential  for  accurate  winter  counts.  Berge- 
rud  provided  many  other  recommendations  on  cen- 
sus design  and  execution.  Siniff  and  Skoog  (1964) 
reported  on  the  use  of  stratified  random  sampling 
with  aerial  transects. 

In  the  early  1960s,  specialized  aerial  census 
techniques  for  estimating  the  size  of  migratory  cari- 
bou populations  were  developed  in  Alaska.  These 
techniques  take  advantage  of  specific  behavioral  attri- 
butes of  migratory  caribou:  their  use  of  calving 
grounds  at  relatively  fixed  locations,  where  almost  all 
adult  females  converge  for  a  brief  calving  season, 
and  the  typical  concentration  of  these  females  and 
other  individuals  into  extremely  dense  aggregations 
during  an  even  briefer  post-calving  period. 

Skoog  (  1962)  pioneered  the  calving-ground 
count  method  and  Lent  ( 1966)  first  described  the 
post-calving  aerial  photography  method.  Because 
adult  males  are  largely  segregated  from  cows  and 
widely  dispersed  at  these  times,  both  methods  re- 
quire later  samples  to  obtain  adult  sex  ratios  needed 
to  determine  total  population  estimates.  Thus,  the 
current  technique  has  commonly  been  referred  to  as 
the  "aerial  photo-direct  count  extrapolation" 
method.  Such  ratios  are  best  obtained  from  samples 
at  several  locations  during  the  rutting  season. 

Modern  refinements  and  analyses  of  these  basic 
methods  have  been  described  in  several  Pittman- 
Robertson  (Federal  Aid)  reports  of  the  Alaska  De- 
partment of  Fish  and  Game  (Pegau  and  Hemming 
1972).  These  reports  are  discussed  in  detail  by 
Doerr(1979). 

Canadian  workers  have  continued  to  make  more 
use  of  winter  and  spring  censuses  (Parker  1975). 
Aside  from  using  such  censuses  for  total  population 
estimates,  they  are  universally  important  in  deter- 
mining relative  use  of  winter  ranges;  estimating  rela- 
tive mortality  rates  (Davis  and  Valkenberg  1979); 
and  assessing  vulnerability  to  hunting  pressure,  much 
of  which  occurs  on  winter  ranges. 

Calef  and  Heard  ( 1980)  used  aerial  transects  to 
sample  and  derive  population  estimates  for  three 
caribou  populations.  They  used  a  25%  correction 
factor  for  animals  overlooked,  citing  empirical  evi- 
dence for  a  similar  correction  factor  of  29% .  They 
were  able  to  compare  the  transect  estimate  with  an 
estimated  calving  ground  count  for  the  same  popula- 
tion. The  two  estimates  were  17,225  and  15,884, 
respectively,  for  the  Lorillard  herd. 


Population  Measurements.  The  first  aerial  cen- 
suses of  caribou  relied  primarily  on  winter  and 
spring  transect  data.  Bergerud  (1963)  presented  one 
of  the  best  early  treatments  of  such  aerial  censuses, 
their  problems  and  uses.  He  noted,  as  did  Lent 
( 1966),  that  a  low  but  continuous,  unbroken  snow 


Discussion.  Many  census  techniques  have  been 
developed  for  obtaining  good  numeric  and  composi- 
tion data  for  caribou  populations,  particularly  for 
those  inhabiting  tundra  and  open  woodland  environ- 
ments. The  ability  to  meaningfully  relate  habitat 
quality  and  trend  to  population  dynamics  of  Rangifer 


542 


Ungulates 


continues  to  be  marred  by  a  general  lack  of  agree- 
ment. Sampling  of  reproductive  and  physiological 
attributes  of  individual  animals  holds  the  most  prom- 
ise for  monitoring  populations  and,  indirectly,  the 
status  of  their  seasonal  ranges. 

Pronghorn 

Pronghorn  antelope  (Antilocapra  americana) 
symbolize  the  wide  open  rangelands  of  the  West. 
The  range  of  pronghorn  during  the  early  1 800s  in- 
cluded most  of  the  Great  Plains,  the  high  sagebrush 
steppes  and  grass  valleys  of  the  Great  Basin  states, 
and  parts  of  south-central  Canada  and  northern  Mex- 
ico (Einarsen  1948).  By  the  1920s,  however,  the  for- 
mer densely  populated  ranges  had  been  drastically 
reduced.  Ranges  occupied  by  pronghorn  today  are 
virtually  identical  to  historical  ranges  except  areas 
along  the  Mississippi  River.  Reintroductions  since  the 
1920s  are  largely  responsible  for  increased  popula- 
tions of  this  swift  ungulate. 

Limiting  factors  for  pronghorn  herds  include 
predation;  severe  winters  with  deep  snow;  fences 
that  preclude  seasonal  movements;  and  neonatal 
mortality,  possibly  caused  by  a  lack  of  adequate  nu- 
trition for  pregnant  does  during  the  last  3  months  of 
pregnancy.  Disease  does  not  appear  to  be  a  signifi- 
cant mortality  factor  in  pronghorn  (Yoakum 
1978:107).  Management  concerns  about  pronghorn 
are  related  to  limiting  factors  and  include  coyote  and 
bobcat  predation  on  newborn  fawns;  road  kills  on 
high-speed  roads;  wire  fences  constructed  in  ways 
that  preclude  pronghorn  movements,  especially  in 
winter;  and  decreasing  availability  of  diversified  vege- 
tative ranges. 

Habitat  Requirements.  Pronghorn  typically  oc- 
cupy habitats  characterized  by  low,  rolling,  wide- 
open,  expansive  terrain.  Sites  with  sparse  stands  of 
conifers  may  be  used  at  certain  times  if  understory 
vegetation  allows  distant  visibility  and  rapid  mobility 
(Yoakum  1978). 

Habitat  requirements  for  pronghorn  are  de- 
scribed in  detail  in  Autenrieth  ( 1978),  Yoakum 
(1980),  Kindschy  et  al.  (1982),  and  Allen  et  al. 
(1984). 

Free-standing  water,  available  year-round,  ap- 
pears necessary  for  producing  and  maintaining  high 
pronghorn  densities.  This  water  should  be  available 
every  1.6  to  8.0  km  (1  to  5  mi.).  Studies  by  Sund- 
strom  (1968)  in  Wyoming  indicated  that  95%  of  the 
observed  pronghorn  were  within  a  4.8  to  6.4  km  (3 
to  4  mi. )  radius  of  water.  Highest  densities  of  prong- 
horn occur  in  habitats  averaging  25  to  38  cm  (10 
to  1 5  in. )  of  precipitation  per  year.  Where  prong- 
horn have  been  transplanted  to  areas  of  higher  pre- 
cipitation, production  and  survival  rates  declined. 


•*4Bi 


Mature  pronghorn  buck. 


Pronghorn  also  occur  in  areas  of  lower  precipitation, 
but  population  densities  are  less  (Yoakum  1978). 

Most  pronghorn  ranges  receive  some  snow; 
however,  when  snow  accumulations  exceed  25  to 
30  cm  (10  to  12  in.),  pronghorns  have  trouble  find- 
ing forage.  Prolonged  winters  with  deep  snow  are 
especially  harmful  when  combined  with  factors  such 
as — 

( 1 )  Low  quantities  or  qualities  of  forage, 

(2)  Excessive  wind,  and 

(3)  Man-made  obstacles  that  impede  or  restrict 
free  movement  to  areas  with  less  snow  cover. 

Temperatures  do  not  seem  to  be  a  significant  limit- 
ing factor,  as  pronghorns  can  adapt  to  hot  deserts  or 
alpine  plateau  conditions. 

Quality  and  quantity  of  vegetation  appear  to  be 
the  major  factors  affecting  pronghorn  population 
densities.  In  the  sagebrush-grassland  areas  of  the 
Great  Basin  states,  the  following  vegetative  charac- 
teristics are  found  on  ranges  preferred  by  pronghorn 
(from  Kindschy  et  al.  1982): 

•  ground  cover  averaging  50%  living  vegetation; 

•  general  range  composition  of  40  to  60%  grass, 
10  to  30%  forbs,  and  5  to  20%  browse; 

•  a  variety  of  species,  within  a  vegetative  commu- 
nity, including  5  to  10  species  of  grasses,  20  to 
40  species  of  forbs,  and  5  to  10  species  of 
shrubs; 

•  succulent  plants,  available  in  spring  and  wet 
summers; 

•  open,  rolling  rangelands  having  a  variety  of  veg- 
etative types  rather  than  monotypic  vegetative 
communities; 


Ungulates 


543 


•  low  vegetation  structure  averaging  38  to  6  lcm 
(15  to  24  in.)  in  height. 

Most  of  these  preferred  habitat  qualities  would  also 
apply  to  pronghorn  ranges  east  of  the  Continental 
Divide. 

West  of  the  100th  Meridian,  pronghorn  typically 
inhabit  ranges  characterized  by  low  rolling,  expan- 
sive terrain  and  are  almost  never  observed  for  more 
than  a  few  minutes  at  a  time  where  their  view  was 
restricted  by  terrain  or  other  natural  features  (Allen 
et  al.  1984).  Kindschy  et  al.  (1982)  indicate  that 
areas  with  slopes  less  than  5%  were  best  for  prong- 
horn in  this  region. 

Microhabitats  provided  by  topographic  relief  in 
this  region  apparently  increase  habitat  quality  during 
winter  by  providing  lower  wind  velocities,  less 
snow,  and  less  dense  snow.  During  the  fall  and  win- 
ter, pronghorn  spent  more  time  in  basins  less  than 
1.6  km  (1  mi.)  in  diameter  than  at  any  other  time  of 
the  year  (Prenzlow  et  al.  1968). 

Pronghorn  typically  forage  heavily  on  browse 
and  forbs  and  only  eat  grass  when  it  is  green  and 
succulent.  Thus,  pronghorn  food  habits  do  not  gen- 
erally overlap  those  of  horses  and  cattle.  However, 
pronghorn  food  habits  are  similar  to  those  of  domes- 
tic sheep,  and  pronghorn  may  compete  with  sheep 
for  available  forage. 

Wheat  ( Triticum  aestivum )  is  a  major  portion 
of  the  diet  of  pronghorn  living  near  winter  wheat 
fields  in  Colorado  (Hoover  et  al.  1959).  Heavy  use  of 
wheat  by  pronghorn  has  also  been  reported  from 
Kansas,  Montana,  and  Alberta  (Allen  et  al.  1984). 
Winter  wheat  near  or  interspersed  with  rangeland  is 
assumed  to  improve  the  winter  food  value  for  prong- 
horn if  shrubs  are  present  at  densities  of  75%  or 
less  (Allen  et  al.  1984). 

Fences  on  pronghorn  ranges  west  of  the  100th 
Meridian  may  restrict  movements  and  cause  direct 
injury  or  mortality  (Allen  et  al.  1984).  Fences  may 
cause  significant  adverse  impacts  when  constructed 
across  migration  routes  or  where  they  interfere  with 
daily  movements  to  and  from  water. 

Population  Measurements.  Most  pronghorn  are 
counted  from  the  air,  although  some  ground  counts 
are  used  on  sparsely  inhabited  ranges.  Ground  sur- 
veys are  time-consuming  and  expensive,  but  they 
provide  accurate  data  on  fawn  survival. 

Aerial  counts,  usually  by  fixed-wing  aircraft, 
cover  larger  areas  in  less  time.  Aerial  surveys  are 
conducted  during  the  summer  and  fall  in  some 
states,  and  during  the  winter  in  others.  Winter 
counts  are  usually  flown  to  obtain  total  numbers,  as 
fawns  are  hard  to  distinguish  from  yearlings  in  win- 


Twin  pronghorn  fawns. 


ter,  and  most  bucks  have  shed  their  horn  sheaths  by 
the  time  winter  aerial  counts  are  made. 

If  sex  and  age  ratio  data  are  needed,  counts  are 
made  in  late  summer  or  early  fall.  Flight  patterns 
should  be  either  sample  strips  (usually  0.8  km 
[0.5  mi.]  on  each  side  of  the  aircraft  and  1.6  km 
[1  mi.]  apart)  or  in  randomly  located  (1.6  km" 
[1  mi."])  quadrats.  Most  flying  for  pronghorn  census 
should  be  approximately  46  to  91  m  (150  to  300  ft.) 
aboveground  with  speeds  of  about  112  to  128  km 
(70  to  80  mi.)  per  hour.  Usually  30%  of  the  area  is 
sampled  by  a  strip  census  while  10%  is  sampled  by 
quadrats.  The  quadrat  sample  method  usually  takes 
twice  as  long  as  the  strip  census  (T.M.  Pojar,  pers. 
commun. ). 

Discussion.  The  pronghorn  is  unique  in  that  it  is 
easy  to  observe  during  all  seasons  in  all  its  preferred 
habitats;  therefore,  censuses  are  relatively  easy.  Cen- 
sus techniques  vary,  however,  from  Great  Plains 
grassland  habitats  to  Great  Basin  sagebrush  habitats. 
Inventory  and  monitoring  of  pronghorn  habitats  is 
also  relatively  easy  since  their  habitat  requirements 
are  relatively  similar  throughout  their  range,  have 
been  well-described,  and  can  be  measured  easily. 

Bison 

The  American  bison  (Bison  bison )  today  is 
being  raised  in  predominantly  domestic  situations. 
With  few  exceptions,  free-rar  ging  herds  of  bison  are 
no  longer  found.  Only  in  the  vicinity  of  Wood  Buf- 
falo National  Park  in  Canada  and  in  Yellowstone 
National  Park  in  Wyoming  can  free-ranging  bison 
populations  be  found  that  are  neither  cropped  nor 
hunted.  Free-ranging  herds  in  Utah  and  Alaska  are 
intensively  managed  and  cropped.  In  recent  years, 
the  commercial  value  of  bison  has  increased.  Numer- 
ous private  herds  are  managed  as  livestock,  with 


544 


Ungulates 


American  bison  (bull,  cow,  and  calf). 


meat  being  sold  as  a  specialty  item,  and  hides,  heads, 
and  other  parts  finding  a  ready  market. 

Natural  limiting  habitat  factors  for  the  American 
bison  are  not  of  great  concern  because  most  of  to- 
day's herds  are  intensively  managed,  which  includes 
disease  control  and  selective  culling.  Historically,  the 
grizzly  bear  and  the  wolf  were  probably  the  only 
natural  predators  that  affected  bison  (Meagher 
1978).  Because  of  the  present-day,  low  population 
levels  of  these  two  large  predators,  they  no  longer 
significantly  threaten  free-ranging  bison  herds. 

Large-scale  mortality  from  adverse  winter 
weather  has  been  reported  historically.  Today,  only 
the  herds  in  Yellowstone  National  Park  suffer  any 
significant  winter  losses  (Meagher  1978).  Winterkill 
usually  results  from  the  combined  effects  of  forage 
availability,  climatic  stress,  and  physical  condition  of 
individual  animals.  These  losses  usually  occur  in 
late  winter  and  early  spring  after  prolonged  severe 
winters.  Some  calf  mortality  has  been  reported  dur- 
ing severe  spring  weather. 

Habitat  Requirements  and  Analysis.  The  primary 
habitat  requirements  for  bison  on  public  lands  are 
adequate  forage,  water,  and  space. 

Bison  are  primarily  grazers.  Grasses  furnish  most 
of  their  food  over  the  year;  however,  sedges  make 
up  a  large  portion  of  their  diet  in  both  Wood  Buffalo 
and  Yellowstone  National  Parks.  Bison  seem  quite 
similar  to  cattle  in  their  forage  preferences  and,  as  a 
result,  these  two  animals  often  directly  compete 
with  each  other.  Peden  et  al.  (1974)  compared  the 
diets  of  bison  and  domestic  cattle  in  a  short-grass 
plains  area  and  found  that  bison  prefer  warm-season 
grasses,  feed  in  different  areas,  and  are  less  selective. 
Bison  appear  to  be  able  to  digest  low-protein,  low- 
quality  forage  and  apparently  can  eat  more  of  this 


type  of  forage.  Bison  appear  to  need  water  every 
day. 

If  bison  are  to  be  maintained  as  free-ranging 
animals,  they  will  need  large  areas  to  roam.  How- 
ever, they  can  also  be  successfully  kept  in  small  pas- 
tures as  semidomestic  animals. 

When  available,  bison  use  forested  areas  for 
shade  and  to  escape  insects  (Meagher  1978).  These 
areas  also  furnish  forage  in  deep,  crusted  snow  be- 
cause snow  is  more  loosely  packed  in  forested  areas. 
During  severe  weather,  forest  areas  and  some  topo- 
graphical features  furnish  cover  (Meagher  1973, 
1976). 

The  bison  can  thrive  in  places  where  no  other 
large  ungulate  can.  This  includes  open  valleys  cov- 
ered with  deep  snow  and  subject  to  frequent,  severe 
wind  and  blizzard  conditions,  lasting  up  to  6  months 
(Meagher  1976). 

Population  Measurements.  Bison  population  lev- 
els are  best  determined  by  aerial  censuses  rather 
than  sample  estimates  (Meagher  1978).  These  counts 
are  most  effective  when  flown  during  the  winter. 
Herd  and  calf  production  were  successfully  counted 
through  the  use  of  aerial  photography  in  Wood  Buf- 
falo National  Park  (Stelfox  1976). 

Discussion.  Free-ranging  American  bison  are  found 
on  only  a  few  areas  of  public  domain  land  (Utah 
and  Alaska).  Present-day  herds  are  small,  easily  lo- 
cated and  tracked,  and  easily  censused.  Any  standard 
vegetation  measurement  technique  can  be  used  to 
determine  range  use  and  to  monitor  condition  and 
trend  of  bison  ranges. 

This  species  is  important  from  a  historical  and 
aesthetic  standpoint.  Free-ranging  populations  will 
never  be  large,  but  proper  management  of  present- 
day,  free-ranging  herds  should  not  be  neglected. 

Mountain  Goat 

The  mountain  goat  (Oreamnos  americanus)  is 
one  of  the  least  understood  of  native  North  Ameri- 
can ungulates.  Relatively  few  studies  have  been  con- 
ducted on  mountain  goats  because  of  their  limited 
geographic  range  and  comparatively  low  economic 
importance.  Further,  the  species  has  been  difficult  to 
study  because  its  habitats  are  severe  and  often 
remote. 

Most  studies  of  mountain  goats  are  recent.  A 
large  portion  of  literature  on  the  goat  can  be  found 
in  the  proceedings  of  a  special  symposium  (Samuel 
and  MacGregor  1977)  and  the  Biennial  Symposium 
of  the  Northern  Wild  Sheep  and  Goat  Council  ( He- 
bert  and  Nation  1978;  Hickey  1980;  Bailey  and 
Schoonveld  1982 ). 


Ungulates 


545 


Rocky  Mountain  goat;  mature  billy. 


Habitat  Requirements  and  Analysis.  The  basic 
habitat  requirements  of  mountain  goats  are — 

•  suitable  topography  for  predator  evasion  along 
with  a  sufficient  microclimate, 

•  forages  for  all  seasons, 

•  salt,  and 

•  water. 


For  management  purposes,  winter  ranges,  summer 
ranges,  kidding  areas,  salt  licks,  and  migration  corri- 
dors need  to  be  identified. 


Mountain  goat  winter  ranges  are  characterized 
by  their  lack  of  persistent  or  melt-crusted  snow 
along  cliffs  and  steep  terrain  interspersed  with  vege- 
tation. Suitable  winter  ranges  may  be  (1)  at  lower 
elevations  (Rideout  1974;  Smith  1977),  including 
coastal  areas  (Hjeljord  1973;  Hebert  and  Turnbull 
1977;  Fox  and  Taber  1981)  where  snow  is  less  abun- 
dant and  persistent,  or  (2)  on  relatively  unforested, 
steep,  mostly  south-facing  slopes  where  snow  sheds 
rapidly  (Brandborg  1955;  Hjeljord  1973).  Mountain 
goats  mostly  use  those  portions  of  winter  ranges  that 
are  on  slopes  exceeding  40°  (Smith  1977;  Smith  and 


Raedeke  1982  ).  Usually,  these  slopes  face  south,  al- 
though Adams  and  Bailey  (1980)  described  some 
goats  in  a  southern  population  wintering  on  north 
aspects.  Overhanging  ledges,  caves,  and  abandoned 
mine  entrances  may  provide  additional  shelter  for 
wintering  goats  (Richardson  1971;  Smith  1977). 

In  winter,  goats  become  sedentary  on  tradition- 
ally used  winter  ranges  (Smith  1976;  Adams  and 
Bailey  1980;  Chadwick  1983),  with  forages  on  ledges 
or  benches  within  cliff  areas  or  on  windswept  ridges. 
One  small  population  of  goats  in  a  forested  area 
rarely  foraged  more  than  50  m  ( 165  ft)  from  escape 
terrain  (Smith  1982).  This  may  have  been  an  excep- 
tion, related  to  the  small  numbers  of  goats,  the  poor 
visibility  in  their  forested  habitat,  and  the  presence 
of  large  predators.  In  contrast,  Fox  and  Taber  (1981) 
observed  that  wintering  goats  made  little  use  of 
areas  more  than  300  m  (990  ft)  from  steep  broken 
terrain;  Schoen  and  Kirchoff  ( 1981 )  reported  over 
90%  of  observed  goats  were  within  400  m  (  1,320  ft) 
of  cliffs. 

At  low  elevations,  particularly  near  the  Pacific 
Coast,  a  conifer  canopy  may  benefit  wintering  moun- 
tain goats  by  intercepting  and  redistributing  snow 
and  by  providing  forage,  including  arboreal  lichens. 
In  contrast,  goats  wintering  on  interior  ranges, 


546 


Ungulates 


where  snowfall  is  great,  tend  to  avoid  dense  stands 
of  conifers  that  accumulate  snow  (Adams  and  Bailey 
1980). 

When  snow  is  not  an  overriding  problem,  habi- 
tat selection  is  most  often  determined  by  needs  for 
security  from  predation  and  for  abundant,  nutritious 
forage.  Security  is  probably  a  function  of  distance 
from  escape  terrain,  visibility  afforded  by  the  habitat, 
and  group  size  (Adams  et  al.  1982a;  Risenhoover 
and  Bailey  1985a).  Thompson  (1980)  noted  the  aver- 
age distance  of  goats  from  escape  terrain  on  his 
study  area  was  75  m  (248  ft).  Schoen  and  Kirchoff 
(1981)  found  that  distance  to  cliffs  was  the  most  im- 
portant factor  determining  goat  distribution  and  that 
summering  goats  made  little  use  of  foraging  areas 
over  400  m  (1,320  ft)  from  cliffs.  McFetridge  (1977) 
also  noted  that  most  foraging  by  nursery  bands  was 
within  400  m  (1,320  ft)  of  escape  terrain.  Risen- 
hoover and  Bailey  (1985a)  suggest  that  nanny  goats 
form  large  nursery  bands  on  alpine  tundra,  where 
visibility  is  excellent.  This  may  be  a  strategy  for  re- 
ducing the  risk  of  predation  while  exploiting  forages 
far  from  escape  terrain.  This  strategy  is  only  effective 
where  goat  populations  are  relatively  large  and 
wolves  are  absent. 

Goats  will  exploit  phenological  differences 
among  habitats  to  obtain  the  most  nutritious  forage. 
They  often  move  to  lower  elevations  in  spring,  seek 
green  forage  at  higher  elevations  in  summer,  and 
feed  on  north  aspects  in  late  summer  ( Brandborg 
1955;  Rideout  1974;  Smith  1976;  Bailey  and  Johnson 
1977).  On  some  ranges,  moist  meadows  below 
cirques  or  snowfields  are  attractive  foraging  sites  in 
late  summer. 

Visibility  of  surroundings  seems  important  to 
mountain  goats  (Chadwick  1983).  This  may  limit 
their  use  of  dense  stands  of  conifers  far  from  escape 
terrain.  In  contrast,  goats  often  use  sparse  stands  of 
conifers  (Smith  1976;  Adams  and  Bailey  1980)  and 
have  used  conifers  for  avoiding  people  (Singer  1975). 
The  value  of  conifers  on  summer  range  is  unclear, 
but  may  involve  associated  forage,  available  water  or 
salt,  or  an  attractive  thermal  regime. 

Kids  are  born  on  the  steepest,  most  rugged 
areas  of  a  goat  range.  Based  on  a  few  reports,  kidding 
areas  are  usually  within  winter  ranges  and  used  year 
after  year  (Smith  1976;  Adams  1981). 

Judging  by  the  frequency  and  tenacity  with 
which  goats  use  licks,  these  areas  seem  to  be  very 
important  habitat  resources.  Peak  use  of  licks  occurs 
in  spring  and  early  summer  (Singer  1975).  The  prin- 
cipal attractant  at  licks  probably  is  sodium  (Hebert 
and  Turnbull  1977).  On  a  Colorado  study  area,  goat 
movements  seemed  limited  by  licks  occurring  at 
only  one  end  of  the  area.  These  summer  movements 


expanded  after  a  new  salt  lick  was  established  at 
the  other  end  of  the  area  (Bailey,  J.A.,  unpubl.  data). 


Migration  corridors  are  used  between  winter 
and  summer  ranges  and  when  goats  are  visiting  salt 
licks.  When  crossing  forested  areas  without  escape 
terrain,  goats  repeatedly  use  the  same  trails  (Singer 
1975 ).  Presumably  the  risk  of  unexpected  predator 
attack  can  be  great  in  these  areas,  since  visibility 
may  be  poor  and  the  frequent  occurrence  of  goats 
may  attract  predators.  If  this  is  correct,  reducing 
conifer  cover  to  enhance  visibility  could  improve 
goat  habitat  on  migration  corridors. 

Water  is  not  believed  to  be  a  limiting  factor  on 
most  mountain  goat  ranges  as  these  are  either  in 
moist  climates  or  in  areas  with  persistent  snowfields. 
However,  water  availability  may  restrict  goat  move- 
ments and  habitat  selection  in  southern  ranges 
where  the  species  has  been  introduced. 


Food  habits  of  mountain  goats  have  been  re- 
viewed by  Hibbs  (1966),  Wigal  and  Coggins  (1982), 
and  Adams  and  Bailey  (1983).  Dailey  et  al.  (1984) 
used  tame  animals  to  evaluate  forage  preference. 
Mountain  goats  are  such  adaptable  feeders  that  litera- 
ture on  their  food  habits  allows  few  generalizations. 
Reported  diets  have  varied  depending  on  locations 
and  habitats,  season,  and  snow  conditions.  Grasses 
and  sedges,  forbs,  and  browse  have  each  been  re- 
ported as  abundant  in  the  diet  or  as  preferred  in 
both  winter  and  summer  studies.  Conifers  or  mosses 
and  lichens  have  been  important  winter  forages. 

The  ability  of  mountain  goats  to  exploit  a  wide 
variety  of  forages  may  be  an  adaptation  compensat- 
ing for  their  narrow  habitat  preferences,  especially 
during  winter  (Adams  et  al.  1982b).  By  using  all 
available  forages,  goats  are  able  to  exist  in  small 
groups  on  small,  snow-shedding  areas  where  they  are 
protected  from  predators.  Wintering  goats  seem  to 
choose  areas  on  the  basis  of  a  sedentary  life-style 
rather  than  for  abundant  forage.  In  contrast,  during 
summer  when  nutritional  demands  for  growth  and 
lactation  are  high,  goats  benefit  by  choosing  lush 
stands  of  quality  herbaceous  forage  in  alpine  grass- 
lands and  meadows  (Risenhoover  and  Bailey  1985a). 

Habitat  types  used  or  preferred  by  mountain 
goats  on  winter  and  summer  ranges  have  been  de- 
scribed in  several  studies  in  diverse  locations  (Smith 
1976;  Adams  and  Bailey  1980;  Fox  and  Taber  1981; 
Schoen  and  Kirchoff  1981).  In  these  studies,  habitat 
classifications  have  involved  overstory  vegetation, 
understory  vegetation,  and  terrain  characteristics  (as- 
pect, steepness,  brokenness).  While  goats  use  a  vari- 
ety of  vegetation  types,  their  habitat  selection  has 
been  most  consistently  related  to  steep  terrain. 


Ungulates 


547 


Fox  et  al.  (1982)  developed  a  model  for  predict- 
ing presence/absence  of  mountain  goats  in  a  coastal 
range  of  southeast  Alaska.  The  principal  predictor 
was  distance  to  cliffs.  This  model  apparently  has  not 
been  tested  in  other  areas.  Useful  models  of  goat 
habitat  quality,  yet  to  be  developed,  will  probably 
depend  on  the  following  factors,  in  order  of  impor- 
tance: 

(  1 )    Distance  to  cliffs  or  slope  characteristics; 

(2)  Elevation  and  aspect,  including  available 
ranges  of  these  variables;  and 

(3)  Vegetation  type. 

The  influences  of  these  factors  vary  among  areas 
within  the  geographic  range  of  the  species.  Another 
important  factor  may  be  presence  and  location  of 
salt  licks.  While  such  a  model  may  predict  the  quan- 
tity and  quality  of  seasonal  goat  ranges,  additional 
consideration  may  be  needed  to  identify  migration 
corridors. 

Population  Measurements. 


Relative  Density.  Trend  counts,  especially 
from  aircraft,  are  the  most  commonly  used  methods 
for  monitoring  mountain  goat  populations.  In 
repeated  counts,  more  goats  have  been  observed 
from  helicopters  than  from  fixed-wing  aircraft 
(Ballard  1977;  Nichols  1980). 

Trend  counts  provide  more  than  annual  popula- 
tion indexes.  They  provide  herd  productivity  (age 
ratios)  and  minimum  population  size  indexes  that 
may  be  used  as  bases  for  conservative  harvest  strate- 
gies. However,  the  precision  of  trend  counts  has 
not  been  evaluated.  Many  successive  surveys  of  one 
population  in  one  season  still  need  to  be  done.  With- 
out measuring  precision,  the  significance  of  a  differ- 
ence between  trend  counts  over  2  years  cannot  be 
evaluated.  Likewise,  the  probability  of  detecting  a 
significant  change  in  population  size  from  trend 
counts  in  consecutive  years  is  uncertain. 

A  subjective  appraisal  of  Nichols'  (1980)  data 
from  replicated  aerial  counts  of  local  populations, 
from  July  to  mid-September  and  with  consistent 
weather  conditions,  suggests  that  the  90%  confi- 
dence limits  for  a  single  count  from  a  fixed-wing  air- 
craft might  vary  by  ±  18%  of  the  count.  If  this  is 
correct,  there  is  only  a  25%  probability  of  detecting, 
with  90%  confidence,  an  18%  change  in  population 
size,  based  on  trend  counts  from  consecutive  years. 
This  statement  must  be  used  with  caution  because 
Nichols'  (1980)  study  was  not  designed  to  measure 
precision  in  this  way,  and  precision  may  vary  greatly 
among  local  populations.  However,  Nichols'  data 
illustrate  that  aerial  trend  counts  of  mountain  goats 
may  not  be  sensitive  to  fairly  significant  changes 


in  goat  abundance,  even  when  flying  and  observing 
conditions  are  selected  for  consistency.  The  possibil- 
ity that  ground-based  counts  will  be  equally  impre- 
cise should  be  evaluated. 

In  mostly  forested  areas,  early  spring  counts  of 
goats  on  winter  ranges  have  been  favored  ( Smith 
1976;  Bone  1978).  Counts  are  best  made  after  most 
snow  has  melted  and  goats  are  easier  to  observe,  but 
before  the  animals  leave  their  winter  ranges.  Succes- 
sive trend  counts  may  be  difficult  within  this  short 
period. 

Midsummer  has  been  preferred  for  trend  counts 
in  areas  where  goats  form  large  bands  on  alpine  tun- 
dra (Ballard  1977;  Hall  1977;  Nichols  1980).  Large 
bands  may  develop  after  the  kidding  season  and  can 
more  easily  be  observed  after  most  snowfields  have 
melted.  Nichols  (1980)  preferred  overcast  skies,  soft 
light,  and  no  turbulence  for  these  aerial  counts. 
Adult  male  goats,  which  are  more  solitary  than  other 
sex-age  classes,  may  be  under-represented  in  counts 
taken  in  midsummer  ( Risenhoover  and  Bailey  1982a; 
Foster  1982). 

Census.  A  known  minimum  population  of 
mountain  goats  may  be  determined  by  using  the 
largest  number  of  goats  seen  in  several  successive 
counts.  If  animals  are  classified,  the  cohort-comple- 
tion method  may  provide  a  larger  known  minimum 
population.  With  this  method,  the  largest  number  of 
goats  seen  in  each  age  class  (rarely,  sex-age  class) 
is  determined  and  the  numbers  summed  (Smith 
1976).  Known  minimum  populations  are  used  as 
bases  for  conservative  harvest  strategies. 

Use  of  the  highest  count  from  several  successive 
counts  within  a  season  may  enhance  precision  for 
detecting  among-years  trends  in  goat  numbers  (year- 
to-year  variation  among  highest  counts  may  be  less 
than  among  single  counts  made  each  year).  In  prac- 
tice, the  precision  of  the  annual  highest  count 
method  has  been  evaluated  subjectively.  Accuracy  is 
unknown,  but  actual  populations  must  be  underesti- 
mated in  almost  every  case. 

Adams  and  Bailey  (1982)  used  marked  goats  and 
the  Peterson  Estimator  to  census  a  Colorado  herd. 
With  this  method,  accuracy  will  be  enhanced  if  ani- 
mals from  all  herd  segments  are  marked.  Adult  males 
and  females  should  be  marked  proportionally  to 
their  occurrence  in  the  herd  (Risenhoover  and  Bai- 
ley 1982a).  The  possibility  of  marking  goats  in  more 
than  one  geographic  segment  of  the  herd  should  also 
be  considered.  The  second  sample  used  to  estimate 
the  proportion  of  marked  goats  in  the  herd  should 
be  based  on  observations  throughout  the  study  area. 

Marking  a  large  proportion  of  a  goat  herd  will 
seldom  be  possible.  Consequently,  the  precision  of  a 
population  estimate  will  be  poor  and  confidence 


548 


Ungulates 


limits  (Bailey  1951)  high  whenever  a  single  second 
sample  is  used.  Precision  can  be  enhanced  indefi- 
nitely by  using  numerous  second  samples,  resulting 
in  replicated  population  estimates.  With  replications, 
the  population  is  estimated  by  the  mean  of  succes- 
sive estimates,  and  confidence  limits  are  based  on 
the  number  and  variance  of  the  estimates,  using 
common  t-statistics. 

Animal  Condition.  Foster  (1978)  reviewed  the 
concepts  of  animal  condition  and  population  quality 
in  relation  to  mountain  goats.  Presumably,  measures 
of  animal  condition  commonly  used  for  ungulates, 
such  as  body  size,  growth  rates,  skeletal  ratios,  and 
measures  of  fat  reserves,  can  be  applied  to  goats. 
However,  these  methods  have  seldom  been  used, 
partly  because  large  samples  of  goat  carcasses  have 
seldom  been  available.  It  would  be  useful  to  obtain 
and  report  such  data  so  that  inter-  and  intrapopula- 
tion  variation  can  be  evaluated. 

Foster  (1978)  evaluated  horn  dimensions  as 
possibly  useful  measurements  of  population  quality 
in  mountain  goats.  Results  were  largely  inconclusive. 
The  most  useful  horn  dimension  measurements  may 
be  (1)  horn  growth  in  the  first  2  years  of  life  (which 
can  be  measured  in  goats  over  2  years  old  because 
the  second  horn  annulus  is  distinct)  and  (2)  a  ratio 
of  horn  length  to  ear  length  for  kids  and  yearlings, 
when  data  are  collected  in  the  same  month  each 
year.  For  2-year  old  goats,  at  least,  horn  lengths  are 
different  between  sexes.  Thus,  populations  should  be 
compared  by  using  data  from  one  sex  for  each  com- 
parison. Lastly,  intrapopulation  variation  of  horn 
growth  may  be  so  large  that  the  horn  measurements 
for  comparing  populations  will  be  invalid. 


distinguished  by  their  advanced  molt  in  midsummer 
(Nichols  1980)  but  may  be  under-represented  in 
samples  taken  during  this  time  (Foster  1982;  Risen- 
hoover  and  Bailey  1982a). 

Small,  randomly  taken  samples  from  a  goat  herd 
will  provide  highly  variable  estimates  of  age  struc- 
ture. Thus,  biologists  should  not  rely  on  classification 
data  unless  a  large  proportion  of  the  herd  can  be 
classified. 

When  data  are  limited,  the  simplest  age  ratio  is 
the  kid  to  older  goat  ratio  (where  older  goats  in- 
clude yearlings,  2-year  olds,  and  adults  of  both 
sexes).  This  ratio  is  not  a  precise  indicator  of  cur- 
rent reproductive  success  since  variation  in  the  sex 
ratio  and  in  the  numbers  of  nonbreeding  older  goats 
will  distort  the  ratio.  However,  this  ratio  has  re- 
flected herd  age  (recently  transplanted  herds  having 
higher  ratios)  and  winter  weather  (Bailey  and  John- 
son 1977;  Adams  and  Bailey  1982).  The  ratio  has  also 
been  used  to  estimate  numbers  of  kids  in  estimated 
herd  sizes  (Adams  and  Bailey  1982). 

With  more  detailed  classifications  of  goats  in 
local  herds,  population  models  may  be  constructed. 
Nichols  (1980)  used  aerial  surveys  in  spring  to  esti- 
mate the  proportion  of  yearlings  and  aerial  surveys 
in  summer  to  estimate  herd  size  and  the  proportions 
of  kids  and  adult  males.  He  then  used  data  from  suc- 
cessive years  to  simulate  herd  dynamics.  Simulations 
provided  estimates  of  the  numbers  of  2 -year  olds  in 
a  herd  each  year.  By  subtraction,  the  number  of 
breeding-age  females  was  estimated,  and  a  kid  to 
breeding-age  nanny  ratio  determined.  Nichols  sug- 
gested this  ratio  will  be  correlated  with  population 
trend.  The  accuracy  of  this  approach  was  untested. 


Population  Structure.  Mountain  goat  females 
are  not  expected  to  breed  before  they  are  2V2  years 
old,  nor  kid  before  they  are  3  years  old.  Conse- 
quently, classification  by  sex  and  by  four  age  classes 
is  desirable,  especially  if  a  ratio  of  kids  to  breeding 
nannies  is  to  provide  an  index  of  current  production. 
However,  such  detailed  classification  is  practical 
only  in  intensive,  local  studies.  Not  only  are  year- 
lings and,  especially,  2-year  olds  difficult  to  distin- 
guish in  some  situations,  but  a  lack  of  strong  sexual 
dimorphism  further  complicates  classification  Ni- 
chols (1980),  Chadwick  (1983),  and  Smith  (unpubl. 
ms.)  have  provided  descriptions  of  the  eight  sex-age 
classes  of  goats  and  Foster  ( 1978 )  illustrated  their 
horn  characteristics. 

The  difficulty  in  distinguishing  age  classes  of 
goats  varies  with  season  and  with  the  distance  and 
duration  of  observation.  Often,  midsummer  is  best 
for  distinguishing  kids  and  yearlings  from  older 
goats.  Kids  usually  become  more  visible  in  July,  and 
yearlings  are  especially  easily  distinguished  by  size 
alone  until  about  September.  Adult  males  may  be 


*j&£ 


Rocky  Mountain  goat;  nanny  and  kid. 


Ungulates 


549 


Behavioral  Indexes.  Petocz  (1973)  suggested 
that  conflicts  (agonistic  behavior)  among  mountain 
goats  may  indicate  the  degree  of  resource  depriva- 
tion. Risenhoover  and  Bailey  (1985a)  speculated  that 
average  group  size  may  be  correlated  with  forage 
abundance  and  continuity. 

Discussion.  A  primary  management  concern  is 
overharvest  of  mountain  goats  when  access  to  their 
habitats  has  been  improved  by  new  roads  (Foster 
1977;  Kuck  1977a,b;  Adams  and  Bailey  1982;  Chad- 
wick  1983).  Appropriate  levels  of  harvest  could  be 
achieved  if  (1)  annual  data  on  population  size  and 
productivity  are  known,  (2)  harvest  levels  can  be 
regulated  on  a  local  basis,  and  ( 3 )  illegal  kill  can  be 
controlled.  However,  these  requirements  are  often 
impractical. 

The  effects  of  removing  vegetation  from  moun- 
tain goat  habitat,  such  as  by  logging  or  fire,  are  un- 
clear. In  addition,  the  effects  may  differ  between 
coastal  and  interior  ranges  (Hebert  and  Turnbull 
1977),  especially  on  winter  ranges,  where  detrimen- 
tal effects  would  most  likely  occur.  Near  the  Pacific 
Coast,  for  example,  forest  removal  may  be  detrimen- 
tal to  winter  ranges. 

Interior  winter  ranges  often  support  sparse 
stands  of  trees  or  shrubs  that  are  used  for  forage. 
The  steep  slopes  of  these  winter  ranges  are  often 
used  by  mountain  goats  for  their  snow-shedding 
characteristics  (Kuck  1977a;  Adams  and  Bailey 
1980).  Removing  forage  in  these  areas  may  affect 
forage  resources,  shelter,  or  snow-shedding 
characteristics. 

Some  goat  populations  are  small  and  isolated 
(Smith  and  Raedeke  1982;  Smith  1982),  and  gene  flow 
among  such  populations  seems  infrequent.  Any  hu- 
man activities  (e.g.,  harvest,  habitat  alteration)  may 
inhibit  goat  movements,  especially  among  males, 
thus  increasing  possibilities  for  inbreeding. 

Mountain  goats  are  known  to  travel  long  dis- 
tances to  visit  natural  licks  or  artificially  established 
salt  licks  (Singer  1975;  Hebert  and  Turnbull  1977; 
Thompson  and  Guenzel  1978;  Adams  et  al.  1982). 
The  possibility  that  a  sparse  distribution  of  licks  may 
limit  the  distribution  or  productivity  of  goats  has 
not  been  tested. 

Muskox 

The  muskox  (Ovibos  moschatus)  is  a  boreal 
species  whose  numbers  and  distribution  were  greatly 
reduced  in  the  1800s  and  early  1900s.  Since  then, 
the  management  of  muskoxen  has  been  a  conserva- 
tion success  story.  Hunting  prohibitions  and  success- 
ful transplants  have  restored  populations  to  the  point 
where  limited  harvests  are  now  possible  in  Alaska, 
Canada,  and  Greenland.  The  species  has  also  been  a 


focus  of  study  in  efforts  to  produce  a  new  domestic 
animal  for  rural  economies  of  circumpolar  nations 
(Lent  1971,  1978;  Coady  and  Hinman  1984;  Wilkin- 
son and  Teal  1984). 

Muskoxen  live  20  years  or  more  and  were  once 
thought  to  have  extremely  low  reproductive  rates. 
New  evidence  demonstrates  that  while  rates  are 
relatively  low  compared  with  ungulates  that  may 
produce  twins  or  triplets,  they  are  considerably 
higher  than  was  once  thought  (Lent  1978). 

Habitat  Requirements  and  Analysis.  The  native 
range  of  muskoxen  is  primarily  in  the  high  arctic, 
a  zone  distinguished  by  very  low  precipitation,  shal- 
low snow  cover,  and  patchy  distribution  of  snow 
from  winds.  Many  areas  where  muskoxen  have  been 
introduced  are  subject  to  much  greater  winter  snow 
accumulation  or  maritime  climates,  having  icing 
conditions  in  winter  or  spring  (Lono  I960;  Spencer 
and  Lensink  1970;  Lent  1974b,  1978).  Thus  identifica- 
tion, monitoring,  and  protection  of  suitable  winter 
habitat  have  been  important  and  will  be  increasingly 
important  as  introduced  populations  expand. 

Riparian  areas  with  shrubs  are  preferred  summer 
habitats  in  Alaska  (Lent  1978;  Robus  1984).  Wet  hab- 
itats with  sedges  (Carex  sp.)  also  receive  heavy  use 
in  summer.  Both  Parker  and  Ross  (1976)  and  Thing 
(1984)  contend  that  such  meadows  are  important  to 
this  species  for  weight  gain  and  development  of  fat 
reserves  in  summer.  In  general,  caribou  and  musk- 
oxen  rarely  compete  in  areas  where  they  occur  to- 
gether (Wilkinson  et  al.  1976;  Thomas  and  Edmonds 
1984). 


Musk-ox  herd  on  tundra. 


550 


Ungulates 


Few  measurements  of  muskox  habitat  have  con- 
clusively related  to  habitat  quality.  The  importance 
of  wet  meadows  and  riparian  strips  has  already  been 
referred  to.  In  the  arctic,  remote  sensing  techniques 
provided  opportunities  to  map  such  features  on  a 
large  scale  (U.S.  Department  of  the  Interior,  Fish  and 
Wildlife  Service  1983).  Home  ranges  that  provide  a 
variety  of  habitats  in  summer  seem  important  be- 
cause muskoxen  can  take  advantage  of  successive 
peaks  in  forage  quality. 

In  winter,  forage  availability  is  the  key  habitat 
requirement.  Regions  with  low  precipitation  and  low 
and  stable  temperature  regimes  are  crucial.  Methods 
for  monitoring  and  mapping  tundra  vegetation  and 
for  mapping  and  measuring  snow-cover  features  have 
been  treated  in  Chapter  8,  Tundra. 


Cementum  annuli  in  incisors  have  been  used  to 
determine  age  at  death.  Lent  ( 1 974c )  concluded  that 
this  technique  was  fairly  reliable.  However,  Gron- 
quist  and  Dinneford  (1984),  working  with  18 
known-age  specimens,  half  of  which  were  from  ani- 
mals raised  in  captivity,  concluded  that  the  tech- 
nique underestimated  true  age  and  had  low 
reliability.  Lent  (1974c)  and  Parker  et  al.  (1975) 
have  used  femur  marrow  lipid  analyses  to  assess  lev- 
els of  malnutrition  and  starvation  in  populations. 

Discussion.  The  most  important  management  con- 
sideration for  muskoxen  is  the  identification  and 
protection  of  suitable  winter  habitat.  This  must  be 
accomplished  before  muskoxen  are  transplanted. 
Protection  of  such  areas  of  stable  winter  conditions 
and  forage  availability  is  crucial  to  long-term  man- 
agement success  (Thomas  et  al.  1981). 


Population  Measurements.  Since  the  early  studies 
by  Vibe  (1958,  1967)  and  Tener  (1965),  through  the 
work  of  Spencer  and  Lensink  (1970),  up  to  the  most 
recent  censuses  (Lassen  1984),  the  total  count 
method  has  been  generally  used  in  obtaining  muskox 
population  estimates.  The  reasons  for  this  are  sev- 
eral: muskoxen  occur  in  relatively  low  numbers; 
they  are  large  and  conspicuous  in  typically  treeless 
habitats;  and  in  winter,  they  are  highly  clumped  in 
their  distribution  and  comparatively  sedentary.  In  a 
few  cases,  sampling  on  transects  has  been  used  to 
estimate  populations  (Gunn  et  al.  1984). 

Total  counts  provide  little  opportunity  to  assess 
sample  error  except  through  successive  counts  over 
short  periods.  Spencer  and  Lensink  (1970),  working 
from  snow  machines  on  Nunivak  Island  where  snow 
conditions  lead  to  extreme  clumping  of  the  popula- 
tion, believed  they  counted  essentially  the  entire 
population. 

Such  ground  counts  and  even  low-level  aerial 
surveys,  especially  with  helicopters,  are  also  useful  in 
obtaining  sex  and  age  composition  data.  After  some 
training,  an  observer  can  accurately  identify  the  sex 
of  all  animals,  1  year  and  older,  and  determine,  by 
year,  the  age  of  animals  up  to  4  years  old  (Allen  1913; 
Smith  1976).  Low-level  aerial  photography  has  also 
been  used  for  this  purpose  (U.S.  Department  of  the 
Interior,  Fish  and  Wildlife  Service  1983). 

Sex  and  age  structure  of  groups  may  be  deter- 
mined in  the  field  by  experienced  observers.  This  is 
done  by  referencing  horn  development  with  the 
sexual  dimorphism  associated  with  horn  boss  and 
curvature.  Adults  may  be  aged  up  to  the  fifth  year. 
Line  drawings  in  Smith  ( 1976)  and  photographs 
in  Pohle  (  1981 )  serve  as  useful  points  of  reference. 
Calf  and  yearling  counts  can  generally  be  obtained 
with  ease,  although  tight  bunching  behavior  may  be 
a  problem  when  observing  young  calves. 


Mountain  Sheep 

The  mountain  sheep  comprise  two  species,  Ovis 
canadensis  and  O.  dalli,  the  bighorn  and  thinhorn 
sheep,  respectively  (Geist  1971a).  Mountain  sheep 
are  distributed  in  mountains  and  rugged  terrain  from 
Mexico  to  Alaska.  Their  recent  research  and  manage- 
ment is  reported  in  the  annual  proceedings  of  the 
Desert  Bighorn  Council  and  the  Biennial  Symposium 
of  the  Northern  Wild  Sheep  and  Goat  Council. 

Habitat  Requirements  and  Analysis.  The  evolu- 
tionary history,  anatomy,  and  behavior  of  mountain 
sheep,  especially  bighorn  sheep,  suggest  that  they 
are  best  adapted  to  living  in  moderately  large  social 
groups  in  open  habitats,  on  or  near  steep,  rugged 
terrain.  Specific  habitat  resources  for  mountain  sheep 
are — 

•  forage  of  adequate  quality  and  of  adequate  den- 
sity and  continuity  to  support  a  large  group; 

•  water; 

•  salt  licks; 

•  good  visibility  for  predator  detection  and  for 
visual  communication; 

•  steep,  usually  rocky,  escape  terrain;  and 

•  a  suitable  thermal  environment. 

Favorable  combinations  of  these  resources  exist  in  a 
patchy  distribution.  Consequently,  the  annual  ranges 
of  mountain  sheep  populations  tend  to  consist  of 
up  to  seven  or  more  seasonal  ranges  (Geist  1971a) 
and  their  connecting  migration  corridors  (Ough  and 
deVos  1984).  Commonly,  an  annual  range  includes 
separate  winter,  spring  (green-up),  and  summer 
ranges  for  rams  and  ewe-juvenile  groups,  and  a  lamb- 
ing area.  Seasonally  critical  areas,  generally  within 
these  ranges,  provide  salt  licks  or,  in  arid  country, 
water  sources.  When  seasonal  ranges  are  large  and 


Ungulates 


551 


dispersed  over  a  large  range  of  elevation,  sheep  have 
options  for  avoiding  harassment  and  for  selecting 
elevations  and  aspects  to  improve  their  foraging  and 
thermal  environments. 

Mountain  sheep  are  adaptable  foragers.  Their 
food  habits  vary  greatly  among  regions  and  seasons 
(Todd  1972;  Browning  and  Monson  1980;  Cooper- 
rider  et  al.  1980).  In  northern  areas,  grasses  tend  to 
dominate  sheep  diets,  especially  during  spring,  sum- 
mer, and  fall,  and  modest  amounts  of  browse  are 
eaten  in  winter  when  snow  inhibits  use  of  herbs.  By 
contrast,  desert  bighorn,  and  Rocky  Mountain  big- 
horn at  low  elevations,  may  browse  abundantly  on 
twigs  or  leaves  in  any  season.  While  species  compo- 
sitions of  mountain  sheep  diets  are  inconsistent, 
three  characteristics  are  common  to  quality  forage 
resources:  abundance,  continuous  distribution,  and 
low  stature.  Forage  abundance  and  continuity  are 
needed  to  support  at  least  modest  group  sizes;  to  al- 
low individuals  within  groups  to  be  dispersed,  thus 
alleviating  social  interference  with  foraging  and  en- 
hancing predator  detection;  and  to  allow  efficient 
rates  of  forage  intake  (Dale  and  Bailey  1982).  Low- 
growing  forage  does  not  interfere  with  vision,  thus 
allowing  for  predator  detection  and  visual  communi- 
cation among  sheep.  Grasses  are  the  epitome  of 
abundant,  continuous,  low-growing  forage,  but  dense 
stands  of  forbs  or  low  shrubs  may  also  be  excellent 
forage. 


by  varying  their  daily  activity  pattern  and  by  select- 
ing favorable  environments.  They  may  use  elevation, 
aspect,  terrain  wind-shadow,  tree  shade  (especially 
in  open  forests  near  or  on  steep  terrain),  and  rock 
overhangs  or  caves  to  advantage  (Simmons  1969). 
Often,  selection  of  elevation  or  aspect  provides  ad- 
vantages of  forage  quality  or  snow  scarcity,  as  well  as 
thermal  advantages  (Hebert  1973;  Shannon  et  al. 
1975). 

Land  areas  are  evaluated  as  habitat  for  mountain 
sheep  when  ranges  are  not  occupied  by  sheep,  but 
are  candidates  for  a  transplanted  herd  or  candidates 
for  habitat  improvement  to  encourage  range  expan- 
sion by  an  established  herd.  Land  areas  can  also  be 
evaluated  when  ranges  are  occupied  by  sheep  if 
range  condition  and  trend  are  to  be  monitored  or 
the  qualities  of  ranges  are  to  be  compared  so  disrup- 
tive developments  or  habitat  degradations  can  be 
directed  toward  less  valuable  sheep  range. 

If  the  seasonal  ranges,  water  sources,  mineral 
licks,  lambing  areas,  and  migration  corridors  for  an 
existing  herd  are  unknown,  first  consideration  should 
be  given  to  a  telemetry  study  for  locating  these 
areas.  This  information  will  be  needed  if  range  con- 
dition is  to  be  monitored.  It  will  also  be  needed  if 
good  range  condition  is  to  be  maintained  or  pro- 
tected from  development  or  habitat  manipulation. 


Water  and  mineral  licks  are  seasonally  critical 
resources.  Sodium  and  magnesium  seem  to  be  key 
components  of  licks  used  by  ungulates,  especially 
during  plant  growth  season  in  moist  climates  (Jones 
and  Hanson  1985).  Water  sources  and  mineral  licks 
can  be  created  and  maintained  in  sheep  manage- 
ment. Mountain  sheep  are  especially  favored  when 
these  sources  are  located  in  open  areas  near  escape 
terrain,  when  human  activity  is  kept  at  a  distance 
and  livestock  are  fenced  from  the  site. 

The  predator-evasion  strategy  of  mountain  sheep 
requires  visibility  and  escape  terrain  (Risenhoover 
and  Bailey  1985b).  These  habitat  characteristics  are 
especially  important  to  ewes  with  lambs.  Rams  are 
more  prone  to  use  areas  lacking  these  security  fac- 
tors. In  addition,  sheep  may  forego  security  to  use 
critical  water  sources,  salt  licks,  or  necessary  migra- 
tion corridors.  Visibility  and  escape  terrain  have 
compensating  effects  on  security.  Poor  visibility  is 
acceptable  on  steep  terrain;  sheep  will  forage  farther 
from  escape  terrain  when  visibility  is  good. 

Lambing  areas  are  usually  the  largest  areas  of 
rugged,  steep  terrain  within  the  annual  range  of  a 
population.  They  are  often  near  water. 

In  both  hot  and  cold  climates,  mountain  sheep 
will  avoid  extreme  temperatures  and  wind  chills 


Range  Assessments.  Hansen  ( 1980),  Grunigen 
(1980),  and  Holl  (1982)  published  methods  for  eval- 
uating land  areas  as  habitat  for  bighorn  sheep.  These 
methods  have  not  been  verified  with  test  applica- 
tions outside  the  areas  where  they  were  developed. 
In  these  methods,  habitat  components  indicating 
high  value  as  sheep  range  include — 

( 1 )  Abundance  of  steep  escape  terrain; 

(2)  Abundance  of  certain  vegetation  types  that 
are  low-growing  or  open  and  known  to  pro- 
vide sheep  forage,  especially  where  these 
types  are  on  or  near  escape  terrain; 

(3)  Abundance  of  southerly  aspects  (this  may 
duplicate  information  from  vegetation  types); 

(4)  Annual  precipitation  and  presence  of  reliable 
water  sources  in  arid  areas; 

(5)  Absence  of  livestock  and  other  human  activi- 
ties and  development  that  would  threaten 
sheep  or  sheep  habitat. 

Ratings  of  such  habitat  components  would  not 
likely  be  combined  in  a  simple  additive  model.  Ab- 
sence of  a  key  factor,  such  as  escape  terrain  or 
water,  may  severely  limit  the  value  of  a  range,  re- 
gardless of  high  ratings  for  other  factors. 


552 


Ungulates 


Desert  bighorn  sheep  on  escape  terrain. 


When  the  objective  of  range  assessment  is  to 
monitor  habitat  condition  and  trend,  data  on  habitat 
condition  should  be  obtained  primarily  from — 

( 1 )  Areas  known  to  be  used  by  sheep; 

(2)  Areas  on  or  within  250  m  (825  ft)  of  steep, 
rocky  terrain; 

(3)  The  most  likely  migration  corridors  between 
seasonal  ranges  (shortest  routes,  with  good 
visibility,  over  escape  terrain  or  along  ridges); 
and 

(4)  Areas  including  potential  water  sources  or 
mineral  licks. 

Habitat  Component  Assessments.  Escape 
terrain  is  the  most  consistent  feature  of  mountain 
sheep  habitat.  However,  most  biologists  have  identi- 
fied escape  terrain  subjectively,  usually  describing 
it  as  steep  (>  80%  slope),  broken,  or  rocky  terrain, 
usually  including  cliffs.  Beasom  et  al.  (1983)  offered 
a  method  for  quantifying  land  ruggedness.  Mc- 
Collough  et  al.  (1980)  found  the  minimum  height 
and  length  of  cliffs  used  as  escape  terrain  were  8  m 
(26  ft)  and  220  m  (726  ft),  respectively.  Lambing 
areas  are  steeper  and  larger  than  other  acceptable  es- 
cape terrain.  The  value  of  escape  terrain  will  be  di- 
minished by  visibility-obstructing  vegetation  or  by 
accumulation  of  snow  on  north  aspects. 


The  most  important  habitats  for  mountain  sheep 
are  on  or  near  escape  terrain.  Rams,  rather  than 
ewes  and  juveniles,  and  sheep  in  large  groups,  rather 
than  small,  will  use  areas  farther  from  escape  terrain. 
However,  out  of  1 3  studies  of  bighorn  sheep  sur- 
veyed, only  one  showed  more  than  10%  of  the  ob- 
served groups  of  sheep  over  250  m  (825  ft)  from 
escape  terrain  (Wakelyn  1984).  This  review  indi- 
cates that  sheep  are  more  restricted  to  escape  ter- 
rain than  suggested  by  Van  Dyke  et  al.  ( 1983).  Until 
more  definitive  studies  are  reported,  range  assess- 
ment should  emphasize  areas  on  or  within  250  m 
(825  ft)  of  escape  terrain. 

Visibility  is  an  important  habitat  component  for 
mountain  sheep.  It  allows  for  predator  detection, 
visual  communication  among  sheep,  and  efficient  for- 
aging. In  addition,  open  habitats  with  good  visibility 
may  provide  abundant  forage.  In  assessing  condition 
and  trend  of  sheep  habitats,  some  measurement  or 
index  of  visibility  should  be  monitored.  However, 
few  objective  methods  for  measuring  visibility  have 
been  applied  to  sheep  habitat,  and  no  standards  for 
acceptable  levels  of  visibility  have  been  developed. 
Most  often,  visibility  has  been  evaluated  by  rating 
vegetation  types  or  tree  or  shrub  crown  cover  (with 
the  more  open  types  or  overstories  rated  as  better 
sheep  habitat).  Risenhoover  and  Bailey  (1985b)  esti- 
mated visibility  as  the  percentage  of  the  compass 


Ungulates 


553 


over  which  an  object,  1  m  (3.3  ft)  tall,  could  be 
seen  at  40  m  ( 132  ft).  A  more  objective,  but  more 
expensive,  method  would  involve  using  a  density 
board  (Gysel  and  Lyon  1980). 

In  arid  ranges,  the  distribution  of  water  may 
limit  bighorn  habitat.  Best  water  sources  are  near  es- 
cape terrain  in  areas  allowing  good  visibility,  located 
or  fenced  so  livestock  competition  or  human  activity 
did  not  discourage  use  by  sheep.  In  dry  seasons, 
most  desert  bighorn  have  been  observed  within  1.6 
to  2.4  km  (0.96  to  1.4  mi.)  of  water  (Leslie  and 
Douglas  1979;  Elenowitz  1984). 

Forage  resources  for  mountain  sheep  have  been 
evaluated  at  two  levels.  In  the  simplest  approach, 
vegetation  types  have  been  numerically  rated  as 
forage  resources,  based  on  local  experience  (Hansen 
1980;  Van  Dyke  et  al.  1983).  These  ratings  may  be 
influenced  not  only  by  forage  values,  but  also  by 
expected  visibility  and  proximity  to  escape  terrain 
for  each  vegetation  type. 

At  a  higher  level  of  resolution,  forage  resources 
may  be  evaluated  by  measuring  abundance,  species 
composition,  and  plant  use  on  selected  areas.  Inter- 
pretation of  species  composition  data  will  require 
a  classification  of  forage  species  based  on  local  stud- 
ies of  sheep  food  habits  and  preferences.  In  the  west- 
ern U.S.,  monitoring  forage  condition  and  trend  on 
critical  ranges  of  wild  ungulates  in  this  way  has  been 
traditional.  These  detailed  measurements  of  forage 
condition  and  trend  can  be  useful  in  managing  sheep 
when  the  numbers  and  productivity  of  sheep  are 
limited  by  forage  resources  and  forage  conditions  are 
determined  by  manageable  factors  (either  the  den- 
sity of  foraging  ungulates  or  plant  succession).  In 
contrast,  sheep  populations  may  be  controlled  by 
factors  such  as  predation,  disease,  or  accidents  and 
forage  conditions  may  be  controlled  largely  by  the 
density-independent  influences  of  weather.  Expen- 
sive systems  for  detailed  measurement  of  forage  con- 
dition and  trend  should  not  be  established  without 
considering  the  potential  value  of  the  information  to 
be  obtained. 

Population  Measurements. 


Relative  Density.  Trends  in  sheep  populations 
have  long  been  evaluated  by  counting  animals  on 
key  areas  (winter  ranges,  at  water  holes)  or  along 
transects  observed  from  aircraft,  vehicles,  or  on  foot. 
Resulting  data  provide  minimum  population  sizes 
and  estimates  of  population  sex-age  structure.  On 
key  areas  or  with  small  herds,  counts  of  animals  may 
include  nearly  the  entire  population,  though  this 
assumption  is  best  tested  with  marked  animals. 

Trend  counts  have  not  usually  been  considered 
for  year-to-year  population  monitoring.  Repeated 


Ram,  ewe,  and  lamb  herd. 


counts  within  years  will  be  needed  to  evaluate  sam- 
pling variations  (measure  the  standard  count  devia- 
tions) so  trends  occurring  over  a  few  years  can  be 
detected.  Repeated  counts  should  be  taken  several 
days  apart  to  allow  for  independence.  Selection  of 
standard  weather  conditions  for  making  counts  may 
reduce  variation  among  counts  and  improve  the 
ability  to  detect  trends.  Without  repeated  counts 
within  years,  only  long-term  trends  may  be  evaluated 
statistically  (Harris  1986). 

Trend  counts  may  be  biased  if  changes  in  ani- 
mal-observability are  correlated  with  population 
trend.  This  could  occur  if  some  animals  alter  their 
habitat  selection  or  their  distribution  in  relation 
to  counting  areas,  in  response  to  changing  popula- 
tion size. 

Census.  On  key  areas,  repeated  classifications 
of  sheep  may  be  used  to  produce  a  known,  mini- 
mum population  size.  Since  sheep  may  be  counted 
by  several  sex-age  classes,  the  cohort-completion 
method  (summing  the  maximum  unduplicated  count 
for  each  class)  may  be  advantageous.  Largest  counts 
are  obtained  when  several  key  areas,  such  as  lambing 
areas,  winter  ranges  or  rutting  areas,  or  waterholes, 
are  well-known  for  the  herd. 

With  a  known  number  of  marked  sheep  in  the 
population,  a  herd  may  be  estimated  by  a  Petersen 
estimate  or  a  similar  method  (Furlow  et  al.  1981). 
Unless  a  large  proportion  of  the  herd  can  be  marked, 
the  precision  of  a  Petersen  estimate  is  best  con- 
trolled by  using  several  second  samples  of  the 
marked  to  unmarked  ratio.  Variances  among  several 
resulting  estimates  may  be  used  to  calculate  confi- 
dence limits.  Limits  will  become  narrower  as  the 
number  of  second  samples  increases. 

Most  often,  all  sex-age  classes  of  sheep  cannot 
be  marked  at  random  nor  in  proportion  to  their 
numbers  in  the  herd.  Consequently,  the  accuracy  of 


554 


Ungulates 


a  Petersen  estimate  will  depend  on  representative 
second  samples.  These  samples  must  be  distributed 
over  the  entire  area  used  by  the  herd.  For  many 
herds,  rams  and  ewes  may  not  be  equally  sampled 
except  during  the  rutting  or  winter  seasons.  An  alter- 
native solution  may  be  to  estimate  the  number  of 
ewes  by  the  Petersen  method  and  to  estimate  the 
numbers  of  lambs,  yearlings,  and  rams  by  lamb  to 
yearling  to  ewe  and  ram  to  ewe  ratios  obtained  dur- 
ing seasons  when  unbiased  ratios  are  most  likely 
(as  during  the  rut). 

Holl  and  Bleich  (1983)  used  a  variation  of  the 
Petersen  Estimator  requiring  concurrent  ground- 
based  and  helicopter-based  counts  of  sheep  to  iden- 
tify duplicate  sightings.  With  this  variation,  many 
ground-based  counters  were  necessary. 

Population  Structure.  Mountain  sheep  may  be 
classified  by  eight  sex-age  categories:  lambs,  yearling 
males,  yearling  females,  ewes,  and  four  classes  of 
rams  based  on  V*  curl  horn  increments.  However, 
yearlings,  especially  yearling  females,  may  be  difficult 
to  distinguish  from  ewes  when  they  are  viewed  from 
a  distance  or  from  aircraft.  Desert  bighorns  have 
relatively  asynchronous  lambing,  and  yearling  fe- 
males may  vary  by  4,  6,  or  more  months.  The  oldest 
of  these  yearling  females  will  be  difficult  to  distin- 
guish from  ewes.  Consequently,  in  desert  sheep  liter- 
ature, many  reported  lamb  to  ewe  ratios  are  based 
on  ewe  classes,  including  nonparous  yearlings  with 
ewes.  In  contrast,  lamb  to  yearling  to  ewe  ratios  are 
routinely  reported  for  northern  mountain  sheep. 

Mountain  sheep  are  social  animals  even  though 
rams  and  ewe-juvenile  groups  use  separate  ranges 
during  most  seasons.  This  creates  special  problems 
in  obtaining  a  representative  sample  of  population 
sex-age  structure.  Ram  to  ewe  ratios  are  best  ob- 
tained during  rutting  or  perhaps  on  winter  ranges  or 


near  water  sources  in  dry  seasons,  when  the  distribu- 
tions of  rams  and  ewes  are  usually  less  independent. 
Sampling  the  population  requires  at  least  one  season- 
ally used  range  to  be  well-known.  Due  to  the  social 
patterns  of  mountain  sheep,  individuals  are  unlikely 
to  be  classified  at  random.  Consequently,  either  the 
social  group  or  all  sheep  seen  on  a  census  route 
or  area  should  be  the  sampling  unit  (Bowden  et  al. 
1 984 ).  Whenever  a  few  large  groups  of  sheep  have 
been  classified,  efforts  should  continue  to  find 
smaller,  peripheral  groups  having  differing  sex-age 
structures.  If  a  large  proportion  of  a  herd  is  classi- 
fied, confidence  limits  for  the  resulting  sex-age  ratios 
may  be  narrowed  by  using  statistical  methods  appro- 
priate for  finite  populations. 

Goodson  ( 1978)  reviewed  23  studies  of  Rocky 
Mountain  bighorn,  seeking  a  correlation  between 
population  age  structure  and  population  trend.  Lamb 
to  ewe  ratios  showed  no  correlation,  but  yearling 
to  ewe  ratios  were  loosely  correlated.  Yearling  to 
ewe  ratios  (mean,  range/ 100  ewes)  were  9  (0-20) 
for  decreasing  herds;  28  (15-56)  for  stable  herds;  31 
(20-43)  for  slightly  increasing  herds;  and  52  (41- 
61)  for  strongly  increasing  herds.  Since  mortality 
rates  are  seldom  known,  these  generalizations  must 
be  used  with  caution. 

Discussion.  Habitat  loss  from  expanding  human 
activities  is  the  most  widespread  threat  to  mountain 
sheep.  In  the  past,  impacts  of  habitat  degradation  and 
expanding  human  activity  have  rarely  been  miti- 
gated. There  may  be  frequent  opportunities  to  miti- 
gate for  habitat  impacts  by  scheduling  threatening 
activities  to  avoid  critical  periods  (e.g.,  lambing)  or 
by  improving  and  expanding  off-site  habitat.  Mitiga- 
tion, however,  remains  largely  untested.  Further- 
more, with  habitat  impacts,  opportunities  for 
poaching  are  often  increased  and  cannot  be 
mitigated. 


.  :t.      r  ■  : 


Bighorn  sheep  in  open  habitat. 


Ungulates 


555 


Harvest  of  mountain  sheep  is  usually  conserva- 
tive, with  only  older  rams  being  legal.  However, 
Heimer  et  al.  ( 1984)  and  Geist  (1971b)  questioned 
the  harvest  of  rams  smaller  than  full  curl. 

Disease  is  a  common  problem  with  bighorn 
sheep.  Rocky  Mountain  and  California  bighorn  have 
suffered  spectacular  all-age  dieoffs  from  pneumonia 
(Stelfox  1971;  Feuerstein  et  al.  1980;  Foreyt  and 
Jessup  1982;  Spraker  et  al.  1984;  Onderka  and  Wis- 
hart  1984);  scabies  caused  an  all-age  dieoff  in  desert 
bighorn  (Lange  et  al.  1980).  Chronic  lamb  mortality 
is  common  in  Rocky  Mountain  and  desert  bighorn 
(Spraker  and  Hibler  1977,  1982;  DeForge  et  al. 
1982).  Chronic  sinusitis  and  contagious  ecthyma  are 
additional  problems  (Bunch  et  al.  1978;  Samuel  et 
al.  1975;  Lance  et  al.  1981).  Factors  predisposing 
sheep  to  diseases  have  been  proximity  to  livestock, 
especially  domestic  sheep  (Goodson  1982),  and 
various  stressors  resulting  from  habitat  loss,  concen- 
tration of  animals,  and  harassment  from  human  activ- 
ities in  critical  areas  or  small  ranges.  In  addition, 
mountain  sheep  may  not  have  had  a  long  evolution- 
ary history  with  many  diseases  they  now  face,  espe- 
cially in  southern  latitudes  and  where  livestock  or 
exotic  ungulates  have  been  introduced  (Geist  1985). 

Lack  of  water  is  another  limiting  factor  for  des- 
ert bighorn  herds  and  may  also  influence  the  distri- 


bution of  Rocky  Mountain  bighorn  at  low  elevations. 
Development  of  artificial  water  sources  is  prominent 
in  desert  bighorn  management  (Graf  1980). 

In  the  forested  ranges  of  Rocky  Mountain  big- 
horn, succession  has  gradually  degraded  many  big- 
horn ranges  (Wakelyn  1984;  Risenhoover  and  Bailey 
1985).  Encroachment  by  trees,  combined  with  man- 
made  habitat  losses,  has  reduced  the  sizes  of  seasonal 
ranges,  interrupted  migration  corridors,  and  isolated 
herds.  The  result  has  been  many  small  sedentary 
herds  with  few  options  to  move  in  response  to 
weather  or  human  activity.  Such  small  isolated  herds 
may  also  suffer  from  inbreeding  and  depression  of 
reproduction  and  survival  (Soule  and  Wilcox  1980; 
Sausman  1982).  In  forested  regions,  there  should  be 
abundant  opportunities  to  expand  sheep  ranges  and 
to  mitigate  for  habitat  losses  by  vegetation  manage- 
ment, especially  with  prescribed  fire. 

Much  effort  is  expended  trapping  and  trans- 
planting bighorn  into  historic  ranges  from  which 
they  have  been  extirpated  (Schmidt  et  al.  1978;  Wil- 
son et  al.  1980;  Dodd  1983;  Fuller  1984;  Bates  et 
al.  1985).  Rocky  Mountain  bighorn  are  also  baited 
and  treated  for  lungworms  (Schmidt  et  al.  1979). 
However,  long-term  effects  of  these  management  ac- 
tivities have  not  always  been  monitored. 


Lamb/cwc  group. 


556  Ungulates 


LITERATURE  CITED 


ADAMS,  L.  1981.  Ecology  and  population  dynamics  of 
mountain  goats,  Sheep  Mountain-Gladstone  Ridge, 
Colorado.  M.S.  thesis,  Colorado  State  Univ.,  Fort  Col- 
lins. 189pp. 

ADAMS,  LG.  and  J.A.  BAILEY.  1980.  Winter  habitat  selec- 
tion and  group  size  of  mountain  goats,  Sheep  Moun- 
tain-Gladstone Ridge,  Colorado.  Pages  465-481  in 
Hickey,  W.O.,  ed.  Proc.  Biennial  Symp.,  Northern  Wild 
Sheep  and  Goat  Council.  Salmon,  ID. 

and .  1982.  Population  dynamics  of  moun- 
tain goats  in  the  Sawatch  Range,  Colorado.  J.  Wildl. 
Manage.  46:1003-1009. 

and .  1983-  Winter  forages  of  mountain 


goats  in  central  Colorado.  J.  Wildl.  Manage.  46:1237- 
1243. 

— ,  MA.  MASTELLER,  and  J.A.  BAILEY.  1982a.  Move- 
ments and  home  range  of  mountain  goats,  Sheep 
Mountain-Gladstone  Ridge,  Colorado.  Pages  391-405 
in  Bailey,  J.A.  and  G.G.  Schoonveld,  eds.  Proc.  Third 
Biennial  Symp.,  Northern  Wild  Sheep  and  Goat  Coun- 
cil. Fort  Collins,  CO. 

— ,  KL.  RISENHOOVER,  and  J.A.  BAILEY.  1982b.  Eco- 


logical relationships  of  mountain  goats  and  Rocky 
Mountain  bighorn  sheep.  Pages  9-22  in  Bailey,  J.A. 
and  G.G.  Schooneld,  eds.  Proc.  Third  Biennial  Symp., 
Northern  Wild  Sheep  and  Goat  Council.  Fort  Collins, 
CO. 

ALLEN,  A.W.,  J.G.  COOK,  and  M.J.  ARMBRUSTER.  1984. 
Habitat  suitability  index  models:  Pronghorn.  U.S.  Dep. 
Inter.,  Fish  and  Wildl.  Serv.  FWS/OBS-82/ 10.65. 
WELUT,  Fort  Collins,  CO.  23pp. 

ALLEN,  E.O.  1977.  A  new  perspective  for  elk  habitat  man- 
agement. Proc.  Western  Assoc.  Fish  and  Game  Comm. 
195-205. 

ALLEN,  J.A.  1913-  Ontogenetic  and  other  variations  in 
muskoxen  with  a  systematic  review  of  the  muskox 
group,  recent  and  extinct.  Memoirs  Am.  Mus.  Nat. 
Hist.,  Vol.  1,  Part  IV:  103-226. 

ANDERSON,  A.E.,  D.E.  MEDIN,  and  DC.  BOWDEN    1972. 
Mule  deer  numbers  and  shrub-yield-utilization  on 
winter  range.  J.  Wildl.  Manage.  36:571-578. 

AUTENRIETH,  R.,  ed.  1978.  Guidelines  for  the  manage- 
ment of  pronghorn  antelope.  Pages  472-526  in  Bar- 
rett, M.W.,  ed.  Proc.  Eighth  Biennial  Pronghorn 
Antelope  Workshop.  Alberta  Recreation,  Parks,  and 
Wildlife,  Fish  and  Wildlife  Division. 

BAILEY,  J.A.  1984.  Principles  of  wildlife  management.  John 
Wiley  and  Sons,  New  York,  NY.  373pp. 

and  AY.  COOPERRIDER.  1982.  Final  Report: 

Trickle  Mountain  Research  Study.  U.S.  Dep.  Inter., 
Bur.  Land  Manage.,  Denver  Service  Center,  Denver, 
CO.  137pp. 

and  B.K  JOHNSON.  1977.  Status  of  introduced 

mountain  goats  in  the  Sawatch  Range  of  Colorado. 
Pages  54-63  in  Samuel,  W.  and  W.G.  MacGregor,  eds. 
Proc.  First  Int.  Mountain  Goat  Symp.  Kalispell,  MT. 

and  G.G.  SCHOONVELD,  eds.  1982.  Northern  wild 


sheep  and  goat  council.  Proc.  Third  Biennial  Symp. 

Fort  Collins,  CO.  405pp. 
BAILEY,  N.T.J.  1951.  On  estimating  the  size  of  mobile 

populations  from  recapture  data.  Biometrika  38:293- 

306. 
BALLARD,  W.  1977.  Status  and  management  of  the  moun- 


tain goat  in  Alaska.  Pages  15-23  in  Samuel,  W.  and 
W.G.  MacGregor,  eds.  Proc.  First  Int.  Mountain  Goat 
Symp.  Kalispell,  MT.  243pp. 

BANFIELD,  A.W.F.  1962.  A  revision  of  the  reindeer  and 
caribou,  Genus  Rangifer.  Natl.  Mus.  Canada  Bull.  177. 
Ottawa.  137pp. 

BARTMANN,  R.M.  1974.  Piceance  deer  study — population 
density  and  structure.  Colorado  Div.  Wildl.  Game 
Res.  Rep.,  July.  Part  2:363-380. 

BATES,  J  W,  Jr.,  J.  W.  BATES,  and  J.G.  GUYMON.  1985. 
Comparison  of  drive  nets  and  darting  for  capture  of 
desert  bighorn  sheep.  Wildl.  Soc.  Bull.  1 3:73-76. 

BEAR,  G.D.  and  R.A.  GREEN.  1980.  Elk  population  and 
ecology  studies.  Colorado  Div.  Wildl.,  Wildl.  Res. 
Rep,  July.  Part  2:221-313. 

BEASOM,  S.L.,  E.P.  WIGGERS,  and  JR.  GIARDINO.  1983.  A 
technique  for  assessing  land  surface  ruggedness.  J. 
Wildl.  Manage.  47:1163-1166. 

BELOVSKY,  G.E.  1978.  Diet  optimization  in  a  generalist 
herbivore:  The  moose.  Theoretical  Population  Biol. 
14:105-134. 

BERGERUD,  AT  1963  Aerial  winter  census  of  caribou.  J. 
Wildl.  Manage.  27:438-449. 

.  1978.  Caribou.  Pages  83-101  in  Schmidt,  J. L.  and 

D.L.  Gilbert,  eds.  Big  Game  of  North  America:  Ecology 
and  Management.  The  Wildl.  Manage.  Inst.,  Washing- 
ton, DC.  and  Stackpole  Books,  Harrisburg,  PA. 

.  1979.  Caribou.  Pages  83-102  in  Schmidt,  J.L.  and 

D.L.  Gilbert,  eds.  Big  Game  of  North  America:  Ecology 
and  Management.  Stackpole  Books,  New  York,  NY. 
494pp. 

BISSONETTE,  J.A.  1976.  The  relationship  of  resource 

quality  and  availability  to  social  behavior  and  organi- 
zation in  the  collared  peccary.  PhD  dissertation,  Univ. 
Michigan,  Ann  Arbor.  137pp. 

.  1982.  Collared  peccary.  Pages  841-850  in  Chap- 
man, J.A.  and  G.A.  Feldhamer,  eds.  Wild  Mammals 
of  North  America.  The  Johns  Hopkins  Univ.  Press,  Bal- 
timore, MD. 

BLACK,  H.,  R.J.  SCHERZINGER,  and  J.W.  THOMAS.  1976. 
Relationships  of  Rocky  Mountain  elk  and  Rocky 
Mountain  mule  deer  habitat  to  timber  management  in 
the  Blue  Mountains  of  Oregon  and  Washington.  Proc. 
Elk-Logging-Roads  Symp.  Univ.  Idaho,  Moscow:  1 1-31. 

BONE,  J.N.  1978.  Status  of  the  mountain  goat  {Oreamnos 
americanus)  of  the  Similkameen  River,  British  Co- 
lumbia. Biennial  Symp.,  Northern  Wild  Sheep  and 
Goat  Council  1:123-130. 

BOWDEN,  DC,  A.E.  ANDERSON,  and  D.E.  MEDIN.  1984. 
Sampling  plans  for  mule  deer  sex  and  age  ratios.  J. 
Wildl.  Manage.  48:500-509. 

BOYD,  R.J.  1958.  Comparison  of  air  and  ground  deer  and 
elk  counts.  Colorado  Dep.  Game  and  Fish  Quarterly 
Progress  Rep.  Oct:  145- 146. 

1970.  Elk  of  the  White  River  Plateau,  Colorado. 

Colorado  Game,  Fish,  and  Parks  Dep.  Tech.  Publ.  25. 
126pp. 

1978.  American  elk.  Pages  10-29  in  Schmidt,  J.L. 

and  D.L.  Gilbert,  eds.  Big  Game  of  North  America: 
Ecology  and  Management.  The  Wildl.  Manage.  Inst., 
Washington,  DC.  and  Stackpole  Books,  Harrisburg,  PA. 

,  T.M.  POJAR,  and  B.D.  BAKER.  1975.  Deer  and  elk 


management  study.  Colorado  Div.  Wildl.  Game  Res. 
Rep,  July.  Part  2:365-412. 

BRANDBORG,  S.M.  1955.  Life  history  and  management  of 
the  mountain  goat  in  Idaho.  Idaho  Dep.  Fish  and 
Game  Wildl  Bull.  2.  l42pp. 


Ungulates 


557 


BRASSARD,  J.M.,  E.  AUDY,  M.  CRETE,  and  P.  GRENIER. 
1974.  Distribution  and  winter  habitat  of  moose  in 
Quebec.  Naturaliste  Canadien  101:67-80. 

BROWNING,  B.M.  and  G.  MONSON.  1980.  Food.  Pages  80- 
99  in  Monson,  G.  and  L.  Sumner,  eds.  The  Desert 
Bighorn.  Univ.  Arizona  Press,  Tucson.  370pp. 

BUNCH,  T.D.,  S.R.  PAUL,  and  HE.  McCUTCHEN.  1978. 

Chronic  sinusitis  in  desert  bighorn  sheep  (Ovis  cana- 
densis nelsoni).  Trans.  Desert  Bighorn  Council 
22:16-20. 

BURNHAM,  K.P.,  DR.  ANDERSON,  and  J.L.  LAAKE.  1980. 
Estimation  of  density  from  line  transect  sampling 
of  biological  populations.  Wildl.  Monogr.  72.  202pp. 

CALEF,  G.  and  DC.  HEARD.  1980.  The  status  of  three 

tundra  wintering  caribou  herds  in  northeastern  main- 
land Northwest  Territories.  Pages  582-594  in  Rei- 
mers,  E.,  E.  Gaare,  and  S.  Skjenneberg,  eds.  Proc. 
Second  Int.  Reindeer/Caribou  Symp.  Dir.  for  Vilt  og 
Ferskvannsfisk,  Trondheim. 

CAMERON,  R.D.  and  KR.  WHITTEN.  1979.  Seasonal 

movements  and  sexual  segregation  of  caribou  deter- 
mined by  aerial  survey.  J.  Wildl.  Manage.  43:626-633- 

CAUGHLEY,  G.  1977.  Analysis  of  vertebrate  populations. 
John  Wiley  and  Sons,  New  York,  NY.  234pp. 

.  1979.  "What  is  this  thing  called  carrying  capacity?" 

Pages  2-8  in  Boyce,  M.S.  and  L.D.  Hayden-Wing,  eds. 
North  American  Elk:  Ecology,  Behavior,  and  Manage- 
ment. Univ.  Wyoming,  Laramie.  294pp. 

CHADWICK,  D.H.  1983-  A  beast  the  color  of  winter.  Sierra 
Club,  San  Francisco,  CA.  208pp. 

COADY,  J  W.  and  RA.  HINMAN.  1984.  Management  of 
muskoxen  in  Alaska.  Pages  47-51  in  Klein,  DR.,  R.G. 
White,  and  S.  Keller,  eds.  Proc.  First  Int.  Muskox 
Symp.  Spec.  Rep.,  Biol.  Pap.  Univ.  Alaska,  Fairbanks. 

COLLINS,  W.B.  and  P.J.  URNESS.  1979.  Elk  pellet  group 
distributions  and  rates  of  deposition  in  aspen  and 
lodgepole  pine  habitats.  Pages  140-144  in  Boyce, 
M.S.,  and  L.D.  Hayden-Wing,  eds.  North  American  Elk: 
Ecology,  Behavior,  and  Management.  Univ.  Wyoming, 
Laramie.  294pp. 

CONLEY,  W.  1978.  Population  modeling.  Pages  305-320 
in  Schmidt,  J.L.  and  D.L.  Gilbert,  eds.  Big  Game  of 
North  America:  Ecology  and  Management.  Stackpole 
Books,  New  York,  NY.  494pp. 

CONNOLLY,  G.E.  1981.  Trends  in  populations  and  har- 
vests. Pages  225-243  in  Wallmo,  O.C,  ed.  Mule  and 
Black-tailed  Deer  of  North  America.  Univ.  Nebraska 
Press,  Lincoln.  605pp. 

COOPERRIDER,  AY.  1982.  Forage  allocation  for  elk  and 
cattle.  Pages  142-149  in  Britt,  T.L.,  and  DP.  Theobald, 
eds.,  Proc.  Western  States  Elk  Workshop,  February 
22-24,  1982.  Flagstaff,  AZ.  166pp. 

and  J.A.  BAILEY.  1984.  A  simulation  approach  to 

forage  allocation.  Pages  525-559  in  Developing  Strate- 
gies for  Rangeland  Management.  Rep.  prepared  by 
the  Committee  on  Developing  Strategies  for  Range- 
land  Manage.  Natl.  Res.  Council/Natl.  Acad.  Sciences. 
Westview  Press,  Boulder,  CO.  2022pp. 

,  S.A.  McCOLLOUGH,  and  J.A.  BAILEY.  1980.  Varia- 


tion in  bighorn  sheep  food  habits  as  measured  by 
fecal  analysis.  Biennial  Symp.,  Northern  Wild  Sheep 
and  Goat  Council.  Salmon,  ID.  2:29-41. 

DAILEY,  T.V.,  NT.  HOBBS,  and  T.N.  WOODARD.  1984. 
Experimental  comparisons  of  diet  selection  by  moun- 
tain goats  and  mountain  sheep  in  Colorado.  J.  Wildl. 
Manage.  48:799-806. 

DALE,  A.R.  and  J.A.  BAILEY.  1982.  Application  of  optimal 


foraging  theory  for  bighorn  sheep  habitat  evaluation. 
Biennial  Symp.,  Northern  Wild  Sheep  and  Goat  Coun- 
cil 3:254-261. 

DAVIS,  J.L.  and  P.  VALKENBERG  1979.  Caribou  distribu- 
tion, population  characteristics,  mortality  and  re- 
sponses to  disturbance  in  northwest  Alaska.  Chapter  2 
in  Lent,  PC,  ed.  Studies  of  Selected  Wildlife  and 
Fish  Populations  on  and  adjacent  to  the  National  Pe- 
troleum Reserve,  Alaska.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.,  Anchorage,  AK  226pp. 

DeFORGE,  JR.,  DA.  JESSUP,  CW.  JENNER,  and  J.E. 

SCOTT.  1982.  Disease  investigations  into  high  lamb 
mortality  of  desert  bighorn  sheep  in  the  Santa  Rosa 
Mountains,  California.  Desert  Bighorn  Council  Trans. 
65-76. 

DODD,  N.L.  1983-  Ideas  and  recommendations  for  maxi- 
mizing desert  bighorn  transplant  efforts.  Desert  Big- 
horn Council  Trans.  12-16. 

DOERR,  J.  1979.  Population  analysis  and  modeling  of  the 
western  arctic  herd  with  comparisons  to  other  Alas- 
kan Rangifer  populations.  Unpubl.  M.S.  thesis,  Univ. 
Alaska,  Fairbanks.  34  lpp. 

DONALDSON,  B.  1967.  Javelina.  Pages  88-94  in  New 
Mexico  Wildlife  Management.  New  Mexico  Dep. 
Game  and  Fish,  Santa  Fe. 

EDDY,  T.A.  1961.  Food  and  feeding  patterns  of  the  col- 
lared peccary  in  southern  Arizona.  J.  Wildl.  Manage. 
25(3>248-257. 

EINARSEN,  AS.  1948.  The  pronghorn  antelope  and  its 

management.  Stackpole  Books,  Harrisburg,  PA.  235pp. 

ELENOWITZ,  A.  1984.  Group  dynamics  and  habitat  use 
of  transplanted  desert  bighorn  sheep  in  the  Peloncillo 
Mountains,  New  Mexico.  Desert  Bighorn  Council 
28:1-8. 

ELLISOR,  J.E.  and  W.F.  HARWELL.  1979.  Ecology  and 

management  of  javelina  in  south  Texas.  FA.  Rep.,  Se- 
ries 16.  Texas  Parks  and  Wildl.  Dep.,  Austin.  25pp. 

ERIKSSON,  O.  1980.  A  method  for  range  appraisal  using 
small  aircraft  for  sampling  of  vegetation  data.  Pages 
41-46  in  Reimers,  E.,  E.  Gaare,  and  S.  Skjenneberg, 
eds.  Proc.  Second  Int.  Reindeer/Caribou  Symp.  Dir.  for 
Vilt  og  Ferskvannsfisk,  Trondheim. 

FEUERSTEIN,  V.,  RL.  SCHMIDT,  C.P.  HIBLER,  and  W.H. 
RUTHERFORD.  1980.  Bighorn  sheep  mortality  in  the 
Taylor  River-Almont  Triangle  Area,  1978-1979.  A  case 
study.  Colorado  Div.  Wildl.  Spec.  Rep.  48.  19pp. 

FICHTER,  E.  1974.  On  the  bedding  behaviour  of  prong- 
horn  fawns.  Pages  352-355  in  Geist,  V.  and  F. 
Walther,  eds.  The  Behaviour  of  Ungulates  and  Its 
Relation  to  Management.  IUCN  Publ.  New  Series  24. 
94  lpp. 

FOREYT,  W.J.  and  DA.  JESSUP.  1982.  Fatal  pneumonia  of 
bighorn  sheep  following  association  with  domestic 
sheep.  J.  Wildl.  Diseases  18:163-168. 

FOSTER,  BR.  1977.  Historical  patterns  of  mountain  goat 
harvest  in  British  Columbia.  Pages  147-159  in  Samuel, 
W.  and  W.  MacGregor,  eds.  Proc.  First  Int.  Mountain 
Goat  Symp.  Kalispell,  Ml 

.  1978.  Horn  growth  and  quality  management  for 

mountain  goats.  Pages  200-226  in  Hebert,  D.M.  and 
M.  Nation,  eds.  Proc.  1978  Northern  Wild  Sheep  and 
Goat  Conf.  Penticton,  British  Columbia. 

1982.  Observability  and  habitat  characteristics  of 


the  mountain  goat  in  west-central  British  Columbia. 
M.S.  thesis,  Univ.  British  Columbia,  Vancouver.  134pp. 
FOWLER,  CW.  and  ID.  SMITH,  eds.  1981.  Dynamics  of 
large  mammal  populations.  John  Wiley  and  Sons,  New 


558 


Ungulates 


York,  NY.  477pp. 

FOX,  J.L  and  R.D.  TABER.  1981.  Site  selection  by  moun- 
tain goats  wintering  in  forest  habitat.  U.S.  Dep.  Agric., 
For.  Serv.,  Pacific  Northwest  For.  and  Range  Experi- 
ment Sta,  Juneau,  AK  55pp. 

,  KJ.  RAEDEKE,  and  C.A.  SMITH.  1982.  Mountain 

goat  ecology  on  Cleveland  Peninsula,  Alaska,  1980- 
1982.  U.S.  Dep.  Agric,  For.  Serv.,  Pacific  Northwest 
For.  and  Range  Experiment  Sta.,  Juneau,  AK.  4  lpp. 

FRANZMANN,  A.W.  1978.  Moose.  Pages  67-82  in  Schmidt, 
J.L.  and  D.L.  Gilbert,  eds.  Big  Game  of  North  America. 
Stackpole  Books,  Harrisburg,  PA. 

,  A.  FLYNN,  and  P.D.  ARNESON.  1975.  Levels  of 

some  mineral  elements  in  Alaskan  moose  hair.  J. 
Wildl.  Manage.  39:374-378. 

,  RE.  LeRESCHE,  R.A.  RAUSCH,  and  J.L.  OLDE 

MEYER.  1978.  Alaskan  moose  measurements  and 
weights  and  measurement-weight  relationships.  Cana- 
dian J.  Zoology  56:298-306. 

FRASER,  D.,  E.R.  CHAVEZ,  and  J.L.  PALOHEIMO.  1984. 
Aquatic  feeding  by  moose:  Selection  of  plant  species 
and  feeding  areas  in  relation  to  plant  chemical  com- 
position and  characteristics  of  lakes.  Canadian  J.  Zool- 
ogy 62:80-87. 

FREDDY,  D.J.  and  DC.  BOWDEN.  1983a.  Efficacy  of  per 
manent  and  temporary  pellet  plots  in  juniper-pinyon 
woodland.  J.  Wildl.  Manage.  47:512-516. 

and .  1983b.  Sampling  mule  deer  pellet- 
group  densities  in  juniper-pinyon  woodland.  J.  Wildl. 
Manage.  47:476-485. 

FULLER,  A.F.  1984.  Drop  net  capture  of  bighorn  sheep  in 
Arizona.  Desert  Bighorn  Council  Trans.  28:39-40. 

FURLOW,  R.C.,  M.  HADERLIE,  and  R.  VAN  DEN  BERGE. 
1981.  Estimating  a  bighorn  population  by  mark-recap- 
ture. Desert  Bighorn  Council  Trans.  25:31-33 

GASAWAY,  W.D.,  S.D.  DUBOIS,  and  S.  HARBO.  1985. 

Biases  in  aerial  transect  surveys  for  moose  during  May 
and  June.  J.  Wildl.  Manage.  49:777-784. 

GEIST,  V.  1971a.  Mountain  sheep:  A  study  in  behavior  and 
evolution.  Univ.  Chicago  Press,  IL.  383pp. 

1971b.  Bighorn  sheep  biology.  The  Wildl.  Soc. 


News  136:61. 
— .  1981.  Behavior- 


-adaptive  strategies  in  mule  deer. 


Pages  157-224  in  Wallmo,  O.C.,  ed.  Mule  and  Black- 
tailed  Deer  of  North  America.  Univ.  Nebraska  Press, 
Lincoln.  605pp. 

1985.  Pleistocene  bighorn  sheep:  Some  problems 


of  adaptation,  and  relevance  to  today's  American 
megafauna.  Wildl.  Soc.  Bull.  13:351-359. 

GILBERT,  JR.  1978.  Estimating  population  characteristics. 
Pages  297-304  in  Schmidt,  J.L.  and  D.L.  Gilbert,  eds. 
Big  Game  of  North  America:  Ecology  and  Manage- 
ment. Stackpole  Books,  New  York,  NY.  494pp. 

GILL,  R.B.,  L.H.  CARPENTER,  and  DC.  BOWDEN.  1983. 
Monitoring  large  animal  populations — the  Colorado 
experience.  Trans.  North  Am.  Wildl.  and  Nat.  Res. 
Conf.  48:330-341. 

GOODSON,  N.J.  1978.  Status  of  bighorn  sheep  in  Rocky 
Mountain  National  Park.  M.S.  thesis,  Colorado  State 
Univ.,  Fort  Collins.  190pp. 

.  1982.  Effects  of  domestic  sheep  grazing  on  big- 
horn sheep  populations:  A  review.  Biennial  Symp., 
Northern  Wild  Sheep  and  Goat  Council  3:287-313. 

GRAF,  W.  1980.  Habitat  protection  and  improvement. 

Pages  310-319  in  Monson,  G.  and  L.  Sumner,  eds.  The 
Desert  Bighorn.  Univ.  Arizona  Press,  Tucson.  370pp. 

GRONQUIST,  R.M.  and  W.B.  DINNEFORD.  1984.  Age 


determination  of  muskoxen  from  dental  cementum 
annuli.  Pages  67-68  in  Klein,  DR.,  R.G.  White,  and  S. 
Keller,  eds.  Proc.  First  Int.  Muskox  Symp.  Spec.  Rep., 
Biol.  Pap.  Univ.  Alaska,  Fairbanks. 

GROSS,  J. E.  1970.  Program  ANPOP.  A  simulation  modeling 
exercise  of  the  Wichita  Mountains  National  Wildlife 
Refuge.  Colorado  Coop.  Wildl.  Res.  Unit  Prog.  Rep. 
Colorado  State  Univ.,  Fort  Collins.  133pp 

GRUNIGEN,  R.E.  1980.  A  system  for  evaluating  potential 
bighorn  sheep  transplant  sites  in  northern  New  Mex- 
ico. Biennial  Symp.,  Northern  Wild  Sheep  and  Goat 
Council  2:211-228. 

GUNN,  A.,  R.  DECKER,  and  T.W.  BARRY.  1984.  Pages  67- 
68  in  Klein,  DR.,  R.G.  White,  and  S.  Keller,  eds.  Proc. 
First  Int.  Muskox  Symp.  Spec.  Rep.,  Biol.  Pap.  Univ. 
Alaska,  Fairbanks. 

GYSEL,  L.W.  and  L.J.  LYON.  1980.  Habitat  analysis  and 
evaluation.  Pages  305-327  in  Schemnitz,  S.D.,  ed. 
Wildlife  Management  Techniques  Manual.  The  Wildl. 
Soc,  Washington,  DC.  686pp. 

HALL,  E.R.  and  KR.  KELSON.  1959.  The  mammals  of 
North  America.  Vol.  11:547-1083.  The  Ronald  Press, 
New  York,  NY. 

HALL,  W.K  1977.  Status  and  management  of  the  Rocky 
Mountain  goat,  Oreamnos  americanus,  in  the  prov- 
ince of  Alberta.  Pages  8-14  in  Samuel,  W.  and  W.G. 
MacGregor,  eds.  Proc.  First  Int.  Mountain  Goat  Symp. 
Kalispell,  MT.  243pp. 

HALLS,  L.K  1984.  White-tailed  deer:  Ecology  and  Manage- 
ment. Stackpole  Books,  Harrisburg,  PA.  870pp. 

HANKS,  J.  1981.  Characterization  of  population  condition. 
Pages  47-74  in  Fowler,  C.W.  and  T.D.  Smith,  eds. 
Dynamics  of  Large  Mammal  Populations.  John  Wiley 
and  Sons,  New  York,  NY.  477  pp. 

HANSEN,  C.G.  1980a.  Habitat.  Pages  64-79  in  Monson,  G. 
and  L.  Sumner,  eds.  The  Desert  Bighorn — Its  Life 
History,  Ecology,  and  Management.  Univ.  Arizona 
Press,  Tucson.  370pp. 

.  1980b.  Habitat  evaluation.  Pages  320-335  in  Mon- 
son, G.  and  L.  Sumner,  eds.  The  Desert  Bighorn — Its 
Life  History,  Ecology,  and  Management.  Univ.  Arizona 
Press,  Tucson.  370pp. 

HANSON,  W.R.  1963  Calculation  of  productivity,  survival 
and  abundance  of  selected  vertebrates  from  sex  and 
age  ratios.  Wildl.  Monogr.  9.  60pp. 

HARRIS,  R.B.  1986.  Reliability  of  trend  lines  obtained 
from  variable  counts.  J.  Wild.  Manage.  50:165-171. 

HEBERT,  D.M.  1973.  Altitudinal  migration  as  a  factor 
in  the  nutrition  of  bighorn  sheep.  PhD  dissertation, 
Univ.  British  Columbia,  Vancouver.  355pp. 

and  M.  NATION,  eds.  1978.  Proceedings  of  the 

1978  northern  wild  sheep  and  goat  conference.  Pen- 
ticton,  British  Columbia.  412pp. 

and  W.G.  TURNBULL.  1977.  A  description  of  south- 
ern interior  and  coastal  mountain  goat  types  of  British 
Columbia.  Pages  126-159  in  Proc  First  Inter.  Moun- 
tain Goat  Symp.  Kalispell,  MT. 

HEIB,  S.R.,  ed.  1976.  Proceedings  of  the  elk-logging-roads 
symposium.  Univ.  Idaho,  Moscow.  142pp. 

HEIMER,  W.E.,  S.M.  WATSON,  and  T.C.  SMITH.  1984. 

Excess  ram  mortality  in  a  heavily  hunted  Dall's  sheep 
population.  Biennial  Symp.,  Northern  Wild  Sheep 
and  Goat  Council  4:425-432. 

HESSELTON,  W.T.  and  R.M.  HESSELTON.  1982.  White- 
tailed  deer.  Pages  878-901  in  Chapman,  J.A.  and  G.A. 
Feldhamer,  eds.  Wild  Mammals  of  North  America.  The 
Johns  Hopkins  Univ.  Press,  Baltimore,  MD. 


Ungulates 


559 


HIBBS,  L.D.  1966.  A  literature  review  on  mountain  goat 
ecology.  Colorado  Dep.  Game,  Fish,  and  Parks  Spec. 
Rep.  8.  23pp. 

HICKEY,  W.O.,  ed.  1980.  Proceedings  of  the  biennial 
symposium  of  the  northern  wild  sheep  and  goat 
council.  Salmon,  ID.  668pp. 

HJELJORD,  O.  1973.  Mountain  goat  forage  and  habitat 

preference  in  Alaska.  J.  Wildl.  Manage.  37(3):353-362. 

HOBBS,  NT.,  D.L.  BAKER, J.E.  ELLIS,  DM.  SWIFT,  and 
R.A.  GREEN.  1982.  Energy-  and  nitrogen-based  esti- 
mates of  elk  winter-range  carrying  capacity.  J.  Wildl. 
Manage.  46:12-21. 

HOLL,  S.A.  1982.  Evaluation  of  bighorn  sheep  habitat. 
Desert  Bighorn  Council  Trans.  24:47-49. 

and  V.C.  BLEICH.  1983.  San  Gabriel  mountain 

sheep:  Biological  and  management  considerations.  U.S. 
Dep.  Agric,  For.  Serv.,  San  Bernardino  National  Forest, 
San  Bernardino,  CA. 

HOOVER,  R.L.,  C.E.  TILL,  and  S.  OGILVIE.  1959.  The  ante- 
lope of  Colorado.  Colorado  Dep.  of  Game  and  Fish 
Tech.  Bull.  4.  Denver,  CO.  110pp. 

HOREJSI,  B.L.  1976.  Suckling  and  feeding  behavior  in 

relation  to  lamb  survival  in  bighorn  sheep  (Ovis  can- 
adensis canadensis  Shaw).  Unpubl.  PhD  dissertation, 
Univ.  Calgary.  265pp. 

HOUSTON,  D.B.  1974.  Aspects  of  the  social  organization 
of  moose.  Pages  690-696  in  Geist,  V.  and  F.  Walther, 
eds.  The  Behaviour  of  Ungulates  and  Its  Relation  to 
Management.  IUCN  Publ.  New  Series  24.  94  lpp. 

JENNINGS,  W.S.  and  J.T.  HARRIS.  1953.  The  collared  pec- 
cary in  Texas.  Texas  Game  and  Fish  Comm.  FA.  Rep. 
Series  12.  31pp. 

JONES,  R.L.  and  H.C.  HANSON.  1985.  Mineral  licks,  geo- 
phagy,  and  biogeochemistry  of  North  American  ungu- 
lates. Iowa  State  Univ.  Press.  302pp. 

JORDAN,  PA.,  D.B.  BOTKIN,  AS.  DOMINSKI,  H.S.  LOW 
ENDORF,  and  G.E.  BELOVSKY.  1973-  Sodium  as  a 
critical  nutrient  for  the  moose  of  Isle  Royale.  North 
Am.  Moose  Conf.  9:13-42. 

JUNIPER,  I.  1980.  Problems  in  managing  an  irrupting 

caribou  herd.  Pages  722-724  in  Reimers,  E.,  E.  Gaare, 
and  S.  Skjenneberg,  eds.  Proc.  Second  Int.  Reindeer/ 
Caribou  Symp.  Dir.  for  Vilt  og  Ferskvannsfisk,  Tron- 
dheim. 

KEISS,  RE.  1969.  Comparison  of  eruption-wear  patterns 
and  cementum  annuli  as  age  criteria  in  elk.  J.  Wildl. 
Manage.  33(1)175-180. 

KELSALL,  J.P.  1965.  The  migratory  barren-ground  caribou 
of  Canada.  Dep.  of  Indian  Affairs  and  Northern  Devel- 
opment. Canadian  Wildl.  Serv.,  Ottawa.  339pp. 

.  1968.  The  migratory  barren-ground  caribou  of 

Canada.  Canadian  Wildl.  Serv.  Monogr.  3-  Ottawa,  On- 
tario. 340pp. 

.  1 969.  Structural  adaptation  of  moose  and  deer  for 


reindeer  on  St.  Matthew  Island.  J.  Wildl.  Manage. 

32:350-367. 
— .  1982.  Fire,  lichens  and  caribou.  J.  Range  Manage. 

35:390-395. 
— ,  R.G.  WHITE,  and  S.  KELLER,  eds.  1984.  Proceed- 


snow.  J.  Mammal.  50:302-310. 

and  E.S.  TELFER.  1974.  Biogeography  of  moose 


with  particular  reference  to  western  North  America. 
Naturaliste  Canadien  101:117-130. 

KERR,  R.  1979.  Mule  deer  habitat  guidelines.  U.S.  Dep. 
Inter.,  Bur.  Land  Manage.  Tech.  Note  366.  Service 
Center,  Denver,  CO. 

KINDSCHY,  R.R.,  C  SUNDSTROM,  and  J.D.  YOAKUM. 
1982.  Wildlife  habitats  in  managed  rangelands — the 
Great  Basin  of  southeastern  Oregon.  U.S.  Dep.  Agric, 
For.  Serv.,  Pacific  Northwest  For.  and  Range  Experi- 
ment Sta.  Gen.  Tech.  Rep.  145.  18pp. 

KLEIN,  DR.  1968.  The  introduction,  increase  and  crash  of 


ings  of  first  international  muskox  symposium.  Biol. 
Pap.  Univ.  Alaska,  Fairbanks. 

KNIPE,  T.  1957.  The  javelina  in  Arizona.  Arizona  Game 
and  Fish  Dep.  Wildl.  Bull.  2. 

KONKEL,  G.W.,  ed.  1980.  Terrestrial  habitat  evaluation 
criteria  handbook — Alaska.  Div.  Ecol.  Serv.,  U.S.  Dep. 
Inter.,  Fish  and  Wildl.  Serv.,  Anchorage,  AK  (Original 
not  seen;  cited  and  partially  reproduced  in  Mule 
1982). 

KRAUSMAN,  PR, JR.  MORGART,  and  M.  CHILELLI.  1984. 
Annotated  bibliography  of  desert  bighorn  sheep  liter- 
ature, 1897-1983.  Southwest  Nat.  Hist.  Assoc,  Phoe- 
nix, AZ.  204pp. 

KUCK,  L  1977a.  Status  and  management  of  the  mountain 
goat  in  Idaho.  Pages  37-40  in  Samuel,  W.  and  W.O. 
MacGregor,  eds.  Proc.  First  Int.  Mountain  Goat  Symp. 
Kalispell,  MT. 

.  1977b.  The  impact  of  hunting  on  Idaho's  Pahsime- 

roi  mountain  goat  herd.  Pages  114-125  in  Samuel, 
W.  and  W.O.  MacGregor,  eds.  Proc  First  Int.  Moun- 
tain Goat  Symp.  Kalispell,  MT. 

KUFELD,  R.C.,  J.H.  OLTERMAN,  and  DC.  BOWDEN.  1980. 
A  helicopter  quadrat  census  for  mule  deer  on  Uncom- 
pahgre  Plateau,  Colorado.  J.  Wildl.  Manage.  44:632- 
639. 

KUROPAT,  P.  and  J.P.  BRYANT.  1980.  Foraging  behavior 
of  cow  caribou  on  the  Utukok  calving  grounds  in 
Northwestern  Alaska.  Pages  64-70  in  Reimers,  E.,  E. 
Garre,  and  S.  Skjenneberg,  eds.  Proc.  Second  Int. 
Reindeer/Caribou  Symp.  Dir.  for  Vilt  og  Ferskvanns- 
fisk, Trondheim. 

KURT,  F.  1968.  Das  Socialverhalten  des  Rehes.  Eines  Feld- 
studie.  P.  Parey-Verlag,  Berlin.  102pp. 

LANCE,  W„  W.  ADRIAN,  and  B.  WIDHALM.  1981.  An 

epizootic  of  contagious  ecthyma  in  Rocky  Mountain 
bighorn  sheep  in  Colorado.  J.  Wildl.  Diseases  17:601- 
603. 

LANGE,  RE.,  A.  SANDOVAL,  and  W.P.  MELANEY.  1980. 
Psoroptic  scabies  in  bighorn  sheep  in  New  Mexico.  J. 
Wildl.  Diseases  16:77-82. 

LaPERRIERE,  A.J.,  PC.  LENT,  W.C.  GASAWAY,  and  FA. 
NODLER.  1980.  Use  of  Landsat  data  for  moose-habitat 
analyses  in  Alaska.  J.  Wildl.  Manage.  44:881-887. 

LASSEN,  P.  1984.  Muskox  distribution  and  population 
structure  in  Jameson  Land,  northeast  Greenland, 
1981-1983.  Pages  19-24  in  Klein,  DR.,  R.G.  White, 
and  S.  Keller,  eds.  Proc.  First  Int.  Muskox  Symp.  Spec. 
Rep.,  Biol.  Pap.  Univ.  Alaska,  Fairbanks. 

LEADER- WILLIAMS,  N.  1980.  Population  ecology  of  rein- 
deer on  South  Georgia.  Pages  664-676  in  Reimers, 
E.,  E.  Gaare,  and  S.  Skjenneberg,  eds.  Proc.  Second  Int. 
Reindeer/Caribou  Symp.  Dir.  for  Vilt  og  Ferskvanns- 
fisk, Trondheim. 

LENT,  PC.  1966.  The  caribou  of  northwestern  Alaska. 
Pages  481-517  in  Reimers,  E.,  E.  Gaare,  and  S.  Skjen- 
neberg, eds.  Proc.  Second  Int.  Reindeer/Caribou 
Symp.  Dir.  for  Vilt  og  Ferskvannsfisk,  Trondheim. 

.  1971.  Muskox  management  controversies  in  North 

America.  Biol.  Conserv.  3:255-263. 

.  1974a.  Mother-infant  relationships  in  ungulates. 

Pages  14-55  in  Geist,  V.  and  F.  Walthers,  eds.  The  Be- 
havior of  Ungulates  and  Its  Relation  to  Management. 


560 


Ungulates 


Vol.  I,  IUCN  Publ.  New  Series  24.  Morges,  Switzer- 
land. 940pp. 

— .  1974b.  A  review  of  rutting  behavior  in  moose. 
Naturaliste  Canadien  101:307-323. 

— .  1974c.  Ecological  and  behavioral  study  of  the 
Nunivak  Island  muskox  population.  Unpubl.  Rep.  U.S. 
Dep.  Inter.,  Fish  and  Wildl.  Serv.,  Anchorage  and 
Bethel,  AK.  90pp. 

— .  1978.  Muskox.  Pages  135-147  in  Schmidt,  J.L.  and 
D.L.  Gilbert,  eds.  Big  Game  of  North  America:  Ecology 
and  Management.  The  Wildl.  Manage.  Inst.,  Washing- 
ton, DC.  and  Stackpole  Books,  Harrisburg,  PA. 

1980.  Synoptic  snowmelt  patterns  in  arctic  Alaska 


in  relation  to  caribou  habitat  use.  Pages  71-83  in 
Reimers,  E.,  E.  Gaare,  and  S.  Skjenneberg,  eds.  Proc. 
Second  Int.  Reindeer/Caribou  Symp.  Dir.  for  Vilt  og 
Ferskvannsfisk,  Trondheim. 

LeRESCHE,  RE.  1974.  Moose  migrations  in  North  Amer- 
ica. Naturaliste  Canadien  101:393-415. 

and  A.  RAUSCH.  1974.  Accuracy  and  precision  of 

aerial  moose  censusing.  J.  Wildl.  Manage.  38:175-182. 

,  U.S.  SEAL,  P.D.  KARNS,  and  AW.  FRANZMANN. 


1974.  A  review  of  blood  chemistry  of  moose  and 
other  Cervidae  with  emphasis  on  nutritional  assess- 
ments. Naturaliste  Canadien  101  263-290. 

LESLIE,  DM,  Jr.  and  C.L.  DOUGLAS.  1979.  Desert  bighorn 
sheep  of  the  River  Mountains,  Nevada.  Wildl.  Monogr. 
66:1-56. 

LONGHURST,  W.M.,  H.K  OH,  MB.  JONES,  and  RE.  KEP- 
NER.  1 968.  A  basis  for  the  palatability  of  deer  forage 
plants.  Trans.  North  Am.  Wildl.  and  Nat.  Resour.  Conf. 
33:181-189. 

LONO,  O.  I960.  Transplantation  of  the  muskox  in  Europe 
and  North  America.  Meddel.  Norsk  Polarinst.  84:3-25. 

LOW,  W.A.  1970.  The  influence  of  aridity  on  reproduction 
of  the  collared  peccary  {Dicotyles  tajacu  [  Linn  ] )  in 
Texas.  PhD  dissertation,  Univ.  British  Columbia,  Van- 
couver. 170pp. 

LYON,  L.J.,  T.N.  LONNER,  J.P.  WEIGAND,  C.L.  MARCUM, 
WD.  EDGE,  J.D.  JONES,  D.W.  McCLEEREY,  and  L.L. 
HICKS.  1985.  Coordinating  elk  and  timber  manage- 
ment. Final  Rep.  Montana  Coop.  Elk-Logging  Study, 
1970-1985.  Montana  Dep.  Fish,  Wildlife,  and  Parks, 
Helena. 

and  A.L.  WARD.  1982.  Elk  and  land  management. 

Pages  443-478  in  Thomas,  J.W.  and  D.E.  Toweill,  eds. 
Elk  of  North  America:  Ecology  and  Management. 
Stackpole  Books,  Harrisburg,  PA,  698pp. 

MACKIE,  R.J.,  K.L.  HAMLIN,  and  D.F.  PAC.  1982.  Mule 
deer.  Pages  862-877  in  Chapman,  J.A.  and  G.A.  Feld- 
hamer,  eds.  Wild  Mammals  of  North  America.  The 
Johns  Hopkins  Univ.  Press,  Baltimore,  MD. 

MacPHERSON,  S.,  F.K.  MARTINSON,  and  AY.  COOPER- 
RIDER.  1982.  Forage  allocation  between  big  game 
and  domestic  livestock — new  approaches.  Proc.  West- 
ern Assoc.  Fish  and  Wildl.  Agencies  62:120-129. 

MAUTZ,  WW.  1978.  Nutrition  and  carrying  capacity. 

Pages  321-348  in  Schmidt,  J.L.  and  D.L.  Gilbert,  eds. 
Big  Game  of  North  America:  Ecology  and  Manage- 
ment. The  Wildl.  Manage.  Inst.,  Washington,  DC.  and 
Stackpole  Books,  Harrisburg,  PA. 

McCOLLOUGH,  S.A.  1982.  Impact  of  cattle  grazing  on 
bighorn  sheep,  Trickle  Mountain,  Colorado.  M.S.  the- 
sis, Colorado  State  Univ.,  Fort  Collins.  1 19pp. 

.,  AY.  COOPERRIDER,  and  J.A.  BAILEY.  1980.  Im- 
pact of  cattle  grazing  on  bighorn  sheep  at  Trickle 
Mountain,  Colorado.  Biennial  Symp.,  Northern  Wild 


Sheep  and  Goat  Council  2:42-58. 

McEWAN,  EH.  and  P.E.  WHITEHEAD.  1970.  Seasonal 

changes  in  the  energy  and  nitrogen  intake  in  reindeer 
and  caribou.  Canadian  J.  Zoology  48:905-913 

McFETRIDGE,  R.J.  1977.  Strategy  of  resource  use  by 
mountain  goats  nursery  groups.  Pages  169-173  in 
Samuel,  W.  and  W.G.  MacGregor,  eds.  Proc.  First  Int. 
Mountain  Goat  Symp.  Kalispell,  MT. 

MEAGHER,  M.M.  1973.  The  bison  of  Yellowstone  National 
Park.  U.S.  Dep.  Inter.,  Natl.  Park  Serv.  Sci.  Monogr.  1. 
161pp. 

.  1976.  Winter  weather  as  a  population  regulating 

influence  on  free-ranging  bison  in  Yellowstone  Na- 
tional Park.  Pages  29-38  in  Research  in  the  Parks. 
Trans.  Natl.  Park  Centennial  Symp.,  AAAS,  December 
28-29,  1971.  Series  1.  U.S.  Govt.  Printing  Office. 
232pp. 

1978.  Bison.  Pages  123-133  in  Schmidt,  J.L.  and 


D.L.  Gilbert,  eds.  Big  Game  of  North  America:  Ecology 
and  Management.  The  Wildl.  Manage.  Inst.,  Washing- 
ton, DC.  and  Stackpole  Books,  Harrisburg,  PA.  494pp. 

MONSON,  G.  and  L.  SUMNER,  eds.  1980.  The  desert  big- 
horn. Univ.  Arizona  Press,  Tucson.  370pp. 

MOULD,  E.  1977.  Habitat  relationships  of  moose  in  north- 
ern Alaska.  North  Am.  Moose  Conf.  13:144-156. 

.  1979.  Seasonal  movements  related  to  habitat  of 

moose  along  the  Colville  River,  Alaska.  Murrelet 
60:6-11. 

MULE,  R.S.  1982.  An  assessment  of  a  wildlife  habitat  evalu- 
ation methodology  for  Alaska.  M.S.  thesis,  Univ. 
Alaska,  Fairbanks.  215pp. 

MURIE,  O.J.  1954.  A  field  guide  to  animal  tracks.  Hough- 
ton Mifflin  Co.,  Boston,  MA.  375pp. 

NAGY,  J.G.,  H.W.  STEINHOFF,  and  GM.  WARD.  1964. 
Effects  of  essential  oils  of  sagebrush  on  deer  rumen 
microbial  function.  J.  Wildl.  Manage.  28:785-790. 

NEAL,  B.J.  1959.  A  contribution  on  the  life  history  of  the 
collared  peccary  in  Arizona.  Am.  Midi.  Nat.  61:177- 
190. 

NEFF,  D.J.  1968.  The  pellet  group  count  technique  for  big 
game  trend,  census,  and  distribution:  A  review.  J. 
Wildl.  Manage.  32:597-614. 

NELSON,  JR.  1984.  A  modeling  approach  to  large  herbi- 
vore competition.  Pages  491-524  in  Developing  Strat- 
egies for  Rangeland  Management.  Rep.  prepared  by 
the  Committee  on  Developing  Strategies  for  Range- 
land  Manage.,  Natl.  Res.  Council/Natl.  Acad.  Sciences. 
Westview  Press,  Boulder,  CO.  2022pp. 

NELSON,  M.E.  1979.  Home  range  location  of  white-tailed 
deer.  U.S.  Dep.  Agric,  For.  Serv.  Res.  Pap.  NC-173- 
10pp. 

and  L.C.  MECH.  1981.  Deer  social  organization  and 

wolf  predation  in  northeastern  Minnesota.  Wildl. 
Monogr.  77.  53pp. 

NICHOLS,  L,  Jr.  1980.  Aerial  census  and  classification  of 
mountain  goats  in  Alaska.  Pages  523-589  in  Hickey, 
W.O.,  ed.  Proc.  Biennial  Symp.,  Northern  Wild  Sheep 
and  Goat  Council.  Salmon,  ID. 

NPR-A  (NATIONAL  PETROLEUM  RESERVE— ALASKA) 
TASK  FORCE.  1978.  Fish  and  wildlife  resources.  Val- 
ues and  Resource  Analysis,  Section  6,  Vol.  3,  Study- 
Rep.  2.  U.S.  Dep.  Inter.  224pp. 

.  1979.  Ecological  profile.  National  Petroleum  Re- 
serve in  Alaska.  Study  25,  Rep.  4.  U.S.  Dep.  Inter. 

OLDEMEYER,  J.L.  1977a.  Nutritive  value  of  moose  forage. 
Naturaliste  Canadien  101:217-226. 

.  1 977b.  Impact  of  LeTourneau  tree  crushers  on 


Ungulates 


561 


moose  habitat  on  the  Kenai  National  Moose  Range. 
Proc.  28th  Annual  Conf.,  Northwest  Sections.  The 
Wildl.  Soc,  Kalispell,  MT. 

— ,  W.J.  BARMORE,  and  D.L.  GILBERT.  1971.  Winter 
ecology  of  bighorn  sheep  in  Yellowstone  National 
Park.  J.  Wildl.  Manage.  35:257-269. 
— ,  AW.  FRANZMANN,  A.L.  BRUNDAGE,  P.D.  ARNE 


SON,  and  A.  FLYNN.  1977.  Browse  quality  and  the 
Kenai  moose  population.  J.  Wildl.  Manage.  41:533- 
542. 

ONDERKA,  D.K  and  WD.  WISHART.  1984.  A  major  big- 
horn sheep  die-off  from  pneumonia  in  southern  Al- 
berta. Biennial  Symp.,  Northern  Wild  Sheep  and  Goat 
Council.  Salmon,  ID.  4:356-363. 

OUGH,  WD.  and  J.C.  DE  VOS,  Jr.  1984.  Intermountain 
travel  corridors  and  their  management  implications 
for  bighorn  sheep.  Desert  Bighorn  Council  Trans. 
32-36. 

PARKER,  G.R.  1975.  A  review  of  aerial  surveys  used  for 
estimating  the  numbers  of  barren-ground  caribou  in 
northern  Canada.  Polar  Record  17:627-638. 

.  1981.  Physical  and  reproductive  characteristics  of 

an  expanding  woodland  caribou  population  in  north- 
ern Labrador.  Canadian  J.  Zoology  59:1929-1940. 

and  R.K  ROSS.  1976.  Summer  habitat  use  by  musk- 
oxen  (Ovibos  moschatus)  and  Peary  caribou  (Rangi- 
fer  tarandus pearyi)  in  the  Canadian  High  Arctic. 
Polarforschung  46:12-25. 

,  DC.  THOMAS,  E.  BROUGHTON,  and  DR.  GRAY. 


1975.  Crashes  of  muskox  and  Peary  caribou  popula- 
tions in  1973-74  on  the  Parry  Island,  Arctic  Canada. 
Canadian  Wild.  Serv.  Progress  Note  56:1-10. 

PEDEN,  D.G.,  GM.  VAN  DYNE,  R.W.  RICE,  and  R.M.  HAN- 
SEN. 1974.  The  trophic  ecology  of  Bison  bison  L. 
on  shortgrass  plains.  J.  Applied  Ecol.  1 1 :489-498. 

PEEK,  J.M.  1974.  On  the  nature  of  winter  habitats  of 
Shiras  moose.  Naturaliste  Canadien  101:131-141. 

,  M.D.  SCOTT,  L.J.  NELSON,  and  DJ.  PIERCE.  1982. 

Role  of  cover  in  habitat  management  for  big  game 
in  northwestern  United  States.  Trans.  North  Am. 
Wildl.  Nat.  Resour.  Conf.  47:363-373- 

PEGAU,  RE.  and  J.E.  HEMMING.  1970.  Caribou  report. 
Alaska  Dep.  of  Fish  and  Game,  Federal  Aid  in  Wildl. 
Restoration  Project  Rep.,  Vol.  12.  Juneau.  224pp. 

PETOCZ,  R.G.  1973.  The  effect  of  snow  cover  on  the  so- 
cial behaviour  of  bighorn  rams  and  mountain  goats. 
Canadian  J.  Zoology  51:987-993. 

POHLE,  C.  1981.  Trergartnerische  Beobachtungen  bei  der 
Haltung  and  Zucht  von  Moschusochsen  in  Tierpark 
Berlin.  Zool.  Gart.  51:289-322. 

POJAR,  T.M.  1981.  A  management  perspective  of  popula- 
tion modeling.  Pages  241-261  in  Fowler,  C.W.  and 
T.D.  Smith,  eds.  Dynamics  of  Large  Mammal  Popula- 
tions. John  Wiley  and  Sons,  New  York,  NY.  477pp. 

and  D.  STRICKLAND,  eds.  1979.  A  workshop  on 

the  status  and  application  of  big  game  population 
modeling.  Colorado  Div.  Wildl.,  Fort  Collins.  53pp. 

POLEQUIN,  A.,  B.  SCHERRER,  and  R.  JOYAL.  1977.  Char 
acteristics  of  winter  browsing  areas  of  moose  in  west- 
ern Quebec  as  determined  by  multivariate  analysis. 
North  Am.  Moose  Conf.  13:128-143. 

PRENZLOW,  E.J.,  D.L.  GILBERT,  and  FA.  GLOVER.  1968. 
Some  behavior  patterns  of  the  pronghorn.  Colorado 
Dep.  Game,  Fish,  and  Parks  Spec.  Rep.  17.  16pp. 

QUIMBY,  D.C.  and  J.E.  GAAB.  1957.  Mandibular  dentition 
as  an  age  indicator  in  Rocky  Mountain  elk.  J.  Wildl. 
Manage.  21(4):435-451. 


REIMERS,  E.,  DR.  KLEIN,  and  R.  SORUMGORD.  1983. 

Calving  time,  growth  rate  and  body  size  of  Norwegian 
reindeer  on  different  ranges.  Arctic  and  Alpine  Res. 
15:107-118. 

RENECKER,  LA.  and  R.J.  HUDSON.  1986.  Seasonal  forag- 
ing rates  of  free-ranging  moose.  J.  Wildl.  Manage. 
50:143-147. 

RICHARDSON,  AH.  1971.  The  Rocky  Mountain  goat  in 
the  Black  Hills.  South  Dakota  Dep.  Game,  Fish,  and 
Parks  Bull.  2.  25pp. 

RIDEOUT,  C.B.  1974.  A  radio  telemetry  study  of  the  ecol- 
ogy and  behavior  of  the  Rocky  Mountain  goat  in 
western  Montana.  PhD  thesis,  Univ.  Kansas,  Lawrence. 
1 46pp. 

RINEY,  T.  1982.  Study  and  management  of  large  mammals. 
John  Wiley  and  Sons,  New  York,  NY.  552pp. 

RISENHOOVER,  KL.  and  J.A.  BAILEY.  1982.  Social  dynam- 
ics of  mountain  goats  in  summer:  Implications  for 
age  ratios.  Pages  364-373  in  Bailey,  J.A.  and  G.G. 
Schoonveld,  eds.  Proc.  Third  Biennial  Symp.,  Northern 
Wild  Sheep  and  Goat  Council.  Fort  Collins,  CO. 

and .  1985a.  Relationships  between  group 

size,  feeding  time,  and  agonistic  behavior  of  mountain 
goats.  Canadian  J.  Zoology  63:2501-2506. 

and .  1985b.  Foraging  ecology  of  mountain 

sheep:  Implications  for  habitat  management.  J.  Wildl. 
Manage.  49:797-804. 

ROBBINS,  C.T.  1973.  The  biological  basis  for  the  determi- 
nation of  carrying  capacity.  PhD  dissertation,  Cornell 
Univ.,  Ithaca,  NY.  239pp. 

ROBINETTE,  W.L.,  CM.  LOVELESS,  and  DA.  JONES.  1974. 
Field  tests  of  strip  census  methods.  J.  Wildl.  Manage. 
38:81-96. 

ROBUS,  M.  1984.  Summer  food  habits  of  muskoxen  in 
northeastern  Alaska.  Pages  81-85  in  Klein,  D.R.,  R.G. 
White,  and  S.  Keller,  eds.  Proc.  First  Int.  Muskox 
Symp.  Spec.  Rep.,  Biol.  Pap.  Univ.  Alaska,  Fairbanks. 

RYEL,  L.A.  1971.  Evaluation  of  pellet  group  surveys  for 
estimating  deer  populations  in  Michigan.  PhD  disser- 
tation, Michigan  State  Univ.,  Lansing.  237pp. 

SAMUEL,  W.,  G.A.  CHALMERS,  J.G.  STELLOX,  A.  LOWEN, 
and  J.J.  THOMSEN.  1975.  Contagious  ecthyma  in 
bighorn  sheep  and  mountain  goat  in  western  Canada. 
J.  Wildl.  Diseases  11:26-31. 

and  W.  MacGREGOR,  eds.  1977.  Proceedings  of  the 

first  international  mountain  goat  symposium.  Kalispell, 
MT. 

SAUSMAN,  K  1982.  Survival  of  captive-born  Ovis  cana- 
densis in  North  American  zoos.  Desert  Bighorn  Coun- 
cil Trans.  26-31. 

SCHMIDT,  J.L.  and  D.L.  GILBERT.  1978.  Big  game  of  North 
America:  Ecology  and  Management.  The  Wildl.  Man- 
age. Inst.,  Washington,  DC  and  Stackpole  Books,  Har- 
risburg,  PA.  494pp. 

SCHMIDT,  R.L.,  C.P.  HIBLER,  T.R.  SPRAKER,  and  W.H. 

RUTHERFORD.  1979.  An  evaluation  of  drug  treatment 
for  lungworm  in  bighorn  sheep.  J.  Wildl.  Manage. 
43:461-467. 

,  W.H.  RUTHERFORD,  and  F.M.  BODENHAM.  1978. 

Colorado  bighorn  sheep-trapping  techniques.  Wildl. 
Soc.  Bull.  6:159-163. 

SCHOEN,  J.W.  and  M.D.  KIRCHOFF  1981.  Habitat  use  by 
mountain  goats  in  southeast  Alaska.  Alaska  Dep.  Fish 
and  Game,  PR  Project  Rep.  W-17-10,  11  and  W-21-1, 
2.  67pp. 

SERGEANT,  DE.  and  D.H.  PIMLOTT.  1959.  Age  determi- 
nation in  moose  from  sectioned  incisor  teeth.  J.  Wildl. 


562 


Ungulates 


Manage.  23:315-321. 

SHANNON,  N.H.,  R.J.  HUDSON,  V.C.  BRINK,  and  WD. 
KITTS.  1975.  Determinants  of  spatial  distribution  of 
Rocky  Mountain  bighorn  sheep.  J.  Wildl.  Manage. 
39:387-401. 

SIMMONS,  N.M.  1969.  Heat  stress  and  bighorn  behavior  in 
the  Cabeza  Prieta  Game  Range,  Arizona.  Desert  Big- 
horn Council  Trans.  13:55-63- 

SINGER,  F.J.  1975.  Behavior  of  mountain  goats,  elk  and 
other  wildlife  in  relation  to  U.S.  Highway  2,  Glacier 
National  Park.  Completion  Rep.  U.S.  Dep.  Inter.,  Natl. 
Park  Serv.,  West  Glacier,  MT.  98pp. 

SINIFF,  D.B.  and  R.O.  SKOOG.  1964.  Aerial  censusing  of 
caribou  using  stratified  random  sampling.  J.  Wildl. 
Manage.  28:39-40. 

SKILES  J.W.,  P.T.  KORTOPATES,  and  G.M.  VAN  DYNE. 
1980.  Optimization  models  for  forage  allocation  to 
combinations  of  large  herbivores  for  grazing  land 
situations:  A  critical  use  of  proper  use  factors.  Rep.  to 
U.S.  Dep.  Inter.,  Bur.  Land  Manage.  86pp. 

SKOOG,  R.O.  1962.  Method  for  estimating  caribou  herds. 
Alaska  Dep.  of  Fish  and  Game  Information  Leaflet 
20.,  Juneau.  6pp. 

SKOVLIN,  J.M.  1982.  Habitat  requirements  and  evalua- 
tions. Pages  368-413  in  Thomas,  J.W.  and  D.E.  Tow- 
eill,  eds.  Elk  of  North  America:  Ecology  and 
Management.  Stackpole  Books,  Harrisburg,  PA.  698pp. 

SMITH,  B.L.  1976.  Ecology  of  Rocky  Mountain  goats  in 
the  Bitterroot  Mountains,  Montana.  M.S.  thesis,  Univ. 
Montana.  203pp 

.  1977.  Influence  of  snow  conditions  on  winter 

distribution,  habitat  use,  and  group  size  of  mountain 
goats.  Pages  1 74- 1 89  in  Samuel,  W.  and  W.G.  Mac- 
Gregor,  eds.  Proc.  First  Int.  Mountain  Goat  Symp. 
Kalispell,  MT. 

SMITH,  C.A.  and  K.J.  RAEDEKE.  1982.  Group  size  and 
movements  of  a  dispersed,  low  density  goat  popula- 
tion, with  comments  on  inbreeding  and  human  im- 
pacts. Pages  54-67  in  Bailey,  J.A.  and  G.G. 
Schoonveld,  eds.  Proc.  Third  Biennial  Symp.,  Northern 
Wild  Sheep  and  Goat  Council.  Fort  Collins,  CO. 

SMITH,  KG.  1982.  Winter  studies  of  forest-dwelling 
mountain  goats  of  Pinto  Creek,  Alberta.  Pages  374- 
390  in  Bailey,  J.A.  and  G.G.  Schoonveld,  eds.  Proc. 
Third  Biennial  Symp.,  Northern  Wild  Sheep  and  Goat 
Council.  Fort  Collins,  CO. 

SMITH,  T.E.  1976.  Reproductive  behavior  and  related 

social  organization  of  the  muskox  on  Nunivak  Island. 
M.S.  thesis,  Univ.  Alaska,  Fairbanks.  1 38pp. 

SOULE,  ME.  and  B.A.  WILCOX,  eds.  1980.  Conservation 
biology,  an  evolutionary-ecological  perspective.  Sin- 
auer  Assoc,  Sunderland,  MA.  395pp. 

SOWLS,  L.K  1978.  Collared  peccary.  Pages  191-205  in 

Schmidt,  J.L.  and  D.L.  Gilbert,  eds.  Big  Game  of  North 
America:  Ecology  and  Management.  The  Wildl.  Man 
age.  Inst.,  Washington,  DC.  and  Stackpole  Books. 
Harrisburg,  PA. 

1984.  The  peccaries.  Univ.  Arizona  Press,  Tucson. 

251pp 
SPENCER,  D.L.  and  C.  LENSINK  1970.  The  muskox  of 
Nunivak  Island.  J.  Wildl.  Manage.  34:1-15. 

SPRAKER,  T.R.  and  C.P.  HIBLER  1977.  Summer  lamb 
mortality  of  Rocky  Mountain  bighorn  sheep.  Desert 
Bighorn  Council  Trans.  21:11-12. 

and .  1982.  An  overview  of  the  clinical 

signs,  gross  and  histological  lesions  of  the  pneumonia 
complex  of  bighorn  sheep.  Proc.  Biennial  Symp., 


Northern  Wild  Sheep  and  Goat  Council.  Fort  Collins, 
CO.  3:163-172. 
— , ,  G.G.  SCHOONVELD,  and  W.S.  ADNEY. 


1984.  Pathologic  changes  and  microorganisms  found 
in  bighorn  sheep  during  a  stress-related  die-off.  J. 
Wildl.  Diseases  20:319-327. 

STELFOX,  J.G.  1971.  Bighorn  sheep  in  the  Canadian  Rock- 
ies: A  history  1800-1970.  Canadian  Field-Naturalist 
85:101-122. 

.  1976.  Wood  Buffalo  National  Park,  bison  research 

1972-76.  Section  A-I  in  1976  Annual  Report.  Cana- 
dian Wildl.  Serv.  and  Parks,  Ottawa. 

SUNDSTROM,  C.  1968.  Water  consumption  by  pronghorn 
antelope  and  distribution  related  to  water  in  Wyo- 
ming's Red  Desert.  Proc.  Biennial  Antelope  States 
Workshop  3:39-46. 

SWEENEY,  J.M.  and  H.W.  STEINHOFF.  1976.  Elk  move- 
ments and  calving  as  related  to  snow  cover.  Pages 
415-436  in  Steinhoff,  H.W.  and  J.D.  Ives,  eds.  Ecologi- 
cal Impacts  of  Snowpack  Augmentation  in  the  San 
Juan  Mountains,  Colorado.  Colorado  State  Univ.,  Fort 
Collins. 

TABER,  R.D.  and  R.F.  DASMANN.  1958.  The  black-tailed 
deer  of  the  chaparral — its  life  history  and  manage- 
ment in  the  North  Coast  Range  of  California.  Califor- 
nia Dep.  Fish  and  Game,  Game  Bull.  8.  163pp. 

TANKERSLEY,  N.G.  and  W.C  GASAWAY.  1983.  Mineral 
lick  use  by  moose  in  Alaska.  Canadian  J.  Zoology 
61:2242-2249. 

TELFER,  E.S.  1974.  Logging  as  a  factor  in  wildlife  ecology 
in  the  boreal  forest.  Forest  Chronical  50:186-190. 

TENER,  J.S.  1965.  Muskoxen  in  Canada.  Canadian  Wildl. 
Monogr.  Series  2.  Dep.  Northern  Affairs  and  Nat. 
Resour.  Queens  Printer,  Ottawa.  166pp. 

THING,  H.  1984.  Food  and  habitat  selection  by  muskoxen 
in  Jameson  Land,  Northeast  Greenland:  A  preliminary 
report.  Pages  69-74  in  Klein,  D.R.,  R.G.  White,  and 
S.  Keller,  eds.  Proc.  First  Int.  Muskox  Symp.  Spec. 
Rep.,  Biol.  Pap.  Univ.  Alaska,  Fairbanks. 

and  B.  CLAUSEN.  1980.  Summer  mortality  among 

caribou  calves  in  Greenland.  Pages  434-437  in  Rei- 
mers,  E.,  E.  Gaare,  and  S.  Skjenneberg,  eds.  Proc.  Sec- 
ond Int.  Reindeer/Caribou  Symp.  Dir.  for  Vilt  og 
Ferskvannsnsk,  Trondheim. 

THOMAS,  D.C.  and  J.  EDMONDS.  1984.  Pages  93-100  in 
Klein,  D.R.,  R.G.  White,  and  S.  Keller,  eds.  Proc.  First 
Int.  Muskox  Symp.  Spec.  Rep.,  Biol.  Pap.  Univ.  Alaska, 
Fairbanks. 

,  F.L.  MILLER,  R.H.  RUSSELL,  and  G.R.  PARKER. 

1981.  The  Bailey  Point  Region  and  other  muskox 
refugia  in  the  Canadian  Arctic:  A  short  review.  Arctic 
34:34-36. 

THOMAS,  J.W.,  H.  BLACK  Jr.,  R.J.  SCHERZINGER,  and  R.J. 
PEDERSEN  1979.  Deer  and  elk.  Pages  104-127  in 
Thomas,  J.W.,  ed.  Wildlife  Habitats  in  Managed  For- 
ests: The  Blue  Mountains  of  Oregon  and  Washington. 
U.S.  Dep.  Agric,  For.  Serv.  Handbook  553-  Washing- 
ton, DC. 

and  D.E.  TOWEILL.  1982.  Elk  of  North  America: 

Ecology  and  Management.  Stackpole  Books,  Harris- 
burg, PA.  698pp. 

THOMPSON,  R.W.  1980.  Population  dynamics,  habitat 

utilization,  recreational  impacts  and  trapping  of  intro- 
duced Rocky  Mountain  goats  in  the  Eagles  Nest  Wil- 
derness Area,  Colorado.  Pages  459-464  in  Hickey, 
W.O.,  ed.  Proc.  1980  Biennial  Symp.,  Northern  Wild 
Sheep  and  Goat  Council.  Salmon,  ID.  668pp. 


Ungulates 


563 


and  R.J.  GUENZEL.  1978.  Status  of  the  introduced 

mountain  goats  in  the  Eagles  Nest  Wilderness  Area, 
Colorado.  Pages  175-197  in  Hebert,  D.M.  and  M. 
Nation,  eds.  Proc.  1978  Biennial  Symp.  Northern  Wild 
Sheep  and  Goat  Conf.  Penticton,  British  Columbia. 

TIMMERMANN,  H.R.  1974.  Moose  inventory  methods:  A 
review.  Naturaliste  Canadien  101:615-629. 

TODD,  J. W.  1972.  A  literature  review  on  bighorn  sheep 
food  habits.  Colorado  Div.  Game,  Fish,  and  Parks 
Spec.  Rep.  27.  21pp. 

TREFETHEN,  J.B.,  ed.  1975.  The  wild  sheep  in  modern 
North  America.  Winchester  Press,  New  York,  NY. 
302pp. 

.  1975.  The  Wild  Sheep  in  Modern  North  America. 

Proc.  Workshop  on  the  Manage.  Biol.  North  American 
Wild  Sheep.  Univ.  Montana,  Missoula,  June  18-20, 
1974.  Boone  and  Crockett  Club  and  the  Winchester 
Press,  New  York,  NY.  302pp. 

TURNER,  J.C  and  RA.  WEAVER.  1980.  Water.  Pages  100- 
112  in  Monson,  G.  and  L.  Sumner,  eds.  The  Desert 
Bighorn — Its  Life  History,  Ecology,  and  Management. 
Univ.  Arizona  Press,  Tucson. 

U.S.  DEPARTMENT  OF  THE  INTERIOR,  FISH  AND 
WILDLIFE  SERVICE.  1983.  Initial  report,  baseline 
study  of  the  fish,  wildlife  and  their  habitats.  Arctic 
Natl.  Wildl.  Refuge  Coastal  Plain  Resour.  Assessment. 
U.S.  Dep.  Inter.,  Fish  and  Wild.  Serv.,  Anchorage,  AK 
507pp. 

VAN  DAELE,  L.J.  and  DR.  JOHNSON.  1983.  Estimation  of 
arboreal  lichen  biomass  available  to  caribou.  J.  Wildl. 
Manage.  47:888-890. 

VAN  DYKE,  W.A.,  A.  SANDS,  J.  YOAKUM,  A.  POLENZ,  AND 
J.  BLAISDELL.  1983.  Wildlife  habitats  in  managed 
rangelands,  the  Great  Basin  of  southeastern  Oregon: 
Bighorn  sheep.  U.S.  Dep.  Agric,  For.  Serv.,  Pacific 
Northwest  Experiment  Sta.,  Gen.  Tech.  Rep.  PNW- 
159.  37pp. 

VIBE,  C.  1958.  The  muskox  in  east  Greenland.  Mammalia 
22:168-174. 

.  1967.  Arctic  animals  in  relation  to  climatic  fluctua- 
tions. Meddel  om  Gronland  171:1-227. 

WAKELYN,  LA.  1984.  Analysis  and  comparison  of  existing 
and  historic  bighorn  sheep  ranges  in  Colorado.  M.S. 
thesis,  Colorado  State  Univ.,  Fort  Collins.  274pp. 

WALLMO,  O.C.,  ed.  1981.  Mule  and  black-tailed  deer  of 
North  America.  Univ.  Nebraska  Press,  Lincoln.  605pp. 

.  1981.  Mule  and  black-tailed  deer  distribution  and 

habitats.  Pages  1-25  in  Wallmo,  O.C.,  ed.  Mule  and 
Black-tailed  Deer  of  North  America.  Univ.  Nebraska 
Press,  Lincoln. 

,  L.C.  CARPENTER,  W.L.  REGELIN,  R.B.  GILL,  and 

D.B.  BAKER.  1977.  Evaluation  of  deer  habitat  on 
a  nutritional  basis.  J.  Range  Manage.  30:122-127. 

,  R.B.  GILL,  L.H.  CARPENTER,  and  D.W.  REICHERT. 


1973-  Accuracy  of  field  estimates  of  deer  food  habits. 
J.  Wildl.  Manage.  37:556-562. 

and  J.W.  SCHOEN.  1981.  Forest  management  for 


deer.  Pages  434-448  in  Wallmo,  O.C.,  ed.  Mule  and 
Black-tailed  Deer  of  North  America.  Univ.  Nebraska 
Press,  Lincoln.  605pp. 

WHITE,  G.C.,  DR.  ANDERSON,  KP.  BURNHAM,  and  D.L. 
OTIS.  1982.  Capture-recapture  and  removal  methods 
for  sampling  closed  populations.  Los  Alamos  Natl. 
Laboratory,  Los  Alamos,  NM.  235pp. 

WIGAL,  RA.  and  V.L.  COGGINS.  1982.  Mountain  goat. 
Pages  1008-1020  in  Chapman,  J.A.  and  GA.  Feld- 
hamer,  eds.  Wild  Mammals  of  North  America.  Johns 
Hopkins  Univ.  Press,  Baltimore,  MD.  1 147pp. 

WILKINSON,  P.F.,  C.C.  SHANK  and  D.E.  PENNER.  1976. 
Muskox-caribou  summer  range  on  Banks  Island, 
Northwest  Territories.  J.  Wildl.  Manage.  40:151-162. 

and  P.N.  TEAL.  1984.  The  muskox  domestication 

project:  An  overview  and  evaluation.  Pages  162-166 
in  Klein,  DR.,  R.G.  White,  and  S.  Keller,  eds.  Proc. 
First  Int.  Muskox  Symp.  Spec.  Rep.,  Biol.  Pap.  Univ. 
Alaska,  Fairbanks. 

WILSON,  L.W.,  J.  BLAISDELL,  G.  WELSH,  R.  WEAVER,  R. 
BRIGHAM,  W.  KELLY,  J.  YOAKUM,  M.  HINKS,  J. 
TURNER,  and  J.  DeFORGE.  1980.  Desert  bighorn 
habitat  requirements  and  management  recommenda- 
tions. Trans.  Desert  Bighorn  Council  24:1-7. 

WISDOM,  M.J.,  L.R.  BRIGHT,  C.G  CAREY,  WW.  HINES, 
R.J.  PEDERSEN,  DA.  SMITHEY,  J.W.  THOMAS,  and 
F.W.  WITMER.  1986.  A  model  to  evaluate  elk  habitat 
in  western  Oregon.  U.S.  Dep.  Agric,  For.  Serv.,  Pacific 
Northwest  Region.  36pp. 

WITMER,  G.W.,  M.  WISDOM,  E.P.  HARSHMAN,  R.J.  AN- 
DERSON, C.  CAREY,  MP.  KUTTEL,  ID.  LUMAN,  J.A. 
ROCHELLE,  R.W.  SCHARPF,  and  D.  SMITHEY.  1985. 
Deer  and  elk.  Pages  231-258  in  Brown,  E.R.,  ed.  Man- 
agement of  Wildlife  and  Fish  Habitats  in  Forests  of 
Western  Oregon  and  Washington.  U.S.  Dep.  Agric, 
For.  Serv.  Pub.  R6-F&WL-192-1985.  Portland,  OR. 

WOOD,  J.E.,  T.S.  BICKLE,  W.  EVANS,  J.C.  GERMANY,  and 
V.W.  HOWARD,  Jr.  1970.  The  Fort  Stanton  mule  deer 
herd  (some  ecological  and  life  history  characteristics 
with  special  emphasis  on  the  use  of  water).  New 
Mexico  State  Univ.,  Agric.  Experiment  Sta.  Bull.  567. 
32pp. 

YOAKUM,  J.D.  1978.  Pronghorn.  Pages  103-122  in 

Schmidt,  J.L.  and  D.L.  Gilbert,  eds.  Big  Game  of  North 
America:  Ecology  and  Management.  The  Wildl.  Man- 
age. Inst.,  Washington,  DC.  and  Stackpole  Books, 
Harrisburg,  PA.  494pp. 

.  1980.  Habitat  management  guides  for  the  Ameri- 
can pronghorn  antelope.  U.S.  Dep.  Inter.,  Bur.  Land 
Manage.  Tech.  Note  347.  Denver  Federal  Center, 
Denver,  CO.  77pp. 


564 


Ungulates 


IV  HABITAT  MEASUREMENTS 

26  Soils 

27  Terrestrial  Physical  Features 

28  Aquatic  Physical  Features 

29  Hydrologic  Properties 

30  Water  Quality 

31  Vegetation 

32  Macroinvertebrates 


m 


■■■ 


26 

SOILS 


James  E.  Stone 

U.S.  Bureau  of  Land  Management 
Washington,  DC  20240 


". . .  there  is  a  remarkable  correlation  between  game 
supply  and  soil  fertility  throughout  North  America." 

—  Aldo  Leopold,  from  Game  Management 


Editor's  Note:  Aldo  Leopold  noted  the  relationship 
between  soils  and  wildlife  populations.  Since  then, 
many  wildlife  biologists  and  soil  scientists  have 
observed  the  linkage  between  soils,  vegetation,  and 
wildlife  populations. 

While  the  concept  of  relating  soils  to  vegetation 
and  wildlife  is  well  established,  the  predictive  rela- 
tionships are  not.  This  chapter  brings  together  in- 
formation available  on  these  relationships. 


INTRODUCTION 

Soil  is  a  vital  part  of  wildlife  habitat.  Indirectly, 
soil  affects  animals  through  its  influence  on  the  com- 
position, amount,  palatability,  and  nutritive  value  of 
vegetation.  In  addition,  soil  properties  directly  affect 
such  things  as  the  ease  of  digging,  burrow  stability 
and  depth,  maintenance  of  body  temperature  and 
moisture  levels,  availability  of  suitable  reproductive 
and  resting  sites,  mobility,  and  body  coloration. 

Until  recently,  the  consideration  of  soils  on 
wildlands  has  been  hampered  by  the  lack  of  available 
soil  information,  but  this  situation  is  rapidly  chang- 
ing. Millions  of  acres  of  lands  administered  by  the 
U.S.  Bureau  of  Land  Management  and  the  U.S.  Forest 
Service  have  now  been  surveyed,  and  ongoing  pro- 
grams are  in  place.  Acquiring  the  soil  information, 
however,  is  only  the  beginning.  Putting  this  informa- 
tion to  use  is  the  ultimate  goal. 

Several  years  ago,  Klemmedson  ( 1970)  wrote  of 
the  communication  gap  between  those  who  make 
soil  surveys  and  those  who  could  and  should  be 
using  the  information.  Although  progress  has  been 
made,  the  problem  still  exists.  Effective  application 
of  soils  data  in  habitat  management  requires  a  coop- 
erative effort  on  the  part  of  both  soil  scientists  and 
wildlife  biologists.  Soil  scientists  must  recognize  the 
needs  of  other  disciplines  and  tailor  surveys  and 
associated  soil  interpretations  to  meet  those  needs. 
Wildlife  biologists  in  turn  must  acquire  at  least  a 
working  knowledge  of  soil  terminology,  kinds  and 
sources  of  soil  information,  and  soil-animal  relation- 
ships so  as  to  fairly  assess  application  opportunities 
as  well  as  constraints. 


Soils 


567 


The  intent  of  this  chapter  is  to  provide  a  work- 
ing-level discussion  of  the  use  of  soil  information 
in  habitat  characterization  and  management. 

SOIL  DESCRIPTION 

Soil  is  a  mixture  of  solids  and  pores.  The  solid 
fraction  consists  of  mineral  particles  from  the  weath- 
ering of  rocks  and  organic  material  from  the  deposi- 
tion and  decay  of  plant  and  animal  tissue.  The 
intervening  pore  spaces  contain  either  air  or  an 
aqueous  solution  with  substances  either  dissolved  or 
in  suspension.  The  nature  and  relative  amounts  of 
these  constituents  are  described  according  to  spe- 
cific standards  set  forth  in  the  Soil  Survey  Manual 
(U.S.  Soil  Conservation  Service  1981 ).  A  representa- 
tive soil  description  is  given  in  Figure  1;  descriptive 
terminology  and  standards  are  summarized  below  for 
selected  soil  characteristics. 


Parent  Material 

Parent  material  is  the  unconsolidated  and  more 
or  less  chemically  weathered  mineral  material  from 
which  soils  develop.  Parent  materials  may  have  been 
formed  in  place  or  transported  by  water,  wind,  ice, 
or  gravity.  In  any  event,  parent  material  exerts  a 
strong  influence  on  soil  development  and  associated 
plant  and  animal  communities.  For  example,  soils 
derived  from  serpentine  parent  materials  are  gener- 
ally low  or  lacking  in  several  essential  plant  nutrients 
and  are  therefore  highly  selective  for  native  plant 
species  (Kruckeberg  1969).  Notable  plant-soil  rela- 
tionships are  also  evident  on  other  parent  materials 
(Marchand  1973).  In  addition,  the  nature  of  the  par- 
ent rock  itself  influences  the  availability  of  special- 
ized habitats  such  as  caves,  lava  tubes,  and  talus 
(Maser  et  al.  1979). 


DEPTH,  DRAINAGE 
PARENT  MATERIAL 


HORIZONATION 
ROOTS 

TEXTURE 
STRUCTURE 
ROCK  CONTENT 

COLOR 
CONSISTENCE 

REACTION  (pH) 


Kech  series 

The  Kech  series  consists  of  shallow,  well-drained  soils  that  formed 
in  place  and  from  locally  transported  sediment  from  sandstone  and 
interbedded  shale.  Kech  soils  are  on  upland  hills  and  ridges.  Slopes 
range  from  3  to  40%.  Average  annual  precipitation  is  about  14  inches, 
and  average  annual  air  temperature  is  about  44°F. 

Kech  soils  are  similar  to  Progresso  and  Potts  soils  and  are  near 
Scholle,  Mesa,  and  Agua  Fria  soils.  Progresso  soils  have  sandstone 
bedrock  at  a  depth  of  20  to  40  inches.  Potts,  Scholle,  Mesa,  and  Agua 
Fria  soils  are  more  than  60  inches  deep.  Mesa  soils  are  dry.  Agua 
Fria  soils  have  more  than  35  percent  clay  in  the  B  horizon. 

Typical  pedon  of  Kech  loam,  in  an  area  of  Kech-Progresso  loams,  3 
to  15  percent  slopes,  SW1/4  Sec.  13,  T.  15  S.,  R.  93  W.,  in  Delta 
County: 

Al — 0  to  4  inches:  brown  (7.5YR  5/2)  loam,  dark  brown  (7.5YR  4/2) 
moist;  moderate  very  fine  granular  structure;  soft,  friable,  slightly 
sticky  and  slightly  plastic;  few  fine  and  medium  roots:  5  to  1 5% 
sandstone  channers;  neutral;  clear  smooth  boundary .  (3  to  5 
inches  thick). 

B2t — 4  to  12  inches;  brown  (7.5YR  5/-t )  clay  loam,  dark  brown 
(7.5YR  4/4)  moist;  moderate  medium  subangular  blockv 
structure  parting  to  moderate  fine  subangular  blocky;  hard, 
friable,  sticky  and  plastic;  few  fine  and  very  fine  roots;  moderate 
continuous  clay  films  on  peds;  5  to  10%  sandstone  channers: 
noncareous;  mildly  alkaline;  clear  wavy  boundary.  (4  to  9  inches 
thick). 

B3ca — 12  to  19  inches;  pinkish  white  (7.5YR  8/2)  channery  loam, 
pink  (7.5YR  7/4)  moist;  weak  medium  subangular  blocky 
structure  parting  to  weak  fine  subangular  blocky;  hard,  firm, 
slightly  sticky  and  slightly  plastic;  25  to  30%  sandstone  channers; 
few  thin  patchy  clay  films;  soft  masses  of  secondary  lime;  mildy 
alkaline;  clear  irregular  boundary  (4  to  8  inches  thick) 

R — 19  inches;  partially  weathered  calcareous  sandstone. 

The  A  horizon  has  value  of  4  to  6  dry  and  3  to  5  moist  and  has 
chroma  of  2  or  3  dry  and  moist.  The  B  horizon  has  hue  of  7.5YR  or 
10YR,  value  of  5  or  6  dry  and  4  or  5  moist,  and  chroma  of  3  or  4  dry 
and  moist.  Texture  of  the  B  horizon  is  clay  loam.  Clay  content  ranges 
from  27  to  35%  in  this  horizon,  and  sandstone  channers  range  from 
0  to  15%.  Reaction  is  neutral  or  mildly  alkaline.  The  B3  horizon 
contains  visible  secondary  lime  and  has  as  much  as  30%  sandstone 
channers.  Weathered  sandstone  bedrock  is  at  a  depth  of  10  to  20 
inches. 


h 


SETTING 


ASSOCIATED 
SOILS 


TYPE 
LOCATION 


PROFILE 
DESCRIPTION 


RANGE  OF 
CHARACTERISTICS 


Figure  1.     Soil  description. 


568 


Soils 


Horizonation 

Vertical  sections  through  the  soil  (soil  profile) 
often  reveal  the  presence  of  distinct  layers  or  hori- 
zons (Figure  2).  The  nature,  arrangement,  and  thick- 
ness of  horizons  (or  their  absence)  are  important 
for  differentiating  soils  and  for  interpreting  soil  be- 
havior. Each  soil  horizon,  designated  by  a  particular 
symbol  (e.g.,  Al — 0  to  4  in.),  is  described  separately. 


••>    ri--       '■%<&*$;>    '  "J,  '-     i       -'r'    *     '  ■    ;i  i   -~ 


mm* 


Clay(%) 


Silt(%) 


Sand(%) 
Figure  3.     Soil  texture. 

pendence  upon  soil  texture  of  18  races  of  rodents. 
Feldhamer  (1979)  reported  a  direct  relationship 
between  the  population  density  of  pocket  mice  (Per- 
ognathus  parvus)  and  the  percentage  of  sand  in  the 
soil;  an  inverse  relationship  existed  with  the  percent- 
age of  clay. 


Figure  2.     Soil  horizonation. 


Texture 

Texture  refers  to  the  size  distribution  of  mineral 
particles  within  the  soil.  Those  particles  smaller  than 
2  mm  (0.08  in.)  in  diameter,  termed  the  fine  earth 
fraction,  are  described  separately  from  larger  rock 
fragments.  Soils  are  assigned  textural  class  names 
(e.g.,  sandy  loam,  silty  clay)  based  on  the  amounts  of 
sand-,  silt-,  and  clay-sized  particles  in  the  fine  earth 
fraction  (Figure  3). 

Small  ground-dwelling  animals  often  show  an 
affinity  for  particular  soil  textures.  Stuart  (1932) 
related  the  distribution  of  several  lizard  species  to 
soil  texture;  Hardy  (1945)  observed  a  varying  de- 


Mojave  fringed-toe  lizard.  Completely  restricted  to  fine, 
loose,  windblown  sand. 


Rock  Fragments 

Rock  fragments,  those  mineral  particles  greater 
than  2  mm  (0.08  in.)  in  diameter,  are  described 
according  to  their  size  and  shape  (Figure  4).  The 
adjective  form  of  a  rock  fragment  class  is  often  used 
to  modify  the  textural  class  name  (e.g.,  channery 
loam ).  Rock  content  in  the  profile  is  described  as  a 
percentage  by  volume. 

Rock  fragments  in  the  soil  profile  reduce  mois- 
ture storage  as  well  as  the  volume  of  soil  material 


Soils 


569 


Fragment 


Centimeters 


Diameter 


Length 


Inches 


(f\p 


Gravel 


0.2-7.6 


0.1-3 


Cobble 


7.6-25 


3-10 


Channer 
(flat) 


0.2-15 


0.1-6 


~^  Flagstone 
(flat) 


15-38 


6-15 


Stone 


25-60 


10-24 


Boulder 


60 


24 


aA  single  fragment  of  this  size  is  called  a  "pebble. 
Figure  4.     Rock  fragments. 


that  plant  roots  and  animals  can  occupy.  Fragments 
on  the  surface  can  provide  specialized  habitats  for 
animals  in  areas  that  might  otherwise  be  unsuitable. 

Structure 

Soil  structure  refers  to  the  organization  of  indi- 
vidual mineral  particles  into  clumps  or  aggregates. 
Some  soils,  such  as  loose  sands,  are  structureless. 
Structured  soils  are  described  by  the  stability,  size, 
and  shape  of  the  aggregates  (e.g.,  moderate,  medium, 
subangular  blocky  structure  [Figure  5]).  Structure 
affects  the  size  of  pores  and  the  total  amount  of  pore 
space  in  the  soil,  thus  influencing  air  and  water  rela- 
tions. Unlike  texture,  soil  structure  can  be  altered 
rapidly  by  use  and  management  practices,  thereby  al- 
tering habitat  suitability  for  some  animals. 

Consistence 

Soil  consistence  characterizes  the  cohesion 
among  soil  particles  and  the  adhesion  between  soil 


and  other  substances.  One  element  of  consistence  is 
soil  strength,  the  degree  of  resistance  to  breaking 
or  crushing  when  force  is  applied.  The  standard 
terms  used  to  describe  soil  strength  are  specific  to 
moisture  condition  (Table  1). 

Soil  strength  is  important  when  considering 
habitat  suitability  for  small  burrowing  animals.  For 
example,  Feldhamer  ( 1979)  found  a  direct  correla- 
tion between  the  population  density  of  chipmunks 
(Eutamias  minimus)  and  soil  strength. 

Color 

Soil  color  is  described  by  comparison  with 
standard  Munsell  color  charts  (Munsell  1975).  The 
Munsell  System  uses  three  elements  (hue,  value, 
chroma)  to  make  up  a  specific  color  notation.  For 
example,  7.5  YR  8/2  is  the  hue  and  value/chroma 
notation  for  pinkish  white  (Figure  6).  Colors  of  both 
moist  and  dry  soil  are  commonly  recorded  in  the 
soil  description. 


570 


Soils 


Structural  Type 


General  Connotation 


Granular 


Found  in  surface  horizons;  formation  promoted  by  organic 
matter;  encourages  air  and  water  transmission. 


Platy 


Found  in  surface  and  subsurface  horizons;  severely 
restricts  air  and  water  transmission. 


Blocky 


Found  in  fine-textured  subsurface  horizons;  most  common 
in  humid  climates;  restricts  air  and  water  transmission. 


Prismatic 


Found  in  fine-textured  subsurface  horizons;  most  common 
in  arid  and  semiarid  climates;  restricts  air  and  water 
transmission. 


Columnar 


Found  in  fine-textured  subsurface  horizons;  most  common 
in  arid  and  semiarid  climates;  often  indicates  high  sodium 
levels;  restricts  air  and  water  transmission. 


Figure  5.     Soil  structure. 


Table  1.     Soil  consistence  (U.S.  Soil  Conservation  Service  1981,  Chap.  4). 


Descriptive  Term  (by  moisture  condition) 

Connotation 

Air  Dry 

Field  Capacity  (moist) 

Loose 

Loose 

No  specimen  can  be  obtained.1 

Soft 

Very  friable 

Crushes  or  breaks  under  slight  force  ap- 
plied by  thumb  and  forefinger.2 

Slightly  hard 

Friable 

Crushes  or  breaks  under  moderate  force  by 
thumb  and  forefinger. 

Slightly  hard 

Firm 

Crushes  or  breaks  under  moderate  force  by 
thumb  and  forefinger. 

Hard 

Very  firm 

Crushes  or  breaks  under  strong  force  by 
thumb  and  forefinger. 

Very  hard 

Extremely  firm 

Cannot  be  crushed  or  broken  by  thumb 
and  forefinger;  can  be  broken  by  squeezing 
slowly  between  hands. 

Extremely  hard 

Extremely  hard 

Cannot  be  crushed  or  broken  in  hands;  can 
be  crushed  or  broken  underfoot 

1A  specimen  for  testing  strength  is  normally  a  cube  approximately  1  cm  on  a  side 

2Under  laboratory  conditions,  strength  classes  are  defined  quantitatively  by  measured  force  in  newtons 


Soils 


571 


MUNSELL-    SOIL  COLOR  CHART 


5YR 


3/ 


2.5/ 


■  ■■■ 


/I  /  /3  A  /6  /8 

CHROMA >■ 


Figure  6.     Soil  color. 

Several  researchers  have  found  a  distinct  rela- 
tionship between  soil  color  and  the  pelage  colora- 
tion of  various  mammals  (Dice  and  Blossom  1937; 
Hardy  1945).  For  woodrats  (Neotoma  sp. )  and  chip- 
munks (Eutamias  sp.)  in  the  South  Dakota  badlands, 
the  color  relationship  was  strongest  on  sites  largely 
devoid  of  vegetative  cover  (Stebler  1939).  Soil  color 
can  also  exert  a  strong  influence  over  surface 
temperatures. 

Depth 

Soil  depth  is  measured  from  the  surface  to  a 
restricting  or  contrasting  layer.  The  nature  of  the 
layer  in  question  is  generally  specified  (e.g.,  depth  to 
hardpan ).  Where  the  nature  of  the  layer  is  not  given, 
it  is  understood  to  be  consolidated  bedrock.  Stand- 
ard classes  for  depth  are  given  in  Figure  7. 

Soil  depth  and  soil  texture  influence  the  size 
and  distribution  of  pocket  gophers  (Thomomys  sp., 
Geomys  sp.)  in  the  western  U.S.  (Davis  1938;  Davis 
et  al.  1938).  Hardy  (1945)  contrasted  the  soil  depth 


Chipmunk. 


Depth  Class 


Very 
shallow 


Shallow 


Moderately 
deep 


Deep 


Very 
ik    deeP 


Centimeters         Inches 


<  25 


25-50 


50-100 


100-150 


>  150 


<  10 


10-20 


20-40 


40-60 


>  60 


Figure  7.     Soil  depth. 

requirements  of  two  species  of  kangaroo  rats  (Dipo- 
domys  sp.).  Additionally,  soil  depth  can  affect  ani- 
mals indirectly  through  its  influence  on  the  kind  and 
amount  of  vegetation  ( McColley  and  Hodgkinson 
1970;  Passey  et  al.  1982). 


572 


Soils 


, 


Table  2.     Root  quantity  and  size  classes  (U.S.  Soil 
Conservative  Service  1981,  Chap.  4). 


Ord's  kangaroo  rat. 


Roots  and  Animal  Traces 

The  quantity  and  size  of  plant  roots  in  each  soil 
layer  are  routinely  recorded  in  soil  descriptions  (Ta- 
ble 2  ).  Evidences  of  animal  activity  are  also  de- 
scribed when  apparent,  although  no  specific 
descriptive  standards  have  been  developed.  Such 
evidences  range  from  surface  features  (e.g.,  ant 
mounds,  burrow  openings)  to  actual  structures 
within  the  soil  (e.g.,  tunnels,  krotovinas ).  Krotovinas 
are  irregular,  tubular  streaks  or  spots  in  the  soil 
caused  when  abandoned  animal  tunnels  become 
filled  with  contrasting  materials. 

The  presence  or  absence  of  roots  and  animal 
traces  is  useful  for  assessing  habitat  suitability  for 
plants  and  animals.  Root  abundance  can  be  used  to 
make  inferences  about  soil  moisture  conditions  as 
well  as  aeration  and  temperature  (Weaver  1977; 
Lunt  et  al.  1973;  Daubenmire  1972).  The  abrupt  ter- 
mination of  roots  at  a  particular  level  generally  de- 
fines a  physical  or  chemical  barrier.  That  which 
restricts  root  growth  will  likely  inhibit  many  burrow- 
ing animals  and  other  biological  activity. 


Root  Quantity 

Descriptive  Term 

<  1  per  unit  area1 
1-5  per  unit  area 
>  5  per  unit  area 

Few 

Common 

Many 

Very  Fine 
Fine 
Medium 
Coarse 

Root  Size 

(diameter) 
<  1  mm 
1-2  mm 
2-5  mm 
>  5  mm 

Unit  area  for  very  fine  and  fine  roots  is  1  square  centimeter; 
unit  area  for  medium  and  coarse  roots  is  1  square  decimeter 


Drainage 

The  times  and  depths  at  which  a  soil  is  wet 
exerts  a  strong  influence  on  biological  activities  as 
well  as  chemical  processes.  Soil  drainage  classes 
(Table  3)  serve  to  summarize  soil  wetness.  As  gener- 
ally defined,  however,  drainage  classes  reflect  an 
agricultural  and  a  geographic  (humid-temperate) 
bias.  Local  refinements  are  often  necessary  to  adapt 
the  drainage  class  concept  to  other  climates  and  land 
uses.  Drainage  is  a  product  of  climate,  slope,  and 
landscape  position  along  with  soil  characteristics. 


Reaction 

Soil  reaction  is  expressed  as  pH.  The  pH  influ- 
ences the  presence  of  toxic  ions  in  the  soil  as  well 
as  nutrient  availability  to  plants.  Plants  growing  in 
soils  with  a  pH  less  than  5.2  are  likely  to  be  affected 
by  aluminum  toxicity.  Soils  with  a  pH  greater  than 
8.5  often  have  excessive  amounts  of  sodium  which 
can  destroy  soil  structure  and  thereby  reduce  soil 
moisture  and  aeration.  In  addition,  sodium  can  be 
toxic  to  some  plants. 


Table  3.     Soil  drainage  classes  (adapted  from  U.S.  Soil  Conservation  Service  1981,  Chap.  4). 


Descriptive  Term 

General  Connotation 

Excessively  drained,  somewhat  excessively 
drained 

These  soils  have  high  to  very  high  rates  of  water  transmission 
and  retain  little  moisture  for  plant  use. 

Well-drained 

These  soils  have  intermediate  water  transmission  rates  and  reten- 
tion capacities. 

Moderately  well-drained,  somewhat  poorly 
drained 

These  soils  are  commonly  wet  near  the  surface  for  at  least  a 
portion  of  the  year. 

Poorly  drained,  very  poorly  drained 

These  soils  are  commonly  wet  at  or  near  the  surface  for  a  con- 
siderable part  of  the  year. 

Soils 


573 


Figure  8.     Soil  reaction  and  nutrient  availability. 

Several  essential  nutrients  become  less  available 
to  plants  as  the  pH  decreases  below  about  6  or  in- 
creases above  about  7  (Figure  8).  This  effect  can 
indirectly  affect  wildlife  through  the  forage  they 
consume.  For  example,  many  arid  soils  of  the  West 
have  pH  values  between  7  and  8.4.  Within  this 
range,  phosphorus  availability  is  reduced  through  the 
formation  of  insoluble  calcium  phosphates.  Short 
( 1979)  suggested  that  declines  in  herds  of  mule  deer 
(Odocoileus  hemionns)  in  the  southwestern  U.S. 
might  be  caused  in  part  by  the  low  available  phos- 
phorus levels  in  the  soils. 

Soil  pH  and  associated  potassium  levels  partially 
account  for  variation  in  peak  densities  in  Microtus 
pennsylvanicus  (Krebs  et  al.  1971).  The  distribution 
of  several  species  of  salamanders  has  also  been  re- 
lated to  soil  pH  (Mushinsky  and  Brodie,  Jr.  1975; 
Batson  1965). 

Other  Soil  Characteristics 

Additional  soil  properties  and  characteristics, 
although  not  specifically  stated  in  a  standard  soil 
description,  are  nevertheless  commonly  determined 
during  the  course  of  a  soil  survey.  Many  of  these 
characteristics  are  not  directly  observed  on  a  soil 
profile,  but  can  be  inferred  from  other  properties  or 
measured  through  laboratory  analyses  of  soil 
samples. 


Available  Water  Capacity.  Within  the  plant-rooting 
zone,  the  amount  of  water  that  can  be  held  between 
1/3  bar  and  15  bars  of  tension  (1  bar  =  0.99  atm) 
is  the  water  retention  difference.  The  amount  of 
moisture  at  1/3  bar,  termed  the  field  capacity, 
approximates  the  moisture  that  would  remain  in  the 
soil  after  natural  drainage  occurred  in  response  to 
gravity  alone.  At  this  point,  the  larger  soil  pores 
would  be  empty  while  the  smaller  pores  would  be 
filled  with  water.  Soil  water  held  at  tensions  of  more 
than  15  bars  (wilting  point)  is  unavailable  to  many 
plants;  hence,  the  moisture  content  between  1/3  and 
15  bars  is  commonly  referred  to  as  the  available 
water  capacity  (AWC).  Some  plants,  however,  most 
notably  greasewood,  saltbush,  shadscale,  and  other 
salt  desert  species,  can  extract  soil  moisture  at 
tensions  far  exceeding  1 5  bars. 

The  available  water  capacity  varies  with  soil 
texture,  structure,  rock  fragment  content,  and  other 
soil  properties.  In  general,  however,  fine-textured 
soils  such  as  clays,  clay  loams,  or  silt  loams  have  a 
high  AWC  whereas  coarse-textured  soils  such  as 
sands  or  sandy  loams  have  a  low  AWC.  The  AWC 
does  not  reflect  actual  soil  moisture  content,  but 
rather  the  potential  to  retain  moisture  within  the  de- 
fined tension  limits.  Thus,  a  sandy  loam  in  New  Mex- 
ico and  New  York  might  well  have  similar  available 
water  capacities,  yet  contain  vastly  different  amounts 
of  moisture  at  any  given  time  due  to  different  pre- 
cipitation regimes.  Nevertheless,  the  AWC  can  be 
useful  in  evaluating  wildlife  habitat  on  a  local  basis 
(Figure  9). 


So/7  Series 
LEAPS 

PARLIN 
DUFFSON 

SPRING 
CREEK 

FOLA 

AVAILAE 

0-10" 
10-48" 

silty  clay  loam 
silty  clay 

0-11" 

11-31" 
31-48" 

channery  loam 
channery  clay  loam 
very  stony  loam 

0-8" 

8-16" 

16-30" 

30" 

loam 

clay  loam 
loam 
bedrock 

0-9" 

9-19" 

19" 

St 

b€ 

Dny  loar 
avelly  Ic 
sdrock 

n 

•am 

0-6" 
6-48" 
I       I 

cc 

ve 

I 

bbly  sandy  loam 
ry  cobbly  sandy  loam 
I       1      I       I       I       I 

012345678 

3LE  WATER  CAPACITY  (inches) 

Figure  9. 


The  potential  of  selected  soils  to  retain 
moisture. 


574 


Soils 


Seasonal  Water  Table.  A  seasonal  water  table  is 
marked  by  the  upper  limit  of  the  soil  that  is 
saturated  with  water  for  a  specified  period  during 
most  years.  For  example,  a  particular  soil  may  be 
characterized  as  having  a  seasonally  high  water  table 
at  a  depth  of  2  feet,  commonly  during  May  and  June. 
The  presence  or  absence  of  a  seasonally  high  water 
table  can  alter  the  suitability  of  a  site  for  burrowing 
animals.  It  can  also  affect  the  vertical  distribution 
within  the  soil  of  certain  animals  such  as 
salamanders  (Taub  1961). 

Hydrologic  Soil  Group.  The  hydrologic  soil  group 
provides  an  interpretation  of  soil  behavior  in  terms 
of  water  infiltration,  transmission,  and  runoff. 
Infiltration  is  the  rate  at  which  water  enters  the  soil 
surface,  transmission  is  the  rate  of  water  movement 
within  the  soil,  and  runoff  is  the  amount  of  water 
that  moves  over  the  surface  to  depressions  or 
defined  channels  without  entering  the  soil.  For 
standardized  conditions  of  bare  soil  after  prolonged 
wetting,  individual  soils  are  assigned  to  one  of  four 
hydrologic  soil  groups,  A,  B,  C,  or  D  (Figure  10). 


be  suitable  for  the  Merriam  kangaroo  rat  (Dipodo- 
mys  merriatni).  He  speculated  that  this  was  due  to  a 
salt-derived  surface  crust. 

Sodium  Adsorption  Ratio.  The  sodium  adsorption 
ratio  (SAR)  relates  the  amount  of  sodium  in  the  soil 
to  the  amount  of  calcium  and  magnesium.  As  noted 
earlier,  high  sodium  can  be  toxic  to  plants  as  well  as 
harmful  to  the  physical  condition  of  the  soil.  The 
SAR  is  particularly  useful  in  assessing  reclamation 
potential  of  mined  lands  or  other  severely  disturbed 
areas.  Most  State  and  Federal  reclamation  guidelines 
for  topsoil  and  overburden  suitability  include  the 
SAR  as  an  evaluation  factor.  In  general,  SAR  values  of 
12  or  more  suggest  unsuitability  or  at  least  severe 
limitations  for  successful  reclamation  (Fisher  and 
Deutsch  1983). 

Erodibility.  Soil  erosion  is  influenced  by  several 
non-soil  factors  such  as  slope,  precipitation,  wind 
velocity,  and  vegetative  cover.  Nevertheless,  some 
soils  erode  more  readily  than  others  even  when  all 
other  factors  are  the  same. 


Group 

Water 
Infiltration 

Water 
Transmission 

Water 
Runoff 

A 

(HIGH) 

(HIGH) 

(LOW) 

B 

C 

D 

\ 
(LC 

)W) 

(LC 

)W) 

(Hl( 

3H) 

Figure  10.     Hydrologic  soil  groups. 


Soils  in  Group  A  with  low  runoff  potential  and 
correspondingly  high  infiltration  rates  consist  chiefly 
of  deep,  well  to  excessively  drained  sands  and  grav- 
els. In  contrast,  soils  in  Group  D,  having  high  runoff 
potential  and  low  infiltration,  are  often  clayey,  have  a 
permanent  high  water  table,  or  are  shallow  over 
impervious  or  nearly  impervious  material. 

Salinity.  Salinity  refers  to  the  amount  of  salts  in  the 
soil,  the  standard  measure  of  which  is  the  electrical 
conductivity  (EC)  of  an  extract  of  the  soil  solution. 
In  many  situations,  however,  salinity  is  inferred  and 
expressed  as  salinity  classes  covering  a  range  of  EC 
values. 

Soil  salinity  exerts  a  strong  influence  over  vege- 
tation; as  might  be  expected,  the  higher  the  salinity, 
the  greater  the  effect  (Table  4).  Salts  in  the  soil  can 
also  affect  certain  animals  directly.  Hardy  (1945) 
noted  the  lack  of  digging  activity  on  sites  thought  to 


The  soil  erodibility  factor  (K)  reflects  this  inher- 
ent susceptibility  of  a  soil  to  erode  under  the  action 
of  raindrop  impact  and  water  flowing  over  the  sur- 
face (sheet  and  rill  erosion).  The  wind  erodibility 
group  ( WEG)  reflects  the  susceptibility  of  a  soil  to 
erode  under  wind  action.  Although  inherent  soil 
erodibility  varies  with  several  different  soil  proper- 
ties, a  rough  relationship  exists  between  soil  texture 
and  both  the  K  factor  and  the  WEG  (Figure  11). 

Erosion  presents  dual  problems  in  habitat  man- 
agement. First,  soil  loss  can  result  in  reduced  forage 
production  because  of  the  removal  of  fertile  topsoil. 
Second,  soil  lost  from  one  site  is  eventually  depos- 
ited elsewhere.  During  transport,  suspended  soil 
particles  and  attached  ions  can  degrade  water  qual- 
ity. Upon  deposition,  eroded  soil  material  can  further 
affect  aquatic  habitats.  For  example,  the  distribution 
of  two  species  of  frogs  (Rana  sp. )  varied  with  soil 
texture,  apparently  due  to  different  erodibilities  and 
associated  sediment  burdens  in  adjacent  streams 
(Lynch  1978). 

Soil  Fertility.  Soil  fertility  is  the  inherent  capability 
of  a  soil  to  supply  nutrients  to  plants  in  adequate 
amounts  and  in  suitable  proportions  (Buckman  and 
Brady  1969).  As  discussed  earlier,  soil  pH  serves 
as  an  indicator  of  nutrient  availability.  Soil  fertility  is 
more  specifically  characterized  through  laboratory 
analyses  of  extractable  bases  (e.g.,  Ca,  Mg,  Na,  K), 
phosphorus,  nitrogen,  and  determinations  of  cation 
exchange  capacity,  and  percentage-base  saturation 
(see  Sources  of  Soil  Information  section). 

Soil  fertility  has  long  been  known  to  influence 
the  palatability  of  plants.  For  example,  the  volatile  oil 


Soils 


575 


Table  4.     Soil  salinity  classes  (U.S.  Soil  Conservation  Service  1983,  Part  603). 


Electrical 

Conductivity 

(millimhos/cm) 

Salinity  Class 

Connotation1 

0-2 

Nonsaline 

No  effect  on  plants. 

2-4 

Very  slightly  saline 

Effect  on  plants  is  mostly  negligible;  only  most  sensitive 
plants  affected  adversely. 

4-8 

Slightly  saline 

Many  plants  are  adversely  affected  but  diversity  of  adapted 
species  is  guite  high  (e.g.,  four-wing  saltbush,  winterfat, 
galleta). 

8-16 

Moderately  saline 

Plant  communities  are  dominated  by  salt-tolerant  species 
(e.g.,  greasewood,  mat  saltbush,  western  wheatgrass). 

>  16 

Strongly  saline 

Few  plants  are  adapted  (e.g.,  shadscale,  iodine  bush,  alkali 
sacaton). 

1U.S.  Salinity  Laboratory  Staff  1954,  McArthur  et  al.  1978;  Thornburg  1982 


content  of  sagebrush  leaves  seems  to  vary  with  soil 
depth  and  fertility  among  other  non-soil  factors 
(Powell  1970;  Nagy  1966).  This  is  of  particular  im- 
portance because  high  volatile  oil  content  inhibits 
the  activity  of  rumen  microorganisms  (Nagy  et  al. 
1964).  Some  of  the  more  dramatic  palatability  effects 
are  observed  when  commercial  fertilizers  are  added 
to  nutrient  deficient  soils.  Gessel  and  Orians  (1967) 
observed  that  certain  rodents  selectively  fed  on  ni- 
trogen-fertilized trees  while  avoiding  adjacent  unfer- 
tilized trees  or  those  fertilized  with  potassium. 


In  a  series  of  studies  spanning  several  years  in 
Missouri,  the  abundance,  distribution,  health,  and 
size  of  several  animals  including  wild  turkeys,  rab- 
bits, raccoons,  muskrats,  opossums,  and  squirrels 
were  linked  to  soil  fertility  (Denny  1944;  Crawford 
1950;  Albrecht  1957).  Low  soil  fertility,  particularly 
levels  of  nitrogen  and  phosphorus,  has  previously 
been  noted  as  a  possible  cause  of  the  decline  in 
herds  of  mule  deer  in  the  Southwest  (Short  1979). 


Relationships  between  animals  and  soil  fertility 
are  not  always  predictable.  For  example,  a  pocket 
gopher  (Thomomys  bottae)  has  been  found  in  abun- 
dance on  serpentine-derived  soils  in  contrast  to  adja- 
cent non-serpentine  soils  (Proctor  and  Whitten 
1971 ).  As  noted  previously,  serpentine-derived  soils 
are  generally  low  or  lacking  in  several  essential  plant 
nutrients.  This  apparent  anomaly  was  explained  by 
the  presence  of  Brodiaca  sp.,  a  serpentine-tolerant 
plant  whose  corm  serves  as  a  primary  food  source 
for  the  gophers. 


SOIL  CLASSIFICATION 

Soil  taxonomy,  although  foreign  to  most  biolo- 
gists, is  as  much  a  scientific  necessity  as  the  taxon- 
omic  systems  used  to  classify  plants  and  animals.  The 
soil  taxonomic  system  used  in  the  U.S.  groups  soils 
according  to  measurable  characteristics  such  as 
depth,  texture,  temperature  and  moisture  regimes, 
and  chemical  properties.  Like  botanical  and  zoologi- 
cal systems,  soil  taxonomy  provides  a  uniform  means 
to  identify  a  particular  soil  and  facilitates  the  transfer 
of  both  soils  information  and  that  which  can  be  re- 
lated to  soils. 

The  system  is  hierarchical,  consisting  of  six  lev- 
els or  categories  including  order,  suborder,  great 
group,  subgroup,  family,  and  series.  At  the  highest  or 
most  general  level  there  are  10  soil  orders.  At  the 
lowest  and  most  specific  level,  over  12,000  series 
have  been  identified  in  the  U.S.  alone,  and  the  num- 
ber is  increasing  annually.  Table  5  shows  the  classifi- 
cation at  all  levels  for  one  particular  soil.  The 
nomenclature  applied  conveys  progressively  more 
information  from  the  order  through  the  family  level. 
Each  term  is  specifically  defined  in  a  handbook  out- 
lining soil  taxonomy  (Soil  Survey  Staff  1975).  At 
the  lowest  level,  the  series  name  itself  has  no  spe- 
cific technical  meaning,  often  being  derived  from  a 
place  or  local  landmark  near  where  the  soil  was  first 
described.  A  series  description,  however,  refines 
family  attributes  by  further  narrowing  the  allowable 
range  of  individual  soil  properties  ( Figure  1 ). 

Soil  taxonomy  is  useful  in  making  habitat  gener- 
alizations. For  example,  Kantrud  and  Kologiski 


576 


Soils 


.10 


.17 


.20 


K  FACTOR 

.24  .32 


.37 


.43 


.64 


WATER 

ERODIBIUTY 


=> 


Sand 

Loamy 

Clay 

Sandy  Loam 

Sandy  Clay, 

Loam, 

Silt  Loam 

Silt 

Sand 

Silty  Clay 

Clay  Loam, 
Sandy  Clay 
Loam 

Silty  Clay 
Loam 

LOW 


■►HIGH 


WIND 

ERODIBIUTY 


=> 


Wet  Soils, 

Silt 

Clay  Loam 

Loam, 

Clay, 

Sandy 

Loamy 

Sand 

Rocky 

Silt  Loam 

Sandy  Clay, 

Silty  Clay, 

Loam 

Sand 

Soils 

Sandy  Clay 
Loam 

Silty  Clay 
Loam 

6  5  4  3 

WIND  ERODIBIUTY  GROUP 


Figure  11.     Soil  erodibility. 


Table  5.     Taxonomic  system  for  soil  classification  (Soil  Survey  Staff  1975). 


Taxonomic 
Category 


Number  in 
System' 


Example 


Order 


10 


Aridisol 


Suborder 


45 


Argid 


Great  Group 


187 


Haplargid 


Subgroup 


990 


Lithic        Ustollic        Haplargid 


Family 


5,603 


Loamy,        mixed,        mesic,        Lithic        Ustollic        Haplargid 


Series 


12,002 


<  35%  clay  in  zone  of  clay 
accumulation 


"iixea  mineral' 
mineral  other  than  quartz) 


vo  of  any  one 


mean  annual  soil  temp,  between  8°  and  15°  C 
(at  50  cm  or  bedrock,  whichever  shallower) 


depth  to  bedrock  is  <  50  cm  (20  in.) 


is  wetter  and  has  more  organic  matter  than  normal  for 
the  great  group 


has  no  cemented  pan  or  high  sodium  content 
has  a  horizon  of  silicate  clay  accumulation 


is  dry  in  plant-root  zone  for  at  least  half  of  growing  season 


as  of  October  1980 


Soils 


577 


(1982)  found  that  average  bird  density  and  species 
richness  were  slightly  higher  on  soils  of  the  order 
Mollisol  than  on  soils  of  the  order  Aridisol.  Within 
the  order  Mollisol,  significantly  different  species 
densities  were  found  at  the  soil  subgroup  level.  In 
another  study,  Leopold  and  Dalke  (1941)  found  that 
a  single  soil  series  supported  over  90%  of  the  wild 
turkey  population  in  Missouri. 

SOIL  MAPPING 

Soil  surveys  are  designed  and  conducted  at  dif- 
ferent intensities  to  meet  different  needs.  Accord- 
ingly, not  all  surveys  provide  information  at  the  same 
level  of  detail  nor  of  equal  applicability  in  habitat 
management.  Failure  to  recognize  this  fact  can  and 
has  led  to  the  misuse  of  soil  survey  information. 
Potential  users  of  soil  information  should  be  in- 
volved in  the  design  of  soil  surveys.  Where  soil  sur- 
veys are  already  completed,  users  of  the  data  should 
be  sufficiently  knowledgeable  to  recognize  both  the 
strengths  and  the  limitations  of  the  soil  map. 

Mapping  Variables 

Four  main  variables,  map  unit  components, 
kinds  of  map  units,  field  procedures,  and  map  scales, 
are  used  in  designing  soil  surveys.  The  same  varia- 
bles are  useful  in  assessing  the  applicability  of  exist- 
ing soil  survey  information. 

Map  unit  components  are  bodies  of  soil  that  can 
be  identified  as  being  a  member  of  some  class 
(taxon)  in  soil  taxonomy  or  bodies  of  nonsoil  that 
can  be  identified  as  any  of  22  types  of  miscellaneous 
areas  (e.g.,  rock  outcrop,  badland,  riverwash;  U.S.  Soil 
Conservation  Service  1981).  In  practice,  soil  taxons 
and  miscellaneous  areas  are  often  further  refined 
by  one  or  more  phase  modifiers.  Phases  reflect  var- 
ious use-related  soil  and  nonsoil  attributes  such  as 
surface  texture  and  slope  (see  examples  in  Figure 
12). 

The  kind  of  map  unit  depends  on  the  number 
and  arrangement  of  the  individual  components  on 
the  landscape  (Figure  12).  Single-component  map 
units,  termed  consociations,  are  dominated  by  one 
soil  taxon  or  miscellaneous  area.  Multicomponent 
map  units,  termed  either  associations  or  complexes, 
contain  two  or  more  dominant  taxons  or  miscella- 
neous areas.  In  an  association,  the  named  compo- 
nents generally  lie  in  a  regular  geographic  pattern 
and  can  be  located  in  the  field  by  visible  landscape 
features  (e.g.,  hill  slopes  vs.  valley  bottoms,  north 
slopes  vs.  south  slopes).  In  contrast,  the  named  com- 
ponents in  a  complex  either  lie  in  a  very  intricate 
pattern  that  cannot  be  determined  by  visible  land- 
scape features  or  are  simply  too  small  to  be  sepa- 
rated. The  choice  of  map  unit  depends  on  the  soil 
itself  as  well  as  the  purpose  of  the  survey  and  the 


level  of  detail  needed.  Regardless  of  the  map  unit, 
however,  there  will  always  be  some  areas  of  soil  or 
nonsoil  material  that  do  not  conform  to  the  named 
components.  Such  areas,  termed  inclusions,  occur 
because  of  map  scale  limitations  or  errors  in  the 
placement  of  map  unit  boundaries.  As  a  general  rule, 
inclusions  that  would  differ  significantly  in  use  and 
management  from  the  named  components  should 
not  exceed  15%  of  the  map  unit. 

The  initial  step  in  a  soil  survey  is  often  to  de- 
velop a  preliminary  soil  map  based  on  ancillary  data 
(e.g.,  geology,  landform,  adjacent  soil  maps)  and  a 
knowledge  of  soil  formation  and  its  relationship  to 
other  landscape  features.  These  preliminary  map 
units  are  then  field  checked  and  revised  as  necessary. 
The  specific  procedures  used  in  the  field  work  deter- 
mine the  precision  with  which  map  unit  components 
are  identified  and  described,  and  the  accuracy  of 
map  unit  boundaries.  Common  field  procedures  in- 
clude explicit  soil  sampling  schemes,  observations  of 
exposed  soils  (e.g.,  roadcut),  landscape  features  on 
the  ground,  and  interpretations  of  aerial  photographs 
or  other  remotely  sensed  images. 

Map  scale  determines  the  smallest  sized  unit 
that  can  be  practically  delineated  on  a  map.  Gener- 
ally this  is  a  1/4  x  1/4  in.  square  or  a  circular  area  of 
about  1/16  square  inch.  By  using  these  guidelines, 
the  smallest  unit  that  can  be  delineated  at  a  scale  of 
1:24,000  (7  1/2'min  quadrangle)  is  about  6  acres 
(Table  6).  Smaller  areas,  if  important,  can  be  shown 
by  spot  symbols.  Such  a  practice  should  be  used 
to  indicate  special  habitat  features  such  as  small 
ponds,  or  wet  depressions,  springs,  and  escarpments. 


Table  6.     Guide  to  map  scales  and  corresponding 
minimum  size  delineations  (U.S.  Soil 
Conservation  Service  1981,  Chap.  2). 


Minimum 

Delineation 

Scale 

Inches/Mile 

(acres) 

1 

7,920 

8.0 

0.6 

1 

15,840 

4.0 

2.5 

1 

24,000 

2.6 

6.0 

1 

31,680 

2.0 

10.0 

1 

63,360 

1.0 

40.0 

1 

126,720 

0.5 

160.0 

Mapping  Intensity 

Based  on  the  four  variables  above,  soil  surveys 
can  be  grouped  by  levels  of  intensity  and  general 
application.  Five  such  levels  are  commonly  recog- 
nized ranging  from  Order  1  surveys,  the  most  inten- 


578 


Soils 


sive,  to  Order  5  surveys,  the  least  intensive.  Any 
given  survey,  however,  need  not  be  of  a  single  inten- 
sity. In  fact,  a  mixture  of  intensities  will  often  best 
meet  user  needs  most  economically.  Soil  surveys  on 
BLM-administered  lands  often  contain  elements  of 
Order  2,  3,  and  4  mapping  intensities.  Figure  13 
shows  how  a  single  area  might  differ  when  mapped 
at  each  of  these  intensity  levels. 


SOURCES  OF  SOIL  INFORMATION 

Perhaps  the  most  obvious  source  of  soils  infor- 
mation is  a  soil  scientist.  Currently,  the  Bureau  of 
Land  Management  employs  about  75  soil  scientists 
distributed  among  its  various  field  offices,  the  BLM 
Service  Center,  and  the  Headquarters  Office  in 
Washington,  DC. 


map  units 


single-component 


multi-component 


components 
separable 


components 
not  separable 


CONSOCIATION 


ASSOCIATION 


COMPLEX 


CONSOCIATIONS 

Avalon  loam,  0  to  3  percent  slopes 

(series  taxon  with  texture  and  slope  phase  modifiers) 
Torrifluvents,  frequently  flooded 

(great  group  taxon  with  flooding  phase  modifier) 
Gullied  land 

(miscellaneous  area  with  no  phase  modifier) 

ASSOCIATIONS 

Progresso — Potts  association,  1  to  6%  slopes 

(series  taxons  with  slope  phase  modifier) 
Typic  Haplargrids — Typic  Camborthids  association 

(subgroup  taxons  with  no  phase  modifier) 

COMPLEXES 

Lazear — Rock  outcrop  complex,  3  to  30%  slopes 

(series  taxon  and  miscellaneous  area  with  slope  phase  modifier) 
Fughes — Curecanti  loams,  10  to  40%  slopes 

(series  taxons  with  texture  and  slope  phase  modifiers;  where  textural 

phase  is  same  among  components,  plural  texture  phase  is  used  without 

the  term  complex) 
Orthents — Rock  outcrop  complex 

(suborder  taxon  and  miscellaneous  area  with  no  phase  modifier) 


Figure  12.     Kinds  of  soil-map  units. 


Soils 


579 


1:24,000 


ORDER  2 

|  Rock  outcrop 

2  Alpha  gravelly  loam 
]  Beta  gravelly  loam 


ORDER  3 

\^  Alpha-rock  outcrop 
association 


1 :63,360 


~]  Beta  gravelly  loam 


ORDER  4 
XI 


b^ 


Alpha-beta  rock 
outcrop  association 


1:126,720 


Figure  13.     Soil  mapping  intensity. 

Published  soils  information  is  concentrated  in 
soil  survey  reports.  Such  reports  are  generally  pub- 
lished by  the  U.S.  Soil  Conservation  Service  (SCS)  as 
part  of  a  formal  series,  often  in  cooperation  with 
State  Agricultural  Experiment  Stations  and  other 
State  and  Federal  agencies.  These  reports  have  a 
standardized  format,  the  primary  elements  of  which 
are  listed  in  Table  7.  A  nationwide  list  of  published 
soil  surveys  is  prepared  and  periodically  updated  by 
the  SCS  (U.S.  Soil  Conservation  Service  1982). 

Interim  reports,  of  variable  content  and  formal- 
ity, are  often  prepared  before  formal  publication. 
This  is  true  on  much  of  the  public  land  administered 
by  the  BLM  where  field  mapping  is  still  in  progress 
or  only  recently  completed.  These  reports  are  avail- 
able in  the  respective  District  or  Resource  Area  Of- 
fices. Special  purpose  surveys,  often  of  limited  areal 
extent  (e.g.,  mining  and  mineral  exploration  tracts, 
research  areas),  may  also  be  available  through  pri- 
vate companies,  universities,  or  other  Federal  agen- 
cies. 

Soil  laboratory  data  in  much  greater  detail  than 
that  commonly  found  in  soil  survey  reports  are  pub- 
lished as  Soil  Survey  Investigations  Reports  (SSIR). 
They  have  been  issued  for  many  individual  States 
and  groups  of  States  ( U.S.  Soil  Conservation  Service 
1972).  Similar  information  is  also  published  by  State 
Agricultural  Experiment  Stations  in  various  forms 


(e.g.,  Science  Monographs,  Miscellaneous  Reports, 
Special  Resource  Series). 


APPLICATION  OF  SOIL  INFORMATION 

A  knowledge  of  soils  and  their  variability  across 
the  landscape  can  aid  in  delineating  habitat  units, 
in  assessing  habitat  suitability,  and  in  developing  and 
evaluating  habitat  management  alternatives. 

Habitat  Delineation 

An  initial  objective  in  wildlife  inventory  is  to 
delineate  land  units  with  the  same  environmental  at- 
tributes. The  "habitat  site"  constitutes  this  unit  in 
the  BLM's  Integrated  Habitat  Inventory  and  Classifi- 
cation System  (IHICS;  U.S.  Bureau  of  Land  Manage- 
ment 1982).  Once  delineated,  similar  although  not 
necessarily  adjacent  habitat  sites  are  aggregated  into 
so-called  standard  habitat  sites  (SHS).  Data  on  animal 
occurrence  from  a  few  representative  habitat  sites 
are  then  extrapolated  to  characterize  an  entire  SHS. 
The  reliability  of  this  procedure  rests  on  the  assump- 
tion that  any  given  habitat  site  is  uniform  in  terms 
of  those  site  factors  that  influence  animal 
occurrence. 

By  definition,  a  habitat  site  is  an  area  of  homoge- 
neous (present)  vegetation,  landform,  soils,  and  cli- 
mate (U.S.  Bureau  of  Land  Management  1982).  In 
practice,  however,  habitat  sites  and  standard  habitat 
sites  are  often  delineated  almost  exclusively  on  the 
present  vegetation.  This  approach  can  result  in  non- 
homogeneous  sites  which  in  turn  can  lead  to  erro- 
neous conclusions  on  animal  occurrence  and  distri- 
bution. For  example,  overgrazing  or  other  land-use 
practices  can  produce  relatively  uniform  vegetation 
over  broad  areas.  With  vegetation  as  the  main  habitat 
criterion,  the  illusion  of  habitat  uniformity  exists 
even  though  the  underlying  soils  may  be  variable. 
For  animals  sensitive  to  soil  variability,  subsequent 
habitat  characterizations  are  incomplete  at  best. 

In  most  instances,  neither  soils  nor  vegetation 
( nor  any  other  single  factor )  can  adequately  describe 
habitat.  When  all  relevant  factors  are  considered 
together,  however,  they  are  complementary  and  pro- 
vide an  added  dimension  to  the  understanding  of 
the  habitat  unit. 


Habitat  Suitability 

The  suitability  of  a  standard  habitat  site  for  a 
particular  wildlife  species  depends  on  whether  exist- 
ing habitat  features  meet  the  needs  of  the  animal. 
When  soil  information  is  used  in  the  delineation  and 
aggregation  of  habitat  sites,  suitability  assessments 


580 


Soils 


Table  7.     Major  sections  of  a  soil  survey  report  (U.S.  Soil  Conservation  Service  1983,  Part  605). 


Section 

General  Content 

General  nature  of  the  survey  area 

Commonly  contains  a  discussion  of  the  history,  climate,  physiography, 
drainage,  and  natural  resources  of  the  survey  area 

General  soil  map  and  map  unit 
description 

Includes  a  general  soil  map  (common  scale  of  1:250,000  or  smaller)  and 
associated  map  unit  descriptions;  narrative  descriptions  commonly  give  a 
brief  characterization  of  the  included  soils,  their  relative  proportions,  and 
major  inclusions,  along  with  information  on  elevation,  landform,  topogra- 
phy, average  precipitation  and  air  temperature,  plant  communities,  and 
dominant  use  and  management  considerations. 

Use  and  management 

Provides  interpretive  information,  mainly  in  tabular  form,  on  soil  productive 
capability,  potential  plant  communities,  and  suitability  and/or  limitations 
for  major  land  uses  and  practices. 

Soil  properties 

Commonly  includes  tabular  data  on  soil  texture,  rock  content,  water  reten- 
tion difference,  pH,  salinity,  wind  and  water  erosion  factors,  hydrologic 
soil  group,  and  other  features. 

Classification  of  soils 

Gives  the  taxonomic  classification  for  each  soil. 

Soil  series  and  morphology 

Contains  the  detailed  descriptions  for  each  soil  series  (or  other  category  if 
applicable);  Figure  1  was  taken  from  this  section  of  a  soil  survey  report. 

can  be  strengthened.  To  do  so,  relationships  be- 
tween animals  and  soils  must  be  characterized.  More 
specifically,  the  soil-related  affinities  and  tolerances 
must  be  identified  for  individual  animal  species  or 
groups  of  species. 

Applicable  information  on  soil-animal  relation- 
ships does  exist,  although  it  is  widely  scattered  and 
poorly  developed  for  actual  application.  The  results 
of  a  limited  literature  review  are  shown  in  Table  8. 
The  task  of  accumulating  and  arraying  such  informa- 
tion in  a  usable  form  should  be  a  joint  effort  be- 
tween soil  scientists  and  wildlife  biologists.  The 
logical  first  step  would  be  an  exhaustive  review  of 
pertinent  literature.  The  next  step  would  be  to  de- 
velop a  data  format  to  enable  practical  application.  A 
sample  format  patterned  after  traditional  soil  inter- 
pretations is  shown  in  Table  9.  Once  identified,  rela- 
tionships between  soil  properties  and  animals  can 
be  used  to  interpret  soil  map  units  or  components  of 
map  units.  For  example,  the  Kech  loam,  3  to  15% 
slopes,  is  a  soil  map  unit  whose  primary  component 
is  the  Kech  soil  series.  If  the  Kech  series  exhibits 
properties  that  are  suited  to  a  particular  animal  spe- 
cies, all  areas  mapped  as  such  could  be  considered  at 
least  potentially  suitable  for  that  species.  In  conjunc- 
tion with  the  evaluation  of  other  factors  (e.g.,  vegeta- 
tion), such  analyses  can  be  useful  in  predicting 
animal  presence  or  absence,  determining  probable 
ranges,  and  selecting  optimum  sampling  sites  for 
supporting  studies. 

Even  when  relationships  between  soils  and  ani- 
mals are  unknown,  there  are  opportunities  to  make 
use  of  existing  soils  information.  For  example,  where 


estimates  of  wildlife  occurrence  have  been  estab- 
lished through  previous  sampling  or  other  studies, 
presumed  distribution  patterns  can  be  superimposed 
on  soil  maps.  The  objective  is  to  identify  any  appar- 
ent associations  between  wildlife  distribution  and 
soil  map  units.  At  this  level  of  analysis,  it  is  not  nec- 
essary to  know  the  exact  nature  of  the  relationship, 
but  simply  that  a  relationship  is  probable.  This 
knowledge  alone  can  aid  in  extrapolating  data  from 
sampled  to  unsampled  areas. 


For  certain  animal  species,  particularly  large 
ungulates  with  broad  ranges,  habitat  suitability  may 
be  closely  related  to  the  intermingling  of  various 
soils  and  associated  vegetation  producing  so-called 
edge  effects  or  ecotones.  Wertz  ( 1966)  provided  an 
example  of  such  an  analysis  from  Wisconsin.  Individ- 
ual soils  were  characterized  by  vegetation  type  and 
cover.  Natural  edge  conditions  were  identified  where 
contrasting  soils  occurred  together  on  the  landscape. 
Ratings  based  on  the  number  of  miles  of  soil-associ- 
ated natural  edge  were  then  used  to  assess  the  suita- 
bility of  different  management  areas  for  various 
wildlife  species. 


Searching  for  soil-animal  relationships  by  analyz- 
ing several  layers  of  data  is  a  tedious  job.  This,  in 
part,  accounts  for  the  general  lack  of  species-specific 
soil  interpretations  for  wildlife  habitat  management. 
With  the  increasing  availability  of  computerized  soils 
and  vegetational  data  bases  along  with  the  analytic 
capabilities  of  geographic  information  systems  (GIS), 
such  analyses  should  be  greatly  facilitated. 


Soils 


581 


Table  8.     Relationships  between  soils  and  wildlife. 


Soil  Property 

Wildlife  Characteristic 

Wildlife  Species 

Reference 

Texture 

Distribution 

Kangaroo  rats  (Dipodomys  sp.) 

Dale  1939 

Texture 

Distribution 

Pocket  mice  (Perognathus  sp.) 

Hardy  1945 

Texture 

Distribution 

Deer  mouse  {Peromyscus  manicula- 

tus) 
S.  grasshopper  mouse  {Unchomys 

Hardy  1945 

Texture 

Distribution 

Hardy  1945 

torridus) 

Texture 

Distribution 

Various  lizard  species 

Stuart  1932 

Texture 

Population  density 

G.  Basin  pocket  mouse  (Perogna- 
thus parvus) 

Feldhamer  1979 

Texture 

Population  density 

Least  chipmunk  (Tamias  minimus) 

Feldhamer  1979 

Texture 

Dust-bathing  behavior 

Bobwhite  quail  (Colinus  virginianus) 

Borchelt  and  Overmann 
1974 

Texture 

Denning  habits 

Red  fox  (Vulpes  vulpes) 

Sheldon  1950 

Texture 

Burrow  location 

Woodchuck  (Marmota  monax) 

Moss  1940 

Texture  &  erodibility 

Distribution 

Leopard  frogs  (Rana  sp.) 

Lynch  1978 

Texture  &  depth 

Distribution,  animal  size 

Pocket  gophers  {Thomomys  sp.) 

Davis  1938 

Texture  &  depth 

Distribution,  animal  size 

Pocket  gophers  (Geomys  sp.) 

Davis  et  al.  1938 

Depth 

Distribution 

Kangaroo  rats  (Dipodomys  sp.) 

Hardy  1945 

Depth 

Distribution 

W.  harvest  mouse  (Reithrodontomys 

Hardy  1945 

megalotis) 

Hardy  1945 

Depth  &  strength 

Population  density 

Least  chipmunk  (Tamius  minimus) 

Feldhamer  1979 

Strength 

Various  desert  rodents 

Rosenzweig  and  Winakur 
1969 

Drainage 

Burrowing  habits 

Black-tailed  prairie  dog  (Cynomys 
ludovicianus) 

Sheets  et  al.  1971 

Water  table  depth 

Vertical  distribution 

Red-backed  salamander  (Plethodon 
cinereus) 

Taub  1961 

Moisture 

Distribution 

Voles  (Microtus  sp.) 

Hodgson  1972 

Moisture 

Distribution 

Sagebrush  vole  (Lagurus  curtatus) 

O'Farrell  1972 

Moisture 

Distribution 

Ground  squirrels  (Spermophilus  sp.) 

Turner  1972 

Moisture 

Distribution 

Kangaroo  rats  (Dipodomys  sp.) 

Bienek  and  Grundmann 

1971 
Spight  1967 

Moisture 

Rehydration  (water 

Various  salamander  species 

balance) 

Color 

Pelage  coloration 

Various  small  mammals 

Dice  and  Blossom  1937 

Color 

Pelage  coloration 

Various  small  mammals 

Stebler  1939 

Reaction  (pH) 

Substrate  preference 

Various  salamander  species 

Mushinsky  &  Brodie,  Jr, 
1975 

Reaction  (pH) 

Population  density 

Meadow  vole  (Microtus  pennsylvan- 
icus) 

Krebs  et  al.  1971 

Nitrogen  level 

Feeding  habits 

Red-backed  vole  (Clethnonmys 

sp.) 
Mule  deer  (docoileus  hemionus) 

Gessel  and  Orians  1967 

Available  phosphorus 

Fawn  recruitment 

Short  1979 

Sodium  level 

Lick  preference 

Mountain  goat  (Oreamnos  amen- 

canus) 
Elk  (Cervus  elaphus) 

Stockstad  et  al.  1953 

Sodium  level 

Lick  preference 

Stockstad  et  al.  1953 

Sodium  level 

Lick  preference 

Mule  deer  (Odocoileus  hemionus) 

Stockstad  et  al.  1953 

Sodium  level 

Lick  preference 

White-tailed  deer  (Odocoileus  virgi- 

Stockstad et  al.  1953 

nianus) 

Stockstad  et  al.  1953 

Sodium  level 

Peak  population  density 

Microtine  rodents 

Aumann  1965 

Calcium  availability 

Feeding  habits 

Ring-necked  pheasant  (Phasianus 
colchicus) 

Harper  and  Labisky  1964 

Overall  fertility 

Distribution,  animal  weight 

Raccoon  (Procyon  lotor) 

Crawford  1950 

Overall  fertility 

Pelt  size  and  quality 

Opossum  (Didelphis  virginianus) 

Crawford  1950 

Overall  fertility 

Animal  weight,  bone 
strength 

E,  cottontail  (Sylvilagus  flondanus) 

Crawford  1950 

Overall  fertility 

Fecundity 

E.  cottontail  (Sylvilagus  flondanus) 

Williams  and  Caskey  1965 

Overall  fertility 

Litter  size 

E.  cottontail  (Sylvilagus  flondanus) 

Hill  1972 

Overall  fertility 

Breeding  potential 

White-tailed  deer  (Odocoileus  virgi- 
nianus) 

Crawford  1953 

Parent  material  (Ser- 

Distribution 

Botta's  pocket  gopher  (Thomomys 

Proctor  and  Whitten  1971 

pentine) 

bottaie) 

Soil  classification 

series 
Soil  classification 

Distribution 

Turkey  (Meleagns  gallopavo) 

Leopold  and  Dalke  1943 

Forage  preference 

Black-tailed  deer  (Odocoileus  Hem- 

Whitaker 1965 

series 

ionus  columbianus) 

Subgroup 

Species  richness 

Various  bird  species 

Kantrud  and  Kologiski 
1982 

582 


Soils 


Table  9.     Sample  format  for  soil  interpretations  of  habitat  suitability  for  a  particular  animal  species. 


Soil  Property  or  Soil- 
Related  Habitat  Feature 

HABITAT  SUITABILITY 

Optimum 

Marginal 

Unsuitable 

Texture 

Loamy  sand,  Sandy  loam, 
Sandy  clay  loam,  Loam 

Silt  loam,  Silt,  Silty  clay 
loam,  Clay  loam 

Clay,  Silty  clay,  Sandy  clay, 
Sand 

Depth  (inches) 

>40 

2CM0 

<20 

Drainage 

Well-drained 

Moderately  well-drained, 
Somewhat  poorly  drained 

Poorly  drained,  Very  poorly 
drained,  Excessively  drained, 
Somewhat  excessively 
drained 

Consistence  (moist) 

Very  friable,  Friable 

Firm 

Loose,  Very  firm,  Extremely 
firm,  Extremely  hard 

Salinity  (mmhos/cm) 

0-4 

4-8 

>8 

Depth  to  high  water  table  (ft) 

— 

— 

<2 

Rock  fragments  3  in.  (wt  %) 

>15 

— 

— 

Hydrologic  soil  group 

B 

C 

A,  D 

Salt-derived  surface  crust 

— 

Present 

Habitat  Management 

Habitat  managers  are  naturally  concerned  with 
existing  habitat  conditions.  Management  objectives, 
however,  may  well  call  for  more  than  maintenance 
of  the  status  quo.  Effective  management  often  neces- 
sitates change  and  with  it  the  need  to  predict  the 
likelihood  of  success,  risks  involved,  and  any  special 
practices  that  may  be  desirable.  This  requires  a 
knowledge  of  the  present,  but  equally  important,  a 
knowledge  of  habitat  capability.  Habitat  capability  as 
embodied  in  the  concepts  of  habitat  type  (Dauben- 
mire  1968;  Dyksterhuis  1983),  and  ecological 
(range)  site  (U.S.  Soil  Conservation  Service  1976; 
Society  for  Range  Management  1983)  is  expressed  as 
the  potential  plant  community.  Where  sites  have 
been  subjected  to  little  disturbance,  potential  can 
often  be  described  directly  or  deduced  from  the 
existing  plant  community.  As  a  result  of  past  land 
use,  however,  present  vegetation  may  bear  little 
resemblance  to  the  potential.  In  such  situations,  soils 
may  provide  the  only  identifiable  link  between  pres- 
ent and  potential  (Munn  et  al.  1978;  Passey  et  al. 
1982).  With  soil  as  the  common  denominator, 
knowledge  of  potential  plant  communities  can  be 
extrapolated  from  a  few  relict  sites  to  numerous 
disturbed  or  altered  sites.  In  addition,  arraying  vege- 
tational  data  from  several  sites  on  the  same  or  simi- 
lar soils  reveals  the  variety  of  developmental  stages, 
presumably  leading  to  the  same  potential  plant  com- 
munity. Comparison  of  the  present  with  the  poten- 
tial provides  a  valuable  clue  as  to  site  responsiveness 
and  thereby  aids  in  evaluating  opportunities  and 
choosing  among  various  management  alternatives. 


The  concept  of  site  capability  goes  beyond  an 
interest  in  potential  plant  communities.  Equally  im- 
portant are  soil-related  limitations  in  the  selection  of 
habitat  improvement  methods  and  materials  and  in 
the  establishment  of  reasonable  use  restrictions  to 
prevent  habitat  degradation.  Some  soils  have  few 
limitations.  They  are  often  deep,  flat-lying,  well- 
drained,  and  are  free  of  surface  stones  or  boulders. 
Other  soils  may  have  yearlong  or  perhaps  seasonal 
limitations  due  to  wetness.  Surface  stones  may  pre- 
clude the  use  of  certain  types  of  seeding  equipment 
and  brush  choppers.  Hardpans  and  soils  shallow  to 
bedrock  can  obviously  affect  projects  requiring  exca- 
vation. Failure  to  observe  such  limitations  can  dam- 
age equipment,  increase  construction  and 
maintenance  costs,  reduce  the  effective  life  of  an 
improvement,  and  result  in  unnecessary  habitat 
degradation. 


Soil  survey  reports  commonly  rate  soils  for  shal- 
low excavations  (e.g.,  pipelines),  pond  and  reservoir 
areas,  embankments,  and  as  sources  of  construction 
materials  (e.g.,  sand,  gravel,  topsoil).  Many  other 
interpretations  applicable  to  habitat  management  are 
being  developed  although  there  has  been  insufficient 
testing  and  validation  to  warrant  widespread  use. 
Examples  include  soil  limitations  for  rangeland 
equipment  (e.g.,  drills,  discs),  tree  planting,  fencing, 
prescribed  burning,  and  the  use  of  off-road  vehicles. 
A  strong  commitment  to  testing  and  refining  inter- 
pretations is  essential  if  they  are  to  be  applied  with 
confidence. 


Soils 


583 


LITERATURE  CITED 


ALBRECHT,  W.A.  1957.  Soil  fertility  and  biotic  geography. 
Geog.  Rev.  47(1):86-105. 

AUMANN,  G.D.  1965.  Microtine  abundance  and  soil  so- 
dium levels.  J.  Mammal.  46(4):594-604. 

BATSON,  J.D.  1965.  Some  comparative  behavior  studies 
on  three  genera  of  salamanders.  Trans.  Ky.  Acad.  Sci. 
25:120-128. 

BIENEK,  G.K  and  AW.  GRUNDMANN.  1971.  Burrowing 
habits  of  two  subspecies  of  Dipodomys  merriami 
in  California  and  Utah.  Great  Basin  Nat.  3 1(3):  190- 
192. 

BORCHELT,  PL.  and  S.R.  OVERMANN.  1974.  Develop- 
ment of  dust-bathing  in  bobwhite  quail:  I.  Effects  of 
age,  experience,  texture  of  dust,  and  social  facilita- 
tion. Dev.  Psychobiol.  7(4):305-313. 

BUCKMAN,  HO.  and  N.C.  BRADY.  1969.  The  nature  and 
properties  of  soils.  The  Macmillian  Company.  New 
York,  NY.  567pp. 

CRAWFORD,  W.T.  1950.  Some  specific  relationships  be- 
tween soils  and  wildlife.  J.  Wildl.  Manage.  14(2):1 15- 
123. 

.  1953-  Relationships  of  soils  and  wildlife.  Pages  10- 

18  in  Nagel,  W.O.  ed.  Wildlife  and  the  Soil.  Missouri 
Conservation  Commission,  Jefferson  City,  MO. 

DALE,  F.H.  1939-  Variability  and  environmental  responses 
of  the  kangaroo  rat,  Dipodomys  heermanni  saxatilis. 
Amer.  Midi.  Nat.  22:703-731. 

DAUBENMIRE,  R.  1968.  Plant  communities:  A  textbook  of 
plant  synecology.  Harper  and  Row.  New  York,  NY. 
300pp. 

.  1972.  Annual  cycles  of  soil  moisture  and  tempera- 
ture as  related  to  grass  development  in  the  steppe 
of  eastern  Washington.  Ecol.  53(3):4l9-424. 

DAVIS,  W.B.  1938.  Relation  of  size  of  pocket  gophers  to 
soil  and  altitude.  J.  Mammal  19:338-342. 

,  R.R.  RAMSEY,  and  J.M.  ARENDALE.  1938.  Distribu- 
tion of  pocket  gophers  (Geotnys  breviceps)  in  rela- 
tion to  soils.  J.  Mammal.  19:412-418. 

DENNY,  AH.  1944.  Wildlife  relationships  to  soil  types. 
Trans.  N.  Am.  Wildl.  Conf.  9:316-322. 

DICE,  L.R.  and  P.M.  BLOSSOM.  1937.  Studies  of  mammal- 
ian ecology  in  southwestern  North  America  with 
special  attention  to  the  colors  of  desert  mammals. 
Carnegie  Inst.  Wash.  Publ.  No.  485:1-129. 

DYKSTERHUIS,  E.J.  1983.  Habitat-type:  A  review.  Range- 
lands  5(6):270-271. 

FELDHAMER,  GA.  1979.  Vegetative  and  edaphic  factors 
affecting  the  abundance  and  distribution  of  small 
mammals  in  southeast  Oregon.  Great  Basin  Nat. 
39(3):207-218. 

FISHER,  S.  and  P.  DEUTSCH.  1983.  The  soil  resource:  Its 
importance  in  the  West  and  its  role  in  coal  develop- 
ment and  reclamation  in  Coal  Development:  Col- 
lected Papers.  Papers  presented  at  Coal  Development 
Workshops  in  Grand  Junction,  Colorado,  and  Casper, 
Wyoming.  Sponsored  by  U.S.  Dep.  Inter.,  Bur.  Land 
Manage.  Vol.  11:845-984. 

GESSEL,  S.P.  and  G.H.  ORIANS.  1967.  Rodent  damage  to 
fertilized  silver  fir  in  western  Washington.  Ecol. 
48(4):694-697. 

HARDY,  R.  1945.  The  influence  of  types  of  soil  upon  the 
local  distribution  of  small  mammals  in  southwestern 
Utah.  Ecol.  Monog.  15(  1  ):7 1-108. 


HARPER,  J.A.  and  R.F.  LABISKY.  1964.  The  influence  of 
calcium  on  the  distribution  of  pheasants  in  Illinois.  J. 
Wildl.  Manage.  28:722-731. 

HILL,  E.P.,  HI.  1972.  Litter  size  in  Alabama  cottontails  as 
influenced  by  soil  fertility.  J.  Wildl.  Manage.  36:1199- 
1209. 

HODGSON,  J.R.  1972.  Local  distribution  of  Microtus  mon- 
tanus  and  M.  pennsylvanicus  in  southwestern  Mon- 
tana. J.  Mammal.  53(3):487-499. 

KANTRUD,  HA.  and  R.L.  KOLOGISKI.  1982.  Effects  of 
soils  and  grazing  on  breeding  birds  of  uncultivated 
upland  grasslands  of  the  Northern  Great  Plains.  U.S. 
Dep.  Inter.,  Fish  and  Wildl.  Serv.  Wildl.  Res.  Rep.  15. 
Washington,  DC.  33pp. 

KLEMMEDSON,  JO.  1970.  Needs  for  soil  information  in 
the  management  of  range  resources.  J.  Range  Manage. 
23(2):  139- 143. 

KREBS,  C.J.,  B.L.  KELLER,  and  J.H.  MYERS.  1971.  Microtus 
population  densities  and  soil  nutrients  in  southern 
Indiana  grasslands.  Ecol.  52(4):660-663. 

KRUCKEBERG,  A.R.  1969.  Soil  diversity  and  the  distribu- 
tion of  plants  with  examples  from  western  North 
America.  Madrono  20:129-154. 

LEOPOLD,  AS.  and  P.D.  DALKE.  1943.  The  1942  status  of 
wild  turkeys  in  Missouri.  J.  Forest.  4l(6):428-435. 

LUNT,  OR.,  J.  LETEY,  and  SB.  CLARK.  1973-  Oxygen 
requirements  for  root  growth  in  three  species  of 
desert  shrubs.  Ecol.  54(6):1356-1362. 

LYNCH,  J.D.  1978.  The  distribution  of  leopard  frogs  {Rana 
blairi  and  R.  pipiens)  (Amphibia,  Anura,  Ranadae)  in 
Nebraska.  J.  Herpetol.  12(2):  157- 162. 

MARCHAND,  D.E.  1973.  Edaphic  control  of  plant  distribu- 
tion in  the  White  Mountains,  eastern  California.  Ecol. 
54(2):233-250. 

MASER,  C,  J.M.  GEIST,  DM.  CONCANNON,  R.  ANDER- 
SON, and  B.  LOVELL.  1979.  Wildlife  habitat  in  man- 
aged rangelands — the  Great  Basin  of  southeastern 
Oregon:  geomorphic  and  edaphic  habitats.  U.S.  Dep. 
Agric,  For.  Serv.  Gen.  Tech.  Rep.  PNW-99.  Portland, 
OR.  84pp. 

MCARTHUR,  ED.,  A.P.  PLUMMER,  and  J.N.  DAVIS.  1978. 
Rehabilitation  of  game  range  in  the  salt  desert.  Pages 
23-50  in  Wyoming  Shrublands,  Proc.  Seventh  Wyo- 
ming Shrub  Ecology  Workshop,  Rock  Springs,  WY. 
Range  Manage.  Div.,  Univ.  WY.  Laramie. 

MCCOLLEY,  P.D.  and  H.S.  HODGKINSON.  1970.  Effect  of 
soil  depth  on  plant  production.  J.  Range  Manage. 
23(3):  189- 192. 

MOSS,  A.E.  1940.  The  woodchuck  as  a  soil  expert.  J.  Wildl. 
Manage.  4:441-443 

MUNN,  L.C.,  GA.  NIELSEN,  and  W.F.  MUEGGLER.  1978. 
Relationships  of  soils  to  mountain  and  foothill  range 
habitat  types  and  production  in  western  Montana.  Soil 
Sci.  Soc.  Amer.  J.  42:135-139. 

MUNSELL  COLOR.  1975.  Munsell  Soil  Color  Charts.  Mac- 
beth Division,  Kollmorgen  Corp.  Baltimore,  MD. 
21218. 

MUSHINSKY,  H.R.  and  E.D.  BRODIE,  Jr.  1975.  Selection  of 
substrate  pH  by  salamanders.  Amer.  Midi.  Nat. 
93(2):440-443. 

NAGY,  J.G  1966.  Volatile  oils  and  antibiosis  of  Artemisia. 
PhD.  dissertation.  Colorado  State  Univ.  Ft.  Collins. 
73pp. 

,  H.W.  STEINHOFF,  and  G.W.  WARD.  1964.  Effects 

of  essential  oils  of  sagebrush  on  deer  rumen  microbial 
function.  J.  Wildl.  Manage.  28:785-790. 


584 


Soils 


O'FARRELL,  T.P.  1972.  Ecological  distribution  of  sage- 
brush voles,  Lagurus  curtatus,  in  south-central  Wash- 
ington. J.  Mammal.  53(3):632-636. 

PASSEY,  H.B.,  V.K.  HUGIE,  E.W.  WILLIAMS,  and  D.E.  BALL. 
1982.  Relationships  between  soil,  plant  community, 
and  climate  on  rangelands  of  the  Intermountain  West. 
U.S.  Dep.  Agric,  Soil  Conserv.  Serv.  Tech.  Bull.  1669- 
123pp. 

POWELL,  J.  1970.  Site  factor  relationships  with  volatile 
oils  in  big  sagebrush.  J.  Range  Manage.  23(  1  ):42-46. 

PROCTOR,  J.  and  K.  WHITTEN.  1971.  A  population  of  the 
valley  pocket  gopher  (Thomomys  bottae)  on  a  ser- 
pentine soil.  Amer.  Midi.  Nat.  85(  2  ):5 17-521. 

ROSENZWEIG,  M.L.  and  J.  Winakur.  1969.  Population 
ecology  of  desert  rodent  communities:  Habitats  and 
environmental  complexity.  Ecol.  50(4):558-572. 

SHEETS,  R.G.,  R.L.  LINDER,  and  R.B.  DAHLGREN.  1971. 
Burrow  systems  of  prairie  dogs  in  South  Dakota.  J. 
Mammal.  52(2):451-453. 

SHELDON,  W.G.  1950.  Denning  habits  and  home  range  of 
red  foxes  in  New  York  State.  J.  Wildl.  Manage. 
14:33-42. 

SHORT,  H.L.  1979.  Deer  in  Arizona  and  New  Mexico: 

Their  ecology  and  a  theory  explaining  recent  popula- 
tion decreases.  U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech. 
Rep.  RM-70.  25pp. 

SOCIETY  FOR  RANGE  MANAGEMENT.  1983.  Guidelines 
and  terminology  for  range  inventories  and  monitor- 
ing. Report  of  the  Range  Inventory  Standardization 
Committee  (RISC).  13pp. 

SOIL  SURVEY  STAFF.  1975.  Soil  taxonomy:  A  basic  sys- 
tem of  soil  classification  for  making  and  interpreting 
soil  surveys.  U.S.  Dep.  Agric,  Soil  Conserv.  Serv., 
Agric.  Handb.  436.  754pp. 

SPIGHT,  T.  M.  1967.  The  water  economy  of  salamanders: 
Exchange  of  water  with  the  soil.  Biol.  Bull. 
132(1):  126- 132. 

STEBLER,  A.M.  1939.  An  ecological  study  of  the  mammals 
of  the  Badlands  and  the  Black  Hills  of  South  Dakota 
and  Wyoming.  Ecol.  20:382-393. 

STOCKSTAD,  D.S.,  M.S.  MORRIS,  and  EC.  LORY.  1953. 
Chemical  characteristics  of  natural  licks  used  by  big 
game  animals  in  western  Montana.  Trans.  N.  Am. 
Wildl.  Conf.  18:247-257. 

STUART,  L.C.  1932.  The  lizards  of  the  middle  Pahvant 
Valley,  Utah:  Materials  for  a  study  in  saurian  distribu- 
tion. Occ  Papers  Mus.  Zool.,  Univ.  Mich.  244:1-33- 

TAUB,  F.B.  1961.  The  distribution  of  the  red-backed  sala- 
mander, Plethodon  C.  cinereus,  within  the  soil.  Ecol. 
42(4):681-698. 

THORNBURG,  AA.  1982.  Plant  materials  for  use  on  sur- 


face-mined lands  in  arid  and  semiarid  regions.  U.S. 
Dep.  Agric,  Soil  Conserv.  Serv.  SCS-TP-157.  88pp. 

TURNER,  L.W.  1972.  Habitat  differences  between  Spermo- 
philus  beldingi  and  Spermophilus  columbianus  in 
Oregon.  J.  Mammal.  53(4):9l4-917. 

U.S.  BUREAU  OF  LAND  MANAGEMENT.  1982.  Inte 
grated  Habitat  Inventory  and  Classification  System. 
U.S.  Dep.  Inter.,  Bur.  Land  Manage.  Manual  6602. 
Washington,  DC. 

U.S.  SALINITY  LABORATORY  STAFF.  1954.  Diagnosis 
and  improvement  of  saline  and  alkali  soils.  U.S.  Dep. 
Agric,  Agric.  Handb.  60.  Washington,  DC.  160pp. 

U.S.  SOIL  CONSERVATION  SERVICE.  1972.  Soil  Survey 
Laboratory  Methods  and  Procedures  for  Collecting 
Soil  Samples.  Soil  Survey  Investigation  Report  (SSIR) 
No.  1.  U.S.  Dep.  Agric,  Soil  Conserv.  Serv.  Washing- 
ton, DC.  20013- 

Additional  reports,  containing  soil  survey  labora- 
tory data  for  specific  states  include: 
SSIR  No.     7  Montana 
SSIR  No.     8  Wyoming 
SSIR  No.  10  Colorado 
SSIR  No.  1 1  Oklahoma 
SSIR  No.  22  Alabama  and  Florida 
SSIR  No.  23  Nevada 
SSIR  No.  24  California 
SSIR  No.  28  Arizona 

.  1976.  National  range  handbook.  NRH-1.  U.S.  Dep. 

Agric,  Soil  Conserv.  Serv.  Washington,  DC.  20013- 

.  1981.  Soil  survey  manual  SSM-430-V.  U.S.  Dep. 

Agric,  Soil  Conserv.  Serv.  Washington,  DC.  20013 

.  1982.  List  of  published  soil  surveys:  January  1982. 

U.S.  Dep.  Agric,  Soil  Conserv.  Serv.  Washington,  DC. 
20013. 

.  1983.  National  soils  handbook.  NSH-430-VI.  U.S. 

Dep.  Agric,  Soil  Conserv.  Serv.  Washington,  DC. 
20013. 

WEAVER,  T  1977.  Root  distribution  and  soil  water  re- 
gimes in  nine  habitat  types  of  the  northern  Rocky 
Mountains.  Marshall,  J.  ed.  The  Below  Ground  Ecosys- 
tem. Range  Sci.  Dep.  Sci.  Series  26.  Colorado  State 
Univ.,  Fort  Collins.  351pp. 

WERTZ,  W.A.  1966.  Interpretation  of  soil  surveys  for  wild- 
life management.  Amer.  Midi.  Nat.  75(1  ):2 11-224. 

WHITAKER,  G.A.  1965.  Deer-soil  relationships  in  the  oak- 
grasslands.  M.S.  Thesis,  Humboldt  State  College.  Ar- 
eata, CA.  87pp. 

WILLIAMS,  C.E.  and  A.L.  CASKEY.  1965.  Soil  fertility  and 
cottontail  fecundity  in  southeastern  Missouri.  Amer. 
Mid.  Nat.  74(  1  ):21 1-224. 


Soils 


585 


27 

TERRESTRIAL 

PHYSICAL 

FEATURES 

Allen  Cooperrider 

U.S.  Bureau  of  Land  Management 
Service  Center 
Denver,  CO  80225 


"With  a  broad  awareness  and  understanding  of  par- 
ticular physiographic  features,  one  begins  to  appreci- 
ate them  as  wildlife  habitat,  their  relationship  to 
land  mangement,  and  the  need  to  account  for  them 
in  land-use  planning." 


— Chris  Maser  et  al.,  Geomorphic  and  Edaphic 
Habitats 


Editor's  Note:  The  wildlife  biologist  instinctively 
recognizes  the  importance  of  physical  features  for 
wildlife — the  nesting  cliff  for  falcons,  the  talus 
slope  for  marmots,  the  south-facing  snow-free  win- 
ter range  for  mule  deer.  However,  habitat  invento- 
ries and  evaluations  often  concentrate  exclusively 
on  vegetation  and  ignore  these  important  habitat 
components  and  attributes.  Formal  systems  for 
classifying  and  naming  such  features  as  wildlife 
habitat  are  virtually  non-existent  for  terrestrial 
features.  This  chapter  covers  the  classification  and 
measurement  of  terrestrial  features. 


INTRODUCTION 

Physical  features  of  the  environment  are  just  as 
important  as  vegetation  in  determining  animal  abun- 
dance and  distribution.  Many  wildlife  species  have 
adapted  to  using  certain  physical  features  of  the  en- 
vironment to  the  point  where  they  cannot  survive 
and  reproduce  without  them.  Consider  such  obligate 
relationships  as  use  of  cliffs  for  breeding  by  prairie 
falcons  and  many  other  raptors,  or  use  of  steep,  rug- 
ged topography  for  escape  cover  by  bighorn  sheep 
(Ovis  canadensis). 

Biologists  and  naturalists  have  long  recognized 
these  dependencies  and  use  of  physical  features  by 
individual  species.  However,  little  work  has  been 
done  to  develop  general  methodology  for  categoriz- 
ing or  quantifying  the  features  themselves.  Several 
reasons  account  for  this  neglect.  First,  some  of  the 
most  important  attributes  of  physical  features  are 
spatial  or  geographic.  For  example,  the  most  impor- 
tant attribute  about  a  feature,  such  as  a  seep  in  a 
limestone  outcropping,  is  its  location.  If  biologists 
have  marked  the  location  on  a  map,  they  have  re- 
corded the  data  needed  to  predict,  for  instance,  that 
a  species  of  salamander  may  be  present.  Measure- 
ments of  height,  slope,  etc.,  may  be  quite  irrelevant. 
Even  with  features  such  as  cliffs,  where  attributes 
such  as  height  are  important,  broad  categories,  such 
as  "10  to  30  m"  (33  to  99  ft)  and  "over  30  m"  (99 
ft),  are  adequate  for  most  purposes. 

Second,  most  biologists  only  work  with  a  lim- 
ited number  of  species  and  they  know  from  experi- 
ence what  a  suitable  physical  feature  looks  like.  For 
example,  bighorn  sheep  biologists  know  that  "steep, 
rugged,  rocky  terrain"  is  an  essential  component  of 
sheep  habitat.  However,  since  they  easily  recognize 
such  habitat,  few  bother  to  attempt  to  quantify 
"steep,"  "rugged,"  or  "rocky."  This  lack  of  quantifica- 
tion of  terrestrial  physical  features  sharply  contrasts 
with  aquatic  features,  for  which  well-developed 
measurement  systems  have  been  in  use  for  years. 

However,  the  need  for  more  quantification  or  at 
least  more  standardized  categorization  has  become 
evident.  Biologists  must  deal  with  many  more  spe- 


Terrestrial  Physical  Features 


587 


cies  than  in  the  past,  and  it  is  useful  to  have  quanti- 
tative measures  or  standard  categories  of  physical 
features  that  they  can  use  with  unfamiliar  species. 
Furthermore,  biologists  must  document  their  work 
more  thoroughly.  The  subjective  judgment  that  "it 
looks  like  bighorn  habitat  to  me"  is  being  replaced 
or  corroborated  by  simple  measurements  that  can  be 
compared  to  a  standard  description  to  confirm  such 
judgment. 

The  limited  literature  on  terrestrial  physical 
features  is  confusing  due  to  slightly  different  termi- 
nology and  varying  approaches  in  determining  physi- 
cal features  from  biological  ones.  Is  a  snag,  for 
example,  physical  or  biological?  A  concrete  dam  is 
certainly  a  man-made  physical  feature,  but  what  is  a 
beaver  dam?  Similarly,  physical  structures,  such  as 
playas,  are  sometimes  considered  landforms  (Peter- 
son 1981)  and  other  times  considered  geomorphic 
features  (Maser  et  al.  1979a). 

Much  of  this  confusion  is  caused  by  biologists 
borrowing  terms  and  definitions  from  other  fields 
such  as  soil  science  and  geology.  These  definitions 
have  been  developed  for  quite  different  purposes. 
For  the  purposes  of  wildlife  biology,  how  a  feature  is 
classified  (e.g.,  landform  or  geomorphic  feature)  is 
probably  not  too  important,  as  long  as  biologists 
understand  what  the  terms  mean  and  know  how  to 
use  such  features  to  predict  species  occurrence.  If 
biologists  understand  what  a  playa  is  and  what  ani- 
mal species  depend  on  such  a  feature,  it  does  not 
matter  if  they  call  it  a  landform  because  of  its  large 
size  or  a  geomorphic  feature  because  it  is  a  small, 
distinct,  localized  habitat. 

For  the  purposes  of  this  chapter,  terrestrial 
physical  features  are  all  features  of  the  terrestrial 
environment,  including  man-made,  that  are  not  com- 
posed of  living  vegetation,  but  are  useful  in  predict- 
ing animal  abundance  and  distribution.  Therefore, 
this  chapter  identifies  those  features  that  are  com- 
monly termed  landforms  (Peterson  1981 ),  geo- 
morphic and  edaphic  habitats  (Maser  et  al.  1979a), 
special  habitat  features  (Hamilton  et  al.  1983),  man- 
made  habitats  (Maser  et  al.  1979b),  and  features 
derived  from  vegetation  as  snags  and  litter.  Soils  also 
fall  into  this  category.  However,  because  of  the  im- 
portance of  soils  in  determining  both  plant  and  ani- 
mal distribution,  an  entire  chapter  in  this  publication 
has  been  devoted  to  this  relationship  (Chapter  26). 
Soils  are  only  briefly  discussed  in  this  chapter.  This 
chapter  emphasizes  "wildland"  environments,  those 
habitats  that  are  not  used  for  urban,  suburban,  or 
agricultural  purposes,  but  which  may  be  managed  for 
forestry,  livestock  production,  or  other  purposes. 

The  dearth  of  literature  on  the  subject  has  been 
mentioned.  Several  publications  are  available,  how- 
ever, that  must  be  considered  as  pioneering  efforts 
in  the  field,  and  they  will  be  cited  and  quoted  fre- 


quently in  the  following  discussions.  The  biologist 
who  is  seriously  involved  in  measuring  terrestrial 
physical  features  should  obtain  copies  of  these 
publications. 

The  paper  by  Maser  et  al.  (1979a)  on  "Geo- 
morphic and  Edaphic  Habitats"  is  one  of  the  first  at- 
tempts to  list,  categorize,  and  quantify  these  habitats. 
Although  it  focuses  on  the  Great  Basin  of  southeast- 
ern Oregon,  much  of  the  information  and  suggested 
approaches  can  apply  to  different  regions.  Similarly, 
the  paper  by  Maser  et  al.  (1979b)  on  "Man-Made 
Habitats"  published  in  the  same  series  may  have 
widespread  application.  Biologists  have  been  aware 
for  years  of  the  benefits  of  man-made  habitat  features 
such  as  nest  boxes  and  gallinaceous  guzzlers  that 
were  developed  specifically  for  wildlife.  In  the  latter 
paper,  the  importance  to  many  species  of  wildlife 
of  certain  man-made  features,  such  as  mine  shafts, 
abandoned  buildings,  and  telephone  poles,  that  were 
developed  for  quite  different  reasons,  is  also 
described. 


A  paper  by  Short  and  Burnham  ( 1982)  repre- 
sents one  of  the  first  attempts  to  systematically  de- 
fine structural  components  of  habitat  that  are  used 
for  feeding  and  breeding.  They  use  both  physical 
strata  and  vegetation  structure  for  this  purpose.  The 
former  provides  one  of  the  first  systematic  efforts 
to  provide  a  comprehensive  listing  of  physical  fea- 
tures used  by  vertebrate  species. 

Finally,  Hays  et  al.  (  1981 )  have  provided  a  com- 
prehensive and  systematic  listing  of  measurement 
techniques  for  habitat  variables.  Many  procedures, 
such  as  estimating  height,  can  be  applied  to  a  variety 
of  habitat  features,  such  as  cliffs,  trees,  power  poles, 
etc.  I  refer  to  this  publication  frequently  for  guid- 
ance on  making  these  measurements  or  estimates, 
and  for  information  on  considerations  such  as  preci- 
sion, accuracy,  cost,  efficiency,  equipment,  and  train- 
ing associated  with  different  sampling  techniques. 


FEATURES  AND  THEIR  MEASUREMENT 

Physical  features  are  discussed  here  under  seven 
general  categories:  ( 1 )  landforms,  (2)  geomorphic 
habitats,  (3)  soils,  (4)  edaphic  habitats,  (5)  vegeta- 
tion-derived features,  (6)  animal-made  features,  and 
(7)  man-made  features.  These  categories  represent  a 
convenient  grouping  based  primarily  on  two  factors: 
(1 )  derivation  of  the  feature  and  (2)  size  of  the  fea- 
ture. Some  features  such  as  a  beaver  dam,  can  fall 
easily  into  two  categories:  vegetation  derived  and  an- 
imal made.  It  may  be  classified  as  either  one  or  both 
depending  on  the  purpose  of  the  investigation. 
Where  authors  have  categorized  or  defined  features 
slightly  differently,  I  point  out  such  differences. 


588 


Terrestrial  Physical  Features 


Landforms 

A  landform  is  "a  three-dimensional  part  of  the 
land  surface,  formed  of  soil,  sediment,  or  rock,  that  is 
distinctive  because  of  its  shape,  that  is  significant  for 
land  use  or  to  landscape  genesis,  that  repeats  in  var- 
ious landscapes,  and  that  also  has  fairly  consistent 
position  relative  to  surrounding  landforms"  (Peter- 
son 1 98 1 ).  Alluvial  fans  and  rock  pediment  are  ex- 
amples of  landforms.  Landforms  are  the  basic 
building  blocks  of  the  landscape  and,  thus,  have  a 
dominating  influence  on  the  soils,  vegetation,  and  an- 
imal life  that  develop  on  their  surfaces.  For  this  rea- 
son, landform  is  quite  useful  in  predicting  animal 
occurrence. 

The  following  two  aspects  of  landforms  are  usu- 
ally considered: 

Classification.  Classification  of  landforms  is  the 
process  of  delineating  an  area  on  a  map  as  a 
landform  type.  A  landform  is  thus  always  something 
with  area;  this  contrasts  with  most  other  physical 
features  used,  which  tend  to  be  small  and  localized 
and  typically  treated  as  points  on  maps.  Landform 
can  usually  be  classified  by  simple  visual  obser- 
vation or  by  interpretation  of  appropriate  aerial 
photography  by  a  biologist  who  is  familiar  with  the 
concept  and  the  local  landforms.  Unfortunately,  few 
if  any  landform  classifications  have  been  developed 
for  use  by  wildlife  biologists.  In  fact,  landform 
classifications  for  any  purpose  are  not  available  for 
most  areas.  Therefore,  in  many  cases,  classifications 
will  have  to  be  developed. 


Peterson  (1981 )  developed  a  landform  classifica- 
tion for  the  Basin  and  Range  Province  for  purposes 
of  soil  surveys  (Table  1 ),  and  this  has  been  used 
by  many  range  conservationists,  botanists,  and  wild- 
life biologists  for  their  own  purposes.  This  is  a  hier- 
archical classification  in  which  major  landforms  are 
broken  down  into  component  landforms,  landform 
elements,  and  slope  components.  A  biologist  not 
wanting  to  use  all  levels  of  the  classification  may  find 
the  lower  levels  of  classification  ( landform  elements, 
slope  components)  useful  in  predicting  animal  oc- 
currence. However,  Peterson's  classification  should 
be  useful  in  developing  classifications  for  other  areas. 

The  U.S.  Department  of  Interior,  Bureau  of  Land 
Management's  (BLM's)  Integrated  Habitat  Inventory 
and  Classification  System  (IHICS)  (1982)  lists  land- 
forms  used  by  biologists;  these  represent  ad  hoc 
definitions,  and  new  categories  are  being  added  con- 
tinuously. However,  most  of  the  initial  definitions 
came  from  Peterson  (  1981 ). 

Attributes.  Biologists  frequently  find  that  attributes 
of  landforms,  such  as  slope  and  aspect,  are  useful 
in  predicting  animal  occurrence.  These  types  of 
measurements  can  be  made  anywhere  and  treated  as 
point  measurements  (i.e.,  a  characteristic  of  an  exact 
location).  Since  measurements  such  as  slope  are 
usually  averaged  and  used  to  categorize  a  landform 
or  some  other  area  of  land,  they  are  treated  here 
as  landform  attributes.  The  three  most  common  such 
measurements  are  slope,  aspect,  and  elevation.  Hays 
et  al.  (  1981 )  describe  the  first  two  measurements. 


1 

LU 
00 

E 
< 

o 

h- 
IX 

LU 

>   < 

^4  5° 

HORIZONTAL  RUN 

SLOPE  OF  45°OR    100% 

SLOPE 

SOUTH 
ASPECT 


ASPECT 


ELEVATION 

Terrestrial  Physical  Features 


589 


Table  1.     Landform  Classification  System  for  the  Basin  and  Range  Province  (from  Peterson  1981). 


Landtorms 

Parts  of  Landforms 

1 

II 

III 

IV 

V 

Major 

Component 

Physiographic  Part 

Major  Landform 

Landform 

Landform  Element 

Slope  Component 

Bounding  mountains 

(not  defined) 

Piedmont  slope 

Mountain-valley 

Erosional  fan 

Summit 

fan 

remnant 

QiH£iQlr*r~i£> 

Shoulder 

OlUt-olU[Jt; 

Backslope 

Footslope 

Partial  ballena 

Crest 
Shoulder 
Backslope 
Footslope 

Inset  fan 

Channel 
Channel 

Rock  pediment 

Rock-pediment 

Qi  i mmit  or    _ 

Crest 

OUilllllll  Ul 

remnant 

sideslope 

Channel 

Shoulder 
Backslope 

Footslope 

Ballena 

Crest 
Shoulder 

Backslope 

Footslope 

Channel 

Inset  fan 

Channel 

Alluvial  fan 

Fan  collar 

Channel 

Erosional  fan 

Summit 

remnant 

QiHoclnno 

Shoulder 

OlUcblupt; 

Backslope 

Footslope 

Partial  ballena 

Crest 
Shoulder 
Backslope 
Footslope 

Channel 

Inset  fan 

Channel 

Fan  pediment 

Erosional  fan 

Summit 

remnant 

QiHoclnno  - 

Shoulder 

OlUcolUJJt; 

Backslope 

Footslope 

Partial  ballena 

Crest 

Shoulder 

Backslope 

Channel 

Inset  fan 

Channel 

Fan  apron 

Channel 

Nonburied  fan 

Channel 

remnant 

Beach  terrace 

Pediment 

Pediment  remnant 

Summit 

QiHoclnno  - 

Shoulder 

OIUfc.olU|JC 

Backslope 

Footslope 

Channel 

Fan  okirt 

Channel 

I    c*l  I    or\l I  I 

Beach  terrace 

590 


Terrestrial  Physical  Features 


Table  1.     Landform  Classification  System  for  the  Basin  and  Range  Province  (concluded). 


Landforms 

Parts  of  Landforms 

1 

Major 

Physiographic  Part 

II 
Major  Landform 

III 

Component 

Landform 

IV 
Landform  Element 

V 
Slope  Component 

Basin  floor  (Bolsom 
or  semi-Bolsom 
floor) 

Alluvial  flat 

Relict  alluvial  flat 
Recent  alluvial  flat 

Channel 
Channel 

Alluvial  plain 

Sand  sheet 

Sand  dune 
(Parna  dune) 

Interdune  flat 

Beach  plain 
(Bolson  floor  only) 

Offshore  bar 
Barrier  bar 
Lagoon 

Channel 

Lake  plain 
(Bolson  floor  only) 

Lake-plain  terrace 

Channel 

Playa  (Bolson  floor 
only) 

Floodplain  playa 

Channel 

Axial-stream  flood- 
plain 

Floodplain  playa 
Stream  terrace 

Channel 
Summit 

Shoulder 

Backslope 

Footslope 

Slope  is  the  angle  between  the  horizontal  and 
the  plane  of  the  ground  surface.  It  is  usually  ex- 
pressed in  either  angle  degrees  or  percent.  In  the 
latter  case,  percent  is  the  vertical  rise  per  horizontal 
run.  Thus,  a  45-degree  angle  would  be  100%  so 
that  slopes  of  over  100%  can  occur.  Slope  can  be 
measured  using  a  clinometer  (Hays  et  al.  1981:87). 
Usually  an  area  will  vary  quite  a  bit  in  slope  due 
to  microtopography.  Since  a  biologist  is  usually  inter- 
ested in  the  average  slope  of  the  entire  area,  many 
slope  measurements  are  usually  taken  and  then  aver- 
aged. Certain  types  of  "badland"  areas  may  have 
very  steep  slopes  even  though  there  is  little  net 
slope  in  any  direction  from  one  boundary  of  the  area 
to  another.  In  such  cases,  average  slope  would  be 
close  to  zero  and  thus  meaningless.  However,  the 
variance  of  slope  measurements  may  be  a  useful 
measure  of  the  ruggedness  of  the  topography. 

Aspect  is  the  downhill  direction  of  a  slope  or 
the  direction  a  slope  faces.  It  is  usually  measured  in 
conjunction  with  slope  by  using  a  compass  (Hays 
et  al.  1981 :87 ).  As  with  slope,  average  aspect  for  an 
area,  rather  than  a  point  measurement,  is  usually 
determined.  On  flat  or  relatively  flat  areas  of  little 
measurable  slope,  aspect  is  of  negligible  interest  and 
is  often  not  measured. 


Elevation  is  the  height  above  sea  level.  Although 
elevation  can  be  measured  with  an  altimeter,  since 
most  areas  in  the  lower  48  states  have  already  been 
surveyed,  the  elevation  can  be  read  from  a  topo- 
graphic map. 


Slope,  aspect,  and  elevation  are  extremely  im- 
portant determinants  of  animal  distribution.  For  ex- 
ample, in  northern  areas,  south-facing  slopes  are 
substantially  warmer  in  winter  and  contain  less 
snow;  therefore,  animals  congregate  in  these  areas. 
Similarly,  the  distribution  of  many  animal  species 
is  limited  by  elevation.  However,  if  one  has  topo- 
graphic maps,  measurement  of  these  attributes  is  not 
necessary  since  they  can  be  read  (elevation)  or 
measured  (slope  and  aspect)  from  such  maps  (Hays 
et  al.  1981:85).  Furthermore,  if  information  is  being 
stored  in  a  geographic  information  system  (GIS), 
then  digitized  topographic  information  for  many 
areas  is  available  on  tapes.  This  information  can  be 
purchased  from  the  National  Cartographic  Informa- 
tion Center  (NCIC)  of  the  U.S.  Geological  Survey. 
Most  GIS  software  has  commands  for  making  such 
calculations,  so  the  biologist  does  not  need  to  do  the 
process  manually. 

Soils 

Soils  provide  a  substrate  for  vegetation  to  grow 
and  are  a  major  determinant  of  the  type  and  abun- 
dance of  vegetation  that  can  grow  on  a  site.  How- 
ever, soil  also  provides  habitat  for  many  species  of 
fossorial  animals  such  as  moles,  ground  squirrels, 
pocket  gophers,  badgers,  and  many  species  of  am- 
phibians and  reptiles.  Soils  provide  habitat  for — 

•  feeding  (e.g.,  American  woodcock  [Scolopax 
minor],  starnose  mole  \Condylura  aistata]. 
California  legless  lizard  [Aniella  pulchra)); 


Terrestrial  Physical  Features 


591 


•  breeding  or  nesting  (e.g.,  desert  tortoise  [Go- 
pherus  agassizii],  white-tailed  prairie  dog  [Cy- 
nomys  leucurus],  burrowing  owl  [Athene 
cunicularia]); 

•  cover  (e.g.,  California  ground  squirrel  [Spermo- 
philus  beecheyi],  Coachella  Valley  fringe-toed 
lizard  [Utna  notata}). 

The  type,  texture,  and  other  properties  of  the 
soil  will  all  determine  the  suitability  of  an  area  for 
animals  such  as  those  mentioned  above.  For  more  in- 
formation see  Soils,  Chapter  26  in  this  book;  for  fur- 
ther reading,  see  Bailey  (1984:68-78)  and  Robinson 
and  Bolen  (1984.125-146). 


contrast  markedly  with  the  surrounding  area."  They 
described  one  edaphic  habitat  that  supports  a  desert 
shrub  community  of  saltbush  (Atriplex  sp.),  sage- 
brush {Artemisia  sp.),  and  rabbitbrush  (Chrysotham- 
nus  sp. ).  Although  they  recognize  that  edaphic 
habitats  have  universal  occurrence,  they  stated  that 
it  is  of  special  interest  in  the  northern  Great  Basin 
because  it  is  the  primary  habitat  of  the  white-tailed 
antelope  squirrel  (Ammospermophilus  leucurus) 
and  the  only  habitat  of  several  native  endemic  plants. 

Edaphic  habitats  form  the  basis  for  many  wildlife 
habitat  classifications.  As  such,  no  particular  measure- 
ments are  required  but  rather  edaphic  habitats  are 
identified  as  either  areas  or  points  on  maps  and  class- 
ified accordingly. 


Edaphic  Habitats 

An  edaphic  habitat  is  one  in  which  existing  or 
potential  vegetation  is  determined  primarily  by  soils 
rather  than  climate.  To  a  certain  extent,  this  is  true 
of  all  habitats.  Thus,  Odum  (1959)  states  that  for 
a  given  region  it  is  convenient  to  recognize  ( 1 )  a 
single  climatic  climax,  which  is  in  equilibrium  with 
the  general  climate,  and  ( 2  )  a  varying  number  of 
edaphic  climaxes,  which  are  modified  by  the  local 
conditions  of  the  substrate.  Theoretically,  given 
enough  time,  the  edaphic  climaxes  would  converge 
upon  the  climatic  climax.  In  practice,  however,  biol- 
ogists work  with  edaphic  climaxes  in  terms  of  identi- 
fying potential  natural  vegetation.  Thus  the  concept 
of  edaphic  habitat  is  used  commonly  in  vegetation 
and  wildlife  habitat  classifications. 

Edaphic  habitats  have  been  used  in  a  more  re- 
strictive sense,  however,  by  Maser  et  al.  (1979b). 
They  considered  edaphic  features  (1979a:2)  as  "local 
distinctive  soils  that,  along  with  their  vegetation, 


Geomorphic  Habitat  Features 

Geomorphic  habitat  features  are  products  of 
geologic  or  geomorphic  processes  and  include  cliffs, 
caves,  talus,  lava  flows,  sand  dunes,  and  playas.  There 
is  some  overlap  between  what  some  consider  a  land- 
form  and  a  geomorphic  feature.  For  example,  a  playa 
is  considered  a  geomorphic  feature  by  Maser  et  al. 
( 1979a)  and  a  landform  by  Peterson  (  1981 ).  In  gen- 
eral, however,  geomorphic  features  represent  parts 
of  the  larger  landform  and  do  not  necessarily  occur 
in  a  predictable  position  relative  to  surrounding 
landforms  or  geomorphic  features.  For  example,  low 
cliffs  may  be  scattered  throughout  a  larger  landform 
such  as  a  fan  piedmont.  They  do  not  necessarily  bear 
any  consistent  relationship  to  each  other  or  to  other 
landforms.  Similarly,  geomorphic  features  such  as 
caves  are  often  relatively  small  and  localized  and 
treated  as  points  on  maps  rather  than  features  with 
areas.  Landforms,  on  the  other  hand,  always  have 
area. 


m     v'ili;: 
mmSKmm 


Sand  dunes  are  examples  of  geomorphic  habitats. 


Playas  are  considered  either  a  geomorphic  feature  or  a 
landform. 


592 


Terrestrial  Physical  Features 


Maser  et  al.  (1979a)  described  six  geomorphic 
features  of  particular  use  in  predicting  the  occur- 
rence of  vertebrate  species  in  the  Great  Basin  of 
eastern  Oregon:  cliffs,  caves,  talus,  lava  flows,  sand 
dunes,  and  playas.  Other  geomorphic  features  are 
used  in  the  BLM  IHICS  efforts.  In  both  systems,  most 
of  these  features  are  treated  as  point  data,  that  is, 
the  only  information  recorded  during  an  inventory  is 
the  location.  The  area  of  a  rock  shelter,  for  example, 
is  considered  to  be  negligible.  Others  are  treated  as 
point  data  or  an  area  depending  on  size.  For  exam- 
ple, a  cliff  that  is  very  localized  and  occupies  less 
than  a  hectare  (2.5  a.),  may  be  conveniently  treated 
as  a  point  on  a  map,  whereas  in  other  situations, 
where  a  series  of  cliffs  are  found  for  many  kilometers 
along  a  river  system,  they  are  more  usefully  re- 
corded on  the  map  as  polygons  with  area. 

Most  descriptions  of  geomorphic  features  for 
purposes  of  wildlife  habitat  surveys  are  limited  to 
mere  notation  of  occurrence.  However,  systems  of 
classification  or  measurement  have  been  developed 
for  several  of  these  features.  These  systems  are  de- 
scribed below. 

Cliffs.  Maser  et  al.  (1979a)  described  features  of 
igneous  rock  cliffs  such  as  facial  cracks  or  joints, 
ledges  and  occasional  caves,  and  the  shelflike 
features  and  holes  or  pockets  in  sedimentary  rock 
cliffs  as  being  very  important  for  wildlife  species  in 
the  Great  Basin.  Cliffs  are  equally  important  in  other 
regions. 


In  addition  to  noting  presence  and  possibly  area, 
cliffs  can  be  classified  according  to  parent  rock  and 
resistance  (Table  2)  and  according  to  structural 
characteristics  of  height  class,  length  class,  joint 
spacing,  and  joint  width  (Table  3).  Any  of  these  vari- 
ables such  as  height  can  be  measured  according  to 
methods  described  in  Hays  et  al.  (1981). 

Table  2.  General  resistance  of  cliffs  to  weathering 
in  Great  Basin  Region  of  southwestern  Oregon 
(from  Maser  et  al.  1979a). 


Parent  Rock 

Group 

1 

II 

Basalt 

• 

Rhyolite 

• 

Andesite 

• 

Nonwelded  tuff 

o 

Welded  tuff 

• 

Sedimentary  rock 

o 

(weakly  consoli- 

dated) 

more  resistant 


0=  less  resistant 


Caves.  Caves  are  extremely  important  for  many 
species  of  vertebrates  because  they  provide  shelter 
from  weather  extremes,  stable  environments, 
darkness,  and  seclusion.  Generally  caves  are 
recorded  as  points  on  maps.  Like  cliffs,  they  can  be 


Igneous  rock  cliff's  are  important  to  wildlife  because  they 
have  long-lived  features  such  as  cracks,  ledges,  and  deep 
caves. 


Caves  in  igneous  rock  are  more  important  to  cave  dwell- 
ing animals  than  sedimentary  rock  cliffs  because  they  are 
usually  deeper  and  more  resistant  to  weathering. 


Terrestrial  Physical  Features 


593 


classified  according  to  the  type  of  parent  rock,  type 
of  formation  ( Figure  1 ),  as  well  as  by  structural 
characteristics  such  as  diameter  of  opening  and 
horizontal  depth  (Table  4). 

Talus.  Talus  is  the  accumulation  of  rocks  on  or  at 
the  base  of  steep  slopes,  usually  cliffs.  Talus  affords 
protection  for  reproduction  and  hibernation  and 
includes  a  stable  environment.  Talus  slopes  are 
frequently  recorded  as  point  locations,  although 
Maser  et  al.  ( 1979a)  indicate  that  length  and  width 
of  deposits  as  well  as  depth  are  important 
determinants  of  animal  use.  A  slope  can  be  noted  as 
a  point  location,  while  measurements  are  recorded 
as  attributes.  In  addition,  talus  can  be  classified  by 
structural  characteristics  such  as  type  and  size 
classes  of  rocks  (Table  5  and  Figure  2). 

Lava  Flows.  Lava  flows  have  been  classified  into  two 
general  categories,  "old"  and  "new,"  by  Maser  et  al. 
(1979a).  New  flows  are  distinguished  from  old  flows 
by  the  lack  of  soil  and  vegetation  development.  Lava 
flows,  particularly  old  flows,  create  areas  of  rock 
rubble  of  a  size  similar  to  talus  and  would  afford  the 
same  advantages  to  various  wildlife  species  as  talus 
areas. 

Sand  Dunes.  Sand  dunes  form  unique  habitats  for 
many  species  of  vertebrates.  Because  the  habitat 
is  usually  unique  and  dunes  tend  to  be  rather 


localized  and  few  in  number,  many  threatened, 
endangered,  and  sensitive  species  of  vertebrates  as 
well  as  plants  and  invertebrates  occur  on  sand 
dunes.  They  are  usually  mapped  out.  In  addition, 
Maser  et  al.  (1979a)  distinguish  between  active  and 
stabilized  dunes. 

Playas.  Playas  are  shallow  desert  basins  into  which 
water  drains  after  snow  melt  or  rain  storms.  They  do 
not  have  natural  drainageways  and  water  is 
accumulated  seasonally.  These  areas  that  contain 
water  are  important  feeding  and  nesting  habitat  for 
some  water  bird  species.  They  also  form  important 
resting  areas  for  migrating  waterfowl  in  the  spring 
(Maser  et  al.  1979a).  Because  of  salts  left  after 
evaporation,  playas  have  distinctive  vegetation 
communities.  Playas  are  usually  mapped  as 
intermittent  water  areas  on  maps  and  labeled  as 
playas. 


Vegetation-Derived  Habitat  Features 

Vegetation-derived  habitat  features  refer  here  to 
any  habitat  feature  that  is  composed  of  dead  vegeta- 
tion. For  convenience,  these  features  are  frequently 
measured  or  located  during  the  course  of  vegetation 
surveys.  For  example,  density  of  snag  trees  can  usu- 
ally be  measured  during  the  course  of  a  timber  in- 
ventory without  much  additional  effort.  Similarly, 


Table  3.     Classification  of  structural  characteristics  of  cliffs  that  produce  habitat  attributes  that 
can  be  exploited  by  adapted  species  of  wildlife  (from  Maser  et  al.  1979a). 


Structural  Characteristic 

Size  Class  or  Description 

Height  class 

3  m  (10  ft) 
3-10  m  (10-33  ft) 
10-30  m  (33-100  ft) 
30  m  (100  ft) 

Length  class 

30  m  (100  ft) 
30-100  m  (100-330  ft) 
100-500  m  (330-1 ,640  ft) 
500-1 ,500  m  (1 ,640-4,920  ft) 
1,500  m  (4,920  ft) 

Joint  spacing 

0.3  m  (1  ft) 

(called  netted  jointing — fine  network  of  small  cracks) 

0.3-1.5  m  (1.5  ft) 

1 .5-7.5  m  (5-25  ft) 

7.5-30  m  (25-100  ft) 

30  m  (100  ft) 

Joint  width  (opening  size) 

0.15  m  (0.5  ft) 
0.15-0.30  m  (0.5  ft) 
0.3-0.6  m  (1-2  ft) 
0.6-1.5  m  (2.5  ft) 
1 .5  m  (5  ft) 

594 


Terrestrial  Physical  Features 


CO 

CD 

a 
>^ 

i- 

co 

CD 

co 

D 

"O 

c 
CO 

CO 

c 

CO 
LL 

L^2>^ 

t/if  ^ 

fillip® 

Shallow  caves 

Cliff-face  caves 

Lava  tubes 

Mine  shafts 

Mice 
Wood  rats 

Bats 

Bats 

Bats 

Shelter,  reproduction, 
nesting 

Roosting,  reproduction, 
hibernation 

Roosting,  reproduction, 
hibernation 

Roosting,  reproduction, 
hibernation 

Coyotes 
Bobcats 

Mice 
Wood  rats 

Mice 
Wood  rats 

Shelter,  feeding 

Shelter,  reproduction, 
nesting 

Shelter,  reproduction 
nesting 

Figure  1.     Types  of  caves  and  some  of  the  wildlife  species  that  use  them. 


Table  4.     Structural  characteristics  of  caves  that 
produce  habitat  attributes  which  can  be 
exploited  by  adapted  species  of  wildlife  (from 
Maser  et  al.  1979a). 


Structural 
Characteristic 

Size  Class  or  Description 

Opening 
(diameter) 

0.3  m  (1  ft) 
0.3-1  m  (1-3.3  ft) 
1  m  (3.3  ft) 

Depth  of  cave 
(horizontal) 

1  m  (3.3  ft) 

1-3  m  (3.3-10  ft) 

3  m  (10  ft) 

Origin 

Natural 

Man-made 
— abandoned  mine  shaft 
— abandoned  railroad 
tunnel 

Table  5.  Talus  class,  based  on  the  predominant  or 
most  common  rock  size  in  the  talus  field  (from 
Maser  et  al.  1979a). 


Talus  Class 

Rock  Size  (Diameter) 

1 

II 
III 
IV 

0.5  m  (<  1.6  ft) 
0.5-1  m  (1.6-3.3  ft) 
1-2  m  (3.3-6.5  ft) 

2m(<  6.5  ft) 

density  or  thickness  of  litter  is  often  estimated  dur- 
ing range  surveys.  The  features  are  described  here 
because  they  are  sometimes  omitted  from  vegetation 
surveys;  those  conducting  such  surveys  are  primarily 
concerned  with  live  vegetation  and  often  do  not 
understand  the  importance  of  dead  material  to 
wildlife. 


Dead  vegetation  as  a  habitat  feature  is  extremely 
important  to  many  species.  Furthermore,  many  vege- 
tation management  practices  such  as  livestock  graz- 
ing, timber  harvesting,  fire  management,  and 
firewood  cutting  and  collecting  are  done  in  a  man- 
ner that  tends  to  minimize  the  amount  of  dead  vege- 
tation left  in  the  habitat.  The  quantity  and  quality  of 
vegetation-derived  habitat  features  may  thus  change 
drastically  in  a  few  years.  By  contrast,  physical  fea- 
tures such  as  cliffs,  caves,  and  other  geomorphic 
habitats  are  largely  unaffected  by  human  activities 
and  are  primarily  quantified  during  inventories  but 
rarely  monitored.  Measurement  of  dead  vegetation 
should  not  be  overlooked  or  dismissed  in  inventory- 
ing or  monitoring  habitat  features. 


There  are  literally  thousands  of  habitat  features 
formed  by  dead  vegetation  that  are  relevant  to  indi- 
vidual vertebrate  species.  Three  major  categories  are 
discussed  here:  ( 1 )  litter  and  mulch,  (2)  dead  and 
down  woody  material,  and  (3)  snags. 


Terrestrial  Physical  Features 


595 


1 

ZONE    OF 

ZONE   WITH 

LITTLE    PLANT    LIFE 

SOME    PLANT    LIFE 

»| 

~~, 

*                                                      * 

"1 

>    m 
-    < 

<   w 
in    z 

, 

_l 

v      ^S&SSrc-S^fc'-  V>_ 

>   -> 

—    CQ 

<  < 

UJ     ^ 

W^^jmUL '"  *'  JiNHw 

CC 

_ 

ZONE    OF 

ZONE    OF 

LEAST    ANIMAL    USE 

GREATEST    ANIMAL    USE 

Figure  2.     Structure  of  a  talus  field  and  its  use  by 
wildlife. 


Litter  and  Mulch.  Litter  or  mulch  is  the  dead 
material  from  herbaceous  or  woody  plants,  primarily 
grasses,  forbs,  and  shrubs.  Range  conservationists 
make  the  distinction  between  persistent  and  non- 
persistent  litter.  Persistent  litter  is  defined  as 
"undecomposed  organic  debris  on  or  near  the  soil 
surface  with  expected  decomposition  rates 
exceeding  2  years"  (U.S.  Department  of  the  Interior, 
Bureau  of  Land  Management  1985). 

Non-persistent  litter  is  "undecomposed  organic 
debris  on  or  near  the  soil  surface  with  expected 
decomposition  rates  of  2  years  or  less"  (U.S.  Depart- 
ment of  the  Interior,  Bureau  of  Land  Management 
1985).  It  is  composed  primarily  of  herbaceous  mate- 
rial. 


Litter  is  commonly  measured  during  range  vege- 
tation surveys  through  step-point  transects  or  other 
techniques.  It  is  usually  expressed  as  percent  cover. 
In  certain  areas,  particularly  the  Mediterranean,  an- 
nual grasslands  of  California,  and  deciduous  riparian 
habitats,  litter  is  so  abundant  and  important  that 
thickness  is  important  and  is  measured.  The  term 
mulch  is  often  used  for  this  type  of  litter  in  annual 
grasslands. 


Dead  and  Down  Woody  Material.  Dead  and  down 
woody  material  refers  to  fallen  trees  in  various  stages 
of  decay.  The  importance  of  such  material  as  wildlife 
habitat  in  forests  has  received  considerable  attention 
in  recent  years  (Maser  et  al.  1979c;  Maser  and 
Trappe  1984),  because  many  management  practices, 
such  as  chipping  and  slash-burning,  tend  to  reduce 
the  amount  of  such  material. 

Biologists  have  rarely  measured  amounts  of  dead 
and  down  woody  material.  However,  Maser  et  al. 
(1979c)  have  suggested  several  approaches  that 
might  be  useful  and  practical.  The  number  of  dead 
and  down  logs  in  an  area  can  be  sampled  and  ex- 
pressed as  number  per  unit  area  without  any  mea- 
surements of  characteristics  of  individual  logs.  In 
addition,  if  a  more  refined  approach  is  desired,  they 
have  suggested  classifying  of  logs  in  terms  of  their 
stage  of  decomposition,  which  is  relevant  to  their 
value  for  wildlife  species  (Table  6). 

Snags.  A  snag  has  been  defined  for  wildlife  purposes 
as  "any  dead  or  partly  dead  tree  at  least  10.2  cm  (4 
in.)  in  diameter  at  breast  height  (DBH)  and  at  least 
1.8  m  (6  ft)  tall"  (Thomas  et  al.  1979).  This 
definition  is  based  on  the  minimum  diameter  and 
height  of  trees  for  cavity  nesting  birds.  Snags  are 
important  for  many  species  of  wildlife  not  only  for 
nest  or  den  sites  but  for  many  other  biological 
functions  such  as  feeding  and  roosting  (Davis  et  al. 
1983;  Thomas  et  al.  1979). 

Snags  in  an  area  are  generally  counted  or  esti- 
mated by  using  a  standard  sampling  procedure  and 
expressed  as  a  density  (i.e.,  the  number  of  snags  per 
hectare  or  acre).  Thomas  et  al.  (1979)  have  sug- 
gested that  snags  may  be  usefully  classified  as  hard 
or  soft,  depending  on  the  degree  of  decay  and  dete- 
rioration. Hard  and  soft  snags  can  be  classified  in  the 
field  by  striking  them  with  an  axe  (Gale  1973).  If 
the  axe  sinks  in  with  difficulty,  the  snag  is  hard, 
whereas  if  it  penetrates  the  wood  easily,  the  snag  is 
soft.  Hard  snags  usually  have  many  dead  branches 
and  an  intact  top,  whereas  soft  snags  usually  have 
broken  tops  and  few  limbs  (Gale  et  al.  1973). 

Snags  can  be  further  classified  based  on  DBH. 
The  U.S.  Fish  and  Wildlife  Service  uses  many  varia- 
tions of  snag  measurements  in  their  habitat  suitabil- 
ity index  (HSI)  models. 


Animal-Made  and  Man-Made  Features 


Litter  is  an  important  habitat  component  for 
many  species  of  small  mammals,  reptiles,  and  am- 
phibians. It  provides  habitat  for  many  invertebrates 
that  are  food  sources  for  vertebrates,  and  it  provides 
thermal  and  hiding  cover  for  many  of  these  species. 


In  many  cases,  animals  create  distinctive  fea- 
tures in  a  habitat  which  are  then  used  by  the  same 
or  other  species.  Many  of  these  are  small  and  local- 
ized, such  as  the  tree  cavities  created  by  woodpeck- 
ers that  may  subsequently  be  used  by  birds  lacking 


596 


Terrestrial  Physical  Features 


Table  6.     A  classification  of  decomposed  logs  (from  Maser  et  al.  1979c). 


Decomposition 
Class 

Log  Characteristic 

Bark 

Twigs 

3  cm  (1.18  in.) 

Texture 

Shape 

Color  of 
Wood 

Portion  of  Log 
on  Ground 

1 

Intact 

Present 

Intact 

Round 

Original 
color 

Log  elevated 
on  support 
points 

2 

Intact 

Absent 

Intact  to 
partly  soft 

Round 

Original 
color 

Original  log 
elevated  on 
support  points 
but  sagging 
slightly 

3 

Trace 

Absent 

Hard,  large 
pieces 

Round 

Original 
color  to 
faded 

Log  is  sag- 
ging near 
ground 

4 

Absent 

Absent 

Small,  soft, 

blocky 

pieces 

Round 
to  oval 

Light  brown 
to  faded 
brown  or 
yellowish 

All  of  log  on 
ground 

5 

Absent 

Absent 

Soft  and 
powdery 

Oval 

Faded  to 
light  yellow 
or  gray 

All  of  log  on 
ground 

Decaying  logs  furnish  cover  for  many  species  of  wildlife. 


Snags  are  used  by  many  species  of  cavity 
using  animals. 


Terrestrial  Physical  Features 


597 


the  ability  to  excavate  a  cavity.  However,  some  ani- 
mal-made features  are  large  and  important  enough  to 
be  noted  during  habitat  surveys.  Examples  of  these 
types  of  features  are  beaver  dams  and  elk  wallows. 
These  are  usually  noted  as  present  or  absent  without 
any  measurement  or  further  categorization  of  the 
feature. 


The  importance  of  man-made  features  for  wild- 
life has  received  increased  recognition.  Wildlife  man- 
agers have  developed  structures  such  as  guzzlers 
and  nesting  platforms  for  wildlife  for  many  years  and 
their  value  to  wildlife  is  understood.  In  recent  years, 
the  value  of  structures  built  for  other  purposes  has 
received  attention.  These  include  features  such  as 
bridges  and  culverts  used  for  nesting  by  swallows 
and  other  birds,  mine  shafts  used  by  many  bat  spe- 
cies, and  rock  walls  used  by  many  species  of  reptiles 
and  small  mammals.  However,  rarely  has  a  man-made 
feature  been  solely  responsible  for  the  occurrence 
of  a  species  in  an  area. 

On  the  other  hand,  many  man-made  features 
such  as  roads,  powerlines,  and  pipelines,  have  ob- 
vious detrimental  effects  on  wildlife.  These  detrimen- 
tal features  need  to  be  recorded  during  wildlife 
habitat  inventories  as  they  are  also  an  important 
component  of  the  habitat.  Maser  et  al.  (1979b)  de- 
scribes in  detail  important  man-made  features  for  the 
Great  Basin  of  eastern  Oregon.  However,  few  fea- 
tures in  other  regions  have  been  described. 

Both  animal-  and  man-made  features  are  gener- 
ally not  measured  but  simply  recorded  as  being  pres- 
ent in  habitat  inventories.  Since  these  types  of 
features  change  very  slowly  over  time,  monitoring 
them  is  rarely  a  high  priority. 


SPECIAL  HABITAT  FEATURES  AND  THEIR 
USE 

The  above  description  of  terrestrial  physical 
features  categorizes  habitat  components  according  to 
their  origin  in  a  series,  ranging  from  those  caused 
by  geological  processes  to  those  created  by  humans. 
Although  useful  in  subdividing  a  large  set  of  compo- 
nents, in  practice,  the  origin  of  a  habitat  feature  is 
probably  of  little  concern  to  the  animals  that  use 
them.  If  the  environmental  conditions  are  otherwise 
similar,  an  abandoned  mine  shaft  is  equally  accept- 
able as  a  natural  cave  to  a  roosting  bat.  Thus  a  biolo- 
gist conducting  an  inventory  of  such  features  does 
not  need  to  be  overly  concerned  with  categorizing 
origin. 

The  BLM's  IHICS  makes  extensive  use  of  "spe- 
cial habitat  features,"  including  any  of  the  above 


categories  (geomorphic,  edaphic,  soils,  vegetation- 
derived,  animal-made,  man-made).  In  most  cases, 
these  are  habitat  features  that  are  too  small  to  be 
mapped  as  units  with  area,  i.e.,  that  are  mapped  as 
points  on  maps.  In  addition,  these  features  include 
small  patches  of  vegetation  that  are  too  small  to  be 
mapped  as  separate  vegetation  types,  such  as  a  small 
stand  of  trees  around  a  desert  spring  (Table  7). 

Since  the  number  of  special  habitat  features  is 
almost  unlimited,  time  and  money  do  not  permit 
a  complete  inventory  of  them.  Thus,  the  biologist 
must  choose  those  that  are  relevant  to  wildlife  man- 
agement. BLM  biologists  have  found  this  approach 
of  recording  terrestrial  physical  features  to  be  quite 
useful. 


The  special  habitat  features  listed  in  Table  7  for 
IHICS  are  those  that  were  in  use  as  of  July  1985. 
However,  these  are  not  fixed.  Biologists  may  add 
other  features  as  required.  Undoubtedly  features 
listed  in  Table  7  that  are  used  in  other  systems  or 
described  in  other  publications,  will  be  added  as 
necessary.  As  with  all  habitat  components,  the  biolo- 
gist must  determine  which  ones  are  most  useful 
based  on  the  objectives  of  the  inventory  or  monitor- 
ing effort. 


DISCUSSION 

Wildlife  use  physical  features  in  many  different 
ways  with  varying  degrees  of  dependence.  First, 
some  species  in  an  area  increase  their  abundance 
through  the  presence  or  addition  of  a  feature.  For 
example,  many  lizard  populations  will  increase 
through  the  addition  of  a  wooden  fence,  but  the 
populations  do  not  depend  on  fences  for  survival. 
Most  species  respond  to  man-made  features  of  this 
type. 

Second,  some  species  that  are  mobile  will  use  a 
site  if  a  necessary  feature  is  there,  but  the  feature 
may  not  be  limiting  the  population.  For  example, 
many  cliff-nesting  raptors  use  a  limited  number  of 
appropriate  cliff  sites;  presence  of  an  appropriate  site 
may  determine  local  use,  but  may  not  limit  popula- 
tion density  or  productivity  in  the  region.  If  the  site 
is  destroyed,  the  birds  simply  use  alternative  nest 
sites  with  no  detrimental  effects  on  the  population. 
Unfortunately,  many  man-made  structures,  including 
some  that  have  been  constructed  deliberately  for 
wildlife,  fall  into  this  category.  For  example,  many 
gallinaceous  guzzlers  have  been  constructed  for  quail 
and  chukars  in  desert  areas  that  were  already  fairly 
well-watered.  Although  quail  will  concentrate  around 
such  waters  during  dry  periods  and  drink  from  them, 
there  is  no  evidence  that  they  increase  population 
densities  in  such  areas. 


598 


Terrestrial  Physical  Features 


Table  7.     Special  habitat  features  used  in  the  Bureau  of  Land  Management's  Integrated  Habitat  Inventory  and 
Classification  System  (IHICS). 


Natural  Features 

Man-Made  Features 

Avalanche-slide  area 

Bridge 

Cave 

Fence 

Cave  ice 

Underpass 

Cave  lava 

Salting  area 

Cliff 

Goose  nesting  platforms 

Cone,  volcanic 

Artificial  nesting  boxes 

Dike,  volcanic 

Small  seedlings 

Dune,  sand 

Buffer  strip 

Insect  mounds 

Building 

Overhang 

Bird  ramp 

Salting  area 

Berm 

Seep 

Culvert 

Cold  spring 

Dock 

Sinkhole 

Dredged  area 

Snag  or  group  of  snags 

Exclosure,  study 

Talus  slope 

Fish  migration  barrier 

Talus  field 

Gauging  station,  water 

Elk  wallow 

Mining  activity 

Waterfall 

Poles,  electrical/telephone 

Wasteland 

Perches 

Island  (too  small  to  be  typed  as  habitat  site) 

Road 

Log  jam 

Trail 

Downed  timber 

Stream  improvement  structure 

Bluff 

Railroad 

Beaver  dam 

Stream  crossing 

Muskrat  house 

Shelter  (overnight) 

Cataracts  (stream) 

Recreation  area 

Barren  lands 

Feeding  stations 

Hot  springs 

Fire  break 

Blowouts 

Seismographic  trail 

Mudflow 

Oil  sump  pit 

Temporary  pond 

Windmill 

Small  natural  pond 

Irrigation  diversion  or  ditch 

Small  grp  trees  or  shrubs 

Water  gap 

Small  grp  trees  or  riparian 

Stock  water  pond 

Dry  meadow  (not  typed  as  vegetation  type) 

Corral  or  loading  chute 

Dry  wash 

Artificial  wildlife  water 

Stream  bank  gravel 

Domestic  water  source 

Raptor  nest  tree 

Artesian  well 

Buffalo  wallow 

Oil  well 

Boulder  or  rock  outcrop 

Gas  well 

Rodent  colony 

Pipeline 

Beaver  lodge 

Material  site 

Otter  slide 

Airfield 

Perennial  snowfield 

Breakwater 

Rocky  crags 

Dam 

Alpine  fell  field 

Wilderness  camp 

Pingo 

Winter  trail 

Gravel  bar 

Burn 

Sand  bar 

Mine  shaft 

Ocean  cliff 

Mine  tunnel 

Stack 

Stock  water  tank 

Glacier 

Disposal  site  (active) 

Spit 

Disposal  site  (inactive) 

Barren 

Wrecked  ships 

Burn 

Abandoned  homesites 

Booming,  dancing,  or  strutting  ground 

Relay  stations 

Wet  meadow  (not  typed  as  vegetation  type) 

Pump  jack 

Brushy  openings  (too  small  to  be  typed  as  habitat  site) 

Brush  pile/rows 

Terrestrial  Physical  Features 


599 


Table  7.     Special  habitat  features  used  in  the  Bureau  of  Land  Management's  Integrated  Habitat  Inventory  and 
Classification  System  (IHICS)  (concluded). 


Natural  Features 

Man-Made  Features 

Snake  den 

Roost 

Kipuka 

Stream  (too  small  to  be  mapped  as  habitat  site) 

Mineral  spring 

Rock  formation  raptor  nests 

Raptor  cliff  nest-stick 

Raptor  cliff  nest-scrape  (no  nest  material) 

Raptor  cliff  perch 

Raptor  nest  on  pinnacle  (rock  or  earthen) 

Raptor  nest  (ground  or  hillside) 

Raptor  nest  (shrub) 

Raptor  nest  (ground  burrow) 

Rookery 

Oxidation  ponds  (oxidation  ponds  and  evaporation 

playa) 
Spring,  wildlife  (developed  &  useable  by  wildlife  only) 
Spring,  livestock  (developed  &  useable  by  livestock 

only) 
Spring,  wildlife/livestock  (developed  &  useable  by 

wildlife  &  livestock) 
Exclosure  (vegetation  protection  for  wildlife) 
Vegetation  Manipulation  (for  wildlife;  too  small  to  be 

typed  as  a  habitat  site 
Artificial  catchment 
Reservoir 

Gallinaceous  guzzler 
Rain  gauge 
Raptor  nest  (artificial) 
Raptor  nest  (earth  cut) 

A  third  category  involves  species  that  are  totally 
dependent  on  a  feature  within  an  area.  Examples  of 
this  relationship  are  species  of  salamanders  found 
only  at  limestone  seeps  and  the  many  desert  fishes 
found  at  isolated  desert  springs.  Most  of  these  spe- 
cies are  sedentary,  such  as  amphibians  and  reptiles. 
Many  species  are  remnant  or  relict  species  from 
earlier  ages  when  conditions  favorable  to  their  exis- 
tence were  more  widespread.  Desert  pupfish  (Cypri- 
nodon  sp. )  are  good  examples  of  such  a  situation. 

Biologists  should  consider  these  three  types  of 
relationships  when  planning  and  setting  priorities  to 
inventory  or  measure  physical  features.  Priority 
should  normally  be  given  to  inventory  features  in 
the  third  category  and  subsequently  verify  species 
presence  around  them.  The  second  category  requires 
that  the  biologist  determine  the  extent  to  which  the 
features  are  limiting  and  at  what  level  (site,  regional, 
statewide,  etc.).  Returning  to  the  raptor  example, 
the  biologist  must  determine  if  suitable  nest  sites  are 
a  factor  limiting  the  population  at  the  state  or  re- 
gional level.  If  it  is  not,  then  attention  to  inventory 
and  management  of  such  sites  is  not  warranted  un- 


less the  management  objective  is  to  attract  birds 
to  the  local  site.  Special  features  of  the  first  type  are 
not  normally  a  high  priority  for  inventory  or  man- 
agement, unless  they  increase  the  abundance  of  high 
priority  species,  such  as  threatened,  endangered,  or 
game  species. 


SUMMARY 

Physical  features  are  important  habitat  compo- 
nents for  many  wildlife  species.  These  include  land- 
forms,  geomorphic  and  edaphic  features,  soils, 
vegetation-derived  features,  animal-made  features, 
and  man-made  features.  Formal  systems  for  invento- 
rying and  describing  such  features  have  not  been 
developed  to  the  same  degree  as  systems  for  survey- 
ing vegetation.  In  many  cases,  simply  noting  or  map- 
ping their  presence  is  adequate.  For  some  special 
features,  measurements  may  be  required;  for  others, 
classifying  them  by  size  or  other  characteristics  is 
useful.  The  biologist  must  determine  which  physical 
features  to  inventory,  describe,  or  measure  depend- 
ing on  the  objectives  of  the  project. 


600 


Terrestrial  Physical  Features 


LITERATURE  CITED 


BAILEY,  J. A.  1984.  Principles  of  wildlife  management.  John 
Wiley  &  Sons.  New  York,  NY  373pp. 

DAVIS,  J.W.,  G.A.  Goodwin,  and  R.A.  Ockenfels,  tech. 

coords.  1983  Snag  habitat  management:  Proceedings 
at  the  symposium.  U.S.  Dep.  Agric,  For.  Serv.,  Gen. 
Tech.  Rep.  RM-99.  226pp. 

GALE,  R.M.  1973-  Snags,  chainsaws  and  wildlife:  One  as- 
pect of  habitat  management.  Page's  97-1 12  in 
Yoakum,  J.  ed.  Cal-Neva  Wildlife  1973,  Transactions. 
Reno 

,  W.F.  KELLEY,  and  J.A.  LORENZANA.  1973.  Snag 

management:  Coordination  guidelines  for  wildlife 
habitat.  U.S.  Dep.  Agric,  For.  Serv.,  California  Region. 
13pp. 

HAMILTON,  C.KL,  R.M.  KERR,  and  LA.  PETERSON.  1983. 
IHICS,  the  Bureau  of  Land  Management's  habitat  in- 
ventory system.  Presented  at  National  Workshop 
on  Computer  Uses  in  Fish  and  Wildlife  Programs — a 
State  of  the  Art  Review.  Virginia  Polytechnic  Institute 
and  State  University,  Blacksburg,  VA.  1 3pp 

HAYS,  R.L.,  C.  SUMMERS,  and  W.  SEITZ.  1981.  Estimating 
wildlife  habitat  variables.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.,  FWS/OBS-81/47.  111pp. 

MASER,  C,  R.G.  ANDERSON,  K.  CROMACK,  Jr.,  J.T.  WIL- 
LIAMS, and  RE.  MARTIN.  1979c.  Dead  and  down 
woody  material.  Pages  78—95  in  Thomas,  J.W.  ed. 
Wildlife  Habitats  in  Managed  Forests — The  Blue 
Mountains  of  Oregon  and  Washington.  U.S.  Dep. 
Agric,  For.  Serv.,  Agric.  Handbook  553   512pp. 

,  J.M.  GEIST,  DM.  CONCANNON,  R.  ANDERSON, 

and  B.  LOVELL.  1979a.  Wildlife  habitats  in  managed 
rangelands — The  Great  Basin  of  Southeastern  Ore- 


gon— Geomorphic  and  edaphic  habitats.  U.S.  Dep. 
Agric,  For.  Serv.,  Gen.  Tech.  Rep.  PNW-99.  84pp. 

— ,  J.W.  THOMAS,  ID.  LUMAN,  and  R.  ANDERSON. 
1979b.  Wildlife  habitats  in  managed  rangelands — The 
Great  Basin  of  southeastern  Oregon — man  made  habi- 
tats. U.S.  Dep.  Agric,  For.  Serv.,  Gen.  Tech.  Rep.  PNW- 
86.  84pp. 

—  and  J.M.  TRAPPE,  eds.  1984.  The  seen  and  unseen 


world  of  the  fallen  tree.  U.S.  Dep.  Agric,  For.  Serv., 
Gen.  Tech.  Rep.  PNW-164.  56pp. 

ODUM,  E.P.  1959.  Fundamentals  of  ecology.  2nd  Edition, 
W.B.  Saunders  Co.,  Philadelphia,  PA.  546pp. 

PETERSON,  F.F.  1981.  Landforms  of  the  Basin  and  Range 
Province.  University  of  Nevada,  Tech.  Bull.  28,  Reno. 
52pp. 

ROBINSON,  W.L  and  E.G.  BOLEN.  1984.  Wildlife  ecology 
and  management.  MacMillan  Publishing  Co.,  New 
York,  NY.  478pp. 

SHORT,  H.L.  and  K.P.  BURNHAM.  1982.  Techniques  for 
structuring  wildlife  guilds  to  evaluate  impacts  on 
wildlife  communities.  U.S.  Dep.  Inter.,  Fish  and  Wildl. 
Serv.,  Spec.  Sci.  Rep. — Wildlife  244.  34pp. 

THOMAS,  J.W.,  R.G.  ANDERSON,  C  MASER,  and  E.  BULL. 
1979.  Snags.  Pages  60-77  in  Thomas,  J.W.  ed.  Wildlife 
Habitats  in  Managed  Forests — the  Blue  Mountains  of 
Oregon  and  Washington.  U.S.  Dep.  Agric,  For.  Serv., 
Agric.  Handbook  553-  512pp. 

U.S.  DEPARTMENT  OF  THE  INTERIOR,  BUREAU  OF 

LAND  MANAGEMENT.  1982.  Integrated  habitat  inven- 
tory and  classification  system.  U.S.  Dep.  Inter.,  Bur. 
Land  Manage.,  Manual  6602,  unnumbered. 

U.S.  DEPARTMENT  OF  THE  INTERIOR,  BUREAU  OF 

LAND  MANAGEMENT.  1985.  Rangeland  monitoring— 
analysis,  interpretation,  and  evaluation.  U.S.  Dep. 
Inter.,  Bur.  Land  Manage.,  Tech.  Reference 
4400-7.  69pp. 


Terrestrial  Physical  Features 


601 


28 

AQUATIC 

PHYSICAL 

FEATURES 

Paul  Cuplin 


U.S.  Bureau  of  Land  Management 
Service  Center 
Denver,  CO  80225 


Editor's  Note:  Although  biologists  have  done  little 
to  develop  standardized  methods  or  measurements 
for  terrestrial  physical  features,  there  has  been  more 
work  with  aquatic  systems.  Fisheries  biologists  and 
limnologists  have  long  recognized  the  importance 
of  physical  features  of  streams  such  as  streamflow, 
stream  gradient,  and  water  temperature.  Whereas 
the  terrestrial  biologist  must,  for  the  most  part, 
develop  ad  hoc  systems  for  quantifying  physical 
features,  the  aquatic  biologist  has  some  reasonably 
well -developed  measurements  and  classification 
systems  for  aquatic  systems.  This  chapter  describes 
physical  features  that  are  important  determinants 
or  predictors  of  the  quality  of  the  aquatic  habitat 
for  fish  and  other  vertebrates,  and  systems  for  meas- 
uring or  classifying  such  features. 


INTRODUCTION 

Aquatic  biologists  have  collected  many  types  of 
data  to  describe  aquatic  habitat — sometimes  with 
uncertainty  as  to  their  value.  Accuracy,  precision, 
and  bias  were  not  generally  considered;  usually  the 
ease  of  data  collection  was  most  important.  More 
attention  is  now  being  directed  toward  repeatable 
results,  recognizing  fewer  variables  that  can  be  meas- 
ured with  higher  precision  and  accuracy.  Examples 
of  these  variables  are  water  temperature,  stream 
width,  stream  depth,  lake  depth,  surface  acreage, 
stream  gradient,  and  turbidity.  Examples  of  variables 
that  cannot  be  measured  with  high  precision  are 
pool  quality,  stream  channel  stability,  streambank  sta- 
bility, and  the  amount  of  various-sized  particles  in 
the  streambed. 


HABITAT  FEATURES  AND  MEASUREMENT 
TECHNIQUES 

The  variables  that  are  most  important  for  de- 
scribing stream  and  lake  habitat  are  discussed  in  this 
chapter,  regardless  of  the  precision  with  which  they 
can  be  measured. 

Precision  and  accuracy  of  measurements  of 
aquatic  physical  features  vary  considerably.  Platts 
(1981)  rates  the  precision  that  can  be  applied  to 
variables  related  to  stream  habitat  (Table  1 ).  Hamil- 
ton and  Bergersen  (1984)  discuss  precision  and  accu- 
racy of  many  of  these  variables. 

Streamflow  Pattern 

Characterization  of  streamflow  pattern  as  peren- 
nial, intermittent,  or  ephemeral  streamflow  can  be 
determined  by  observing  the  stream  and  measuring 
streamflow  during  high  and  low  precipitation  pe- 
riods. It  can  also  be  determined  from  U.S.  Geological 


Aquatic  Physical  Features 


603 


Survey  (USGS)  topographic  maps  as  different  sym- 
bols are  used  tor  perennial,  intermittent,  and  ephem- 
eral stream  reaches. 

Streamilow 

Streamflow  is  measured  to  determine  stream 
size  for  baseline  information.  Streamflow  volume  re- 
lates to  the  amount  of  living  space  available  for  fish 
and  aquatic  life. 

Historic  records  of  streamflow  (mean,  record 
high  and  low,  etc.)  can  be  obtained  from  the  U.S. 
Geological  Survey. 

Streamflow  can  be  measured  by  using  the  for- 
mula R  =  WDCV  (Robins  and  Crawford  1954): 

where   R  =  flow  in  cubic  feet  per  second  ( cfs ) 
W  =  average  width  of  stream  in  feet 
D  =  average  depth  of  stream  in  feet 
C  =  a  constant  for  stream  bottom 
0.8  -  rough  bottom 
0.9  -  smooth  bottom 
V  =  velocity  in  feet/second 

Two  people  are  generally  required  to  make 
these  measurements  at  each  sampling  station.  Ap- 
proximately half  an  hour  is  required  to  measure 
stream  depths,  widths,  and  velocities  which  are 
needed  to  calculate  streamflow  in  cubic  feet  per 
second.  Measurement  of  each  of  these  is  described 
separately.  The  minimum  equipment  required  is  a 
watch  with  second  hand,  yardstick  or  meter  stick, 
float  stick,  and  100-foot  measuring  tape. 


Installing  a  streamflow  measuring  gauge. 


Table  1.     The  average  expected  repeatability,  precision,  and  confidence  intervals  of  water  column  and 

streambank  measurement  means  from  six  selected  streams  in  Idaho  and  Nevada.  Confidence  intervals  are 
at  the  95%  level  and  expressed  as  percent  of  the  mean  (from  Platts  1981). 


Confidence 

Repeatability 

Measurement 

Interval 

Precision 

over  Time 

Water  Column 

Stream  width 

5.4 

© 

© 

Stream  depth 

8.2 

© 

©-• 

Streamside  water  depth 

16.6 

© 

©-© 

Pool  (percent) 

10.3 

© 

O 

Pool  (quality) 

8.0 

© 

o-© 

Riffle  (percent) 

12.5 

© 

o 

Sun  arc  angle 

1.1 

• 

© 

Bank  to  bank  width 

Very  wide 

o 

o 

High  water  stream  width 

Very  wide 

O 

o 

Streambank 

Soil  alteration 

12.3 

© 

©-© 

Vegetative  stability 

3.1 

• 

©-© 

Undercut 

18.5 

© 

© 

Angle 

4.4 

• 

© 

Rock  content 

Very  wide 

o 

o 

Excellent 


Good         Q=  Fair        Q=  poor 


604 


Aquatic  Physical  Features 


Stream  Velocity.  Stream  velocity  can  be  measured 
in  numerous  ways.  The  simplest  method,  and  the 
one  commonly  used  in  conjunction  with  streamflow 
measurement,  is  the  float  method. 

A  straight  section  of  stream,  100  feet  long,  is 
measured.  A  6-inch  wooden  stick  or  table  tennis  ball 
is  used  as  a  float.  Some  people  recommend  using  an 
object  with  the  specific  gravity  of  water  (1.0),  such 
as  an  orange  or  water-filled  prescription  bottle. 

Streamflow  velocity  is  more  precisely  measured 
by  using  a  flow  meter  to  determine  feet-per-second 
flow.  A  minimum  of  20  measurements  are  taken 
across  a  stream  transect  and  averaged. 

For  a  more  detailed  discussion  of  velocity  and 
its  measurement,  see  Hamilton  and  Bergersen 
(1984). 

Stream  Width.  Stream  width  is  the  average  distance 
between  water's  edges  transversely  across  the 
stream.  On  smaller  streams,  measurements  can  be 
made  with  a  tape  or  estimated.  Several 
measurements  should  be  taken  and  the  average 
calculated  for  a  given  stream  reach. 

For  larger  streams,  an  optical  range  finder  can 
be  used  in  the  field  and  a  map  measurer  can  be  use- 
ful in  making  these  measurements  in  the  office.  Hays 
et  al.  (1981)  describe  use  of  these  techniques. 

Stream  Depth.  Water  depth  is  the  vertical  distance 
from  the  surface  to  the  bottom.  Water  depth 
measurements  may  be  required  for  either  streams  or 
lakes;  however,  the  techniques  used  are  often 
different. 

In  addition  to  its  use  in  streamflow  calculations, 
measuring  the  depth  of  a  body  of  water  provides  a 
check  on  siltation  rate  and  a  basis  for  classifying  and 
typing  lakes  and  reservoirs. 

For  streams,  water  depth  can  be  measured  with 
a  calibrated  rod.  Several  depth  recordings  are  used 
to  calculate  average  stream  depth. 

Lakes,  ponds,  and  reservoirs  and  larger  streams 
or  rivers  can  be  sounded  with  a  weighted  line  or 
recorded  with  depth-sounding  equipment  from  a 
boat.  A  sounding  line,  marked  in  5-foot  increments, 
will  provide  adequate  depth  information  for  cross 
sections  and  verification  and  correction  of  topo- 
graphic maps  for  small  lakes  and  reservoirs. 

A  depth  recorder,  with  the  transducer  mounted 
on  the  boat  keel,  provides  an  accurate  timesaving 
depth-measuring  tool.  The  recorder  should  be  cali- 
brated in  feet  or  meters  and  equipped  with  a  record- 
ing tape.  A  minimum  of  four  depth-sounding  cross 


sections  are  located  on  the  topographic  map — a  tran- 
sect from  the  backwater  to  the  outlet  in  the  center 
of  the  lake  and  three  or  four  transects  perpendicular 
to  the  backwater-outlet  transect.  Irregularly  shaped 
reservoirs  and  lakes  will  require  additional  transect 
soundings.  The  depth-sounding  tapes  keyed  to  the 
topographic  map  can  be  processed  by  an  engineer  to 
develop  a  new  or  current  topographic  map. 

Hamilton  and  Bergersen  (1984)  provide  a  de- 
tailed description  of  techniques  for  measuring  water 
depth. 


Instream  Flow  Needs 

The  amount  of  water  needed  to  maintain  stream 
habitat  on  a  year-round  basis  is  termed  "instream 
flow  needs."  Instream  flow  needs  are  assessed  to 
identify  streamflow  requirements  for  resident  aquatic 
life  and  identify  the  amount  of  habitat  loss  that 
would  occur  at  reduced  rates  of  streamflow. 

Instream  flow  needs  can  be  determined  by  sev- 
eral methods  as  summarized  by  Cuplin  et  al.  (1979). 
Where  detailed,  precise,  legally  defensible  informa- 
tion is  required,  the  multiple  transect — incremental 
flow  method — is  recommended.  This  method  is  de- 
scribed in  detail  in  Bovee  and  Milhous  (1978). 

For  rapid  assessment,  a  single  transect  method  is 
available  (Cuplin  et  al.  1979).  This  technique  re- 
quires less  time  for  field  work  and  data  analysis  but 
it  produces  less  precise  results  than  the  incremental 
flow  methods. 

Computer  programs  have  been  developed  for 
analyzing  stream  channel  cross-section  survey  data 
collected  in  conjunction  with  instream  flow  assess- 
ments (Parsons  and  Hudson  1985).  A  single  transect 
method  can  be  used  for  rapid  assessment  using  ADP 
programs.  A  computer  analysis  program  is  available 
for  U.S.  Bureau  of  Land  Management  offices  through 
the  Denver  Service  Center,  Division  of  Resources 
(Parsons  and  Hudson  1985). 


Streambank  and  Shoreline  Stability 

Streambank  and  shoreline  stability  refers  to  the 
resistance  of  bank  materials  to  erosion  from  flowing 
water.  Streambank  and  shoreline  stability  measure- 
ments are  used  to  classify  stream/riparian  habitat 
quality  and  to  monitor  changes  in  stability  caused  by 
the  land  use. 

Streambank  stability  can  be  classified  into  four 
condition  classes  from  excellent  to  poor  (Table  2). 
This  classification  is  designed  for  addressing  stability 
as  a  function  of  livestock  grazing.  Other  land-use 
impacts  can  also  be  assessed  by  ocular  surveys.  All 


Aquatic  Physical  Features 


605 


Table  2.     Condition  classes  for  streambanks  and  shorelines  (U.S.  Department  of  the  Interior,  Bureau  of  Land 
Management  1979). 


Category 

Rating 

Description 

1 

• 

No  negligible  use/damage;  well-rooted  vegetation, 
primarily  grasses,  sedges,  and  forbs;  sod  intact; 
very  little  if  any  erosion  from  vegetation  areas;  less 
than  5%  bare  soil  showing  along  shoreline. 

II 

9 

Some  use/damage;  vegetation  generally  well-rooted; 
sod  mostly  intact;  soil  showing  in  places,  6  to  1 5% 
bare  soil  showing  overall;  some  surface  erosion 
evident. 

III 

9 

Use  or  damage  close  to  sod;  vegetation  shallowly 
rooted;  moderate  surface  erosion,  16  to  25%  bare 
soil  showing  overall. 

IV 

O 

Heavy  to  severe  use/damage;  vegetation  generally 
grazed  down  to  the  soil;  considerable  soil  showing, 
over  25%  with  serious  sod  damage;  active  surface 
erosion  is  a  serious  problem. 

I  =  Excellent         Q  =  Good         Q  =  Fair        Q  =  Poor 


estimates  are  subjective  and  thus  the  accuracy  of  this 
method  depends  on  the  ability  and  experience  of 
the  surveyor. 

Stream  Channel  Stability 

Stream  channel  stability  is  rated  as  part  of  the 
overall  stream  habitat  condition.  Existing  natural 
conditions  and  land  use  govern  stream  channel 
stability. 

Stream  channel  stability  is  evaluated  by  ocular 
classification.  One  method  that  can  be  used  classifies 
channel  stability  in  four  categories — excellent,  good, 
fair,  and  poor  (Table  3).  This  method  has  limited 
application  to  desert  streams  that  are  generally  ero- 
sional  in  nature. 

Pfankuch  ( 1975)  and  Duff  and  Cooper  ( 1978) 
describe  a  very  detailed  channel  stability  evaluation 
procedure.  This  method  requires  visually  categoriz- 
ing numerous  bank  and  bottom  attributes  individ- 
ually and  then  assigning  them  a  point  value  (Figure 
1 ).  An  overall  score  is  determined  by  summing 
scores  for  individual  attributes.  This  system  works 
best  for  timbered  mountainous  streams;  I  have  had 
poor  results  trying  to  apply  it  to  arid  land  sedimen- 
tary streams.  For  the  latter,  the  simpler  visual  system 
(Table  3)  works  better. 

Stream  Gradient 

Stream  gradient  is  a  general  slope  or  rate  of 
change  in  vertical  elevation  per  unit  of  horizontal 
distance  of  water  surface  of  a  flowing  stream.  Stream 


gradient  can  be  measured  in  the  field  using  a  clino- 
meter or  Abney  level  or  calculated  in  the  office  from 
measurements  on  topographic  maps. 

In  the  field,  gradient  is  estimated  by  measuring 
the  angle  from  the  horizontal  on  two  selected  points 
on  the  ground  with  a  clinometer  or  Abney  level. 
Use  a  rod  with  a  mark  at  the  same  height  as  the 
measuring  crew  member's  eye  level  to  measure  an- 
gle from  the  horizontal.  One  crew  member  holds  the 
marked  rod  at  one  of  the  sample  points  and  the 
other  measures  the  angle  of  the  target  from  the  hori- 
zontal using  the  clinometer.  One  person  can  com- 
plete the  required  measurements,  but  it  is  more 
efficient  with  two.  Hays  et  al.  (1981:87-90)  describe 
this  technique  in  detail. 

In  the  office,  gradient  can  be  calculated  off  USGS 
topographic  maps  using  the  following  formula: 

Em  (map  scale  fraction) 
G  ~  d 

where  G  =  gradient  (in  percent) 

E  =  difference  (in  feet)  in  elevation  on 

ground  between  two  points 
m  =   the  map  scale,  decimal  fraction 
d  =  distance  (in  inches)  on  the  map 

between  the  two  points 

Hays  et  al.  (1981:85-87)  describe  this  technique  in 
detail. 

If  digital  elevation  model  tapes  are  available  and 
a  stream  course  has  been  digitized,  then  gradient 
can  be  calculated  automatically  for  any  stream  reach 
using  a  geographic  information  system  (GIS). 


606 


Aquatic  Physical  Features 


Table  3.     Stream  channel  stability  (from  Cuplin  1978). 


Category 

Rating 

Description 

1 

• 

None  or  negligible  lateral  channel  movement  and 
bank  erosion  (cutting)  (5%  or  less),  scour,  or 
changing  channels. 

II 

Q 

Some  lateral  channel  movement  and  bank  erosion 
(5  to  10%);  minor  channel  scour  or  changing 
channel  within  streambed. 

III 

© 

Frequent  lateral  channel  movement  (10  to  15%); 
moderate  channel  scour  or  channel  change  within 
streambed. 

IV 

O 

More  than  20%  lateral  channel  movement  and  bank 
cutting;  changing  channels  and  severe  scour 
evident;  source  of  extreme  sedimentation. 

I  =  Excellent        Q  =  Good        Q  =  Fair       Q  =  Poor 


Pool  Quality 

Pools  are  the  deeper,  slow-moving  sections  of 
streams  which  provide  resting  areas  for  fish.  A  vari- 
ety of  pool  classes  will  provide  the  greatest  diversity 
and  fulfill  the  requirements  for  various  stages  of  fish 
life. 

Pool  quality  is  an  estimate  of  the  pool's  capabil- 
ity to  provide  habitat  for  the  various  life  stages  of 
fishes.  A  method  of  rating  pools  has  been  developed 
that  considers  pool  size,  depth,  and  cover  ratings 
(Table  4).  By  using  these  measurements,  pools  are 
categorized  into  five  pool  classes  of  which  one  is  the 
best.  Nevertheless,  pools  with  low  ratings  may  have 
some  value.  For  example,  a  pool  that  receives  a  low 
rating  of  1  for  depth  and  1  for  size  may  still  provide 
essential  habitat  for  the  survival  of  young  fish.  Hamil- 
ton and  Bergersen  (1984)  describe  pool  rating  sys- 
tems in  detail. 


Riffles 

Riffles  are  the  steeper  gradient,  higher  velocity 
stream  sections.  They  generally  consist  of  coarse 
gravel,  rubble,  or  larger  substrate.  Riffles  are  used  for 
spawning  and  also  produce  macroinvertebrates — a 
primary  source  of  fish  food.  Headwater  streams  are 
often  primarily  composed  of  riffles  with  few  pools. 

Riffles  are  not  usually  rated  the  way  pools  are. 
However,  riffle  length  and  width  can  be  measured 
with  a  tape  or  ocular  estimate.  Pool-to-riffle  ratio  is 
often  used  as  an  indicator  of  stream  habitat 
condition. 


Streambed 

Streambed  material  can  be  classified  by  particle 
size  according  to  Table  5.  A  stream  transect  or  cross 


Meandering,  low-gradient  stream. 


Steeper  gradient,  as  indicated  by  riflles. 


Aquatic  Physical  Features 


607 


Table  4.     Pool  quality  rating  system  (after  Hamilton  and  Bergersen  1984). 


Category 

Rating1 

Description 

Pool  size 

3 
2 
1 

Pool  larger  or  wider  than  average  width  of  stream 
Pool  as  wide  or  long  as  average  width  of  stream 
Pool  shorter  or  narrower  than  average  width  of  stream 

Depth 

First  through  Fourth 
Order  streams 

Fifth  Order  streams 
or  larger 

3 
2 

1 

3 
2 

1 

Over  3  feet  deep 
2-3  feet  deep 
Under  2  feet  deep 

Depth  20%  of  average  stream  width  or  greater 
Depth  1 0-20%  of  average  stream  width 
Depth  less  than  10%  of  average  stream  width 

Cover 

3 
2 
1 

Abundant  cover 
Partial  cover 
Exposed 

1Pool  size,  depth,  and  cover  ratings  are  summed  to  provide  an  overall  pool  rating  according  to  the  following  scheme: 


Total  Ratings 

Pool  Class 

8-9 

1 

7 

2 

5-62 

3 

4-5 

4 

3 

5 

2Sum  of  5  must  include  2  for  depth  and  2  for  cover. 


Table  5.     Classification  of  streambed  material  (from 
Duff  and  Cooper  1978). 


An  electronic  depth  finder. 


Material 

Particle  Size 

Bedrock 

Exposed  solid  rock 

Boulder 

+  12  in.  (305  mm) 

Large  rubble 

6-12  in.  (152-305  mm) 

Small  rubble 

3-6  in.  (76-152  mm) 

Coarse  gravel 

1-3  in.  (25-76  mm) 

Fine  gravel 

0.1-1.0  in.  (2.5-25  mm) 

Sand 

(0.074-2.5  mm) 

Clay 

(0.074  mm) 

Organic  muck 
(sapropel) 

Decomposed  organic 
matter 

Organic  debris  (detritus) 

Undecomposed  organic 
matter 

608 


Aquatic  Physical  Features 


oo  c  F 

3     O    CO 


C     <=  c 

flj    (1)  0) 

5-  E  - 

g^-6  E 

"  «  p 


03  CD 

<»  TT 

-C  >> 

o  c 

*-"  co 

(D  c 

5  E 

CD  O 

■O  "O 

O  CD 


TO 


"  _5 

cd  o  ° 
5  O  m 
0)  Q.]£ 
"*~    en   CO 

en  2 

3    MT3 

q.  y  " 

-tr  E  to 
</>  .  3 
o  § 

-f| 

o  go  o  <" 
V  ca  -d  E 


a) 


■9  ° 
cd  to 
>  ■- 
OQ 


CT° 
CD    E 

73  E 


c 

CD 

E    « 
0)  "> 

£  o 

8^? 


C    Q_  cj 
03    03    o 


£  ™  E 

en    <D 


CD  ^ 

CD 

o* 

3  u 

>.  C 

o 

a -SC- 
SI'S 

w    C 

CM 

03    03 

V 

CD  -C 

tfl 

LL  "D 

>.  O 

to  o 

P.    0) 


3  ■«  § 

§"-  5- 

—  ■*  5 

to  CM  <5 

o  ,_  € 

F    B  o 

4?  >  > 

<    O  O 


y  e 

tr  o_ 

—    03  O 

o  g.  aj 


Q.   >•  CO 


LU    Q.< 


E  <D 

O  to 

~°  8 

Q.  CO 


8    S| 
-I    O 


CD    >■ 

"O  ^ 

>  03 

CD    CD 
O)  w 

c  cr. 

■5    CD 

o  E 

03  tr 
Q.  o 

°    CO 

Z    03 


2  .05 

5  2 

13  <? 

T3 

CD  CD 

-£  23 

03  03 
2  CO 


*n 

^ 

CO    GO    cr 

o 

O 

CD       -    CD 

Q.  c   oo 

X 

>fl)    U) 

) 

—    CD    J- 

CO 

**- 

m 

«  B3a 

o 

c 
o 

C     i    CO 

c   g  -Q 

b 

2 
to 

CO 
13 

Pere 
Yello 
may 

^ 

CO 

>s 

t£    CD 


cd  x: 

Sj  to  a> 

O"  o  sz 

CD    Q.  a) 

-    to  c 

CD    5  § 

03  2  -o 

®    CD  Jr, 

•D    p  2 

■5    °  § 


CD    CD 


>  £  "O    oo 

i  s « « 

o  o  5  E 


CD    5    03    O 

T)  or    3 


I      T]     O     O 

o  c  M  « 

IT)    03    03  "O 


CO  oo 
CD  T3 
Q.  o 

*-    O 


if    OQ 

™  2  5 


Eo0£ 


gS  03 

o  "5 

6  2 

CM  "O 


CO 

=1  o 

cr  d 

CD  >= 

.£=  oo 


5  E 

Sj  o 

o  JX   v> 

E  £  o 

^  03    O 

oo  X3    Q. 

I?° 

CD  00    O) 

=  =3    C 

CD  CO  = 

"D  O  ^z 


i2    2  CD 

3  "E  T3 

o  g>  > 

„  o  <" 

c  _  ci) 

CO    CO  c 

^Er 

^   „  D) 

g>  o  o 

co  tr  go 


5  CO  >= 

—  oo  CO 

M  cd  ■Q 

CD  CO  J 

X)  O  c 

CD  °  g 

?!?  § 

CD  CD  00 

73  £  T3 

o  CO  u 

5  O)  to 


§.9 

CD    " 

08 .1 

00  X3 
CD    O 

E  5 


E 

oo  M 

00  »- 

CO  CD 


O    CD 


£6 

CD  CM 
O)  jo 
05  « 
o    CD 


D 

n 

o 

n 

u 

10 

CO 

u 

JJ 

- 

c  CO 


CO 


T3  -Q 

O    CO 

^55 


■=  o 

to    C3) 


o  ff  o 

Q.  03  *" 

CO    ,_  QJ 

-Q  I  E 

^  -^  CO 

¥  ^  E 


cp 

_z     13 
CO     =3 

i; 

CD 

2  5 


«^ 

„    CO 

C     CD  _ 

CD  SI  ro 

CD    00  CD 

•£■§  ° 


rr 

Li 

-i 

i ) 

c/> 

o 

m 

* 

U 

n 

ni 

13) 

<u 

ii 

0) 

u 

00 

5 

o 

CD   UJ 

>  TT 

O  co 


03   "O 

ol  co 


cs  o  o)  S 
cic^g 

Is-    CO    03    t 


n 

c 

iz= 

01 

CD 

n 

CO 

CO 

a 

r 

r 

c 

00 

u 

E 
"a 

o 

tj 

3 

c 
to 

u 

CO 

CO 

ni 

^_r 

(71 

iJ 

< 

C_ 
CD 

c 
CD 

u 

CD 

C 

U) 

c/) 

i 

LI 

<) 

0) 

r 

CO 
CO 

o 

o 

u 
CB 

i ) 

u 

(I) 

U) 

o 

Q.-D 

en 

CO 

CO 

i> 

o 

c_ 

o 

CNI 

c 

u 

CO 

t— 

■ 

(- 

>. 

y 

o 

0) 

o 

r 

CO 

u. 

01 

en 

J 

(1) 

E 

o 

CD 

xJ 

o 

c 

oo 

o 
F 

<u 

m 

5 

c 

o 

c_ 

03 

E 

CD 

C 

CO 

CD 

"1 

c 

CD 

CO 

I'll 

b 

o 

b 

c 

> 

m 

E8 

8| 

S^ 
5  =) 

cr  oo 


CO 


-5  ^ 

to  ^5 


CO     CL 
CD  i? 


4=   O 

Z  a? 

"3 

o10 

~   u) 

£  S 

«  to 
Q   E 


■=  en  5 


=3  CD  CO 

O  _c  O 

O  >  Q. 

w  >  a> 

•a  T3 

■O  c  ni 

CD  CO  <K 

t3  CO  k 

0)  c  o 

-*-  n  C/J 


3=    O 


c»o 
o  S  o 


S  i»  w 

O    to    S 
—     O    CD 

.E  5  « 

«    00     g 

co  -5   oo 
<   Q.  c 

coa  * 
°>-8 

E  8ff 
CJ^r: 


i;    O) 


CO    CO 
CO    CO 

a-  5 


CD    CO 

y  e 


CD  ^    CI 


C  C 
CD  C 
CO    CO 

-Q  sz 
CO  o 
_>^  CD 

to  ffl 
^  T3 


□3     . 

s  °- 


"-1  oo  co 

_    CD    p 


+    >•  CD 

SS   SJ   c? 

O    CO    03 


oo  v 

CO 

JZ)     CO  O 

5  J  2 

c  —  ^-~ 

CD  j:  Q 

co    03  S 

a;   cd  § 

O.Q-  — 

i^    CD  CD 

«,    OT  CZ 

I  8  § 


03    E 


5  oo 

,      CD 

+  -n 


CD  *- 

c    «    O 

E  an 

™  si 


SSa 


S  co^ 
o  E 

cc    CD  ofl 


cu 

r 

> 

CO 

_>. 

CD 

^ 

CB 

CD 

m 

Hi 

(  ) 

C 

cu 

D) 

i 

!  1 

Ml 

CU 

CD 

r 

P    m 


JO    C 

c  o 

CD    Q. 


—    CO 


Cl  CD 


o  91 


3    CZ 
T3    CD 

-CD 

CO 

SI  ~o 


3    2 
C/)     00 


E 

Q  <= 


g>  c 


00     CD 
T3    § 


<     CO 


O)  5 

SE 

r:   a) 

<->  X5 
O    ™ 

Z  co 


00  CD  Q- 
CD  3Z  CD 
_l    CO  "O 


O)     -  2 

CO    5  5 

r.   2  * 

2   ^  co 

CDS  c 

CO  ^£  = 

c  M  s. 

23    00  CD 

<    E  Q. 


n 

to 

C) 

x 

00 

^ 

E 

ea 

o 

cr 

o 

Lll 

c 

Q 

_l 

O 

O)  "■ 
SO. 

oo  -^ 
to  00 
CD    3 


in  o 

O  CD 

E  ° 

CO  O 

c?2 

J3  to 

CD  O 


to 
o_ 

<?    CO 

o  -t 

ill 

)  oo  E 
i  ~°  -o 

)    5    CD 

:    O  W 


o  .5? 

3    CD 

2  co 


P  -o 


II 


g)fo 


Aquatic  Physical  Features 


609 


section  can  be  inventoried  for  particle  size  by  per- 
cent of  the  total  cross  section,  and  the  transect  site 
marked  by  steel  stakes  for  future  monitoring  (Duff 
and  Cooper  1978).  The  amount  of  the  various  parti- 
cle sizes  is  estimated  by  ocular  estimates.  Estimates 
are  not  reproducible  to  a  high  degree  of  accuracy 
even  by  the  same  stream  surveyor.  This  variable 
does  not  have  a  high  degree  of  precision  from  one 
biologist  to  another  or  from  one  year  to  the  next. 
However,  it  is  valuable  to  describe  existing  habitat 
conditions  on  the  date  of  the  survey,  which  can  be 
used  for  general  trend  comparisons  on  a  continuing 
inventory  basis.  Hamilton  and  Bergersen  (1984)  de- 
scribe substrate  sampling  in  detail. 


Surface  Acreage 

Surface  acreage  can  be  determined  from  a  map 
or  aerial  photograph  using  an  acreage  dot  grid  or 
a  polar  planimeter.  An  average  of  three  readings 
from  the  map  or  photograph  should  be  used  for 
acreage  computation.  The  256-dots-per-square-inch 
dot  grid  should  be  adequate  for  lake  and  reservoir 
measurement.  Increased  precision  can  be  achieved 
by  using  the  microdot  grid  with  1,024  dots  per 
square  inch.  Hamilton  and  Bergersen  (1984)  describe 
this  technique  and  others  for  calculating  acreage. 


Shoreline  Miles 

The  number  of  shoreline  miles  can  also  be  de- 
termined using  a  planimeter.  This  measurement  is 
used  to  calculate  an  index  of  shoreline  development 
described  below. 

Shore  Development 

Shore  (or  shoreline)  development  is  a  measure 
of  the  convolution  of  the  shoreline.  An  irregular 
shoreline  provides  more  littoral  zones  for  the  pro- 
duction of  aquatic  fish  food  organisms  and  more 
spawning  and  nursing  areas  for  fish.  An  index  to 
shoreline  development  can  be  calculated  using  the 
ratio  of  actual  shoreline  length  to  shoreline  length  of 
a  circular  lake  of  the  same  acreage,  as  follows: 


shore  development  = 


2     A 


where  S  equals  the  length  of  shoreline  and  A  equals 
the  surface  area  of  the  lake  in  acres. 


Water  Temperature 

Water  temperature  can  be  accurately  recorded 
by  a  pocket  thermometer,  maximum-minimum  ther- 
mometer, or  thermograph.  The  temperature  variable 


A  surface  scatter  turbidimeter  being  used  to  monitor  stream  turbidity-  during  a  controlled 
reservoir  release  designed  to  flush  fine  sediments  from  the  stream  bed. 


610 


Aquatic  Physical  Features 


Table  6.     Average  turbidities  found  to  be  fatal  to  fish  (from  Federal  Water  Pollution  Control 
Administration  1968). 


Length  of 

Turbidity 

Species 

Exposure  (days) 

(mg/l) 

Largemouth  bass  (Micropterus  salmoides) 

7.6 

101,000 

Pumpkin  seed  sunfish  (Lepomis  gibliosus) 

13.0 

69,000 

Channel  catfish  (Ictalurus  punctatus) 

9.3 

85,000 

Black  bullhead  (/.  melas) 

17.0 

222,000 

Golden  shiner  (Notemigonus  crysoleucas) 

7.1 

166,000 

is  reproducible  with  a  high  degree  of  accuracy  when 
instruments  are  accurately  calibrated.  Hamilton  and 
Bergersen  (1984)  describe  temperature  measure- 
ments in  greater  detail. 

Turbidity 

Turbidity  is  an  expression  of  the  optical  prop- 
erty of  a  sample  of  water  which  causes  light  to  be 
scattered  and  absorbed  rather  than  transmitted  in 
straight  lines  through  a  sample  (Federal  Water  Pollu- 
tion Control  Administration  1968).  Turbidity  is 
caused  by  the  presence  of  suspended  material  such 
as  clay,  silt,  finely  divided  organic  matter,  plankton, 
and  other  microscopic  organisms.  Turbidity  is  mea- 
sured by  turbidimeters  in  Jackson  Turbidity  Units 
(JTUs)  or  Nephelometric  Turbidity  Units  (NTUs). 

Turbidity  of  not  more  than  25  JTUs  (25  NTUs) 
is  recommended  for  cold-water  streams  and  not 
more  than  50  JTUs  (50  NTUs)  for  warm -water 
streams.  High  turbidities  can  cause  fish  mortality  as 
shown  in  Table  6.  The  examples  given  are  unusually 
high  for  extended  periods  of  time.  Instream  activity 
such  as  dredging  and  placer  mining  can  cause  high 
turbidities  that  may  not  kill  fish  directly  but  cause 


other  detrimental  effects  such  as  streambed  siltation. 
Mining  can  cause  high  turbidities  that  may  not  kill 
fish  but  may  cover  the  streambed  and  suffocate  fish 
eggs  and  macroinvertebrates. 

Hamilton  and  Bergersen  (1984)  discuss  turbidity 
in  detail. 


SUMMARY 

Inventory  and  monitoring  should  be  directed  to 
the  variables  that  have  the  greatest  possibility  of 
repeated  results  by  various  workers.  The  variables 
that  can  be  measured  with  the  greatest  accuracy 
(i.e.,  water  temperature,  stream  width,  stream  depth, 
lake  depth,  surface  acreage,  stream  gradient,  and 
turbidity)  should  be  used  in  monitoring  change  in 
habitat  conditions  related  to  specific  land-use  ac- 
tions. Inventory  and  monitoring  should  measure 
those  variables  that  limit  distribution  and  abundance 
of  aquatic  organisms.  Measured  variables  should 
serve  as  indicators  of  effectiveness  of  management 
practice.  The  ultimate  measure  of  the  effectiveness 
or  impact  of  management  should  be  productivity  of 
the  fish  or  other  aquatic  species. 


Aquatic  Physical  Features 


611 


LITERATURE  CITED 


ARMOUR,  C.L.,  K.P.  BURNHAM,  and  W.S.  PLATTS.  1983. 
Field  methods  and  statistical  analysis  for  monitoring 
small  salmonid  streams.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  FWS/OBS-83/33.  200pp. 

BOVEE,  K.D.  and  R.  MILHOUS.  1978.  Hydraulic  simulation 
in  instream  flow  studies:  Theory  and  techniques. 
IFIP  5,  Cooperative  Instream  Flow  Serv.  Group,  U.S. 
Dep.  Inter.,  Fish  and  Wildl.  Serv.,  Ft.  Collins,  CO. 
130pp. 

CUPLIN,  P.  1978.  The  use  of  large  scale  color  infrared 
photography  for  stream  habitat  inventory.  U.S.  Dep. 
Inter.,  Bur.  Land  Manage.  Tech.  Note  TN-325.  Denver, 
CO.  11pp. 

,  R.  BOROVICKA,  J.  ERDMANN,  B.  VAN  HAVEREN, 

L.  LEE,  R.  MCQUISTEN,  and  M.  WHITTINGTON.  1979. 
Instream  flow  guidelines.  U.S.  Dep.  Inter.,  Bur.  Land 
Manage.,  Service  Center.  Denver,  CO.  57pp. 

DUFF,  DA.  and  J.L.  COOPER.  1978.  Techniques  for  con- 
ducting stream  habitat  surveys  on  National  Resource 
Land.  U.S.  Dep.  Inter.,  Bur.  Land  Manage.,  Tech.  Note 
TN-283.  72pp. 

FEDERAL  WATER  POLLUTION  CONTROL  ADMINIS- 
TRATION. 1968.  Water  quality  criteria.  Washington, 
DC.  234pp. 


HAMILTON,  K.  and  E.P.  BERGERSEN.  1984.  Methods  to 
estimate  aquatic  habitat  variables.  Colorado  Coop. 
Fishery  Res.  Unit.  Colorado  State  Univ.,  Fort  Collins. 

HAYS,  R.L.,  C.  SUMMERS,  and  W.  SEITZ.  1981.  Estimating 
wildlife  habitat  variables.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  FWS/OBS-81-47.  111pp. 

PARSONS,  S.C.  and  S.  HUDSON.  1985.  Stream  channel 
cross-section  surveys  and  data  analysis.  U.S.  Dep. 
Inter.,  Bur.  Land  Manage.  TR-4341-1.  Denver,  CO. 
48pp. 

PFANKUCH,  D.J.  1975.  Stream  reach  inventory  and  chan- 
nel stability  evaluation.  U.S.  Dep.  Agric,  For.  Serv. 
Northern  Region.  R 1-75-002.  26pp. 

PLATTS,  W.S.  1981.  Stream  inventory:  Acquisition  and 
utilization  of  aquatic  habitat  inventory  information. 
Pages  75-84  in  Proc.  Symp.  Western  Div.  Am.  Fish. 
Soc,  Portland,  Oregon,  October  1981. 

,  W.F.  MEGAHAN,  and  G.W.  MINSHALL  1983-  Meth- 
ods for  evaluating  stream,  riparian,  and  biotic  condi- 
tions. U.S.  Dep.  Agric.  For.  Serv.  Gen.  Tech.  Rep.  INT- 
138.  70pp. 

ROBINS,  C.R.  and  R.W.  CRAWFORD.  1954.  A  short  accu- 
rate method  for  estimating  the  volume  of  streamflow. 
J.  Wildl.  Manage.  18:366-369. 

U.S.  DEPARTMENT  OF  THE  INTERIOR,  BUREAU  OF 
LAND  MANAGEMENT.  1979.  Wetland-riparian  area 
protection  and  management.  Manual  6740.  Washing- 
ton, DC.  16pp. 


612 


Aquatic  Physical  Features 


29 

HYDROLOGIC 
PROPERTIES 

Bruce  Van  Haveren 


U.S.  Bureau  of  Land  Management 
Service  Center 
Denver,  CO  80225 


Editor's  Note:  Biologists  know  that  animals  do  not 
exist  in  isolation  and  that  the  habitat  in  which 
they  live — the  ecosystem  of  which  they  are  a  part — 
must  be  understood.  This  interrelationship  is  fre- 
quently the  first  principle  ignored  by  students  as 
they  eagerly  pursue  a  narrow  disciplinary  ap- 
proach. Water  is  the  single  most  important  life 
requirement  for  all  plant  and  animal  life.  Thus,  the 
movement  and  properties  of  water  in  an  animal's 
habitat  should  be  of  concern  to  the  biologists  con- 
ducting a  habitat  inventory  and  monitoring  effort, 
particularly  in  the  arid  western  United  States.  Yet 
hydrology,  the  study  of  water  in  the  natural  envi- 
ronment, is  virtually  ignored  by  many  biologists. 
This  chapter  describes  some  of  the  common  meas- 
urements of  hydrologic  properties  useful  to  the 
biologist. 


INTRODUCTION 

Water,  one  of  the  vital  wildlife  habitat  elements 
on  publie  lands,  is  the  lifeblood  of  all  ecosystems, 
terrestrial  as  well  as  aquatic.  Its  natural  occurrence 
and  circulation  are  especially  vital  to  the  functioning 
and  vigor  of  wetland  and  riparian  systems.  Biologists 
must  remember  that  all  terrestrial  and  aquatic  eco- 
systems are  tied  to  hydrologic  and  geomorphic  sys- 
tems. For  example,  to  manage  riparian  habitat,  one 
must  consider  the  surface  and  groundwater  systems 
that  are  hydrologically  related. 

Fishery  and  wildlife  biologists  should  understand 
that  subtle  changes  in  land  use  can  produce  not-so- 
subtle  alterations  in  hydrologic  conditions,  which 
can  directly  influence  fish  and  wildife  habitat.  Fur- 
thermore, hydrologic  changes  may  occur  several 
miles  from  the  habitat  site  that  is  or  could  be 
affected. 

This  chapter  focuses  on  the  measurement  and 
use  of  certain  hydrologic  properties  important  to 
wildlife  habitat.  The  hydrologic  properties  of  both 
lentic  (lakes,  ponds,  bogs,  meadows)  and  lotic  (riv- 
ers, streams,  springs)  habitats  are  discussed  in  terms 
of  measurement  techniques  and  data  applications. 

Hydrologic  Terminology 

Biologists  should  be  familiar  with  common 
terms  used  in  land-use  hydrology'  to  facilitate  com- 
munication with  their  hydrologist  colleagues.  Precip- 
itation, evaporation,  runoff,  soil  water  storage,  and 
sediment  transport  are  all  examples  of  hydrologic 
processes.  Hydrologic  variables,  such  as  instanta- 
neous stream  discharge,  monthly  groundwater  levels, 
or  daily  lake  evaporation,  quantitatively  describe 
hydrologic  processes.  Hydrologic  variables  are  nearly 
always  continuous  in  time  and,  therefore,  must  be 
observed  or  measured  at  given  intervals  to  convert 
them  to  discrete  variables.  When  discrete  variables 


Hydrologic  Properties 


613 


are  described  statistically,  they  are  called  hydrologic 
parameters.  Mean  annual  precipitation,  the  100-year 
peak  river  discharge,  and  the  coefficient  of  variation 
of  monthly  suspended  sediment  are  all  examples 
of  hydrologic  parameters. 

The  era  of  the  Environmental  Impact  Statement 
(EIS)  has  brought  with  it  a  new  vocabulary.  In  the 
field  of  hydrology,  terms  such  as  hydrologic  changes, 
effects,  and  impacts  are  all  used  in  environmental 
reports,  often  interchangeably.  However,  used  cor- 
rectly, hydrologic  change  refers  to  the  increase  or 
decrease  (statistically  at  a  given  level  of  probability) 
in  a  hydrologic  variable  or  parameter.  For  example,  a 
100-mg/L  increase  (significant  at  the  0.80  level  of 
probability)  in  mean  daily  suspended  sediment  rep- 
resents a  hydrologic  change. 

The  term  hydrologic  effect  invokes  a  cause  and 
effect  relationship.  A  significant  hydrologic  effect 
would  be  road  construction  causing  the  100-mg/L 
increase  in  mean  daily  suspended  sediment. 

Finally,  a  distinction  must  be  made  between 
effects  and  impacts.  From  a  hydrologic  or  water  re- 
sources standpoint,  an  impact  occurs  only  when  the 
value  of  that  resource  diminishes  (negative  impact) 
or  increases  (positive  impact).  Referring  back  to  our 
suspended  sediment  example,  the  100-mg/L  increase 
due  to  road  construction  cannot  be  called  an  impact 
unless  the  effect  has  resulted  in  a  change  in  a  related 
resource  value. 


Watershed  Stability:  The  Concept 

Watershed  stability  has  an  inherent  relationship 
to  the  stability  and  productivity  of  wildlife  habitat. 
Watershed  stability  includes  the  following  elements: 
soil  productivity,  hydrologic  response  to  precipita- 
tion, quality  of  surface  and  shallow  groundwater, 
movement  of  nutrients  by  water,  and  structure  and 
function  of  wetland-riparian  systems. 

These  may  be  viewed  as  components  as  well  as 
partial  indexes  of  watershed  stability.  They  are 
highly  interrelated  and  directly  influence  habitat 
variables. 

If  a  simple  equation  could  be  used  for  measur- 
ing the  present  watershed  stability,  it  would  look 
like  this: 

Watershed  stability  =  watershed  constants  + 
climate  variables  +  present  watershed  condition  + 
land-use  variables. 

The  watershed  constants  that  influence  stability 
include  geology,  soils,  latitude  and  elevation,  topog- 
raphy (relief,  slope,  and  aspect),  and  vegetation  type, 


both  present  and  potential.  These  factors  are  static 
over  time  and  easily  determined. 

Watershed  condition  and  trend  are  determined 
semi-quantitatively  by  evaluating  soil  erodibility  (see 
Soils  chapter),  mass  wasting  potential  (potential  for 
mass  slope  failures),  stream  channel  stability  (see 
section  on  Stream  Channel  Stability  and  Sediment 
Transport,  this  chapter),  and  watershed  infiltration- 
storage  capacity. 

The  climate  factors  important  to  watershed  sta- 
bility are  limited  to  the  frequency  and  intensity  of 
runoff-producing  precipitation,  seasonality  of  precipi- 
tation, and  the  frequency  and  duration  of  drought. 

Land  use  refers  to  past  and  present  land-use 
factors  that  influence  watershed  stability,  such  as 
road  construction,  timber  harvesting,  excessive  live- 
stock trampling,  overutilization  of  forage,  mining, 
and  mineral  or  energy  exploration  activities.  Also  in- 
cluded would  be  natural  events  that  may  affect  wa- 
tershed stability  such  as  storms,  wildfires,  floods,  and 
insect  or  disease  attacks  on  vegetation. 

All  of  these  factors  are  evaluated  in  the  present, 
assigned  an  appropriate  weight,  and  then  combined 
into  an  overall  watershed  stability  rating.  Trend  in 
watershed  stability  can  be  inferred  by  comparing 
past  and  present  watershed  stability  ratings  and  by 
considering  present  and  near-future  land-use  activi- 
ties that  influence  watershed  stability. 

Watershed  stability  ratings  are  subjective.  Rating 
systems  are  nearly  always  tailored  to  fit  specific  lo- 
cales and  applications.  For  this  reason,  information 
on  watershed  stability  is  generally  not  available  from 
other  sources  and  must  be  obtained  directly  from 
field  surveys.  Before  field  work,  the  area  in  question 
must  be  stratified  on  the  basis  of  hydrologic  re- 
sponse units  (homogeneous  areas  that  respond  simi- 
larly to  a  given  hydrologic  event). 

Because  land  management  activities  can  affect 
stability,  the  specialist  must  determine  and  present 
to  management  the  acceptable  versus  unacceptable 
levels  of  watershed  stability.  The  consequences  of 
unacceptable  watershed  stability  must  be  thoroughly 
explained  in  terms  of  impact  magnitude  and  proba- 
bility of  occurrence. 


PRECIPITATION,  SNOW,  AND 
EVAPORATION 

Wildlife  biologists  may  on  occasion  have  need 
for  hydrometeorological  data  such  as  precipitation, 
evaporation,  and  snow  depth.  Precipitation  and  evap- 
oration have  an  indirect  influence  on  habitat  through 
the  fluctuations  of  lake  and  pond  levels  and  the 
water  balance  of  wetland  ecosystems.  Snow,  in  terms 


614 


Hydrologic  Properties 


of  depth  and  areal  distribution,  affects  animal  move- 
ment and  food  availability,  and  thus  the  health  of 
many  wildlife  species. 

Measurement  of  Precipitation 

Precipitation  includes  rain,  snow,  and  all  other 
forms  of  precipitated  atmospheric  water.  The  biolo- 
gist should  first  ascertain  the  type  and  quality  of  data 
needed.  Are  data  needed  on  a  year-round  basis  or 
just  seasonally?  What  is  the  desired  frequency  of  the 
data:  monthly,  daily,  hourly?  Will  snowfall  data  be 
needed  as  well  as  rainfall  data?  These  questions  must 
first  be  answered  before  a  gage  and  sample  schedule 
are  selected.  For  most  wildlife  applications,  monthly 
precipitation  totals  collected  year-round  would  suf- 
fice, particularly  for  U.S.  Bureau  of  Land  Management 
(BLM)  special  wildlife  management  areas  where 
water  supply  is  a  concern.  Many  national  wildlife 
refuges,  for  example,  routinely  collect  this  type  of 
precipitation  data. 

Cumulative  rain  is  measured  in  a  storage  can 
having  an  orifice  diameter  of  at  least  20.3  cm  (8  in.). 
Many  rain  gages  are  available  that  will  store  up  to  a 
year's  accumulation  of  precipitation.  Lightweight 
oil  is  added  to  retard  evaporation. 

Recording  gages  are  necessary  if  one  needs  to 
know  rainfall  intensity  or  the  time  of  occurrence  of 
rainfall.  There  are  several  commercial  types  available. 
Agricultural  Handbook  224  (Brakensiek  et  al.  1979) 
includes  a  good  discussion  on  precipitation  record- 
ing and  characteristics  of  various  gage  types. 

Measurement  of  precipitation  falling  as  snow  is 
much  more  difficult.  Most  precipitation  gages  used 


currently  in  the  U.S.  underestimate  snowfall.  Effi- 
ciency of  catch  depends  on  the  size  of  orifice,  gage 
siting,  and  whether  or  not  the  gage  has  a  windshield 
(Figure  1).  The  efficiency  of  catch  of  snowfall  at 
this  site  has  been  at  least  90% .  Catch  efficiency  is 
checked  by  measuring  the  water  content  of  snow  on 
the  ground  at  the  gage  and  comparing  this  to  the 
gage  catch  for  the  same  period  of  time. 

Measurement  of  Evaporation 

Evaporation  is  also  a  difficult  variable  to  meas- 
ure. Lake  or  pond  evaporation  is  usually  inferred 
from  measurements  of  pan  evaporation.  The  National 
Weather  Service's  1.2-m  (4-ft)  diameter  class  A  evap- 
oration pan  (Figure  2)  is  the  standard  instrument 
for  evaporation  measurements  in  the  U.S.  The  U.S. 
Agriculture  Handbook  224  contains  a  good  discus- 
sion on  siting  and  operation  of  a  class  A  pan. 


Figure  1.  A  precipitation  storage  gage  with  an 
"alter  shield"  at  an  ideal  gage  site  in  a  forest 
clearing. 


Figure  2.     Class  A  evaporation  pan. 


Measurement  of  Snow  on  the  Ground 

Depth  of  snow  on  the  ground  is  important  be- 
cause it  may  limit  access  to  food  and  make  move- 
ment difficult  for  big  game  animals.  Depth  of  snow  is 
measured  with  a  yardstick  or  graduated  rod.  Snow 
depth,  because  of  its  extreme  variability,  must  be 
sampled  carefully  and  systematically.  Stratified  sam- 
pling, considering  elevation,  vegetation  type,  and 
slope  aspect  strata,  is  recommended. 


Hydrologic  Properties 


615 


Measurement  of  the  water  content  of  snow  is  of 
interest  for  calculating  water  balances;  these  data 
may  also  supplement  or  verify  winter  precipitation 
gage  catch.  The  U.S.  Soil  Conservation  Service  (SCS), 
for  example,  operates  a  Westwide  network  of  snow 
surveys  used  for  forecasting  spring  and  summer  run- 
off for  irrigators,  reservoir  operators,  and  hydro- 
power  companies.  There  is  no  reason  why  the  same 
techniques  could  not  be  used  to  forecast  wildlife 
water  supplies  in  certain  situations. 

The  standard  federal  snow  sampler  is  used  for 
measuring  both  depth  and  water  content  in  moun- 
tain snowpacks  (Figure  3).  Agriculture  Handbook 
224  gives  a  brief  description  of  snow  surveying.  The 
SCS  State  snow  survey  supervisor  (located  in  the 
SCS  State  Office  of  each  western  State  except  Califor- 
nia) can  be  contacted  for  more  information  and 
assistance. 

Snow  density,  the  ratio  of  snow  water  content 
to  snow  depth,  can  be  computed  directly  from  the 
measurements  data  taken  with  the  federal  snow  sam- 
pler. Snow  density  is  a  good  indicator  of  the  matu- 
rity or  ripeness  of  a  seasonal  snowpack.  New  snow 
densities  vary  between  0.07  and  0.25  g/cm  .  Late 
season  snowpacks  range  between  0.30  and  0.45  g/ 
cm3.  As  snowpack  density  approaches  0.50  g/cm  , 
the  snowpack  is  said  to  be  "ripe"  and  is  ready  for 
snowmelt  release. 


Figure  3.     A  standard  federal  snow  sampler  kit  for 
measuring  snow  depth  and  water  content. 


ators,  U.S.  Forest  Service,  and  Agricultural  Research 
Service  experiment  stations.  Snow  data  are  available 
for  selected  NWS  stations  and  from  the  SCS. 

Hydrometeorological  data  are  scarce  for  public 
lands  in  general.  The  State  climatologist  is  generally 
a  good  source  of  information  about  data  availability 
for  specific  areas  within  the  State.  The  State  climatol- 
ogist can  also  assist  with  data  extrapolation  and 
interpretation. 


STREAMFLOW  MEASUREMENT  AND 
CHARACTERISTICS 

Wildlife  biologists,  and  particularly  fishery  biolo- 
gists, occasionally  need  streamflow  data.  The  antici- 
pated use  of  these  data  will  determine  the  data  type 
and  desired  accuracy.  For  example,  for  some  pur- 
poses such  as  water  quality  grab-sampling,  an  instan- 
taneous discharge  measurement  will  suffice  (and 
should  always  accompany  a  water  quality  sample). 
For  long-term  characterization  of  streamflow,  dis- 
charge data  continuously  recorded  over  several  years 
are  necessary.  To  characterize  flow  and  channel  hy- 
draulic conditions  for  a  given  channel  reach,  one 
would  need  only  a  few  discharge  measurements.  The 
BLM  Instream  Flow  Guidelines  (Cuplin  and  Van  Hav- 
eren  1979)  should  be  consulted  for  advice  on  the 
measurement  of  stream  discharge  in  instream  flow 
studies. 

A  decision  on  the  desired  accuracy  of  these  data 
is  important  because  it  directly  relates  to  the  mea- 
surement method  chosen  and  corresponding  cost. 
Eight  streamflow  measurement  methods  can  be  used 
for  typical  small  stream  applications: 

( 1 )  Float  method 

(2)  Volumetric  measurements 

(3)  Current-meter  and  velocity-head  measure- 
ments 

(4)  Portable  weirs  and  flumes 

(5)  Dye-dilution  method 

(6)  Continuous  streamflow  recording — natural 
control 

(7)  Continuous  streamflow  recording — artificial 
control 

(8)  Indirect  methods 

Float  Method 


Sources  of  Published  Data 

Published  precipitation,  evaporation,  and  snow 
data  are  available  for  the  more  populated  areas  in 
the  West.  Sources  of  precipitation  and  evaporation 
data  include  the  National  Weather  Service  (NWS), 
State  agricultural  experiment  stations,  reservoir  oper- 


The  float  method  is  the  least  expensive,  but 
least  accurate,  of  the  available  methods.  A  partially 
submerged  float  (water-logged  stick,  orange,  piece  of 
ice,  or  half-full  beer  can)  is  timed  through  a  meas- 
ured reach  to  estimate  average  water  velocity  (the 
average  of  several  runs  is  computed ).  A  cross  section 
of  the  wetted  portion  of  the  channel  is  then  meas- 
ured. The  stream  discharge  is  therefore  the  product 


616 


Hydrologic  Properties 


of  the  average  velocity  and  the  cross  section.  If  a 
good  estimate  of  velocity  is  obtained,  the  method  is 
accurate  within  ±  10%.  Under  conditions  where 
velocity  cannot  be  estimated  adequately,  the  error 
could  be  ±25%. 


Volumetric  Measurements 

Volumetric  measurements  are  taken  whenever 
the  flow  is  sufficiently  small  and  is  concentrated 
or  can  be  concentrated  so  that  all  of  it  may  be 
caught  in  a  container.  Time  required  to  fill  a  known 
volume  is  measured  with  a  stopwatch.  The  method 
is  fairly  accurate,  with  the  error  averaging  ±  2  to  3% 
or  less.  Volumetric  measurements  can  often  be  made 
at  culverts  by  using  a  19-L  or  38-L  (5-gal.  or  10-gal.) 
bucket  for  small  flows  or  a  208-L  (55-gal.)  drum 
for  larger  flows.  I  have  often  carried  a  piece  of  thin 
sheet  metal  in  my  vehicle  for  forming  temporary 
flumes  to  facilitate  volumetric  measurements. 

At  culvert  sites  where  the  flow  is  too  large  for  a 
volumetric  measurement,  the  float  method  may  be 
used  with  reasonable  accuracy. 


Current-Meter  and  Velocity-Head 
Measurements 

Although  several  types  of  current  meters  are  in 
use  at  this  time,  the  Price  AA  and  pygmy  vertical-axis 
meters  (Figure  4)  are  used  most  commonly  by  the 
U.S.  Geological  Survey,  Water  Resources  Division. 
Because  current  meters  measure  velocity  at  a  point, 
several  velocity  measurements  are  taken  at  a  stream 
cross  section.  The  number  of  measurements  in  the 
vertical  direction  is  chosen  according  to  the  accu- 
racy desired.  The  one-point  procedure,  where  a  sin- 
gle velocity  measurement  is  taken  (for  each 
horizontal  point  on  the  cross  section)  at  18.3-cm 
(0.6-ft)  depth  below  the  surface,  is  most  commonly 
used  for  small  streams.  In  larger  streams  and  where 
more  accuracy  is  desired,  velocity  measurements  are 
taken  at  the  6.1-  and  24.4-cm  (0.2-  and  0.8-ft)  depths 
below  the  surface.  The  gain  in  accuracy  is  about  one 
percentage  point.  Additional  accuracy  may  be  ob- 
tained with  the  three-point  method  at  6-,  18-,  and 
24-cm  (0.2-,  0.6-,  and  0.8-ft)  depths  where  stream 
depths  exceed  76.2  cm  (2.5  ft).  Spacing  of  measure- 
ment points  across  the  channel  depends  on  the 
width  of  the  stream.  Normal  spacing  is  within  0.6  or 


Figure  4.     Price  AA  and  pygmy  current  meters  for  measuring  stream  velocity. 


Hydrologic  Properties 


617 


1.5  m  (2  or  5  ft);  however,  the  points  should  be 
spaced  so  that  no  subsection  has  more  than  10%  of 
the  total  discharge. 

Selection  of  a  suitable  site  for  current  meter 
measurements  should  follow  these  criteria  (see 
Figure  5): 

( 1 )  Cross  section  lies  within  a  straight  reach  and 
flow  lines  are  parallel. 

(2)  Velocities  exceed  15  cm/sec  (0.5  ft/see)  and 
maximum  depths  exceed  15.2  cm  (0.5  ft). 

(3)  Streambed  is  relatively  uniform,  free  of  large 
boulders,  organic  debris,  and  aquatic  growth. 

(4)  Flow  is  uniform,  free  of  eddies,  slack  water, 
and  excessive  turbulence. 

(5)  The  stream  can  be  waded  from  bank  to  bank. 

If  a  stream  is  too  deep  or  velocities  too  high 
for  wading,  a  suitable  bridge  site  must  be  located.  In 
this  situation,  a  suspension  cable  is  substituted  for 
the  wading  rod.  Procedures  for  taking  current  meter 
measurements  from  bridges  are  covered  in  detail 
by  Rantz  (1982). 


A  standard  discharge  measurement  form  used  by 
the  U.S.  Geological  Survey  for  all  stream  discharge 
measurements  is  shown  in  Figure  6.  The  "midsection 
method"  (Buchanan  and  Somers  1969;  Rantz  1982) 
is  used  to  compute  stream  discharge  from  this  form. 

Accuracy  of  the  current  meter  method  depends 
on  flow  and  channel  conditions,  number  of  velocity 
measurements  taken  in  the  vertical  direction,  spacing 
of  measurements  across  the  channel,  condition  and 
calibration  of  current  meter,  and  operator  error.  U.S. 
Geological  Survey  provides  the  following  accuracy 
limits: 


Overall  Rating  of 
Measurement 

Excellent 
Good 
Fair 
Poor 


Error 

±2% 
±5% 
±8% 
>8% 


In  spite  of  their  lower  cost  and  ease  of  opera- 
tion, velocity-head  rods  (Figure  7)  have  not  achieved 
the  same  popularity  as  current  meters  for  gaging 
streams.  The  rod  consists  of  a  broad  edge  at  least  2.5 


Figure  5.     A  good  location  for  a  current  meter  measurement. 
618  Hyiirologic  Properties 


cm  (1  in.)  wide  and  a  sharp  edge  opposed  180°  to 
the  broad  edge.  Constructed  of  hardwood  or  alumi- 
num, the  rod  is  scaled  in  English  or  metric  units. 
To  measure  the  velocity  head,  the  sharp  edge  of  the 
rod  is  pointed  upstream  and  the  water  elevation  read 
from  the  rod.  The  rod  is  then  revolved  180°  so  the 
broad  edge  is  perpendicular  to  the  flow  lines.  The 
water  elevation  is  again  read,  and  the  difference  be- 
tween readings  computed.  The  resulting  number 
is  the  velocity  head  expressed  in  feet  or  meters.  Ve- 
locity m/sec  (ft/sec)  is  found  by  the  equation — 

V  =  2gh 

where  g  =  acceleration  due  to  gravity  (ft/sec/sec) 
and  h  the  velocity  head  (ft). 

Under  very  good  channel  conditions,  the  veloc- 
ity-head rod  will  give  discharge  readings  within  a 
few  percentage  points  of  the  true  value  (Wilm  and 
Storey  1944).  It  is  comparable  in  accuracy  to  the 
one-point  method  using  current  meters.  Heede 
(1974)  recommended  use  of  the  velocity-head  rod 
only  if  channel  bottoms  are  fairly  smooth  and  chan- 
nels are  approximately  prismatic.  This  precludes 
their  use  in  boulder-strewn  mountain  streams  with 
highly  turbulent  flows,  unless  the  gaged  section  can 
be  modified  to  fit  the  above  criteria. 

Because  of  the  surging  nature  of  flow  in  natural 
streams,  the  precision  of  head  measurements  is 
around  ±  0.01  m  (  ±  0.05  ft).  Heede  (1974)  recom- 
mended reading  the  water  elevation  that  occurs 
most  frequently,  rather  than  the  average  elevation. 

Portable  Weirs  and  Flumes 

Portable  weir  plates  and  pre-calibrated  flumes 
can  be  used  to  accurately  measure  small  flows.  Both 
rectangular  and  triangular  (30°,  60°,  90°  or  120° 
notches )  weirs  are  used;  the  specific  design  is  based 
on  the  most  probable  range  of  expected  flows  (Van 
Haveren  1986). 

Parshall  flumes  are  the  most  common  pre-cali- 
brated, portable  flumes  in  use  (Figure  8).  They  are 
commercially  available  in  a  variety  of  sizes  and  are 
easily  installed  at  most  locations  (see  Rantz  1982). 

Dye-Dilution  Method 

The  dye-dilution  method  is  fairly  sophisticated 
and  requires  moderately  expensive  sampling  equip- 
ment. A  known  quantity  of  tracer  dye  is  injected 
at  the  upstream  end  of  the  reach  being  measured, 
allowed  to  mix  thoroughly  through  natural  stream 
turbulence,  and  the  resulting  concentration  meas- 
ured at  a  downstream  point.  The  technique  has  defi- 
nite advantages  in  very  turbulent  mountain  streams 
where  more  traditional  stream  gaging  techniques  are 


not  possible.  Intensive  training  is  required  to  effec- 
tively use  the  technique.  Refer  to  Rantz  ( 1982)  for  a 
complete  discussion  of  this  technique. 

Continuous  Streamflow  Measurement — 
Natural  Control 

Continuous  measurement  of  streamflow  gener- 
ally involves  more  expense  and  expertise  than  the 
methods  discussed  above.  The  method  involves  di- 
rect sampling  of  water  level  (stage)  and  the  estab- 
lishment of  a  stage-discharge  relationship  from  which 
streamflow  is  computed.  This  method  is  expensive 
in  terms  of  labor  and  equipment.  As  in  all  stream 
gaging  methods,  site  selection  is  critical.  Where  an 
artificial  control  (e.g.,  bridge,  culvert,  flume,  bank 
revetment)  is  not  used,  a  suitable  natural  control 
must  be  found.  Rantz  ( 1982)  listed  criteria  to  be 
used  in  selecting  a  gage  site  with  a  natural  control. 
Water  level  is  recorded  continuously  with  a  stage 
recorder.  Stream  discharge  is  related  to  stage  height 
by  periodically  making  discharge  measurements  at 
known  stages.  A  stage-discharge  rating  curve  is  estab- 
lished from  these  measurements,  taken  at  a  wide 
range  of  stage  heights.  The  rating  curve  is  checked 
and  modified,  usually  annually.  The  stage-discharge 
curve  will  remain  stable  at  sites  having  good  natural 
channel  control  and  no  net  sediment  accumulation 
or  channel  erosion. 


Continuous  Streamflow  Recording — 
Artificial  Control 

Artificial  controls  are  installed  where  a  good 
natural  control  cannot  be  found,  or  when  a  very 
high  degree  of  accuracy  is  desired.  Artificial  controls 
include  weirs;  flumes;  and  concrete,  rock,  or  treated 
wood  control  structures  that  maintain  a  stable  chan- 
nel geometry.  Further  discussion  of  the  types  of  arti- 
ficial controls  available,  including  their  advantages 
and  disadvantages,  is  beyond  the  scope  of  this  chap- 
ter. Brakensiek  et  al.  (1979)  and  Rantz  ( 1982)  pro- 
vided good  discussions  of  weirs  and  flumes.  Some 
weirs  and  flumes  are  available  with  known  calibra- 
tions, precluding  the  need  for  field  development 
of  the  stage-discharge  rating  curve. 

Discharge  measurements  can  also  be  made  in 
reservoirs  where  presumably  all  inflow  is  contained 
in  the  structure  with  negligible  leakage.  A  water 
level  recorder  is  used  to  measure  water  depth, 
which  is  related  to  inflow  volume  through  a  reser- 
voir stage-capacity  curve. 

Indirect  Methods 

Indirect  methods  of  streamflow  measurement 
include  channel  geometry  relationships  and  the  Man- 
nings equation. 


Hydrologic  Properties 


619 


9-275-G  (Slnffle  5x8) 

(Rev.  5-63) 


Sta.  No. 


UNITED  STATES 

DEPARTMENT  OF  THE  INTERIOR 

GEOLOGICAL  SURVEY 

WATER  RESOURCES  DIVISION 

DISCHARGE  MEASUREMENT  NOTES 


Meu.  No. 


Comp.  by  -. 
Checked  by 


Date ,  19 Party 

Width Area Vel G.  H. 

Method No.  sees. G.  H.  change 


Disch. 


in hrs.     Susp. 

Method  coef.  .. _     Hor.  angle  coef.  Susp.  coef Meter  No.  


GAGE  READINGS 

Date  rated .       Used  rating 

Time 

Recorder 

Inside 

Outside 

for  rod susp.      Meter  . _.  ft. 

above  bottom  or  wt.      lags  checked 

Spin  before  meas after  . 

Meas.  plots  %  diff.  from  .. rating 

Wading,  cable,  ice,  boat,  upstr.,  downstr.,  side 

bridge feet,  mile,  above,  below 

gage,  and 

Check-bar,  chain  found 

Weighted  M.  G.  H..._ 

changed  to at 

Correct 

Correct  M. 

G.  H 

Levels  obtained 

Measurement  rated  excellent  (2  %),  good  (5%),  fair  (8%),  poor  (over  8%),    based   on  following 

conditions:  Cross  section 

Flow Weather 

Other Air °F@ 


Gage Water 


-°F@ 


Record  removed Intake  flushed  fr. 


Observer 


Control 


Remarks 


G.  H.  of  zero  flow ft. 


■fr  GPO  :  1963— O-688360 


Figure  6.     Standard  discharge  measurement  form  used  by  the  U.S.  Geological  Survey. 


620 


Hydrologic  Properties 


.0               .10              .20                .30                 .40                    .50                          .60                                   .70                      .75 

River  at — 

—  'o 
< 

Dist. 
from 
initial 
point 

Width 

Depth 

>  & 
0-2 

Rev- 
olu- 
tions 

Time 

in 
sec- 
onds 

VELOCITY 

Adjusted 
for  hor. 
angle  or 

Area 

.80 
Discharge 

At 
point 

Mean 
in  ver- 
tical 

.85 

.90 

.92 

.94 

.96 

.97 

.98 

.99 

o 

1.00 

.99 

.98 

.97 

.96 

.94 

.92 

.90 

.85 

.80 

.0               .10              .20                .30                 .40                    .50                          .60                                 .70                      .75 

Figure  6.     Standard  discharge  measurement  form  used  by  the  U.S.  Geological  Survey 
(concluded). 


Hydrologic  Properties 


621 


The  Manning  equation  can  be  used  to  estimate 
flows  in  natural  channels: 


1.486AR23S,/2 


Figure  7.     A  velocity-head  rod  for  measuring  stream 
velocity. 


Q  = 

n 

where  A  is  the  cross  section  of  the  channel,  R  the 
hydraulic  radius,  S  the  slope  of  the  energy  line  (usu- 
ally approximated  by  the  water-surface  slope ),  and 
n  the  roughness  coefficient.  All  the  independent 
variables  can  be  measured  in  the  field  except  n, 
which  must  be  estimated.  This  technique  is  often 
used  for  estimating  flood  peaks  if  sufficient  high- 
water  marks  can  be  found.  A  minimum  error  of  10% 
can  be  expected  even  if  the  roughness  coefficient 
chosen  is  within  ±  0.005  of  the  true  value. 

Channel  geometry  can  be  used  to  estimate  mean 
annual  streamflow  and  flood  peaks  of  various  fre- 
quencies. Using  only  active-channel  width,  Hedman 
and  Osterkamp  (1982)  constructed  equations  for 
predicting  the  2-,  5-,  10-,  25-,  50-,  and  100-yr  flood 
discharges  and  the  mean  annual  runoff  for  four  re- 
gions in  the  western  U.S.  State-specific  channel  ge- 
ometry relations  have  been  established  for  most 
States  by  the  respective  Water  Resources  Division 
District  Office  of  the  U.S.  Geological  Survey. 

Field  measurement  of  active-channel  width  re- 
quires very  little  equipment  and  effort.  Training  is 
required  in  order  to  properly  identify  the  active 
channel  elevation. 


KqU 


Figure  8.     A  Parshall  flume  for  the  continuous  measurement  of  streamflow. 
622  Hydrologic  Properties 


Also,  annual,  monthly,  or  even  daily  streamflow 
can  be  estimated  by  using  index  techniques.  A  gaged 
stream  with  characteristics  similar  to  the  ungaged 
stream  of  interest  can  be  used  to  index  or  predict 
some  streamflow  variables.  This  is  accomplished  by 
first  expressing  the  streamflow  data  on  a  "per  square 
mile"  basis.  Adjustments  in  certain  streamflow  vari- 
ables should  then  be  made,  based  on  differences 
in  elevation,  exposure,  geology,  or  other  appropriate 
hydrophysiographic  factors.  This  procedure  is  highly 
subjective  and  should  be  done  in  consultation  with 
an  experienced  hydrologist. 


Sources  of  Streamflow  Data 

Published  streamflow  data  are  available  from  the 
U.S.  Geological  Survey,  Water  Resources  Division, 
on  many  larger  streams  and  rivers  in  the  nation.  Data 
for  small  watersheds  in  remote  areas  in  the  West 
are  more  difficult  to  acquire.  Figure  9  is  taken  from 
Water  Resources  Data  for  Colorado,  1985  (U.S.  Geo- 
logical Survey  1985).  This  is  a  standardized  stream- 
flow  data  format  used  in  every  State.  Maps,  available 
from  the  Water  Resource  Division  Offices,  show 
locations  of  streamflow  gaging  stations. 


HYDROLOGY  OF  LAKES,  PONDS,  AND 
RESERVOIRS 

The  hydrologic  factors  that  relate  to  wildlife 
habitat  characteristics  of  standing  water  bodies  in- 
clude inflow  and  outflow  rates,  volume  (or  capacity), 
mean  and  maximum  depths,  basin  shape,  precipita- 
tion-evaporation balance,  drainage  basin  area,  and 
groundwater  interaction  (both  physical  and 
chemical ). 


Before  undertaking  a  field  inventory  of  a  lake, 
pond,  or  reservoir,  an  attempt  should  be  made  to 
characterize  the  water  body  on  the  basis  of  origin, 
geology,  morphometric  and  physiographic  properties 
(which  can  be  obtained  from  maps),  drainage  basin 
characteristics,  trophic  state,  and  hydrologic  setting. 
Winter  (1977)  discussed  hydrologic  classification 
and  characterization  of  lakes. 


Inflow  and  outflow  characteristics  (including 
quantity  and  quality  of  the  water)  directly  influence 
lakes,  ponds,  and  reservoirs.  The  measurement  of 
inflow  or  outflow  usually  involves  stream  gaging.  Re- 
fer to  "Streamflow  Measurement  and  Characteristics." 


The  State  Engineer's  Office  is  another  potential 
source  of  streamflow  data.  The  local  U.S  Forest  Ser- 
vice hydrologist  should  also  be  consulted  if  national 
forest  lands  are  involved  or  are  close  to  the  area  in 
question. 


Interpretation  of  Streamflow  Data 

Analysis  and  interpretation  of  streamflow  data 
depend  on  the  intended  use  of  these  data.  Discrete 
values  of  streamflow  (hourly,  daily,  monthly)  can  be 
plotted  as  a  "hydrograph"  for  visual  analysis  pur- 
poses. Figure  10  is  an  example  of  an  annual  hydro- 
graph  based  on  monthly  streamflow  data.  Low-flow 
statistics,  such  as  the  minimum  7-day  mean  flow,  can 
be  computed  easily  from  daily  flow  data.  For  U.S. 
Geological  Survey  continuous  streamflow  data,  a 
number  of  different  streamflow  statistics  are  available 
through  the  WATSTORE  data  bank,  accessible 
through  any  U.S.  Geological  Survey  Water  Resources 
Division. 

Streamflow  data  for  other  than  general  interest 
should  be  interpreted  by  an  experienced  hydrologist. 
The  wildlife  biologist  must  tell  the  hydrologist  the 
specific  objectives  or  intended  uses  of  these  data. 

Flow  data  are  usually  expressed  in  units  of  vol- 
ume per  unit  time,  for  example  cubic  feet  per  sec- 
ond or  cfs.  Table  1  contains  a  set  of  factors  for 
converting  different  streamflow  measurement  units. 


For  water  balance  and  certain  limnological  stud- 
ies, lake  volume  often  needs  to  be  determined.  As  a 
rough  estimate,  the  lake  surface  area  can  be  meas- 
ured from  a  map,  the  basin  shape  assumed,  and  the 
volume  computed  from  mensuration  formulas.  A  few 
common  lake  basin  shapes  and  their  volume  formu- 
las are  given  in  Figure  1 1 . 


More  accurate  volume  determinations  are  made 
from  detailed  surveys  and  soundings  of  lake  depths. 
Depths  are  best  measured  in  winter  when  a  standard 
land  survey  can  be  conducted  on  the  frozen  surface. 
Depths  are  easily  measured  through  holes  augered 
in  the  ice. 


Water  level  often  directly  relates  to  habitat  qual- 
ity. Periodic,  visual  readings  of  water  level  can  be 
taken  inexpensively  with  staff  gages  (Figure  12). 
Continuous  recording  of  water  level  requires  the  in- 
stallation of  a  water  level  recorder,  as  in  continuous 
streamflow  measurement. 


The  precipitation-evaporation  ratio  is  important 
in  determining  lake  characteristics  (Winter  1977). 
The  ratio  should  be  calculated  on  an  average  annual 
basis  from  reliable  data.  Precipitation  data  are  usually 
obtained  from  NWS.  Evaporation  data  may  also  be 
available  from  NWS  or  from  a  nearby  reservoir  oper- 
ator. Pan  evaporation  data  should  be  adjusted  to 
represent  free-water  evaporation. 


Hydrologic  Properties 


623 


GREEN 

RIVER 

3A3IN 

09241000   ELK  RIVER 

AT  CLARK 

,  CO 

LOCATION 

.— Lat  40° 

43'03", 

long  106 

°54'55' 

,  in  NWJNWJ 

sec. 27 

,  T.9  N. , 

R.85  W. , 

Routt  County,  Hyd 

rologic 

Unit 

14050001  , 

on  left 

bank  30 

ft  downstream  from 

bridge 

on  State 

Highway 

129,  0.8 

mi  north 

of  Clark, 

and 

2.0  mi  upstream 

from  Co 

ttonwood 

Gulch. 

DRAINAGE 

AREA.--206  mi2. 

PERIOD  OF  RECORD. - 

-May  1910  to  Sep 

tember 

192^ 

(publi: 

hed  as 

"near  CI 

ark"),  Ap 

ril  1930 

to  current  year. 

Monthly 

disch 

arge  only 

for  some 

periods 

,  publishec 

1  in  WSP 

1313. 

REVISED 

RECORDS. -- 

WSP  1733 

:   1956. 

GAGE. — Hater-stage 

recorder.   Datum  of  ga 

ge  is  7,267 

75  ft, 

(State  H 

ighway  Dc 

partment 

bench  mar 

k ) .   May 

191C 

to 

September  1922, 

nonrecording  ga 

ge  at  site 

30  ft  up 

stream 

at  datum 

0.15  ft 

lower.   A 

pr.  23,  1930,  to  S 

ept 

27, 

1934, 

water-sta 

ge  recor 

der  at  p 

resent 

si  te 

at  datum  0.15 

ft  lower 

REMARKS. 

--Estimate 

d  daily 

discharg 

es :  Nov 

.  2C 

to  Apr. 

7.  Records  goo 

i   except 

for  estimated  daily  dischar 

ges 

which 

are  poor. 

Diversions  abo 

ve  stat 

ion 

for  im 

gation 

of  abou 

l  230  acres  above 

and  about 

460  acres  below 

station.   Natural  flow 

of  stream  affected 

by  store 

ge  in 

Lester  Creek  Reservoir  (known  also 

as  Pearl 

Lake) , 

capacity,  5,660 

acre-f t 

since  1963  and 

Steamboat  L 

ake , 

capacity , 

23,060  acre-ft  since  1968. 

Several 

observations  of 

specific  conduc 

tance  and  water  ten 

perature  were 

obtained 

and  are  p 

ublished 

elsewhere 

in 

this 

repor 

t. 

AVERAGE 

DISCHARGE. 

--67  years,  339 

ft3 /s; 

245, 

600  acre 

-ft/yr 

EXTREMES 

FOR  PERIOD  OF  RECORD. --Ma 

ximum  c 

isch 

arge,  4, 

910  ft 

'/s,  May 

>3,  1984, 

gage  hei 

ght,  6.12 

ft;  minimum 

daily 

determine 

d,  22  ft 

/s,  Dec 

.  12,  1963, 

but  a  1 

esser 

discharge 

may  have 

occurred 

during  p 

eriods  of 

no 

gage- 

height  record  prior  to  1939- 

EXTREMES 

FOR  CURRENT  YEAR. 

— Peak  d 

ischarg 

es  e 

ibove  base  of  1 

,900  ft3/ 

5,  and  maximum  (*] 

D 

ischarge 

Gag 

e  height 

Discharge 

Gage 

height 

Date 

Time 

(ft3/s) 

(ft) 

Date 

Time 

(ft3/s) 

(ft) 

May   10 

1900 

2,000 

4.44 

June  8 

2300 

*3, 

180 

»5. 

36 

May   28 

2200 

2,300 

4.68 

Minimum  daily  d 

ischarge 

,  40  ft3 

/s,  Feb 

.  2C 

to  Mar. 

2. 

DISCHARGE, 

IN  CUBIC  FEET 

PER 

SECOND, 

WATER 

YEAR  OCTOBER  1984 

TO  SEPTEMBER  1985 

MEAI\ 

VALUES 

DAY 

OCT 

NOV 

DEC 

JAN 

FEB 

MAR 

APR 

MAY 

JUN 

JUL 

AUG 

SEP 

1 

121 

104 

72 

100 

60 

40 

226 

1030 

1200 

527 

226 

72 

2 

126 

98 

66 

100 

60 

40 

226 

1210 

1180 

501 

221 

75 

3 

118 

107 

70 

98 

60 

42 

225 

1380 

1220 

491 

211 

97 

14 

115 

95 

72 

100 

60 

45 

223 

1500 

1290 

483 

194 

86 

5 

124 

92 

74 

100 

60 

47 

223 

1630 

1330 

463 

185 

78 

6 

131 

104 

74 

98 

60 

50 

223 

1500 

1350 

437 

179 

72 

7 

123 

97 

76 

98 

58 

56 

223 

11)10 

1710 

398 

170 

70 

8 

113 

97 

78 

98 

56 

60 

224 

1580 

2280 

381 

161 

75 

9 

108 

93 

80 

96 

54 

64 

262 

1640 

2410 

377 

167 

71 

10 

104 

104 

80 

96 

50 

66 

258 

1720 

2080 

343 

156 

68 

11 

106 

109 

80 

96 

45 

68 

258 

1620 

1630 

329 

149 

67 

12 

104 

100 

82 

96 

45 

68 

329 

1300 

1410 

358 

181 

86 

13 

108 

102 

82 

96 

42 

70 

41  1 

1120 

1440 

343 

1it8 

72 

14 

111 

105 

84 

96 

42 

70 

500 

1000 

1470 

298 

135 

67 

15 

100 

92 

86 

94 

42 

70 

675 

1030 

1530 

266 

126 

64 

16 

97 

100 

88 

94 

42 

70 

817 

1  110 

1570 

251 

119 

65 

17 

96 

106 

90 

94 

42 

70 

89  3 

1  160 

1520 

233 

111 

62 

18 

1  1  1 

87 

92 

94 

42 

72 

1090 

1  150 

1370 

390 

1  12 

62 

19 

110 

89 

92 

94 

42 

74 

1240 

1140 

1280 

715 

115 

73 

20 

1  12 

90 

92 

90 

40 

76 

1  100 

1  170 

1220 

560 

107 

68 

21 

111 

90 

90 

86 

40 

78 

1060 

1280 

1210 

409 

101 

70 

22 

103 

80 

88 

84 

40 

80 

987 

1310 

1090 

374 

99 

73 

23 

108 

80 

86 

82 

40 

80 

964 

1340 

968 

364 

94 

77 

24 

108 

80 

88 

82 

40 

80 

953 

1500 

892 

408 

89 

72 

25 

97 

80 

94 

80 

40 

82 

789 

1700 

1120 

322 

85 

71 

26 

105 

80 

94 

74 

40 

84 

574 

i860 

1000 

302 

83 

71 

27 

106 

80 

92 

70 

40 

86 

569 

1930 

738 

288 

82 

79 

28 

103 

78 

92 

66 

40 

88 

708 

2000 

631 

251 

83 

92 

29 

108 

76 

94 

62 



90 

832 

1760 

583 

301 

81 

84 

30 

105 

74 

96 

60 



100 

890 

1590 

565 

273 

77 

68 

31 

106 

— 

98 

60 

— 

105 

--- 

1320 



248 

73 



TOTAL 

3398 

2769 

2622 

2734 

1322 

2171 

17952 

43990 

39287 

11684 

4123 

2207 

MEAN 

110 

92.3 

84.6 

88.2 

47.2 

70.0 

598 

1419 

1310 

377 

133 

73.6 

MAX 

131 

109 

98 

100 

60 

105 

1240 

2000 

2410 

715 

226 

97 

MIN 

96 

74 

66 

60 

40 

40 

223 

1000 

565 

233 

73 

62 

AC-FT 

6740 

5490 

5200 

5420 

2620 

4310 

35610 

87250 

77930 

23180 

8180 

4380 

CAL  YR  1984   TOTAL 

183267 

MEAN 

501 

MAX 

4090 

MIN 

53 

AC-FT   363500 

WTR  YR  1985   TOTAL 

134259 

MEAN 

368 

MAX 

2410 

MIN 

40 

AC-FT   266300 

Figure  9-     An  annual  summary  of  dailv  streamflow  data  collected  and  published  by  the  U.S.  Geological  Survey 
(1985). 


624 


Hydrologic  Properties 


Oct      Nov      Dec      J»n       Feb      Mar     Apr      May     Jun       Jul       Aug       Sep 


Figure  10.     An  example  of  an  annual  hydrograph 
based  on  mean  monthly  streamfiow  data,  Elk 
River  at  Clark,  CO,  Station  09241000  (U.S.  Geo- 
logical Survey  1985). 


r^flH  & 

* 

'  :* 

-"*- 

*V 

«3J0^ 

!  3.20  — 

fir 

i             i 

*'<' 

i  ... 

3.10^— 

3.00-^ 
2.90-^- 

1 

n 

2.80  — 
2.70^- 

it? 

pi 

** '  ZKi 

**N|r 

2.60  ~  9 

jMPm* 

*>■'    " 

m 

■■ 

.2.50  ^-M 

0  At\    "  •      ■ 

"'Jr      1 

'.<:'.  '-£L--&st&&!? 

. 

1 

2.30- — 
2.20^— 

?.10^- 
.'.00-^- 

1 

ll 

■<■/.*■ 

I 

1.90  -^  1 
1.80-^ 

^ 

# 

• 

~^»J*3>. 

pi  / 

,1.70  ^-J 
1.60-^ 
1.50^ —  1 

1.40 -^rl 

li 

Figure  12.     A  staff  gage  for  measuring  water 
level. 


Elliptic  depression 


Cone  frustrum 


Volume    :    3    (B^  B2+\JB,B2  ) 

where    B,  =  iir,2  ,    B,  =  iir,2 


Volume  =    9  n  abh 


Sphere  segment 


r  T=VJ 


Volume  =     6  (3a2+  3b2+hJ) 


Elliptic  sinusoid 

T" 


Volume  =  4  abD(l-fr) 


Figure  11.     Common  lake  basin  shapes  and  their  corresponding  volume  formulas. 

Hydrologic  Properties 


625 


Table  1.     Streamflow  unit  conversions. 


To  convert  from 

Multiply  by 

To  find 

acre-ft/day 

0.504167 

cfs 

acre-ft/day 

0.123349 

ha-m/day 

acre-ft/day 

0.325851 

million  gal./day 

acre-ft/mo  (28-day) 

0.018006 

cfs 

acre-ft/mo  (29-day) 

0.017385 

cfs 

acre-ft/mo  (30-day) 

0.016806 

cfs                                     ^^^^^ 

acre-ft/mo  (31  -day) 

0.016263 

cfs 

acre-ft/yr 
acre-in./hr 

0.001370 
2 

acre-in./hr 
acre-ft/day 

acre-in./hr 

1 .008333 

cfs 

acre-in./hr 

1 .027906 

ha-cm/hr 

cfs 

1 .98347 

acre-ft/day 

cfs 

55.53719 

acre-ft/mo  (28-day) 

cfs 

57.52066 

acre-ft/mo  (29-day) 

cfs 

59.50413 

acre-ft/mo  (30-day) 

cfs 

61 .48760 

acre-ft/mo  (31  -day) 

cfs 

723.96694 

acre-ft/yr 

cfs 

725.95041 

acre-ft/leap  year 

cfs 

0.99174 

acre-in./hr 

cfs 

60.45620/area  (a.) 

cm/day 

cfs 
cfs 

2.51 901 /area  (a.) 

cm/hr 

1.69901 

m3/min 

cfs 

0.02832 

m3/sec 

cfs 

0.00021424 

m'3/yr               „«______«__«__. 

cfs 
cfs 

6.66667 
448.83117 

yd3/min 
gal./min 

cfs 

538168.7 

Imp.  gal./day 

cfs 

373.73 

Imp.  gal./min 

cfs 

23.801 65/area  (a.) 

in./day 

cfs 

0.99174/area(a.) 

in./hr 

cfs 

28.31685 

L/sec 

cfs 

0.64632 

million  gal./day 

cfs 

50 

miner's  inches  (ID,  KS,  NE,  SD, 
ND,  NM,  UT,  WA,  S.  CA) 

cfs 

40 

miner's  inches  (AZ,  MT,  OR,  NV, 
N.  CA) 

cfs 

38 

miner's  inches  (CO) 

cfs 

2696.082 

tons  of  water  (50°F)/day 
cm/day 

m3/sec 

864/area  (ha) 

m3/sec 

315360/area  (ha) 

cm/yr 

m3/sec 

35.31467 

ft3/sec  (cfs) 

m3/sec 

8.640 

ha-m/day 

m3/sec 

3153.60 

ha-m/yr 

m3/sec 

3162.24 

ha-m/leap  yr          ^^^^^^ 

m3/sec 

86400010 

Uday 

m3/sec 

60000 

Umin 

m3/sec 
cm/day 

1000.00012 

Usee 

0.01654  x  area  (a.) 

cfs 

cm/day 

0.00116  x  area  (ha) 

m3/sec 

cm/hr 

0.39698  x  area  (a.) 

cfs 

cm/yr 

3.171E-06  x  area  (ha) 

m3/sec 

626 


Hydrologic  Properties 


Table  1.     Streamflow  unit  conversions  (concluded). 


To  convert  from 

Multiply  by 

To  find 

gal./day 
gal./min 

0.00112 
0.00442 
1.61301 

acre-ft/yr 

acre-ft/day 

acre-ft/yr 

gal./min 
gal./min 
gal./min 

0.00223 

0.00006309 

0.06309 

cfs 

m3/sec 

L/sec 

Imp.  gal./min 
million  gal./day 
million  gal./day 

0.00268 
3.06888 
1.54723 

cfs 
acre-ft/day 

ha-m/day 
ha-m/day 
ha-m/day 

8.10710 

0.11574 

115.74075 

acre-ft/day 

m3/sec 

L/sec 

ha-m/yr 
ha-m/yr 
ha-m/yr 

8.10710 

3.17E-04 

0.31710 

acre-ft/yr 
m3/sec 

in./day 

in./hr 

in./yr 

0.04201  x  area  (a.) 
1 .00833  x  area  (a.) 
0.000115  x  area  (a.) 

cfs 
cfs 
cfs 

L/sec 
L/sec 
L/sec 

0.864/area  (ha) 
0.036/area  (ha) 
315.36/area  (ha) 

cm/day 

cm/hr 

cm/yr 

Usee 
L/sec 
L/sec 

0.03531 
0.001 
15.85032 

cfs 

m3/sec 

gal./min. 

miner's  inches  (ID,  KS,  NE,  SD, 
ND,  NM,  UT,  WA,  S.  CA) 
miner's  inches  (AZ,  MT,  OR,  NV, 
N.  CA) 

0.020 
0.025 

cfs 
cfs 

miner's  inches  (CO) 

acre-ft 

acre-ft/mi.2 

0.026 
504.17 
0.04763 

cfs 

Kft3/sec-days 

cm  over  watershed 

acre-ft/mi.2 
acre-ft/mi.2 
cm  over  watershed 

0.04763 

0.01875 

20.99738 

ha-m/km2 

in.  over  watershed 

acre-ft/mi2 

cfs/mi2 
cfs/mi2 
cfs/mi2 

0.01093 
1.04132 
1.07851 

m3/sec/km2 
in./mo  (28  days) 
in./mo  (29  days) 

cfs/mi2 
cfs/mi2 
cfs/mi2 

1.11570 

1.15289 

13.57438 

in./mo  (30  days) 
in./mo  (31  days) 
in./yr  (365  days) 

cfs/mi2 
cfs/mi2 
cfs/mi2 

13.61157 

0.03719 

10.93320 

in./yr  (366  days) 

in./day 

L/sec/km2 

m3/sec/km2 

ha-m/km2 

in.  over  watershed 

91 .46457 
20.99729 
53.33333 

ft3/sec/mi2 
acre-ft/mi2 
acre-ft/mi2 

in.  over  watershed 
Kft3/sec-days/mi2 
Kft3/sec-days 
L/sec/km2 

0.02689 

37.19008 

0.00198 

0.09146 

Kft3/sec-days/mi2 
in.  over  watershed 
acre-ft 
ft3/sec/mi2 

Hydrologic  Properties 


627 


GROUNDWATER  FEATURES  AND  THEIR 
RELATIONSHIP  TO  WILDLIFE  HABITAT 

I  have  taken  the  liberty  in  this  section  to  discuss 
springs,  seeps,  and  related  water  features  in  an  over- 
all groundwater  context,  even  though  many  classify 
springs  as  surface  water  features.  Springs  and  seeps 
are  usually  an  expression  of  the  local  groundwater 
system. 

Wildlife  biologists  may  be  concerned  about  the 
reliability  of  wildlife  water  supplies  from  springs  and 
seep  areas.  Spatial  distribution  as  well  as  discharge 
of  individual  springs  or  seeps  would  be  of  interest 
for  inventory  purposes.  Measurement  of  spring  dis- 
charge is  done  by  the  same  basic  techniques  covered 
in  "Streamflow  Measurement  and  Characteristics" 
and  will  depend  on  size  and  physical  setting  of  the 
spring. 

Groundwater  levels  may  be  of  interest  to  the 
wildlife  biologist,  particularly  where  wildlife  water 
supplies  are  obtained  from  pumping  wells  or  artesian 
wells.  The  Water  Resources  Division  of  the  U.S.  Geo- 
logical Survey  operates  a  network  of  observation 
wells  around  the  U.S.  That  agency  should  be  con- 
tacted for  data  on  water  levels  and  water  quality. 
Nearly  all  completed  water  wells  may  be  used  for 
measuring  water  levels.  However,  the  well  log 
should  be  consulted  to  determine  the  depth  of  the 
well,  whether  or  not  packers  were  included  during 
well  construction,  and  any  other  construction 
details. 


STREAM  CHANNEL  STABILITY  AND 
SEDIMENT  TRANSPORT 

A  very  good  overview  of  stream  dynamics,  from 
a  hydrologic  and  geomorphic  point  of  view,  has 
been  prepared  by  Dr.  Burchard  Heede  of  the  U.S. 
Forest  Service  (Heede  1980).  This  is  highly  recom- 
mended reading  for  the  biologist  planning  to  engage 
in  stream  channel  stability  or  sediment  transport 
investigations,  or  riparian  improvement  projects. 

Stream  Channel  Stability 

The  channel  stability  evaluation  procedure  used 
most  often  by  land  management  agencies  is  that 
presented  by  Pfankuch  (1978).  An  evaluation  form 
for  this  method  is  presented  in  Figure  1 3.  Field  train- 
ing is  required  before  using  this  procedure  because 
of  its  subjectivity.  Furthermore,  the  summary  numer- 
ical rating  system  may  need  adjusting  to  fit  the  gen- 
eral region  of  interest. 


The  interpretation  of  the  results  depends  some- 
what on  the  use  of  the  ratings;  it  is  not  an  aquatic 
habitat  evaluation.  It  can  be  used  as  an  index  to  sedi- 
ment delivery  because  poor  stability  usually  indi- 
cates high  levels  of  sediment  transport  and  good  to 
excellent  stability  correlates  with  low  sediment 
transport. 


Standard  statistical  sampling  guidelines  should 
be  used  to  determine  sample  stratification  and  sam- 
ple size  (Platts  et  al.  1983). 


Suspended  Sediment  and  Bedload 

Sediment  transport  and  deposition  in  streams 
directly  influence  riparian  systems.  For  example, 
aquatic  habitat  condition  deteriorates  with  increasing 
suspended  sediment. 


Field  sampling  of  suspended  sediment  is  a  rela- 
tively easy  exercise  in  wadable  streams.  Williams  and 
Thomas  (1984)  discuss  the  collection  and  analysis  of 
sediment  data  from  a  land  management  perspective. 
Grab  samples  should  be  taken  with  a  DH-48  sampler 
(Figure  14).  Streamflow  must  be  measured  at  the 
time  of  sediment  sampling  to  facilitate  meaningful 
interpretation  of  these  data. 


Continuous  sediment  sampling  is  both  difficult 
and  expensive  and  should  be  left  to  experienced 
hydrologic  technicians. 


Total  sediment  in  a  stream  is  composed  of  both 
suspended  and  bedload  fractions.  Accurate  bedload 
sampling  is  a  difficult  procedure.  Fortunately,  the 
wildlife  biologist  seldom  has  need  for  such  data.  Guy 
and  Norman  (1970)  discuss  bedload  sampling 
procedures. 


MONITORING  HYDROLOGIC  VARIABLES 

Water  in  nature  is  highly  transient  in  time  and 
space.  Monitoring  of  hydrologic  variables  thus  re- 
quires some  special  considerations.  Ponce  (1980a, 
1980b)  and  Jackson  et  al.  (1985)  provide  excellent 
guidance  for  hydrologic  monitoring  in  land  manage- 
ment situations.  Those  references  should  be  con- 
sulted prior  to  engaging  in  any  monitoring  involving 
water  quantity  or  quality. 


628 


Hydrologic  Properties 


B-l  STREAM  REACH  INVENTORY  and  CHANNEL  STABILITY  EVALUATION 
BEACH  LOCATION i   Survey  Date Time Obs. 


Forest 


Streaa 


Rgr.  Dist. 
"P. W.I. 
W/S  No.  _■ 


Reabh  Description  4 
Other  Identification 


Key  j Stability  Indicators  by  Classes  (Fair  and  Poor  on  reverse  BJdTJ 


EXCELLENT 


Bank   slope   gradient  < 30%- 


No  evidence   of   past  or  any 
potential  for  future   aass 
wasting   into  channel. 


Essentially  absent  froa 
imaediate   channel  area. 


10 


9056+  plant  density.      Vigor 
and  variety  suggests  a 
deep,   dense,   soil  binding, 
root  mass . 


Anple  for  present  plus  soae 
increases.  Peak  flows  con- 
ta ined .   .  W/b  ratio   <7. 


6^36+  with  large,   angular 
boulders   12"+  numerous. 


Rocks  and  old   logs   firmly 
embedded.   Flow   pattern  with 
out  cutting  or  deposition. 
Pools  and  riffles  stable. 


HI 


(3) 


(2) 


(3; 


(1) 


(2) 


Little   or  none   evident. 
Infrequent  raw  banks  less 
than  6"   high  generally.  _ 


Little   or  no   enlargement 
of  channel  or  point  bars. 


Sharp  edges  and  corners, 
plane   surfaces  roughened. 


, ,      Surfaces  dull,   darkened,   or 
stained.   Gen,   not  "bright'*. 


12 


13 


Assorted  Bizes  tightly 
packed  and/or  overlapping. 


No  change  in  sizes  evident, 

Stable   materials  80-100%, 


14 


Less   than  5%  of   the   bottoa 
affected   by   scouring  and 
deposi  tion. 


Abundant .     Growth  largely 
15      moss-like,   dark  green,    per- 
I  ennial .    In  swift   water  too. 


(2) 


CO 


CO 


(1) 


(1) 


(2) 


CO 


(6) 


(1) 


EXCELLENT  COLUMN  TOTAL  •*■[    [       GOOD  COLUMN  TOTAL 
Add  values  in  each  column*  and  record  in  spaces  below.  Add  column  scores, 


GO0P 


Bank  slope  gradient  30-40%. 


Infrequent  and/or  very  small, 
Mostly  healed  over.     Low 
future   potential. 


Present  but  ssostly  small 
twigs  and  liabs. 


70-90%  density.  Fewer  plant 
species  or  lower  vigor 
suggests  a  less  dense  or 
deep  root  mass, 


Adequate.  Overbank  flows 
rare.  Width  to  Depth  (w/PJ 
ratio  8  to  15. 


40  to  65%,  mostly  small 
boulders  to  cobbles  6-12". 


Some  present,  causing  erosive 
cross  currents  and  minor  pool 
filling.  Obstructions  and 
deflectors  newer  and  less 
fim. 


Some,   intermittently  at 
outcurves  and  constrictions. 
Raw  bankB  «tay  be  up  to  12". 


Soae  new  increase  in  bar 
formation,  mostly  from 
coarse  gravels. 


Rounded  corners  and  edges, 
surfaces  smooth  and  flat. 


Mostly  dull,  but  may  nave  up 
to  35%  bright  surfaces. 


Moderately  packed  with 
some  ove r 1 apping. 


Distrioution  shift  slight. 
Stable  materials  50-60%. 


5-30^  affected.  Scour  at 
constrictions  and  wnere 
grades  steepen.  Some 
deposition  in  pools. 


Comnon,  Algal  forms  in  low 
velocity  &  pool  areas.  Moss 
here  too  and  swifter  waters. 


HE 


(6; 


CO 


(6) 


(2) 


(4) 


CO 


(8) 


(8; 


(2) 


(2) 


CO 


(8j 


(12; 


(2; 


+  F, 


+  P, 


Total  Reach  Score 
Adjective  ratings :<38=Excellent ,    39-76=Good,   77-114-=Fair,    11 


■Poor* 


•(Scores  above  may   be   locally  adjusted    Dy  Forest  Hydrologist) 

Rl-Form  25OO-5A  Rev. 1-75       Side   1 


Figure  13.     Stream  channel  stability  evaluation  form  used  with  the  evaluation  procedure 
presented  by  Pfankuch  (1978). 


Hydrologic  Properties 


629 


INVENTORY  DATA:   (observed  or  measured  on  this  date) 


Side  2 


Stream  Width ft.X  Ave. Depth ft.X  Ave. Velocity f/s= Flow  cfs 

Reach         Stream     rurbidity      Stream       Sinuosity 

Gradient %,  Order ,   Level  ,  Stage ,   Ratio  , 

Temperature  A^r 

°F  or  C  of:   Water ,  Others 


Key 


10 


11 


12 


13 


14 


15 


Stability  Indicators  by  Glasses 


FAIR 


Bank  slope  gradient  40-60%. 


Moderate  frequency  &  size, 
with  some  raw  spots  eroded 
by  water  during;  high  flows 


Present,  volume  and  size 
are  both  increasing 


30-7 OJb  density .  Lower  vigor 
and  still  fewer  species 
form  a  somewhat  shallow  and 
discontinuous  root  mass. 


Barely  contains  present 
peaks.  Occassional  overbank 
floods.  W/P  ratio  15  to  25. 


5.    V/l 


20  to  40%,   with  most   in 
the   3-6"   diameter  class. 


Moderately   frequent,    moder- 
ately unstable   obstructions 
&  deflectors   move   with  high 
water  causing  bank  cutting 
and  filling  of   pools. 


Significant.     Guts   12"-2^" 
high.   Root  mat  overhangs 
and  sloughing  evident, 


Moderate  deposition  of   new 
gravel  &  coarse   sand  on 
old  and  some  new  bars. 


Corners  &  edges  well  round' 
ed  in  two  dimensions. 


Mixture,    50-50%  dull  and 
bright.  ±15%  ie.    35-65%. 


Mostly  a   loose  assortment 
with  no  apparent  overlap. 


Moderate   change  in  sizes. 
Stable  materials  20-50%. 


30-50%  affected.    Deposits 
4  scour  at  obstructions, 
constrictions,   and   bends. 
Some   filling  of   pools. 


Present  but  spotty,  mostly 
in  backwater  areas.  Season- 
al  blooms  make  rocks  slick. 


FAD*  COLUMN  TOTAL' 


m 


(9) 


(6) 


(9) 


(3) 


[6) 


(6j 


(12; 


(12) 


(3) 


(3) 


(6) 


(12) 


(18) 


(3. 


POOR 


Bank   slope  gradient  6oJfr 


+  . 


Frequent  or  large,  causing 
sediment  nearly  yearlong  OR 
imminent  danger  of  same. 


Moderate  to  heavy  amounts, 
predominantly  larger  sizes, 


<50%  density  plus  fewer 
species  &   less  vigor  indi- 
cate poor,  discontinuous, 
and  shallow  root  mass. 


Inadequate.  Overbank  flows 
common.  W/D  ratio  >  25. 


<  20%  rock  fragments  of 
gravel  sizes,  1-3"  or  less. 


Frequent  obstructions  and 
deflectors  cause  bank  ero- 
sion yearlong.  Sediment 
traps  full,  channel 
migration  occurring. 


Almost  continuous  cuts, 
some  over  24"  high.  Fail- 
ure  of  overhangs  frequent. 


Extensive  deposits  of  pre- 
dominantly fine  particles. 
Accelerated  bar  development. 


Well  rounded  in  all  dimen- 
sions.  surfaces  smooth. 


Predominantely  bright,  65%+, 
exposed  or  scourea  surfaces. 


No  packing  evident.  Loose 
assortment,  easily  moved. 


Marked  distribution  change, 
Stable  materials  0-20%. 


More  than  50%  of  the  bottom 
in  a  state  of  flux  or 
change  nearly  yearlong. 


Perennial  types  scares  or 
absent.  Yellow-green,  short 
term  bloom  may  be  present. 


HI 


(12; 


(8) 


(12) 


CO 


(8) 


(6) 


(16) 


(16; 


CO 


CO 


(8) 


:i6j 


(24; 


C+; 


POOR  COLUMN  TOTAL- 


1. 
2. 

3. 
4. 


Size  Composition  of  Bottom  Materials  (Total  to  100%) 


Exposed  bedrock, 
Large  boulders,  3'+  Dia, 
Small  boulders,  1-3'..., 
Large  rubble ,  6" -12" 


Small  rubble,  3"-6 

6.  Coarse  gravel,  l"-3".. 

7.  Fine  gravel,  0.1-1" 
6.  Sand,  silt,  clay,  muck 


Figure  13-     Stream  channel  stability  evaluation  form  used  with  the  evaluation  procedure 
presented  by  Pfankuch  (1978)  (concluded). 


630 


Hydrologic  Properties 


Figure  14.     Model  DH-48  suspended-sediment  sampler. 


Hydrologic  Properties 


631 


LITERATURE  CITED 


BRAKENSIEK,  D.L.,  H.B.  OSBORN,  and  WJ.  RAWLS, 
coords.  1979-  Field  manual  for  research  in  agricul- 
tural hydrology.  Agriculture  Handbook  224.  U.S.  Dep. 
Agric.  550pp. 

BUCHANAN,  T.J.  and  W.P.  SOMERS.  1969.  Discharge 
measurements  at  gaging  stations.  Techniques  of 
Water-Resources  Investigations,  Book  3,  Chapter  A8. 
U.S.  Dep.  Inter.,  Geol.  Surv.  65pp. 

CUPLIN,  P.  and  B.P.  VAN  HAVEREN,  eds.  1979.  Instream 
flow  guidelines.  U.S.  Dep.  Inter.,  Bur.  Land  Manage. 
57pp. 

GUY,  HP.  and  V.W.  NORMAN.  1970.  Field  methods  for 
measurement  of  fluvial  sediment.  Techniques  of 
Water-Resources  Investigations,  Book  3,  Chapter  C2. 
U.S.  Dep.  Inter.,  Geol.  Surv.  59pp. 

HEDMAN,  E.R.  and  W.R.  OSTERKAMP.  1982.  Streamflow 
characteristics  related  to  channel  geometry  of  streams 
in  western  United  States.  U.S.  Dep.  Inter.,  Geol.  Surv. 
Water-Supply  Pap.  2193.  17pp. 

HEEDE,  B.H.  1974.  Velocity-head  rod  and  current  meter 
use  in  boulder-strewn  mountain  streams.  U.S.  Dep. 
Agric,  For.  Serv.  Res.  Note  RM-271.  4pp. 

.  1980.  Stream  dynamics:  An  overview  for  land 

managers.  U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep. 
RM-72.  26pp. 

JACKSON,  W.L.,  S.  HUDSON,  and  K.  GEBHARDT.  1985. 
Consideration  in  rangeland  watershed  monitoring. 
U.S.  Dep.  Inter.,  Bur.  Land  Manage.  Tech.  Note  369. 
Denver,  CO.  25pp. 


PFANKUCH,  DJ.  1978.  Stream  reach  inventory  and  chan- 
nel stability  evaluation.  U.S.  Dep.  Agric,  For.  Serv., 
Northern  Region,  Missoula,  MT.  26pp. 

PLATTS,  W.S.,  W.F.  MEGAHAN,  and  G.W.  MINSHALL 
1983  Methods  for  evaluating  stream,  riparian,  and 
biotic  conditions.  U.S.  Dep.  Agric,  For.  Serv.  Gen. 
Tech.  Rep.  INT- 138.  70pp. 

PONCE,  S.L.  1980a.  Statistical  methods  commonly  used  in 
water  quality  data  analysis.  U.S.  Dep.  Agric,  For.  Serv., 
Watershed  Systems  Development  Group,  Tech.  Pap. 
WSDG-TP-001.  144pp. 

.  1980b.  Water  quality  monitoring  programs.  U.S. 

Dep.  Agric,  For.  Serv.,  Watershed  Systems  Develop- 
ment Group,  Tech.  Pap.  WSDG-TP-002.  68pp. 

RANTZ,  S.E.  1982.  Measurement  and  computation  of 

streamflow:  Volumes  1  and  2.  U.S.  Dep.  Inter.,  Geol. 
Surv.  Water-Supply  Pap.  2175.  631pp. 

U.S.  GEOLOGICAL  SURVEY.  1985.  Water  resources  data 
for  Colorado,  Part  I — Colorado  River  Basin.  U.S.  Dep. 
Inter.,  Geol.  Surv. 

VAN  HAVEREN,  B.P.  1986.  Water  resource  measurements: 
A  handbook  for  hydrologists  and  engineers.  American 
Water  Works  Assoc,  Denver,  CO.  132pp. 

WILLIAMS,  O.  and  R.B.  THOMAS.  1984.  Guidelines  for 

collection  and  analysis  of  sediment  data.  Draft  report, 
U.S.  Dep.  Agric,  For.  Serv.,  Watershed  Systems  Devel- 
opment Group,  Fort  Collins,  CO.  108pp. 

WILM,  H.G.  and  H.C.  STOREY.  1944.  Velocity-head  rod 
calibrated  for  measuring  stream  flow.  Civil  Engineer- 
ing. l4(ll):475-476. 

WLNTER,  T.C.  1977.  Classification  of  the  hydrologic  set- 
tings of  lakes  in  the  north-central  United  States.  Water 
Res.  13(4):753-767. 


632 


Hydrologic  Properties 


30 

WATER 
QUALITY 

Paul  Cuplin 


U.S.  Bureau  of  land  Management 
Service  Center 
Denver,  CO  80225 


Editor's  Note:  Water  quality,  the  purity  of  water,  is 
a  concern  from  many  perspectives — public  health, 
aesthetics,  agriculture,  and  as  habitat  for  aquatic 
life.  This  chapter  focuses  on  water  quality  require- 
ments for  fishes  and  other  aquatic  vertebrates,  and 
field  methods  to  measure  or  monitor  these 
requirements. 


INTRODUCTION 

Water  availability  is  critical  to  almost  every  re- 
newable natural  resource  including  wildlife  and  veg- 
etation. Therefore,  information  on  all  of  the  water 
that  occurs  on  desert,  forest,  stream,  lake,  and  wet- 
land on  public  lands  is  of  importance  to  land  man- 
agement. Water  found  on  public  lands  represents  all 
of  the  major  water  systems,  i.e.,  marine,  estuarine, 
riverine,  lacustrine,  palustrine,  perennial  ice  or  snow, 
and  terrestrial  systems.  The  chemical  and. physical 
properties  of  water,  as  they  affect  "water"  as  a  habi- 
tat for  aquatic  organisms  and  as  a  substance  for  con- 
sumption by  wildlife  and  domestic  livestock,  are 
the  concerns  of , this  chapter. 

LAND-USE  ACTIVITIES  THAT  AFFECT 
WATER  QUALITY 

Human  activities  can  alter  baseline  water  chem- 
istry, resulting  in  detrimental  effects  on  aquatic  and 
terrestrial  organisms  that  drink  or  dwell  in  water. 
Land-use  activities  such  as  grazing,  mining,  timber 
harvest,  and  recreation  may  increase  sediment  pro- 
duction and  therefore  affect  water  quality.  Mining 
can  also  produce  "acid  mine  waste,"  making  the  pH 
of  water  too  acidic  for  any  organism  to  survive.  Her- 
bicide or  pesticide  spraying  or  the  dumping  of  toxic 
materials  can  significantly  alter  water  quality  and 
cause  high  mortality  rates  in  sensitive  aquatic  life.  To 
protect  aquatic  habitats  and  water  quality,  the  U.S. 
Environmental  Protection  Agency  (EPA)  established 
standards  for  the  introduction  of  toxic  materials  into 
water,  which  were  published  in  "Quality  Criteria 
for  Water"  (U.S.  Environmental  Protection  Agency 
1976). 

The  Water  Quality  Act  of  1965  required  States 
to  set  standards  for  releasing  materials  into  streams 
and  lakes.  The  Federal  Water  Pollution  Control  Act 
and  Amendments  of  1972  (Public  Law  92-500;  86 
Stat.  816),  as  amended  by  the  Clean  Water  Act  of 
1977  (Public  Law  95-217;  91  Stat.  1566),  are  also  perti- 
nent. See  33  U.S.C.,  Chapter  23: 

[Section  208  directs  that  the  Administrator  of 
the  Environmental  Protection  Agency  shall 
promulgate  comprehensive  regulations  in- 
tended to  control  water  pollution  sources,  in- 
cluding nonpoint  sources.  Under  the 
regulations  which  have  been  promulgated,  the 


Water  Quality 


633 


use  of  permits  is  required  for  any  continuing 
pollution  sources.  The  Federal  land  manage- 
ment practices  are  designed  to  prevent  degra- 
dation of  streams  and  lakes  on  the  lands 
under  their  jurisdiction.  ] 

(author's  emphasis) 
[Section  404  authorizes  the  Army  Corps  of 
Engineers  to  issue  permits  for  activities  which 
result  in  earth  disturbances  contributing  to 
stream  sedimentation  or  which  will  result  in 
decreased  stream  flows  because  of  water 
impoundment.] 


UNIQUE  PROPERTIES  OF  WATER 

Water,  a  universal  solvent,  dissolves  many  sub- 
stances, resulting  in  the  alteration  of  water  quality. 
Water  is  the  only  known  substance  which  allows  the 
solid  state  (ice)  to  float  on  the  liquid  state.  If  this 
process  were  not  possible,  lakes  would  freeze  from 
the  bottom  up,  eliminating  aquatic  life  annually. 
Some  deep  lakes  would  remain  frozen  forever.  The 
greatest  water  density  is  reached  at  4°C  (39.6°F), 
allowing  aquatic  life  to  survive  in  an  ice-covered  lake 
unless  light  penetration  is  prevented  for  long  pe- 
riods. Then  the  available  oxygen  may  be  completely 
consumed,  resulting  in  a  "winter  kill"  of  fish. 


WATER  QUALITY  DATA  SOURCES 

Many  sources  provide  water  quality  information. 
The  U.S.  Geological  Survey  (USGS)  annually  pub- 
lishes surface  water  reports,  by  State,  that  contain 
streamflow  and  chemical  analysis  data  for  major 
streams  where  sampling  stations  have  been  estab- 
lished. The  U.S.  Environmental  Protection  Agency 
maintains  two  computerized  data  bases:  (STORET) 
for  chemical  water  information  and  (BIOSTORET) 
for  biological  information  on  aquatic  life.  Data  are 
stored  geographically  by  latitude  and  longitude. 

State  water  quality  control  commissions  and 
State  fish  and  wildlife  agencies  are  also  sources  of 
chemical  and  biological  information  on  water 
quality. 


WATER  SAMPLING 

If  animals  become  distressed  from  being  in  or 
drinking  from  polluted  waters  on  public  lands,  then 
water  sampling  and  analysis  become  urgent  matters 
in  determining  the  cause  and  solution  to  the  prob- 
lem. Reduced  oxygen  or  introduced  toxic  materials 
are  common  causes.  Dissolved  oxygen  (DO)  can 
be  readily  analyzed  in  the  field  with  a  field  DO  kit.  If 
the  dissolved  oxygen  is  5  mg/1  or  higher,  the  prob- 
lem is  more  likely  to  be  a  toxic  material  problem 
than  a  lack  of  oxygen. 


At  least  a  1  -gallon  water  sample  is  collected  and 
carefully  labeled  as  to  time,  date,  location,  and  site 
conditions.  Water  chemical  analysis  is  then  done  by 
a  qualified  chemist.  The  chemist  must  be  informed  of 
the  most  likely  substance  causing  the  problem,  as 
the  chemist  cannot  conduct  a  thorough  analysis  on 
all  elements  which  might  be  present  in  the  water 
sample. 

Preferably,  water  quality  should  be  monitored 
on  site  since  many  parameters,  e.g.  temperature,  DO, 
and  pH,  will  be  altered  during  transportation  to  the 
laboratory.  If  monitoring  is  done  in  the  field,  samples 
need  only  be  collected  for  the  more  detailed  anal- 
yses such  as  trace  metals. 

Beyond  the  crisis  situation  described  above, 
what  chemical  and  physical  characteristics  of  water 
are  useful  for  land  management?  The  list  is  short:  pH, 
DO,  temperature,  alkalinity,  conductivity,  and  nu- 
trients. These  variables  aid  in  describing  habitat  con- 
ditions suitable  for  specific  aquatic  organisms.  In 
addition  to  these  variables,  others  may  be  of  interest 
or  importance.  The  most  important  are  listed  with 
references  where  more  detailed  information  can  be 
found. 

Water  samples  to  be  analyzed  by  a  chemist  (or 
placed  in  storage )  should  contain  identifying  infor- 
mation such  as  sample  or  station  number;  name  of 
water,  stream,  lake  or  reservoir;  time  and  date  of 
collection;  collector's  name  and  agency;  and  a  report 
about  the  chemical  analysis  done  at  the  collection 
site.  Because  storage  of  water  samples  in  certain 
containers  for  extended  periods  may  alter  water 
chemistry,  inquiries  as  to  sampling  procedures  and 
containers  should  be  made  from  local  and  regional 
water  quality  analysts  or  by  consulting  "Standard 
Methods  for  the  Examination  of  Water  and  Wastewa- 
ter" (American  Public  Health  Association  1976)  and 
"Methods  for  Chemical  Analysis  of  Water  and 
Wastes"  (U.S.  Environmental  Protection  Agency 
1979).  The  Federal  Register,  Vol.  44,  No.  244,  pp. 
75028-75052,  December  18,  1979,  describes  ap- 
proved test  procedures,  containers,  and  preservation 
and  holding  times. 

Sampling  Procedures 

Water  sampling  analysis  describes  the  water 
qualities  for  a  particular  location  on  a  special  date,  at 
a  designated  time.  This  is  sometimes  called  a  "grab 
sample"  which  may  be  more  meaningful  if  the  same 
site  is  sampled  over  a  period  of  time. 

Water  can  be  analyzed  with  a  field  chemical  kit 
at  the  sampling  site  or  transported  to  a  chemical 
laboratory  for  more  detailed  analyses.  The  field 
chemical  kit  can  be  used  to  analyze  temperature, 
turbidity,  pH,  hardness,  alkalinity,  sulfate,  ammonia, 
phosphate,  nitrate,  some  metals,  and  dissolved  gases. 


634 


Water  Quality 


In  fact,  dissolved  gases,  such  as  oxygen,  carbon  diox- 
ide, and  hydrogen  sulfide,  must  be  analyzed  at  the 
site.  (Special  care  is  required  to  avoid  altering  the 
levels  of  dissolved  gases  by  aeration. )  All  measure- 
ments should  be  recorded  carefully  in  a  field  note- 
book or  on  appropriate  forms. 

Water  Quality  Criteria 

Water  quality  criteria  specify  concentrations  of 
water  constituents  for  various  uses,  e.g.,  fish,  wildlife, 
recreation,  domestic  water  sources,  and  agriculture 
(irrigation  and  livestock).  Typically,  fish  and  wildlife 
criteria  are  the  most  sensitive.  Criteria,  derived  from 
scientific  research  and  observation,  are  established 
by  each  State  and  may  vary  regionally  as  well  as 
among  States. 

Water  quality  criteria  are  listed  alphabetically  in 
"Quality  Criteria  for  Water"  (U.S.  Environmental 
Protection  Agency  1976 )  with  comments  by  the 
American  Fisheries  Society  (Thurston  et  al.  1979). 
The  issue  of  water  quality  criteria  is  unsettled.  In 
most  cases,  establishing  water  quality  criteria  is  not 
desirable  or  feasible  at  the  local  level.  The  State 
water  quality  criteria  should  therefore  be  used 
where  applicable.  Additional  guidance  can  be  ob- 
tained from  the  literature. 

Water  Measurement  Terms 

Some  common  terms  used  in  water  quality  anal- 
ysis are  described  below. 


mg/1  —  milligrams  of  a  chemical  per  liter  of  water;  a 
way  to  express  the  concentration  of  chemi- 
cals; equivalent  to  parts  per  million  (ppm)  or 
micrograms  per  gram  (u-g/g)  or  milligrams 
per  kilogram  (mg/kg). 

u.g/1  =  micrograms  of  chemical  per  liter  of  water; 

equivalent  to  parts  per  billion  (ppb)  or  nano- 
grams per  gram  (ng/g)  or  micrograms  per 
kilogram  ( u-g/kg). 

LC     =  lethal  concentration;  usually  expressed  as 
a  concentration  of  a  toxicant  that  is  lethal 
( fatal )  to  a  fixed  percentage  of  the  organisms 
in  a  specified  time.  Thus  a  96-hour  LCs„  will 
be  lethal  to  50%  of  the  organisms  within 
96  hours. 


Water  quality  criteria  are  generally  listed  as 
dissolved  (d)  or  total  (t)  depending  on  the  applica- 
tion. This  is  particularly  important  for  metals  where 
the  dissolved  (ionic)  forms  can  be  quite  toxic, 
whereas  the  precipitated,  bound  forms  are  relatively 
nontoxic. 


Quality  Criteria  for  Selected  Substances 

The  following  are  criteria  and  wildlife  and  fish- 
ery considerations  for  some  of  the  more  important 
substances,  considering  that  water  quality  changes 
among  sites,  seasons,  and  even  days.  "EPA  criteria" 
refer  to  criteria  listed  in  "Quality  Criteria  for  Water" 
(U.S.  Environmental  Protection  Agency  1976). 

Alkalinity.  Alkalinity  is  the  sum  total  of  the 
components  in  the  water  that  tend  to  elevate  the  pH 
of  water  above  a  value  of  about  4.5.  It  is  measured 
by  titration  with  standardized  acid  to  a  pH  value 
of  about  4.5  and  expressed  as  mg/1  of  calcium 
carbonate.  Alkalinity  is  therefore  a  measure  of  the 
buffering  capacity  of  water.  The  EPA  criteria  for 
alkalinity  is  20  mg/1  or  more  as  CaCO^  for  fresh- 
water aquatic  life  except  where  natural  concen- 
trations are  less.  Thurston  et  al.  (1979:3)  say  the  EPA 
criteria  of  20  mg/1  are  not  valid.  Controversy  exists 
in  all  stated  criteria  because  of  the  great  variability 
in  regional  water  quality  and  the  aquatic  species  that 
inhabit  the  waters. 

Cadmium.    TTie  EPA  criteria  for  cadmium  are — 

Soft  Water 
0.4  u.g/1  for  cladocerans  and  salmonid  fishes 
4.0  (jig/1  for  other,  less  sensitive  aquatic  life 

Hard  Water 
2.0  u.g/1  for  cladocerans  and  salmonid  fishes 
12.0  u.g/1  for  other,  less  sensitive  aquatic  life 

Thurston  et  al.  (1979:51)  indicate  that  "Toxicological 
data  would  suggest  criteria  for  five  levels  of  hard- 
ness: 0-35,  35-75,  75-150,  150-300,  and  300  mg/1 
CaC03." 

Chlorine.  The  EPA  criteria  for  total  residual 
chlorine  are — 

2.0  u.g/1  for  salmonid  fish 
10.0  |xg/l  for  other  fresh-water  and  marine  organisms 

Thurston  et  al.  (1979:67-72)  feel  these  criteria  are 
not  acceptable  and  state,  "A  single  criterion  of  3- 
5  mg/1  (measured  by  amperometric  titration  in  con- 
junction with  polarograph)  is  more  appropriate  for 
the  protection  of  fresh-water  organisms." 

Chromium.  The  EPA  criterion  is  100  |JLg/l  for  fresh- 
water aquatic  organisms. 

Thurston  et  al.  (1979)  recommend  that  in  no  case 
should  the  criterion  be  higher  than  50  u.g/1  for  any 
aquatic  organisms. 

Conductivity.  Conductivity  is  a  numerical 
expression  of  the  ability  of  an  aqueous  solution  to 
carry  electric  current,  expressed  in  u.mhos/cm 


Water  Quality 


635 


(micromhos/cm)  at  25°  C.  Conductivity  can  be  used 
to  approximate  total  dissolved  solids  in  water  by 
using  the  formula:  Specific  conductance  x  (0.65  ± 
0.05)  =  mg/1  dissolved  solids.  The  formula  should 
be  verified  by  comparing  specific  conductance  in 
u,mhos  with  the  total  dissolved  solids  determined 
empirically  by  filter-evaporation  procedures. 

Copper.  The  EPA  criteria  for  fresh-water  and 
marine  aquatic  life,  0.1  X  a  96-hr  LCs„  value,  is 
determined  through  nonaerated  bioassay  using 
sensitive  aquatic  (resident)  species. 

Thurston  et  al.  (1979:100)  do  not  agree  with  this 
criterion  (which  is  10  times  higher — 0.01  to  0.1 — 
than  the  previous  criterion);  however,  they  do  not 
suggest  another  criterion. 

Cyanide.  The  EPA  criterion  is  5  u.g/1  for  fresh-water 
and  marine  aquatic  life  and  wildlife. 

Thurston  et  al.  (1979:100)  suggest  that — 

"The  criterion  for  cyanide,  presumably  an  up- 
per limit  of  total  cyanide  concentration  at  any 
time  or  place,  is  unacceptable.  A  limit  of  the 
concentration  of  free  cyanide  ( HCN  and  CN_ ) 
or  of  molecular  hydrogen  cyanide  (HCN),  with 
the  same  numerical  value  ( 5  u.g/1 )  can  be  ac- 
cepted as  a  reasonable  water  quality  criterion 
for  general  application  with  some  reservations. 
However,  the  . . .  criterion  evidently  is  a  limit 
for  total  cyanide  . . .  Such  a  criterion  can  leave 
no  sound  toxicological  basis,  because  only  free 
and  not  complexed  cyanide  has  been  shown 
to  be  toxic  at  concentrations  less  than  1  mg/1 
(1,000  u.g/1).  Moreover,  very  low  levels  of  mo- 
lecular HCN  can  now  be  reliably  determined 
by  any  of  several  proven  methods  . . ." 

Hardness.  Hardness  is  reported  as  an  equivalent 
concentration  of  calcium  carbonate  (CaCOs). 


Concentration  of  CaCO, 
(mg/1) 

Description 

0-75 
75-150 

soft  water 
moderately  hard  water 

150-300 
300  and  up 

hard  water 
very  hard  water 

The  EPA  (1976)  suggests  that  the  effects  of  hardness 
on  fresh-water  fish  and  other  aquatic  life  are  related 
to  the  ions  causing  the  hardness  rather  than  hardness 
itself. 

Iron.  The  EPA  criterion  is  1  mg/1  for  fresh-water 
aquatic  life. 

636  Water  Quality 


Thurston  et  al.  (1979)  suggest  the  criterion  for  iron 
is  too  high.  Lethality  for  aquatic  insects  has  been 
observed  at  320  |xg/l. 

Lead.  The  EPA  criterion  is  0.01  times  the  96-hour 
LC50  value,  using  the  receiving  or  comparable  water 
as  the  diluent  and  soluble  lead  measurement  (non- 
filterable  lead  using  0.45  micron  filter),  for  sensitive 
fresh-water  resident  fish. 

Mercury.  The  EPA  criterion  is  0.05  u.g/1  for  fresh- 
water aquatic  and  marine  aquatic  organisms. 

Nickel.  The  EPA  criterion  is  0.01  of  the  96-hour 
LC50  value,  for  fresh-water  marine  aquatic  organisms. 

Nitrates  and  Nitrites.  The  nitrogen  compounds  in 
natural  waters  are  derived  from  the  fixation  of 
atmospheric  nitrogen  or  from  pollution  sources. 
These  compounds  are  available  for  absorption  by 
bacteria,  which  produce  ammonia,  nitrite,  and 
nitrate.  Inorganic  forms  of  nitrogen  are  rapidly 
absorbed  and  concentrated  by  phytoplankton.  Little 
plant  growth  occurs  if  nitrate  nitrogen  is  below  0.3 
mg/1. 

The  EPA  (1976)  states — 

"It  is  concluded  that:  (1)  levels  of  nitrate  nitro- 
gen at  or  below  90  mg/1  would  have  no  ad- 
verse effects  on  warm-water  fish  ...  (2)  nitrite 
nitrogen  at  or  below  5  mg/1  should  be  protec- 
tive of  most  warm-water  fish  . . .  and  (  3 )  nitrite 
nitrogen  at  or  below  0.06  mg/1  should  be  pro- 
tective of  salmonid  fishes  . . .  These  levels  either 
are  not  known  to  occur  or  would  be  likely  to 
occur  in  natural  surface  waters.  Recognizing 
that  concentrations  of  nitrate  or  nitrite  that 
would  exhibit  toxic  effects  on  warm  or  cold 
water  fish  could  rarely  occur  in  nature,  restric- 
tive criteria  are  not  recommended." 

Oil  and  Grease.  The  EPA  (1976)  suggests,  for 
aquatic  life,  that  levels  of  individual  petrochemicals 
in  the  water  should  not  exceed  0.01  of  the  lowest 
continuous  flow  96-hour  LC50  value,  since  each 
species  has  demonstrated  high  susceptibility  to  oils 
and  petrochemicals. 

Oxygen,  Dissolved.  The  EPA  criteria  for  dissolved 
oxygen  (DO)  for  fresh-water  aquatic  life  is  5  mg/1. 
The  amount  of  oxygen  that  will  dissolve  in  water  is 
affected  by  temperature,  elevation,  and  total 
dissolved  solids,  although  certain  desert  fishes  such 
as  pupfish  can  survive  low  oxygen  tension,  e.g.,  0.1  to 
0.4  mg/1. 

pH.  The  abbreviation  "pH"  is  a  measure  of  the 
hydrogen-ion  activity  in  a  water  sample.  It  is 
mathematically  related  to  hydrogen-ion  activity 
according  to  the  expression:  pH  =  -  logi()[H  +  ], 


where  [H  +  ]  is  the  hydrogen-ion  activity.  Stream 
water  in  areas  not  influenced  by  pollution  generally 
has  a  pH  range  between  6.5  and  8.5,  which  is 
acceptable  for  fish.  The  EPA  criterion  is  6.5  to  90 
for  fresh-water  aquatic  life. 

Acid  rain,  the  product  of  fossil  fuel  combustion, 
produces  sulfur  dioxide  and  other  oxides  which 
combine  with  water  to  form  acids.  Water  bodies 
with  low  buffering  capacity  (i.e.,  low  alkalinity)  can- 
not counteract  acid  rain,  and  the  pH  level  lowers 
to  a  lethal  level  of  less  than  4. 

Low  pH  caused  by  acid  precipitation  has  elimi- 
nated fish  populations  in  many  lakes  in  the  Adiron- 
dack Mountains  of  New  York,  the  LaCloche 
Mountains  of  Ontario,  and  the  lakes  and  streams  of 
southern  Norway  and  western  Sweden  ( Fritz  1980 ). 

Phosphates.  Phosphates  occur  in  water  as  a  result 
of  leaching  from  minerals  or  as  one  of  the  stabilized 
products  of  decomposition  of  organic  matter. 
Phosphorus  also  occurs  as  orthophosphate  from  land 
agricultural  application  and  is  essential  to  the  growth 
of  organisms.  It  can  be  the  nutrient  that  limits 
productivity  of  a  water  body.  The  EPA  (1976) 
discusses  effects  of  phosphorus  concentrations  on 
aquatic  life  but  does  not  set  any  criteria  for  fresh- 
water fish  or  aquatic  life. 

Polychlorinated  Biphenyls  (PCB).  The  EPA 

criterion  is  0.001  u.g/1  for  fresh-water  and  marine  life 
and  for  the  consumers  thereof. 

Selenium.  The  EPA  criterion  is  0.01  mg/1  of  the  96- 
hour  LCS()  value,  as  determined  by  bioassay  using 
sensitive  resident  species  for  marine  and  fresh-water 
aquatic  life. 

Silver.  The  EPA  criterion  is  0.01  mg/1  of  the  96-hour 
LCso  value,  as  determined  by  bioassay  using  sensitive 
resident  species  for  marine  and  fresh-water  aquatic 
life. 

Solids  (suspended,  settleable)  and  Turbidity. 

Suspended  or  settleable  solids  include  both  inorganic 
and  organic  materials.  Turbidity  is  an  expression  of 
the  optical  property  that  causes  light  to  be  scattered 
and  absorbed  rather  than  transmitted  in  straight  lines 
through  water.  The  EPA  criterion  for  both  solids 
(suspended,  settleable)  and  turbidity  for  fresh-water 
fish  and  other  aquatic  life  is  that  "settleable  and 
suspended  solids  should  not  reduce  the  depth  of  the 
compensation  point  for  photosynthetic  activity  by 


more  than  10  percent  from  the  seasonally 
established  norm  for  aquatic  life." 

Thurston  et  al.  (1979)  disagree  with  this  criterion 
because  "the  criterion  treats  solids  and  turbidity  as 
synonymous,  which  they  are  not."  They  recommend 
two  sets  of  criteria:  residue  (solids)  measured  in 
mg/1  and  turbidity  measured  in  Nephelometric  Tur- 
bidity Units  (NTU's).  Both  sets  of  measurements 
are  to  be  done  in  accordance  with  "Standard  Meth- 
ods for  the  Examination  of  Water  and  Wastewater 
(American  Public  Health  Association  et  al.  1976). 

Commercial  turbidimeters  measure  light  transmitted 
in  NTU's.  Turbidity  criteria  vary  regionally  and  may 
depend  on  the  type  of  fishery  of  concern,  i.e.  warm- 
or  cold-water. 

Sulfide  (Hydrogen  Sulfide).  The  EPA  criterion  is 
2  u.g/1  undissociated  H2S  for  fish  and  aquatic  life. 

Tainting  Substances.  Tainting  substances  can 
produce  undesirable  flavor  in  the  edible  portions  of 
fish  or  other  aquatic  vertebrates.  Diesel  oil  in  very 
small  amounts  will  taint  fish  flesh. 


WATER  QUALITY  BASELINE  DATA 
COLLECTION— PROJECT  PLANNING 

Land-use  projects  having  major  impacts  on 
water  quality  require  special  attention  in  establishing 
baseline  pre-project  water  quality.  An  example  of 
gathering  water  quality  data  before  surface  mining  is 
displayed  in  Table  1  (Harris  et  al.  1983).  The  field  and 
laboratory  analyses  are  divided  logically.  Baseline 
physical  and  chemical  data  will  provide  a  basis  for 
water  quality  monitoring  during  the  lifetime  of  the 
land-use  project. 


WATER  QUALITY  MONITORING  PROGRAMS 

Water  quality  monitoring  can  be  related  to  man- 
agement objectives  on  a  long-term  basis.  Methods 
for  testing  the  statistical  significance  of  change  in 
water  data  collected  over  a  period  of  time  are  out- 
lined in  several  publications.  Two  publications  pro- 
viding statistical  analysis  procedures  for  field  use  are 
"Water  Quality  Monitoring  Programs"  (Ponce  1980) 
and  "Stream  Monitoring  Techniques"  (Armour  et 
al.  1983). 


Water  Quality 


637 


Table  1.     Water  quality  and  bottom-sediment  parameters  measured  in  the  Tyro  Creek  watershed,  September 
1981  through  September  1982  (from  Harris  et  al.  1983). 


Field  Determinations1 

Laboratory  Analyses1 

Bottom  Sediments 

Discharge  (cfs) 

Suspended  sediment 

Sieve  analyses 

Temperature  (°C) 

Calcium  (Ca) 

Trace  metals  and 

pH  (as  units) 

Chloride  (CI) 

nutrients  (mg/kg) 

Alkalinity  (as  HC03) 

Fluoride  (F) 

Arsenic  (As) 

Specific  conductance 

Potassium  (K) 

Cadmium  (Cd) 

(|jirmhos/cm) 

Magnesium  (Mg) 

Chromium  (Cr) 

Color  (Pt-Co) 

Sodium  (Na) 

Cobalt  (Co) 

Turbidity  (JTU) 

Nitrate  (N)  (dissolved) 

Copper  (Cu) 

Dissolved  oxygen 

Nitrate  (N03) 

Iron  (Fe) 

Silica  (Si02) 

Lead  (Pb) 

Sulfate  (S04) 

Mercury  (Hg) 
Selenium  (Se) 

Trace  metals  (u.g/1) 

Strontium  (Sr) 

Arsenic  (As) 

Vanadium  (V) 

Cadmium  (Cd) 

Zinc  (Zn) 

Chromium  (Cr) 

Chloride  (CI) 

Cobalt  (Co) 

Phosphate  (P04) 

Copper  (Cu) 

Nitrate  (N03) 

Iron  (Fe) 

Lead  (Pb) 

pH  (as  units) 

Vanadium  (V) 

Manganese  (Mg) 

Mercury  (Hg) 

Selenium  (Se) 

Strontium  (Sr) 

Zinc  (Zn) 

All  measurements  should  be  taken  in  mg/l  unless  otherwise  indicated 


LITERATURE  CITED 


AMERICAN  PUBLIC  HEALTH  ASSOCIATION.  1976. 

Standard  methods  for  the  examination  of  water  and 

wastewater,  15th  edition.  Am.  Public  Health  Assoc. 

New  York,  NY.  1134pp. 
ARMOUR,  C.L.,  K.P.  BURNHAN,  and  W.S.  PLATTS.  1983. 

Field  methods  and  statistical  analysis  for  monitoring 

small  salmonid  streams.  U.S.  Dep.  Inter.,  Fish  and 

Wildl.  Serv.  FWS/OBS-83/33.  200pp. 
FRITZ,  E.S.  1980.  Potential  impacts  of  low  pH  on  fish  and 

fish  populations.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv. 

Biological  Services  Program,  National  Power  Plant 

Team,  FWS/OSS-80/40.2.  I4pp. 
HARRIS,  C.H.,  P.E.  O'NEIL,  R.V.  CHANDLER,  M.F.  METEE, 


and  E.J.  MCCULLOUGH.  1983.  Biological  and  hydro- 
logical  impacts  of  surface  mining  for  federal  minerals 
on  the  Tyro  Creek  Watershed,  Alabama,  Phase  1. 
Premining-aquatic  baseline  information.  Contract  per- 
formed for  U.S.  Dep.  Inter.,  Bur.  Land  Manage.  98pp. 

PONCE,  S.L.  1980.  Water  quality  monitoring  programs.  U.S. 
Dep.  Agric,  For.  Serv.,  3825  E.  Mulberry  St.,  Ft.  Col- 
lins, CO  80524.  66pp. 

THURSTON,  R.V.,  R.C.  RUSSON,  CM.  FETTEROLF,  Jr.,  T.A. 
EDSALL,  and  Y.M.  BARBER,  Jr.  1979.  A  review  of  the 
EPA  Red  Book:  Quality  Criteria  for  Water.  Am.  Fish. 
Soc.  Bethesda,  MD.  313pp. 

U.S.  ENVIRONMENTAL  PROTECTION  AGENCY.  1976. 
Quality  criteria  for  water.  U.S.  Environmental  Protec- 
tion Agency,  Washington,  DC  20460.  501pp. 

.  1979.  Methods  for  chemical  analysis  of  water  and 

wastes.  EPA-600/4-79-020.  Washington,  DC. 


638 


Water  Quality 


31 
VEGETATION 


Bertin  W.  Anderson  and  Robert  D.  Ohmart 

Center  for  Environmental  Studies 
Arizona  State  University 
Tempe,  AZ  85287 


Editor's  Note:  Although  vegetation  is  widely  recog- 
nized as  one  of  the  most  important  determinants  of 
wildlife  abundance  and  distribution,  most  systems 
for  measuring  and  classifying  vegetation  have  been 
developed  for  other  purposes.  Development  of  vege- 
tation measurements  for  wildlife  purposes  has 
mostly  focused  on  measuring  forage  supplies  for 
big  game  and  other  herbivores.  Such  techniques  are 
well-described  in  the  literature  and  are  discussed 
briefly  in  Chapter  25,  Ungulates. 

This  chapter  focuses  on  vegetation  measurement 
technicpies  that  have  been  found  useful  for  ( 1 ) 
evaluating  habitat  for  vertebrate  communities,  (2) 
predicting  presence  or  abundance  of  individual 
species  for  a  wide  variety  of  vertebrates,  and  (3) 
predicting  species  richness  of  habitat  areas.  Because 
of  the  diversity  of  possible  approaches,  the  authors 
have  not  attempted  to  provide  a  thorough  review  of 
all  techniques  and  their  relative  strengths.  Rather, 
they  have  described  one  system  that  is  reasonably 
efficient  and  has  worked  well  for  them.  Biologists 
will  undoubtedly  need  to  modify  systems  to  meet 
the  needs  of  their  inventory  or  monitoring  pro- 
grams. Nevertheless,  the  attributes  described  in  this 
chapter  should  be  useful  predictors  of  wildlife  habi- 
tat quality  in  a  wide  variety  of  habitats,  and  the 
measurement  techniques  described  should  be  usable 
or  modifiable  for  a  diversity  of  problems. 


INTRODUCTION 

This  chapter  describes  attributes  of  plant  com- 
munities that  we  found  useful  in  evaluating  habitat 
for  vertebrate  wildlife.  We  describe  methods  for 
measuring  these  attributes  and  show  how  these 
measurements  may  be  useful  in  predicting  species 
presence  or  absence,  species  abundance,  and  wildlife 
species  richness  in  various  communities.  We  devel- 
oped these  primarily  for  riparian  habitat,  but  the 
general  procedures  should  be  applicable  to  other 
habitats.  This  is  not  intended  to  be  an  exhaustive 
catalogue  of  all  or  even  a  majority  of  the  methods 
available,  nor  a  critical  review  of  the  numerous 
methods  for  quantifying  vegetation  attributes.  Re- 
views of  methods  for  measuring  vegetation  can  be 
found  in  U.S.  Department  of  the  Interior,  Bureau  of 
Land  Management  ( 1985);  Daubenmire  (1968); 
Mueller-Dombois  and  Ellenberg  (1974);  and  other 
standard  texts. 

One  of  the  fundamental  reasons  for  measuring 
attributes  of  a  plant  community  is  to  classify  the 
stand  under  consideration.  Classification  facilitates 
communication  by  reducing  the  verbage  required  to 
provide  someone  with  a  clear  and,  hopefully,  unam- 
biguous mental  picture  of  the  community  being  con- 
sidered. We  think  classifying  and  describing  plant 
communities  are  virtually  synonymous. 


Vegetation 


639 


In  our  experience,  time  and  personnel  for  meas- 
uring vegetation  are  always  limited.  Therefore,  we 
have  developed  a  reasonably  small  number  of  field 
methods  that  can  be  done  relatively  quickly  and, 
from  which,  other  plant  community  descriptors  can 
be  generated.  For  example,  we  measure  foliage  den- 
sity in  the  field  and,  from  this,  calculate  patchiness  in 
the  horizontal  dimension,  foliage  height  diversity, 
and  foliage  density  at  various  vertical  layers  (i.e., 
ground,  shrub,  canopy).  We  describe  these  methods 
in  this  chapter.  Other  sets  of  more  traditional  meth- 
ods can  lead  to  the  accumulation  of  data  equally 
useful  in  describing  a  vegetation  community,  but 
their  use  will  probably  be  more  costly  in  terms  of 
time  and  personnel. 

In  classifying  (describing)  vegetation  communi- 
ties, a  two-  or  three-dimensional  approach  should 
be  considered.  In  general,  the  physiognomy  or  struc- 
ture of  the  vegetation  represents  two  dimensions. 
For  example,  a  given  stand  of  vegetation  varies  in 
vertical  and  horizontal  space.  Variation  in  the  verti- 
cal dimension,  whether  single  or  multilayered,  is 
particularly  useful  in  describing  the  stand.  Similarly, 
the  floristics,  i.e.,  species  composition  of  a  stand,  is 
often  invaluable  in  describing  that  stand.  If  structure 
and  floristics  can  be  relatively  quickly  and  easily 
quantified,  limits  of  a  vegetation  type  can  be  clearly 
defined.  Standard  texts  describing  methods  in  syne- 
cology  (Daubenmire  1968)  or  in  mapping  vegetation 
(Kuchler  1967)  should  be  consulted  for  more  details 
on  quantifying  vegetation  variables  and  classifying 
vegetation. 

The  same  characteristics  used  in  quantitatively 
describing  a  stand  of  vegetation  can  be  used  in  de- 
veloping predictive  capabilities  relative  to  the  resi- 
dent wildlife.  We  describe  reasonably  fast  and 
accurate  field  methods  for  quantitatively  classifying 
vegetation  and  correlating  wildlife  associations  with 
various  attributes  of  vegetation  communities. 


VEGETATION  ATTRIBUTES 

Simple  Basic  Variables 

We  refer  to  foliage  density,  plant  species  com- 
position, and  fruit  production  as  simple,  basic  varia- 
bles because  they  are  usually  quantified  in  the  field. 

Foliage  Density.   Foliage  density  refers  to  the 
amount  of  green  foliage  present  or  to  the  amount  of 
leaf-bearing  stems  and  leaves  per  unit  area.  (Foliage 
density  should  not  be  confused  with  "plant  density" 
or  "density,"  which  is  the  number  of  plants  per  unit 
area.)  Usually  foliage  density  is  measured  at  various 
vertical  levels  (e.g.,  every  yard  or  0.1  yard  [meter  or 
0.1  meter]).  Foliage  density  measurements  taken  in 
summer  may  be  useful  in  describing  the  foliage 
density  in  winter,  in  terms  of  the  relative  amounts  of 


leaf-bearing  stems  and  leaves  present,  thus  negating 
the  need  for  remeasurements  in  winter.  This 
procedure  is  most  valid  in  areas  dominated  by  trees 
or  annuals. 

Species  Composition.  This  can  be  easily 
determined  by  counting  individuals  of  each  tree/ 
shrub  species  present.  This  is  not  as  simple  as  it 
sounds;  size  classes  must  be  considered.  Even  then, 
two  trees  of  the  same  height  and  species  can  be 
quite  different.  The  general  health  of  a  tree  or  tree 
density  can  affect  general  structure.  Thus  some 
thought  must  be  given  to  what  will  actually  be 
counted. 

Fruit  Production.   In  stands  of  vegetation  that 
produce  fruit,  especially  fruits  that  are  sought  by 
wildlife,  it  may  be  useful  to  obtain  some  idea  of  the 
total  fruit  produced.  We  have  done  this  when  the 
correlation  between  number  of  trees  present  and 
fruit  production  is  rather  poor.  For  example, 
mistletoe  {Phoradendron  californicum )  along  the 
Colorado  River  parasitizes  honey  mesquite  (Prosopis 
glandulosa)  more  frequently  than  other  tree 
species.  However,  the  proportion  of  trees  parasitized 
varies  widely  from  stand  to  stand.  We  therefore 
obtained  estimates  of  the  number  of  mistletoe 
clumps  in  a  given  stand.  Similarly,  pod  production  by 
individual  honey  mesquite  trees  varies  widely 
between  stands;  thus,  there  is  a  poor  correlation 
between  number  of  trees  present  and  production  of 
pods.  This  is  also  true  of  pod  production  by 
screwbean  mesquite  (P.  pubescens).  At  higher 
elevations,  some  shrub  species  may  be  important  for 
berry  or  seed  production  for  wildlife.  These  shrub 
species  should  be  noted  and  given  important 
management  concern. 


Composite  or  Derived  Variables 

From  these  foliage  density  and  species  composi- 
tion measurements,  a  host  of  other  variables  can  be 
derived.  Foliage  height  diversity  (FHD)  is  calculated 
from  the  foliage  density  measurements  taken  at  var- 
ious vertical  planes.  Horizontal  foliage  diversity 
( patchiness )  can  be  determined  by  considering  foli- 
age density  variation  in  the  horizontal  plane.  In  verti- 
cally complex  (i.e.,  multilayered)  communities,  it 
may  be  desirable  to  calculate  patchiness  for  each  of 
the  vertical  layers,  with  total  patchiness  equal  to  the 
sum  of  patchiness  in  the  vertical  layers.  The  number 
of  vertical  layers  recognized  is  an  arbitrary  matter, 
as  is  the  size  of  the  area  being  called  a  patch.  The 
concept  of  FHD  is  depicted  in  Figure  1  and  patchi- 
ness in  Figure  2. 

Tree  density  estimates  can  be  expressed  as  the 
number  of  each  species  per  unit  area  or  as  the  pro- 
portion of  the  total  tree  species  present.  When  meas- 
uring foliage  density,  one  may  record  the  plant 


640 


Vegetation 


Figure  1.     Diagram  of  foliage  diversity  in  the  vertical  plane.  The  stand  depicts  an  area  of  at  least  25  a. 
(10  ha). 


C? 


Q2  G> 


O 


I  -J  \K    \w 


L,  L^Ok^  JWfi'a* 


gggSj^  i  ^ 


/ 


O 


Figure  2.     Diagram  of  foliage  diversity  (patchiness)  in  the  horizontal  plane  at  each  of  three  vertical  layers. 
The  blocks  represent  patches  of  roughly  5  a.  (2  ha). 

Vegetation  641 


species  contributing  to  the  foliage  density  at  each 
point  where  measurements  are  taken.  The  density  of 
each  tree/shrub  species  can  then  be  given  for  each 
vertical  layer  or  in  terms  of  the  total  foliage  density 
for  the  entire  area. 

Usefulness  of  the  Variables 

We  found  that  the  counts  of  tree/shrub  species 
were  particularly  useful  in  predicting  the  presence 
and  densities  of  many  rodent  and  bird  species  (An- 
derson and  Ohmart  1984a,  in  press-b;  Rice  et  al. 
1983,  1984).  We  found  that  patchiness  and  FHD 
have  useful  predictive  value,  but  are  not  as  good  a 
predictor  as  tree/shrub  counts  (Anderson  and 
Ohmart  1984a,  in  press-b;  Rice  et  al.  1983,  1984). 
Mistletoe  counts  were  associated  with  the  presence 
of  frugivorous  birds  (Anderson  and  Ohmart  1978). 
In  general,  foliage  density  at  low  levels  had  a  high 
association  with  lizards  (Anderson  and  Ohmart 
1982)  and  also  some  rodents  and  birds  (Anderson 
and  Ohmart,  in  press-b ).  At  various  times  and  places, 
desert  mule  deer  (Odocoileus  hemionus)  were  asso- 
ciated with  screwbean  and  honey  mesquite,  but  foli- 
age density  and  diversity  were  also  important  in 
explaining  their  habitat  selection  (Haywood  et  al. 
1984). 

From  this  brief  discussion,  it  seems  that  all  of 
the  vegetation  variables  given  may  be  important  to 
at  least  some  species  of  wildlife.  We  found  consider- 
able seasonal  variation  in  habitat  selection  within  a 
group,  i.e.,  birds  (Anderson  and  Ohmart  1984a). 
Rodents  used  the  vegetation  differently  in  any  given 
season  than  the  majority  of  bird  species.  The  point 
is,  there  is  no  a  priori  means  of  selecting  a  single 
attribute,  or  even  a  few  vegetation  attributes,  that 
will  be  adequate  for  predicting  wildlife  use  of  a 
habitat. 

Findings  from  our  research  lead  us  to  offer  a 
precautionary  note  to  others  planning  habitat  stud- 
ies, particularly  studies  of  avian  habitat  use.  From 
most  literature  sources,  we  would  have  been  led  to 
de-emphasize  tree  species  measurements  completely. 
We  have  found  many  habitat  attributes  to  be  crucial 
to  the  prediction  of  habitat  selection  of  at  least  some 
bird  species.  Had  only  a  small  set  of  attributes  been 
collected,  analyses  would  have  been  successful 
enough  that  one  would  have  been  tempted  to  con- 
clude erroneously  that  the  important  habitat  attri- 
butes and  selection  processes  for  the  entire 
community  had  been  captured  in  the  study. 

Studies  of  habitat  use  should  be  designed  to 
sample  the  annual,  seasonal,  and  spatial  variation  in 
habitat  use  of  the  animal  community.  Within  the 
limits  of  biological  reality,  biologists  should  measure 
as  comprehensive  a  set  of  attributes  as  time  and 
money  allow  and  should  avoid  measuring  only  a  few 
attributes  based  on  knowledge  of  a  few  species,  data 


from  other  areas,  or  information  from  the  literature. 
Studies  providing  data  for  environmental  manage- 
ment plans,  especially  habitat  modification  projects 
with  impacts  on  substantial  portions  of  the  animal 
communities,  should  be  based  on  data  sets  adequate 
for  a  thorough  description  of  the  system.  If  we  had 
studied  selected  subsets  of  our  species,  transects, 
time  periods,  or  habitat  attributes,  we  could  have 
undoubtedly  produced  results  supporting  quite  a 
wide  range  of  invalid  proposals.  Such  actions  are  not 
ecologically  wise  nor  do  they  lead  to  effective 
management. 


MEASURING  VEGETATION  ATTRIBUTES 

Establishing  Study  Sites 

Lines  or  transects  are  established  by  cutting 
swaths  3  ft  ( 1  m )  wide  through  the  middle  of  stands 
encompassing  at  least  25  a.  ( 10  ha),  with  dimensions 
of  at  least  2,461  ft  (750  m)  long  by  164  ft  (50  m) 
wide.  Small  patches  (2.5  a.  [1  ha])  of  vegetation,  dif- 
fering in  species  composition  or  structure  from  the 
major  type  in  the  stand,  should  be  bisected  by  the 
transect  at  right  angles  whenever  possible.  In  no 
case  should  a  transect  be  situated  so  that  vegetation 
differing  from  the  stand  as  a  whole  is  paralleled  by 
the  transect  at  a  distance  closer  than  49  ft  ( 1 5  m ). 
When  vegetation  is  very  much  unlike  the  stand  as  a 
whole,  it  has  an  unduly  large  effect  on  the  vege- 
tation analysis  and  on  estimates  of  bird  and  rodent 
densities. 

Semipermanent  markers  should  be  placed  at  the 
beginning  and  end  of  each  transect.  A  stake  with 
the  distance  from  the  beginning  of  the  transect  in- 
scribed on  it  can  be  driven  into  the  ground  at  inter- 
vals every  492  ft  ( 1 50  m ),  which  can  represent  a 
patch  size.  The  size  of  a  patch  will  depend  on  the 
species  being  studied.  Patches  492  ft  ( 150  m)  long 
x  656  ft  ( 200  m )  wide  represent  average  territory 
size  for  bird  species  in  southwestern  riparian  vegeta- 
tion. A  transect  2,461  ft  (750  m )  long  has  five  sub- 
plots or  patches,  each  492  ft  ( 1 50  m )  long  on  each 
side,  for  a  total  of  10  subplots.  Vegetation  data 
should  be  collected  within  each  subplot.  Each  tran- 
sect can  be  numbered,  and  the  number  and  direc- 
tional orientation  recorded  on  a  map.  A  typical 
transect  is  illustrated  in  Figure  3- 

Vegetation  Measurements 

In  each  subplot  ( patch ),  tree  counts  and  foliage 
density  measurements  are  made.  Tree  counts  are 
made  only  once  on  each  transect  unless  the  area  is 
later  affected  by  some  major  disturbance.  Counting 
is  unaffected  by  amount  of  foliage  present  and  can 
be  done  any  time.  Individuals  of  each  species  of  tree 
or  shrub  within  49  ft  (  1 5  m )  of  the  transect  are 


642 


Vegetation 


/ 


Edge  ot  Community  -  Structural  Type 


Outer  Boundary  ofx Sample  Area  (130  m) 


Subplot   1 


Subplot   2 


W  Subplot   3 


Subplot   4 


Subplot   5 


*      750 

m     > 

600 

m 

450 

m 

300 

m 

150 

m 

* 

s 

0 

m  \ 

Subplot   6 


Subplot   7 


Subplot   8 


Subplot   9 


Subplot    10 


Transect   Path 


Figure  3-     Typical  transect  through  a  relatively 
homogeneous  stand  of  vegetation,  showing 
individual  subplots  (patches)  and  outer 
boundaries. 


counted  in  each  492  ft  ( 1 50  m )  subplot.  Each  indi- 
vidual tree  is  categorized  by  height  (e.g.,  >  or  <  10 
ft  [>  or  <  3  m]),  by  presence  or  absence  of  mistle- 
toe, and  by  its  condition  (alive  or  dead). 

Sometimes  shrubs  or  trees  grow  in  densities  so 
great  that  it  is  not  feasible  or  possible  to  count  indi- 
viduals. Often  densely  packed  individuals  provide 
no  more  ground  cover  than  trees  in  less  dense  areas. 
For  example,  20  trees  in  one  area  could  equate,  in 
terms  of  ground  cover,  to  200  trees  in  another  area. 
To  circumvent  this  problem,  measure  the  height 
and  north-south  crown  diameter  of  30  or  more  indi- 
viduals of  each  tree  species  growing  at  various 
heights  in  uncrowded  conditions.  From  these  meas- 
urements, develop  regression  equations  to  determine 
the  ground  cover  by  an  individual  tree  ( shrub )  of  a 
given  height  ( Figure  4 ).  Thus,  when  a  dense  patch  of 
trees  or  shrubs  is  encountered,  measure  the  area  of 
the  patch  and  obtain  the  average  height  of  the  trees 
in  it.  Then,  divide  the  area  of  the  patch  by  the  area 
occupied  by  the  average  single  tree  of  the  same 
height  growing  in  uncrowded  conditions  to  obtain 
the  equivalent  number  of  trees  or  shrubs  growing 
under  uncrowded  conditions.  This  method  may  be 
applied  to  all  trees  and  shrubs  to  obtain  a  rough 
estimate  of  the  number  of  full-sized  equivalents  of  a 
given  plant  species  in  an  area. 


Figure  4.     (A)  Reference  sample  of  quail  bush.  (B)  Sample  area  with  approximately  30%  cover  by  quail  bush 
or  1,435  ft    (135  m~).  Since  one  shrub  occupies  9  ft"  (0.83  m  ),  the  area  has  approximately  163  individ- 
ual shrubs.  The  area  occupied  by  each  patch  of  quail  bush  was  measured  to  determine  the  ground  cover 
by  quail  bush. 


Vegetation 


643 


Foliage  density  estimates  are  made  in  all  stands 
annually  between  May  and  July;  stands  undergoing 
succession  (i.e.,  burned  and  regenerated  areas)  can 
be  measured  again  in  September  or  October  or  even 
more  frequently  if  desired.  Relative  foliage  density 
estimates  are  made  using  a  cover  board  technique 
(MacArthur  and  MacArthur  1961 ).  Sampling  is  done 
on  each  side  of  a  transect  at  three  points  ( 49  ft 
[15  m],  246  ft  [75  m],  and  432  ft  [135  m])  from  the 
beginning  of  each  subplot  (patch).  Thus  on  a  2,461 
ft  (750  m)  transect,  there  are  15  points  per  side 
for  a  total  of  30  points  (Figure  5). 

At  each  sample  point,  one  observer  paces  one 
step  perpendicular  to  the  transect.  A  second  ob- 
server holds  a  board  (approximately  8  x    16  in.  [20 
X  40  cm])  at  a  given  height  behind  the  nearest 
green,  leafy  vegetation  on  the  appropriate  side  of  the 
transect.  The  first  observer  directs  the  second  ob- 
server to  stop  when  green  foliage  covers  half  the 
board.  Distance  from  observer  to  board  is  measured 
with  a  tape  measure  or  range  finder  ( Figure  6 ).  Foli- 
age density  measurements  are  recorded  in  feet  be- 
cause of  the  scaling  of  the  equipment  we  used;  also, 
all  vegetation  calculations  were  based  on  the  English 
system  of  measures.  All  of  this  could  just  as  easily 
be  done  in  metric  units. 

Round  all  measurements  to  the  nearest  foot 
(0.3  m),  except  the  first.  In  the  first  foot  (0.3  m),  a 
distance  of  2  in.  (  5  cm )  represents  very  dense  vege- 
tation, but  zero  means  that  foliage  is  absent.  Thus, 
any  distance  >  0  but  <  1  ft  (<  0.3  m)  should  be 
regarded  as  1  ft  (0.3  m).  Often  it  is  difficult  to  obtain 
agreement  between  two  observers  for  distances 
<  1  ft  (<  0.3  m),  yet  the  difference  in,  for  example, 


the  foliage  density  estimate  between  1  in.  (2.5  cm) 
and  2  in.  (5  cm)  is  large.  Rounding  to  1  ft  (0.3  m) 
results  in  a  conservative  estimate  of  foliage  density 
in  very  dense  places,  but  yields  reproducible  results. 
Vertical  foliage  density  determinations  can  be  made 
at,  for  example,  0.5  ft  (0.15  m),  2  ft  (0.6  m),  5  ft 
(1.5  m),  10  ft  (3  m),  15  ft  (4.5  m),  20  ft  (6  m), 
25  ft  (7.5  m),  30  ft  (9  m),  and  every  10  ft  (3  m) 
thereafter,  until  no  vegetation  is  present.  Where 
measurements  are  taken  is  up  to  the  judgment  of 
those  in  charge  of  the  project.  Theoretically,  one 
should  use  a  ladder  to  make  measurements  at  higher 
levels.  This  is  impractical.  We  use  a  range  finder  to 
locate  a  point,  for  example,  at  26  ft  (8  m),  then  esti- 
mate as  carefully  as  possible  the  distance  to  a  second 
point  where  leaves  would  cover  half  the  board. 

Distances  are  measured  with  a  tape  measure  to 
the  nearest  foot  within  the  first  10  feet  (3  meters), 
because  the  foliage  density  index  is  more  sensitive 
to  vegetation  located  nearby.  Measurements  beyond 
this  are  estimated  with  the  aid  of  a  range  finder.  In 
sparse  areas,  the  distance  to  the  few  tall  trees  and 
prominent  shrubs  with  foliage  should  be  measured 
from  only  one  sample  point  within  a  subplot — the 
point  to  which  they  are  closest.  These  data  should 
be  recorded;  many  suitable  ways  are  imaginable. 

Frequency  of  Measurements.  It  is  important  to 
take  as  many  measurements  as  necessary  to  obtain  a 
reasonably  accurate  reflection  of  the  true  foliage 
density  and  diversity  of  a  stand.  In  general,  the  more 
measurements  that  are  taken,  the  greater  the 
precision;  however,  for  most  field  workers,  time  and 
personnel  are  restraining  factors.  Beyond  some 
number,  additional  measurements  increase  labor 


75m  ^ 


Figure  5.     Sampling  points  for  foliage  density  measurements  within  each  subplot  along  the  length  of  a 
transect. 


644 


Vegetation 


11  ft 


M^U  K^^ 


■H'V"  mp&m2 


5  ft 


Mir^/ 


Annual 


Atriplex 


Willow 


f  Transect 


Figure  6.     Selection  of  vegetation  for  foliage  density  measurements. 


requirements  more  than  can  be  justified  when  the 
increase  in  precision  is  very  small. 

In  identifying  the  effort  required  to  obtain  rea- 
sonably precise  data,  we  made  foliage  density  meas- 
urements at  1,2,  3,  4,  5,  9,  and  12  points  within 
each  subplot  on  each  side  of  7  transects,  which  in- 
volved 104  subplots.  Foliage  density  calculations 
were  made  using  one  measurement  from  each  sub- 
plot, then  two  measurements,  and  so  on,  until  five 
separate  sets  of  calculations  were  made.  This  was 
repeated  using  9  and  1 2  measurements  per  subplot. 
We  assumed  that  1 2  measurements  per  subplot 
yielded  results  as  close  to  reality  as  could  possibly 
be  obtained  with  our  methods.  The  foliage  density 
and  diversity  results  obtained  for  each  set  of  meas- 
urements were  expressed  as  percentage  deviation 
from  the  results  obtained  with  1 2  measurements  per 
subplot.  The  mean  percentage  difference  and  stan- 
dard deviation  decreases  as  the  number  of  measure- 
ment points  increases  (Figure  7),  until  five 
measurements  were  included  per  subplot.  Precision 
did  not  increase  with  nine  measurements  per  sub- 
plot. We  ultimately  chose  to  make  three  measure- 
ments per  subplot.  Making  five  measurements 
decreases  the  error  rate  by  only  1  % ,  but  increases 
the  work  by  67% .  In  our  judgment,  the  additional 
precision  is  not  warranted  by  the  effort  to  obtain  it. 

Repeatability.  Work  encompassing  a  large  area 
(100,000  a.  [40,000  ha])  will  probably  require 
several  individuals  to  collect  data.  Thus,  all 
personnel  must  be  carefully  trained  and  the 
similarity  of  data  collected  from  the  same  stand  by 


25- 

20- 

cc 

O        15- 

cc 
cc 

LLI 

1— 

z 

LU 

o 

CC        '0- 

LU 

Q. 

z 
< 

LU 

5- 

\ 

V 

1 

1           2           3          4           5          6           7          8           9          10         11         12 

NUMBER  OF  POINTS  PER  STRIP 

Figure  7.     Mean  percentage  error  associated  with 
foliage  density  and  diversity  measurements  when 
measurements  were  made  at  various  numbers  of 
points  in  each  subplot.  The  values  obtained  from 
1 2  measurements  per  subplot  were  assumed  to 
accurately  represent  the  true  values.  This  standard 
was  used  for  comparing  values  obtained  with 
smaller  numbers  of  measurements  per  subplot. 


Vegetation 


645 


different  observers,  determined.  In  general,  we  found 
least  agreement  in  data  from  stands  with  tall 
vegetation  (>  33  ft  [>  10  m]).  Space  precludes  an 
exhaustive  presentation  of  our  data;  however,  we 
describe  where  discrepancies  among  observer  teams 
were  greatest.  For  a  more  thorough  presentation, 
see  Anderson  and  Hunter  (  1983) 

Foliage  measurements  were  taken  by  four  teams 
in  a  relatively  dense  stand  encompassing  about  25 
a.  (10  ha)  and  containing  trees  to  64  ft  (20  m)  tall. 
It  was  assumed  that  some  observer  estimates  would 
be  too  high  and  some  would  be  too  low;  thus  the 
average  for  the  teams  would  approximate  the  "cor- 
rect" estimate  for  each  foliage  variable.  The  devia- 
tion from  the  average  was  then  considered  to  be  the 
error  rate  for  a  team.  The  error  divided  by  the  "cor- 
rect" value  yielded  the  percentage  of  error.  This 
procedure  indicated,  for  all  variables,  a  mean  error 
of  8% .  The  greatest  difficulty  was  in  assessing  vegeta- 
tion density  at  the  highest  levels;  the  average  error 
at  26  ft  (  8  m)  was  33%  (Table  1 ).  The  mean  error 
was  lowest  for  FHD  (2%  )  and  total  patchiness  (4%  ). 
Since  the  error  rates  for  foliage  density  estimates 
were  higher  than  hoped,  we  examined  the  results  in 
greater  detail. 

Although  foliage  density  estimates  had  a  mean 
error  rate  of  15%,  the  profiles  of  vertical  foliage 
distribution  derived  from  the  estimates  were  nearly 
identical  in  three  of  the  four  cases  in  the  first  test 
and,  in  all  four  cases,  the  vegetation  would  be  classi- 
fied the  same.  In  the  second  test,  two  of  the  three 
profiles  would  have  led  to  the  same  classification. 
This  is  of  major  importance  because  the  fact  that  all 
four  teams  in  the  first  set  obtained  results  that  led  to 


classifying  the  stand  in  the  same  way,  should  mini- 
mize the  importance  of  the  observed  variation  in 
data  obtained  by  different  teams.  FHD  and  patchiness 
estimates  obtained  by  different  teams  for  a  given 
stand  were  always  similar. 

Overall,  we  are  satisfied  that  these  methods 
accurately  and  consistently  differentiated  areas  of 
high,  intermediate,  and  low  foliage  density  and  diver- 
sity in  the  vertical  and  horizontal  planes.  However, 
stands  differing  slightly  in  foliage  density  probably 
cannot  be  reliably  separated. 

Training  Personnel.  All  personnel  should  be 
thoroughly  trained  in  field  techniques  before 
collecting  data  on  their  own.  More  experienced 
personnel  should  accompany  less  experienced 
personnel  in  the  field.  Having  people  with 
experience  with  the  technique  is  not  essential,  but 
having  dedicated,  conscientious  personnel  is. 

Need  for  Preliminary  Studies 

In  any  study  where  discovery  of  relationships 
between  animals  and  environmental  variables  is  a 
goal,  there  is  a  need  for  preliminary  studies  (Green 
1979;  Platts  1981;  Platts  et  al.  1983).  Such  studies 
are  essential  because  they  allow  investigators  to  test 
to  see  if  their  methods  satisfactorily  produce  the 
desired  kinds  of  information.  Preliminary  studies  also 
offer  an  opportunity  to  become  thoroughly  familiar 
with  the  study  area  before  beginning  the  study  and 
provide  a  training  period  for  field  personnel.  In  addi- 
tion, preliminary  studies  allow  the  investigators  to 
determine  differences  between  experimental  and 
control  areas  prior  to  treating  the  experimental  area. 


Table  1.     Average  error  associated  with  various  vegetation  variables  collected  by  four  different  teams  from 
the  same  stand.  (The  mean  measurements  by  all  four  teams  were  the  standards  with  which  comparisons 
were  made.) 


Foliage  Density  at  Various  Heights 

Patchiness 

0.0-0.6  m  0.6-4.5  m  4.5-7.5  m  &7.5  m 

0.0-0.6  m 

0.6-4.5  m 

4.5-7.5  m  s=7.5  m 

Team 

(0-2  ft)       (2-15  ft)    (15-25  ft)  (^-25  ft) 

Total 

FHD 

(0-2  ft) 

(2-15  ft) 

(15-25  ft)  ( 

^25  ft) 

Total 

TESTI 

1 

20.5               3.4            20.8          74.2 

31.6 

1.1 

0.9 

0.8 

0.9 

11.6 

1.8 

2 

12.6             13.9               5.9          21.0 

1.7 

4.1 

5.2 

3.7 

8.7 

6.1 

6.7 

3 

16.8               6.1             14.8          27.0 

17.3 

0.9 

2.0 

1.4 

1.5 

6.9 

3.7 

4 

16.2               4.3             11.7          26.3 

15.9 

2.0 

2.3 

1.3 

1.5 

10.7 

5.0 

TEST  II 

1 

8.5               3.2               1.0          13.5 

2.4 

2.8 

1.5 

0.2 

2.4 

0.0 

2.8 

2 

4.6               7.3             10.2          42.0 

1.3 

2.8 

1.3 

0.9 

0.3 

2.0 

2.8 

3 

12.7             10.5             11.0          28.0 

1.6 

0.1 

1.6 

2.1 

1.2 

0.6 

0.1 

Mean 

13.3               7.0             10.8          33.1 

10.3 

2.0 

2.1 

1.5 

2.4 

5.4 

3.3 

SD 

4.9               4.0               6.3          20.0 

11.7 

1.4 

1.4 

1.1 

2.9 

4.7 

2.1 

646 


Vegetation 


For  example,  if  the  objective  is  to  determine  the 
impact  of  light  grazing  on  an  area  in  terms  of 
changes  in  wildlife  densities  and  diversities,  it  is 
critically  important  to  establish  plots  where  all  varia- 
bles, except  the  one  being  tested,  are  controlled. 
This  means  that  the  vegetation  and  wildlife  use  must 
be  analyzed  in  the  control  and  experimental  areas 
before  grazing.  Impacts  can  only  be  evaluated  after 
internal  variation,  before  the  impact,  is  known. 

The  argument  is  often  heard  that  such  prelimi- 
nary studies  are  too  time-consuming  and  expensive 
to  carry  out.  This  is  fallacy;  preliminary  studies  can 
save  time  and  money  and  are  the  only  way  of  obtain- 
ing scientifically  valid  results  (Green  1979).  Without 
them,  results  obtained  after  an  impact  cannot  be 
determined  because  the  lack  of  knowledge  of  the 
variation  before  the  impact  was  unknown.  In  other 
words,  as  Platts  ( 1981 )  correctly  pointed  out,  the 
notion  that  collection  of  inventory  "garbage"  (i.e., 
data  without  preliminary  studies )  leads  to  reliable 
analyses  is  purely  mythical.  Unfortunately,  it  is  a 
pervasive  myth.  If  preliminary  studies  cannot  be 
done,  the  study  should  not  be  undertaken  or  under- 
taken with  the  realization  that  conclusions  based 
on  the  findings  are  not  well-founded  scientifically. 


ANALYTICAL  METHODS 

Foliage  Density 

Each  plant  distance  measurement  is  converted 
to  surface  area  per  cubic  unit  of  space  (i.e.,  foliage 
density )  according  to  the  following  formula: 


K 


log,2        0.693 


These  measurements  will  produce  a  table  of 
density  measurements  such  as  in  Table  2,  which  can 
be  used  for  subsequent  calculations  of  derived  varia- 
bles such  as  vertical  and  horizontal  diversity. 

Table  2.     Sample  foliage  density  estimates  (ft "/ft  ) 
used  for  calculating  patchincss  and  foliage  height 
diversity. 


0.15  m 

0.6  m 

1.5  m 

3  m 

4.6  m 

Plot 

(0.5  ft) 

(2  ft) 

(5  ft) 

(10  ft) 

(15  ft) 

1 

0.16 

0.20 

0.29 

0.10 

0.01 

2 

0.12 

0.15 

0.23 

0.06 

— 

3 

0.08 

0.09 

0.27 

0.09 

0.01 

4 

0.28 

0.15 

0.09 

0.01 

0.00 

5 

0.19 

0.22 

0.09 

0.02 

0.00 

6 

0.18 

0.34 

0.29 

0.10 

0.02 

7 

0.07 

0.31 

0.31 

0.03 

0.01 

8 

0.08 

0.18 

0.31 

0.02 

— 

9 

0.15 

0.16 

0.32 

0.03 

— 

10 

0.23 

0.15 

0.13 

0.01 

— 

Vertical  Diversity  (Foliage  Height  Diversity) 

Vertical  or  foliage  height  diversity  (FHD)  for 
each  transect  is  calculated  according  to  the  informa- 
tion theory  (Shannon  and  Weaver  1949): 


D 


D 


FHD 


-2 


(p,)(l0gnPi) 


Where:  K=  the  foliage  density  and  D 
ured  distance. 


the  meas- 


Foliage  density  per  subplot  is  the  sum  of  the 
average  of  the  three  measurements  taken  at  each 
vertical  plane.  For  example,  foliage  density  at  5  ft 
(1.5  m)  in  one  subplot,  for  which  the  distances  were 
9,  15,  and  2  ft  (2.7,  4.5,  and  0.6  m),  respectively, 
would  be  calculated  as  follows: 


0.693        0.693    ,    0-693 
9  15  2 


■*■  3  =  0.1556 


Foliage  density  at  10  ft  (3  m)  for  distances  of  1,  2, 
and  3  ft  (0.3,  0.6,  0.9  m),  respectively,  would  be — 


0.693        0.693        0.693 
1  2  3 


3  =  0.4217 


The  density  for  the  two  vertical  planes  is  0.1556 
+  0.4217  =  0.5773.  If  no  green  foliage  occurred  at 
a  particular  point,  a  zero  is  used  in  the  calculations. 


Where:  p,  is  the  proportion  of  total  foliage  density 

contributed  by  the  density  at  level  i.  (Sample 
calculations  are  shown  in  Table  3- ) 

Many  stands  of  vegetation  include  a  shrubby 
layer  up  to  about  3  ft  (  1  m).  We  chose  foliage  den- 
sity estimates  from  0.5  ft  (0.15  m)  and  2  ft  (0.6  m) 
to  represent  this  vertical  layer.  Foliage  density  in  this 
layer  is  the  sum  of  the  density  at  0.5  ft  (0.15  m) 
and  2  ft  ( 0.6  m ).  Most  stands  will  probably  have  an- 
other layer  extending,  for  example,  from  2  to  15ft 
(0.6  to  4.5  m).  Foliage  density  estimates  at  5  ft  ( 1.5 
m)  and  10  ft  (3  m)  could  be  used  to  represent  this 
layer.  Many  stands  have  a  third  layer,  usually  poorly 
developed,  extending  above  the  second  layer  for 
an  additional  7  to  10  ft  (2  to  3  m)  or  more,  depend- 
ing on  locality.  We  chose  foliage  density  estimates 
at  15  ft  (4.5  m)  and  20  ft  (6  m)  to  represent  this 
layer.  Foliage  <  7.5  m  (<  25  ft)  could  represent  a 
fourth  layer.  The  concept  of  FHD  with  four  vertical 
layers  is  depicted  in  Figure  1. 


Vegetation 


647 


Table  3.     Example  of  calculation  of  foliage  height  diversity  from  data  in  Table  2. 


Item 

Height  Class 

Mean  total  density 
Proportion  (p,) 
log-toPi 
p,log10p. 

0.15-0.6  m 
(0.5-2  ft) 

1 .5-3.0  m 
(5-10  ft) 

4.6-6.0  m 
(15-20  ft) 

>  7.5  m 
(>  25  ft) 

Total 

0.35 

0.55 

-0.26 

-0.14 

0.28 

0.44 

-0.36 

-0.16 

0.00 

0.01 

-2.20 

-0.01 

0 
0 
0 
0 

0.63 

FHD  =      -  2(p,)(log10p,)  =    -[(-0.14)  +  (0.16)  +  (0.01)  +  (0)] 
=  0.31 

Horizontal  Diversity  (Patchiness) 

Horizontal  diversity  (or  patchiness)  is  a  struc- 
tural feature  of  a  habitat  describing  the  regularity  of 
vegetation  as  it  is  distrihuted  in  the  horizontal  plane 
(Figure  2).  A  citrus  orchard  with  roughly  equally 
sized  and  evenly  spaced  trees  has  little  horizontal  di- 
versity. Patchiness  or  diversity  is  greater  in  a  honey 
mesquite-quail  bush  (Atriplex  lentiformis )  habitat 
with  irregularly  spaced  trees  and  shrubs  of  different 
heights.  Diversity  in  the  horizontal  plane  can  be 
calculated  for  any  vertical  layer  from  which  foliage 
density  estimates  are  made. 

The  variance  associated  with  the  mean  total 
foliage  density  for  each  vertical  plane  across  all  sub- 
plots can  be  used  as  a  measure  of  horizontal  diver- 
sity. For  example,  the  summed  and  averaged  foliage 
densities  for  0.5  ft  (0.15  m)  and  2  ft  (0.6  m)  in  each 
plot  (Table  4)  yields  the  mean  foliage  density  for 
the  layer  0  to  2  ft  (0.0  to  0.6  m),  in  this  case,  0.35. 
Horizontal  diversity  is  the  variance  associated  with 
mean  total  foliage  density,  in  this  case,  0.01.  Variance 
or  standard  deviation  squared  (s")  is  defined  as — 


This  variance  is  calculated  for  each  vertical 
layer.  Total  horizontal  diversity  is  the  sum  of  the 
variances  for  all  layers. 

In  calculating  horizontal  diversity,  we  are  assess- 
ing the  variance  between  subplots.  Therefore,  if  hori- 
zontal diversity  is  thought  of  as  patchiness,  we  are 
defining  a  patch  as  a  unit  492  ft  ( 150  m)  long  and  as 
wide  as  the  distance  from  the  transect  to  the  edge 
of  the  stand  under  study,  usually  420  ft  (128  m). 
Choice  of  this  patch  size  was  based  on  evidence  that 
many  common  birds  in  the  area  use  patches  of  about 
this  size  (Conine  1982).  It  is  possible  that  an  area 
that  rated  very  patchy  on  a  smaller  scale  could  be 
rated  homogeneous  using  this  particular  scale. 

Since  0.00  and  0.69  represent  minimum  and 
maximum  foliage  density  values,  respectively,  maxi- 
mum horizontal  diversity  or  patchiness  for  a  given 
layer  is  0.238.  Since  there  are  four  layers,  maximum 
horizontal  plane  diversity  is  4  x  0.238  or  0.952. 
Since  this  is  close  to  1,  the  sum  of  the  diversity  for 
the  four  layers  closely  represents  the  percentage 
of  maximum  diversity  possible  for  an  area. 


HDI 


2 

s     = 


2    (k,  -  K)2 


n-1 


It  can  be  calculated  more  easily  using  the 
formula — 


HDI 


2 

s     = 


2      k,2   -  (Sk,)2 

i=  1 n 

n-1 


Where:  kj  =  the  foliage  density  of  the  ith  sam- 
pie 
K  =  the  mean  foliage  density  for  the 

sample 
n  =  sample  size 
HDI  =  Horizontal  Diversity  Index 


Another  method  for  calculating  FHD  and  patchi- 
ness is  to  simply  record  the  presence  or  absence  of 
vegetation  at  various  vertical  positions.  This  could  be 
done  with  a  long  pole  and/or  a  range  finder.  More 
stops  would  have  to  be  made,  but  FHD,  relative  den- 
sity values,  and  patchiness  estimates  could  be  made 
on  the  basis  of  the  proportion  of  total  points  at 
which  foliage  occurred.  This  method  may  be 
quicker,  could  reduce  the  amount  of  calculation,  and 
might  be  equally  accurate. 


CLASSIFYING  VEGETATION 

If  the  goal  of  a  field  project  is  to  determine  hab- 
itat associations  for  a  wildlife  group,  such  as  birds, 
over  a  relatively  large  area  (e.g.,  100,000  a.  [40,000 
ha]),  the  area  must  be  sampled  with  sufficient  inten- 
sity so  all  habitats  are  represented  by  at  least  one 


648 


Vegetation 


Table  4.     Example  of  calculation  of  horizontal  diversity  index  ( HDI )  from  data  in  Table  2. 


Item 

Height  Class 

Sum  of  k. 
Sum  of  squares 

kj 
Sum  squared 

(2  k? 

n 

0.15-0.6  m 
(0.5-2  ft) 

1 .5-3.0  m 
(5-10  ft) 

4.6-6.0  m 
(15-20  ft) 

>  7.5  m 
(>  25  ft) 

Total 

3.49 

1 .3073 

12.1801 
10 

2.80 

0.8967 

7.8400 
10 

0.05 

0.0007 

0.0025 
10 

— 

Ik,2  -  (1 
Hni       ~2 

k,)2 
^           o  m 

+      0.01 

+      0.00 

+     — 

=  0.02 

HUI  -  o     -               n_1 

—    U.U  I 

sample  plot  ( transected  area ).  Of  course,  replication 
is  desirable.  If  this  sampling  is  random,  then  the 
number  of  transects  per  habitat  will  be  proportional 
to  the  abundance  of  that  hahitat  in  the  study  area. 
All  transects  should  be  about  the  same  length  and 
should  be  within  a  relatively  homogeneous  stand.  At 
this  point,  some  arbitrary  decisions  may  have  to  be 
made  because  of  the  ambiguity  associated  with  the 
term  "relatively  homogeneous." 

In  general,  experience  and  common  sense  will 
be  immeasurably  helpful.  A  field  biologist  relatively 
familiar  with  an  area  will  generally  know  when  ( s  )he 
leaves  one  habitat  type  and  enters  another.  It  would 
be  poor  judgment  to  establish  a  transect  in  a  mixed 
cottonwood  (Populus  fremontii)  and  willow  (Salix 
sp. )  habitat  and  have  the  transect  cross  990  ft  (  300 
m)  of  honey  mesquite  habitat.  Having  such  a  tran- 
sect cross  a  small  group  of  10  or  15  honey  mesquite 
trees  might  not  be  inappropriate.  A  general  familiar- 
ity with  what  is  available  will  usefully  serve  as  a 
guide.  If  the  decision  is  made  to  include  one  or 
more  small  patches  of  honey  mesquite  in  the  cotton- 
wood-willow  habitat,  this  will  add  heterogeneity  to 
the  data.  Vegetation  often  varies  considerably  over 
small  areas;  thus,  some  heterogeneity  is  not  necessar- 
ily bad  and  is  probably  inevitable. 

The  next  step  is  to  cluster  transects  that  are 
similar  in,  for  example,  vertical  diversity.  This  can  be 
done  by  calculating  the  overlap  in  vertical  foliage 
diversity.  One  way  to  do  this  is  to  use  Horn's  ( 1966 ) 
equation  for  overlap: 


R« 


2(x,  +  y,)log(x,  +  y~j)  -  SxjlogXj  -  Sy.logy, 
(X  +  Y)  log(X  +  Y)  -  X  logY  -  Y  logY 

Where:   x(  and  yt  =  proportion  of  total  foliage 
density  occurring  at  vertical  band  i 
for  stand  X  and  stand  Y. 


X,Y  =  the  total  foliage  density  for  stand  X  and 
stand  Y.  X  and  Y  represent  total  foliage 
density. 

Although  this  equation  appears  formidable,  it  is  ac- 
tually very  simple.  The  calculations  are  laborious  but 
in  this  age  of  computers,  large  numbers  of  overlaps 
can  be  calculated  in  minutes.  The  biologist  must, 
of  course,  grasp  the  fundamental  arithmetic  process. 
Biologically,  high  overlaps  simply  represent  a  mathe- 
matical expression  of  a  high  degree  of  similarity  in 
vertical  configuration.  Once  this  is  understood,  the 
computer  can  take  care  of  the  tedium. 

From  a  matrix  of  overlap  values,  including  all 
possible  two-way  comparisons  between  stands  (tran- 
sects), a  dendrogram,  such  as  that  in  Figure  8,  can 
be  constructed  using  the  following  relationship: 


_   CA 


CB 


C,  AB 


Where:    C,  AB  =  the  overlap  of  stand  C  with 

stands  A  and  B 
CA       =  the  overlap  between  stand 

C  and  stand  A 
CB       =  the  overlap  between  stand 

C  and  stand  B 


Simply  stated,  the  overlap  in  the  vertical  distribution 
of  foliage  in  stand  C  with  stands  A  and  B  is  equal  to 
the  average  overlap  of  C  with  A  and  C  with  B  (Cody 
1974).  The  dendrogram  (Figure  8)  was  interpreted 
as  revealing  six  categories,  based  on  vertical  configu- 
ration without  consideration  of  dominant  vegetation. 
Each  transected  stand  within  a  category  (designated 
I-VI )  has  a  vertical  configuration  more  similar  to 
other  intracategory  stands  than  to  any  stand  within 


Vegetation 


649 


VEGETATION  TYPE 


I     I!     Ml 


V  VI 


Figure  8.     Dendrogram  showing  similarities  in  ver- 
tical profile  between  transects,  based  on  overlap 
in  vertical  distribution  of  foliage. 

any  other  category.  This  determination  is  based  on 
proportional  distribution  of  the  foliage  in  the  vertical 
dimension.  These  categories  can  now  be  subdivided 
according  to  the  dominant  vegetation  present 
(honey  mesquite,  salt  cedar  [Tamarix  chinensis], 
screwbean-mesquite  mix,  quaking  aspen  (Populus 
tremuloides),  sycamore,  etc.).  Criteria  used  to  arrive 
at  these  divisions  is  up  to  the  investigator.  As  a 


guideline,  consider  that  each  habitat  should  be  rep- 
resented by  at  least  two  transected  stands  and  that 
the  number  of  recognized  habitats  should  not  be 
unwieldy.  Stress  similarities  rather  than  differences  in 
attempting  to  reduce  the  overall  number. 

The  categories  could  be  subdivided  again  on  the 
basis  of  foliage  density.  For  example,  if  we  found  a 
stand  dominated  by  willow  to  20  ft  ( 6  m )  containing 
1,250  trees/a.  (500/ha)  and  another  stand  with  the 
same  vertical  configuration  but  only  half  as  many 
trees,  we  might  want  to  recognize  these  stands  as 
distinct  habitats.  Ordinarily,  this  situation  does  not 
arise  because  the  sparser  ( more  open )  area  would 
have  a  more  highly  developed  layer  of  vegetation 
(shrubs,  annuals)  and  would,  therefore,  have  a  differ- 
ent vertical  configuration. 

The  vertical  classification  we  arrived  at  for  ripar- 
ian vegetation  is  depicted  in  Figure  9.  From  this 
information,  a  quick  and  easy-to-use  key  to  the  var- 
ious habitats  can  be  developed  (Table  5). 

Habitat  Heterogeneity 

Including  several  transects  in  one  habitat  can 
increase  within-habitat  variation  considerably.  The 
advantage  of  using  the  habitat  concept  is  that  habi- 
tats can  be  mapped,  data  are  less  cumbersome  to 


Table  5.     User's  guide  to  classifying  vegetation  by  dominant  tree  or  shrub  species  present. 


(1)  A.  Stand  in  which  virtually  100%  of  the  trees  present  are  of  one  species  or  virtually  100% 

arrowweed  Go  to  2 

B.  Trees  within  stand,  clearly  of  mixed  species.  The  different  species  may  occur  as  mixed  individuals  or  as 
small  clumps  Go  to  3 

(2)  A.  Stand  in  which  trees  are  composed  of  nearly  100%  of  some  species  (may  be  occasional,  widely  scat- 

tered individuals  of  one  or  more  species).  Many  large  stands  have  arrowweed  in  patches  encompassing 
5  a.  (2  ha)  or  more.  Honey  mesquite  stands  in  addition  to  or  instead  of  arrowweed  may  have  quail  bush, 
four-winged  salt  bush,  wolfberry,  or  mkweed  Salt  Cedar  I-IV  or  Honey  Mesquite  III-IV 

B.  Stand  composed  of  nearly  100%  arrowweed;  may  be  an  occasional  tree  or  widely  scattered  clump  of 
some  other  shrub  Arrowweed 


(3)  A.  Stand  of  vegetation  is  structural  type  I  and  trees  are  primarily  salt  cedar,  cottonwood  and/or  willow  with 

occasional  widely  scattered  screwbean  or  honey  mesquite  tree  or  clumps  of  trees.  Arrowweed  or  some 
other  shrub  may  occur  in  relatively  widely  scattered  clumps  Salt  Cedar-Cottonwood  Willow  mix 

B.  Vegetation  not  structural  type  I Go  to  4 

(4)  A.  Stand  of  vegetation  is  structural  type  II  or  III  Go  to  5 

B    Stand  not  structural  type  II  or  III  Go  to  6 


650 


Vegetation 


Table  5.     User's  guide  to  classifying  vegetation  by  dominant  tree  or  shrub  species  present  (concluded). 


(5)  A.  Stand  in  which  trees  are  salt  cedar  with  large  numbers  of  Cottonwood  and/or  willow  present;  may  be 

widely  scattered  individuals  or  clumps  of  screwbean  or  honey  mesquite Salt  Cedar-Cottonwood/ 

Willow  mix 

B.  Stand  in  which  trees  are  mainly  salt  cedar  and  screwbean  mesquite;  may  be  an  occasional, 
widely  scattered  clump  or  individual  Cottonwood  and/or  willow  or  honey 
mesquite Salt  Cedar-Screwbean  Mesquite  mix 

(6)  A.  Stand  of  vegetation  in  structural  type  IV  Go  to  7 

B.  Stand  not  structural  type  IV  Go  to  8 

(7)  A.  Stand  composed  mainly  of  salt  cedar  but  with  significant  numbers  of  cottonwood  and/or  willow  present; 

may  be  widely  scattered  individuals  or  clumps  of  screwbean  or  honey  mesquite.  Shrubs,  mainly  ar- 
rowweed,  abundant  and  occurring  in  moderate  to  relatively  large  patches,  sometimes  encompassing  5  a. 
(2  ha)  or  more  Salt  Cedar-Cottonwood/Willow  mix 

B.  Stand  much  as  above  but  with  screwbean  mesquite  or  honey  mesquite  instead  of  cottonwood  and/or 
willow  Salt  Cedar-Screwbean  Mesquite  mix  or  Salt  Cedar-Honey  Mesquite  mix 

(8)  A.  Stand  of  vegetation  is  structural  types  V  or  VI  Go  to  9 

B.   Stand  not  structural  type  V  or  VI  Go  to  3 

(9)  A,   Stand  composed  mainly  of  salt  cedar,  but  with  significant  numbers  of  cottonwood  and/or  willow  occurring 

as  scattered  individuals  or  clumps.  Arrowweed  is  usually  abundant  (occasionally  some  other  shrub 

species  such  as  quail  bush  also  present)  and  occurring  in  patches  encompassing 

several  a.  (ha)  Salt  Cedar-Cottonwood/WUlow  mix 

B.   Stand  composed  primarily  of  salt  cedar  but  with  significant  numbers  of  individuals  or  clumps  of  screw- 
bean or  honey  mesquite.  May  be  widely  scattered  individuals  or  clumps  of  screwbean  or  honey  mes- 
quite. Arrowweed  present  as  in  (9)A Salt  Cedar-Screwbean  Mesquite  mix  or  Salt  Cedar-Honey 

Mesquite  mix 


NOTE:  This  key  can  be  used  to  classify  about  95%  of  the  riparian  vegetation  found  along  the  lower  Colorado  River  By  applying  the 
same  general  principles  used  to  construct  the  key  and  a  little  imagination,  rare  vegetation  types  can  also  be  classified 


work  with,  and  habitats  can  be  communicated  easier 
than  transects.  Furthermore,  habitat  is  usually  man- 
aged as  a  concept.  However,  if  microhabitat  variation 
is  marked,  it  will  be  a  matter  of  concern  to  many 
biologists.  Use  of  the  habitat  concept  should  occur 
after  one  is  thoroughly  familiar  with  the  variation 
that  will  be  concealed  and  its  meaning  when  inter- 
preting subsequent  data  analyses. 


Foliage  Density.   In  separating  transects  into 
structural  types,  having  various  structural  types 
different  from  one  another  at  a  statistically  significant 
(P  <  0.05)  level  for  at  least  one  of  the  recognized 
vertical  layers  could  be  desirable.  Adjacent  structural 
types  may  overlap,  but  as  long  as  some  statistical 


differences  exist,  one  can  legitimately  maintain  that 
they  represent  different  habitats.  Data  from  our 
studies  (Anderson  and  Ohmart  1984b;  Anderson  and 
Hunter  1983)  are  shown  for  foliage  density  in  Figure 
1 0  and  the  proportion  of  foliage  in  each  of  three 
vertical  layers  in  Figure  1 1 .  These  figures  illustrate 
the  range  of  variation  found  among  transects  falling 
into  each  category;  they  also  show  the  mean  and 
two  standard  errors  of  the  mean  for  each  type.  Note 
that  when  using  foliage  density  measurements,  types 
V  and  VI  differed  little  from  each  other,  but  when 
the  proportion  of  the  total  foliage  found  in  each  of 
three  layers  was  considered,  type  VI  had  a 
significantly  greater  proportion  of  its  total  foliage  in 
the  lower  layer  and  significantly  less  in  the  middle 
layer  (  Figure  1 1 ). 


Vegetation 


651 


1   J         v 

AT                     vJ       (r^X 

4                       I] 

I J 1  ^JL 

J*J<( 

^^V*^, 

IV 


/>w£^. 


Figure  9-     Examples  of  vertical  configurations  for  the  vegetation  structural  types  defined  (Figure  1 1 )  in  the 
lower  Colorado  River  Valley.  Type  I  is  open  gallery  forest;  type  II  is  closed  gallery  forest;  type  III  is  scrub 
thicket;  type  IV  is  scrub  thicket  with  patches  of  bushes;  type  V  is  mostly  shrubs  and  bushes  with  scat- 
tered scrubby  thickets;  type  VI  is  bushes,  shrubs,  and  grasses. 


Heterogeneity  in  Tree  Counts.  The  mean  number 
of  trees  of  a  particular  species  can  also  vary 
considerably  among  patches.  For  example,  the  mean 
number  of  salt  cedar  per  salt  cedar  thicket  with 
patches  of  shrubs  was  163  trees,  with  a  very  large 
standard  deviation  (105;  Table  6).  This  could  be 
expected  in  a  very  patchy  habitat.  Some  patches  will 
have  more  trees  than  others,  which  makes  them 
patchy.  However,  habitats  classified  as  honey 
mesquite-woodland  had  very  few  tree  species 
present  other  than  honey  mesquite;  salt  cedar 
thickets  had  only  one  species  present  other  than  salt 
cedar  (Table  6).  Thus,  while  salt  cedar  is  virtually 
the  only  tree  present  in  salt  cedar  habitats,  the  trees 
within  such  habitats  may  be  tall  and  relatively 
homogeneously  distributed  (type  I)  or  scrubby  and 
patchily  distributed  (type  IV),  with  patches  of  shrubs 
intermingled  among  the  salt  cedar. 

Within-habitat  counts  can  also  vary  through 
highly  localized  edaphic  features.  For  example,  the 
soil  moisture  level  in  an  old  oxbow  that  is  inter- 
sected by  the  transect  may  have  a  few  individual 
tree  species  not  found  elsewhere  on  the  transect. 

Local  heterogeneity  in  soil  layering  and  struc- 
ture can  cause  heterogeneity  in  plant  structure.  The 


distribution  of  soil  types  within  a  floodplain  is  typi- 
cally heterogeneous.  A  highly  localized,  dense  clay 
soil  type  could  cause  a  very  local  concentration  of 
soil  electrolytes.  Vegetation  growing  in  such  soil 
often  attains  less  stature  and  biomass  (Anderson  and 
Ohmart  1982)  and,  therefore,  vertical  differentiation 
is  simpler  than  for  adjacent  vegetation.  Such  varia- 
tion may  be  so  frequent  that  it  is  not  feasible  or 
desirable  to  delineate  it  within  otherwise  relatively 
homogeneous  stands. 

Another  source  of  variation  includes  scattered 
individual  trees  that  were  once  more  widely  distrib- 
uted. In  our  study  area,  Cottonwood  and  willow 
trees,  often  occurring  as  widely  scattered  individuals 
or  as  small  clumps  (66  x  66  ft  [20  x  20  m])  of 
trees,  are  relicts  of  gradually  disappearing  habitat 
(Ohmart  et  al.  1977). 

Fire,  another  cause  of  within-stand  heterogene- 
ity, in  varying  degrees  and  at  various  times,  has  af- 
fected nearly  every  stand  of  vegetation  along  the 
lower  Colorado  River.  When  a  stand  is  burned,  not 
all  parts  of  it  burn  with  equal  intensity;  some  corners 
or  clumps  remain  intact.  Parts  of  a  stand  may  burn 
more  than  once,  so  at  any  given  time  not  all  parts 


652 


Vegetation 


CO 

c> 


z 

LU 
Q 

LU 
O 
< 


2.200 
2.000  - 
1.800 
1.600  - 
1.400  • 

1.200  • 

1.000 

0.800 

0.600- 

0.400 

0.200 

0.000 


TYPE  TYPE 

IV 


TYPE  TYPE 

V  VI 


JUL 


a 'B'cl a'b'c Ia'b'c     a'b'c  Ia'b'cIa'b'c 

HEIGHT 


Figure  10.     Variation  in  foliage  density  between  subplots  within  all  structural  types  at  each  of  three  vertical 
levels.  The  proportional  distribution  clearly  differentiates  the  vegetation  types,  but  foliage  density  does 
not.  Horizontal  lines  represent  mean  values;  large  rectangles  represent  ±  1  standard  deviation;  small  rec- 
tangles represent  ±2  standard  errors.  A  =  0-2  ft  (0.0-0.6  m);  B  =  2-15  ft  (0.6-4.5  m);  C  =  >  15  ft 
(>  4.5  m). 


z 

LU 
Q 

LU 
O 
< 


LL 

o 


LU 

o 
<r 

LU 
0- 


Figure  11.     Proportional  distribution  of  vegetation  in  three  vertical  layers  among  subplots  within  various 
stands  of  vegetation,  generally  classified  as  one  vertical  structural  type  (I-VI).  Horizontal  lines  represent 
mean  values;  large  rectangles  represent  ±  1  standard  deviation;  small  rectangles  represent  ±  2  standard 
errors.  A  =  0-2  ft  (0.0-0.6  m);  B  =  2-15  ft  (0.6-4.5  m);  C  >  15  ft  (4.5  m). 


Vegetation 


653 


are  at  precisely  the  same  stage  of  post-fire  recovery. 
Even  when  burned  evenly,  not  all  parts  of  a  stand 
redevelop  at  precisely  the  same  rate.  Thus,  at  some 
level  of  analysis,  considerable  heterogeneity  could  be 
found  within  any  fundamentally  homogeneous  stand. 

Although  such  delineation  of  plant  species  het- 
erogeneity may  be  important  for  understanding  some 
aspects  of  the  distribution  of  vegetation  or  wildlife 


using  the  vegetation,  it  may  be  beyond  the  scope  of 
study.  Such  delineation  could  require  more  time  and 
money  than  is  available.  Such  small  parcels  typically 
range  from  2.5  a.  (1  ha)  to  about  12.5  a.  (5  ha). 
Areas  smaller  than  25  a.  (10  ha)  could  not  be  accu- 
rately plotted  on  map  scales  of  1:9,449  cm.  Availabil- 
ity of  funding  and  consideration  of  the  desired  scale 
are  factors  that  must  be  considered  when  deciding 
how  much  edaphic  variation  should  be  delimited. 


Table  6.     Average  number  of  trees  ( 
the  lower  Colorado  River. 


1  SD)  per  subplot  in  each  of  23  recognized  riparian  habitat  types  along 


Vegetation 
Type 

N 

Number  of  Trees  Per  Subplots 
150  x   15  m  (492  x  50  ft) 

Percent  of 
Subplots 
with  No 
Trees  of 
Dominant 
Species 

Salt 
Cedar 

Cottonwood 

Willow 

Screwbean 
Mesquite 

Honey 
Mesquite 

x          SD 

X 

SD 

X          SD 

x            SD 

x         SD 

Salt  cedar 

1 

II 
III 
IV 
V 
VI 

18 
8 
28 
32 
109 
20 

95         20 
47          19 
74         25 
163       105 
133       146 
31          50 

0 
0 
0 
0 
0 
0 

0 
0 
0 
0 
0 
0 

0           0 
0            0 
0           0 
0           0 
0            0 
0           0 

0              0 
0               0 
7            13 

0  0 

1  3 
0               0 

2         20 
0           0 
0           0 
0           0 
0           0 
0           0 

0 
0 
0 
0 
1 
0 

Salt  cedar- 
cotton  wood/ 
willow 

1 

II 
III 
IV 
V 
VI 

18 
10 
62 
52 
30 
22 

52          13 

129  46 

130  147 
38          53 
44          49 
19          32 

59 

38 

19 

0 

0 

1 

27 
22 

44 
0 
0 
1 

87          23 
49          34 
54          66 
29          17 
17          21 
1          26 

0               0 
0               0 
13            23 
7             15 
0               0 
0               0 

0           0 
0           0 
6            7 
0           0 
0           0 
0           0 

0-0 
0-0 
0-6 
3-8 
0-0 
0-50 

Salt  cedar- 
screwbean 
mesquite 

II 
III 
IV 

V 
VI 

10 
40 
78 
84 
18 

63          24 
49          43 
60          58 
45          39 
45          55 

2 
0 
0 
0 
0 

4 
0 
0 
0 
0 

1             1 
0            0 
4          25 
0            0 
0            0 

96             17 
18             15 
39            31 
44            62 
6               6 

0           0 
0           0 
0           0 
0           0 
0           0 

0-0 
0-8 
1-6 
0-8 
0-22 

Salt  cedar- 
honey  mes- 
quite 

IV 

38 

41          53 

0 

0 

0            0 

0               0 

35         68 

2-6 

Honey  mes- 
quite 

III 
IV 
V 
VI 

24 

122 

56 

52 

0            0 
0            0 
0            0 
0            0 

0 
0 
0 
0 

0 
0 
0 
0 

0            0 
0            0 
0            0 
0            0 

1* 

0               0 
0               0 

1* 

93          50 

31          42 

12            7 

9            7 

0 
1 
2 
2 

'Standard  deviation  not  calculated  where  x  <  1. 


NOTE:  N  =  number  of  subplots 


654 


Vegetation 


Intrahabitat  Heterogeneity  among  Patches.  The 

169  yd  (150  m)  subplots  along  transects  traversing 
stands  of  a  given  structural  type  will  reveal  much  of 
the  heterogeneity  within  a  habitat.  This  hetero- 
geneity reflects  differences  between  subplots 
(patches)  in  vertical  foliage  distribution  and  foliage 
density.  If  the  manager  decides  that  a  patch  is  any  5 
a.  (2  ha)  area,  the  classification  system  and  any 
subsequent  wildlife  use  analyses  will  be  entirely 
unsuitable  for  any  stand  less  than  5  a.  (2  ha).  The 
system  becomes  more  suitable  as  stand  size 
approaches  25  a.  (10  ha;  larger  sample  of  patches). 
For  example,  within  vegetation  classified  as  scrub 
habitat  with  patches  of  shrubs,  vertical  configuration 
in  the  various  subplots  more  frequently  resembles 
this  habitat  type  (Table  7)  than  any  other  habitat 
type.  However,  in  the  strictest  sense,  it  is  not 
appropriate  to  give  structural  type  designations  to 
patches  constituting  a  certain  structural  type. 
Structural  types  were  defined  on  the  basis  of 
transects  with  similar  average  vertical  foliage 
distributions.  The  foliage  distributions  were 
determined  on  the  basis  of  measurements  taken  in 
subplots.  That  it  is  wrong  to  classify  subplots  (or  any 
area  25  a.  [10  ha])  can  be  seen  from  the  following 
analogy. 

Suppose  that  all  books  in  six  private  libraries 
were  measured  and  mean  heights  were  found  to  be 
8.7,  9.1,  9.4,  9.8,  10.2,  and  10.6  in.  (22,  23,  24,  25, 
26,  and  27  cm)  for  collections  I  through  VI,  respec- 
tively. It  would  be  improper  to  conclude  from  exam- 
ining the  books  in  collection  IV  that  those  books 
<  94  in.(<  24  cm)  in  height  came  from  collections 
I,  II,  or  III,  or  that  all  books  >  9.8  in.  (>  26  cm)  in 
height  came  from  collections  V  and  VI.  Desirable 
as  it  might  be  to  know  the  origin  of  the  books  in 
collection  IV,  such  a  determination  simply  cannot  be 
made  from  the  evidence  presented. 

We  present  the  data  in  Table  7  merely  to  em- 
phasize that  ( 1 )  there  is  heterogeneity  between  sub- 
plots and  (2)  it  is  not  valid  to  obtain  foliage 
measurements  for  a  5  a.  (2  ha)  plot  and  then  be  able 


to  determine  its  vegetation  type.  This  would  be  anal- 
ogous to  having  a  book  10  in.  (26.2  cm)  in  height 
and  concluding  that  it  came  from  collection  V.  It 
really  could  have  come  from  any  of  the  collections. 
Desirable  as  it  may  be  to  classify  a  5  a.  ( 2  ha )  stand 
and  assess  the  wildlife  use  associated  with  it,  such 
a  determination  is  not  possible  with  the  data  pre- 
sented. This  must  be  compatible  with  the  objectives 
of  the  study.  A  classification  at  a  smaller  scale  will 
lead  inexorably  to  a  proliferation  of  vegetation  types 
recognized.  This  will  be  incompatible  with  the  ob- 
jective of  emphasizing  elements  of  similarity  be- 
tween stands  rather  than  differences.  More 
importantly,  we  have  learned  that  classification  at  a 
smaller  scale  can  lead  to  a  cloudy  or  erroneous  pic- 
ture of  how  wildlife  uses  vegetation. 

Cloudiness  begins  to  appear  at  a  scale  of  about 
50  a.  (20  ha;  Anderson  and  Ohmart,  unpubl.  data), 
and  an  opaqueness  emerges  from  analyses  at  a  scale 
of  about  5  a.  (2  ha;  Rosenberg  1980;  Engel-Wilson 
1982;  Anderson  and  Ohmart,  unpubl.  ms),  i.e.,  only 
weak  wildlife  use  patterns  are  discernible  af  a  scale 
of  <  50  a.  (<  20  ha),  and  either  wrong  impressions 
or  no  impression  emerges  at  a  scale  of  <  10  a. 
(<  2  ha).  Investigators  working  in  other  habitats 
have  reported  similar  findings  (Wiens  1981;  Wiens 
and  Rotenberry  1981a,  b). 

Some  of  our  conclusions  about  wildlife  use  of 
riparian  vegetation  along  the  lower  Colorado  River 
at  the  vegetation  (habitat)  type  scale  have  been  ex- 
perimentally tested  and  confirmed  (Meents  et  al. 
1982;  Anderson  and  Ohmart,  in  press-a).  Other  tests 
are  in  progress.  It  is  critical  to  clearly  define  the 
purpose  for  conducting  a  study  and  to  decide  on  the 
appropriate  scale.  A  classification  made  for  one  pur- 
pose will  probably  fail  when  used  for  other  than 
the  intended  purpose. 

Our  discussion  thus  far  has  centered  on  hetero- 
geneity within  habitats.  Next,  we  consider  the  differ- 
ences among  a  group  of  habitats. 


Table  7.     Subplots  of  six  recognized  structural  types. 


Structural  Type 


Subplots 


Number 


Percent  of  Subplots  of  Structural  Type 


II 


IV 


VI 


I 


IV 
V 
VI 


Open  gallery 

Closed  gallery 

Scrub  thickets 

Scrub  thickets  with  patchy  shrubs 

Shrubs-scrub  thickets 

Shrubs,  annuals,  grasses  with  scattered  trees 


36 
30 
154 
366 
291 
279 


38 

31 

9 

9 

13 

0 

0 

63 

20 

17 

0 

0 

6 

9 

66 

11 

9 

0 

1 

4 

23 

38 

19 

15 

0 

0 

8 

4 

59 

30 

0 

0 

0 

8 

18 

75 

NOTE:  Variation  in  vertical  configuration  among  subplots  within  each  structural  type  is  indicated  by  the  following  data.  See  text  for 
discussion  of  tautology  in  this  type  of  analysis. 


Vegetation 


655 


ANALYZING  HETEROGENEITY  AMONG 
HABITATS 

Although  many  of  the  differences  between  two 
habitats  may  be  obvious  to  the  observer  (a  patchy 
salt  cedar-scrub  thicket  is  obviously  different  in 
many  ways  from  a  cottonwood-willow  gallery  for- 
est), the  differences  need  to  be  quantified.  Although 
one  may  adequately  describe  the  differences  be- 
tween two  habitats,  such  a  description  may  require 
several  pages  and  cannot  be  readily  used  in  statistical 
treatments.  Therefore,  differences  must  be  expressed 
quantitatively.  Among  the  community  attributes 
measured,  several  may  be  intercorrelated,  i.e.,  as  the 
value  for  one  increases,  the  value  for  another  also 
increases.  When  colinearity  exists  between  variables, 
determining  the  extent  to  which  either  variable  is 
associated  with  wildlife  will  not  be  possible  without 
additional  experimentation  or  without  data  from 
some  area  where  these  variables  are  not  confounded. 
In  such  situations,  a  species  or  group  of  species 
could  be  significantly  associated  with  both  variables. 
In  reality,  one  of  the  variables  may  be  attracting 
the  species  and  the  other  one  may  not. 

Such  situations  are  probably  inevitable  with  any 
relatively  large  data  set,  and  little  can  be  done,  ex- 
cept to  recognize  that  some  constellations  of  com- 
munity attributes  are  positively  associated  with 
particular  species  or  groups  of  species.  Only  care- 
fully designed  experiments  will  divulge  which  attri- 
butes among  the  constellation  are  really  attracting 
wildlife.  Analysis  of  a  data  set  can  be  greatly  simpli- 
fied, however,  if  a  constellation  of  intercorrelated 
variables  can  be  quickly  recognized  and  treated  as  a 
single  variable. 

Principal  component  analysis  (PCA)  is  a  statisti- 
cal tool  that  combines  intercorrelated  variables  into 
new  derived  variables.  We  have  used  the  technique 
extensively;  details  are  given  elsewhere  (Anderson 
and  Ohmart  1984b).  The  derived  variables  can  usu- 
ally be  readily  interpreted,  and  they  can  be  treated 
as  independent  variables  in  subsequent  analyses. 
Each  habitat  receives  a  score  from  roughly  —  3  to 
+  3  on  each  derived  variable.  In  our  study,  PCA 
yielded  four  derived  variables.  The  first  included  foli- 
age density  and  diversity  measures  above  the  lowest 
layer  and  FHD.  Wildlife  associated  with  such  a  de- 
rived variable  is  most  abundant  in  habitats  with 
dense  foliage  that  is  horizontally  and  vertically  di- 
verse. Managing  wildlife  associated  with  this  derived 
variable  involves  creating  dense  and  diverse  areas. 

The  second  derived  variable  was  bipolar,  i.e., 
the  number  of  honey  mesquite  per  unit  area  was  as- 
sociated positively  with  this  component,  and  the 
number  of  salt  cedar  was  associated  negatively.  Spe- 
cies associated  with  this  derived  variable  were  posi- 
tively associated  with  honey  mesquite  but  negatively 
with  salt  cedar.  Negative  associations  indicated  a 


positive  association  with  salt  cedar,  but  no  associa- 
tion or  a  negative  one  with  honey  mesquite.  Manage- 
ment, in  the  first  instance,  might  involve  planting 
honey  mesquite  trees.  The  other  two  principal  com- 
ponents were  interpreted  in  similar  ways. 

PCA  can  be  used  to  compress  a  large  and  com- 
plex set  of  measurements  (vegetation  community 
attributes)  into  a  small  set  of  derived  variables  that 
can  be  used  as  independent  variables.  Associations 
between  wildlife  and  the  attributes  of  the  habitats 
can  be  determined  by  using  the  wildlife  populations 
associated  with  various  habitats  in  conjunction  with 
the  score  of  that  habitat  for  each  of  the  derived  vari- 
ables. Techniques  such  as  analysis  of  variance,  simple 
linear  correlation,  and  multiple  regression  are  appro- 
priate for  quantifying  the  extent  of  such  associations. 

Cluster  analyses  can  be  used  to  group  transects 
with  similar  vertical  configurations.  These  clusters 
can  be  further  subdivided  according  to  the  numeri- 
cally dominant  vegetation  present.  By  recognizing 
relatively  few  vertical  configurations  (stress  similari- 
ties rather  than  differences,  for  example,  6)  and  rela- 
tively few  subdivisions  by  dominant  vegetation 
(again,  6),  one  can  define  up  to  36  different  habitats, 
all  of  which  differ  from  one  another  by  vertical  con- 
figuration, dominant  vegetation,  or  both.  Recognizing 
relatively  few  vertical  configurations  (3  to  9)  and 
having  relatively  few  different  dominant  vegetation 
types  (3  to  9)  are  aids  to  the  memory.  Most  people 
can  keep  five  or  six  definitions  in  mind.  Thus,  when 
looking  at  a  stand  of  vegetation,  the  manager  needs 
to  answer  only  two  questions  to  classify  the  stand: 


( 1 )  What  is  the  vertical  configuration  of  the  vege- 
tation, i.e.,  is  it  four-layered,  three-layered, 
etc.? 

(2)  What  plant  species  appear  to  be  numerically 
dominant  in  the  stand? 

Thus,  in  a  short  time,  with  a  classification 
scheme  such  as  the  one  described  here,  the  manager 
can  acquire  enough  general  information  about  the 
stand  to  describe  it  in  detail  in  terms  of  the  classifi- 
cation system.  This  augments  communication  and 
may  be  useful  in  making  management  decisions 
when  such  decisions  need  to  be  made  quickly. 


DETERMINING  WILDLIFE-HABITAT 
ASSOCIATIONS 

One  can  look  for  relationships  between  a  given 
species  and  the  vegetation  attributes.  Conceivably, 
every  species  in  the  habitats  studied  could  be  in- 
volved in  such  analyses.  This  information  is  often 
useful  when  the  manager  has  collected  information 
on  a  single  species  that  is  of  particular  management 


656 


Vegetation 


interest.  For  example,  the  Yuma  clapper  rail  (Rallus 
longirostris  yumanensis),  which  is  on  the  endan- 
gered species  list,  is  of  great  interest  to  biologists 
along  the  Colorado  River.  We  found  (Anderson  and 
Ohmart  1985)  that  in  spring,  the  Yuma  clapper  rail 
was  linearly  associated  with  the  first  principal  com- 
ponent of  an  analysis  of  marsh  habitat  attributes.  The 
marsh  habitats  were  analyzed  according  to  the 
method  described  above.  The  habitat  breadth  of 
various  species  can  be  used  to  identify  habitat  spe- 
cialists or  those  species  with  narrow  habitat 
breadths.  The  data  in  Tables  8  and  9  identify  Bell's 
vireo  (Vireo  bellii),  summer  tanager  (Piranga 
rubra),  and  yellow-billed  cuckoo  (Coccyzus  ameri- 
canus)  as  habitat  specialists  and  cottonwood-willow 
woodlands  as  their  "preferred"  habitats  (Meents  et 
al.  1984). 


Variables  such  as  bird  species  diversity  (BSD) 
and  species  richness  may  also  be  related  to  the  de- 
rived variables.  We  found  significant  relationships 
during  most  seasons  for  BSD  and  species  richness 
using  various  multivariate  techniques  (Rice  et  al. 
1983).  Densities  of  birds  in  various  species  groups, 
such  as  permanent  resident  insectivores,  may  also  be 
related  to  the  derived  variables  during  one  or  more 
seasons. 


Here  we  add  a  cautionary  note.  Analyses  should 
be  conducted  over  several  seasons  and  years.  We 
have  found  that  analyses  over  just  one  season  or 
over  several  seasons  for  only  1  or  2  years  can  lead  to 
egregious  errors  in  assessing  habitat  importance  to 
wildlife  (Rice  et  al.  1980,  1983,  1984). 

Quite  frequently,  multiple  linear  regression  is 
used  as  an  analytical  tool.  However,  not  all  meaning- 
ful relationships  between  wildlife  and  habitat  attri- 
butes would  be  linear,  i.e.,  as  the  amount  of  the 
habitat  variable  increased,  the  wildlife  community 
attribute  would  increase  or  decrease.  For  example, 
we  found  that  curvilinear  relationships  accounted  for 
a  significant  amount  of  the  variance  for  19  common 
species  of  riparian  birds. 

We  quantified  deer  use  of  habitats  by  finding 
deer  use  areas  (foraging,  resting,  fawning,  etc.) 
and  analyzing  the  vegetation  in  plots  99  x  99  ft 
(30  x  30  m).  These  were  subdivided  into  four  sub- 
plots. Vegetation  measurements  and  tree  counts 
were  taken  within  these  subplots  in  the  manner  de- 
scribed above.  In  addition,  measurements  were  taken 
in  a  series  of  randomly  selected  plots.  Attributes  of 
the  vegetation  in  the  deer  use  plots  were  then  com- 
pared statistically  with  the  randomly  selected  control 
plots  (Haywood  et  al.  1984). 


Pinyon- juniper/sagebrush  association  showing  deer  feeding  area  next  to  cover  or  resting  areas. 


Vegetation  657 


> 

u 

o 

"S 

Ul 

o 

— ^ 

o 
u 

I 

u 
•S 

"8 

c 

3 

E 

a 

o 
<J 

c 
a 


a 

T3 


O 
O 


.C 
O 


c 
u 

T3 

U 

SP 

U 
3 


X 

1 


3 

rr 

CD 

> 

^ 

>^ 

> 

CD 

c 

o 

X 

— 

CO 
CD 
O 
Q) 
Q. 
CO 


m 


•<t      ncoowo      -t  co  co 
oj       *-  cvj  r-~  cm 


o 
cvj 


^  ^  cm  n-  o      ■>-  t-  cm 

CM  t-  t- 


CD 


*3- 


CO 
CO 


i-  co  co  in  o 

T-  ■* 


ONCDlO 

T-  ■* 


IflNOOOOr- 
t-  CO 


CM  CM  CO  CD  CO 

1-  i-  T-  CM 


CO  ID  CM 
LO  CM 


LO  Tt  O 

"fr  CM 


(COXO 
O  CM 


CO  h-  t- 
*-  <tf  CO 


CM  CO  CM  N-  N- 
CM  CM 


r*«  ■*  i-  1^.  N. 
CM  CM 


t-    Tt  tJ-  ■■-  |-~  |^ 


CM 


CD 
CO 


CO 
CO 


O 

CO 


lO  h-  CM  CM  CM 
CM  CM 


CM  CM 


>»  t-  i-  CD  CD 
CO  CO 


-3-  CD  CM 

CM  i- 


t-  LT)  O 

CO  tt- 


CM  ■*  CD 
CO 


CO  CM  CM 


LO  CD  in 

CM  r- 


COO)* 


O) 

h»  oo  co  ■«*•  i— 

CO 

oso 

t-  ■*  CM 

CD 

■T— 

•*  r^  •>-  co  f- 

co  co  co 

CM  CM 

CM    t- 


C0  CD  CO  CM  O    CDCMCO 


CO 


O  CO 


co       coun-^-T-o       m  in  o 
CM  t-  ^  ^t 


CO  CO  ^  -r-  CM  IflCDN 

CO  i—  CO  i— 


in        CM  CO  N-  CD  CO        -^  r-  co 

cm  t-  wr- 


en        CMin-r-l^-CO         LO  h-  CM 
CO  CM  CO  t- 


CO 

sconot 

CM 

CO  O  t- 
r-  00  CM 

CO 

in 

I"--  CD  CD  i—  00 

T-    CO    T- 

CM  •'J  CM 

o 

O  CM  t-  CM  CO 

x-  CO  ** 

CO  CO  CO 

CM 

CO 

o  o  cm  i-  in 

CO 

CO  i-  CD 

CM  CM 

>-  CD 

CD         ~  ^  "O    i_    CD 

dj        g3-c        ra  o  id  c)  > 
2  S  £-  nuo  ±  -o.e 

£o  c  «  ~  c  >-o  o  !=t5 


|  i£Msi*liii 


t-CDOCOOO         O         t-         O 

T-  CM     T-  ■* 


■tf  CO  O  CM  CM  O         O         Is-         O 


oiootco 


O  O  O  CO  CO  o 


COCJO^-i-O         O         CD         O 
t-  CM  t-  CM 


TTWi-como      m      o      co 

CM  "tf  CM  CO 


osomioo 

T~ 

o 

CM 

co  co  o  m  cm  o 

o 

CO 

CO  t-  O  CO  CO  o 

o 

CO 

o 

CD  CO  O  CO  CM  T- 
CM               r- 

'r~ 

CD 
CD 

COCOOtlfiO 

T-                ■*     T- 

o 

co 

CM 

CO  CO  O  CO  CD  CO 

^-       1- 

CM 

r^ 

** 

c 

CD 

w 

CD 

c 
o 


CD 
J»    CD-o 
-    O    ro    CD         "O 

CD  -c    C   ~    ,_    CD 

O  2    ££    g> 


cDcoocoino      in      cm      o 

t-  t-  CM 


NCO  r-  CDCDO         t-         CM         O 


CMOOCDCOO         CM  CO 

-r-  t-  y-  t-  ^ 


CONONNO         CD         in 

T-  T"  t^ 


o-^-incocoo       CO       CO       CM 
cm  t-  t-  ,-       r» 

co 


cDcoommo       ■>-       in       o 

T-  T-  CM 


CMCDOincDO       ^r       ■*       t- 

T-  T-     T-  CM 


CM  CM  CO  N-  O  CM    CO    r)" 
t-  t-    i-  CM  CO 


r^  co  co  m  cd  co   o   m   in 

x-  x-     t-  CM        t-     CO 


t-tj-ooocdo       co       co       in 

t-   t-  <M    i-   t-  y-  CD 


inTt-O-r-Oh-  O         CD         CD 

CO  t-  CM         'vf 


m    *-    <D  CD 

2  V&  o  o 

CO    CD  -S 

•      CD     j 

-    CO 


ro^tocD-e^^2f 
ooo.^coEcd0 

^50O5>5CDpc|: 

<        CQ  CO  _l  ZC0  CO 


O    CD 


■D 

CD    i  ^ 


w    <    o    " 


"°  "CD 

>- 


Tl 

, 

CD 
0) 

TO 
"D 

5 

o 

o 

co 

TO 

w 

II 

II 

rr 

I 

< 

(Xi 

658 


Vegetation 


Table  9.     Habitat  breadth  of  summer  bird  species  in 
riparian  communities  of  the  lower  Colorado 
River  Valley. 


Habitat 

Percent  of  Maximum 

Bird  Species 

Breadth* 

Habitat  Breadth 

Resident 

Abert  towhee 

1.28 

0.94 

Black-tailed 

gnatcatcher 

1.28 

0.94 

Crissal  thrasher 

1.30 

0.96 

Cactus  wren 

1.20 

0.88 

Gambel's  quail 

1.30 

0.96 

Gila  woodpecker 

0.95 

0.70 

Ladder-backed 

woodpecker 

1.20 

0.88 

Mourning  dove 

1.23 

0.90 

Verdin 

1.32 

0.97 

Nonresident 

Ash-throated 

flycatcher 

1.31 

0.96 

Blue  grosbeak 

1.33 

0.98 

Bell's  vireo 

0.62 

0.46 

Lucy  warbler 

1.26 

0.93 

Northern  oriole 

1.26 

0.93 

Summer  tanager 

0.66 

0.49 

Brown-crested 

flycatcher 

0.98 

0.72 

White-winged  dove 

1.12 

0.82 

Yellow-billed 

cuckoo 

0.86 

0.63 

"Habitat  breadth  calculated  using  thenequation: 
HB  =   -Sp,  log10  p, 


where  p,  = 

Habitat  Breadth  Mean  = 
Standard  Deviation  = 


the  proportion  of  a  species' 
population  occurring  in  community. 
1.137 
0226 


We  studied  lizard  use  of  the  landscape  over  a 
heterogeneous  75  a.  (30  ha)  area  by  scattering  plots 
measuring  10  x   10ft(3  x  3m).  These  were  subdi- 
vided into  1   x    1  yd  (m)  subplots.  Attributes  of  the 
vegetation  were  then  determined  by  the  methods 
described  above.  The  nature  of  the  substrate  (sand, 
hardpan,  etc. )  was  also  noted.  These  plots  were  vis- 
ited daily  at  the  time  of  peak  lizard  activity;  the  num- 
ber of  each  species  detected  was  recorded.  The 
characteristics  of  the  landscape  could  then  be  associ- 
ated with  greatest  numbers  of  detections  of  various 
lizard  species  (Anderson  and  Ohmart  1982).  Data 
obtained  from  can  traps  were  used  to  corroborate  or 
refute  observational  data. 

Rodent  association  with  various  habitats  was 
determined  using  multiple  regression  analysis  where 
relative  densities  of  each  rodent  species  were  the 
dependent  variables  and  the  vegetation  factor  scores 
for  each  habitat  (determined  from  PCA)  were  the 
independent  variables.  Curvilinear  relationships  were 
also  considered  (Anderson  and  Ohmart  1984a). 

Finally,  we  determined  the  impact  of  clearing 
vegetation  from  areas  by  first  obtaining  the  vegeta- 
tion attributes  of  control  and  experimental  areas 
(those  to  be  cleared)  before  clearing.  The  impact  of 
clearing,  according  to  different  patterns  and 
amounts,  was  determined  for  each  vegetation  attrib- 
ute separately  (FHD,  foliage  density,  etc.).  The  con- 
trols were  used  to  indicate  the  extent  of  change 
when  no  clearing  was  done.  We  also  noted  the 
change  in  principal  component  factor  scores  for 
each  of  the  habitats  affected.  Bird  and  rodent  num- 
bers were  obtained  before  and  after  clearing  in  ex- 
perimental and  control  areas  in  order  to  assess 
impacts  on  wildlife  in  various  habitats  (Anderson  and 
Ohmart,  in  press-a). 


Vegetation 


659 


LITERATURE  CITED 


ANDERSON,  B.W.  and  R.D.  OHMART.  1978.  Phainopepla 
utilization  of  honey  mesquite  forests  in  the  Colorado 
River  Valley.  Condor  80:334-338. 

and .  1982.  Revegetation  for  wildlife  en- 
hancement along  the  lower  Colorado  River.  Final  Rep. 
to  U.S.  Dep.  Inter.,  Bur  Reclamation,  Lower  Colorado 
Region,  Boulder  City,  NV. 

and .  1984a.  A  vegetation  management  study 

for  enhancement  of  wildlife  along  the  lower  Colorado 
River.  Final  Rep.  to  U.S.  Dep.  Inter.,  Bur.  Reclamation, 
Lower  Colorado  Region,  Boulder  City,  NV. 

and .  1984b.  Vegetation  community  type 

maps:  Lower  Colorado  River.  U.S.  Dep.  Inter.,  Bur.  Rec- 
lamation, Lower  Colorado  Region,  Boulder  City,  NV. 

and .  1985.  Habitat  use  by  clapper  rails  in 

the  lower  Colorado  River  Valley.  Condor  87:1 16-126. 

and .  In  press-a.  Evaluation  of  the  impact 

of  vegetation  removal.  Final  Rep.  to  U.S.  Dep.  Inter., 
Bur.  Reclamation,  Lower  Colorado  Region,  Boulder 
City,  NV. 

and .  In  press-b.  Riparian  vegetation  as  a 


mitigating  process  in  stream  and  river  restoration  in 
Gore,  J.A.,  ed.  The  Restoration  of  Rivers  and  Streams. 
Butterworth  Publ.,  Stoneham,  MA. 

and  W.C.  HUNTER.  1983-  Quantifying  variables  for 


classifying  desert  riparian  vegetation.  Pages  34-44  in 
Moir,  W.H.  and  E.L.  Hendzel,  tech.  coords.  Proc. 
Workshop  on  Southwestern  Habitat  Types.  U.S.  Dep. 
Agric,  For.  Serv.,  Albuquerque,  NM. 

CODY,  ML.  1974.  Competition  and  the  structure  of  bird 
communities.  Monogr.  Population  Biol.  7,  Princeton 
Univ.  Press,  Princeton,  NJ. 

CONINE,  K.  1982.  Avian  use  of  honey  mesquite,  interior 
and  agricultural  edge  habitat  in  the  lower  Colorado 
River.  M.S.  Thesis,  Arizona  State  Univ.  Tempe. 

DAUBENMIRE,  R.  1968.  Plant  communities:  A  textbook  of 
plant  synecology.  Harper  and  Row,  Publishers.  New 
York,  NY.  300pp. 

ENGEL-WILSON,  R.W.  1982.  Comparison  of  bird  commu- 
nities in  salt  cedar  and  honey  mesquite  along  the 
lower  Colorado,  Verde,  and  Gila  rivers.  M.S.  Thesis, 
Arizona  State  Univ.,  Dep.  Zoology,  Tempe. 

GREEN,  R.H.  1979.  Sampling  design  and  statistical  meth 
ods  for  environmental  biologists.  John  Wiley  &  Sons, 
New  York,  NY. 

HAYWOOD,  R.J.,  B.W.  ANDERSON,  and  R.D.  OHMART. 
1984.  Habitat  use  by  four  radio-collared  mule  deer 
along  the  lower  Colorado  River.  U.S.  Dep.  Inter.,  Bur. 
Reclamation,  Lower  Colorado  Region,  Boulder  City, 
NV. 

HORN,  H.S.  1966.  Measurement  of  "overlap"  in  compara- 
tive ecological  studies.  Am.  Nat.  100:419-424. 

KUCHLER,  AW.  1967.  Vegetation  mapping.  Ronald  Press, 
New  York,  NY. 

MACARTHUR,  R.H.  and  J.W.  MACARTHUR    1961.  On  bird 
species  diversity.  Ecology  42:594-598. 


MEENTS,  J.K.,  B.W.  ANDERSON,  and  R.D.  OHMART.  1982. 
Vegetation  relationships  and  food  of  sage  sparrows 
wintering  in  honey  mesquite  habitats.  Wilson  Bull. 
94:129-138. 

, ,  and .  1984.  Sensitivity  of  riparian 

birds  to  habitat  loss.  Pages  619-625  in  Warner,  RE. 
and  K.M.  Hendrix,  eds.  California  Riparian  Systems: 
Ecology,  Conservation,  and  Productive  Management. 
Univ.  California,  Berkeley. 

MUELLER-DOMBOIS,  D.  and  H  ELLENBERG.  1974.  Aims 
and  methods  of  vegetation  ecology.  John  Wiley  and 
Sons,  New  York,  NY.  547pp. 

OHMART,  R.D.,  WO.  DEASON,  and  C.  BURKE.  1977.  A 

riparian  case  history:  The  Colorado  River.  Pages  35-47 
in  Johnson,  R.R.  and  DA.  Jones,  tech.  coords.  Impor- 
tance, Preservation,  and  Management  of  Riparian 
Habitats:  A  Symposium.  U.S.  Dep.  Agric,  For.  Serv. 
Gen.  Tech.  Rep.  RM-43- 

PLATTS,  W.S.  1981.  Stream  and  inventory  garbage  in — 
reliable  analysis  out:  Only  in  fairy  tales.  Pages  75-85 
in  Armantrout,  N.B.,  ed.  Acquisition  and  Utilization  of 
Aquatic  Inventory  Information.  Western  Div.,  Amer. 
Fish.  Soc. 

,  W.F.  MEGAHAM,  and  G.W.  MARSHALL.  1983. 

Methods  for  evaluating  stream,  riparian,  and  biotic 
conditions.  U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep. 
138. 

RICE,  J.,  B.W.  ANDERSON,  and  R.D.  OHMART.  1980.  Sea- 
sonal habitat  selection  by  birds  in  the  lower  Colorado 
River  Valley.  Ecology  61:1402-141 1. 

, ,  and .  1983-  Habitat  selection  attri- 
butes of  an  avian  community:  A  discriminant  investi- 
gation. Ecol.  Monogr.  53263-290. 

, ,  and .  1 984.  Comparison  of  the 


importance  of  different  habitat  attributes  to  avian 
community  organization.  J.  Wildl.  Manage.  48:895- 
911. 

ROSENBERG,  K.V.  1980.  Breeding  bird  community  orga- 
nization in  a  desert  riparian  forest.  M.S.  Thesis,  Ari- 
zona State  Univ.,  Dep.  Zoology',  Tempe. 

SHANNON,  C.E.  and  W.  WEAVER.  1949.  The  mathematical 
theory  of  communication.  Univ.  Illinois  Press,  Urbana. 
125pp. 

U.S.  DEPARTMENT  OF  THE  INTERIOR,  BUREAU  OF 

LAND  MANAGEMENT.  1985.  Rangeland  Monitoring— 
Trend  Studies.  U.S.  Dep.  Inter.,  Bur.  of  Land  Manage. 
Technical  Reference  4400-4.  1 30pp. 

WHITTAKER,  R.H.  1975.  Communities  and  systems,  2nd 
edition.  MacMillan  and  Co.  New  York,  NY. 

WIENS,  J.A.  1981.  Scale  problems  in  avian  censusing.  Pages 
513-521  in  Ralph,  C.J.  andJ.M.  Scott,  eds.  Estimating 
Numbers  of  Terrestrial  Birds.  Studies  in  Avian 
Biology. 

and  J.T.  ROTENBERRY.  1981a.  Censusing  and  the 

evaluation  of  avian  habitat  occupancy.  Pages  522-532 
in  Ralph,  C.J.  and  J.M.  Scott,  eds.  Estimating  Numbers 
of  Terrestrial  Birds.  Studies  in  Avian  Biology. 

and .  1981b.  Habitat  associations  and  com- 


munity structure  of  birds  in  shrub-steppe  environ- 
ments. Ecol.  Monogr.  51:21-41. 


660 


Vegetation 


32 

MACRO- 
INVERTEBRATES 

Fred  Mangum 


U.S.  Forest  Service 
Aquatic  Laboratory 
Brigham  Young  University 
Provo,  UT  84602 


Editor's  Note:  Macroinvertebrate  surveys  are  one  of 
the  more  specialized  habitat  analysis  techniques. 
Macroinvertebrate  analyses  could,  in  theory,  be 
used  on  any  type  of  habitat;  in  practice,  however, 
the  technique  is  used  primarily  for  stream  habitat. 
In  aquatic  habitats,  macroinvertebrate  analyses 
have  proven  to  be  good  indicators  of  habitat  condi- 
tion and  water  quality.  In  the  future,  such  analyses 
may  be  used  for  other  purposes,  such  as  analyzing 
suitability  of  marshes  for  waterfowl. 


INTRODUCTION 

Macroinvertebrates  are  those  invertebrates  that 
can  be  detected  with  the  unaided  eye.  Although 
macroinvertebrates  are  an  important  habitat  compo- 
nent for  many  terrestrial  vertebrates,  such  as  insecti- 
vorous birds,  terrestrial  macroinvertebrates  are 
rarely  sampled  as  part  of  terrestrial  habitat  inventory 
and  monitoring  programs. 

On  the  other  hand,  macroinvertebrates  in  the 
aquatic  environment  provide  a  link  in  the  food  chain 
between  microscopic,  multicelled  organisms  and 
fish  populations.  They  are  essential  to  the  growth 
and  production  of  fish  and,  because  of  their  strict 
habitat  requirements,  are  very  useful  indicators  of 
aquatic  habitat  changes.  The  number,  size,  and  spe- 
cies of  aquatic  invertebrates  are  important  to  fisher- 
ies habitat,  as  they  are  the  primary  food  source  for 
most  salmonids  and  warm-water  fish.  Aquatic  inverte- 
brates include  insects,  such  as  mayflies,  stonefiles, 
caddisflies,  and  diptera  (two-winged  flies);  crusta- 
ceans, such  as  crayfish  and  shrimp;  mollusks,  such  as 
snails  and  clams;  and  fresh-water  earthworms. 

Macroinvertebrate  data  and  many  other  types  of 
scientific  information  should  be  considered  by  land 
managers  in  the  decisionmaking  process.  Inventory 
data  covering  physical,  chemical,  and  biological 
water  quality  and  aquatic  habitat  factors  can  be  ana- 
lyzed and  interpreted  in  a  manner  that  can  be  under- 
stood by  interdisciplinary  team  members  and 
planners.  Current  environmental  disturbances  should 
be  identified  as  well  as  potential  problems  that  could 
develop  if  new  management  activities  adversely  af- 
fect a  stream.  When  linked  to  other  factors,  macroin- 
vertebrate data  can  assist  the  decisionmaker  in 
assessing  impacts  to  stream  ecosystems  from  land 
management  activities. 

Positive  or  negative  effects  from  almost  every- 
thing that  occurs  in  a  drainage  show  up  in  streams 
and  affect  the  macroinvertebrate  community  compo- 
sition. The  community  composition  can  reveal  the 
health  of  an  ecosystem,  provide  a  warning  system,  or 
show  the  relative  intensity  of  pollution  problems. 
Macroinvertebrate  analysis  often  identifies  types  of 
pollution,  which  is  valuable  information  for  consci- 
entious land  managers.  To  make  use  of  all  available 


Macroinvertebrates 


661 


physical  and  biological  water  quality  and  aquatic 
habitat  data,  there  must  be  full  coordination  among 
specialists  and  management. 

Data  about  macroinvertebrates  are  site-specific. 
Aquatic  ecosystems  cannot  be  managed  on  the  basis 
of  average  values  over  large  diverse  areas.  Particu- 
larly for  grazing  prescriptions,  time  and  space  con- 
siderations within  a  watershed  must  be  considered 
in  the  planning  process  if  high  standards  of  aquatic 
habitat  and  water  quality  are  to  be  maintained. 

The  U.S.  Bureau  of  Land  Management  (BLM)  and 
the  U.S.  Forest  Service  have  used  macroinvertebrate 
data  to  assess  habitat  quality  and  nutrients  available 
for  resident  and  anadromous  fisheries.  Effects  from 
livestock  grazing,  dam  building,  mining,  timber  har- 
vest, rotenone  projects,  instream  improvement  struc- 
tures and  exclosures,  and  sanitary  facilities  in 
campgrounds  can  be  measured. 

This  chapter  describes  procedures  for  sampling 
aquatic  macroinvertebrates  and  analyzing  resulting 
data.  These  procedures  are  primarily  used  for  stream 
surveys;  however,  parts  of  the  procedure  may  be 
used  for  lake,  pond,  and  reservoir  surveys. 


MACROINVERTEBRATE  SURVEYS  FOR 
STREAMS 

Procedure 

Surveys  of  macroinvertebrates  consist  of  distinct 
phases:  field  sampling,  laboratory  analysis,  and  data 
analysis.  The  field  biologist  usually  does  the  field 
sampling  and  sends  samples  to  an  aquatic  laboratory 
for  analysis.  Field  sampling  methods  are  described 
here  as  well  as  data  analysis  and  interpretation  since 
many  of  the  latter  techniques  are  specific  to  aquatic 
macroinvertebrates. 

Sampling  Design.  When  designing  a  stream  survey 
for  macroinvertebrates,  consult  the  aquatic 
specialists  responsible  for  habitat  management.  This 
will  ensure  that  the  data  and  information  generated 
will  be  useful  in  answering  pertinent  management 
questions. 

A  minimum  of  three  stratified,  random  aquatic 
macroinvertebrate  samples  should  be  taken  at  each 
station,  as  described  in  the  techniques  section.  A 
station  is  defined  as  a  stream  reach  up  to  30  m  (100 
ft)  long.  The  samples  generally  provide  desired  com- 
munity representation  and  data  that  can  be  statisti- 
cally analyzed.  Four  samples  may  be  needed  for 
some  streams,  depending  on  the  statistical  reliability 
desired. 

The  macroinvertebrate  community  can  reveal 
the  quality  of  habitat  components  essential  to 


aquatic  fauna,  such  as  water  quality,  substrate  com- 
position, riparian  habitat  quality,  ecosystem  stability, 
and  past  history. 

Criteria  for  sampling  station  selection,  duration, 
and  frequency  differ  depending  on  the  basis  of  in- 
ventory or  monitoring  emphasis.  For  example,  if  an 
action  plan  was  developed  to  evaluate  a  new  grazing 
prescription  and  an  exclosure,  three  stations  should 
be  established — above,  within,  and  below  the  pro- 
posed exclosure  area.  The  "above"  station  should  be 
within  50  yards  of  the  exclosure;  the  "within"  sta- 
tion, on  the  lower  end  of  the  exclosure;  and  the 
"below"  station,  with  gradient  and  other  natural 
physical  characteristics  similar  to  the  "within"  sta- 
tion, generally  within  1.6  km  (1  mi.)  of  the  exclo- 
sure. Spring,  summer,  and  fall  samples  should  be 
taken  before  installing  the  exclosure;  sampling  on 
those  dates  should  be  repeated  for  3  to  5  years  fol- 
lowing exclosure  installation. 

When  inventorying  physical,  chemical,  and  bio- 
logical parameters,  I  recommend  establishing  a  mini- 
mum base  period  of  3  years  to  show  effects  of 
natural  extremes  in  physical  and  water  chemistry 
influences  on  an  ecosystem,  and  a  minimum  fre- 
quency of  three  dates  annually — spring,  summer,  and 
late  fall.  Additional  samples  may  be  taken  to  cover 
projected  heavy-use  periods. 

For  monitoring  or  surveillance  of  the  effects  of  a 
management  activity,  a  minimum  of  1  year  is  neces- 
sary, but  up  to  3  or  5  years  can  be  required  depend- 
ing on  the  activity  or  pollution  source  being 
monitored.  The  location  and  number  of  sampling 
stations  depend  on  the  management  activity  being 
monitored  and  its  duration  or  projected  commence- 
ment. Monitoring  requires  a  minimum  of  two  sta- 
tions— one  above  (a  control)  and  one  below  the 
activity  monitored.  A  third  station  downstream  helps 
define  downstream  impacts  or  recovery. 

To  more  specifically  locate  a  source  of  stream 
stress,  stations  can  be  added  within  the  management 
area  initially  or  upon  detection  of  stress  conditions. 
Point  or  nonpoint  pollution  sources  may  be  moni- 
tored. To  show  the  quality  of  the  habitat  and  the 
water  as  it  leaves  public  land,  management  plans  for 
some  streams  may  be  adequately  served  by  a  single 
station.  Control  stations  should  be  placed  in  a  riffle, 
if  possible,  between  45  and  180  m  (50  and  200  yd) 
above  the  upper  boundaries  of  the  suspected  pollu- 
tion sources. 

Experimental  stations  should  include  at  least 
one  station  below  the  lower  boundaries  of  the  sus- 
pected source.  Stations  may  be  placed  in  two  cate- 
gories depending  on  their  projected  use:  ( 1 )  long- 
term  (permanent)  stations  established  on  the  main 
stream  or  major  tributary  in  each  drainage  to  moni- 
tor long-term  projects  or  continuing  land-use  activi- 


662 


Macroinvertebrates 


ties  that,  by  their  nature,  could  cause  on-site  or 
downstream,  accumulated  impacts  or  stress  or  (2) 
short-term  (temporary)  stations  established  to  moni- 
tor projects  such  as  timber  harvest,  prescribed 
burns,  insect  spray  projects,  or  others  having  impacts 
over  a  definite  time  period.  All  such  monitoring 
should  include  predisturbance  (before),  operation 
(during),  and  post-operation  (after)  sampling. 

Substrate  Sampled.  Samples  are  taken  in  riffle  areas 
with  streambed  substrate  of  gravel  or  rubble.  A 
majority  of  the  species  occurring  in  the  stream's 
aquatic  macroinvertebrate  community  will  be  found 
there.  Replicate  samples  should  be  taken  within  a 
100-foot  stretch  of  a  stream,  with  each  sample  being 
taken  in  locations  where  velocity  and  depth  are  as 
similar  as  possible. 

Equipment.  Equipment  needed  for  macroin- 
vertebrate sampling  is  listed  below: 

•  Modified  Surber  Net 

•  250-micron  sieve 

•  Three  plastic  bottles  per  station  with  strip  of 
masking  tape  attached  for  identification  pur- 
poses 

•  Preserving  solution 

•  Hip  boots 

•  Saturated  saltwater  solution 

•  Two  aluminum  bread  pans 

•  Waterproof  gloves 

•  Laundry  pen  (waterproof  marker) 


Field  Sampling. 


Modified  Surber  Net  Samples.  The  modified 
Surber  square -foot  sample  net  (  Figure  1 )  has  proven 
to  have  a  correlation  coefficient  equal  to  or  better 
than  other  currently  used  sampling  devices.  I  recom- 
mend using  a  280-micron  mesh,  nylon  net  when 
sampling  all  macroinvertebrates.  Three  steps  should 
be  followed  when  using  this  modification. 

1.  The  foot-square  Surber  frame  is  placed  over  the 
gravel-rubble  substrate  in  the  stream  with  the 
net  downstream.  As  the  rocks  within  the  frame 
are  hand  scrubbed,  macroinvertebrates  are  car- 
ried into  the  net  by  the  flowing  water.  The  sub- 
strate underlying  the  gravel-rubble  is  also  stirred 
by  hand  to  a  depth  of  7  to  10  cm  (3  to  4  in.), 

if  possible. 

2.  After  allowing  the  water  to  drain  from  the  net, 
the  net  is  inverted  into  an  aluminum  pan  con- 
taining a  saturated  saltwater  solution.  As  the 
saltwater  is  poured  into  a  second  pan,  the  or- 
ganic materials  thus  floated  are  caught  in  a  250- 
micron  mesh  sieve.  The  saltwater  is  then  poured 
back  into  the  first  pan,  the  contents  are  again 
vigorously  stirred,  and  the  floating  materials  and 
specimens  are  poured  for  a  second  time  through 
the  sample  sieve.  The  sample  may  require  siev- 
ing two  or  three  times. 

3.  The  sample  in  the  sieve  is  then  washed  from  the 
sieve  pan  into  the  sample  bottle  with  an  alcohol 
solution.  The  alcohol  solution  is  mixed  at  a  ratio 


Nitex  280  micron  mesh  netting 


Viny 


1"x  3/16"  aluminum  frame 


Figure  1.     Modified  Surber  sampler. 


Macroinvertebrates 


663 


of  1  cup  of  10%  formalin  to  38  L  (  1  gal.)  of  70% 
ethanol.  Enough  alcohol  should  be  added  to 
the  sample  bottle  to  cover  the  sample.  Caddis- 
flies  in  their  cases  must  be  handpicked  from  the 
pan  and  added  to  the  sample. 

Artificial  Substrate  Basket  Samples.  The  use 

of  artificial  substrate  sampling  baskets  is  also  recom- 
mended by  the  U.S.  Environmental  Protection 
Agency  because  of  their  comparatively  low  coeffi- 
cient of  variation  in  collection  of  macroinvertebrate 
individuals  and  taxa  (see  Figure  2). 

The  following  methods  are  recommended  for 
basket  sampling: 

1 .  Placement.  The  baskets,  filled  with  two  layers  of 
baseball-sized  rubble  gathered  from  the  immedi- 
ate stream  area,  are  worked  down  into  the  sub- 
strate in  a  riffle  area  among  rocks  of  nearly  the 
same  size,  if  possible.  The  rocks  may  be  col- 
lected along  the  bank  of  the  stream  or  in  the 
stream  and  placed  in  the  basket.  If  the  rocks  are 
gathered  from  the  stream  substrate,  they  should 
be  scrubbed  clean  to  remove  any  existing  in- 
sects. For  statistical  purposes,  three  baskets  are 
placed  at  each  station.  Each  station  should  not 
exceed  a  100-foot  distance  between  the  up- 
stream and  downstream  baskets. 

2.  Retrieval.  After  a  1 -month  (not  to  exceed  33 
days)  colonization  period,  samples  are  collected 
from  the  baskets.  The  baskets  should  be  ap- 
proached from  their  downstream  side,  the  Sur- 
ber  net  placed  over  the  basket,  and  the  basket 
pulled  into  the  net.  With  one  hand  on  the  bas- 
ket and  the  other  holding  the  net,  both  are 
pulled  from  the  stream  in  an  upstream  motion. 
The  rocks  from  the  baskets  are  placed  in  a 
bucket  half  filled  with  water.  The  Surber  net  is 
inverted  and  washed  in  the  bucket  of  water,  and 
the  rocks  are  individually  scrubbed  with  a 
brush.  After  scrubbing,  the  rocks  are  placed 
back  into  the  basket  that  is  then  repositioned 
into  the  stream.  The  basket  should  be  covered 
with  water  and  camouflaged,  if  possible,  to  avoid 
vandalism.  The  water  and  sample  in  the  bucket 
are  poured  into  a  250-micron  mesh  sieve  to 
remove  the  water.  If  a  lot  of  debris  and  sand  are 
present,  the  sample  can  be  placed  in  a  saturated 
saltwater  solution  that  will  float  all  organisms 
out  of  the  sand  and  gravel.  The  sample  is 
washed  into  the  sample  bottle  from  the  sieve 
with  alcohol.  Alcohol  is  added  to  the  bottle 
until  it  just  covers  the  sample. 

Supporting  Data.  Several  types  of  supporting 
data  are  commonly  collected  with  macroinvertebrate 
samples.  Water  chemistry  and  physical  habitat  data 
are  integrated  into  the  elevation  to  bring  the  total 


Figure  2.     Artificial  substrate  basket. 


biological  analysis  into  perspective  for  the  biologist 
and  land  manager.  In  particular,  data  on  alkalinity, 
sulfates,  gradient,  and  substrate  composition  are  nec- 
essary for  calculating  a  Biotic  Condition  Index  (BCI). 
The  following  data  are  commonly  collected  with 
macroinvertebrate  data: 

1 .    Specific  Conductivity.  Specific  conductivity, 
measured  as  umhos/cm  at  25°C  (77°F),  is  partic- 
ularly important  for  streams  with  high  alkalinity 
and  hardness. 


Alkalinity.  Alkalinity,  measured  as  HC03  con- 
centration in  mg/1,  is  closely  linked  to  stream 
productivity  and  is  necessary  to  determine  the 
BCI.  Alkalinity  correlates  with  community  den- 
sity, diversity,  and  biomass,  and  also  affects  the 
primary  production  in  a  stream. 


3.  Water  Temperature.  Temperature  is  a  limiting 
factor  on  certain  sites. 

4.  Nitrate  Nitrogen.  Nitrate  nitrogen,  measured  as 
mg/1,  may  indicate  organic  enrichment. 

5.  Orthophosphate.  Orthophosphate,  measured  as 
mg/1,  may  indicate  organic  enrichment. 


6.   Sulfate.  Sulfate,  measured  as  mg/1,  may  indicate 
sources  of  accretion  flows.  It  is  a  factor  affecting 


664 


Macroinvertebrates 


community  composition  and  is  necessary  to 
determine  the  BCI.  An  increase  in  sulfate 
generally  indicates  a  deterioration  of  natural 
water  quality.  Many  poor  quality  waters  in 
the  western  U.S.  have  high  sulfate  levels. 


7.    pH.  High  or  low  pH  also  limits  biological 
productivity. 


Gradient.  Gradient,  measured  as  a  percent,  is 
related  to  a  stream's  ability  to  maintain  sub- 
strate quality.  It  affects  community  composi- 
tion and  is  used  in  BCI  analyses  because  of  its 
positive  correlation  with  macroinvertebrate 
community  diversity. 


Elevation.  Elevation  is  used  to  define  ranges 
for  aquatic  species. 


10.    Substrate  Composition.  Substrate  composi- 
tion, measured  as  the  relative  composition 
(ranking)  of  streambed  boulder,  rubble, 
gravel,  and  sand-silt  materials,  is  necessary  to 
determine  the  BCI.  Rubble  substrates  are 
preferred  by  the  greatest  number  of  species 
because  they  provide  a  greater  diversity  of 
microhabitat  than  smaller  substrates. 


11.    Streamside  Vegetation.  Streamside  vegetation, 
measured  as  the  relative  dominance  (rank) 
of  grass,  shrubs,  brush,  deciduous  trees,  and 
evergreen  trees  present  on  streambank  ripar- 
ian areas,  affects  stream  stability  and  nutrient 
sources. 


12.    Flow  Rates.  Mean  summer  and  mean  low-flow 
rates,  measured  in  ft  /sec,  significantly  affect 
macroinvertebrate  community  structure. 
Methods  for  measuring  each  of  these  varia- 
bles are  described  in  Chapters  28  and  30. 


Sample  Analysis.  Analyses  are  based  on  several 
factors  such  as  the — 

•  Diversity  and  taxa  index  (DAT); 

•  Biomass  of  the  macroinvertebrates  per  sample; 

•  Numbers  of  organisms  per  species; 

•  Species  and  trophic  balance  in  the  community; 

•  Abundance  and  degree  of  dominance  among 
indicator  species;  and 

•  Tolerances  of  species  in  the  community,  as  indi- 
cated by  the  community  tolerance  quotient 
(CTQ). 

The  BCI  suggests  a  management  strategy  that 
integrates  biological,  chemical,  and  physical  factors 
present  in  and  adjacent  to  the  stream's  environment. 


Analysis  Elements  (Major).  The  following 
three  major  analysis  elements  will  be  found  on  data 
sheets  included  in  reports,  along  with  scales  indicat- 
ing the  significance  of  these  data  values: 

1.   BCI.  The  BCI  measures  a  stream  against  its  own 
potential  and  not  that  of  other  streams.  The 
BCI  is  based  on  mean  community  tolerance, 
composed  of  the  tolerances  of  individual  taxa  or 
species  found  in  the  community,  which  varies 
in  response  to  intensity  of  perturbations  in  the 
ecosystem. 

The  BCI  is  calculated  as  follows: 

a.  Determine 

(  1 )  Gradient  Percentage 

( 2 )  Relative  Substrate  Dominance  given  a 
ranking  from  1  through  4 
Boulder     >  30  cm  (over  12  in.) 

Rubble     7.5-30         cm  (3-12  in.) 
Gravel     0.3-7.5  cm  (1/8-2.99  in.) 

Sand-silt    <  0.3  cm  (<  1/8  in.) 


Analysis 

Laboratory  Analysis.  Normally  samples  are  sent  to 
regional  laboratories  for  processing  and  analysis. 
Samples  should  be  properly  labeled  with  agency, 
administrative  unit  location,  stream,  station,  and  date. 


Laboratory  personnel  analyze  these  data.  These 
analyses  are  described  below  since  the  field  biolo- 
gists are  probably  familiar  with  how  they  are  derived 
and  interpreted  even  if  they  do  not  analyze  the  raw 
data. 


(3)  Total  Alkalinity  (mg/1) 

(4)  Sulfate  (mg/1 ) 

b.  Determine,  using  the  above  information  and 
the  key  in  Table  1 ,  the  predicted  community 
tolerance  quotient  (CTQP)  for  the  reach  on 
the  study  stream. 

c.  List,  using  at  least  three  or  four  macroinverte- 
brate samples  at  a  station,  the  taxa  and  their 
associated  tolerance  quotients  (TQ) 

(Table  2). 


Macroinvertebrates 


665 


d.  Add  the  TQ  values  and  divide  by  the  number 
of  taxa  to  get  the  actual  community  tolerance 
quotient 


The  BCI  is 


STQs 


CTQa 


CTO 
e.  Determine  the  BCI  =        ^p  x 

L*  1  V^a 


100 


The  scale  of  BCI  values  listed  below  can  be 
used  as  indexes  or  values  to  determine  the 
health  of  a  stream  ecosystem  for  which  a  de- 
fined management  strategy  can  be  based  for 
each  stream  reach. 


Scale 


BCI 


Excellent 

Above  90 

Good 

75-90 

Fair 

Below  75 

Poor 

Below  75 

Sensitive  to  all  types  of  environmental  stress; 

Applicable  to  various  types  of  streams; 

Capable  of  giving  a  linear  assessment  from  un- 
stressed to  highly  stressed  conditions; 

Independent  of  sample  size,  providing  the  sam- 
ple contains  a  representative  assemblage  of 
species; 

Based  on  data  readily  available  or  easily  ac- 
quired; and 

Capable  of  meshing  with  and  supporting  exist- 
ing stream  habitat  or  water  quality  management 
programs. 

Dominance  and  Taxa  Diversity  Index  (DAT). 
This  index  combines  a  measure  of  species  domi- 
nance in  the  community  with  the  number  of 
species  present.  Many  other  indexes  measure 
one  or  the  other  of  these  aspects. 


Table  1.     A  key  giving  predicted  community  tolerance  quotients  (CTQp's)  for  various  combinations  of 

gradient  (%),  substrates,  total  alkalinity  (mg/1  CaC03),  and  sulfate  (mg/1  S04)  for  any  given  stream  (from 
U.S.  Department  of  Agriculture,  Forest  Service  1981). 


Combinations 

Goto 
Key  No. 

CTQp 

1 .  Stream  gradient  0.1—1.2  

2 

51 

1 .3-3.0  

15 

>3.0 

2.  Substrate  mostly  boulder  &  rubble 

Gravel  &  rubble 

28 

3 

7 

Sand  &  boulder, 
rubble,  or  gravel 

3.  Total  alkalinity  0-1 99 

11 

4 

200-300  

>300 

4.  Sulfate  0-1 49 

5 

6 

150-300 

71 

>300 

90 

5.  Sulfate  0-1 49 

53 

150-300 

71 

>300 

90 

6.  Sulfate  0-1 49 

90 

150-300 

96 

>300 

108 

7.  Total  alkalinity  0-1 99 

8 

53 

200-300 

>300 

9 

10 

8.  Sulfate  0-149 

150-300 

85 

>300 

103 

666 


Macroinvertebrates 


Table  1.  A  key  giving  predicted  community  tolerance  quotients  (CTQps)  for  various  combinations  of 
gradient  (%  ),  substrates,  total  alkalinity  (mg/1  CaC03),  and  sulfate  (mg/1  S04)  for  any  given  stream 
(continued). 


Combinations 

Goto 
Key  No. 

CTQP 

9.  Sulfate  0-1 49 

55 

150-300 

.    86 

>300 

..103 

10.  Sulfate  0-149 

89 

150-300 

97 

>300 

..108 

1 1 .  Total  alkalinity  0-1 99 

12 

200-300  

13 

>300 

14 

12.  Sulfate  0-149 

60 

150-300 

.    90 

>300 

..108 

13.  Sulfate  0-149 

60 

150-300 

90 

>300 

..108 

14.  Sulfate  0-149 

90 

150-300 

.    99 

>300 

..108 

1 5.  Substrate  mostly  boulder  &  rubble 

16 

50 

Gravel  &  rubble 

Sand  &  boulder, 
rubble,  or  gravel 

1 6.  Total  alkalinity  0-1 99 

20 

24 

17 

200-300 

>300 

17.  Sulfate  0-149  

18 

19 

150-300 

65 

>300 

90 

18.  Sulfate  0-149 

50 

150-300 

65 

>300 

.    90 

19.  Sulfate  0-199 

.    90 

150-300 

.    96 

>300 

108 

20.  Total  alkalinity  0-1 99 

200-300 

>300 

21 .  Sulfate  0-1 49 

21 

22 

230 

50 

150-300 

.    80 

>300 

103 

22.  Sulfate  0-1 49 

55 

150-300 

80 

>300 

108 

23.  Sulfate  0-1 49 

80 

200-300 

96 

>300 

108 

24.  Total  alkalinity  0-1 99 

200-300 

>300 

25 

26 

27 

Macroinvertebrates 


667 


Table  1.  A  key  giving  predicted  community  tolerance  quotients  (CTQp's)  for  various  combinations  of 
gradient  (%  ),  substrates,  total  alkalinity  (mg/1  CaC03),  and  sulfate  (mg/1  S04)  for  any  given  stream 
(concluded). 


Combinations 

Goto 
Key  No. 

CTQP 

25.  Sulfate  0-1 49 

66 

150-300 

88 

>300 

..108 

26.  Sulfate  0-1 49 

65 

150-300 

.    88 

>300 

1 08 

27.  Sulfate  0-1 49 

85 

150-300 

93 

>300 

108 

28.  Substrate  mostly  boulder  &  rubble 

Gravel  &  rubble 

Sand  &  boulder, 

rubble,  or  gravel 

29.  Total  alkalinity  0-1 99 

200-300 

>300 

30.  Sulfate  0-1 49 

29 

33 

37 

30 

31 

32 

50 

150-300 

62 

>300 

100 

31 .  Sulfate  0-1 49 

50 

150-300 

62 

>300 

108 

32.  Sulfate  0-1 49 

85 

150-300 

90 

>300 

108 

33.  Total  alkalinity  0-1 99 

200-300 

>300 

34.  Sulfate  0-1 49 

34 

35 

36 

50 

150-300 

77 

>300 

108 

35.  Sulfate  0-1 49 

50 

150-300 

77 

>300 

108 

36.  Sulfate  0-1 49 

90 

150-300 

99 

>300 

108 

37.  Total  alkalinity  0-1 99 

200-300 

>300 

38.  Sulfate  0-1 49 

25 

26 

27 

80 

150-300 

100 

>300 

1 08 

39.  Sulfate  0-1 49 

80 

150-300 

....  1 00 

>300 

..108 

40.  Sulfate  0-1 49 

100 

150-300 

108 

>300 

1 08 

668 


Macroinvertebmtes 


Table  2.     Tolerance  quotients  (TQs)  of  aquatic  macroinvertebrates  based  on  tolerance  to  alkalinity,  sulfate, 
and  sedimentation  including  low  stream  gradients  (from  U.S.  Department  of  Agriculture,  Forest  Service 
1981). 


Taxa 

TQ 

Family  Tipulidae 

72 

Antocha  monticola 

24 

Dicranota  sp. 

24 

Hexatoma  sp. 

36 

Holorusia  grandis 

72 

Helobia  sp. 

36 

Tipula  sp. 

36 

Family  Psychodidae 

36 

Maruina  sp. 

36 

Psychoda  sp. 

36 

Pericoma  sp. 

36 

Family  Blephariceridae 

2 

Bibiocephala  grandis 

2 

Agathon  sp. 

2 

Family  Deuterophlebiidae 

4 

Deuterophlebia  coloradensis 

4 

Family  Culicidae 

108 

Aedes  sp. 

108 

Culex  sp. 

108 

Anopheles  sp. 

108 

Mansonia  sp. 

108 

Psorophora  sp. 

108 

Culiseta  sp. 

108 

Family  Dixidae 

108 

Dixa  sp. 

108 

Family  Simuliidae 

108 

Family  Chironomidae 

108 

Family  Ceratopogonidae 

108 

Family  Stratiomyidae 

108 

Euparyphus  sp. 

108 

Family  Tabanidae 

108 

Tabanus  sp. 

108 

Family  Rhagionidae 

24 

Atherix  pachypus 

24 

Family  Dolichopodidae 

108 

Family  Empididae 

108 

Hemerodromia  sp. 

108 

Family  Ephydrinae 

108 

Ep hydra  sp. 

108 

Family  Muscidae 

108 

Limnophora  sp. 

108 

Family  Syrphidae 

108 

Chrysogaster  sp. 

108 

Tubifera  sp. 

108 

Helophilus  sp. 

108 

Family  Dytiscidae 

72 

Derovatellus  sp. 

72 

Laccophilus  sp. 

72 

Bidessus  sp. 

72 

Agabus  sp. 

72 

Hygrotus  sp. 

72 

Hydroporous  sp. 

72 

Oreodytes  sp. 

72 

Illy bi us  sp. 

72 

Rhantus  sp. 

72 

Dytiscus  sp. 

72 

Acilius  sp. 

72 

Cybister  sp. 

72 

Deronectes  sp. 

72 

Taxa 

TQ 

Family  Dytiscidae  (continued) 

Thermonectes  sp. 

72 

Coptotomus  sp. 

72 

Family  Hydrophilidae 

72 

Helophorus  sp. 

72 

Hydrochara  sp. 

72 

Berosus  sp. 

72 

Enochrus  sp. 

72 

Hydrophilus  sp. 

72 

Tropisternus  sp. 

72 

Hydrobius  sp. 

72 

Paracymus  sp. 

72 

Crenitus  sp. 

72 

Ametor  sp. 

72 

Helochares  sp. 

72 

Laccobius  sp. 

72 

Enochrus  sp. 

72 

Cymbiodyta  sp. 

72 

Family  Elmidae 

108 

Zaitzevia  sp. 

Narpus  sp. 

Stenelmis  sp. 

Dubiraphia  sp. 

Optioservus  sp. 

108 

Heterlimnius  sp. 

£/m/s  sp. 

Simsonia  sp. 

Microcylloepus  sp. 

Z_ara  sp. 

Family  Gyrinidae 

108 

Gyrinus  sp. 

108 

Family  Amphizoidae 

24 

Amphizoa  sp. 

24 

Family  Hydraenidae 

72 

Order  Diptera 

Family  Hydroptilidae 

Leucotrichia  sp. 

108 

Alisotrichia  sp. 

108 

Mayatrichia  sp. 

108 

Family  Limnephilidae 

108 

Limnephilus  sp. 

108 

Dicosmoecus  sp. 

24 

Hesperophylax  sp. 

108 

Oligophlebodes  sp. 

24 

Apatania  sp. 

18 

Amphicosmoecus  sp. 

18 

Neothremma  sp. 

8 

Lenarchus  sp. 

18 

Chyranda  sp. 

18 

Psychoglypha  sp. 

24 

Ecclisomyia  sp. 

24 

Homophylax  sp. 

18 

Allocosmoecus  sp. 

18 

Asynarchus  sp. 

108 

Clistoronia  sp. 

108 

Grammotaulius  sp. 

108 

Imania  sp. 

48 

Neophylax  sp. 

24 

Onocosmoecus  sp. 

18 

Pycnopsyche  sp. 

72 

Macroinvertebrates 


669 


Table  2.     Tolerance  quotients  (TQs)  of  aquatic  macroinvertebrates  based  on  tolerance  to  alkalinity,  sulfate, 
and  sedimentation  including  low  stream  gradients  (continued). 


Taxa 

TQ 

Family  Leptoceridae 

54 

Oecetis  sp. 

54 

Leptocella  sp. 

54 

Triaenodes  sp. 

54 

Mystacides  sp. 

54 

Ceraclea  sp. 

54 

Family  Lepidostomatidae 

18 

Lepidostoma  sp. 

18 

Family  Brachycentridae 

24 

Brachycentrus  sp. 

24 

Micrasema  sp. 

24 

Oligoplectrum  sp. 

24 

Amiocentrus  sp. 

24 

Family  Helicopsychidae 

18 

Helicopsyche  borealis 

18 

Family  Polycentropodidae 

72 

Polycentropus  sp. 

72 

Nyctiophylax  sp. 

72 

Family  Sericostomatidae 

72 

Gumaga  sp. 

72 

Order  Coleoptera 

Family  Haliplidae 

54 

Brychius  sp. 

54 

Haliplus  sp. 

54 

Peltodytes  sp. 

54 

Family  Periodidae 

Pictetiella  expansa 

18 

Diura  knowltoni 

24 

Isoperla  sp. 

48 

/.  ebria 

24 

1.  fulva 

48 

1.  mormona 

48 

1.  quinquepunctata 

48 

Family  Chloroperlidae 

24 

Family  Perlidae 

24 

Acroneuria  abnormis 

6 

Claassenia  sabulosa 

6 

Hesperoperla  pacifica 

18 

Perlesta  placida 

24 

Doroneuria  theodora 

18 

Order  Trichoptera 

Family  Rhyacophilidae 

18 

Rhyacophila  sp. 

18 

Atopsyche  sp. 

18 

Himalopsyche  sp. 

18 

Family  Glossosomatidae 

32 

Glossosoma  sp. 

24 

Anagapetus  sp. 

24 

Protoptila  sp. 

32 

Culoptila  sp. 

32 

Family  Philopotamidae 

24 

Chimarra  sp. 

24 

Dolophilodes  (sortosa) 

24 

Wormaldia  sp. 

24 

Family  Psychomyiidae 

108 

Polycentropus  sp. 

108 

Nyctiophylax  sp. 

108 

Psychomyia  sp. 

108 

Ti nodes  sp. 

108 

Taxa 

TQ 

Family  Hydropsychidae 

108 

Hydropsyche  sp. 

108 

Cheumatopsyche  sp. 

108 

Arctopsyche  sp. 

18 

Smicridea  sp. 

72 

Diplectrona  sp. 

48 

Macronema  sp. 

48 

Parapsyche  sp. 

6 

Family  Hydroptilidae 

108 

Hydroptila  sp. 

108 

Agralyea  sp. 

108 

Ochrotrichia  sp. 

108 

Stactobiella  sp. 

108 

Neotrichia  sp. 

108 

Ithytrichia  sp. 

108 

Oxyethira  sp. 

108 

Order  Ephemeroptera 

Family  Gerridae 

72 

Gem's  sp. 

72 

Rheumatobates  sp. 

72 

Family  Naucoridae 

72 

Ambrysus  mormon 

72 

Pelocoris  sp. 

72 

Family  Notonectidae 

108 

Notonecta  sp. 

108 

Buenoa  sp. 

108 

Family  Veliidae 

72 

Microvelia  americana 

72 

Rhagovelia  distincta 

72 

Family  Mesoveliidae 

72 

Mesovelia  sp. 

72 

Family  Macroveliidae 

72 

Macrovelia  sp. 

72 

Order  Plecoptera 

Family  Nemouridae 

36 

Amphinemura 

6 

Malenka 

36 

Prostoia  besametsa 

24 

Podmosta 

12 

Zapada 

16 

Nemoura 

24 

Family  Capniidae 

32 

Capnia 

32 

Eucapnopsis 

18 

Isocapnia 

24 

Mesocapnia  frisonl 

32 

Utacapnia 

18 

Family  Taeniopterygidae 

48 

Taenionema 

48 

Doddsia 

24 

Oemopteryx 

48 

Family  Leuctridae 

18 

Paraleuctra  sp. 

18 

Perlomyia  sp. 

18 

Family  Pteronarcyidae 

24 

Pteronarcella  badia 

24 

Pteronarcys  californica 

18 

P.  princeps 

24 

Family  Periodidae 

48 

670 


Macroinvertebrates 


Table  2.     Tolerance  quotients  (TQs)  of  aquatic  macroinvertebrates  based  on  tolerance  to  alkalinity,  sulfate, 
and  sedimentation  including  low  stream  gradients  (concluded). 


Taxa 

TQ 

Family  Perlodidae  (continued) 

Megarcys  signata 

24 

Skwala  parallela 

18 

Cultus  aestivalis 

12 

Isogenoides  sp. 

24 

/.  elongatus 

24 

1.  zionensis 

24 

Kogotus  modestus 

18 

Order  Megaloptera 

72 

Family  Sialidae 

72 

Sialis  sp. 

72 

Family  Corydalidae 

72 

Corydalus  cognata 

72 

Order  Lepidoptera 

72 

Family  Pyralidae 

72 

Paragyractis  kearfottalis 

72 

Order  Ephemeroptera 

72 

Family  Siphlonuridae 

72 

Ameletus  sp. 

48 

Siphlonurus  occidentalis 

72 

Isonychia 

48 

Family  Baetidae 

72 

Baetis  sp. 

72 

Callibaetis  sp. 

72 

Pseudocloeon  sp. 

72 

Centroptilum  sp. 

36 

Dactylobaetis  sp. 

36 

Paracloeodes  sp. 

72 

Family  Oligoneuriidae 

36 

Lachlania  saskatchewanensis 

36 

Homoeoneuria  sp. 

36 

Family  Heptageniidae 

48 

Heptagenia  sp. 

48 

Stenonema  sp. 

48 

Cinygmula  sp. 

21 

Rhithrogena  sp. 

21 

Epeorus  sp. 

21 

Anepeorus  sp. 

48 

Family  Leptophlebiidae 

36 

Paraleptophlebia  sp. 

24 

Leptophlebia  sp. 

24 

Choroterpes  sp. 

36 

Traverella  sp. 

36 

Family  Tricorythidae 

108 

Tricorythodes  sp. 

108 

Leptohyphes  sp. 

72 

Family  Ephemerellidae 

48 

Ephemerella  sp. 

48 

E.  grandis 

24 

E.  doddsi 

48 

E.  coloradensis 

18 

E.  tibialis 

24 

Taxa 

TQ 

Family  Ephemerellidae  (continued) 

E.  inermis 

48 

E.  infrequens 

48 

E.  spinifera 

24 

Family  Ephemeridae 

36 

Ephemera  simulans 

36 

Hexagenia  limbata 

36 

Phylum  Coelenterata 

108 

Class  Hydrozoa 

108 

Phylum  Aschelminthes 

108 

Class  Nematoda 

108 

Phylum  Mollusca 

108 

Class  Gastropoda 

108 

Family  Lymnaidae 

108 

Lymnaea  sp. 

108 

Family  Physidae 

108 

Physa  sp. 

108 

Family  Planorbidae 

108 

Phylum  Annelida 

108 

Class  Hirudinea 

108 

Class  Oligochaeta 

108 

Family  Tubificidae 

108 

Tubifex  sp. 

108 

Family  Lumbricidae 

108 

Lumbricus  aquaticus 

108 

Phylum  Platyhelminthes 

Class  Turbellaria 

108 

Order  Tricladida 

108 

Phylum  Arthropoda 

Class  Arachnida 

Suborder  Hydracarina 

108 

Class  Crustacea 

108 

Order  Isopoda 

108 

Family  Asellidae 

108 

Asellus  sp. 

108 

Order  Amphipoda 

108 

Family  Talitridae 

108 

Hyalella  azteca 

108 

Family  Gammaridae 

108 

Gammarus  lacustris 

108 

Order  Decapoda 

108 

Family  Astacidae 

108 

Pacifastacus  gambeli 

108 

Cambarus  laevis 

108 

Order  Cladocera 

108 

Daphnia  sp. 

108 

Order  Copepoda 

108 

Order  Ostracoda 

108 

Class  Insecta 

108 

Order  Collembola 

108 

Family  Poduridae 

108 

Podura  aquatica 

108 

Family  Entomobryidae 

108 

Macroinvertebrates 


671 


GASTROPODA 

Lymnaea 


AMPHIPODA 

Gammarus 


OLIGOCHAETA 

Lumbricus 


HYDRACARINA 

Mideopis 


LEPIDOPTERA 

Paragyractis 


TRICLADIDA 

Planaria 


MEGALOPTERA 

Sialis 


EPHEMEROPTERA 

Stenonema 


Figure  3.     Common  aquatic  macroinvertebrate  orders  used  in  ecosystem  analysis. 
672  Macroinvertebrates 


DIPTERA 

Psychoda 


TRICHOPTERA 

Helicopsyche 


PLECOPTERA 

Isoperla 


COLEOPTERA 

Gyrinus 


DIPTERA 

Simulium 


HEMIPTERA 

Lethocercus 


DIPTERA 

Chironomus 


Figure  3.     Common  aquatic  macroinvertebrate  orders  used  in  ecosystem  analysis  (concluded). 

Macroinvertebrates 


673 


The  procedure  for  determining  DAT  is  as 
follows:  the  dominance  part  of  the  index  is  de- 
termined by  placing  a  sample  of  aquatic  insects 
in  a  petri  dish  with  lines  marked  on  the  bottom. 
Each  species  of  insects  viewed  through  a  dis- 
secting microscope  can  be  identified  if  the  dish 
is  moved  in  a  snake-like  fashion  so  the  same 
organisms  are  not  viewed  twice.  Each  organism 
and  series  can  be  counted.  A  series  is  recorded 
each  time  a  sighted  species  is  different  from  the 
previous  one.  If  the  sample  has  a  sufficient  num- 
ber of  organisms,  300  to  400  should  be  ob- 
served to  determine  the  diversity  index  value. 
The  number  of  series  (s)  is  divided  by  the  num- 
ber of  organisms  (o)  to  determine  the  domi- 
nance value  (Dom.  =  s/o),  which  will  be  a 
fraction  between  0  and  1 .  The  number  of  taxa 
in  the  sample  is  multiplied  by  this  dominance 
value  to  obtain  the  DAT  Diversity  Index 
(Dom.  x  Taxa  =  DAT). 


DAT  Scale 

Stream  Value 

18-26 

Excellent 

11-17 

Good 

6-10 

Fair 

0-5 

Poor 

3.   Standing  Crop.  The  dry  weight  of  the  organisms 
is  determined  for  each  sample.  Small  aluminum 
pans  are  predried  for  8  to  24  hours  and  are 
weighed  to  the  nearest  tenth  of  a  milligram  on  a 
Mettler  balance.  The  samples  placed  in  these 
pans  are  dried  at  a  temperature  of  65°C  (149°F) 
for  8  to  24  hours;  then  the  pan  and  its  contents 
are  weighed.  This  is  important  in  indicating 
whether  a  stream  is  reaching  its  potential  and 
shows  the  stream's  potential  for  supporting  a 
resident  fishery.  Figure  3  shows  some  of  the 
more  common  aquatic  insects  used  in  aquatic 
ecosystem  analysis. 

Analysis  Elements  (Minor).  The  following 
summary  analyses  are  helpful  in  evaluating  aquatic 
ecosystems: 

1.  Total  Number  of  Species.  Total  number  of  spe- 
cies is  an  indication  of  macroinvertebrate  com- 
munity diversity.  It  often  gives  an  initial 
indication  of  the  stability  of  the  environment  or 
of  the  community  within  the  environment. 

2.  Mean  Number  of  Organisms  Per  Taxa  Per 
Square  Meter.  This  value  is  an  indication  of  the 
stability  of  the  community,  habitat,  and  water 
quality.  Taxa  may  be  transient  individuals  or 
resident  populations  that  may  indicate  whether 
or  not  conditions  are  favorable. 


3.  Standard  Error  of  the  Mean.  This  is  a  measure 
of  similarity  in  samples  which  indicates  when 
enough  samples  have  been  taken  to  obtain  the 
desired  reliability.  According  to  the  U.S.  Envi- 
ronmental Protection  Agency  standards  (Weber 
1973;  Green  1979),  this  value  should  not  be 
under  20% ;  on  most  streams,  this  can  be 
achieved  with  three  samples.  Some  streams  re- 
quire four  or  five  samples.  This  value  shows 
when  the  mean  is  within  80%  confidence  limits. 

4.  Coefficient  of  Variation  Coefficient  of  variation 
is  a  value  that  indicates  variability  in  samples, 
independent  of  the  number  of  samples.  It  evalu- 
ates the  sampling  technique  and  effectiveness 
of  the  sample  equipment.  The  U.S.  Environmen- 
tal Protection  Agency  has  set  the  standard  at 
under  50%  for  acceptable  reliability. 

5.  Standard  Deviation.  Standard  deviation  meas- 
ures sample  variation  and  indicates  whether 
sampling  techniques  and  number  of  samples  are 
sufficient  to  effectively  show  community  struc- 
ture. 


DISCUSSION 

Macroinvertebrate  data,  when  combined  with 
data  on  aquatic  features  and  water  quality,  can  be 
useful  to  land  managers  in  managing  aquatic  re- 
sources and  surrounding  areas.  Positive  or  negative 
effects  from  almost  every  activity  that  occurs  in  a 
drainage  are  eventually  reflected  in  the  macroinver- 
tebrate community  composition  of  the  streams  that 
drain  it.  This  community  composition  can  reveal  the 
health  of  the  ecosystem,  provide  a  warning  system, 
or  reflect  the  relative  intensity  of  pollution  problems. 
Macroinvertebrate  analysis  often  can  identify  the 
type(s)  of  pollution,  making  it  a  valuable  tool  for 
conscientious  land  managers. 

Macroinvertebrate  data  are  site-specific  and  av- 
erage values  for  large  diverse  areas  should  not  be 
used  for  decisions.  Time  and  space  considerations 
are  very  important  in  maintaining  high  standards  of 
aquatic  habitat  and  water  quality,  particularly  in  for- 
mulating grazing  prescriptions.  Monitoring  plans 
for  such  prescriptions  need  to  incorporate  the  same 
considerations  of  time  and  space. 

Finally,  the  biologist  and  manager  should  be 
cautioned  against  relying  on  macroinvertebrate  data 
alone.  Macroinvertebrate  data  should  be  used  in 
conjunction  with  other  types  of  data  to  provide 
sound  information  on  the  health  of  aquatic-riparian 
ecosystems. 


674 


Macroinvertebrates 


LITERATURE  CITED  u.s.  department  of  agriculture,  forest  ser 

VICE.  1981.  Aquatic  habitat  surveys  handbook  FSH 

2609. 2  3  for  the  General  Aquatic  Wildlife  System.  U.S. 

Dep.  Agric,  For.  Serv.  Region  4.  Ogden,  UT.  257pp. 

(Draft). 
GREEN,  R.H.  1979.  Sampling  design  and  statistical  meth-  WEBER,  C.  1973-  Biological  field  and  laboratory  methods 

ods  for  environmental  biologists.  John  Wiley  &  Sons.  for  measuring  the  quality  of  surface  waters  and  ef- 

251pp.  fluents.  EPA.  670/4-73-001. 


Macroinvertebrates  675 


V  SPECIAL  STUDIES 


33  Radiotelemetry 

34  Food  Habits 

35  Weather  and  Climate 


33 

RADIO- 
TELEMETRY 

Paul  L.  Hegdal1 


U.S.  Fish  and  Wildlife  Service 
Denver  Wildlife  Research  Center 
Building  16,  Denver  Federal  Center 
Denver,  CO  80225 


Bruce  A.  Colvin 


Editor's  Note:  Within  the  past  25  years,  wildlife 
research  has  moved  from  subjective  field  observa- 
tions to  highly  complex,  objective  measurements  of 
animal  behavior.  The  objective  measurements,  in 
many  cases,  are  accomplished  by  use  of  radiotelem  - 
etry.  This  chapter  provides  an  overview  of  current 
state-of-the-art  uses  of  radiotelemetry  in  locating 
habitats  used  by  subject  animals. 

While  the  urge  to  use  new  "gadgets"  to  learn 
about  animals  is  understandable,  today's  biologists 
should  conform  to  conventional  responsibilities 
toward  experimental  design  and  project  planning. 
"Does  the  end  justify  the  means?"  This  chapter, 
therefore,  also  identifies  current  processes  and 
equipment  needed  to  plan  and  execute  a  telemetry 
program  useful  in  habitat  management,  but  recog- 
nizes the  field  is  continuously  expanding. 


Department  of  Biological  Sciences 
Bowling  Green  State  University 
Bowling  Green,  OH  43403 


INTRODUCTION 

The  development  and  implementation  of  radio- 
telemetry  in  wildlife  research  has  tremendously 
broadened  the  opportunity  to  examine  components 
of  species'  natural  history  and  ecology.  Since  its  in- 
troduction into  wildlife  research  approximately  25 
years  ago  (Marshall  and  Kupa  1963;  Mech  1983), 
radiotelemetry  use  has  vastly  increased,  especially  in 
recent  years  with  the  increasing  attractiveness  of 
"high-tech"  approaches  to  research,  their  usefulness 
and  availability. 

Today,  examples  of  radiotelemetry  research  are 
commonly  found  in  literature.  Radiotelemetry  hard- 
ware, techniques,  and  uses  are  continuously  being 
improved  and  evaluated.  The  number  and  diversity 
of  species  being  studied  with  radiotelemetry  also  are 
continuing  to  grow  and  include  mammals  (Marshall 
et  ai.  1962;  Mech  1977;  Madison  1978;  Barrett 
1984),  birds  (Nicholls  and  Warner  1972;  Cochran 
1975),  reptiles  (Carr  1965;  Schubauer  1981;  Osgood 
1970;  Kenward  et  al.  1982),  amphibians  (Jansen 
1982),  fish  (Winter  et  al.  1978),  and  even  crabs 
(Wolcott  1980)  and  crayfish  (Covich  1977).  Habitats 
where  these  studies  have  been  undertaken  range 
from  the  polar  circle  (Kolz  et  al.  1980)  to  temperate 
regions  (Verts  1963;  Imboden  1975),  tropical  re- 
gions (Bruggers  et  al.  1983),  and  oceans  (Garshelis 
and  Garshelis  1984). 

There  are  several  possible  reasons  why  radiote- 
lemetry may  be  implemented  as  part  of  a  wildlife 
research  or  management  scheme.  Accessibility  of  the 
species  often  is  a  principal  reason.  Physical  charac- 
teristics of  a  species'  habitat,  such  as  rough  terrain  or 
dense  vegetation,  may  limit  the  opportunity  to  seek 

1  Current  affiliation:  Denver  Wildlife  Research  Center,  APHIS/ 
USDA  Building  16,  Denver  Federal  Center,  Denver,  CO  80225. 


Radiotelemetry 


679 


and  observe  a  species.  Also,  a  species  may  be  noctur- 
nal, highly  secretive,  difficult  to  trap  repeatedly, 
capable  of  wide-ranging  movements,  or  even  subter- 
ranean or  aquatic  in  its  habits. 

A  second  advantage  of  radiotelemetry  is  that 
data  can  take  a  continuous  form  rather  than  the  dis- 
crete form  which  occurs  for  example,  through 
trapping  and  marking.  Once  an  operating  radio  trans- 
mitter is  attached  to  an  animal,  that  particular  animal 
can  potentially  be  located  continuously  either  day 
or  night.  Radiotelemetry  also  provides  the  opportu- 
nity to  remotely  follow  or  census  wildlife.  Once 
instrumented,  the  specific  animal  can  be  identified 
and  observed  in  a  non-disruptive  manner  and,  thus,  a 
more  accurate  depiction  of  the  species'  movements, 
habitat  use,  and  ecology  may  be  acquired. 

In  contrast,  when  traps  are  used  to  locate  an 
animal,  one  must  assume  that  a  trap  is  available  to 
capture  the  animal  at  a  particular  location  and  time 
and  that  the  animal  will  enter  the  trap.  Traps  (or 
observers)  are  typically  placed  at  preconceived  (sub- 
jective) locations  where  animals  might  range,  and 
thus  movement  data  generated  from  trapping  (or  di- 
rect observation)  may  be  strongly  biased.  Addition- 
ally, trapping  often  takes  on  a  day  or  night  censusing 
form,  depending  on  the  activity  periods  characteris- 
tic of  the  species,  thus  adding  to  the  discrete  and 
limited  nature  of  observations  that  can  be  made. 
Continuous  observations  also  can  be  made  by  direct 
observation;  however,  without  radiotelemetry,  locat- 
ing target  animals,  distinguishing  individuals,  and 
monitoring  the  activities  of  many  isolated  individuals 
may  be  far  more  difficult. 

The  purpose  of  this  chapter  is  to  discuss  radio- 
telemetry techniques;  types  of  available  equipment; 
and  applications  for  determining  movement,  migra- 
tion, and  habitat  use  by  wildlife.  In  addition,  we  de- 
scribe potential  problems,  limitations,  and  costs. 


RADIO-TRACKING  EQUIPMENT 

Radio  tracking  should  be  considered  as  no  more 
than  a  technique  for  extending  the  range  of  one's 
observational  powers.  Even  the  simplest  equipment 
requires  a  significant  financial  investment.  A  simple 
system  for  tracking  about  10  animals  with  hand- 
carried  receivers  and  antennas  will  cost  $3,000  to 
S4,000  (1986  values). 

Many  improvements  have  been  made  since  the 
initial  successful  studies  in  the  early  1960s,  such 
as  more  efficient  transmitters;  better  encapsulating 
materials;  lighter,  more  energy-efficient  batteries; 
solar  units;  and  more  efficient,  easy-to-use  receivers. 
In  addition,  all  components  and  complete  systems 
now  are  commercially  available  from  several  sources 
(Table  1). 


Table  1.     Possible  sources  of  supply  for  radio- 
tracking  equipment. 


Advanced  Telemetry  Systems,  Inc. 
23859  Northeast  Highway  65 
Bethel,  MN  55005 
(612)434-5040 

AVM  Instrument  Company 
6575  Trinity  Court 
Dublin,  CA  94566 
(415)829-5030 

Cedar  Creek  Bioelectronics  Laboratory 
University  of  Minnesota 
Bethel,  MN  55005 
(612)  434-7361 

CompuCap 

8437  Yates  Avenue  North 

Brooklyn  Park,  MN  55443 

(612)424-2373 

Custom  Electronics 
2009  Silver  Court  West 
Urbana,  IL  61801 
(217)344-3460 

Custom  Telemetry  and  Consulting 
185  Longview  Drive 
Athens,  GA  30605 
(404)548-1024 

L.  L.  Electronics 
Box  247 

Mahomet,  IL  53405 
(217)  586-2132 

Ocean  Applied  Research  Corp. 
10447  Roselle  Street 
San  Diego,  CA  92121 
(619)453-4013 

Smith-Root,  Inc. 

14014  Northeast  Salmon  Creek  Avenue 

Vancouver,  WA  98665 

(206)  573-0202 

Stuart  Enterprises 

Box  310,  124  Cornish  Court 

Grass  Valley,  CA  95945 

(916)273-9188 

Telemetry  Systems,  Inc. 
Box  187 

Mequon,  Wl  53092 
(414)  241-8335 

Telonics 

932  East  Impala  Avenue 
Mesa,  AZ  85204-6699 
(602)  892-4444 


680 


Radiotelemetry 


Table  1.     Possible  sources  of  supply  for  radio- 
tracking  equipment  (concluded). 


Wildlife  Materials,  Inc. 
R.R.  1,  Grant  City  Road 
Carbondale,  IL  62901 
(618)  549-6330 


'Use  of  companies  on  this  list  does  not  imply  Federal 
Government  endorsement. 


There  are  basically  two  components  in  a  radio- 
tracking  system:  a  transmitting  system  and  a  receiv- 
ing system.  The  transmitting  system  consists  of  the 
transmitter  which  is  attached  to  an  animal  by  an 
appropriate  method  such  as  a  radio  collar  on  big 
game  animals,  patagial  transmitters  on  birds,  and  im- 
planted transmitters  in  fish  or  snakes.  The  receiving 
system  consists  of  a  receiver,  a  receiving  antenna, 
and  either  an  operator  or  recorder. 


Patagial  transmitter  attached  to  a  hen  pheasant. 


Transmitters 

The  specific  type  of  transmitter  chosen  will 
depend  on  size,  morphology,  and  behavior  of  the 
animal  under  study;  possible  attachment  methods; 
transmitter  availability  and  cost;  necessary  transmis- 
sion range;  habitat  where  used;  and  the  particular 
data  to  be  collected. 

Almost  any  animal  that  weighs  over  15  g  (0.5 
oz)  can  be  equipped  with  a  radio  transmitter  and 
monitored  for  at  least  a  short  time.  Obviously,  the 
larger  the  animal,  the  larger  the  transmitter  package 
can  be.  Transmitters  are  available  commercially  rang- 
ing from  slightly  over  a  gram  (0.03  oz)  to  several 
kilograms.  The  actual  transmitter  per  se  does  not 


vary  much  in  weight  but  the  power  source,  packag- 
ing material,  and  attachment  material  can  add  sub- 
stantially to  the  weight.  Weight  considerations  for 
the  animal  usually  become  important  only  for  the 
smaller  (<lkg[<2.2  1b])  species.  This  can  be  the 
most  critical  consideration  with  very  small  animals 
(<  20  g  [<  0.7  oz])  and  especially  birds.  Generally, 
the  transmitter  should  be  no  more  than  5%  of  body 
weight.  The  Banding  Office  of  the  U.S.  Fish  and  Wild- 
life Service  recommends  no  more  than  3%  of  body 
weight  for  transmitters  used  on  birds.  Cochran 
(1980)  stated  that  many  species  seem  to  tolerate  a 
package  that  is  4%  of  their  body  weight  and  appear 
to  behave  normally  not  too  long  after  such  a  package 
has  been  attached.  He  also  added  that  there  is  noth- 
ing "magic"  about  4% .  However,  there  may  be  some 
species  that  cannot  be  radio-equipped  with  these 
transmitter-weights,  or  even  lighter  transmitters, 
without  having  significant  behavioral  or  physical  ef- 
fects as  a  result  of  instrumentation. 

Ordinarily,  each  transmitter  used  in  a  particular 
study  area  is  on  a  unique  frequency  (or  channel); 
therefore,  to  tune  in  each  animal,  the  operator 
merely  turns  a  dial  or  activates  switches  on  a  radio 
receiver  to  the  appropriate  frequency  (or  channel). 

Radio  frequency  management  in  the  U.S.  is 
based  on  the  Communications  Act  of  1934.  This 
legislation  divided  frequency  spectrum  users  into 
two  groups  (federal  and  non-federal).  The  National 
Telecommunications  and  Information  Administration 
(NTIA)  controls  the  federal  portions  of  the  spec- 
trum, while  the  Federal  Communication  Commission 
(FCC)  controls  the  non-federal  portions  of  the  spec- 
trum for  use  by  state  and  local  governments  as  well 
as  the  private  sector.  Since  segments  of  the  spectrum 
are  shared  by  federal  and  non-federal  users,  opera- 
tions are  mutually  coordinated  by  the  Interdepart- 
mental Radio  Advisory  Committee  (IRAC).  Each 
Federal  Government  agency  and  the  FCC  are  repre- 
sented on  the  IRAC  to  address  spectrum  manage- 
ment issues.  The  IRAC,  through  NTIA,  has  authority 
to  approve,  disapprove,  and  cancel  any  federal  radio- 
frequency  assignment.  In  effect,  IRAC  is  the  licensing 
authority  for  federal  users  while  the  FCC  licenses 
non-federal  users. 

Two  portions  of  the  spectrum  have  been  set 
aside  for  wildlife  telemetry  use  (40.16  to  40.20  MHz 
and  216  to  220  MHz);  however,  these  frequencies 
have  not  been  widely  used  by  wildlife  researchers  or 
managers.  Radio-frequency  spectrum  managers  are 
becoming  increasingly  concerned  about  unlicensed 
operations  and  the  violations  of  rules  and  regulations 
pertaining  to  radio  communications.  By  law,  teleme- 
try users  must  operate  within  the  rules  and  regula- 
tions because  any  interference  occurring  between 
authorized  and  unauthorized  uses  will  cause  unau- 
thorized use  to  shut  down,  resulting  in  costly  losses 
of  time  and  effort  expended  on  studies.  One  must 


Radiotelemetry 


681 


have  authorization  to  use  a  particular  frequency  be- 
fore any  transmitters  can  be  placed  on  free-ranging 
animals.  For  further  guidance  and  information,  see 
Kolz  (1983). 

Most  transmitters  use  mercury  or  lithium  batter- 
ies to  power  the  transmitter;  however,  solar-powered 
units,  some  with  rechargeable  batteries,  also  are 
available  (Cochran  1980).  Battery  choice  must  be 
tempered  by  the  amount  of  additional  weight  the  an- 
imal can  carry  and  the  length  of  time  the  animal 
needs  to  be  tracked.  If  the  transmitter  is  retrievable 
or  the  animal  recapturable,  transmitters  can  be 
reused  and  their  longevity  extended  by  changing 
batteries. 

Most  simply,  activating  a  transmitter  includes 
soldering  a  final  connection  and  then  potting  (cover- 
ing) that  connection  with  an  appropriate  quick-set- 
ting potting  material  (i.e.,  dental  acrylic,  epoxy 
patch,  or  silicone  rubber);  others  are  activated  by 
cutting  a  wire  that  allows  completion  of  a  circuit. 
Additionally,  with  some  transmitters,  placing  or  re- 
moving a  magnet  near  an  imbedded  switch  turns  the 
transmitter  on  or  off.  A  magnet  system  is  preferred 
as  it  does  not  require  any  soldering  or  potting  by  the 
user  and  is  more  likely  to  be  completely  sealed  from 
moisture.  Normally  transmitters  are  activated  only 
shortly  before  attachment  to  an  animal.  However, 
some  users  routinely  turn  transmitters  on  for  a  few 
days  prior  to  attachment  to  assure  proper  operations. 
The  loss  of  a  few  days  of  battery  life  may  be  critical 
only  for  small,  short-lived  transmitters. 

Most  transmitter  antennas  are  some  sort  of 
tuned  whip.  The  most  efficient  antenna  lengths  are 
in  quarters  of  wave  lengths  of  a  particular  frequency 
(such  as  lA,  V-z,  or  whole  wave).  Except  for  the  high 
frequencies,  most  wildlife  transmitter  antennas  can- 
not be  that  long  (a  full  wave  length  at  164  MHz 
=   1.82  m)  and,  therefore,  are  tuned  with  a  coil  to 
whatever  length  can  be  tolerated  on  the  animal. 
Whip  antennas  are  more  efficient  than  loop  antennas. 
However,  in  some  applications,  the  loop  antenna 
serves  as  an  integral  part  of  the  attachment  collar. 

There  are  a  variety  of  methods  for  attaching 
transmitters  to  animals,  and  before  any  transmitters 
are  attached  to  any  species,  one  should  review  avail- 
able literature  on  the  species,  try  the  attachments, 
and  act  on  the  results  of  those  endeavors.  Neck  col- 
lars are  the  most  common  radio-attachment  method 
for  terrestrial  mammals  (Cochran  1980).  Harness, 
ear  tag,  and  implanted  transmitters  have  been  used 
on  some  mammals  that  are  difficult  to  collar  (Mech 
1983).  For  birds,  harness,  tail-clip,  poncho,  glued-on, 
leg-band,  and  patagial  transmitters  have  been  suc- 
cessfully employed  (Bray  and  Corner  1972;  Fitzner 
and  Fitzner  1977;  Amstrup  1980;  Bruggers  et  al. 
1981 ).  Implantable  or  ingestible  transmitters  are  use- 
ful where  external  attachment  is  impractical,  such 


as  with  fish  or  snakes  (Osgood  1970).  Tethers  have 
proven  useful  for  sea  turtles  and  manatees  (Timko 
and  Kolz  1982). 

A  number  of  factors  need  to  be  considered 
when  designing  or  specifying  the  type  of  collar,  har- 
ness, or  other  attachment  to  be  used.  Shape,  width, 
contouring,  durability,  flexibility,  smoothness  next  to 
the  animal,  size  adjustability,  compactness,  cryptic 
design,  internal  or  external  antenna,  and  ease  of  at- 
tachment all  can  be  important  (Cochran  1980). 
Sharp  edges  and  points  should  be  avoided  or  at  least 
placed  where  there  will  be  minimal  irritation  from 
contact  or  pressure.  Without  at  least  considering 
these  factors,  routine  behaviors  such  as  running, 
flying,  feeding,  mating,  or  even  resting  may  be  ad- 
versely affected.  If  necessary,  one  may  have  to  test 
various  attachments  on  some  species  where  methods 


Radio-collared  opossum. 


-% 


"•»* 


% 


Radio-collared  coyote. 


682 


Radiotelemetry 


and  designs  have  not  been  evaluated  by  others.  Ide- 
ally, the  transmitter  will  be  on  the  animal  only  as 
long  as  needed  to  collect  the  desired  data.  Some 
transmitters  are  designed  to  fall  off  after  some  prede- 
termined time;  however,  there  have  been  some 
problems  with  premature  transmitter  loss.  With  a 
well-designed  attachment,  many  animals  have  not 
had  significant  problems  carrying  the  transmitter  for 
life.  We  have  had  transmitters  on  owls  for  over  5 
years  with  no  apparent  problems. 

Transmitters  for  remote  locations  generally  cost 
between  $100  and  $250  each.  A  variety  of  special- 
purpose  transmitters  have  been  developed  for  de- 
tecting movement  or  mortality  (Kolz  1975);  measur- 
ing heart  rate,  respiration  rate,  temperature,  and 
blood  pressure,  as  well  as  conducting  electrocardi- 
ograms and  encephalograms  (Amlaner  1978);  firing 
darts  on  command  for  recapture  (Mech  et  al.  1984); 
and  some  specific  behaviors  such  as  frequency  of 
urination.  However,  these  transmitters  and  receivers 
are  more  expensive  than  transmitters  for  remote 
locations. 


Receivers 

A  variety  of  receivers  are  available,  ranging  from 
$700  for  the  simpler  receivers  to  about  $2,000  for 
some  of  the  newer,  wider  band  models  with  pro- 
grammable automatic  switching  functions,  often  re- 
ferred to  as  "scanners."  However,  paying  more 
money  for  a  receiver  will  not  guarantee  a  more  sen- 
sitive receiver.  Most  can  be  tuned  to  equal  "state 
of  the  art"  sensitivity  (  ±  150  db) — even  the  $700 
versions. 


When  planning  purchases  of  telemetry  equip- 
ment, at  least  one  extra  receiver  should  be  acquired 
for  the  project  as  a  backup.  There  is  nothing  more 
frustrating  than  having  radio-equipped  animals  and 
no  working  receivers  to  track  them.  Extra  batteries 
should  always  be  available  for  receivers  and  any 
other  equipment  used  in  conjunction  with  them. 

Most  receivers  have  a — 


( 1 )  power  switch,  often  with  a  position  for  inter- 
nal or  external  power; 

(2)  channel  selector  (for  digital  frequency  display 
models,  this  may  be  dials  or  switches  for 
changing  the  displayed  frequency); 

(3)  fine-tuning  frequency  dial  (on  non-digital 
models); 


(4)    gain  (volume)  dial; 


(5)  sweep  switch  that  allows  automatic  sweep 
within  a  channel  or  between  preset  exact 
frequencies;  and 

(6)  switches  for  setting  frequencies  to  be 
searched. 


Most  receivers  also  have  a  variety  of  input  and 
output  plugs  for  antennas,  headphones,  external 
power,  meters,  pulse-interval  counters,  recorders, 
and  other  attachments. 

Earphones  are  not  necessary  in  many  tracking 
situations.  The  inexpensive  models  (  $20  or  less)  are 
adequate  in  most  cases  for  hand  or  vehicle  tracking. 
However,  for  tracking  in  aircraft  in  windy  or  noisy 
conditions,  earphones  are  necessary,  and  it  is  worth 
the  extra  money  to  get  the  best — $100  to  $200 
each.  (Long-continued  use  of  earphones  at  high  au- 
dio levels  has  resulted  in  diminished  hearing  abilities 
among  some  research  biologists.) 


Antennas 

Almost  any  piece  of  metal  connected  to  the 
antenna  plug  of  a  receiver  will  enhance  the  signal, 
but  properly  tuned  and  tested  antennas  should  be 
used  for  maximum  efficiency.  The  yagi  antenna  is  the 
most  commonly  used  antenna  for  radiotelemetry 
studies.  A  good,  single  hand-held  yagi  antenna  costs 
$50  to  $100.  As  frequency  increases,  wave  length 
decreases  and  the  size  of  antennas  also  decreases. 
Yagi  antennas  are  too  large  to  be  very  practical  for 
hand-held  or  vehicle-mounting  at  the  lower  (30  to 
50  MHz)  frequencies  but  can  readily  be  used  on 
vehicles  or  hand-held  at  the  higher  (over  150  MHz) 
frequencies. 

Antennas  can  be  mounted  on  towers,  vehicles, 
or  aircraft.  Two  (or  more)  yagi  antennas  can  be 
combined  in  such  a  manner  (called  stacking)  to  in- 
crease their  range  and  directional  ability.  In  dual 
array  systems,  antennas  can  provide  a  bearing  accu- 
racy of  1  to  2°.  Often,  permanent  tracking  stations 
consisting  of  a  few  strategically  placed  antenna  tow- 
ers or  masts  mounted  on  the  highest  hills  in  the 
study  area  will  yield  adequate  signals.  These  towers 
can  be  similar  to  the  large,  automatically  rotating 
system  described  by  Nicholls  and  Warner  (1972)  or 
relatively  simple  structures  with  a  coaxial  cable  left 
dangling  near  the  base  (Merson  et  al.  1982). 

Permanent  or  temporary  tower  antennas  can 
work  well  for  sedentary  animals;  however,  the  accu- 
racy of  radio  locations  deteriorates  near  the  base  line 
of  the  antennas  (because  the  angle  of  intercept  nears 
zero).  If  additional  towers  or  mobile  tracking  sta- 
tions can  be  employed,  this  problem  can  be 
overcome. 


Radiotelemetry 


683 


Radio-tracking  vehicles  (mobile  tracking  sta- 
tions) can  be  equipped  with  roof-mounted,  dual  yagi 
antenna  systems.  Bray  et  al.  (1975)  described  a  re- 
movable cartop  antenna  system.  However,  for  most 
studies  using  mobile  tracking  stations,  we  strongly 
recommend  vehicles  equipped  with  a  through-the- 
roof  antenna  system  similar  to  that  described  by 
Hegdal  and  Gatz  (1978).  These  antennas  are  rotated 
from  inside  the  vehicles,  and  radio  bearings  are  indi- 
cated on  a  360°,  25-cm  ( 10-in.)  protractor  by  a 
pointer  attached  to  the  antenna  mast.  The  pointer  is 
aligned  to  the  null  of  the  dual  yagi  antennas.  Coaxial 
cables  from  the  dual  yagis  are  attached  to  a  null- 


peak  switchbox  (available  from  commercial  sup- 
pliers), which  allows  switching  from  in-phase  (for 
maximum  signal  strength)  to  out-of-phase  (for  pre- 
cise bearings)  operation.  Additionally,  radio  commu- 
nication, a  plotting  table,  auxiliary  batteries,  and 
extra  lighting  installed  in  the  vehicles  will  expedite 
tracking  studies.  Figure  1  shows  an  inexpensive 
method  for  equipping  vehicles  with  a  roof-mounted, 
dual-beam  antenna  system.  If  properly  balanced, 
antennas  can  be  rotated  easily  and  animals  can  be 
tracked  while  the  vehicle  is  moving  as  fast  as  55 
mph. 


o  o 

o  o 


Jk 





An 


tJi"muiiT  . 


I 


f 

9 


It 


■tt    ? 


j  ,,,  [4 


-t- 


tr,    i. ■-■■V-  i  v.- 


I, I. inn 


I 


Figure  1.     A  roof-mounted,  dual  yagi,  radio-tracking  system. 
684  Radiotelemetry 


On  aircraft,  antennas  are  usually  mounted  on 
wing  struts  (with  certified  antenna  mounts)  in  a 
"side-looking"  fashion,  perpendicular  to  the  fuselage 
(Gilmer  et  al.  1981).  Care  also  must  be  taken  to 
ensure  that  coaxial  cables  are  properly  secured  with 
tape,  not  crimped,  and  that  they  and  the  rest  of  the 
telemetry  equipment  do  not  interfere  with  the  pilot's 
operation  of  the  aircraft. 

Too  often,  new  (and  experienced)  radio-track- 
ers do  not  periodically  test  and  check  their  antenna 
system.  They  may  have  an  extremely  sensitive  re- 
ceiver matched  to  a  poorly  tuned  antenna  and  there- 
fore have  (unknowingly)  a  very  inefficient  receiving 
system. 

Coaxial  cables  are  used  to  connect  antennas  or 
other  equipment,  such  as  recorders,  to  receivers. 
This  is  a  shielded  cable  (a  center  wire  conductor 
surrounded  by  the  ground  wire ).  Care  must  be  taken 
to  avoid  sharp  bends,  twisting,  and  flattening  or 
pinching;  distortion  of  the  cable  shape  can  drastically 
affect  or  eliminate  signal  reception. 


'^Rc~^^  ^ 

A  radio-tracking  vehicle  at  a  marked  tracking  station. 


Other  Equipment 

Pulse  interval  counters  are  available  for  about 
$400  and  can  be  used  in  conjunction  with  several 
receivers.  These  devices  measure  the  time  between 
consecutive,  pulsed  radio  signals  and  usually  display 
that  value  in  thousandths  of  a  second.  They  are  espe- 
cially helpful  for  identifying  pulse  rates  (animals) 
when  there  is  more  than  one  transmitter  operating 
on  the  same  frequency  (or  channel).  Recorders  can 
be  used  in  conjunction  with  automatic  switching 
(scanning)  receivers  to  record  presence  or  absence 
of  radio-equipped  animals  over  time  at  sites  such 
as  dens  or  nests  (Gilmer  et  al.  1971;  Harrington  and 
Mech  1982;  Mech  1983).  However,  as  with  auto- 
matic tracking  systems,  one  must  have  a  good  strong 
signal  for  most  of  the  recorders  and  counters  to 
work  properly.  The  human  ear  can  detect  signals 
from  the  receiver  better  than  any  of  the  available 
auxiliary  equipment. 

Satellite  radiotelemetry  also  has  been  used  suc- 
cessfully on  a  few  wildlife  species  (Kuechle  et  al. 
1979;  Timko  and  Kolz  1982).  However,  transmitters 
and  receiving  equipment  are  more  specialized  and 
considerably  more  expensive  than  those  we  have 
described,  and  it  is  beyond  the  scope  of  this  chapter 
to  fully  discuss  these  aspects  of  radiotelemetry. 

Advice 

We  strongly  recommend  that  the  telemetry  user 
depend  on  electronic  specialists  for  constructing 
and  servicing  transmitters,  receivers,  and  other  spe- 
cialized equipment.  However,  the  biologist  must 
have  a  general  knowledge  of  radio  signal  propaga- 
tion; factors  that  affect  it;  and  some  skills  and  equip- 
ment for  servicing,  checking,  and  repairing  parts  of 
the  telemetry  system.  Users  should  be  able  to  charge 
power  sources  for  receivers  and  other  equipment; 
change  batteries  in  some  transmitters  (some  must  be 
returned  to  the  manufacturer);  replace  coaxial  cables 
and  antenna  elements;  turn  transmitters  on  and  off; 
check  power  sources,  telemetry,  and  antenna  sys- 
tems for  shorts  and  open  circuits;  and  recognize  that 
such  problems  exist.  Minimal  support  equipment 
should  include  a  volt-ohmmeter;  battery  tester;  sup- 
ply of  miscellaneous  wire  connectors  and  terminals; 
soldering  gun;  solder;  wire  strippers;  and  various- 
sized  screwdrivers,  pliers,  and  adjustable  wrenches. 


DATA  SAMPLING 

The  planning  stages  of  a  radiotelemetry  study 
must  include  decisions  on  specific  use  or  need  of  the 
technology,  equipment,  personnel  availability,  and 
cost.  Just  as  important,  however,  are  decisions  on 
how  data  will  be  collected,  analyzed,  and  inter- 
preted. The  recording  of  a  single  data  point  in  the 
field  can  be  a  simple  procedure;  however,  deciding 


Radiotelemetry 


685 


the  value  or  use  of  hundreds  or  thousands  of  coordi- 
nates may  not  be  and  should  not  be  left  until  the 
field  work  is  complete. 

Locating  an  animal  either  on  a  monthly,  weekly, 
daily,  hourly,  or  minute-by-minute  basis  all  may  be 
possible  depending  on  the  available  hardware,  per- 
sonnel, and  funding.  However,  while  a  large  number 
of  radio-equipped  animals  can  be  located  continu- 
ously, rarely  is  that  done.  The  specific  form  and  fre- 
quency of  sampling  should  be  particularly  relevant 
to  the  specific  questions  that  are  being  addressed 
with  radiotelemetry  use.  It  may  be  appropriate  to 
simply  monitor  general  movements  or  continued 
survivorship;  thus,  the  sampling  scheme  may  be  ex- 
tensive or  random  and  the  radio-tracking  results  may 
be  presented  in  a  rather  descriptive  form.  In  con- 
trast, when  evaluating  habitat  selection  or  rates  of 
movement,  for  example,  the  sampling  scheme  may 
have  a  highly  intensive  or  systematic  style  with  data 
recorded  in  more  of  a  quantitative  manner  (e.g., 
effort-  or  time-specific).  In  the  latter  case,  animals 
may  be  located  at  specific  and  independent  time 
intervals  or  continuously  tracked  within  discrete  and 
independent  time  periods. 

Sampling  scheme  decisions  also  may  be  affected 
by  the  species  being  studied.  Animals  that  are  capa- 
ble of  wide-range  movement  or  frequent  habitat 
change  may  have  to  be  located  more  often  to  avoid 
loss  of  contact  than  those  that  are  largely  sedentary. 
The  length  and  timing  of  daily  activity  cycles  of  ani- 
mals, for  example,  nocturnal  versus  diurnal,  also 
will  affect  the  length  and  timing  of  radio-tracking 
efforts;  the  sampling  scheme  for  any  given  individual 
may  have  to  be  compromised  considerably  depend- 
ing on  how  many  other  animals  are  simultaneously 
radio-equipped.  Additionally,  more  frequent  sampling 
may  be  needed  if  the  hardware  being  used  includes 
short  transmission-reception  range  or  a  limited  bat- 
tery or  radio  attachment  life. 

A  sampling  scheme  to  evaluate  the  importance 
of  specific  habitat  types  might  be  based  on  the  total 
time  an  animal  spends  in  each  habitat,  the  number  of 
locations  that  are  randomly  or  systematically  re- 
corded in  each  habitat  type  or,  in  contrast,  the  num- 
ber of  times  an  animal  enters  each  habitat  from 
alternative  habitats.  The  time  that  an  animal  spends 
in  a  particular  habitat  may  not  be  equated  necessar- 
ily to  the  value  of  that  habitat  to  the  species;  fre- 
quent, but  short  visits  to  other  habitat  types  may 
indicate  a  greater  resource  value.  In  addition,  when 
sampling  for  home  ranges,  Smith  et  al.  ( 1981 )  stated 
that  it  is  important  to  define  a  large  portion  of  the 
area  over  a  relatively  short  observation  period.  This 
is  because  home  ranges  are  really  time-specific, 
meaning  they  potentially  can  change  with  season, 
habitat,  or  the  animal's  reproductive  state.  Thus, 
intensive  sampling  efforts  in  a  specific  time  period 


may  be  necessary  when  accurately  defining  a  home 
range. 

One  additional  concern  related  to  sampling  is 
movements  recorded  immediately  following  animal 
instrumentation  and  release.  Such  movements  should 
be  considered  biased  (rather  than  exemplary)  be- 
cause of  the  disruption  created  by  trapping  and  han- 
dling. Therefore,  although  tracking  an  animal 
immediately  following  release  (e.g.,  1  day)  may  be 
appropriate,  there  should  be  a  degree  of  indepen- 
dence used  in  the  initial  observation  of  habitat  or 
home  range  analyses. 


FIELD-TRACKING 

Field-tracking  first  dictates  that  one  must  be- 
come an  expert  in  capturing  live,  uninjured  animals 
of  the  desired  species,  which  can  become  a  major 
effort  in  itself  for  some  species.  Secondly,  we 
strongly  recommend  that  those  new  to  radio-track- 
ing and  those  not  using  it  for  2  to  3  years  visit  and 
obtain  training  from  a  state-of-the-art  ongoing  proj- 
ect. Novice  radio-trackers,  especially,  will  encounter 
many  pitfalls  and  these  may  be  avoided  by  such  help. 
Also,  there  are  new  developments  annually  that  may 
be  useful  in  any  radiotelemetry  project. 

Triangulation  is  the  basic  principle  in  most  ra- 
dio-tracking. It  can  be  accomplished  by  taking  bear- 
ings from  two  locations;  the  animal  is  assumed  to  be 
near  the  point  where  the  bearings  cross  (the  loca- 
tion or  "fix").  Ideally,  bearings  should  cross  at  about 
a  90°  angle  and  be  taken  simultaneously  by  two  ra- 
dio communication  observers  and  as  close  to  the 
animal  as  possible.  As  the  angle  of  interception  be- 
comes more  acute  and  the  distance  from  the  animal 
to  the  observers  increases,  the  locations  become  less 
accurate  and  the  error  polygon  increases.  For  a  more 
thorough  discussion  of  error  polygon  and  means  of 
estimating  radio-location  errors,  see  Mech  (1983), 
Lee  et  al.  (1985),  and  Saltz  and  Alkon  (1985).  Con- 
siderable experience  in  the  field  is  necessary  to  be- 
come skilled  at  rapidly  selecting  tracking  locations 
for  accurate  bearings. 

When  learning  to  radio-track,  one  should  first 
practice  locating  an  activated  transmitter  hidden  by 
someone  else.  Repetition  of  this  "game"  can  most 
readily  acquaint  anyone  with  radio-tracking  tech- 
niques and  the  sensitivity  and  capability  of  the 
equipment. 

Hand  Tracking 

In  its  simplest  form,  field  radio-tracking  is  done 
by  carrying  the  receiver  and  hand-held  antenna  and 
"homing  in"  (walking  out)  on  the  radio-equipped 
animal.  Proper  "homing"  procedures  involve  deter- 
mining the  approximate  direction  of  the  transmitter, 


686 


Radiotelemetry 


tuning  in  the  signal,  reducing  gain  to  minimal  level, 
again  reducing  gain  to  minimal  level,  and  moving 
in  the  direction  of  the  signal.  While  moving  toward 
the  signal,  the  antenna  should  be  waved  in  an  ap- 
proximate 120°  arc,  back  and  forth,  continuing  to  re- 
duce gain  to  a  minimal  level.  When  gain  cannot  be 
reduced  further,  it  indicates  the  transmitter  is 
nearby.  If  the  animal  has  not  been  observed  or  the 
transmitter  found,  it  can  be  located  by  removing  the 
antenna,  turning  the  gain  up  so  it  barely  can  be 
heard  and,  depending  on  signal  strength,  determining 
the  location  by  moving  the  receiver  toward  the 
strongest  signal.  A  small  "locator"  loop  antenna  also 
is  useful  for  radio-tracking  at  close  range.  These  loca- 
tor loops  are  especially  useful  when  digging  trans- 
mitters from  underground  burrows. 

"Homing"  procedures  have  some  obvious  draw- 
backs. For  example,  the  animal  may  not  be  observed 
by  the  biologist  before  it  has  been  "pushed"  or 
flushed  from  roosting,  foraging,  or  loafing  areas,  and 
the  location  recorded  may  not  reflect  the  habitat 
utilized  prior  to  the  investigator's  disturbance.  Care 
must  be  used  to  avoid  these  false  locations  and  artifi- 
cial or  stimulated  movements  caused  by  the  investi- 
gator. Some  animals  (such  as  big  game,  raptors,  or 
animals  that  are  nocturnal)  are  difficult  to  observe  or 
approach  without  disturbing  them.  This  tracking 
method  may  be  relatively  useless  for  determining 
movements  of  nocturnal  animals  during  activity  pe- 
riods, but  may  (with  care)  readily  be  employed  to 
determine  specific  daytime  roosting,  denning,  or 
bedding  areas  during  periods  of  inactivity.  When  ra- 
dio-tracking large  numbers  of  animals,  hand-held 
equipment  becomes  inefficient  as  it  may  take  several 
hours  to  locate  some  animals  by  "walking  them  out." 
This  is  especially  true  in  inaccessible  areas  or  in 
areas  of  dense  vegetation. 

With  some  small,  relatively  immobile  and  fosso- 
rial  species  (such  as  rodents,  small  birds,  and  rep- 
tiles), virtually  all  radio-tracking  may  be  done  with 
hand-held  equipment.  Hand-held  tracking  will  always 
be  necessary  to  locate  and  recover  dead  animals, 
animals  that  must  be  recaptured,  or  transmitters  that 
have  fallen  off  the  carrier  animal. 

Recovery  of  lost  transmitters  or  dead  animals 
can  be  difficult,  sometimes  requiring  considerable 
digging,  use  of  heavy  excavating  equipment,  or  ex- 
plosives. For  example,  black-tailed  prairie  dogs  were 
recovered  in  South  Dakota  at  depths  of  2.5  to  3  m 
(8.25  to  10  ft — deepest  4.3  m  [14  ft]).  Also,  one 
cannot  assume  that  radio-equipped  animals  always 
will  be  in  normal  habitat  or  locations  expected  for 
that  species.  Finding  an  animal  is  often  based  on 
what  a  species  is  "supposed  to  do"  rather  than  what 
it  may  actually  do.  Birds,  rodents,  or  most  any  spe- 
cies can  be  taken  by  predators  as  well  as  legally  or 
illegally  by  humans.  Animals  and  their  transmitters 
have  been  recovered  in  raptor  nests,  burrows,  under 


buildings,  in  car  trunks,  freezers  in  homes,  and  many 
other  unusual  locations  where  the  animal  being 
tracked  would  not  be  expected  to  venture. 


Biologist  tracking  using  a  hand-held  vagi. 


Vehicle  Tracking 

Mobile  tracking  stations  (vehicles)  allow  one  to 
reduce  the  antenna-to-animal  distance  and  take  ad- 
vantage of  topography  in  the  area  for  better  tracking 
points.  Vehicle  tracking  can  only  be  done  if  there  are 
reasonable  access  routes  within  the  study  area.  Mo- 
bile tracking  stations  (two  or  more),  equipped  with 
the  described  antenna  systems,  radio  communication 
equipment,  auxiliary  batteries,  extra  lighting,  and 
plotting  tables,  allow  investigators  to  quickly  search 
areas  and  follow  fast-moving  animals  such  as  migrat- 
ing birds  in  daylight  or  darkness.  During  activity,  it 
may  be  especially  important  to  obtain  simultaneous 
bearings.  Since  many  animals  can  move  rapidly,  false 
interception  points  may  be  recorded  if  there  is  a 
time  lag  between  bearings.  Most  dual-beam  antenna 
systems  mounted  on  a  vehicle  require  a  minimum 
clearance  of  about  3  m  (  10  ft).  If  the  study  area 
contains  wooded  areas  with  low-hanging  branches,  it 
may  be  necessary  to  trim  several  kilometers  of  road 
for  adequate  clearance  of  the  tracking  vehicle.  In 
areas  where  this  is  not  practical,  a  single  yagi  should 
be  mounted  in  a  horizontal  plane.  This  will  lower 
clearance  requirements  and  still  give  fairly  reasona- 
ble bearings,  provided  the  distance  from  the  tracking 
vehicle  to  the  animal  is  short  when  bearings  are 
taken. 


Radiotelemetry 


687 


Turning  directional  yagi  antenna  from  inside  the  tracking 
van 


Recording  signals  on  plotter  system. 


Aerial  Tracking 

Usually,  aerial  tracking  starts  with  searching  the 
area  for  the  animal's  last  known  location,  increasing 
altitude  to  3,000  m  (9,900  ft),  and  searching  in  an 
enlarging  circle  or  flying  swaths  (20-  to  50-km  [12.4- 
to  3 1  -mi.  ]  wide )  for  complete  coverage  of  the  areas 
(Gilmer  et  al.  1981).  In  heavily  populated  areas  near 
cities,  there  may  be  too  much  radio  interference  to 
fly  at  higher  altitudes  (over  500  to  1,000  m  [310 
to  620  ft]).  Once  a  signal  is  detected,  aerial  tracking 
simply  can  become  "homing,"  similar  to  "walking 
out"  a  radio-equipped  animal  (reducing  gain  and 
lowering  altitude  as  the  source  of  the  signal). 

The  use  of  aircraft  is  especially  important  in 
large  or  inaccessible  areas  and  with  wide-ranging  or 


migrating  species.  Additionally,  aerial  searches  for 
missing  or  lost  animals  are  usually  much  more  effi- 
cient than  ground  searches.  The  additional  height 
provided  by  aircraft  greatly  increases  reception 
range.  For  example,  our  vehicle  tracking  system  has 
a  range  of  about  3  to  4  km  ( 1.8  to  2.5  mi.)  with  a 
7  g  (0.24  oz)  transmitter  on  a  bird,  while  with  air- 
craft, a  35-km  (21.7-mi.)  receiving  range  can  be 
achieved. 

The  first  step  in  aerial  tracking  is  consulting 
with  a  certified  aircraft  mechanic  to  make  sure  that 
the  antenna  mounting  system  is  certified  by  the  Fed- 
eral Aviation  Administration  (FAA).  The  biologist 
should  be  familiar  with  areas  to  be  searched  and  re- 
view the  flight,  search  area,  and  procedures  with 
the  pilot  before  the  flight.  Exact  tracking  procedures 


688 


Radiotelemetry 


depend  on  terrain  and  habitat,  as  well  as  the  mobil- 
ity of  the  species  being  followed.  Experience  of  the 
observer  and  pilot  also  can  influence  the  efficiency 
of  aerial  tracking  (Hoskinson  1976). 


Yagi  antenna  mounting  system  on  aircraft. 


Data  Recording 

Radio  locations  need  to  be  plotted  initially  on 
maps  of  an  appropriate  scale  (U.S.  Geological  Survey 
1:24,000)  or  aerial  photographs  (for  many  species, 
a  l:7,920-scale  seems  best  [1  km  =   12.7  cm, 
1  mi  =  8  in.]).  A  clearly  defined  grid  coordinate 
system  should  be  superimposed  on  the  maps  or  pho- 
tos, using  acetate  (5  mil  thickness  works  well)  over- 
lays. Permanent  ink  felt  pens  should  be  used  for 
marking  the  grid  and  any  other  permanent  features, 
while  temporary  markers  are  best  for  plotting  on  the 
acetate.  Our  experience  has  indicated  that  black, 
red,  green,  blue,  and  purple  are  excellent  colors  to 
use.  After  the  locations  are  recorded  on  data  sheets, 
the  temporary  markings  can  be  wiped  off  the  acetate 
with  a  damp  cloth.  Our  experience  has  shown  that 
the  Universal  Transverse  Mercator  System  (UTM;  U.S. 
Department  of  the  Army  1958)  is  ideal  for  most 
radiotelemetry  studies.  UTM  coordinates  are  noted 
on  U.S.  Geological  Survey  maps  and  can  be  trans- 
ferred to  aerial  photographs  to  create  the  grid  coor- 
dinate system.  Data  then  can  be  digitally  recorded  to 
the  nearest  km  (0.62  mi.),  0.1  km  (0.06  mi.),  or 
even  0.001  km  (0.006  mi.)  as  appropriate  for  the 
species.  Additionally,  the  coordinates  can  be  entered 
directly  into  a  computer  terminal  for  data  analysis. 

Figure  2  illustrates  tracking  forms  used  for  re- 
cording radiotelemetry  data.  Observations  are  se- 
quentially noted  and  include  date,  time,  location,  and 
any  pertinent  behavioral  or  habitat  descriptions. 
Time  is  recorded  as  0000  to  2400  hrs  to  prevent  any 
possibility  of  errors.  UTM  coordinates  include  nu- 
merical and  directional  values,  for  example,  2760N, 


824E,  as  explained  by  the  U.S.  Department  of  the 
Army  (1958). 

Problems 

Radio  interference  can,  in  some  locations,  create 
difficulties.  Bearings  should  be  taken  from  natural 
high  points  in  the  area.  Power  lines,  fences,  and  large 
buildings  should  be  avoided  since  they  can  "bounce" 
radio  signals  and  readily  produce  false  and  highly 
confusing  bearings. 

The  number  of  biological  studies  using  radiote- 
lemetry has  been  increasing  each  year.  For  example, 
in  1985,  over  20,000  transmitters  were  placed  on 
wildlife  species  in  the  U.S.  and  a  considerable  por- 
tion of  these  were  not  authorized.  In  addition,  some 
of  the  unauthorized  frequencies  being  used  are  in 
the  range  of  police  or  other  users  who  are  emitting 
very  strong  signals  (compared  to  most  wildlife  trans- 
mitters) that  can  drastically  interfere  with  anyone 
trying  to  track  wildlife  on  these  frequencies.  These 
facts  stress  the  importance  of  having  the  proper 
clearances  for  use  of  particular  radio  frequencies  and 
the  need  to  coordinate  activities  with  telemetry 
users  in  the  same  area.  This  point  becomes  critical 
when  tracking  migrating  birds.  While  tracking  owls 
and  hawks,  problems  have  been  encountered  with 
other  researchers  using  unauthorized  transmitters  on 
black  bear.  Obviously,  one  could  be  dangerously 
surprised  "walking  out"  a  bear  while  thinking  it  was 
an  eastern  screech  owl. 


Radiotelemetry 


689 


■o 

c 

12 
u 


■o 
si 


690 


Radiotelemetry 


No  matter  how  well  individuals  or  equipment 
perform,  contact  with  some  transmitters  could  be 
lost.  To  help  locate  these  missing  animals,  aerial 
tracking  may  be  necessary.  In  addition,  receiving 
equipment  should  be  left  on  at  all  times  (tuned  to 
the  appropriate  missing  frequencies )  while  traveling 
in  vehicles;  missing  animals  are  often  located  when 
least  expected  (frequently  near  their  known  home 
range ).  A  bent  transmitter  antenna;  the  orientation  of 
an  animal's  body  and,  thus,  antenna  orientation;  or 
location  of  an  animal  in  a  low  or  structurally  se- 
cluded area  (such  as  underground)  may  drastically 
reduce  reception  range  and  thus  cause  the  animal  to 
essentially  "disappear."  Some  transmitters  will  come 
off  animals  (or  be  taken  off  by  them)  and,  depending 
where  they  end  up,  can  be  difficult  to  locate.  Unfor- 
tunately, some  lost  transmitters  are  never  located 
and  no  biological  conclusions  can  be  made  in  most 
of  these  situations. 


DATA  ANALYZING 

Radio-tracking  data  may  be  reported  in  several 
different  ways.  Initial  description  of  results  usually 
includes  the  length  of  time  that  the  animal  carried 
an  operating  transmitter,  fate  of  the  radio-equipped 
individual  (i.e.,  mortality,  lost  contact,  dead  battery), 
the  number  of  data  points  recorded,  and  possibly 
the  number  of  tracking  periods  or  days.  These  data 
are  helpful  in  relating  the  kind  and  number  of  obser- 
vations made  to  the  amount  of  tracking  time  and 
effort  spent. 

Enumeration  data,  reported  in  either  descriptive 
or  tabular  form,  include  the  frequency  at  which  var- 
ious events  occurred  or  when  observations  were 
made — for  example,  the  number  of  times  that  var- 
ious types  of  mortality  were  documented  or  the 
frequency  that  animals  were  located  at  each  type  of 
roost  or  habitat  type.  Statistical  comparisons  then 
can  be  made  between  the  frequency  of  habitat  use 
and  the  availability  of  those  same  habitats  within  the 
range  of  the  animal,  resulting  in  determining  habitat 
preferences  (Johnson  1980). 

Measurement  data  might  include  rates  of  move- 
ment (e.g.,  km/h),  rates  of  habitat  interception  (e.g., 
number  of  locations  per  habitat  type  per  hour),  time 
spent  in  each  habitat  type  (e.g.,  number  of  hours  or 
percentage  of  total  tracking  time ),  distances  between 
nest  sites  and  foraging  areas,  and  home  range  size. 
Home  range  size  may  be  calculated  several  different 
ways.  The  recommended  method  is  reviewing  home 
range  analysis  techniques  before  deciding  on  any 
one  method  (Jennrich  and  Turner  1969;  Dunn  and 
Gipson  1977;  Dixon  and  Chapman  1980;  MacDonald 
et  al.  1980;  Anderson  1982;  Hackett  and  Trevor- 
Deutsch  1982).  However,  most  often  radiotelemetry 
home  ranges  have  been  described  as  some  form  of 
a  minimum  area  or  convex  polygon  created  by  en- 


compassing the  recorded  outermost  points  (see  Bar- 
rett 1984;  Garshelis  and  Garshelis  1984).  Also, 
computer  programs  can  be  used  to  calculate  home 
range  size  (Ford  and  Krumme  1979;  Anderson 
1982). 

Presentation  of  radiotelemetry  data  in  graphic 
form  is  common  practice  and  may  offer  a  reader 
a  perceptive  view  of  movements.  To  prepare  a  figure 
showing  radiotelemetry  results,  tracing  paper  can 
be  placed  over  a  grid  pattern  of  UTM  coordinates 
marked  to  scale,  and  UTM  coordinates  recorded  in 
the  field  can  then  be  plotted  on  tracing  paper.  Var- 
ious landmarks  or  habitat  types  can  be  added  appro- 
priately to  emphasize  relationships  between 
locations,  movement  distances,  and  environmental 
features.  The  completed  art  work  then  can  be  repro- 
duced photographically.  Computer-assisted  graphics 
also  can  be  generated. 

When  diagramming  radiotelemetry  data,  one 
should  consider  showing  only  those  data  that  illus- 
trate a  particular  point,  rather  than  trying  to  present 
all  or  much  of  the  data  since  that  may  not  be  neces- 
sary and  can  result  in  a  confusing  illustration.  Move- 
ment data  presented  in  graphic  form  may  only 
exemplify  the  movement  observed  during  any  one 
tracking  period. 

Maps  showing  tracking  data  may  show  specific 
routes  or  sites  where  events  such  as  initial  capture, 
migration,  nesting,  or  mortality  occurred  (Figure  3). 
Other  types  of  graphics  can  include  "scatter"  dia- 
grams that  show  the  range  of  locations  and  centers 
of  activity  (Figure  4).  Also,  computer  graphics  can 
be  particularly  effective  in  relating  centers  of  activity 
or  utilization  distributions  (Tarter  and  Kronmal 
1976;  Anderson  1982),  especially  if  the  graphics  are 
in  three-dimensional  form  (Figure  5).  Range  overlap 
among  conspecifics,  or  predator  range  versus  that 
of  prey,  can  be  diagrammed  and  home  range  sizes 
presented  clearly  (Figure  6).  Continuous  movement 
patterns  can  illustrate  the  rate  and  directionality  of 
movement  (Figure  7),  and  the  same  radio-tracking 
data  imposed  on  habitat  types  can  show  distinctly 
the  relationships  between  movement  patterns  and 
particular  habitat  types  (Figure  8).  Active  video- 
graphic  techniques  allow  viewing  distribution  and 
movements  of  many  animals  in  a  continual,  spatial, 
and  temporal  dispersion  (White  1979). 

Once  locations  are  tabulated  or  graphically  plot- 
ted, a  critical  aspect  of  a  radiotelemetry  study  must 
follow,  that  is,  interpreting  the  findings  and  answer- 
ing why  an  animal  was  at  a  particular  place  at  a  par- 
ticular time.  Sanderson  ( 1966)  made  an  important 
point  when  he  stated  that  researchers  must  shift 
their  emphasis  from  concern  over  the  movements  of 
animals  to  the  reasons  for  the  movements.  Sanderson 
(1966)  continued  by  stating  that  movement  patterns 
are  established  and  regulated  by  the  density  of  the 


Radiotelemetry 


691 


LEGEND 
A  Release  site 
▼  Aircraft  location 
•  Satellite  location 
O  Visual  (Halo) 


92° 


~r 

91° 


90° 


89° 


Figure  3.     Movement  of  a  loggerhead  turtle  radio-tracked  from  October  16,  1979  to  June  8,  1980  in  the  Gulf 
of  Mexico  (from  Timko  and  Kolz  1982). 


N* 


•\ 


\ 


I  J\V.  . 


V  *  J>A 


N 


*** 


A 


O 


^ 


\ 


1  km 


y 


/ 


\ 


_./ 


y 


/ 


6>\ 


D  Experimental  Nest 
O  Treated  Farmstead 

♦  Day  Roost  VTimes  Observed 

•  Night  Location 


Figure  4.     Range,  night  locations,  and  daytime  roost  sites  recorded  for  a  female  common  barn  owl  with 
young,  3-10  weeks  old,  in  southwest  New  Jersey.  The  owl  was  radio-tracked  randomly  from  June  2  to 
July  29,  1982.  Dashed  lines  encompass  outermost  locations  recorded  (approximately  977  ha  [2,250  a.]; 
from  Colvin  1984). 


692 


Radiotelemetry 


o 

z 

I- 

0) 
LU 

cr 

l- 
z 

UJ 
O 

<r 
u 
a. 


NELSON  BAY 


Figure  5.     Distribution  of  resting  locations  (N  =   1,163)  of  radio -tagged  sea  otters  in  Nelson  Bay,  a  male  area 
in  Prince  William  Sound,  Alaska,  1979-1981.  Preferred  rest  area  was  near  the  center  of  the  Bay  (from 
Garshelis  and  Garshelis  1984). 


tN 


/" 


/ 


7 


\ 


/ 


/> 


s 


/ 


/ 


/       ^ 

s  / 

-'  / 
/     


i/ 


/ 


/ 


1  km 


\/ 


/ 


□  Nest  90 
—  Female 
(}  Nest  108 

Female 

___  Male 


Figure  6.     Home  range  overlap  of  three  adult,  common  barn  owls,  radio-tracked  in  southwest  New  Jersey, 
June-July  1982  (from  Colvin  1984). 


Radiotelemetry 


693 


d  Nest  Site,  Day  Roost 
•  Night  Location 

1  km 


Figure  7.     Pattern  showing  rate  of  movement  of  an 
adult,  common  male  barn-owl  between  its  nest 
site  (3  young,  7.5  weeks  old)  and  foraging  habi- 
tats in  southwest  New  Jersey.  The  owl  was 
radio-tracked  continuously  2043-2253  h  on 
12  August  and  2107-2330  h  on  13  August  1982 
(total  tracking  time  =  4  h,  33  min)  (from  Col- 
vin  1984). 


species,  food  supply,  reproductive  activity,  quality 
and  physiographic  arrangement  of  the  habitat  and, 
likely,  many  other  factors.  In  other  words,  the  total 
life-history  strategy  of  a  species  (Stearns  1976)  and 
both  the  biotic  and  abiotic  components  of  its  envi- 
ronment may  have  to  be  considered  when  interpret- 
ing the  significance  of  a  movement  pattern  or  the 
habitat  use  disclosed  by  radiotelemetry.  Therefore, 
background  information  on  a  species'  natural  history 
and  ecology  become  critical  in  accurately  interpret- 
ing movements.  Additionally,  detailed  data  on 
weather;  availability  of  selected  habitats  or  food  re- 
sources; population  density;  and  animal  age,  sex,  and 
reproductive  state  all  may  be  essential  in  evaluating 
movements  and  successfully  completing  the  project. 
Certainly,  radiotelemetry  locations  can  be  plotted 
and  presented;  however,  the  researcher  who  inter- 
prets the  data  in  context  with  species'  ecology  will 
get  the  most  out  of  the  time,  effort,  and  finances 
invested  in  radiotelemetry  studies. 


DISCUSSION 

Although  there  are  many  positive  reasons  for 
incorporating  radiotelemetry  as  part  of  a  research  ef- 
fort, a  researcher  should  not  first  decide  to  use  radio- 
telemetry and  then  seek  reasons  to  justify  its  use.  A 
researcher  should  first  rigorously  consider  the  spe- 
cific questions  being  addressed  or  investigated  and 
decide  whether  those  questions  can  be  answered 
without  radiotelemetry.  Whether  radiotelemetry  is 
truly  needed  too  often  becomes  a  subjective  proce- 
dure because  of  the  attractiveness  of  radiotelemetry 
technology.  Therefore,  we  strongly  recommend  that 
the  objectives  of  the  study,  as  well  as  the  reasons  for 
using  radiotelemetry,  be  well-defined  and  associated 
costs  and  time  be  evaluated  in  relation  to  the  type  of 
data  needed.  Many  other  techniques  for  marking 
and  studying  the  movements  of  animals  are  available 
and  may  be  more  appropriate.  These  include  tattoos, 
brands,  fur  removal,  aluminum  or  plastic  tags  or 
bands,  neck  collars,  fluorescent  bone  marker,  radio- 
active markers,  microtaggents,  and  dyes  (Day  et  al. 
1980).  Trap  recapture  or  direct  observation,  in  com- 
bination with  one  of  the  above  marking  techniques, 
may  prove  more  appropriate  than  radiotelemetry 
given  specific  research  designs,  budgets,  and  person- 
nel constraints. 

To  the  novice,  radiotelemetry  may  appear  to  be 
an  efficient,  simple,  and  exciting  way  to  study  wild- 
life. However,  when  considering  initiation  of  a  radio- 
telemetry project,  even  a  researcher  experienced  in 
radiotelemetry  is  reminded  of  radiotelemetry's  limi- 
tations and  the  time,  cost,  and  technical  problems 
that  can  be  involved. 

Radiotelemetry  has  been  used  in  many  specific 
ways  in  wildlife  studies.  Investigations  of  home  range 
and  general  movement  most  commonly  have  been 
performed,  and  these  studies  actually  are  often  the 
base  process  in  the  many  types  of  wildlife  research 
that  may  incorporate  radiotelemetry.  Home  range 
analysis  with  radiotelemetry  repeatedly  has  resulted 
in  a  much  expanded  view  of  species'  home  range 
and  potential  for  movement  than  previously  detected 
by  trapping  or  direct  observation  (Taylor  and  Quy 
1978).  Basic  studies  of  movement  may  focus  on  how 
movement  patterns  and  home  range  size  change  as 
a  function  of  sex  and  age,  population  density,  sea- 
sons of  the  year,  time  of  day,  or  available  habitat 
types  or  habitat  diversity  (Mech  et  al.  1966;  Trent 
and  Rongstad  1974;  Hemker  et  al.  1984). 

Beyond  the  initial  documentation  of  movement 
are  important  opportunities  to  investigate  habitat  use 
and  habitat  requirements  and  to  find  migration 
routes,  wintering  areas,  nesting  sites,  and  foraging 
areas  (Marshall  et  al.  1962;  Curtis  and  Braun  1983; 
Loft  et  al.  1984).  Also,  the  positive  or  negative  im- 
pact of  various  land  management  or  land-use  prac- 
tices on  wildlife  may  be  discernible  from  radio- 


694 


Radiotelemetry 


Figure  8.  Movement  pattern  and  habitats  intercepted  by  an  adult  male  common  barn-owl  with  3  young,  7.5 
weeks  old,  in  southwest  New  Jersey.  The  owl  was  continuously  radio-tracked  2043-2253  h  on  12  August 
and  2107-2330  h  on  13  August  1982  (total  tracking-time  =  4  h,  33  min.)  (from  Colvin  1984). 


equipped  individuals  that  are  subjected  to  a  chang- 
ing environment.  When  radio-tracking  and  habitat 
analysis  are  conducted  in  combination,  a  perceptive 
view  of  discrete  habitat  requirements  and  relation- 
ships between  habitat  and  population  maintenance 
may  be  achieved  (Kohn  and  Mooty  1971;  Jenkins 
and  Starkey  1984;  Pierce  and  Peek  1984;  Riley  and 
Dood  1984). 

Examination  of  population  dynamics,  including 
age-specific  survival  rates  and  mortality  factors,  has 
been  studied  often  with  radiotelemetry  (Stoddart 
1970;  Cook  et  al.  1971;  Barrett  1984).  A  radio  trans- 
mitter allows  continued  survivorship  to  be  docu- 
mented or,  conversely,  mortality  to  be  documented 
essentially  when  it  occurs.  Thus,  for  example,  a  more 
accurate  representation  of  the  occurrence  of  various 
mortality  factors  in  a  population  can  be  determined, 
compared  to  when  radiotelemetry  is  not  used,  and 
emphasis  is  placed  on  those  mortality  factors  that  are 


most  easily  identified  (e.g.,  car  collision).  Predator- 
prey  relationships  also  can  be  studied  and  provide 
additional  insight  into  population  dynamics  (Mech 
1967;  Kolenosky  1972;  Franzmann  et  al.  1980;  Fuller 
and  Keith  1980). 

Radiotelemetry  also  provides  great  advantages  in 
endangered  species  research  because  radio-equip- 
ping a  single  individual  can  potentially  locate  con- 
specifics  in  an  efficient  and  non-disruptive  manner 
(Mech  1977;  Fagerstone  et  al.  1985).  In  addition, 
because  endangered,  rare,  or  threatened  wildlife 
often  have  highly  specific  habitat  requirements,  these 
"micro"  habitats  may  be  more  clearly  and  quickly 
identified  with  radiotelemetry  and  thus  protected. 

Other  uses  of  radiotelemetry  have  included 
monitoring  the  status  and  movements  of  animals  in- 
volved in  translocations  or  reintroductions  (Fritts 
et  al.  1984).  Also,  population  censusing  has  been 


Radiotelemetry 


695 


conducted  by  determining  the  proportion  of  radio- 
equipped  animals  not  observed  in  a  direct  count  and 
adjusting  the  total  population  count  upwards  propor- 
tionately (Floyd  et  al.  1979).  Additionally,  vertebrate 
pest  research  often  has  included  radiotelemetry  in 
evaluation  of  movements  of  pest  species,  efficacy  of 
wildlife  control  procedures  on  target  species,  and 
toxic  hazards  to  non-target  species  (Taylor  1978; 
Fagerstone  et  al.  1981;  Hegdal  and  Blaskiewicz  1984; 
Heisterberg  et  al.  1984).  In  a  more  specialized  form, 
radiotelemetry  has  been  used  to  investigate  the  phys- 
iological adaptation  of  free-ranging  animals  to  their 
environment  by  remotely  monitoring,  for  example, 
heart  rate  (Kanwisher  et  al.  1978;  Follmann  et  al. 
1982). 

From  a  review  of  the  literature,  it  may  appear 
that  uses  of  radiotelemetry  are  limited  only  by  the 
creativity  of  researchers  and  available  hardware. 
As  the  technology  continues  to  improve,  so  will  the 
opportunities  to  attempt  new  uses  and  to  further 
address  complex  wildlife  and  ecological  issues. 


CONCLUSION 

There  are  four  major  components  of  a  radiotelem- 
etry study: 


( 1 )  Justified  use  of  the  technology  based  on  a 
specific  research  design; 


(2)    Consideration  of  hardware,  personnel,  cost, 
and  sampling  strategies; 


(3)    Implementation  of  field  research  and  data 
collection;  and 


(4)    Analysis  and  interpretation  of  data. 

Each  of  the  above  components  is  equally  important 
in  the  planning  and  success  of  a  radiotelemetry 
study.  Often,  because  of  the  specialization  of  these 
components  and  the  time  involved,  a  team  approach 
is  essential  to  adequately  plan,  conduct,  and  evaluate 
radiotelemetry  studies.  The  experience  of  the  "team" 
will  strongly  affect  the  efficiency,  accuracy  and,  thus, 
the  outcome  of  the  project.  In  final  form,  discussion 
of  radiotelemetry  should  be  well-integrated  with 
information  on  species  biology  and  ecology  to  best 
understand  and  illustrate  the  role  that  movement, 
migration,  and  habitat  use  play  in  the  life-history 
strategy  of  a  species. 


Radio-tagged  mourning  dove. 


Burn  owl  instrumented  with  a  transmitter  and  whip  an- 
tenna. 


696 


Radiotelemetry 


LITERATURE  CITED 

AMLANER,  C.J.,  Jr.  1978.  Biotelemetry  from  free-ranging 
animals.  Pages  205-228  in  Stonehouse,  B.,  ed.  Animal 
Marking:  Recognition  Marking  of  Animals  in  Research. 
Macmillan  and  Company,  London. 

AMSTRUP,  S.C.  1980.  A  radio-collar  for  game  birds.  J. 
Wildl.  Manage.  44:214-217. 

ANDERSON,  D.J.  1982.  The  home  range:  A  new  nonpara- 
metric  estimation  technique.  Ecology  63:103-112. 

BARRETT,  M.W.  1984.  Movements,  habitat  use,  and  preda- 
tion  on  pronghorn  fawns  in  Alberta.  J.  Wildl.  Manage. 
48:542-550. 

BRAY,  O.E.  and  G.W.  CORNER.  1972.  A  tail  clip  for  attach- 
ing transmitters  to  birds.  J.  Wildl.  Manage.  36:640- 
642. 

,  RE.  JOHNSON,  and  A.L.  KOLZ.  1975.  A  removable 

cartop  antenna  system  for  radio-tracking  birds.  Bird- 
banding  46:15-18. 

BRUGGERS,  R.,  J.  ELLIS,  J.  SEDGWICK,  and  J.  BOURASSA. 
1981.  A  radio  transmitter  for  monitoring  the  move- 
ments of  small  passerine  birds.  Proc.  Int.  Conf.  Wildl. 
Biotelemetry  3:69-79. 

,  MM.  JAEGER,  and  J.B.  BOURASSA.  1983.  The  ap- 
plication of  radiotelemetry  for  locating  and  control- 
ling concentrations  of  red-billed  quelea  in  Africa. 
Tropical  Pest  Manage.  29:27-32. 

CARR,  A.  1965.  The  navigation  of  the  green  turtle.  Sci. 
Am.  212:79-86. 

COCHRAN,  W.W.  1975.  Following  a  migrating  peregrine 
from  Wisconsin  to  Mexico.  Hawk  Chalk  14:28-37. 

.  1980.  Wildlife  telemetry.  Pages  507-520  in  Schem- 

nitz,  S.D.,  ed.  Wildl.  Manage.  Techniques  Manual. 
Wildl.  Soc,  Inc.,  Washington,  DC. 

COLVIN,  B.A.  1984.  Barn  owl  foraging  behavior  and  sec- 
ondary poisoning  hazard  from  rodenticide  use  on 
farms.  Ph.D.  dissertation.  Bowling  Green  State  Univ., 
Bowling  Green,  OH.  326pp. 

COOK,  R.S.,  M.W.  WHITE,  DO.  TRAINER,  and  W.C. 
GLAZENER.  1971.  Mortality  of  young  white-tailed 
deer  fawns  in  south  Texas.  J.  Wildl.  Manage.  35:47-56. 

COVICH,  A.  1 977.  Shapes  of  foraging  areas  used  by  radio- 
monitored  crayfish.  Am.  Zoology  17:205  (Abstract). 

CURTIS,  P.D.  and  C.E.  BRAUN.  1983.  Radiotelemetry  loca- 
tion of  nesting  band-tailed  pigeons  in  Colorado.  Wil- 
son Bull.  95:464-466. 

DAY,  G.I.,  S.D.  SCHEMNITZ,  and  R.D.  TABER.  1980.  Cap- 
turing and  marking  wild  animals.  Pages  61-68  in 
Schemnitz,  S.D.,  ed.  Wildl.  Manage.  Techniques  Man- 
ual. Wildl.  Soc,  Inc.,  Washington,  DC. 

DKON,  KR.  and  J.A.  CHAPMAN.  1980.  Harmonic  mean 
measure  of  animal  activity  areas.  Ecology  61:1040- 
1044. 

DUNN,  J.E.  and  P.S.  GIPSON.  1977.  Analysis  of  radiotelem- 
etry data  in  studies  of  home  range.  Biometrics  33:85- 
101. 

FAGERSTONE,  KA.,  D.E.  BIGGINS,  and  T.M.  CAMPBELL, 
III.  1985.  Marking  and  tagging  of  black-footed  ferrets 
(Mustela  nigripes).  Pages  10.1-10.10  in  Anderson, 
S.H.  and  D.B.  Inkley,  eds.  Proc.  Black-footed  Ferret 
Workshop,  Sept.  18-19,  Univ.  Wyoming,  Laramie. 

,  G.H.  MATSCHKE,  and  D.  ELIAS.  1981.  Radiotelem- 
etry to  evaluate  effectiveness  of  a  new  fumigant  car- 
tridge for  controlling  ground  squirrels.  Proc.  Int. 
Conf.  on  Wildl.  Biotelemetry.  Univ.  Wyoming,  Lara- 
mie. 3:20-25. 

FITZNER,  RE.  and  J.N.  FITZNER.  1977.  A  hot  melt  glue 


technique  for  attaching  radio  transmitter  tail  packages 
to  raptorial  birds.  North  Am.  Bird  Bander  2:56-57. 

FLOYD,  T.J.,  L.D.  MECH,  and  ME.  NELSON.  1979.  An 
improved  method  of  censusing  deer  in  deciduous- 
coniferous  forests.  J.  Wildl.  Manage.  43:258-261. 

FOLLMANN,  E.H.,  A.E.  MANNING,  and  J.L.  STUART.  1982. 
A  long-range  implantable  heart  rate  transmitter  for 
free-ranging  animals.  Biotelemetry  Patient  Monitoring 
9:205-212. 

FORD,  R.G.  and  D.W.  KRUMME.  1979.  The  analysis  of 
space  use  patterns.  J.  Theoretical  Biol.  76:125-155. 

FRANZMANN,  A.W.,  C.C.  SCHWARTZ,  and  R.O.  PETER- 
SON. 1980.  Moose  calf  mortality  in  summer  on  the 
Kenai  Peninsula,  Alaska.  J.  Wildl.  Manage.  44:764-768. 

FRITTS,  S.H.,  W.J.  PAUL,  and  L.D.  MECH.  1984.  Move- 
ments of  translocated  wolves  in  Minnesota.  J.  Wildl. 
Manage.  48:709-721. 

FULLER,  T.K  and  LB.  KEITH.  1980.  Wolf  population 
dynamics  and  prey  relationships  in  northeastern  Al- 
berta. J.  Wildl.  Manage.  44:583-602. 

GARSHELIS,  D.L.  andJ.A.  GARSHELIS.  1984.  Movements 
and  management  of  sea  otters  in  Alaska.  J.  Wildl. 
Manage.  48:665-678. 

GILMER,  D.S.,  L.M.  COWARDIN,  R.L.  DUVAL,  L.M.  MECH- 
LIN, C.W.  SHAIFFER,  and  V.B.  KUECHLE.  1981.  Proce- 
dures for  the  use  of  aircraft  in  wildlife  biotelemetry 
studies.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  Resour. 
Publ.  140.  19pp. 

,  V.B.  KUECHLE,  and  I.J.  BELL,  Jr.  1971.  A  device  for 

monitoring  radio-marked  animals.  J.  Wildl.  Manage. 
35:829-832. 

HACKETT,  D.F.  and  B.  TREVOR-DEUTSCH.  1982.  Radiote- 
lemetric  assessment  of  grid-trapping  techniques  in  a 
study  of  the  eastern  chipmunk  {Tatnias  striatiis  L. ). 
Biotelemetry  Patient  Monitoring.  9:213-226. 

HARRINGTON,  F.H.  and  L.D.  MECH.  1982.  Patterns  of 
homesite  attendance  in  two  Minnesota  wolf  packs. 
Pages  81-105  in  Harrington,  F.H.  and  P.C.  Paquet,  eds. 
Wolves  of  the  World.  Noyes,  Park  Ridge,  NJ. 

HEGDAL,  PL.  and  R.W.  BLASKIEWICZ.  1984.  Evaluation 
of  the  potential  hazard  to  barn  owls  of  Talon  {brodi- 
facoum  bait)  used  to  control  rats  and  house  mice. 
Environmental  Toxicological  Chemistry  3:167-179- 

and  T.A.  GATZ.  1978.  Technology  of  radio-tracking 

for  various  birds  and  mammals.  Pages  204-206  in 
Symp.  on  Application  of  Remote  Sensing  Data  to 
Wildlife  Management.  PECORA  IV.  Sioux  Falls,  SD. 

HEISTERBERG,  J.F.,  C.E.  KNITTLE,  O.E.  BRAY,  D.F.  MOTT, 
andJ.F.  BESSER.  1984.  Movements  of  radio-instru- 
mented blackbirds  and  European  starlings  among  win- 
ter roosts.  J.  Wildl.  Manage.  48:203-211. 

HEMKER,  T.P.,  F.G.  LINDZEY,  and  B.B.  ACKERMAN.  1984. 
Population  characteristics  and  movement  patterns  of 
cougars  in  southern  Utah.  J.  Wildl.  Manage.  48:1275- 
1284. 

HOSKINSON,  R.L.  1976.  The  effect  of  different  pilots  on 
aerial  telemetry  error.  J.  Wildl.  Manage.  40:137-139. 

IMBODEN,  C.  1975.  A  brief  radiotelemetry  study  on 
moreporks.  Notornis  2:221-230. 

JANSEN,  D.K  1982.  A  new  potting  material  for  radiotelem- 
etry packages.  Copeia  1982:189. 

JENKINS,  KJ.  and  E.E.  STARKEY.  1984.  Habitat  use  by 

Roosevelt  elk  in  unmanaged  forests  of  the  Hoh  Valley, 
Washington.  J.  Wildl.  Manage.  48:642-646. 

JENNRICH,  R.L  and  F.B.  TURNER.  1969.  Measurement  of 
non-circular  home  range.  J.  Theoretical  Biol.  22:227- 
237. 


Radiotelemetry 


697 


JOHNSON,  D.H.  1980.  The  comparison  of  usage  and  avail- 
ability measurements  for  evaluating  resource  prefer- 
ence. Ecology  61:65-71. 

KANWISHER,  J.W.,  T.C.  WILLIAMS,  J.M.  TEAL,  and  KO. 

LAWSON,  Jr.  1978.  Radiotelemetry  of  heart  rates  from 
free-ranging  gulls.  Auk  95:288-293- 

KENWARD,  RE.,  G.J.  HIRONS,  and  F.  ZIESEMER.  1982. 
Devices  for  telemetering  the  behavior  of  free-living 
birds.  Pages  127-129  in  Cheesman,  C.L.  and  R.B. 
Mitson,  eds.  Telemetric  Studies  of  Vertebrates.  Zoo- 
logical Soc.  London  Symp.  49. 

KEUCHLE,  V.B.,  DP.  DEMASTER,  and  D.B.  SINIFF.  1979. 
State  of  the  art  of  and  needs  of  the  earth  platform. 
Proc.  Int.  Symp.  Remote  Sensing  Environment  13505- 
518. 

KOHN,  BE.  and  J.J.  MOOTY.  1971.  Summer  habitat  of 

white-tailed  deer  in  north-central  Minnesota.  J.  Wildl. 
Manage.  35:476-487. 

KOLENOSKY,  G.B.  1972.  Wolf  predation  on  wintering 

deer  in  east-central  Ontario.  J.  Wildl.  Manage.  36:357- 
369. 

KOLZ,  A.L.  1975.  Mortality-sensing  wildlife  transmitters. 
Int.  ISA  Biomedical  Sci.  Instrument  Symp.  12:57-60. 

,  J.W.  LENTFER,  and  H.G.  FALLEK  1980.  Satellite  ra- 
dio-tracking of  polar  bears  instrumented  in  Alaska. 
Pages  743-752  in  Amlaner,  C.J.,  Jr.  and  D.W.  Mac- 
Donald,  eds.  A  Handbook  on  Biotelemetry  and  Radio- 
Tracking.  Pergamon  Press,  Oxford. 

LEE,  J.E  ,  G.C.  WHITE,  R.A.  GARROTT,  R.M.  BARTMANN, 
and  AW.  ALLDREDGE.  1985.  Accessing  accuracy  of  a 
radiotelemetry  system  for  estimating  animal  locations. 
J.  Wildl.  Manage.  49:658-663. 

LOFT,  E.R.,  J.W.  MENKE,  and  T.S.  BURTON.  1984.  Seasonal 
movements  and  summer  habitats  of  female  black- 
tailed  deer.  J.  Wildl.  Manage.  48:1317-1325. 

MACDONALD,  D.W.,  F.G.  BALL,  and  N.G.  HOUGH.  1980. 
The  evaluation  of  home  range  size  and  configuration 
using  radio-tracking  data.  Pages  405-424  in  Amlaner, 
C.J.,  Jr.  and  D.W.  MacDonald,  eds.  A  Handbook  on 
Biotelemetry  and  Radio-Tracking.  Pergamon  Press, 
Oxford. 

MADISON,  DM.  1978.  Movement  indicators  of  reproduc- 
tive events  among  female  meadow  voles  as  revealed 
by  radiotelemetry.  J.  Mammal.  59:835-843. 

MARSHALL,  W.H.,  G.W.  GULLION,  AND  R.G.  SCHWAB. 
1962.  Early  summer  activities  of  porcupines  as  deter- 
mined by  radio-positioning  techniques.  J.  Wildl.  Man- 
age. 26(l):75-79. 

and  J.J.  KUPA.  1963.  Development  and  use  of  radio- 
telemetry techniques  for  ruffed  grouse  studies.  Trans. 
North  Am.  Wildl.  Nat.  Resour.  Conf.  28:443-456. 

MECH,  L.D.  1967.  Telemetry  as  a  technique  in  the  study 
of  predation.  J.  Wildl.  Manage.  31:492-496. 

.  1977.  Productivity,  mortality,  and  population  trend 

in  wolves  from  northeastern  Minnesota.  L  Mammal. 
58:559-574. 

.  1983.  Handbook  of  animal  radio-tracking.  Univ. 

Minnesota  Press,  Minneapolis.  107pp. 

,  R.C.  CHAPMAN,  WW.  COCHRAN,  L.  SIMMONS, 

and  U.S.  SEAL.  1984.  A  radio-triggered  anesthetic-dart 
collar  for  recapturing  free-ranging  mammals.  Wildl. 
Soc.  Bull.  12:69-74. 

,  KL.  HEEZEN,  and  D.B.  SINIFF.  1966.  Onset  and 


cessation  of  activity  in  cottontail  rabbits  and 


snowshoe  hares  in  relation  to  sunset  and  sunrise. 
Animal  Behavior  14:410-413- 

MERSON,  M.H.,  RE.  BYERS,  and  L.D.  LETA.  1982.  A  porta- 
ble antenna  base  for  fixed-station,  radio-tracking. 
Wildl.  Soc.  Bull.  10:44-45. 

NICHOLLS,  T.H.  and  D.W.  WARNER.  1972.  Barred  owl 
habitat  use  as  determined  by  radiotelemetry.  J.  Wildl. 
Manage.  36:213-224. 

OSGOOD,  D.W.  1970.  Thermoregulation  in  water  snakes 
studied  by  telemetry.  Copeia  1970:568-571. 

PIERCE,  J.D.  and  J.M.  PEEK.  1984.  Moose  habitat  use  and 
selection  patterns  in  north-central  Idaho.  J.  Wildl. 
Manage.  48:1335-1343- 

RILEY,  S.J.  and  A.R.  DOOD.  1984.  Summer  movements, 
home  range,  habitat  use,  and  behavior  of  mule  deer 
fawns.  J.  Wildl.  Manage.  48:1302-1310. 

SALTZ,  D.  and  P.U.  ALKON.  1985.  A  simple  computer- 
aided  method  for  estimating  radio-location  error.  J. 
Wildl.  Manage.  49:664-668. 

SANDERSON,  G.C.  1966.  The  study  of  mammal  move- 
ments— a  review.  J.  Wildl.  Manage.  30:215-235. 

SCHUBAUER,  J.P.  1981.  A  reliable  radiotelemetry  tracking 
system  suitable  for  studies  of  Chelonians.  J.  Herpetol- 
ogy  15:117-120. 

SMITH,  G.J.,  JR.  CARY,  and  O.J.  RONGSTAD.  1981.  Sam- 
pling strategies  for  radio-tracking  coyotes.  Wildl.  Soc. 
Bull.  9:88-93. 

STEARNS,  S.C.  1976.  Life-history  tactics:  A  review  of  the 
ideas.  Quarterly  Review  Biol.  51:1-45. 

STODDART,  L.C.  1970.  A  telemetric  method  for  detecting 
jackrabbit  mortality.  J.  Wildl.  Manage.  34:501-507. 

TARTER,  ME.  and  R.A.  KRONMAL.  1976.  An  introduction 
to  the  implementation  and  theory  of  nonparametric 
density  estimation.  Am.  Statistician  30:105-112. 

TAYLOR,  K.D.  1978.  Range  of  movement  and  activity  of 
common  rats  {Rattiis  norvegicus)  on  agricultural 
land.  J.  Applied  Ecol.  1 5:663-677. 

and  R.J.  QLIY.  1978.  Long-distance  movements  of  a 

common  rat  {Rattus  norvegicus)  revealed  by  radio- 
tracking.  Mammalia  42:63-71. 

TIMKO,  RE.  and  A.L.  KOLZ.  1982.  Satellite  sea  turtle 
tracking.  Marine  Fish.  Review  44:19-24. 

TRENT,  T.T.  and  O.J.  RONGSTAD.  1974.  Home  range  and 
survival  of  cottontail  rabbits  in  southwestern  Wiscon- 
sin. J.  Wildl.  Manage.  38:459-472. 

U.S.  DEPARTMENT  OF  THE  ARMY.  1958.  Universal 
Transverse  Mercator  Grid.  Technical  Manual  TM  5- 
241-8.  U.S.  Govt.  Printing  Office  1969-390-947/404. 
66pp. 

VERTS,  B.J.  1963.  Equipment  and  techniques  for  radio- 
tracking  striped  skunks.  J.  Wildl.  Manage.  27:325-339 

WHITE,  G.C.  1979.  Computer-generated  movies  to  display 
biotelemetry  data.  Proc.  Int.  Conf.  Wildl.  Biotelemetry 
2:210-214. 

WINTER,  J.D.,  V.B.  KUECHLE,  D.B.  SINIFF,  and  JR. 

TESTER.  1978.  Equipment  and  methods  for  radio- 
tracking  freshwater  fish.  Misc.  Rep.  152.  Agric.  Experi- 
ment Station,  Univ.  Minnesota. 

WOLCOTT,  T.G.  1980.  Optical  and  radio  optical  tech- 
niques for  tracking  nocturnal  animals.  Pages  333-338 
in  Amlaner,  C.J.,  Jr.  and  D.W.  MacDonald,  eds.  A 
Handbook  on  Biotelemetry  and  Radio-Tracking.  Per- 
gamon Press,  Oxford. 


698 


Radiotelemetry 


34 

FOOD  HABITS 


Allen  Y.  Cooperrider 

U.S.  Bureau  of  Land  Management 
Service  Center 
Denver,  CO  80225 


Editor's  Note:  An  adequate  food  supply  is  one  of  the 
most  basic  habitat  requirements  of  all  species. 
Therefore,  some  understanding  of  food  habits  is  a 
prerequisite  to  evaluating  habitat  and  providing 
good  habitat  management.  Similarly,  forage  utili- 
zation is  an  important  factor  in  habitat  monitor- 
ing. The  two  closely  related  topics  are  discussed 
in  this  chapter,  emphasizing  herbivore  food  habits 
and  vegetation  utilization. 


"Food  is  the  burning  question  in  animal  society,  and 
the  whole  structure  and  activities  of  the  community 
are  dependent  upon  questions  of  food  supply." 


— Charles  Elton,  from  Animal  Ecology 


INTRODUCTION 

Measuring  food  habits  and  forage  utilization  may 
appear  to  be  rather  specialized  monitoring  tech- 
niques; however,  such  information  is  frequently  re- 
quired in  monitoring  studies,  particularly  when 
working  with  herbivores.  These  animals  are  capable 
of  destroying,  damaging,  or  altering  their  habitat 
through  overgrazing  or  overbrowsing.  Such  impacts 
may  affect  not  only  the  species  itself  but  many  oth- 
ers as  well.  Similarly,  the  species  responsible  for  the 
damage  may  be  either  a  wild  or  domestic  ungulate, 
or  combinations  of  one  or  more  of  these.  On  public 
lands  of  western  North  America,  where  many  species 
of  wild  and  domestic  ungulates  coexist  on  ranges 
being  managed  for  multiple  uses,  information  on 
forage  utilization  and  food  habits  can  be  quite 
valuable. 


Furthermore,  the  need  for  an  understanding  of 
food  habits  of  species  other  than  ungulates  will  prob- 
ably increase.  Twenty  years  ago,  Gullion  (1966) 
suggested  the  importance  of  understanding  food  hab- 
its of  game  birds.  Today,  biologists  are  charged  with 
managing  hundreds  of  vertebrate  species,  of  which 
the  majority  of  food  habits  are  largely  unknown. 

In  this  chapter,  I  describe  methods  for  measur- 
ing food  habits  and  forage  utilization,  discuss  prob- 
lems associated  with  such  measurements,  and 
present  and  evaluate  methods  of  analyzing  such  data. 
This  chapter  emphasizes  techniques  for  herbivores 
since  these  are  more  commonly  required  in  monitor- 
ing studies  by  management  biologists.  Techniques 
for  omnivores  and  carnivores  are  covered  briefly  and 
reference  is  made  to  publications  with  more  detailed 
information  on  this  subject.  Similarly,  techniques  for 
measuring  forage  utilization  will  be  described  in 
less  detail  than  those  for  food  habits  since  utilization 
techniques  are  well-described  in  numerous,  readily 
available  publications. 

In  the  literature  on  subjects  such  as  food  habits, 
forage  selection,  preference,  and  utilization  have 
been  used  quite  loosely  and  even  interchangeably. 
More  seriously,  this  has  often  resulted  in  biologists 


Food  Habits 


699 


confusing  concepts  and  stating  erroneous  conclu- 
sions. In  particular,  workers  have  confused  food  hab- 
its or  diet  with  forage  preference.  To  avoid  further 
confusion,  I  therefore  begin  with  brief  definitions  of 
each  of  these  measurements. 

Food  habits  or  diets  refer  to  the  actual  foods 
that  are  eaten  by  one  or  more  animals  of  the  same 
species  during  a  given  season  or  time  period  within 
a  defined  geographic  area.  Thus,  food  habits  are  char- 
acteristic of  an  animal  species,  but  only  for  a  given 
time  and  place;  extrapolation  of  food  habit  data  to 
other  times  and  regions,  without  supplementary  data, 
is  frequently  misleading.  Food  habits  are  typically 
expressed  as  percentages  of  the  overall  diet,  which 
consists  of  each  individual  food  item  (plant  or  ani- 
mal species  or  species  group;  Figure  1 );  this  is  the 
most  useful  and  easily  interpretable  measure.  The 
percentages  may  be  by  weight  or  volume  depending 
on  the  technique.  Usually  the  two  are  pretty  well 
correlated;  however,  food  habits  are  also  sometimes 
reported  as  a  frequency  (percentage  of  samples  con- 
taining each  food  item)  or,  more  rarely,  as  the 
weight  of  each  food  item  in  the  daily  (or  seasonal) 
diet. 


Forage  utilization  refers  to  the  amount  of  vege- 
tation (or  food  supply)  removed  by  foraging  animals 
from  a  defined  area  of  land  during  a  defined  period. 
Utilization  measurements  are  rarely  made  except 
on  plant  forages,  although  the  same  principles  apply. 
They  are  thus  attributes  of  the  land,  vegetation,  or 
food  supply  and  not  of  an  animal  or  species  of  ani- 
mal. Utilization  measurements  are  usually  expressed 
as  the  percentage  of  the  available  forage  or  annual 
forage  production  removed  during  a  given  season  or 
year  (Figure  2).  They  may  be  broken  down  by  plant 
species  or  group  and  the  percentage  may  be  by 
weight  or  volume. 


Forage  utilization  may  also  be  expressed  in  a 
variety  of  other  ways  such  as  the  percentage  of 
plants  or  twigs  that  are  grazed  or  browsed,  or  the 
height  of  grazed  down  plants.  In  practice,  this  utiliza- 
tion may  be  attributed  to  one  species  of  animal  or 
to  several.  When  several  species  of  large  herbivores 
occupy  the  same  range  at  the  same  time,  determin- 
ing the  relative  amount  used  by  each  animal  could 
be  difficult  or  even  impossible. 


Food  habits  of  mule  deer  on  Trickle  Mountain, 
Colorado,  Spring  1978. 


tr 

03 

-C 

o 
b_ 


Food  habits  of  bighorn  sheep  on  Trickle 
Mountain,  Colorado,  Spring  1978. 


"Grass  and  grasslike  plants 

Muhly  (Muhlenbergia  sp.)  16% 

Fescue  (Festuca  sp.)  11% 

Sedge  (Cyperace)  8% 

Other  17% 

Forbs 

Sagebrush  (Artemisia  sp.)  22% 

L       Other  19% 


52% 


7% 


Estimated  food  habits  of  elk, 
Trickle  Mountain,  Colorado,  1978. 


a    GRASS 

□    FORBS  JAN  I 

■-E3    BROWSE        I978 


100% 


JULY 
1978 


Figure  1.     Food  habits  refer  to  the  percentage  of  each  forage  species,  items,  or  classes  in  the  diet  for  a 
given  period  and  location.  These  habits  may  be  depicted  in  a  pie  chart,  table,  or  graph. 


700 


Food  Habits 


Mountain 
Mahogany 
30% 


Wildrye 


1% 


Western 
Wheatgrass 
85% 


Figure  2.     Forage  utilization  refers  to  the  percentage  of  annual  production  used  by  one  or  more  animals. 


The  total  amount  of  forage  produced  on  an  area 
during  a  given  year  or  season  is  called  forage  pro- 
duction. This  production  may  be  used  synonymously 
with  availability  or  corrected  for  forage  physically 
or  seasonally  unavailable  to  animals  for  various  rea- 
sons. Measurement  of  forage  production  and  availa- 
bility are  not  described  in  this  book  in  any  detail 
because  they  are  covered  elsewhere  (U.S.  Depart- 
ment of  the  Interior,  Bureau  of  Land  Management 
1984).  However,  they  have  been  introduced  here  to 
clarify  the  concept  of  forage. 

In  contrast,  forage  preference  is  the  relative 
degree  to  which  animals  select  one  forage  species 
over  another.  It  is  thus  an  attribute  of  an  animal  or 
species.  However,  to  quantify  this  variable,  both  food 
supply  and  food  habits  need  to  be  measured.  The 
simplest  measurement  of  forage  preference  is  the 
percentage  ratio  of  a  given  food  item  in  the  diet  to 
the  entire  food  supply  (available  forage)  that  is  pres- 
ent (Table  1 ).  Clearly,  both  measurements  must 
occur  at  the  same  time  and  in  the  same  geographical 
area.  (Food  items  with  preference  values  greater 
than  1  are  preferred.)  The  important  concept  to  re- 
member is  that  food  habits  are  not  synonymous  with 
forage  preference  even  though  they  are  often  treated 
the  same  in  much  of  the  older  wildlife  and  range 
management  literature.  Put  simply,  forage  preference 
is  what  an  animal  likes  to  eat;  food  habits  are  what 
it  actually  eats.  For  further  discussion  on  this  see 
Petrides(1975). 


MEASUREMENT  OF  FORAGE  UTILIZATION 

Two  basic  procedures  are  available  for  deter- 
mining forage  utilization:  ( 1 )  difference  methods  and 
(  2 )  grazed-plant  methods.  Since  many  variations  on 
these  techniques  have  been  developed  and  used, 
no  single  technique  can  be  prescribed  as  "best." 


Rather,  the  investigator  will  have  to  determine  the 
most  appropriate  and  efficient  method  for  the  prob- 
lem at  hand. 


Difference  Methods 

Difference  methods  require  estimating  the 
amount  of  forage  or  herbage  produced  with  and 
without  grazing.  The  difference  then  is  used  to  esti- 
mate the  amount  used.  Estimates  can  be  obtained 
in  two  different  ways:  single  plots,  before  and  after 
grazing,  and  paired  plots,  one  grazed  the  other  not. 

In  the  before-and-after  approach,  the  amount  of 
forage  is  estimated  or  measured  on  an  area  before 
and  after  a  grazing  period  or  season.  Any  acceptable 
technique  for  measuring  or  estimating  forage  pro- 
duction can  be  used;  however,  variances  in  the  esti- 
mates of  differences  can  be  reduced  substantially 
by  using  permanent  or  marked  plots  for  both  meas- 
urements. In  this  case,  the  first  (before)  measure- 
ment cannot  be  made  by  using  a  removal  method 
such  as  clipping.  To  reduce  variances  in  utilization 
estimates,  yet  still  use  clipping  methods,  many  work- 
ers have  used  paired  plots.  Using  this  method,  a  se- 
ries of  paired  plots  are  selected  that  appear  to  be 
similar  in  species  composition  and  total  forage.  One 
plot  is  then  selected  randomly  for  clipping  prior  to 
grazing;  the  other  is  clipped  at  the  end  of  the  season. 

In  the  paired  plot  approach,  cages  or  other  ex- 
closures  are  used  to  estimate  ungrazed  forage.  Before 
the  grazing  season  or  period,  paired  plots  with  simi- 
lar species  composition  and  total  forage  production 
are  selected,  and  one  or  the  other  plot  is  randomly 
chosen  to  be  caged  or  excluded  from  grazing.  At  the 
end  of  the  grazing  period,  the  amount  of  forage  on 


Food  Habits 


701 


Table  1.     Forage  production,  food  habits,  and  forage  preference  of  bighorn  sheep  on  Trickle  Mountain, 
Colorado,  1978  (from  Bailey  and  Cooperrider  1982). 


Percent  of 

Percent  in 

Forage 

Diet  of 

Preference  Index 

Vegetation 

Production 

Bighorn  Sheep 

for  Bighorn  Sheep 

GRASS 

Arizona  fescue 

12 

7 

0.58 

(Festuca  arizonica) 

Blue  grama 

15 

4 

0.26 

(Boutelova  gracilis) 

Muhly 

11 

10 

0.91 

{Muhlenbergia  sp.) 

Sedge 

4 

5 

1.25 

(Cyperaceal) 

Other 

11 

12 

0.92 

FORBS 

41 

6 

0.15 

BROWSE 

Rabbitbrush 

3 

* 

<0.01 

(Chrysothamnus  sp.) 

Sagebrush 

1 

25 

25.00 

(Artemisia  sp.) 

Saltbrush 

* 

11 

>  100.00 

{Atriplex  sp.) 

True  Mountain  Mahogany 

* 

6 

>  100.00 

(Cercocarpus  sp.) 

Other 

2 

14 

7.00 

Note:  The  preference  index  is  calculated  by  dividing  percent  in  diet  by  percent  of  forage  production. 


trace 


both  the  ungrazed  and  grazed  plot  are  measured 
or  estimated  through  clipping  or  other  appropriate 
methods. 

Both  types  of  difference  methods  have  problems 
and  pitfalls.  Obtaining  good  estimates  of  forage  pro- 
duction is  time-consuming  and  expensive.  If  meas- 
urements have  to  be  made  twice,  then  the  costs 
in  time  and  money  increase.  Furthermore,  forage 
production  estimates  will  yield  even  less  precision. 
Forage  production  estimates  are  rarely  better  than 
within  ±  10%.  Therefore,  difference  methods 
should  not  be  used  unless  utilization  is  expected  to 
be  high  (50%  or  more).  Difference  methods  thus 
may  be  good  for  determining  high  rates  of  utilization 
(50%  to  100%  ),  or  for  determining  if  utilization  is 
high  or  not.  These  methods,  however,  are  not  partic- 
ularly good  for  determining  small  differences  in  utili- 
zation, for  example,  distinguishing  between  40% 
and  50%  utilization. 


Grazed-Plant  Methods 

Grazed-plant  methods  are  used  for  both  herba- 
ceous plants  and  woody  plants,  but  plants  are  exam- 


ined only  at  the  end  of  the  grazing  season.  These 
methods  fall  into  three  general  categories: 

( 1 )  frequency  methods, 

(2)  height  or  length  conversion  methods,  and 

(3)  form-class  or  ocular  estimates. 

Frequency  methods  require  the  biologist  to  count  a 
number  of  plants  to  determine  how  many  have  been 
grazed  or  not.  Previously  developed  regression  tables 
can  then  be  used  to  convert  the  percentage  of 
grazed  plants  to  percentage  of  utilization  (Cook  and 
Stubbendieck  1986).  Alternatively,  the  average 
weight  of  grazed  portions  can  be  determined  and 
used  to  calculate  utilization  (U.S.  Department  of  the 
Interior,  Bureau  of  Land  Management  1984:21).  This 
technique  can  also  be  used  with  browse  species  by 
counting  the  number  of  unbrowsed  and  browsed 
leaders. 

Height  or  length  conversion  methods  require 
the  biologist  to  determine  the  average  height  that 
herbaceous  species  have  been  grazed  or  the  average 
remaining  leader-length  for  browse  species.  These 
measurements  are  then  converted  into  percentage  of 


702 


Food  Habita 


Measuring  browse  plant  leader-length. 


utilization  through  previously  developed  tables  or 
regression  equations  (Cook  and  Stubbendieck  1986; 
U.S.  Department  of  the  Interior,  Bureau  of  Land  Man- 
agement 1984). 

Form-class  or  ocular  measurements  are  used 
primarily  with  browse  species.  They  are  subjective, 
requiring  judgments  about  the  degree  of  past  or 
current  browsing  of  individual  browse  plants  based 
on  their  shapes.  These  judgments  are  then  used  to 
infer  the  degree  of  utilization. 

Problems  with  Utilization  Techniques 

Techniques  for  determining  forage  utilization 
are  not  precise,  even  though  many  methods  are  rela- 
tively time-consuming  and  expensive.  Furthermore, 
the  accuracy  of  most  methods  is  limited  because 
of  factors  such  as  regrowth  of  grazed  plants,  loss  of 
forage  from  factors  other  than  grazing,  and  many 
other  causes.  For  a  discussion  of  some  of  these  fac- 
tors see  Martin  (1970),  Cook  and  Stubbendieck 
(1986),  and  Pieper  (1978). 

Because  of  the  high  variances,  many  biologists 
prefer  utilization  measurements  of  only  a  few  plant 
species  known  to  be  key  forages.  Good  estimates  on 
a  few  key  species  are  usually  preferable  to  poor  esti- 
mates on  many,  particularly  if  the  key  species  are 
chosen  carefully  and  judiciously.  See  U.S.  Depart- 
ment of  the  Interior,  Bureau  of  Land  Management 
( 1984)  for  a  discussion  on  choice  of  key  species. 

In  general,  utilization  measurements  may  be 
useful  in  answering  general  questions  such  as  "Are 
perennial  grasses  being  overused  on  summer 
ranges?";  "Is  bitterbrush  being  used  heavily  in  the  fall 
by  mule  deer  (Odocoileus  bemionus)?";  or  "Are  the 
elk  (Cervus  elaphns)  using  grass  or  browse  on  the 
winter  range?"  They  are  not,  however,  very  useful  in 


determining  slight  variations  in  levels  of  utilization. 
Furthermore,  some  questions  such  as  the  last  one 
above  can  be  answered  more  easily  and  cheaply  by 
food  habit  studies  rather  than  utilization  studies.  The 
biologist  must  determine  on  a  case-by-case  basis 
which  type  of  study  and  particular  method  will  be 
most  appropriate  and  efficient  in  terms  of  time  and 
money. 


MEASUREMENT  OF  FOOD  HABITS 

This  section  concentrates  on  techniques  ap- 
plicable to  estimating  herbivore  diets.  However,  par- 
allel techniques  for  carnivores  are  mentioned  as 
appropriate. 

Food  habits  of  wild  animals  can  be  determined 
through  three  basic  methods: 


( 1 )  direct  or  indirect  observation  of  the  animal, 

(2)  stomach  or  rumen  analysis,  and 

(3)  fecal  analysis. 

Direct  or  Indirect  Observation 

Direct  observation  of  animals  has  been  used 
with  varying  success  for  many  years  to  estimate  diets 
of  free-ranging  or  tame  animals  (Neff  1974;  Wallmo 
et  al.  1973;  Riney  1982).  Feeding  on  individual  food 
items  is  quantified  as  "bite  counts"  or  "feeding  min- 
utes," resulting  in  data  which  are  then  converted, 
with  or  without  correction  factors,  to  percentage  in 
the  diet.  It  requires  equipment  no  more  complicated 
than  binoculars  (or  spotting  scopes)  and  a  field  note- 
book. This  method  can  be  extremely  time-consum- 
ing, however,  particularly  if  animals  are  difficult  to 
locate  or  difficult  to  approach  without  flushing  them. 
The  technique  works  best  with  diurnal  or  crepuscu- 
lar animals  that  occupy  relatively  open  habitat  be- 
cause biologists  can  approach  them  close  enough  to 
observe  their  feeding  habits  without  causing  them  to 
leave.  Few  animals  meet  all  these  criteria,  although 
the  technique  has  been  successful  with  animals  like 
bighorn  sheep  (Ovis  sp.)  and  in  National  Parks  and 
other  areas  where  animals  become  accustomed  to 
the  near  presence  of  human  observers. 

Because  of  the  difficulty  in  locating  and  observ- 
ing animals,  some  biologists  have  preferred  to  work 
with  tame,  semi-tame,  or  constrained  animals.  Typi- 
cally, these  animals  are  raised  in  captivity  so  they  are 
accustomed  to  humans.  To  determine  food  habits, 
the  animals  are  taken  to  an  area  and  allowed  to  for- 
age. The  biologist  then  stays  near  the  animal  and 
observes  what  it  eats.  There  are  two  potential  prob- 
lems with  using  tame  animals:  differences  between 
tame  and  wild  animals  and  cost. 


Food  Habits 


703 


Food  habits  of  tame  animals  may  be  different 
from  those  of  wild  animals  due  to  (  1 )  previous  die- 
tary experience;  (2)  nutritional  status  immediately 
before  a  foraging  trial  (e.g.,  if  the  animal  has  a  full 
stomach  or  has  been  starved  before  the  trial);  and 
(3)  choice  of  a  feeding  site/foraging  area  by  the  cap- 
tors rather  than  the  animal.  Some  believe  these  dif- 
ferences can  be  overcome.  However,  use  of  tame 
animals  remains  an  expensive  method  for  measuring 
food  habits  and  will  probably  be  too  expensive 
in  most  management  (as  opposed  to  research) 
situations. 

The  precision  and  accuracy  of  data  from  direct 
observation  of  animals  have  been  questioned  by 
some  biologists.  However,  some  believe  that  direct 
observation  is  not  only  adequate  for  management 
but  is  adequate  for  measuring  food  habits  and  estab- 
lishing a  standard  to  measure  the  accuracy  of  alterna- 
tive techniques  (Gill  et  al.  1983). 

An  alternative  method  of  animal  observation  is 
the  "feeding  site"  examination.  This  technique  is 
similar  in  many  ways  to  the  utilization  methods  pre- 
viously described,  except  the  measurement  obtained 
is  food  habits  rather  than  forage  utilization.  Basically, 
animals  are  located  in  the  field;  then  the  exact  loca- 
tion where  the  animal  has  been  feeding  is  examined 
and  the  plants  or  twigs  grazed  are  tallied.  The  num- 
ber of  plants  of  each  species  that  have  been  grazed  is 
then  converted  into  a  percentage  to  provide  an  esti- 
mate of  the  food  habits  of  the  animal.  Such  feeding 
can  be  relatively  easy  to  observe  in  some  situations, 
such  as  when  animals  are  pawing  through  fresh  snow 
to  eat  plants.  This  can  also  be  done  while  trailing 
animals.  At  other  times,  accurately  determining  what 
the  animal  has  eaten  can  be  extremely  difficult,  if 
not  impossible.  Like  other  observation  methods, 
feeding  site  examinations  can  be  very  time- 
consuming. 

Direct  observation  or  feeding-site  observations 
are  used  primarily  with  herbivores.  However,  Riney 
(1982:132)  suggests  that  direct  observation  is  also 
easy  with  larger  carnivores.  In  North  America,  large 
carnivores  are  rarely  found  in  high  enough  densities 
such  that  direct  observation  would  be  feasible.  Simi- 
lar problems  arise  with  smaller  mammals  as  well  as 
most  bird,  reptile,  and  amphibian  species.  A  notable 
exception  is  diurnal  raptors  at  nest  sites.  In  this  case, 
forages  brought  to  nest  sites  by  parent  birds  can  be 
observed  relatively  easily  and  recorded,  providing  an 
estimate  of  food  habits  during  the  important  nesting 
season. 

Stomach  or  Rumen  Analyses 

Stomach  or  rumen  analyses  have  frequently 
been  used  to  determine  food  habits.  The  technique 
has  been  used  on  virtually  every  major  group  of 
vertebrates  ranging  from  songbirds  and  small  rodents 


to  moose  and  elephants.  Stomach  or  rumen  samples 
are  taken  from  animals  that  have  been  killed,  or  ob- 
tained in  other  ways;  the  contents  are  then  analyzed 
in  the  field  or  sent  to  a  laboratory  for  later  analysis. 
Medin  (1970)  and  Korschgen  (1980)  describe  the 
basic  technique  in  detail,  including  preservation 
of  materials  and  identification  of  food  items.  Food 
habits  can  then  be  determined  and  quantified  by  ( 1 ) 
tabulating  numbers  of  each  food  item,  (2)  tabulating 
frequency  of  occurrence  of  food  items,  (  3  )  measur- 
ing volume,  or  (4)  measuring  weight. 

Microscopic  techniques,  including  the  microhis- 
tological  technique  of  Sparks  and  Malechek  (1968) 
and  the  microscope-point  technique  of  Heady  and 
Van  Dyne  (1965)  have  also  been  used,  primarily 
with  herbivores. 

Stomach  samples  create  major  problems  in  that 
they  normally  require  shooting  or  otherwise  sacrific- 
ing animals.  In  addition  to  being  time-consuming, 
shooting  animals  for  the  sole  purpose  of  studying 
their  food  habits  is  rarely  warranted,  at  least  with  the 
larger  or  more  economically  important  animals  or 
with  any  rare  or  endangered  species. 

Recently,  fistula  techniques  have  been  devel- 
oped to  remove  forage  samples  from  the  rumen  or 
esophagus  of  captive  ungulates  without  killing  them. 
This  allows  multiple  samples  to  be  taken  from  the 
same  animal.  These  techniques  were  developed  pri- 
marily by  range  and  animal  scientists  working  with 
livestock,  but  have  been  used  successfully  with  only 
a  few  species  of  wildlife.  Although  the  technique  is 
useful  in  research,  the  time  and  money  required 
to  raise  and  keep  fistulated  animals  prohibit  this 
technique  from  being  used  to  obtain  information  for 
routine  land  and  wildlife  management. 


Fecal  Analysis 

Fecal  analysis  has  been  used  more  for  estimating 
range  herbivore  food  habits  in  the  past  10  years 
than  any  other  procedure  (Holechek  et  al.  1982a). 
This  happened  in  spite  of  the  fact  that  the  accuracy 
of  the  technique  has  been  seriously  questioned.  The 
procedure  is  similar  to  stomach  analyses  except  that 
a  fecal  sample  rather  than  a  stomach  sample  is 
collected. 

In  the  laboratory,  microscopic  techniques  are 
normally  used  to  sample  herbivores,  whereas  macro- 
scopic techniques  are  used  for  carnivores.  The  tech- 
niques for  identifying  food  items  in  the  laboratory 
are  the  same  ones  used  for  stomach  samples  and, 
therefore,  have  the  same  advantages  and  disadvan- 
tages in  terms  of  time,  cost,  precision,  and  accuracy. 
Typically,  the  management  biologist  need  not  learn 
how  to  perform  the  laboratory  analyses,  since  non- 
profit regional  laboratories  will  perform  the  analyses 


704 


Food  Habits 


at  cost.  As  with  many  types  of  specialized  laboratory 
techniques,  it  is  usually  more  efficient  for  the  man- 
agement biologist  to  pay  for  such  work  than  to  try 
to  do  it  independently. 

Cooperrider  et  al.  (1982)  describe  field  proce- 
dures for  herbivores,  and  Korschgen  ( 1980)  de- 
scribes the  macroscopic  techniques  for  carnivores  as 
well  as  herbivores. 

Even  though  the  fecal  analysis  technique  for  use 
with  herbivores  has  been  criticized  (see  Holechek 
et  al.  1982a;  Gill  et  al.  1983),  the  technique  has 
some  important  advantages.  Holechek  et  al.  (1982a) 
list  the  following  seven  advantages: 

•  It  does  not  interfere  with  the  normal  habits  of 
animals. 

•  It  permits  practically  unlimited  sampling. 

•  It  places  no  restriction  on  animal  movement. 

•  It  has  particular  value  where  animals  range  over 
mixed  communities. 

•  It  is  the  only  feasible  procedure  to  use  when 
studying  secretive  and/or  endangered  species. 

•  It  can  be  used  to  compare  the  diets  of  two  or 
more  animals  at  the  same  time. 

•  Actual  sampling  requires  very  little  equipment. 

In  addition  to  these  important  considerations,  Coo- 
perrider et  al.  (1982)  suggest  that  the  technique 
is  relatively  cost-effective  and  is  a  practical  tech- 
nique for  management  biologists  working  in  govern- 
ment agencies. 

Holechek  et  al.  ( 1982a),  reviewing  the  many 
criticisms  of  the  technique  to  that  date,  compiled  a 
similar  list  of  disadvantages  of  the  fecal  analysis  tech- 
nique for  herbivores  and  concluded  that  inaccuracy 
is  the  greatest  overall  limitation.  More  recently,  Gill 
et  al.  (1983)  reiterated  this  criticism  and  suggested 
that  the  technique  be  critically  evaluated. 

Fecal  analysis  techniques  are  commonly  used 
with  carnivores,  both  large  and  small.  Such  use  has 
generated  far  less  criticism  and  controversy,  al- 
though the  accuracy  of  such  procedures  is  not  well 
known  either.  Before  discussing  accuracy,  I  will  dis- 
cuss the  nature  of  diets  of  wild  animals  since  this  has 
an  important  bearing  on  both  accuracy  and  preci- 
sion, as  well  as  the  way  such  data  are  used. 

VARIABILITY  IN  FOOD  HABITS  OF  WILD 
ANIMALS 

A  common  misconception  of  novice  biologists 
and  lay  persons  is  that  food  habits  of  wild  animals 
are  rigid  or  fixed.  Statements  like  "mule  deer  eat 
bitterbrush  in  winter"  or  "elk  are  grass-eaters"  are 


quite  common.  Many  similar  statements  can  be 
found  in  scientific  literature.  Such  statements  tend  to 
oversimplify  the  real  situation. 

Most  vertebrate  animals  are  opportunistic  in 
their  feeding  habits  and  are  capable  of  eating  and 
surviving  on  a  wide  variety  of  foods.  This  is  in  con- 
trast to  insects,  for  example,  many  of  which  have 
become  adapted  to  feeding  on  only  one  or  a  few 
closely  related  plant  or  animal  species. 

Animals  have  limitations  imposed  by  the  struc- 
ture of  their  digestive  tracts  and  their  ability  to  lo- 
cate, capture,  or  ingest  adequate  quantities  of  certain 
foods.  Coyotes  cannot  survive  on  dry  grass,  and 
horses  do  not  thrive  by  eating  crickets.  Nevertheless, 
a  herbivore  such  as  Rocky  Mountain  mule  deer  has 
been  reported  in  99  separate  studies  to  eat  over  700 
different  plant  species  (Kufeld  et  al.  1973). 

Given  the  opportunistic  nature  of  most  verte- 
brates, particularly  of  most  North  American  herbi- 
vores, and  the  diversity  of  habitats  and 
environmental  conditions  that  these  animals  are  ex- 
posed to  over  their  geographic  ranges,  it  is  not  sur- 
prising that  food  habits  vary  greatly  in  time  and 
space.  Cooperrider  et  al.  ( 1980),  for  example,  have 
demonstrated  that  food  habits  of  one  bighorn  sheep 
herd  vary  significantly  between  seasons,  sites,  and 
years.  Differences  between  regions  are  likely  to  be 
even  greater.  Although  these  sorts  of  phenomena 
have  not  been  analyzed  explicitly  for  many  carni- 
vores, the  diversity  of  the  food  habits  data  reported 
suggests  that  the  same  is  true  for  many  of  these  spe- 
cies. Food  habits  of  animals  are  not  an  attribute  of 
a  species,  but  rather  a  species  in  a  given  locality 
during  a  specific  time.  Extrapolation  of  food  habit 
data  to  other  populations,  localities,  or  times  may  be 
very  misleading. 

On  the  other  hand,  vertebrate  feeding  is  far 
from  random.  Overwhelming  data  suggest  that  ani- 
mals of  the  same  species  have  similar  if  not  identical 
preferences  for  food  items.  Thus  forage  preference  as 
opposed  to  food  habits,  to  a  certain  extent,  is  an 
attribute  of  an  animal  species.  The  prime  determi- 
nants of  food  habits,  what  the  animal  actually  eats, 
are  forage  availability  and  forage  preference. 

Forage  availability  is  quite  different  from  site  to 
site;  and  changes  over  time  from  removal,  phenol- 
ogy, weather,  and  other  factors  account  for  much  of 
the  variation  in  food  habits.  Given  that  forage  prefer- 
ence is  an  attribute  of  an  animal  species  and  that  it 
together  with  forage  availability  are  the  prime  deter- 
minant of  food  habits,  two  important  conclusions 
can  be  drawn.  First,  if  forage  preferences  and  forage 
availability  are  known  or  estimated,  a  biologist 
should  be  able  to  predict,  albeit  very  crudely,  food 
habits  for  a  given  site  or  range.  Second,  if  food  habits 
for  a  species  are  known  for  an  area,  then  food  habits 


Food  Habits 


705 


should  be  the  same  in  areas  with  similar  forage  avail- 
ability. Both  types  of  inferences  require  some  initial 
estimation  of  food  habits,  however. 


ACCURACY  AND  PRECISION  OF  FOOD 
HABIT  MEASUREMENTS 

The  question  of  accuracy  and  precision  of  food 
habit  techniques  continues  to  be  argued,  yet  few 
investigators  give  the  subject  enough  priority  or  ex- 
pend much  effort  in  testing  the  techniques.  How- 
ever, in  fairness,  testing  the  accuracy  of  a  technique 
is  extremely  difficult  without  having  a  "standard" 
technique  that  all  investigators  agree  is  both  accu- 
rate and  precise  and  from  which  other  techniques 
can  be  evaluated.  Even  the  obvious  solution  of  feed- 
ing animals  by  hand  can  introduce  biases,  in  addition 
to  being  costly  and  time-consuming.  Therefore,  there 
are  few  conclusions  about  accuracy  and  precision  of 
given  techniques  that  all  biologists  will  agree  upon. 
The  reader  who  wants  or  needs  to  pursue  the  sub- 
ject further  is  encouraged  to  investigate  the  litera- 
ture. Holechek  et  al.  (1982a,  1984)  provide  a  good 
review  of  techniques  for  herbivores. 

All  three  techniques  (observation,  stomach  or 
rumen  analysis,  and  fecal  analysis)  have  limited  accu- 
racy. Direct  observations  can  be  deceptive  because 
the  number  of  "bites"  or  the  time  spent  feeding  on  a 
given  plant  relative  to  the  total  bites  or  time  may 
not  be  directly  proportional  to  the  percentage  by 
weight  of  a  species  that  is  ingested.  Problems  with 
tame  animals  have  already  been  discussed.  For  fur- 
ther discussion  see  Wallmo  et  al.  (1973)  and  Wallmo 
and  Neff  (1970). 

Stomach  and  fecal  analyses  have  limited  accu- 
racy for  basically  the  same  reasons.  The  items 
counted  or  sampled  from  stomachs  or  feces  may  not 
reflect  the  proportion  by  weight  in  the  diet.  In  all 
three  cases,  correction  factors  have  been  developed 
to  adjust  these  figures.  The  development  and  use 
of  these  correction  factors,  however,  continues  to  be 
more  of  an  art  than  a  science  since  they  are  fre- 
quently not  written  down,  recorded,  or  reported  in 
the  literature.  Thus  it  is  difficult  to  replicate  results. 

Technical  error,  of  course,  is  always  a  concern 
and  can  be  a  problem  with  any  of  these  techniques. 
Field  biologists  have  tried  to  at  least  check  technical 
or  other  types  of  errors  with  stomach  and  fecal  anal- 
yses by  sending  in  replicate  samples,  i.e.,  two  sam- 
ples drawn  from  the  same  stomach,  rumen,  pellet 
group,  etc.  This  is  a  good  recommended  practice; 
however,  the  biologist  should  be  cautious  about 
interpreting  such  information  when  it  is  returned. 
Many  biologists  have  sent  replicate  samples  to  a 
laboratory  and,  after  receiving  results  that  were  not 
"the  same,"  have  rejected  the  technique. 


Determining  if  results  are  "the  same"  requires 
knowledge  of  the  normal  variation  among  samples. 
Whenever  sampling  from  a  "population"  (in  the  sta- 
tistical sense),  results  are  expected  to  be  different. 
Most  food  habit  techniques,  including  microhistolog- 
ical,  involve  some  sort  of  sampling;  two  analyses  of 
the  same  sample  should  not  be  expected  to  be  ex- 
actly the  same.  Depending  on  the  precision  of  the 
technique  and  sampling  error,  results  may  be  quite 
different.  There  are  statistical  procedures  to  deter- 
mine the  probability  that  two  samples  are  really 
"different"  providing  one  has  estimates  of  variance 
or  can  obtain  them.  In  summary,  a  basis  for  deter- 
mining whether  replicate  samples  are  "the  same"  or 
"not  the  same"  should  be  established  before  a  tech- 
nique is  rejected  or  accepted,  based  upon  such 
evidence. 

The  problem  of  accuracy  will  remain  until  suffi- 
cient studies  are  devoted  to  quantifying  the  problem. 
In  the  meantime,  biologists  must  judge  subjectively 
or  intuitively  if  accuracy  of  a  technique  is  within 
reasonable  limits.  Management  biologists  must  also 
judge  if  the  technique  will  be  accepted  or  if  the 
results  will  be  deemed  credible  by  the  managers, 
agencies,  courts,  etc.,  that  will  be  using,  reviewing, 
or  making  decisions  based  on  such  information.  The 
accuracy  of  the  fecal  analysis  technique  for  herbi- 
vores has  been  criticized  more  than  others,  yet  many 
of  the  most  vocal  critics  have  continued  to  use  it 
and  publish  papers  based  on  its  use.  As  mentioned 
earlier,  it  is  also  the  most  widely  used  method,  and 
results  from  such  analyses  have  stood  up  in  court 
hearings. 


Knowledge  of  the  precision  of  most  techniques 
is  also  quite  limited,  although  it  is  easier  to  quantify 
and  evaluate.  The  subject  is  complicated  by  consid- 
erations of  the  level  of  precision  needed.  More  specif- 
ically, the  basic  sampling  unit  is  usually  one  stomach 
sample,  one  pellet  group,  or  one  day  or  session  of 
bite  counts.  All  of  these  more  or  less  estimate  the 
diet  of  an  animal  for  1  or  2  days.  This  daily  diet  can 
usually  be  estimated  with  good  precision;  however, 
the  "daily  diet"  of  one  animal  is  not  usually  of  much 
interest  to  management. 

An  animal's  diet  may  change  drastically  from  day 
to  day,  drainage  to  drainage,  animal  to  animal,  etc. 
Therefore,  what  is  usually  of  interest  is  the  seasonal 
or  monthly  food  habits  of  a  herd  or  some  other  logi- 
cal grouping  of  individuals.  Reasonable  precision  in 
estimating  this  can  usually  be  obtained  if  the  biolo- 
gist can  obtain  enough  samples.  With  fecal  samples, 
this  is  usually  not  a  problem,  but  with  techniques 
that  require  killing  or  fistulating  animals,  it  may  be 
difficult  or  even  impossible.  Dearden  et  al.  (1975), 
Cooperrider  et  al.  (1982),  and  Holechek  and  Vavra 
(1983)  discuss  precision  of  microhistological  and 
fistula  estimates  of  herbivore  food  habits. 


706 


Food  Habits 


In  deciding  if  accuracy  and  precision  are  within 
acceptable  limits,  the  biologist  must  keep  clearly  in 
mind  the  reason  data  are  being  collected  and  the 
purpose  of  these  data.  Precision  and  accuracy  re- 
quired or  feasible  for  management  generally  are  not 
as  stringent  as  those  required  for  research.  Similarly, 
precision  and  accuracy  required  to  obtain  prelimi- 
nary insight  into  range  utilization  and  foraging  dy- 
namics may  be  much  less  than  for  a  statistically 
sound  monitoring  program.  For  example,  a  few  fecal 
samples  of  deer  and  elk  may  be  collected  and  ana- 
lyzed to  select  plant  species  for  intensive  monitoring 
of  utilization  on  winter  range.  Insight  gained  from 
limited,  slightly  inaccurate  data  may  be  quite  helpful 
in  focusing  a  monitoring  project  on  relevant 
problems. 


USE  AND  INTERPRETATION  OF  FOOD 
HABITS  AND  FORAGE  UTILIZATION  DATA 

Because  food  habits  or  forage  utilization  data  are 
rarely  useful  by  themselves,  the  ways  such  data  are 
used  (and  misused)  in  monitoring  programs  and 
in  influencing  management  decisions  are  discussed 
below. 

Forage  utilization  data  are  principally  used  to 
determine  whether  key  plant  species  are  being  over- 
used. If  such  a  determination  is  made  and  more  than 
one  major  herbivore  is  present  on  the  range,  then 
further  study  will  be  required  to  determine  relative 
use  by  various  species.  This  can  be  done  with  food 
habit  studies  or,  if  the  animals  are  present  at  differ- 
ent times  or  seasons,  it  can  be  done  with  more  de- 
tailed utilization  studies. 


quality  of  the  forage  and  the  nutritional  require- 
ments of  the  animal.  In  some  cases,  subjective 
knowledge  may  be  adequate.  For  example,  if  winter 
food  habits  indicate  bighorn  sheep  are  eating  primar- 
ily winterfat  and  fringed  sagebrush,  and  these  are 
both  known  to  be  good  nutritious  winter  forages, 
then  the  biologist  can  reasonably  infer  that  the  diet 
is  adequate.  In  other  cases,  measurements  of  nutri- 
tional content  of  the  plants  may  have  to  be  obtained. 
The  whole  area  of  nutritional  analysis  is  beyond  the 
scope  of  this  chapter;  for  herbivores,  a  reader  inter- 
ested in  pursuing  the  subject  is  referred  to  several 
good  books  and  publications  on  the  subject  re- 
viewed and  listed  in  Zarn  ( 1981 )  and  Holechek  et  al. 
(1982b). 

Whether  a  diet  is  good  or  adequate  does  not 
mean  that  range  conditions  are  good.  Animals  could 
be  receiving  an  adequate  diet  while  severely  overus- 
ing many  plant  species  on  the  range.  To  determine 
this,  the  biologist  needs  information  on  animal  densi- 
ties, forage  production,  or  forage  utilization. 

Similarly,  food  habit  data  alone  are  not  adequate 
to  determine  competition  for  forage.  Forage  compe- 
tition refers  to  common  use  of  forage  by  two  or 
more  animal  species,  and  dual  use  of  forages  that  are 
limiting  factors  for  one  or  more  of  the  animal  spe- 
cies. Food  habits  of  sympatric  species  are  commonly 
compared  by  investigating  dietary  overlap  or  similar- 
ity. Dietary  overlap  is  quantified  using  similarity  in- 
dexes such  as  shown  in  Table  2.  High  dietary 
overlap  is  often  used  as  evidence  of  severe  competi- 
tion for  forage,  and  similarly  low  overlap  is  used  to 
suggest  absence  of  forage  competition.  These  infer- 
ences may  be  unwarranted  depending  on  additional 
data. 


In  some  cases,  biologists  attempt  to  use  utiliza- 
tion measurements  to  make  inferences  about  food 
habits.  This  procedure  can  be  awkward  and  inaccu- 
rate for  several  reasons,  even  where  only  one  princi- 
pal herbivore  is  present  on  the  range  (Martin  1970; 
Cook  and  Stoddart  1953).  For  example,  annual  forbs 
may  constitute  a  significant  portion  of  the  herbi- 
vores' diet,  but  if  the  forbs  have  already  dried  up  and 
disappeared  at  the  time  utilization  studies  are  con- 
ducted, such  studies  will  not  detect  them  and  infer- 
ences about  food  habits  will  be  inaccurate.  If  food 
habits  need  to  be  determined,  a  technique  for  meas- 
uring food  habits  should  be  chosen. 

Food  habits  data  have  two  principal  uses:  ( 1 ) 
determination  of  the  adequacy  of  the  habitat  to  sup- 
ply food  of  sufficient  quality  and  (  2  )  analysis  of  com- 
petition for  forage  between  two  or  more  sympatric 
species. 

Biologists  cannot  determine  the  adequacy  of  the 
forage  supply  from  knowledge  of  food  habits  alone. 
They  must  know  something  about  the  nutritional 


High  overlap  in  forage  use  in  a  noncompetitive 
situation  can  occur  on  a  range  with  an  abundance  of 
forage  preferred  by  two  species.  Conversely,  low 
overlap  with  severe  competition  can  occur  where 
one  animal  species  uses  preferred  and  limited  for- 
ages in  an  area  before  a  second  animal  species  ar- 
rives. This  situation,  sequential  competition,  can 
occur  when  cattle  occupy  a  range  in  summer  that  is 
used  by  wild  ungulates  in  winter.  If  the  cattle  re- 
move the  preferred,  presumably  highest  quality,  for- 
age for  the  wild  ungulates,  overlap  in  food  habits  can 
be  low  while  competition  for  forage,  extreme. 

Both  of  these  uses  of  food  habit  data  point  to 
the  value  of  understanding  forage  preference.  A  biol- 
ogist can,  without  great  risk,  infer  that  forage  prefer- 
ence is  well -correlated  with  nutritional  value.  Those 
species  that  are  preferred  generally  contain  the  high- 
est levels  of  basic  nutrients  such  as  digestible  energy 
or  digestible  protein.  If  an  animal  is  primarily  eating 
forages  that  are  known  to  be  highly  preferred  by  that 
species,  a  biologist  may  infer  that  the  diet  is  good. 
Similarly,  if  two  animal  species  are  present  on  the 


Food  Habits 


707 


Table  2.     Dietary  overlap  between  bighorn  sheep  and  cattle,  Trickle  Mountain,  Colorado,  1978  (from  Bailey 
and  Cooperrider  1982). 


Percent  of 

Vegetation 

Percent  of 
Forage  in 
Cattle  Diet 

Forage  in 
Bighorn  Sheep 
Diet 

Dietary 
Overlap 

GRASSES  AND  GRASSLIKE 

PLANTS 

Fescue 

17 

7 

7 

(Festuca  sp.) 

Muhly 
(Muhlenbergia  sp.) 

14 

10 

10 

Sedge 
(Cyperaceal) 

16 

5 

5 

Wheatgrass 
(Agropyron  sp.) 

21 

0 

0 

Other 

19 

11 

11 

FORBS 

4 

6 

4 

BROWSE 

Sagebrush 
{Artemisia  sp.) 

2 

25 

2 

Saltbrush 

0 

11 

0 

(Atriplex  sp.) 

Other 

7 

14 

7 

Note:  Dietary  overlap  determined  by  taking  the  lesser  of  the  two  figures  in  each  row  as  circled;  total  dietary  overlap  is  46%. 


708 


Food  Habits 


same  range  and  are  using  large  quantities  of  a  forage 
item  that  is  in  short  supply  and  that  is  a  preferred 
forage  for  at  least  one  animal  species,  then  a  biolo- 
gist may  infer  that  forage  competition  is  taking  place. 
A  thorough  determination  that  the  species  are  really 
competing  requires  more  information,  particularly 
evidence  that  populations  of  one  or  the  other  spe- 
cies are  limited  by  availability  of  the  preferred  forage 
species. 

Determination  of  forage  preference  requires 
information  on  both  food  habits  and  forage  availabil- 
ity. This  information  is  used  in  a  variety  of  ways  to 
calculate  preference  indexes — the  simplest  of  which, 
and  the  only  one  with  a  straightforward  interpreta- 
tion, is  the  ratio  of  a  given  forage  in  the  diet  to  the 
ratio  of  that  same  forage  in  the  habitat.  Other  prefer- 
ence indexes  include  information  on  such  factors  as 
frequency  on  the  range  and  frequency  in  the  diet,  an 
interpretation  that  is  unclear  at  best  (Chamrad  1979; 
Krueger  1972).  For  a  thorough  review  of  forage 
preference  of  ungulates,  see  Skiles  (1984). 

Food  habits,  as  described  earlier,  can  be  consid- 
ered to  be  primarily  a  function  of  forage  preference 
and  forage  availability.  Spatial  arrangement  of  forage 
is  of  course  important,  but  is  difficult  to  quantify. 
All  forage  availability  techniques,  therefore,  treat 
available  forage  as  if  it  is  equally  available.  Neverthe- 
less, if  forage  preferences  are  known,  then  a  biologist 
should  be  able  to  predict,  albeit  very  crudely,  the 
food  habits  of  an  animal  species  in  that  area.  Many 
biologists  are  capable  of  doing  this,  through  knowl- 
edge of  forage  preference  and  forage  availability 


gained  by  experience  and  undocumented  observa- 
tions. However,  attempts  to  use  forage  preference  in- 
dexes and  forage  availability  information  in 
quantitative  models  to  predict  actual  diets  have  not 
been  very  successful,  at  least  with  herbivores 
(Loehle  and  Rittenhouse  1982).  Hopefully,  future 
work  in  this  area  will  result  in  preference  indexes 
and  quantitative  models  that  will  allow  the  biologist 
to  predict  diets.  Developing  more  powerful  and 
useful  predictive  models  will  probably  depend  on 
development  of  methods  to  quantify  the  spatial 
arrangement  of  forage  in  a  way  that  can  be  measured 
simply  and  efficiently,  yet  is  still  useful  for  predic- 
tion. Until  such  models  are  available,  biologists  will 
need  to  continue  to  measure  food  habits  directly. 


SUMMARY 

Measurement  of  forage  utilization  and  food  hab- 
its and  the  use  of  such  data  in  monitoring  programs 
are  often  required.  However,  forage  selection  and 
utilization  by  wild  animals  is  a  complex  process,  and 
most  measurement  techniques  have  limited  accuracy 
and  precision.  In  addition,  many  techniques  require 
time,  money,  and  equipment  that  are  generally  not 
available  in  some  agencies  or  cannot  be  justified 
by  the  problems  being  addressed.  Nevertheless, 
measurements  of  food  habits  and  forage  require- 
ments will  continue  to  be  needed  for  special  situa- 
tions. Biologists  should  carefully  choose  appropriate 
techniques,  considering  the  type  of  data  needed, 
the  precision  and  accuracy  required,  the  acceptabil- 
ity of  these  data  to  those  who  must  use  it,  and  the 
relative  cost  in  terms  of  money  and  personnel. 


Food  Habits 


709 


LITERATURE  CITED 

CHAMRAD,  AD.  1979-  Preference  ratings  aid  in  food  habit 
evaluations  and  interpretations.  Pages  273-276  in 
Proc.  First  Welder  Wildl.  Foundation  Symp. 

COOK,  C.W.  AND  L.A.  STODDART.  1953  The  quandary  of 
utilization  and  preference.  J.  Range  Manage. 
6:329-335. 

AND  J.  STUBBENDIECK  1986.  Methods  of  measur- 
ing herbage  and  browse  utilization.  Pages  1 20- 121 
in  Cook,  C.W.  and  J.  Stubbendieck,  eds.  Range  Re- 
search: Basic  Problems  and  Techniques.  Soc.  Range 
Manage.,  Denver,  CO.  317pp. 

COOPERRIDER,  AY.,  J.A.  BAILEY,  AND  R.M.  HANSEN. 
1982.  Cost  efficient  methods  of  estimating  ungulate 
food  habits  by  fecal  analysis.  Pages  399-406  in  Arid 
Land  Resource  Inventories:  Developing  Cost-Efficient 
Methods.  U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech.  Rep. 
WO-28.  620pp. 

,  S.A.  MCCOLLOUGH,  AND  J.A.  BAILEY.  1980.  Varia- 
tion in  bighorn  sheep  food  habits  as  measured  by 
fecal  analysis.  Pages  29-41  in  Proc.  Biennial  Symp. 
Northern  Wild  Sheep  and  Goat  Council,  Salmon,  ID. 
668pp. 

DEARDEN,  B.L.,  RE.  PEGAU,  AND  R.M.  HANSEN.  1975. 
Precision  of  microhistological  estimates  of  ruminant 
food  habits.  J.  Wildl.  Manage.  39:402-407. 

GILL,  R.B.,  L.H.  CARPENTER,  R.M.  BARTMANN,  D.L. 

BAKER,  AND  G.G.  SCHOONVELD.  1983.  Fecal  analy- 
sis to  estimate  mule  deer  diets.  J.  Wildl.  Manage. 
47:902-915. 

GULLION,  G.W.  1966.  A  viewpoint  concerning  the  signifi- 
cance of  studies  of  game  bird  food  habits.  Condor 
68:372-376. 

HEADY,  HE  AND  G.M.  VAN  DYNE.  1965.  Prediction  of 
weight  composition  from  point  samples  on  clipped 
herbage.  J.  Range  Manage.  18:144-149- 

HOLECHEK,  J.L.  AND  M.  VAVRA.  1983.  Fistula  sample 
numbers  required  to  determine  cattle  diets  on  forest 
and  grassland  ranges.  J.  Range  Manage.  36:323-326. 

, ,  AND  R.D.  PIEPER.  1982a.  Botanical  com- 
position determination  of  range  herbivore  diets:  A 
review.  J.  Range  Manage.  35:309-315. 

, ,  AND .  1982b.  Methods  for  deter- 
mining the  nutritive  quality  of  range  ruminant  diets:  A 
review.  J.  An.  Sci.  54:363-376. 

, ,  AND .  1984.  Methods  for  determin- 


ing the  botanical  composition,  similarity  and  overlap 
of  range  herbivore  diets.  Pages  425-471  in  Develop- 
ing Strategies  for  Rangeland  Management.  Westview 
Press,  Boulder,  CO.  2022pp. 
KORSCHGEN,  L.J.  1980.  Procedures  for  food-habits  anal- 
yses. Pages  113-127  in  Schemnitz,  S.D.,  ed.  Wildl. 


Manage.  Techniques  Manual,  4th  Ed.,  The  Wildl.  Soc, 
Washington,  DC.  686pp. 

KRUEGER,  W.C.  1972.  Evaluating  animal  forage  prefer- 
ence. J.  Range  Manage.  25:471-475. 

KUFELD,  R.C.,  O.C  WALLMO,  AND  C.  FEDDEMA.  1973. 
Foods  of  the  Rocky  Mountain  mule  deer.  U.S.  Dep. 
Agric,  For.  Serv.,  Res.  Pap.  RM-111.  31pp. 

LOEHLE,  C.  AND  L.R.  RITTENHOUSE.  1982.  An  analysis  of 
forage  preference  indices.  J.  Range  Manage. 
35:316-319. 

MARTIN,  S.C.  1970.  Relating  vegetation  measurements  to 
forage  consumption  by  animals.  Pages  93-100  in 
Range  and  Wildlife  Habitat  Evaluation — A  Research 
Symposium.  U.S.  Dep.  Agric,  For.  Serv.,  Misc.  Publ. 
1147.  220pp. 

MEDIN,  D.E.  1970.  Stomach  content  analyses:  Collections 
from  wild  herbivores  and  birds.  Pages  133-145  in 
Range  and  Wildlife  Habitat  Evaluation — A  Research 
Symposium.  U.S.  Dep.  Agric,  For.  Serv.,  Misc.  Publ. 
1147.  220pp. 

NEFF,  D.J.  1974.  Forage  preferences  of  trained  mule  deer 
on  the  Beaver  Creek  watershed.  Arizona  Game  and 
Fish  Dep.,  Spec.  Rep.  4.  6lpp. 

PETRIDES,  G.A.  1975.  Principal  foods  versus  preferred 
foods  and  their  relations  to  stocking  rate  and  range 
condition.  Biol.  Conserv.  7: 161-169. 

PIEPER,  R.D.  1978.  Methods  for  measuring  utilization. 

Pages  123-147  in  R.D.  Pieper,  ed.  Measurement  Tech- 
niques for  Herbaceous  and  Shrubby  Vegetation.  Dep. 
Animal  and  Range  Science,  New  Mexico  State  Univ., 
Las  Cruces.  147pp. 

RINEY,  T.  1982.  Study  and  management  of  large  mammals. 
John  Wiley  and  Sons,  New  York,  NY.  552pp. 

SKILES,  J.W.  1984.  A  review  of  animal  preference.  Pages 
153-213  in  Developing  Strategies  for  Rangeland  Man- 
agement. Westview  Press,  Boulder,  CO.  2022pp. 

SPARKS,  DR.  AND  J.C.  MALECHEK  1968.  Estimating  per- 
centage dry  weight  in  diets  using  a  microscopic  tech- 
nique. J.  Range  Manage.  21:264-265. 

U.S.  DEPARTMENT  OF  THE  INTERIOR,  BUREAU  OF 

LAND  MANAGEMENT.  1984.  Rangeland  Monitoring— 
Utilization  studies.  U.S.  Dep.  Inter.,  Bur.  Land  Manage., 
Tech.  Ref.  4400-3.  105pp. 

WALLMO,  O.C,  R.B.  GILL,  L.H.  CARPENTER,  AND  D.W. 
REICHERT.  1973-  Accuracy  of  field  estimates  of  deer 
food  habits.  J.  Wildl.  Manage.  37:556-562. 

AND  D.J.  NEFF.  1970.  Direct  observation  of  tamed 

deer  to  measure  their  consumption  of  natural  forage. 
Pages  105-109  in  Range  and  Wildlife  Habitat  Evalua- 
tion— A  Research  Symposium.  U.S.  Dep.  Agric,  For. 
Serv.,  Misc.  Publ.  1147.  220pp. 

ZARN,  M.  1981.  Wild  ungulate  forage  requirements — A 
review.  U.S.  Dep.  Inter.,  Bur.  Land  Manage.  103pp. 


710 


Food  Habits 


35 

WEATHER  AND 
CLIMATE 

James  A.  Bailey 


Department  of  Fishery  and  Wildlife  Biology 
Colorado  State  University 
Fort  Collins,  CO  80523 


"Climate  plays  an  important  part  in  determining  the 
average  numbers  of  species,  and  periodical  seasons 
of  extreme  drought  or  cold  I  believe  to  be  the  most 
effective  of  all  checks." 

Charles  Darwin,  from  The  Origin  of  the 


Species 


"Climate  is  what  we  expect,  weather  is  what  we 
get." 

Robert  Heinlein 


Editor's  Note:  Weather  is  an  extremely  important 
factor  to  consider  in  inventories  or  monitoring 
studies.  In  the  western  U.S.,  extreme  fluctuations  in 
weather  are  quite  normal.  Weather  affects  wildlife 
directly  and  indirectly:  it  can  alter  population  size 
or  cause  changes  in  animal  behavior  that  will  bias 
study  results. 

This  chapter  identifies  weather  factors  and  describes 
how  they  are  used  to  enhance  the  precision  of  stud- 
ies and  promote  early  detection  of  trends  in  wild- 
life populations  and  habitats. 


INTRODUCTION 

Weather  affects  wild  animals  directly  by  causing 
mortality,  or  by  influencing  behavior  and  physiology 
in  ways  that  influence  reproduction  or  mortality. 
Weather  also  affects  wildlife  indirectly  by  altering 
the  production  or  availability  of  habitat  resources 
and  by  influencing  competitors,  predators,  and  dis- 
ease organisms.  Consequently,  much  year-to-year 
variation  in  wildlife  abundance  and  in  the  productiv- 
ity or  condition  of  wildlife  and  wildlife  habitat  is  due 
to  stochastic  (random)  variation  in  weather. 

In  addition,  weather  may  affect  the  observability 
or  activity  of  wildlife  and  thus  may  affect  data  ob- 
tained for  wildlife  monitoring.  The  objective  of  mon- 
itoring is  to  detect  trends  of  improvement  or  decline 
in  wildlife  or  in  wildlife  habitat.  However,  the  sto- 
chastic direct  and  indirect  effects  of  weather  on 
wildlife,  and  the  stochastic  effects  of  weather  on 
monitoring  data,  will  reduce  precision  of  the  moni- 
toring system  and  may  delay  detection  of  directional 
trends  in  wildlife  populations  or  in  habitat. 

The  objectives  of  this  chapter  are  to  describe 
some  effects  of  weather  on  wildlife  and  on  monitor- 
ing data,  and  to  promote  the  use  of  weather  data  as  a 
covariant  in  analyzing  wildlife-monitoring  data.  Use 
of  covariant  analysis  can  enhance  precision  and  early 
detection  of  directional  trends  in  wildlife  popula- 
tions and  habitats. 


WEATHER  AND  WILDLIFE 

Direct  and  Indirect  Effects  of 
Weather  on  Wildlife 

A  variety  of  direct  and  indirect  impacts  of 
weather  on  wild  vertebrates  has  been  reported  (Ta- 
ble 1 ).  The  smaller  species,  in  particular,  may  die 
from  acute  exposure  to  severe  weather.  Mortality  of 
larger  species  during  severe  weather  may  result  from 
a  combination  of  direct  exposure  of  the  animals  or 
indirect  effects  on  the  availability  of  habitat  re- 
sources. Drought  or  severe  runoff  events  may  drasti- 
cally alter  current  velocity,  turbidity,  and  amount  of 


Weather  and  Climate 


711 


Table  1.  Some  reported  effects  of  weather  upon  wildlife  and  on  wildlife-monitoring  data. 


Mammals 

Birds 

Amphibians,  Reptiles 

Fish 

1.  Direct  effects: 

Jacobs  and  Dixon 

Yocum  1950 

Larimore  1959 

2       1.  Acute  mortality 

1981 

Jones  1952 

Smith  1960 

Jehl  and  Hussell  1966 

Stout  and  Cornweil 

1976 
Graber  and  Graber 

1979 
Stout  and  Cornweil 

1976 
Graber  and  Graber 

1979 

Jojhn  1964 
Tramer  1977 
Matthews  and  Styron 
1981 

2.  Exposure  with 

Martmka  1967 

Nehring  and 

habitat  loss 

Meslow  and  Keith 

1971 
Connolly  1981 

Anderson  1984 

3.  Effects  on 

Wight  and  Conaway 

Hammond  and 

Starrett  1951 

reproduction 

1961 
Picton  1984 

Johnson  1984 

Gagnark  and  Bakkala 

1960 
John  1963 
Crecco  and  Savoy 

1984 

4.  Effects  on 

Lovaas  1958 

Graber  and  Graber 

habitat 

Richens  1967 

1979 

selection 

5  Effects  on 

Gilbert  et  al.  1970 

Rutherford  1970 

Obert  1976 

Stehr  and  Branson 

movement 

Bruns  1977 

1938 

patterns 

Starrett  1951 
Rinne  1975 
May  et  al.  1976 
Ottaway  and  Clark 
1981 

II.  Indirect  effects: 

Smoliak  1956 

Gullion  1960 

Hoddenbach  and 

1.  Upon  habitat 

Gallizioli  1965 

Francis  1970 

Turner  1968 

Shaw  1965 

Brown  and  Smith 

Turner  et  al.  1969, 

Beatley  1969,  1976 

1980 

1976 

Day  1971 

Medica  et  al.  1975 

Shiflet  and  Dietz  1974 

Whitford  and  Creusere 

Turkowski  and  Vahle 

1977 

1977 

Nixon  et  al.  1980 

Stephenson  and 

Brown  1980 

Mueggler  1983 

2.  Upon 

Longhurst  and 

Vaught  et  al.  1967 

competition 

Douglas  1953 

Graber  and  Graber 

or  disease 

1979 

712 


Weather  and  Climate 


Table  1.  Some  reported  effects  of  weather  upon  wildlife  and  on  wildlife-monitoring  data  (concluded). 


Mammals 

Birds 

Amphibians,  Reptiles 

Fish 

III.  Effects  on 

monitoring 

data: 

Lovaas  1958 

Diem  and  Lu  1960 

Obert  1976 

1 .  Changed 

location 

Gilbert  et  al.  1970 

Rutherford  1970 

affects 

Graber  and  Graber 

duration 

1979 
Best  1981 

2.  Persistence  or 

Wallmo  et  al.  1962 

Emlen  and  DeJong 

detectability  of 

1981 

sign 

3.  Activity  and 

Newman  1959 

Gullion  1966 

Lillywhite  1982 

detectability  of 

Morse  and  Balser 

Robel  et  al.  1969 

animals 

1961 
Progulske  and  Duerre 

1964 
LeResche  and 

Rausche  1974 

Brown  1971 
LaPerriere  and 

Haugen  1972 
Porter  1973 
Morrison  and  Slack 

1977 
Shields  1979 
Conner  and  Dickson 

1980 
Bock  and  Root  1981 
Dawson  1981 
Robbins  1971 

4.  Trap  results 

Bailey  1969 
Phillips  et  al.  1972 

Raveling  1966 

Vogt  and  Hine  1982 

Kirkland  1962 
May  et  al.  1976 
Hubert  and  Schmitt 
1982  a  b 

5.  Observer 

Bock  and  Root  1981 

diligence 

Dawson  1981 

suitable  fish  habitat  in  streams  and  rivers,  often  caus- 
ing acute  mortality  or  downstream  transport  of  fish. 
Weather  may  affect  animal  behavior  in  ways  that  in- 
fluence reproduction,  habitat  selection,  or  movement 
patterns  and  distribution.  Effects  of  weather  on  ani- 
mal condition  may  persist  for  several  months  after  a 
weather  event  (Clutton-Brock  and  Albon  1983). 

Impacts  of  weather  on  the  production  or  availa- 
bility of  habitat  resources,  especially  forage,  have 
been  described.  In  arid  areas,  great  variation  in  pre- 
cipitation, and  consequently  in  the  quality  and  abun- 
dance of  forage,  has  been  correlated  with  reproduc- 
tive success  of  mammals,  birds,  and  reptiles.  Weather 
may  also  influence  wildlife  indirectly  by  altering 
competition  or  by  influencing  exposure  to  disease. 

The  direct  and  indirect  effects  of  weather  on 
wildlife  tend  to  be  more  acute  (capable  of  quick  ef- 
fect) and  severe  for  small  animals  than  for  large  ani- 
mals, and  for  populations  living  near  the  edges  of 
their  geographic  ranges  (Siivonen  1956;  Birch  1958; 
Graber  and  Graber  1979). 


Winter-stressed  mule  deer  (when  the  ears  are  collapsed, 
the  animal  cannot  be  saved). 


Weather  and  Climate 


713 


Effects  of  Weather  on  Wildlife  Data 

Numerous  reports  exist  on  weather  conditions 
influencing  the  results  of  wildlife  monitoring  activi- 
ties (Table  1 ).  Weather  may  influence  the  concentra- 
tion of  wildlife  and  thus  the  probability  that  animals 
will  be  detected  and  counted.  Weather  may  alter  the 
locations  of  animals  among  habitats  or  among  re- 
gions of  the  geographic  range.  Consequently,  animals 
may  move  into  habitats  where  they  are  less  detecta- 
ble or  into  habitats  or  regions  away  from  the  areas 
being  monitored. 

Weather  factors  have  modified  the  persistence 
of  wildlife  sign  and  the  transmission  of  sounds  used 
in  census  methods.  Activity  patterns  of  animals  have 
been  altered  in  ways  that  influence  detection  and  ob- 
servability (Figure  1 ).  Success  of  trapping  animals 
and  even  sex-age  structures  of  trapped  samples  of 
animals  may  be  determined  by  weather.  Wildlife  ob- 
servers may  become  less  diligent  during  poor 
weather. 


In  response  to  these  sources  of  bias,  ornitholo- 
gists have  stressed  a  need  to  schedule  bird  surveys 
so  that  data  are  collected  only  during  suitable 
weather.  However,  this  option  is  not  available  in  all 
wildlife  monitoring.  A  few  authors  (Robel  et  al. 
1969;  Best  1981;  Dawson  1981 )  have  suggested  that 
"adjustment  factors"  might  be  developed  to  account 
for  weather  bias.  This  approach  is  proposed  in  this 
chapter. 


40- 

o 
EC 

30  - 

..••' 

0) 

a 

-1^<^^         ' ' 

•o 
m 

20 

.  •   *   ' 

c 

CO 

a 

10 

I 

I           I           I            I 

f40 

45 

I           I           I           I 
50       55       60      65       70       75 

Temperature  (Fahrenheit) 

80° 

Figure  1.  Effect  of  air  temperature  1/2  hour  before 
sunrise  on  numbers  of  cardinals  {Cardinalis 
cardinalis)  counted  on  standard  breeding-bird 
survey  routes.  Dashed  lines  indicate  95% 
confidence  limits  for  data  in  each  5°-interval  set 
(  after  Robbins  1981 ). 


WEATHER  AND  WILDLIFE  MONITORING 

Wildlife  monitoring  produces  time-series  data. 
These  data  provide  information  on  trends  in  popula- 
tion size,  density,  or  distribution;  in  population  qual- 
ity; or  in  habitat  condition  (Hocutt  and  Stauffer 
1980;  Worf  1980).  Trends  are  usually  measured  over 
several  years.  Whereas  the  objective  of  wildlife  mon- 
itoring is  to  detect  persisting  trends  in  wildlife  or 
habitats,  weather  causes  temporary  changes  in  wild- 
life or  habitats,  or  temporary  bias  in  the  monitoring 
data,  that  vary  stochastically  (improvement  or  de- 
cline) from  year  to  year. 

These  temporary  effects  of  weather  may  obscure 
long-term  trends,  especially  if  a  biologist  hopes  to 
detect  trends  from  a  relatively  short  time-series  of 
wildlife  or  habitat  data.  For  example,  1  or  2  years  of 
favorable  weather  may  enhance  a  wildlife  population 
even  though  the  long-term  trend  in  habitat  condition 
or  population  is  downward.  Consequent  delay  in  de- 
tecting the  downward  trend  may  result  in  several 
years'  delay  in  developing  a  management  response. 

This  problem  can  be  diminished  if  effects  of  one 
or  a  few  important  weather  factors  on  wildlife  can 
be  determined  and  measured  during  the  early  years 
of  monitoring.  Once  the  wildlife -weather  relation- 
ships have  been  quantified,  the  temporary  effects  of 
weather  can  be  isolated  from  long-term  trends  in 
time-series  data.  For  example,  although  the  indicated 
trend  in  wildlife  or  habitat  condition  for  the  most  re- 
cent year  or  two  may  not  be  downward,  the  recent 
conditions  may  be  poor  when  compared  with  only 


714 


Weather  and  Climate 


those  previous  years  having  similar  weather 
conditions. 

In  the  same  manner,  positive  responses  of  wild- 
life or  habitat  to  management  programs  can  be  de- 
tected more  quickly  if  year-to-year  effects  of  variable 
weather  can  be  identified. 


Selecting  the  Key  Weather  Factor 

There  are  no  rules  to  follow  in  selecting  key 
weather  factors  for  use  in  data  analysis.  Seven  sug- 
gestions are- 

( 1 )  Use  local  or  regional  "conventional  wisdom." 

( 2 )  Consider  those  weather  factors  known  to  affect 
production  or  availability  of  limiting  habitat 
resources  . 

(  3 )  Remember  that  autumn  populations  of  species 
with  high  biotic  potentials  may  be  determined 
primarily  by  spring  and  summer  weather  affect- 
ing reproductive  success,  rather  than  by  winter 
weather  affecting  overwinter  survival  and 
spring-breeding  densities. 

(4)  Consider  weather  factors  known  to  bias  the  dis- 
tribution or  observability  of  animals  and  there- 
fore the  accuracy  of  the  monitoring  system. 

(5)  Consider  the  amount  of  seasonal  precipitation 
in  arid  areas  as  a  key  factor. 

(6)  Consider  depth,  persistence,  and  quality  of 
snow  in  northern  and  alpine  areas  as  key 
factors. 

(7)  Note  that  plants  often  respond  to  cumulative 
weather  effects,  whereas  small  wildlife  at  the 
edges  of  their  geographic  ranges  are  most  apt  to 
be  directly  affected  by  acute  weather 
phenomena. 

Key  factors  will  vary  among  species,  regions, 
and  even  sites  (Mueggler  1971,  1983),  and  among 
monitoring  systems.  Selecting  the  key  factor  will  be 
a  local  decision. 

Conventional  Wisdom.  Much  local  and  regional 
"conventional  wisdom"  exists  regarding  effects  of 
weather  on  many  species  of  wildlife.  These  widely 
accepted  but  often  unquantified  impacts  of  weather 
on  wildlife  are  frequently  used  to  determine  the  tim- 
ing of  yearly  wildlife  studies  so  as  to  gather  data  un- 
der similar  weather  conditions  each  year.  When 
these  attempts  to  standardize  weather  conditions  fail, 
the  assumed  weather  effects  are  used  to  subjectively 
interpret  the  resulting  trends  in  wildlife  data.  For  ex- 
ample, in  northern  regions,  population  indices  based 
on  "trend  counts"  of  ungulates  along  fixed  routes 
(aerial  or  ground  counts)  are  often  obtained  during 
that  part  of  each  winter  when  snow  is  deepest  and 
most  widespread.  When  this  strategy  is  successful, 
animals  are  presumed  to  be  equally  concentrated 


and  therefore  equally  distributed  for  being  counted 
each  year.  When  the  strategy  fails,  as  in  a  winter 
without  much  snow,  trends  in  the  counts  are  usually 
presumed  to  be  due  largely  to  the  weather  effect 
and  are  therefore  ignored.  Thus,  failure  to  quantify 
the  weather  effect  results  in  complete  loss  of  1 
year's  data. 

Limiting  Habitat  Resources.  Each  vertebrate  spe- 
cies has  its  own  set  of  habitat  requirements,  and  hab- 
itat resources  vary  temporally  and  geographically. 
Consequently,  there  is  much  taxonomic,  temporal, 
and  geographic  variation  in  which  habitat  resources 
limit  wildlife  populations.  However,  if  the  limiting 
resources  are  known  for  a  local  population,  then — by 
definition — the  performance  of  the  population  (e.g., 
abundance,  reproductive  success,  or  survival )  will  be 
correlated  with  variation  in  the  limiting  resources. 
Furthermore,  population  performance  may  be  corre- 
lated with  weather  factors  that  influence  the  produc- 
tion or  availability  of  the  limiting  resources.  These 
weather  factors,  perhaps  precipitation  influencing 
forage  production  or  snow  depth  determining  forage 
availability,  may  then  be  used  in  analyzing  trends  of 
population  performance. 

Biotic  Potential.  Species  with  high  biotic  potentials, 
such  as  small  game  or  songbirds,  have  high  rates  of 
population  turnover.  In  a  favorable  season  of  repro- 
duction, these  species  may  produce  large  autumn 
populations,  regardless  of  variation  in  spring-breed- 
ing abundance  (Steen  1944).  A  biologist  monitoring 
autumn  populations  may  use  weather  factors  affect- 
ing the  breeding  season  and  reproductive  success  in 
analyzing  trends  in  the  population.  However, 
weather  factors  that  determine  overwinter  survival 
and  spring-breeding  density  will  not  be  useful  when 
spring-breeding  density  is  not  well  correlated  with 
autumn  abundance. 

Bias.  Weather  will  influence  the  distribution,  habitat 
selection,  movement  patterns,  and  activity  of  wild- 
life, and  consequently  may  influence  the  results  of 
monitoring  efforts.  Effects  of  snow  conditions  on 
trend  counts  of  ungulates  are  described  above.  Nu- 
merous studies  have  shown  effects  of  weather  on  re- 
sults of  bird  surveys  (Table  1 ),  despite  the  fact  that 
these  surveys  are  usually  conducted  within  a  narrow 
range  of  weather  conditions.  Weather  may  be  known 
to  affect  observability  of  animals,  or  even  sex-age 
classes  of  animals,  and  thus  may  affect  measures  of 
abundance  or  of  population  composition.  Such 
weather  factors  are  likely  candidates  for  use  in 
analyses. 

Seasonal  Precipitation.  Several  studies  have  shown 
wildlife  data  to  be  correlated  with  seasonal  precipita- 
tion, especially  in  arid  areas  (Shaw  1965;  Gallizioli 
1965;  Day  1971;  Medica  et  al.  1975;  Turner  et  al. 
1976;  Whitford  and  Creusere  1977;  Brown  and 
Smith  1980)  where  primary  production  is  stimulated 


Weather  and  Climate 


715 


by  thresholds  of  soil  moisture  and  temperature  (Beat- 
ley  1974).  Most  authors  have  used  total  precipitation 
for  6  or  more  months  preceding  the  wildlife  survey 
as  a  weather  factor  in  data  analysis. 

Snow.  In  northern  or  alpine  areas,  snow  depths, 
quality,  and  persistence  have  been  correlated  with 
wildlife  data  (Severinghaus  1947;  Loveless  1967;  Gil- 
bert et  al.  1970;  Bartman  and  Steinert  1981;  Adams 
and  Bailey  1982).  When  information  on  daily  snow 
depths  is  available,  a  good  measure  of  snow  impacts 
on  animals  and  on  habitat  is  the  total  number  of  win- 
ter days  with  more  than  a  critical  depth  of  snow  on 
the  ground.  Critical  snow  depths  for  deer  have  been 
determined  to  be  38-50  cm  (15-20  in.),  about  the 
height  of  a  deer's  brisket  (Severinghaus  1947;  Bart- 
man and  Steinert  1981 ).  Similar  depths  may  be  esti- 
mated for  other  ungulates.  However,  for  non-ungu- 
lates, and  for  study  areas  with  modest  snow 
accumulation,  this  criterion  may  not  discriminate 
among  years  very  well.  (There  may  be  very  few  days 
with  >  38  cm  [15  in.  ]  of  snow  on  the  ground  in  any 
year;  yet  there  may  be  considerable  important  varia- 
tion among  years  in  persistence  of  snow  depth  <38 
cm  [15  in.]).  In  these  situations,  the  mean  snow 
depth  for  all  winter  days  is  an  arbitrary  criterion  that 
will  discriminate  among  years. 


Cumulative  vs.  Acute  Effects.  When  monitoring 
production  of  a  habitat  resource,  such  as  forage,  or  a 
wildlife  variable  closely  dependent  on  production  of 
a  habitat  resource,  consider  selecting  a  key  weather 
factor  based  on  cumulative  weather  effects.  For  ex- 
ample, forage  production  may  depend  upon  cumula- 
tive rainfall  over  2-to-several  months  before  and  dur- 
ing the  growing  season  (Fisher  1924).  Negative 
effects  of  early-season  weather  upon  forage  produc- 
tion may  be  completely  overridden  by  positive  late- 
season  effects.  Consequently,  forage  production 
would  not  be  correlated  with  weather  data  for  a  cer- 
tain month,  but  may  be  correlated  with  cumulative 
weather  data  for  2  or  more  months. 

In  contrast,  when  monitoring  a  wildlife  popula- 
tion that  is  expected  to  be  directly  affected  by 
weather  during  a  short  period,  the  key  weather  fac- 
tor correlated  with  population  data  is  apt  to  be  ( 1 ) 
attainment  of  a  threshold  of  weather  severity  at  any 
time  during  a  long  season,  or  ( 2 )  weather  during  a 
short  season  when  the  animals  are  especially  vulner- 
able each  year.  The  wildlife  involved  are  apt  to  be 
small-bodied  species,  and  populations  near  the  edges 
of  their  geographic  ranges  (Graber  and  Graber 
1979).  For  example,  a  severe  period  any  time  during 
the  winter  may  acutely  depress  populations  of  small 


Snow  depths  greater  than  18  inches  (45.7  cm)  hamper  mule  deer  and  pronghorn  movements. 


716  Weather  and  Climate 


birds,  and  a  mild  late-winter  period  cannot  compen- 
sate once  birds  have  died. 

In  this  respect,  large  wild  mammals  are  interme- 
diate between  plant  and  small  wildlife.  Large  mam- 
mals draw  upon  substantial  body  reserves  during  se- 
vere weather  and  are  less  subject  to  acute  effects  of 
weather  than  are  small  wildlife.  Population  data  for 
large  mammals  are  often  correlated  with  weather 
data  based  on  cumulative  effects. 


Local  Verification  of  Key  Weather  Factors 

Selection  of  a  key  weather  factor  for  use  in  ana- 
lyzing wildlife  monitoring  data  should  be  based  upon 
a  demonstrated  correlation  between  the  weather  fac- 
tor and  the  local  wildlife  or  habitat  data.  One  may 
assume  that  correlations  based  on  conventional  wis- 
dom, on  assumptions  about  limiting  habitat  re- 
sources, or  on  results  from  other  areas  will  be  cor- 
rectly applicable  to  a  new  monitoring  project.  But 
this  assumption  should  be  tested  with  local  time- 
series  data  before  assuming  there  is  much  confidence 
in  the  results. 

When  a  time  series  of  wildlife  data  is  available, 
it  is  tempting  to  assess  numerous  weather  conditions 
for  possible  correlations  with  the  wildlife  data.  Very 
many  intercorrelated  weather  parameters  can  be 
generated  from  weather  records.  For  instance,  snow 
can  be  represented  by  several  measurements  relating 
to  thresholds,  persistence,  and  constancy,  and  each 
of  these  measurements  can  be  tabulated  for  several 
separate  as  well  as  overlapping  periods  each  winter. 
Correlations  of  each  of  these  resulting  parameters 
with  the  monitoring  data  can  be  assessed  easily  with 
computers.  Some,  often  several,  "statistically  signifi- 
cant" correlations  usually  are  found.  Some  of  these 
may  represent  real  direct  or  indirect  relationships 
between  weather  and  wildlife.  Other  correlations  are 
probably  due  only  to  chance,  since  statistical  signifi- 
cance is  usually  based  on  an  error  rate  (probability 
of  rejecting  the  hypothesis  of  no  correlation  when 
it  is  true )  that  applies  separately  to  each  weather 
parameter. 

Can  future  wildlife  monitoring  data  be  evaluated 
by  using  any  of  these  correlated  weather  parameters 
as  covariants?  Meslow  and  Keith  ( 1971 )  discussed 
some  aspects  of  this  problem.  If  there  is  a  biologi- 
cally plausible  hypothesis  regarding  a  correlation  be- 
tween weather  and  the  wildlife  data,  if  the  hypothe- 
sis is  developed  before  any  correlation  in  the  data  is 
evaluated  subjectively  or  statistically,  if  only  a  few 
weather  measurements  are  used  to  represent  the 
weather  parameter,  and  if  a  conservative  (small)  er- 
ror rate  is  used,  there  should  be  little  risk  in  accept- 
ing the  correlation  as  a  basis  for  evaluating  future 
monitoring  data.  Otherwise,  any  strong  correlations 
between  a  weather  parameter  and  the  monitoring 


data  should  be  viewed  as  a  new  hypothesis  to  be  tested 
( 1 )  for  biological  plausibility  and  (  2 )  with  several  years 
of  new  data,  before  the  indicated  relationship  is  used  in 
assessing  the  results  of  monitoring. 


COVARIANT  ANALYSIS 

In  the  examples  below,  linear  correlation  and 
regression  are  used  to  estimate  relationships  be- 
tween wildlife  characteristics  and  weather  factors.  In 
reality,  relationships  such  as  these  may  be  curvilin- 
ear, and  abrupt  thresholds  may  exist.  With  suitable 
statistical  methods,  non-linear  relations  between 
weather  and  wildlife  may  be  used  in  covariance  anal- 
ysis of  monitoring  data.  However,  detection  and  veri- 
fication of  non-linear  relationships  will  usually  re- 
quire abundant  and  precise  estimates  of  both 
variables,  and  these  are  not  common  in  wildlife  mon- 
itoring. Consequently,  most  analyses  will  involve  lin- 
ear methods. 

Example  1 — Browse  Utilization  and  Snow  Depth 

During  the  1960s,  the  Montana  Fish  and  Game 
Department  measured  winter  utilization  of  service- 
berry  (Amelanchier  alnifolia )  browse  by  white- 
tailed  deer  {Odocoileus  virgmianus)  in  the  Goat 
Creek  drainage  of  the  Swan  Valley.  Utilization  was 
measured  on  several  marked  plants  at  the  end  of 
each  winter.  Presumably,  browse  utilization  was 
being  monitored  as  an  index  to  the  relative  balance 
between  numbers  of  deer  and  the  availability  of 
browse.  A  trend  toward  higher  levels  of  utilization 
would  indicate  either  increasing  numbers  of  deer  or 
a  declining  availability  of  browse  and  might  require  a 
management  response.  The  purpose  here,  however, 
is  only  to  address  the  problem  of  promptly  detecting 
significant  trends  in  levels  of  utilization  when  those 
levels  fluctuate  up  and  down  each  year  ( Figure  2 ). 
These  data  are  extracted  from  Hildebrand  (  1971 ). 

In  1965,  74%  of  scrviceberry  twigs  on  the  tran- 
sect had  been  browsed — the  highest  level  of  utiliza- 
tion in  6  years  (Figure  2).  But  had  a  significant 
change  in  utilization  occurred? 

In  the  North,  wintering  white-tailed  deer  are 
very  sensitive  to  snow  conditions  and  tend  to  con- 
centrate especially  in  conifer  stands,  as  snow  be- 
comes deep.  Browse  transects  were  typically  placed 
in  these  areas  of  concentration,  often  called  "key 
areas."  Consequently,  among-years  variation  in  utili- 
zation of  browse  on  key  areas  often  represented  vari- 
ation in  concentration  of  deer  caused  by  annual  vari- 
ation of  snow  conditions. 

The  number  of  days  with  >48  cm  (19  in.)  of 
snow  on  the  ground  each  winter,  recorded  at  the 
nearby  Seeley  Lake  Ranger  Station,  was  significantly 


Weather  and  Climate 


717 


correlated  with  percentage  utilization  of  serviceberry 
for  the  years  1960-64  (Figure  3).  Based  on  the 
regression  for  those  years,  the  expected  utilization  of 
serviceberry  in  1965,  when  there  had  been  69  days 
with  > 48  cm  ( 19  in.)  of  snow  on  the  ground,  was 
52.5  ±  12.3%  (95%  conf.  limits).  The  observed  uti- 
lization was  74% .  Considering  the  regression  from 
1960  to  1964,  the  probability  of  this  74%  utilization 
being  due  only  to  sampling  error  was  <0.03.  (I  have 
calculated  probabilities  using  untransformed  utiliza- 
tion data,  justified  for  1960-64  when  percentages 
were  near  50.)  This  was  strong  evidence  that  some- 
thing had  changed  in  the  balance  between  deer 
numbers  and  browse  resources  at  Goat  Creek.  As  it 
turned  out,  1965  was  the  first  of  several  years  that 
exhibited  comparatively  high  levels  of  utilization  at 
Goat  Creek  (Figure  3) 

In  1965,  a  biologist  monitoring  browse  utiliza- 
tion in  the  Swan  Valley  may  have  considered  74% 
utilization  on  serviceberry  as  a  signal  for  a  closer 
look  at  conditions  in  the  Valley.  Certainly  in  1966, 
when  utilization  was  78%,  with  only  25  days  of  deep 
snow,  the  biologist  should  have  been  convinced  that 
something  was  amiss;  and  use  of  data  on  snow 
depths  as  a  covariant  in  analyzing  a  time-series  of 
browse-utilization  data  would  have  enhanced  prompt 
detection  of  the  change. 


100 


o\ 

V— ' 

> 
cr 
oc 

LU 
CO 

LU 

O 
> 

DC 
LU 
C/) 

LL 

o 


< 

N 


K 
3 


50 


1960     1961     1962     1963     1964     1965 


Tagged  branch  of  browse  plant. 


Figure  2.  Utilization  of  serviceberry  browse  by 
white-tailed  deer  on  the  Goat  Creek  transect, 
Swan  Valley,  Montana  (Data  from  Hildebrand 
1971) 

Furthermore,  use  of  this  covariant  analysis  of 
monitoring  data  would  have  enhanced  detection  of 
an  impact  of  land-use  practices  on  a  wildlife  popula- 
tion. Clear-cutting  had  begun  on  private  lands  in  the 
Swan  Valley  in  1955.  Winter  range  was  being  elimi- 
nated, concentrating  deer  on  smaller  amounts  of 
range  (Hildebrand  1971). 

Example  2 — Age  Ratios  and  Snow  Depth 

Mountain  goats  (Oreamnos  americanus)  were 
introduced  to  the  Sawatch  Range,  Colorado,  in  1948 
and  onto  Sheep  Mountain  in  that  range  in  1950  (Ad- 
ams and  Bailey  1982).  Age  ratios  of  the  Sheep  Moun- 
tain herd  have  been  observed  as  an  index  to  monitor 
reproductive  success  since  1966.  Age  ratios  varied 
greatly  among  years.  Bailey  and  Johnson  (1977) 
noted  that  these  ratios  for  1966-74  were  correlated 
with  the  depth  of  snow  each  previous  May  at  a  high- 
elevation  location  observed  by  the  U.S.  Soil  Conser- 
vation Service  (Figure  4).  It  is  biologically  plausible 
that  depth  and  persistence  of  snow  on  winter  ranges 
will  increase  demands  upon  nanny  goats  and  de- 
crease forage  availability,  and  that  these  impacts  dur- 
ing late  gestation  might  influence  reproductive  suc- 
cess, especially  survival  of  neonates. 

The  Sheep  Mountain  population  of  mountain 
goats  grew  steadily  during  1965-7-4.  Most  rapid 
growth  apparently  occurred  after  1970  (Adams  and 
Bailey  1982).  Based  on  concepts  of  density -depen- 
dence, especially  with  introduced  ungulates  (Caugh- 
ley  1970),  a  decline  in  reproductive  success  (repro- 
duction +  neonatal  survival )  was  expected.  A 


718 


Weather  and  Climate 


100- 

£ 

1967*                  .1969 

> 
cr 
en 

UJ 
CD 
UJ 

o 

1968 

1965. 

1964.  ^*-^i 

OF  SERV 
o 

1960*           ^ ■***"            ^\ 

^*~ 1960-1964 

Z 

O 
H 
< 

N 

_J 
H 

. -- '           1 963                          Y  =  36.8  +  0.23  X  (p  <  0.03) 

1961 

0- 

c 

i                         i                         i                        i 

1                       25                     50                      75                    100 

NO.   OF   DAYS  WITH  >  48  CM  SNOW  ON  GROUND 

Figure  3.  Use  of  snow-depth  at  Seeley  Lake, 

Montana,  as  a  covariant  in  analysis  of  browse- 
utilization  data.  A  significant  change  in  the 
relationship  began  in  1965. 


decline  in  age  ratios  was  first  suggested  in  1976,  and 
age  ratios  for  1976-79  were  significantly  different 
from  age  ratios  for  1966-75  (Figure  5;  Adams  and 
Bailey  1982).  To  test  this  hypothesis,  an  "artificial 
variable"  representing  the  two  time  periods  was  used 
in  a  multiple-regression  analysis  (X  =  0  for  1966-75; 
X  =   1  for  1976-79).  At  a  mean  snow  depth  of  97 
cm  (39  in.)  for  the  14  years,  age  ratios  had  declined 
by  47%  in  1976-79  compared  with  1966-75. 

Note  that  the  lowest  age  ratios  had  occurred  in 
1968  and  1970,  before  the  decline  indicated  above. 
But  1968  and  1970  were  very  severe  winters.  It  is 
only  after  using  a  weather  variable  in  covariant  anal- 
ysis, so  that  winters  of  similar  severity  may  be  com- 
pared, that  the  decline  in  age  ratios  becomes 
obvious. 


OBTAINING  WEATHER  DATA 

Existing  Records 

Finklin  (1983)  described  the  sources  of  climatic 
data  available  to  wildland  managers.  I  have  ab- 
stracted his  suggestions. 

Year-round  weather  data  are  published  by  the 
U.S.  Weather  Bureau  and  its  successor  agencies  in- 
cluding the  National  Oceanic  and  Atmospheric 


0.8  - 

CO 

N^ 

_i 

N.              • 

<     0.7  - 

\'66 



\                     • 

<     0.6  - 

•                \              '74 

cr 

LU 

'67                   K 

'69  \ 

Q     0.5- 

^v 

_i 

^v 

O 

Ny 

co     0.4. 

•               N. 

Q 

'72              \. 

X. 

\^ 

0.3  - 

N. 

O 

>. 

h- 

>v 

<     0.2- 

•            *              \ 

oc 

'68       '70 

_l 

_l 

<     0.1  - 

LL 

0.0  - 

1           1          1           1           1           1           1 

0          10        20       30       40        50       60       70 

SNOW   DEPTH  ON    1   MAY    Cinches) 

Figure  4.  Relationship  between  1  May  snow  depth 
from  the  previous  winter  at  Monarch  Pass  and 
kidrolder  animal  ratios  from  fall  aerial  surveys  of 
mountain  goats  on  Sheep  Mountain  and  vicinity 
(Bailey  and  Johnson  1977). 


.'66 

• 1966-75 

07. 

< 

2 
z 

.'67 

< 

cr 

06. 

Q 

'69      v^                    < 

74 

O 

CO 

05. 

•  72         \ 

•75 

o 

•  77 

H 

0.4. 

< 

cr 

UJ 

'•<76 

O 

< 

cc 

• 

LU 
2 

78 

2 

Z> 

02. 

CO 
LU 

79 

•68     -70 

H 

5 

01  . 
0.0 

( 

)                  30 

60                 90                120 

150               180 

1  MAY  SNOW  DEPTH  (cm) 

Figure  5.  Relationship  between  1  May  snow  depth 
from  the  previous  winter  at  Monarch  Pass  and 
kid:older  animal  ratios  of  mountain  goats  on 
Sheep  Mountain-Gladstone  Ridge.  Sawatch 
Range,  Colorado,  1966-79  (Adams  and  Bailey 
1982). 


Weather  and  Climate 


719 


Administration  (Hatch  1983).  These  data  are  avail- 
able in  many  major  libraries  or  from  the  National 
Climatic  Center,  Federal  Building,  Asheville,  NC 
28801.  Information  from  local  stations  may  be  found 
in  monthly  and  annual  "Climatological  Data"  sum- 
maries for  each  state.  Each  station  provides  daily  pre- 
cipitation and  maximum  and  minimum  temperatures. 
Some  stations  also  record  evaporation,  soil  tempera- 
tures, monthly  windspeed,  relative  humidity,  sun- 
shine, snowfall,  and  snow  depth. 

Fire -weather  observations,  maintained  by  the 
U.S.  Forest  Service,  provide  an  additional  data  base, 
but  for  the  fire  season  only.  These  include  afternoon 
temperature,  relative  humidity,  wind,  daily  precipita- 
tion, and  daily  maximum  and  minimum  tempera- 
tures. Such  data  are  archived  on  tapes  at  the  National 
Fire  Weather  Data  Library,  Fort  Collins,  CO,  and  are 
available  via  access  to  the  USDA  computer  at  the 


Fort  Collins  Computer  Center.  (See  Finklin  1983  for 
additional  sources.) 

Snow  survey  data  for  1 1  western  states  are  avail- 
able from  the  USDA  Soil  Conservation  Service,  Port- 
land, OR  97209  or  from  state  offices  of  the  Soil  Con- 
servation Service.  Periodically,  state  or  regional  data 
are  compiled  into  a  Summary  of  Snow  Survey 
Measurements. 

Stream  flow  records,  especially  for  major  rivers, 
may  be  obtained  from  state  or  regional  offices  of  the 
U.S.  Geological  Survey.  These  are  compiled  for  states 
or  for  major  watersheds  within  states  and  are  in 
some  libraries.  Records  include  historic  average 
rates,  extreme  rates,  and  daily  rates. 

See  Finklin  ( 1983)  for  other,  minor  sources  of 
weather  data. 


Solar-powered  weather  station. 


Climatological  data  summaries. 


720 


Weather  and  Climate 


Obtaining  Local  Data 

It  may  be  desirable  to  monitor  one  or  a  few 
weather  variables  on  a  study  area  as  part  of  the  mon- 
itoring program.  This  would  occur  if  existing 
weather  stations  are  not  nearby  of  if  they  do  not  ad- 
equately represent  the  study  area  because  of  differ- 
ences in  elevation,  aspect,  habitat,  or  topography. 
Also,  key  weather  factors  may  not  be  recorded  at  the 
nearby  cooperating  Weather  Bureau  station.  (For  ex- 
ample, not  all  stations  maintain  records  of  snow 
depths. )  Descriptions  of  commonly  used  weather  in- 
struments, of  weather  stations,  and  of  weather  data 
collection  and  analysis  are  available  from  Day  and 
Sternes  (1970)  and  Critchneld  (1974). 


Wildlife  biologists  have  developed  a  few  special 
instruments  for  measuring  impacts  of  weather  on  an- 
imals. Although  deep  snow  may  hinder  movements 
of  large  mammals,  snow  quality — in  particular,  crust- 
ing— may  diminish  or  exacerbate  problems  of  mov- 
ing in  snow,  depending  on  the  density  and  location 
of  crusts  in  the  snowpack.  Hepburn  ( 1978)  de- 
scribed a  penetrometer  for  measuring  the  effective 
depth  of  snow  for  large  mammals  when  the  snow- 
pack  is  crusted.  Energy  demands  of  wintering  big 
game  may  depend,  in  part,  on  the  combined  effects 
of  temperature,  wind,  and  radiation  exchange  with 
the  environment  (Moen  1968,  1973).  Verme  (1968) 
has  described  a  simple  device  for  measuring  these 
combined  effects  upon  energy  demands. 


Weather  and  Climate 


721 


LITERATURE  CITED 

ADAMS,  L.G.  and  J.A.  BAILEY.  1982.  Population  dynamics 

of  mountain  goats  in  the  Sawatch  Range,  Colorado.  J. 

Wildl.  Manage.  46:1003-1009. 
BAILEY,  J  A.  1969.  Trap  responses  of  wild  cottontails.  J. 

Wildl.  Manage.  33:48-58. 
and  B.K  JOHNSON.  1977.  Status  of  introduced 

mountain  goats  in  the  Sawatch  Range  of  Colorado. 

Pages  54-63  in  Samuel,  W.  and  W.G  Macgregor,  eds. 

Proc.  First  Int.  Mountain  Goat  Symp.  Kalispell,  MT. 
BARTMANN,  R.M.  and  S.F.  STEINERT.  1981.  Distribution 

and  movements  of  mule  deer  in  the  White  River 

Drainage,  Colorado.  Colorado  Div.  Wildl.  Spec.  Rept. 

51.  12pp. 
BEATLEY,  J.C.  1969.  Dependence  of  desert  rodents  on 

winter  annuals  and  precipitation.  Ecology  50:721-724. 
.  1974.  Phenological  events  and  their  environmental 

triggers  in  Mohave  desert  ecosystems.  Ecology  55:856- 

863. 

1976.  Rainfall  and  fluctuating  plant  populations  in 


relation  to  distributions  and  numbers  of  desert  ro- 
dents in  southern  Nevada.  Oecologia  24:21-42. 

BEST,  LB.  1981.  Seasonal  changes  in  detection  of  individ- 
ual bird  species.  Studies  in  Avian  Biology  6:252-261. 

BIRCH,  L.C.  1958.  The  role  of  weather  in  determining  the 
distribution  and  abundance  of  animals.  Cold  Spring 
Harbor  Symposia  in  Quantitative  Biology  22:203-215. 

BROWN,  D.E.  and  R.H.  SMITH.  1980.  Winter-spring  pre- 
cipitation and  population  levels  of  blue  grouse  in  Ari- 
zona. Wildl.  Soc.  Bull.  8:136-141. 

BROWN,  W.H.  1971.  Winter  population  trends  in  the  red- 
shouldered  hawk.  Am.  Birds  25:813-817. 

BRUNS,  EH.  1977.  Winter  behavior  of  pronghorns  in  rela- 
tion to  habitat.  J.  Wildl.  Manage.  41:560-571. 

CAUGHLEY,  G.  1970.  Eruption  of  ungulate  populations, 
with  emphasis  on  Himalayan  thar  in  New  Zealand. 
Ecology  51:53-72. 

CLUTTON-BROCK,  T.H.  and  S.D.  ALBON.  1983-  Climatic 
variation  and  body  weight  of  red  deer.  J.  Wildl.  Man- 
age. 47:1197-1201. 

CONNER,  R.N.  and  J.G.  DICKSON.  1980.  Strip  transect 
sampling  and  analysis  for  avian  habitat  studies.  Wildl. 
Soc.  Bull.  8:4-10. 

CONNOLLY,  G.E.  1981.  Limiting  factors  and  population 
regulation.  Chapter  7  in  Wallmo,  O.C.,  ed.  Mule  and 
Black-tailed  Deer  of  North  America.  Univ.  of  Nebraska 
Press,  Lincoln. 

CRECCO,  V.A.  and  T.F.  SAVOY.  1984.  Effects  of  fluctua- 
tions in  hydrographic  conditions  on  year-class 
strength  of  American  shad  (Alosa  sapidissima)  in  the 
Connecticut  River.  Can.  J.  Fish.  Aquatic  Sci.  41:1216- 
1223- 

CRITCHFIELD,  H.J.  1974.  General  climatology  (3rd  Edi- 
tion). Prentice-Hall,  Inc.  446pp. 

DAWSON,  D.G.  1981.  Counting  birds  for  a  relative  mea- 
sure (index)  of  density.  Studies  in  Avian  Biology  6:12- 
16. 

DAY,  G.I.  1971.  Wildlife  research  in  Arizona,  1970-71. 
Pages  153-157  in  Wildlife  Research  in  Arizona.  Ari- 
zona Game  and  Fish  Dept.  Phoenix. 

DAY,  J  A.  and  G.L.  STERNES.  1970.  Climate  and  Weather. 
Addison-Wesley  Publ.  Co.  Reading,  MA.  407pp. 

DIEM,  KL.  and  KH.  LU.  I960.  Factors  influencing  water- 
fowl censuses  in  the  Parklands,  Alberta,  Canada.  J. 
Wildl.  Manage.  24:113-133. 


EMLEN,  J.T.  and  M.J.  DEJONG.  1981.  The  application  of 
song  detection  threshold  distance  to  census  opera- 
tions. Studies  in  Avian  Biology  6:346-352. 

FINKLIN,  A.I.  1983.  Summarizing  weather  and  climatic 
data — a  guide  for  wildland  managers.  U.S.  Dep.  Agric, 
For.  Serv.  Gen.  Tech.  Rep.  INT- 148.  43pp. 

FISHER,  R.A.  1924.  The  influence  of  rainfall  on  the  yield 
of  wheat  at  Rothamsted.  Royal  Soc.  London  Phil. 
Trans.,  Series  B.  213:89-142. 

FRANCIS,  W.J.  1970.  The  influence  of  weather  on  popula- 
tion fluctuations  in  California  quail.  J.  Wildl.  Manage. 
34:249-266. 

GAGNARK,  H.A.  and  R.G.  BAKKALA.  I960.  A  comparative 
study  of  unstable  and  stable  ( artificial  channel )  spawn- 
ing streams  for  incubating  king  salmon  at  Mill  Creek. 
California  Fish  and  Game  46:151-164. 

GALLIZIOLI,  S.  1965.  Quail  research  in  Arizona.  Arizona 
Game  and  Fish  Dep.,  Phoenix.  12pp. 

GILBERT,  P.F.,  O.C.  WALLMO,  and  R.B.  GILL.  1970.  Effect 
of  snow  depth  on  mule  deer  in  Middle  Park,  Colorado. 
J.  Wildl.  Manage.  34:15-23. 

GRABER,  J.W.  and  R.R.  GRABER.  1979.  Severe  winter 

weather  and  bird  populations  in  southern  Illinois.  Wil- 
son Bull.  91:88-103. 

GULLION,  G.W.  I960.  The  ecology  of  Gambel's  quail  in 
Nevada  and  the  arid  southwest.  Ecology  41:518-536. 

.  1966.  The  use  of  drumming  behavior  in  ruffed 

grouse  population  studies.  J.  Wildl.  Manage.  30:717- 
729. 

HAMMOND,  M.C.  and  D.H.  JOHNSON.  1984.  Effects  of 
weather  on  breeding  ducks  in  North  Dakota.  U.S.  Dep. 
Inter.,  Fish  and  Wildl.  Serv.  Tech.  Rep.  1.  17pp. 

HATCH,  W.L.  1983-  Selective  guide  to  climatic  data 

sources.  Key  to  Meteorological  Records  Documenta- 
tion No.  4.11.  National  Oceanic,  Atmospheric  Admin. 
Asheville,  NC.  338pp. 

HEPBURN,  R.L.  1978.  A  snow  penetration  gauge  for  stud- 
ies of  white-tailed  deer  and  other  northern  mammals. 
J.  Wildl.  Manage.  42:663-667. 

HILDEBRAND,  PR.  1971.  Biology  of  white-tailed  deer  on 
winter  ranges  in  the  Swan  Valley,  Montana.  M.S.  The- 
sis, Univ.  Montana.  Missoula.  91pp. 

HOCUTT,  C.H.  andJ.R.  STAUFFER,  JR.,  eds.  1980.  Biologi- 
cal monitoring  of  fish.  Lexington  Books.  Lexington, 
KY.  41 6pp. 

HODDENBACH,  G.A.  and  F.B.  TURNER.  1968.  Clutch  size 
of  the  lizard  Ufa  stansburiana  in  southern  Nevada. 
Amer.  Midi.  Nat.  80:262-265. 

HUBERT,  W.A.  and  D.N.  SCHMITT.  1982a.  Factors  influ- 
encing catches  of  drifted  trammel  nets  in  a  pool  of  the 
Upper  Mississippi  River.  Proc.  Iowa  Acad.  Sci.  89:153- 
154. 

and .  1982b.  Factors  influencing  hoop  net 

catches  in  channel  habitats  of  Pool  9,  Upper  Missis- 
sippi River.  Proc.  Iowa  Acad.  Sci.  89:84-88. 

JACOBS,  D.  and  KR.  DIXON.  1981.  Breeding-season  pre- 
cipitation and  the  harvest  of  cottontails.  J.  Wildl.  Man- 
age. 45:101 1-1014. 

JEHL,  J.  and  D.  HUSSELL.  1966.  Effects  of  weather  on  re- 
productive success  of  birds  at  Churchill,  Manitoba. 
Arctic  19:185-191. 

JOHN,  KR.  1963.  The  effect  of  torrential  rains  on  the  re- 
productive cycle  of  Rhinichthys  osculus  in  the  Chiri- 
cahua  Mountains,  Arizona.  Copeia  2:286-291. 

.  1964.  Survival  of  fish  in  intermittent  streams  of  the 

Chiricahua  Mountains,  Arizona.  Ecology  45:112-119. 


722 


Weather  and  Climate 


JONES,  G.  1952.  Hail  damage  to  wildlife  in  southwest 
Oklahoma.  Wilson  Bull.  64:166-167. 

KIRKLAND,  L.  1962.  A  tagging  experiment  on  spotted  and 
largemouth  bass  using  an  electric  shocker  and  the  Pe- 
tersen disc  tag.  Proc.  Southeast.  Assoc.  Game  and  Fish 
Comm.  16:424-432. 

LA  PERRIERE,  A  J.  and  A.O.  HAUGEN.  1972.  Some  factors 
influencing  calling  activity  of  wild  mourning  doves.  J. 
Wildl.  Manage.  36:1193-1199. 

LARIMORE,  R  W.  1959.  Destruction  and  re-establishment 
of  stream  fish  and  invertebrates  affected  by  drought. 
Trans.  Am.  Fish.  Soc.  88:261-285. 

LE  RESCHE,  RE.  and  R.A.  RAUSCH.  1974.  Accuracy  and 
precision  of  aerial  moose  censusing.  J.  Wildl.  Manage. 
38:175-182. 

LILLYWHITE,  H.B.  1982.  Tracking  as  an  aid  in  ecological 
studies  of  snakes.  Pages  181-191  in  Scott,  NJ.,  Jr.,  ed. 
Herpetological  Communities.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  Wildl.  Res.  Rep.  1 3. 

LONGHURST,  W.M.  andJ.R.  DOUGLAS.  1953.  Parasite  in- 
terrelationships of  domestic  sheep  and  Columbian 
black-tailed  deer.  Trans.  North  Am.  Wildl.  Nat.  Resour. 
Conf.  18:168-188. 

LOVAAS,  A.L  1958.  Mule  deer  food  habits  and  range  use, 
Little  Belt  Mountains,  Montana.  J.  Wildl.  Manage. 
22:275-283. 

LOVELESS,  CM.  1967.  Ecological  characteristics  of  a  mule 
deer  winter  range.  Colorado  Div.  Game,  Fish  and 
Parks.  Tech.  Publ.  20.  Denver,  CO.  1 24pp. 

MARTINKA,  C.J.  1967.  Mortality  of  northern  Montana 
pronghorns  in  a  severe  winter.  J.  Wildl.  Manage. 
31:159-164. 

MATTHEWS,  W.J.  and  J.T  STYRON,  JR.  1981.  Tolerance  of 
headwater  vs.  mainstream  fishes  for  abrupt  physico- 
chemical  changes.  Am.  Midi.  Nat.  105:149-158. 

MAY,  N.,  L.  TRENT,  and  P.J.  PR1STAS.  1976.  Relation  offish 
catches  in  gill  nets  to  frontal  periods.  Fishery  Bull. 
74:449-453. 

MEDICA,  PA.,  R.B.  BURY  and  F.B.  TURNER.  1975.  Growth 
of  the  desert  tortoise  ( Gopherus  agassizi )  in  Nevada. 
Copeia  4:639-643- 

MESLOW,  E.C.  and  LB.  KEITH.  1971.  A  correlation  analy- 
sis of  weather  versus  snowshoe  hare  population  pa- 
rameters. J.  Wildl.  Manage.  35:1-15. 

MOEN,  A.N.  1968.  Surface  temperatures  and  radiant  heat 
loss  from  white-tailed  deer.  J.  Wildl.  Manage.  32: 
38-344. 

.  1973-  Wildlife  Ecology.  W.H.  Freeman  and  Co.  San 

Francisco,  CA.  458pp. 

MORRISON,  ML.  and  R.D.  SLACK.  1977.  Population  trends 
and  status  of  the  Olivaceous  Cormorant.  Am.  Birds 
31:954-959. 

MORSE,  M.A.  and  D.S.  BALSER.  1961.  Fox  calling  as  a 
hunting  technique.  J.  Wildl.  Manage.  25:148-154. 

MUEGGLER,  W.F.  1971.  Weather  variations  on  a  mountain 
grassland  in  southwestern  Montana.  U.S.  Dep.  Agric, 
For.  Serv.  Res.  Pap.  INT-99.  25pp. 

.  1983.  Variation  in  production  and  seasonal  devel- 
opment of  mountain  grasslands  in  western  Montana. 
U.S.  Dep.  Agric,  For.  Serv.  Res.  Pap.  INT-316.  16pp. 

NEHRING,  R.B.  and  R.  ANDERSON.  1984.  Recruitment 
and  survival  of  young-of-the-year  brown  trout  {Salmo 
trutta  L. )  in  the  South  Fork  of  the  Rio  Grande  River 
versus  parent  spawner  density,  stream  discharge  and 
fry  habitat.  Proc.  Colorado- Wyoming  Chapt.  Am.  Fish. 
Soc.  19:35-43. 


NEWMAN,  D.E.  1959.  Factors  influencing  the  winter 
roadside  count  of  cottontails.  J.  Wildl.  Manage. 
23:290-294. 

NLXON,  CM.,  M.W.  McCLAIN,  and  LP.  HANSEN.  1980.  Six 
years  of  hickory  seed  yields  in  southwestern  Ohio.  J. 
Wildl.  Manage.  44:534-539. 

OBERT,  H.  1976.  Some  effects  of  external  factors  upon  the 
reproductive  behavior  of  the  grass  frog,  Rana  tempor- 
aria.  Oecologia.  24:43-55. 

OTTAWAY,  E.M.  and  A.  CLARK  1981.  A  preliminary  inves- 
tigation into  the  vulnerability  of  young  trout  {Salmo 
trutta  L. )  and  Atlantic  salmon  {Salmo  salar  L)  to 
downstream  displacement  by  high  water  velocities.  J. 
Fish.  Biol.  19:135-145. 

PHILLIPS,  R.L.,  R.O.  ANDREWS,  G.L.  STORM,  and  R.A. 

BISHOP.  1972.  Dispersal  and  mortality  of  red  foxes.  J. 
Wildl.  Manage.  36:237-248. 

PICTON,  H.D.  1984.  Climate  and  the  prediction  of  repro- 
duction of  three  ungulate  species.  J.  Applied  Ecol. 
21:869-879. 

PORTER,  D.K  1973-  Accuracy  in  censusing  breeding  pas- 
serines on  the  shortgrass  prairie.  M.S.  Thesis,  Colorado 
State  Univ.  Ft.  Collins.  107pp. 

PROGULSKE,  DR.  and  DC.  DUERRE.  1964.  Factors  influ- 
encing  spotlighting  counts  of  deer.  J.  Wildl.  Manage. 
28:27-34. 

RAVELING,  D.G.  1966.  Factors  affecting  age  ratios  of  sam- 
ples of  Canada  geese  caught  with  cannon  nets.  J. 
Wildl.  Manage.  30:682-691. 

RICHENS,  V.B.  1967.  Characteristics  of  mule  deer  herds 
and  their  range  in  northeastern  Utah.  J.  Wildl.  Manage. 
31:651-666. 

RINNE,  J.N.  1975.  Changes  in  minnow  populations  in  a 
small  desert  stream  resulting  from  naturally  and  artifi- 
cially induced  factors.  Southwest.  Nat.  20:185-195. 

ROBBINS,  C.S.  1981.  Bird  activity  levels  related  to 
weather.  Studies  in  Avian  Biology  6:301-310. 

ROBEL,  R.J.,  D.J.  DICK  and  G.F.  KRAUSE.  1969.  Regres- 
sion coefficients  used  to  adjust  bobwhite  quail  whistle 
count  data.  J.  Wildl.  Manage.  33:662-668. 

RUTHERFORD,  W.H.  1970.  The  Canada  geese  of  south- 
eastern Colorado.  Colorado  Div.  Game,  Fish  and  Parks. 
Tech.  Publ.  26.  65pp. 

SEVERLNGHAUS,  CW.  1947.  Relationships  of  weather  to 
winter  mortality  and  population  levels  among  deer  in 
the  Adirondack  region  of  New  York.  Cong.  12:212- 
223. 

SHAW,  H.  1965.  Investigations  of  factors  influencing  deer 
populations.  Pages  125-143  in  Wildlife  Research  in  Ar- 
izona, 1964.  Arizona  Game  and  Fish  Dep.  Phoenix. 

SHIELDS,  W.M.  1979.  Avian  census  techniques:  an  analyti- 
cal review.  Pages  23-51  in  Dickson,  J.G.,  R.N.  Conner, 
RR.  Fleet,  J.C  Kroll,  and  J.A.  Jackson,  eds.  The  Role  of 
Insectivorous  Birds  in  Forest  Ecosystems.  Academic 
Press.  New  York,  NY. 

SHIFLET,  T.N.  and  HE.  DIETZ.  1974.  Relationship  be- 
tween precipitation  and  annual  rangeland  herbage 
production  on  southeastern  Kansas.  J.  Range  Manage. 
27:272-276. 

SHVONEN,  L.  1956.  The  correlation  between  the  fluctua- 
tions of  partridge  and  European  hare  populations  and 
the  climatic  conditions  of  winters  in  southwest  Fin- 
land during  the  last  thirty  years.  Papers  on  Game  Res., 
Finnish  Game  Foundation  17:1-30. 

SMITH,  A.  I960.  Hail:  great  destroyer  of  wildlife.  Audubon 
Mag.  62:170-171,  189. 


Weather  and  Climate 


723 


SMOLIAK,  S.  1956.  Influence  of  climatic  conditions  on  for- 
age production  of  shortgrass  rangeland.  J.  Range  Man- 
age. 9:89-91. 

STARRET,  W.C.  1951.  Some  factors  affecting  the  abun- 
dance of  minnows  in  the  Des  Moines  River,  Iowa. 
Ecology'  32:13-27. 

STEEN,  M.O.  1944.  The  significance  of  population  turn- 
over in  upland  game  management.  Trans.  North  Am. 
Wildl.  Conf.  9:331-335. 

STEHR,  W.C.  and  J.W.  BRANSON.  1938.  An  ecological 
study  of  an  intermittent  stream.  Ecology  19:294- 
310. 

STEPHENSON,  R.L.  and  D.E.  BROWN.  1980.  Snow  cover 
as  a  factor  influencing  mortality  of  Abert's  squirrels.  J. 
Wildl.  Manage.  44:951-955. 

STOUT,  I.J.  and  G.W.  CORNWELL.  1976.  Nonhunting  mor- 
tality of  fledged  North  American  waterfowl.  J.  Wildl. 
Manage.  40:681-693- 

TRAMER,  E.J.  1977.  Catastrophic  mortality  of  stream  fishes 
trapped  in  shrinking  pools.  Am.  Midi.  Nat.  97:469- 
478. 

TURKOWSKI,  F.J.  and  JR.  VAHLE.  1977.  Desert  rodent 
abundance  in  southern  Arizona  in  relation  to  rainfall. 
U.S.  Dep.  Agric,  For.  Serv.  Res.  Note  RM-346.  4pp. 

TURNER,  F.B.,  PA.  MEDICA,  JR.  LANNON,  JR.,  and  G.A. 
HODDENBACH.  1969.  A  demographic  analysis  of 
fenced  populations  of  the  whiptail  lizard,  Cnemido- 
phorns  tigris,  in  southern  Nevada.  Southwest.  Nat. 
14:189-202. 


and  B.W.  KOWALEWSKY.  1976.  Energy 


utilization  by  a  desert  lizard  (Uta  stansburiana). 
US/IBP  Desert  Biome  Monogr.  1.  57pp. 

VAUGHT,  R.W.,  H.C.  MCDOUGLE,  and  H.  H.  BURGESS. 
1967.  Fowl  cholera  in  waterfowl  at  Squaw  Creek  Na- 
tional Wildlife  Refuge.  Missouri.  J.  Wildl.  Manage. 
31:248-253. 

VERME,  L.J.  1968.  An  index  of  winter  weather  severity  for 
northern  deer.  J.  Wildl.  Manage.  32:566-574. 

VOGT,  R.C.  and  R.L  MINE.  1982.  Evaluation  of  techniques 
for  assessment  of  amphibian  and  reptile  populations  in 
Wisconsin.  Pages  201-217  in  Scott,  N.J. ,  Jr.,  ed.  Herpe- 
tological  Communities.  U.S.  Dep.  Inter.,  Fish  and  Wildl. 
Serv.  Wildl.  Res.  Rep.  1 3 

WALLMO,  O.C.,  A.W.JACKSON,  T.L.  IIAILEY,  and  R.L 
CARLISLE.  1962.  Influence  of  rain  on  the  count  of 
deer  pellet  groups.  J.  Wildl.  Manage.  26:50-55. 

WHITFORD,  W.G.  and  F.M.  CREUSERE.  1977.  Seasonal  and 
yearly  fluctuations  in  Chihuahuan  Desert  lizard  com- 
munities. Herpetologica  335-4-65. 

WIGHT,  H.M.  and  C.H.  CONAWAY.  1961.  Weather  influ- 
ences on  the  onset  of  breeding  in  Missouri  cottontails. 
J.  Wildl.  Manage.  25:87-89. 

WORF,  D.L.,  ed.  1980.  Biological  monitoring  for  environ- 
mental effects.  Lexington  Books,  Lexington,  KY.  227 
pp. 

YOCUM,  C.F.  1950.  Weather  and  its  effect  on  hatching  of 
waterfowl  in  eastern  Washington.  Trans.  North  Am. 
Wildl.  Nat.  Resour.  Conf.  15:309-319. 


724 


Weather  a/id  Climate 


VI  ANALYSIS  AND 
PRESENTATION 

36  Data  Management 

37  Statistical  Analysis 

38  Habitat  Evaluation  Systems 

39  Evaluation  and  Interpretation 

40  Economic  Analysis 

41  Written  Communications 

42  Verbal  Presentations 


m mm 


36 

DATA 
MANAGEMENT 

Larry  Peterson   and  Iris  Matney 


U.S.  Bureau  of  Land  Management 
Service  Center 
Denver,  CO  80225 


"A  frequent  predicament  of  field  workers  is  to  accu- 
mulate so  many  notes  that  time  is  lacking  to  analyze 
them,  or  to  have  notes  string  out  over  such  a  long 
period  that  the  earlier  ones  are  lost  or  hard  to  segre- 
gate by  the  time  a  sufficient  volume  are  at  hand  to 
warrant  a  conclusion." 

— Aldo  Leopold,  Game  Management  ( 1948) 


Editor's  Note:  The  field  of  data  management  is  rap- 
idly growing  and  changing.  Computers  are  already 
widely  used  in  business  and  science  and  are  used 
increasingly  in  resource  management.  Furthermore, 
computers  are  no  longer  merely  number  crunchers; 
data  processing  now  also  includes  image  process- 
ing and  word  processing.  While  data  management 
technology  is  quite  sophisticated,  its  application 
is  becoming  more  and  more  user  friendly.  This 
chapter  provides  an  overview  of  the  processes  in- 
volved in  data  management,  but  the  reader  will 
have  to  keep  up  with  specifics.  A  major  obstacle  to 
the  new  user  of  modern  computer  systems  is  the 
jargon  associated  with  it.  This  chapter  introduces 
many  of  the  terms  and  concepts  used  in  the  field,  so 
that  the  biologist  can  communicate  with  the  data- 
processing  specialist  when  necessary. 


INTRODUCTION 

Aldo  Leopold  (1948)  warned  biologists  about 
problems  of  data  management.  This  same  problem 
exists  today.  By  using  computers,  wildlife  biologists 
can  devote  their  attention  to  the  collection  of  data 
and  their  time  to  the  interpretation  and  analysis  of 
information.  They  can  provide  options  or  recommen- 
dations to  management  while  a  computer  performs 
the  tasks  of  adding,  subtracting,  sorting,  and  other 
data  manipulations.  The  computer  is  an  important 
tool  in  the  management  of  our  natural  resources. 

Data  processing  for  the  wildlife  biologist  can 
mean  very  simple  or  complex  procedures.  It  can 
mean  dealing  with  a  whole  new  group  of  people 
who  have  their  own  jargon  and  who  may  not  be  fa- 
miliar with  the  biologist's  needs  or  activities.  This 
chapter  discusses  some  of  the  most  elemental  con- 
cepts of  data  management.  We  have  suggested  some 
basic  references  at  the  end  of  the  chapter  to  expand 
on  the  topics  presented  here. 

The  data  processing  field  is  now  frequently 
called  information  management.  It  is  relatively  new 
and  has  been  developing  rapidly.  The  first  large-scale 
electronic  digital  computer  was  built  in  1946.  It 
contained  18,000  vacuum  tubes  and  was  pro- 
grammed by  connecting  various  wires  between  units 
of  the  computer  and  setting  up  6,000  switches  so 
the  program  would  run  (Shelly  and  Cashman 
1980:2.3)  By  today's  standards,  this  computer  was 
big,  slow,  unreliable,  and  difficult  to  program. 

By  1954,  the  transistor  brought  about  the  sec- 
ond generation  of  computers  which  were  faster, 
smaller,  and  less  expensive.  In  1964,  third  generation 
computers  using  Solid  Logic  Technology  (SLT)  be- 
gan to  appear. 

'Current  Address:  Tckton  Software,  Inc.,  1495  Canyon  Blvd..  Suite 
100,  P.O.  Box  7300,  Boulder,  CO  80306 


Data  Management 


727 


The  minicomputer  and  the  new  era  of  inte- 
grated electronics  began  with  the  use  of  silicon 
chips.  Small  processing  units  have  brought  the  com- 
puter into  many  small  businesses,  homes,  and 
schools.  In  fact,  some  colleges  require  incoming 
students  to  have  microcomputers.  It  is  a  rare  indi- 
vidual, even  a  wildlife  biologist,  who  has  not  been 
affected  in  some  way  by  a  computer.  The  Inter- 
nal Revenue  Service,  utility  companies,  grocery 
stores,  payroll  departments,  and  organizations — all 
use  computers. 

A  relatively  new  development  in  computer  tech- 
nology, of  particular  interest  to  wildlife  biologists 
and  resource  managers,  is  the  use  of  Geographic 
Information  Systems  (GISs).  A  GIS  consists  of  soft- 
ware and  hardware  for  performing  spatial  analyses. 
Thus,  the  operations  that  biologists  do  on  maps,  such 
as  calculating  areas,  amount  of  edge,  distance  from 
waters,  etc.,  can  now  be  automated.  This  technology 


is  not  discussed  in  detail  in  this  chapter;  however, 
the  principles  described  apply  equally  well  to  such 
systems  as  to  the  traditional  systems  that  process 
alphanumeric  data.  For  further  information  on  GISs, 
see  de  Steiguer  and  Giles  ( 1981 ),  Steenhof  ( 1982), 
and  Mayer  (1984). 

A  common  problem  in  thinking  about  data  man- 
agement is  to  consider  data  only  as  numerical  data, 
and  computers  or  automated  systems  as  only  "num- 
ber crunchers"  or  systems  for  dealing  with  numeri- 
cal data.  Spatial  data  or  maps  analyzed  by  GISs  are  as 
much  data  as  text  processed  on  word  processors. 
Unfortunately,  the  technologies  for  automating  such 
data  have  developed  separately  in  many  cases,  such 
that  they  must  be  used  independently.  However,  the 
future  trend  is  to  develop  integrated  software  and 
hardware  so  the  user  can  easily  use  all  three  types  of 
data  in  one  system,  just  as  biologists  now  use  maps, 
tables,  and  written  reports.  The  following  discussion 


MICROPROCESSORS  AND  MICROCOMPUTERS 

A  microprocessor  is  a  one-chip  central  pro- 
cessing unit  (CPU).  That  is,  all  of  the  components  nec- 
essary to  perform  the  CPU  functions  are  contained  on 
one  silicon  chip.  The  small,  relatively  inexpensive  mi- 
croprocessors are  used  in  a  variety  of  products,  includ- 
ing hand-held  calculators,  vending  machines,  and 
microwave  ovens. 


A  microprocessor  together  with  additional  chips 
containing  circuitry  for  timing,  temporary  storage,  and 
interfaces  for  input  and  output  devices  forms  a  micro- 
computer. First  used  in  industrial  automation,  the 
microcomputer  has  moved  into  a  diverse  environment 
ranging  from  business  and  banking  to  transportation 
and  manufacturing.  Most  personal  and  home  computers 
are  microcomputers.  A  microcomputer  will  be  able  to 
fill  many  of  a  biologist's  computing  needs.  However, 
complex  programs  or  large  amounts  of  data  may  re- 
quire larger  computers  to  process  the  data  efficiently. 


MINICOMPUTERS 

Minicomputer  is  a  term  which  was  first  used  to 
describe  a  machine  that  was  smaller  and  less  expensive 
than  mainframe  computers  typically  manufactured  at 
the  time.  Since  that  time,  the  range  in  size,  ability,  and 
cost  of  computers  has  increased  so  much  that  the  term 
has  lost  its  original  significance.  Although  the  term  is 
still  used,  its  meaning  varies  greatly.  A  machine  called  a 
minicomputer  by  some  users  may  be  referred  to  as  a 
small  computer  by  others.  Similarly,  some  people  may 
call  a  microcomputer  a  minicomputer.  The  usual  dis- 
tinction between  a  "micro"  and  a  "mini"  is  whether  or 
not  the  processing  unit  is  on  one  chip  or  many. 


MAINFRAME  COMPUTERS 

Full-scale  computers,  termed  mainframe  com- 
puters, are  capable  of  processing  large  amounts  of  data 
at  very  fast  speeds  with  access  to  billions  of  characters 
of  data.  Depending  on  the  amount  of  processing  done 
at  one  location,  several  operators  working  in  shifts  may 
be  required.  Also,  an  alternate  power  supply  and  ade- 
quate air  conditioning  are  generally  required. 


GEOGRAPHIC  INFORMATION  SYSTEM  (GIS) 

A  Geographic  Information  System  (GIS)  stores, 
displays,  retrieves,  and  analyzes  spatial  data.  It  is  similar 
to  other  automated  information  systems  except  that  it 
handles  spatial  data  instead  of  just  words  and  numbers. 
Since  spatial  data  can  be  associated  with  specific  geo- 
graphic locations,  the  system  can  model  or  simulate 
land  uses  and  resource  values.  Many  different  systems 
are  available  either  as  public  domain  software  (not 
copyrighted)  or  as  proprietary  systems  (copyrighted 
and  sold  under  license)  that  must  be  purchased.  MOSS 
(Map  Overlay  Statistical  System)  is  a  public  domain 
system  that  is  used  by  the  U.S.  Bureau  of  Land  Manage- 
ment and  several  other  agencies.  ARC/INFO,  on  the 
other  hand,  is  a  proprietary  system,  available  only  under 
license  from  the  manufacturer. 


728 


Data  Management 


of  data  management  applies  to  all  types  of  data  even 
though  the  examples  given  deal  mostly  with  alpha- 
numeric data. 

A  computer  consists  of  a  group  of  devices  that 
process  data  without  intervention  by  people  at  every 
step.  The  machinery  that  makes  up  the  computer 
system  is  called  hardware.  Software  or  programming 
is  merely  a  set  of  instructions  telling  the  machine 
what  to  do. 

Computers  can  do  only  a  limited  number  of 
things.  Programs  must  be  written  to  tell  them  ex- 
actly what  to  do  and  when  to  do  it.  Computers  do 
these  functions  quickly  and  reliably  and  are  espe- 
cially good  for  repetitive  tasks,  such  as — 


( 1 )    Mathematical  tasks,  including  addition,  sub- 
traction, multiplication,  and  division; 


(2)  Logical  operations,  which  may  include  com- 
paring two  values  to  determine  whether  one 
is  larger  or  smaller  than  the  other;  and 

(3)  Input/output  operations,  such  as  accepting 
data  for  processing  or  outputting  data  in  a 
printed  report. 

A  wildlife  biologist  does  these  same  tasks.  There 
is  nothing  a  computer  can  do  with  data  that  cannot 
be  done  by  hand.  But  a  computer  does  calculations 
and  comparisons  much  faster  than  wildlife  biologists. 
In  fact,  all  the  data  processing  and  computational 
work  done  by  computers  in  the  world  today  can  no 
longer  be  done  by  hand:  about  400  billion  people 
would  be  needed  to  tackle  this  workload  (Shelly  and 
Cashman  1980:2.1). 

The  computer  is  as  available  as  trucks,  aerial 
photos,  or  helicopters,  and  it  makes  people  more  ef- 
ficient. One  does  not  need  to  know  how  to  manufac- 


WORD  PROCESSING 

One  of  the  major  uses  of  microcomputers  by  many 
businesses  is  word  processing.  Systems  range  from  the 
editing  typewriter  to  communicating  word  processors 
that  provide  file  transfer  or  electronic  mail  capability. 
Some  systems  also  have  the  ability  through  other  soft- 
ware to  perform  some  data  processing  applications  such 
as  payroll  and  accounting. 


HARDWARE,  SOFTWARE,  AND  FIRMWARE 

Computer  functions  are  often  discussed  in  terms 
of  hardware  and  software.  Hardware  is  a  term  used 
for  the  physical  components  of  the  system.  Input  de- 
vices, output  devices,  main  storage,  and  auxiliary  stor- 
age devices  are  all  part  of  the  computer  hardware. 
Software  is  the  collection  of  programs  ( groups  of  writ- 
ten instructions)  that  tell  the  computer  what  to  do. 
Each  computer  system  has  an  operating  system  (pro- 
gram) to  manage  scheduling,  input,  computing,  and 
output  of  the  application  programs  it  processes.  An  ap- 
plication program  performs  a  specific  task  such  as  pro- 
ducing a  wildlife  inventory  report  or  moving  wildlife 
inventory  data  to  storage.  Programs  that  perform  com- 
mon procedures  such  as  sort  routines  are  called  utility 
programs.  These  are  "packaged"  programs  that  per- 
form functions  common  to  many  organizations  such  as 
accounting,  statistics,  and  modeling. 

With  the  extensive  advancement  in  computer 
hardware  and  software,  the  distinction  between  the  two 
is  not  always  easy  to  define.  Therefore,  the  term  firm- 
ware is  used  to  refer  to  a  combination  of  hardware  and 
software  used  to  control  the  operations  of  a  computer. 
Firmware  instructions  (microcode)  are  used  to  control 
some  of  the  permanently  wired  circuits  (called  hard- 
wired circuits). 


COMPUTER  SYSTEM 

A  computer  system  consists  of  input/output  units, 
a  processing  unit,  and  secondary  storage  devices.  Data 
are  entered  into  input  units  such  as  card  readers,  mag- 
netic tape  drives,  or  key-entry  terminals.  Input  units 
convert  data  to  a  form  understandable  to  the  computer. 
The  data  are  then  processed  by  a  Central  Processing 
Unit  (CPU).  The  results  of  this  processing  are  known 
as  output.  Output  may  be  in  a  form  understandable 
to  people,  such  as  printed  reports,  or  it  may  remain  in  a 
form,  readable  only  by  computers,  for  input  to  another 
computer,  such  as  data  stored  on  tape.  Output  is  made 
available  to  people  through  printers,  magnetic  tape 
drives,  or  terminals. 


Secondary 
Storage 


Input 


Processing 


Control 


ALU* 


Main 
Storage 


-Output 


Computer  system 


Arithmetic/logic  unit* 


Data  Management 


729 


ture  computers  or  to  be  a  programmer  any  more 
than  one  needs  to  know  how  to  repair  a  transmis- 
sion or  fly  a  helicopter  to  use  those  tools.  Other 
specialists  are  available  to  assist  with  these  tasks.  It  is 
important,  however,  to  recognize  computer  capabili- 
ties and  limitations. 

Modern  biologists  often  have  computers  or  ter- 
minals on  their  desks  or  in  an  adjoining  room.  To 
benefit  from  these,  the  biologists  need  to  learn  how 
to  turn  them  on,  use  program  applications,  and  exe- 
cute certain  commands. 

Many  computer  programs  are  being  designed  to 
be  "user  friendly."  Knowing  complicated  program- 
ming languages  is  no  longer  necessary.  In  many  in- 
stances, specific  steps  are  outlined  in  user  guides. 

The  computer  is  part  of  a  technology  that  is  be- 
coming as  important  to  the  wildlife  biologist  as  other 


common  tools  in  the  natural  resource  field.  The 
wildlifer  can  use  computers  to  assist  in  analyzing 
large  volumes  of  data  and  make  better  decisions 
regarding  wildlife  habitats.  Many  wildlife  manage- 
ment organizations  already  use  computers  to  process 
wildlife  species  or  habitat  data.  To  get  the  informa- 
tion required,  biologists  must  decide — 

•  what  data  to  collect; 

•  how  much  data  to  collect; 

•  how  to  store  and  maintain  the  data; 

•  how  to  process  the  data. 

An  understanding  of  computers  will  affect  these 
decisions. 

Information  processing  remains  a  rapidly  devel- 
oping field,  whether  it  be  hardware,  laser  printers, 


COMPONENTS 

A  computer  consists  of  five  components: 

( 1 )  Input  units 

(2)  Main  storage 

(3)  Arithmetic/logic  unit  (ALU) 

(4)  Control  unit 

(5)  Output  units 

Input  units  are  used  to  enter  data  into  the  com- 
puter. Data  are  usually  entered  through  a  keyboard, 
optical  scanner,  or  other  device.  They  are  transferred  to 
computer  memory  for  processing  and  may  then  be 
stored  on  media  such  as  magnetic  tape  or  disk. 

The  ALU  and  control  unit  make  up  the  central 
processing  unit  (CPU).  These  are  the  electronic  com- 
ponents that  cause  processing  to  occur  by  interpreting 
instructions,  performing  calculations,  moving  data 
around  in  main  computer  storage,  and  controlling  in- 
put/output operations. 

The  control  unit  directs  and  coordinates  the  entire 
computer  system.  This  includes  the  entry  and  removal 
of  data  from  main  storage.  The  control  unit  may  coordi- 
nate such  activities  as  rewinding  a  tape  reel  or  routing 
data  from  main  storage  to  the  ALU. 

The  ALU  performs  arithmetic  and  logic  operations 
on  data.  The  arithmetic  circuitry  performs  calculations 
and  related  functions  such  as  rounding.  The  logic  cir- 
cuitry performs  logical  operations  such  as  comparing 
two  sets  of  characters  or  testing  a  number  to  see  if  it  is 
positive,  negative,  or  zero. 


Main  storage  (also  called  memory  or  internal  stor- 
age) is  frequently  considered  a  part  of  the  CPU.  It  is 
somewhat  like  an  electronic  filing  cabinet  with  address- 
able locations.  Each  location  is  capable  of  holding  data 
or  instructions.  This  allows  for  certain  locations  in  main 
storage  to  be  used  in  specific  ways,  such  as  for  calcula- 
tion and  instruction  interpretation.  When  main  storage 
is  not  considered  part  of  the  CPU,  it  is  referred  to  as 
one  of  the  secondary  or  auxiliary  kinds  of  storage  avail- 
able. 

Output  units  are  used  to  obtain  information  in 
various  forms.  Some  types  of  output  are  produced  as 
printed  reports  or  microfiche.  Output  is  also  displayed 
on  video  screen  terminals. 


DATA  PROCESSING  CYCLE 

The  data  processing  cycle  consists  of  inputting, 
processing,  and  outputting  data,  which  transforms  data 
into  information  that  is  organized  and  usable. 

The  input  step  identifies  the  data  to  be  processed 
and  the  instructions  to  be  used  for  processing.  During 
processing,  data  may  be  checked  for  completeness 
and  accuracy.  If  they  are  not  valid,  the  output  will  have 
no  value.  This  is  the  basis  for  the  term  "garbage  in — 
garbage  out."  Therefore,  at  this  step,  it  is  important  to 
validate  the  data  entered  so  the  user  can  make  neces- 
sary corrections  before  other  processing  is  completed. 

During  the  processing  step,  manipulations  or 
computations  will  be  performed  according  to  instruc- 
tions. For  example.  State  codes  could  be  expanded  from 
an  input  abbreviation  (NV)  to  output  a  full  name  (Ne- 


730 


Data  Management 


computers  small  enough  to  fit  in  a  hand,  or  better 
software  programs.  Few  people  know  what  the  fu- 
ture holds,  but  if  history  is  any  indication,  significant 
new  developments  will  continue  and  will  affect  how 
people  will  do  their  work. 

Because  data  management  is  a  rapidly  changing 
field,  we  have  kept  this  chapter  very  general;  users 
should  keep  up  with  the  latest  references  or  ask 
computer  experts  for  the  most  up-to-date  informa- 
tion. Not  every  wildlifer  needs  to  be  a  statistician  or 
programmer,  but  he  or  she  should  become  familiar 
enough  to  converse  with  statisticians  and  data  pro- 
cessing personnel  to  know  when  more  highly  trained 
assistance  in  these  fields  is  needed. 

This  chapter  presents  some  factors  to  consider 
when  working  with  data  processing  or  information 
management.  We  have  tried  to  keep  the  main  text  as 
free  from  undefined  jargon  as  possible.  Many  special 


topics  and  concepts  are  described  in  sidebars,  and 
special  terms  are  defined  in  the  glossary.  We  have 
not  discussed  particular  program  or  machine  opera- 
tions, as  these  are  best  covered  in  detailed  manuals 
or  user  guides — and  any  discussion  would  soon  be- 
come outdated. 


USING  THE  COMPUTER 

Accurate,  timely  information  is  vital  in  making 
decisions.  Computers  are  often  used  to  process  data 
into  the  information  required.  Computers  are  helpful 
when — 


( 1 )  A  large  volume  of  data  is  to  be  processed; 

(2)  Complex  comparisons  are  involved; 


vada),  or  a  code  for  an  animal  species  (ODHE)  will 
be  output  as  mule  deer.  An  example  of  a  computation 
would  be  the  accumulation  of  each  animal  species  in 
each  State. 

The  output  step  provides  the  processed  informa- 
tion. During  the  data  processing  cycle,  the  data  are 
organized  so  they  will  be  meaningful  to  those  who  re- 
ceive them.  Because  decisionmaking  occurs  in  many 
areas  and  at  many  levels,  one  person's  information  may 
be  another  person's  data. 


Processing 


Information 


Stored 


X 


Printed 


Data  processing  cycle 

To  illustrate  the  processing  of  data,  consider  a 
hypothetical  study  a  wildlife  biologist  might  conduct  to 
determine  the  number  of  mule  deer,  elk,  and  desert 
bighorn  sheep  tagged  for  a  special  study  in  several 
States.  Our  example  contains  only  10  records  so  the 
relationship  between  input  and  output  will  be  easier  to 
see.  We  used  the  following  codes  for  the  input  data: 
NV  for  Nevada,  NM  for  New  Mexico,  and  UT  for  Utah. 
The  animal  species  codes  are  ODHE  for  mule  deer, 


CEEL  for  elk,  and  OVCAME  for  desert  bighorn  sheep. 
The  data  are  as  follows: 


NV 

ODHE 

UT 

CEEL 

NV 

ODHE 

NV 

CEEL 

NM 

CEEL 

NM 

OVCAME 

UT 

OVCAME 

NV 

ODHE 

UT 

CEEL 

UT 

OVCAME 

Only  after  the  above  data  were  processed  would 
they  become  useful  information.  The  output  might 
appear  as  follows: 


Number   o-f   Vertebrate  Species  Tagged  by  State 
as  oi    1/1/84 


State 

Species 

Num 

Nevada 

Mule  Deer 
Elk 

3 

1 

New  Mexico 

Elk 

Desert   Bighorn   Sheep 

1 
1 

Utah 

Elk 

Desert   Bighorn   Sheep 

2 
2 

Notice  that  the  total  for  all  species  is  10,  which  is 
the  same  as  the  number  of  records  input. 


Data  Management 


731 


(3)  Calculations  need  to  be  performed; 

(4)  Reordering  (resequencing)  is  required. 

Any  time  data  are  processed,  the  following 
series  of  functions  must  be  performed,  regardless  of 
whether  the  process  is  performed  manually  or 
automatically: 

( 1 )  Define  required  data  elements, 

(2)  Identify  data  sources, 

(3)  Develop  a  collection  process, 

(4)  Define  where  data  are  to  be  processed, 

(5)  Define  programming  requirements, 

(6)  Define  data  entry  procedures, 

(7)  Define  input  procedures, 


(8)  Identify  processing  requirements, 

(9)  Define  output  procedures. 

When  data  are  processed  by  a  computer,  these 
functions  can  be  performed  by  a  systems  analyst, 
programmer  analyst,  or  wildlife  biologist.  These 
functions  are  discussed  further  in  the  following 
paragraphs. 

Define  Required  Data  Elements 

To  identify  data  required  for  input,  the  biologist 
should  identify  and  analyze  each  item  on  the  output 
document.  One  way  is  to  list  each  item,  then  analyze 
it  to  determine  if  the  data  are  read  in,  calculated,  or 
generated.  Figure  1  provides  an  example  of  an  out- 
put document.  The  data  elements  requiring  analysis 
are  numbered  1  through  7. 


DATA  REPRESENTATION 

Data  must  be  in  a  form  recognizable  to  the  com- 
puter to  be  accepted  by  a  data  processing  system.  Sym- 
bols are  used  to  represent  data.  The  symbols  on  this 
page  are  recognizable  by  people.  They  represent  a 
language.  Electronic  signals  are  the  symbols  used  in 
computer  language.  When  people  communicate  with  a 
computer,  they  must  use  special  devices  to  translate 
people-language  into  machine-language. 

Data  going  into  a  computer  are  recorded  in  var- 
ious ways,  depending  on  the  type  of  input  device  being 
used.  They  can  be  recorded  as  punched  holes  on  input 
media  such  as  cards,  magnetic  spots  on  magnetic  tape 
and  disk,  characters  and  lines  on  paper  documents,  and 
so  on.  Not  all  data  are  recorded  on  an  intermediate 
input  media  such  as  cards,  magnetic  tape,  or  disk.  On 
some  terminals,  they  may  go  directly  to  main  storage 
with  no  input  medium. 


When  data  are  entered  through  an  input  device, 
such  as  a  card  reader  or  a  terminal,  the  device  trans- 
lates each  character  of  input  into  a  series  of  electronic 
signals.  A  signal,  called  a  bit,  is  in  one  of  two  states — on 
or  off,  just  like  a  light  bulb.  When  individual  signals  are 
grouped  together  as  a  set  of  signals,  they  are  called  a 
byte.  Each  byte  represents  an  addressable  location 
in  the  computer's  main  storage.  When  data  have  been 
transferred  to  a  location  in  main  storage,  they  can  be 
located  and  used  in  calculations  or  manipulations  as 
necessary. 

A  character  also  is  represented  by  one  byte.  Each 
character  has  a  different  set  of  bits.  One  character  is 
different  from  another  according  to  the  bits  which  are 
on  or  off  within  its  set.  The  combination  of  bits  and  the 


number  of  bits  used  to  represent  a  character  make  up 
a  code.  Because  codes  used  by  computers  vary,  a  code 
set  is  used  to  convert  data  from  one  form  of  representa- 
tion to  another.  Codes  that  represent  alphabetic,  nu- 
meric, and  special  characters  form  a  code  set.  Examples 
of  commonly  used  code  sets  are  Extended  Binary 
Coded  Decimal  Interchange  Code  (EBCDIC), 
American  Standard  Code  for  Information  Inter- 
change (ASCII),  and  Binary  Coded  Decimal  (BCD). 


Codes  are  also  used  to  indicate  the  end  of  data. 
The  computer  accepts  all  these  signals  and  stores  them 
in  main  storage.  They  can  then  be  referenced  and  ma- 
nipulated within  the  processing  unit.  When  processing 
is  completed,  the  electronic  signals  can  be  converted 
back  into  characters  readable  by  people  through  output 
devices.  Because  of  the  similarity  between  the  input 
and  output  functions,  some  devices  are  designed  to  per- 
form both  functions. 


The  above  discussion  refers  to  code  sets  used  for 
character  representation  and  should  not  be  confused 
with  coding  of  computer  programs. 


DATA  STRUCTURES 

The  data  used  during  the  execution  of  a  program 
are  contained  in  a  data  structure  known  as  a  file.  Data 
must  be  organized  in  such  a  way  that  they  can  be  iden- 
tified and  located  during  the  execution  of  a  program. 
The  smallest  data  structure  is  a  character  (byte).  A 
group  of  characters  is  a  data  element;  a  group  of  data 
elements  is  a  record;  and  a  group  of  records  is  referred 


732 


Data  Management 


Stream 

River 

River 

Segment 

Mean 

Rip 

arian 

Riparian 

Name 

Mile  (St 

art) 

Mile  (End) 

Lenqth  (mi ) 

m 

dth 

(ft) 

Acreaoe 

North  Creek 

7.2 

9.4 

2.2 

20 

11.1 

South  Creek 

0.0 

5.0 

5.0 

35 

42.4 

East  Creek 

4.1 

4.3 

0.2 

40 

1.9 

West  Creek 

3.7 

4.7 

1.0 

25 

6.1 

TOTALS 

8.4 

61.5 

Figure  1.     Sample  output  document. 


to  as  a  file.  For  example,  a  person's  name,  consisting 
of  a  specific  number  of  characters,  could  represent 
a  data  element.  When  several  data  elements  containing 
information  about  the  person  are  collected,  a  record 
is  formed.  The  data  elements  in  the  record  could  con- 
tain the  State  codes  and  species  codes.  Several  records 
with  different  data  create  a  file.  A  file  may  be  located  on 
any  media  such  as  cards,  tape,  or  disk. 


Each  field  within  a  record  contains  the  same  num- 
ber of  characters  as  the  corresponding  field  on  another 
record.  Because  a  space  is  a  valid  character,  some  fields 
may  contain  more  spaces  than  others.  For  example,  if 
a  field  contains  10  characters,  the  data  "ABCD"  would 
use  4  characters  and  6  spaces  to  complete  the  field. 


MAIN  STORAGE  CAPACITY 

A  computer's  main  storage  capacity  determines 
the  amount  of  data  and  instructions  that  can  be  held 
in  the  system  at  any  one  time.  The  more  main  storage 
available,  the  faster  storage  and  retrieval  of  data  can 
occur.  Although  it  would  be  desirable  to  have  as  much 
main  storage  as  possible,  it  is  expensive,  and  cost  is 
often  a  limiting  factor. 


Computer  storage  is  measured  in  Ks;  one  K  repre- 
sents 1,024  bytes  (characters).  Computer  main  storage 
sizes  range  from  as  little  as  4K  on  a  microcomputer 
to  as  large  as  16.000K  (16  million  bytes)  on  a  large 
computer.  One  million  bytes  is  called  a  megabyte. 


APPLICATION  PROGRAMS 

Application  programs  perform  specific  user 
functions  such  as  payroll,  wildlife  inventory,  or  material 
sales.  Many  programming  languages  are  available.  Three 
of  the  most  commonly  used  are  COBOL  (Common 
Business  Oriented  Language),  FORTRAN  (Formula 
Translator),  and  BASIC  (Beginner's  All-Purpose 
Symbolic  Instruction  Code).  These  languages  make 
essential  and  difficult  computer  programming  tasks 


The  sequence  of  instructions  that  make  up  the 
program  is  read  into  the  computer's  main  storage.  The 
instructions  are  processed  in  sequential  order  and  the 
results  produced.  Once  the  program  has  begun  to  pro- 
cess, it  will  continue,  without  any  human  intervention, 
until  completed.  The  instructions  will  be  processed 
exactly  as  the  program  indicated. 


PACKAGED  PROGRAMS 

An  application  program  may  be  a  generalized,  pre 
written  program  that  performs  a  function  requested 
by  the  user.  This  type  of  program,  known  as  a  pack- 
aged program,  is  increasingly  available  to  the  user. 
Many  such  programs  are  available  for  business  and 
scientific  functions.  They  usually  contain  alternatives 
that  users  may  select  for  their  particular  situation. 
Sometimes  these  programs  require  modifications  to 
meet  the  users'  specific  requirements.  The  cost  for  a 
packaged  program  is  often  less  than  having  a  specific 
program  written.  Consideration  must  be  given  to  the 
computer  system  the  program  will  be  executed  on. 


Data  Management 


733 


Identify  Data  Sources 


Develop  a  Collection  Process 


An  analysis  of  the  data  elements  indicates  that 
data  elements  1,  2,  3,  and  5  are  read  in  from  a  file. 
Data  element  4  is  calculated  by  subtracting  data 
element  2  from  data  element  3-  Data  element  6  is 
calculated  by  converting  data  element  4  to  feet,  cal- 
culating the  area,  and  converting  the  results  to  acres. 
Data  element  7  is  an  accumulation  of  the  indicated 
data  elements. 


The  source  of  the  data  to  be  read  in  must  be 
identified.  The  data  may  be  in  a  form  ready  for  input 
to  the  computer,  such  as  a  file  located  on  disk  or 
magnetic  tape.  The  data  may  be  located  on  source 
documents,  such  as  field  observation  forms,  but  the 
data  must  be  converted  for  input  to  the  computer.  If 
the  data  are  not  readily  available,  arrangements  must 
be  made  to  collect  them. 


A  method  for  collecting  data  must  be  estab- 
lished. A  special  form  may  need  to  be  designed  for 
this  purpose.  Careful  consideration  should  be  given 
to  the  design  of  data  entry  forms. 


Define  Where  Data  are  to  be  Processed 

When  data  are  to  be  processed  manually,  ar- 
rangements must  be  made  as  to  where  they  will  be 
processed  and  who  will  do  the  processing.  If  a  com- 
puter is  to  process  the  data,  arrangements  must  be 
made  with  the  computer  facility  for  processing.  Be- 
cause of  the  various  sizes  and  capabilities  of  com- 
puters, it  is  important  to  know  that  the  computer  to 
be  used  is  capable  of  processing  the  job.  Where  the 
computer  is  located  will  determine  the  method  used 
in  making  the  input  data  available. 


READERS 

Readers  perform  the  input  function  in  the  data 
processing  cycle.  They  read  data  into  the  computer's 
main  storage  where  it  will  be  available  for  processing. 
The  devices  discussed  here  are  only  capable  of  reading 
data.  Punched  cards  were  once  the  most  important 
medium  for  data  entry.  In  some  locations  they  are  still 
used  despite  their  disadvantages  of  slow  input,  suscepti- 
bility to  operator  error,  and  machine  malfunctions. 
Because  of  their  disadvantages,  cards  are  being  replaced 
as  an  input  media  by  direct  entry  to  an  electronic  stor- 
age medium. 

Another  type  of  reader  is  the  paper-tape  reader. 
This  device  senses  data  as  holes  punched  in  paper  tape. 
Another  reader,  the  magnetic-ink  character  reader, 
translates  magnetic-ink  characters  into  electronic  sig- 
nals. It  is  widely  used  by  the  banking  industry.  Optical- 
character-recognition  devices  are  often  used  to  scan 
or  read  special  marks,  bars,  letters,  numbers,  and  char- 
acters from  special  printed  documents  which  are  then 
converted  to  electronic  signals  for  computer  process- 
ing. 

To  transfer  graphics,  drawings  or  images  such  as  a 
map  to  the  computer,  a  digitizer  is  used.  A  cursor 
or  some  other  reading  mechanism  that  can  be  passed 
over  the  surface  converts  the  image  to  digital  data. 
These  data  can  then  be  displayed  on  a  cathode  ray 
tube  (CRT)  screen  or  processed  by  a  computer  system. 


INPUT  CONSIDERATIONS 

Input  data  arc  often  preprocessed  or  edited  to 
ensure  their  validity.  Because  a  process  for  updating  ac- 


counts receivable  will  have  different  edit  requirements 
than  a  wildlife  inventory  update,  there  will  be  different 
edit  criteria  for  each.  However,  the  following  are  gen- 
eral considerations  that  are  usually  included  in  an  edit: 


( 1 )  Tests  to  ensure  numeric  data  are  numbers  and 
alpha  data  are  characters. 

(2)  Tests  for  reasonableness  of  the  data.  For  exam- 
ple, the  months  in  a  year  should  not  exceed  1 2 
or  be  less  than  0. 

(  3  )     Range  tests  to  see  if  data,  such  as  acreage  and 
herd  numbers,  fall  within  a  range  of  acceptable 
values. 

(4)  Checklists  of  acceptable  values  to  see  if  data  are 
valid,  such  as  correct  species  codes. 

(5)  Tests  for  data  consistency.  The  data  entered 
may  be  valid  by  themselves,  but  invalid  in  rela- 
tion to  other  data  entered.  For  example,  when  a 
wildlife  inventory  record  has  an  entry  for  a 
hypothetical  occurrence  and  a  verified  occur- 
rence, only  one  may  be  entered.  An  entry  for 
both,  or  blanks  for  both,  indicates  error. 

(6)  Checks  for  transcription  and  transposition  er- 
rors. These  errors  usually  occur  when  the  data 
are  being  key-entered. 


The  instructions  performed  during  edit  processing 
are  included  in  a  computer  program.  When  input  data 
are  processed,  each  record  is  checked  using  the  in- 
struction in  the  program.  Any  invalid  data  are  identified 
so  that  they  may  be  corrected.  The  process  may  be 
repeated  as  many  times  as  necessary  to  obtain  valid 
data. 


754 


Data  Management 


Define  Programming  Requirements 

Each  processing  step  must  be  performed  specifi- 
cally and  in  the  correct  sequence.  Because  a  com- 
puter can  only  perform  calculations,  comparisons, 
and  data  movements,  all  steps  must  consist  of  these 
types  of  instructions.  The  computer  will  process 
these  instructions  in  sequential  order.  For  example, 
if  data  are  to  be  calculated,  they  must  be  read  in 
before  the  calculation  instruction  is  processed. 

Programming  is  not  as  simple  as  it  first  appears, 
since  every  instruction  has  to  be  given  to  the  com- 
puter. However,  once  these  instructions  have  been 
prepared  by  a  programmer  and  stored  on  storage 
media,  the  user  needs  only  to  issue  a  few  commands 
and  the  instructions  will  be  executed.  Programs  to 
do  these  calculations  may  already  be  available  in 
some  agencies  or  elsewhere.  Furthermore,  many  soft- 
ware packages  for  common  tasks  are  available,  par- 


ticularly for  microcomputers.  These  include 
spreadsheet  and  data  base  management  software. 
Such  software  requires  simple  commands  to  set  up 
and  perform  tasks,  but  does  not  require  knowledge 
of  a  programming  language.  Many  biologists  use  such 
software  without  assistance  from  a  computer  pro- 
grammer or  specialist. 

Define  Data  Entry  Procedures 

The  data  to  be  processed  must  be  prepared  for 
input.  Data  may  need  to  be  transferred  to  a  data 
entry  form  if  it  is  not  on  a  document  that  can  be 
directly  used  for  data  entry.  If  a  form  is  required,  the 
person  filling  out  the  form  and  the  method  of  enter- 
ing the  data  to  be  used  should  be  identified. 

Define  Input  Procedures 

Data  must  be  converted  to  a  form  usable  by  the 


METHODS  OF  PROCESSING 

One  of  several  tasks  performed  by  the  computer 
when  processing  data  is  to  interpret  instructions.  An- 
other is  to  set  up  a  connection  to  the  input  or  output 
devices  indicated  in  the  instructions.  The  total  tasks 
required  to  process  a  given  set  of  data  are  known  as  a 
job. 

Two  basic  processing  methods  are  used  by  a  com- 
puter system:  batch  processing  and  interactive  pro- 
cessing. In  batch  processing,  data  are  collected  into 
groups  (batches)  and  stored  before  being  processed. 
Batch  jobs  are  scheduled  and  prioritized  by  the  com- 
puter operators  or  the  computer's  operating  system, 
depending  on  the  installation.  They  may  be  deferred  to 
complete  the  processing  when  computer  workloads 
are  lower. 

In  interactive  processing,  data  are  entered  directly 
into  the  computer  and  processed  immediately.  The 
user  interacts  with  the  program  while  it  is  processing, 
providing  input  and  receiving  output  on  a  terminal. 


OUTPUT  CONSIDERATIONS 

A  printed  report  is  one  of  the  main  types  of  out- 
put. It  may  be  produced  on  various  output  devices  and 
media.  Printers  can  produce  reports  on  a  variety  of 
forms  or  on  microfiche.  There  are  three  major  cate- 
gories of  reports:  detail  reports,  summary  reports, 
and  exception  reports. 

Detail  reports  imply  that  each  input  record  is  read 
to  determine  if  it  will  be  printed  on  the  report.  A  sum- 


mary report  provides  a  summary  of  the  input  data.  It 
is  usually  used  by  management  for  making  decisions. 


Exception  reports  contain  information  determined 
to  be  an  "exception"  or  other  than  normal.  They  are 
produced  when  specific  information  is  needed.  Their 
advantage  is  that  they  save  time  and  money.  Large  re- 
ports that  required  many  hours  to  produce  and  are 
printed  on  many  pages  do  not  have  to  be  produced 
when  the  required  information  can  be  produced  in  a 
few  minutes  on  a  few  sheets  of  paper. 


INQUIRY 

Output  from  interactive  processing  is  usually  ob- 
tained by  inquiry.  An  inquiry  is  a  request  from  the 
terminal  operator  to  a  computer  system  for  information. 
This  method  provides  the  user  with  the  most  current 
information  in  the  shortest  period  of  time.  Often  a 
printed  copy  of  the  information  on  the  terminal  screen 
can  be  obtained.  Usually  the  screen  will  prompt  the 
user  for  entries  and  provide  status  information.  Inquiry 
systems  are  usually  designed  for  a  wide  variety  of  users 
and  are  easy  to  use.  A  well-designed  inquiry  system 
will  provide  prompts  to  guide  the  user.  It  will  acknowl- 
edge all  the  user's  entries  and  respond  almost  immedi- 
ately. A  wildlife  biologist  might  use  this  type  of  system 
to  determine  the  number  and  type  of  species  in  a  cer- 
tain location.  This  information  can  be  requested  and 
received  from  the  computer  system  by  the  user  without 
the  need  for  a  specific  program  to  be  written.  Often 
programs  have  already  been  written  and  are  available  to 
assist  the  user  in  manipulating  or  extracting  information 
from  their  data. 


Data  Management 


735 


computer.  Various  input  methods  are  available  to 
accomplish  this.  The  amount  of  data,  the  media  the 
data  are  on,  and  the  organization  of  the  data  (fields, 
records)  are  factors  that  should  be  considered.  If 
input  is  directly  through  a  terminal,  computer  access 
procedures  must  be  established. 

Identify  Processing  Requirements 

The  manipulations  that  must  be  performed  to 
convert  the  data  to  information  must  be  identified. 
All  functions  must  be  specified  precisely  and  the 
sequence  in  which  they  are  to  be  performed  must 
be  identified  exactly.  These  are  the  instructions 
that  will  be  followed  during  the  execution  of  the 
program. 

Define  Output  Procedures 

The  method  of  distributing  the  output  and  the 


type  of  output  must  be  identified.  Output  may  be  in 
several  forms.  Due  consideration  should  be  given 
to  the  types  available.  Future  use  of  the  output 
should  also  be  considered. 


SYSTEM  LIFE  CYCLE 

Just  as  computer  technology  has  evolved  from 
vacuum  tubes  to  semiconductor  chips,  so  have  meth- 
ods for  developing  systems.  The  first  computer  appli- 
cations, such  as  payroll,  were  relatively  simple. 
Problems  in  developing  systems  began  to  appear 
when  applications  such  as  wildlife  were  put  on  com- 
puters. These  systems  had  a  wide  variety  of  data 
and  complex  relationships.  System  failures  and  lack 
of  structure  prompted  managers  to  emphasize  con- 
trol of  the  development  process.  During  the  1970s, 
computer  specialists  increasingly  employed  the  sys- 


PRINTERS  AND  PLOTTERS 

Printers  convert  electronic  signals  received  from 
the  computer  system's  processor  to  letters  and  words 
on  paper  (hard  copy)  that  can  be  understood.  Nearly 
every  computer  system  has  a  printer;  many  have  sev- 
eral. Because  printing  requirements  vary  from  user 
to  user,  many  things  must  be  considered  when  select- 
ing and  installing  a  printer.  Some  considerations  are  the 
volume  of  printing,  the  number  of  multiple  copies 
required,  the  use  of  special  forms,  and  the  print  quality 
required. 


One  way  to  classify  printers  is  by  the  manner  in 
which  they  print  information.  Serial  printers  print  one 
character  at  a  time;  line  printers  print  one  line  of  infor- 
mation at  a  time;  and  page  printers  print  a  complete 
page  of  information  at  a  time.  Most  computer  sheets 
(11  x  14  in.)  contain  66  horizontal  lines  printed  6  lines 
per  inch  with  132  characters  in  each  line. 


Another  way  to  classify  printers  is  by  the  way 
information  is  printed.  When  the  printing  elements  are 
pressed  (impacted)  against  the  paper,  the  method  is 
known  as  impact  printing.  Nonimpact  printers  use  ther- 
mal (heat),  chemical,  electrical,  or  optical  techniques 
to  form  images. 


The  way  a  printer  forms  characters  is  another 
method  of  classification.  A  solid  character  is  like  the 
characters  used  by  a  typewriter.  Dot-matrix  characters 
are  formed  with  a  series  of  dots.  Impact  printers  use 
solid  or  dot-matrix  characters.  Nonimpact  printers  use 
only  dot-matrix  techniques. 


Plotters  are  output  devices  that  convert  numerical 
information  to  graphic  form.  The  output  is  produced 
by  the  plotter  through  use  of  pens  or  pencils  or  by 
electrostatic  means.  Many  types  of  statistical  informa- 
tion can  be  graphically  illustrated  by  using  a  plotter. 


MAGNETIC  TAPE 

Magnetic  tape  units  are  used  for  both  input  and 
output.  They  are  capable  of  reading  and  writing  on 
magnetic  tape.  They  can  read  the  equivalent  of  300,000 
punch  cards  a  minute. 


The  typical  magnetic  tape  reel  contains  2,400  ft  of 
0.5-in.,  oxide-coated  mylar  tape.  Data  are  recorded  on 
the  tape  as  magnetized  spots  in  parallel  channels  or 
tracks  along  the  length  of  the  tape.  The  density  or  num- 
ber of  characters  that  can  be  written  on  a  tape  varies 
from  800  bytes  per  inch  (bpi)  to  6,250  bpi.  The  equiv- 
alent of  1.5  million  fully  punched  cards  can  be  stored 
on  one  reel  of  tape. 


Magnetic  tapes  provide  a  compact,  relatively  inex- 
pensive media  for  retaining  data.  They  require  protec- 
tion from  dust,  humidity,  and  temperature  fluctuation. 
To  assure  that  data  stored  on  a  tape  remain  readable, 
they  should  be  transferred  (copied)  to  another  tape 
once  a  year.  It  is  not  always  necessary  to  use  a  new 
tape.  The  data  on  a  tape  can  be  erased,  making  it  possi- 
ble for  the  tape  to  be  reused  many  times. 


736 


Data  Management 


tem  life  cycle  as  a  methodology  for  system  develop- 
ment. It  is  still  widely  used  today. 

Although  it  has  been  developed  for  large-scale 
application,  the  principles  and  processes  are  applica- 
ble to  any  system.  A  wildlife  biologist,  setting  up  a 
system  for  storing  and  retrieving  data  on  annual 
waterfowl,  counts  on  the  office  microcomputer  and 
follows  the  same  basic  procedure. 

The  evolution  of  an  automated  objective,  such 
as  wildlife  inventory,  from  an  initial  idea  to  the  re- 
quired output,  consists  of  several  phases.  These 
phases  (which  may  be  referred  to  by  different  termi- 
nology) consist  of  initiation,  development,  and 
operation. 

During  the  initiation  phase,  the  general  require- 
ments for  meeting  the  objective  are  established. 


The  feasibility  and  cost  of  automating  are  docu- 
mented; necessary  approvals  are  obtained  and  justifi- 
cations prepared;  alternative  processing  procedures 
are  identified;  and  requests  for  computing  resources 
are  made. 


The  development  phase  consists  of  definition, 
design,  programming,  and  testing.  During  the  devel- 
opment phase,  the  user's  requirements  are  defined. 
From  these  requirements,  specifications  are  pre- 
pared, the  system  is  designed,  and  programs  are  writ- 
ten and  tested,  or  appropriate  software  is  located  or 
purchased  and  tested.  Documentation  is  prepared 
and  user  and  operator  guides  provided  as  necessary. 
Testing  results  are  reviewed  by  the  user  to  assure 
the  results  are  adequate.  All  deficiencies  are  identi- 
fied and  corrected.  When  the  results  produced 
meet  the  requirements,  the  development  phase  is 
complete. 


MAGNETIC  DISK  AND  DISKETTE 

Magnetic  disks  and  diskettes  are  often  referred 
to  as  hard  disks  and  floppy  disks.  The  distinction 
refers  to  the  flexibility  of  the  material  from  which  they 
are  made.  In  this  section,  we  refer  to  hard  disks  as  disks 
and  floppy  disks  as  diskettes  or  flexible  disks. 


Magnetic  disk  units  that  read  and  write  on  mag- 
netic disks  are  called  disk  drives.  A  magnetic  disk  is  a 
thin  metal  platter  coated  on  both  sides  with  a  metal 
oxide.  Data  are  recorded  as  magnetized  spots  placed  in 
circles  on  the  surface  of  the  disk.  These  circular  record- 
ing positions  are  called  tracks.  Even  though  the  tracks 
in  the  center  of  a  disk  are  smaller,  they  contain  the 
same  amount  of  data  as  the  outside  tracks.  The  number 
of  tracks  on  one  surface  of  a  disk  varies  from  less  than 
100  on  microcomputers  to  over  800  on  large  main- 
frames. The  two  recording  surfaces  of  a  disk  have  a 
storage  capacity  ranging  from  2  million  to  about  20 
million  characters,  depending  on  the  number  of  charac- 
ters written  on  each  track.  When  several  disks  are  com- 
bined on  a  common  hub,  they  are  known  as  a  disk 
pack. 


There  are  three  categories  of  disk  drives:  ( 1 )  fixed, 
(2)  removable,  and  (3)  sealed.  A  fixed  disk  drive  con- 
tains non-removable  disks.  These  units  provide  large 
amounts  of  quickly  accessible  storage.  A  removable  disk 
drive  requires  a  disk  or  a  disk  pack  to  be  mounted  on 
the  drive's  vertical  shaft  before  the  data  on  the  disk  can 
be  accessed.  The  need  to  increase  data  reliability  re- 
sulted in  the  development  of  a  sealed  cartridge  known 
as  a  Winchester  disk.  The  cartridge  contains  the  disks 
and  other  components  used  in  reading  and  writing 
on  the  disks. 


Another  type  of  disk  drive  reads  and  writes  on 
diskettes  (floppy  disks).  A  floppy  disk  is  sealed  in  a 
square,  plastic  jacket.  The  most  common  sizes  are  8 
inches,  5  1/4  inches,  and  3  inches.  Since  the  slits  in  the 
jacket  expose  the  recording  surface,  users  should  be 
careful  not  to  touch  the  exposed  area;  this  could  dam- 
age the  recording  surface.  Diskettes  provide  convenient, 
low-cost  data-recording  media  that  are  easy  to  handle 
and  can  be  reused. 

Devices  capable  of  storing  large  amounts  of  data 
on  media  such  as  magnetic  tapes  and  disks  are  known 
as  mass  storage  devices.  Other  types  of  mass  storage 
devices  capable  of  performing  both  input  and  output 
functions  include  magnetic  drum  and  data  cartridge. 
These  types  of  storage  are  used  by  large  computer 
systems  because  of  their  efficient  intermediate  storage 
capabilities. 


ACCESS  METHODS 

There  are  two  methods  of  accessing  data:  sequen- 
tial and  direct.  In  sequential  access,  data  are  read  and 
written  in  a  sequential  order.  To  access  data  located 
at  the  end,  all  data  preceding  it  must  be  read.  The  ac- 
cessing procedure  always  starts  at  the  beginning  of  the 
data. 

In  direct  access,  also  called  random  access,  data 
may  be  written  to  or  read  from  any  location  on  the 
medium.  Data  at  the  end  of  a  project  can  be  directly 
accessed.  That  is,  none  of  the  preceding  data  would  be 
accessed.  Devices  capable  of  accessing  data  directly 
are  called  direct-access  storage  devices. 


Data  Management 


737 


The  operation  phase  includes  maintenance,  eval- 
uation, and  modifications,  if  required. 


be  more  involved  for  a  complex  system  than  a  sim- 
ple one. 


The  limitations  of  the  system  life  cycle  have 
sometimes  resulted  in  a  backlog  of  applications  to  be 
placed  on  computers.  Complex  systems  require  a 
year  or  more  to  reach  full  production  if  this  proce- 
dure is  followed.  Many  organizations  use  structured 
analysis  and  design  techniques  and  very  high-level 
languages  to  reduce  the  time  required  to  develop  a 
system.  Using  these  techniques,  a  skeleton  system 
can  be  developed  to  the  point  of  user  involvement 
within  a  short  period  of  time.  Only  a  few  months 
may  then  be  required  to  bring  the  system  to  full 
production.  The  use  of  these  techniques  will  grow  in 
the  future.  The  time  involved  to  develop  a  system 
depends  on  its  complexity.  Such  considerations  as 
identifying  data  elements,  defining  codes  to  be  used, 
and  designing  procedures  for  inputting  data  may 


SUMMARY 

This  chapter  is  not  an  in-depth  discussion  on 
any  of  the  subjects  mentioned.  Additional  informa- 
tion can  be  obtained  from  a  number  of  good  publica- 
tions, such  as  those  listed  in  the  Literature  Cited 
section  at  the  end  of  this  chapter.  Another  good 
source  of  information  is  through  computer  courses 
offered  at  most  colleges  and  universities.  For  infor- 
mation on  particular  applications,  many  organizations 
have  individuals  who  provide  training  and  assistance. 

Computers  are  sophisticated  tools  that  require 
various  amounts  of  training  in  their  use.  The  infor- 
mation they  provide  is  a  very  valuable  resource. 
As  with  other  resources,  information  must  be  man- 


TERMINALS 

Entering  data  directly  into  a  computer's  main 
storage  is  possible  from  a  terminal.  Receiving  output 
(information)  from  a  computer  system  is  also  possible 
at  a  terminal.  A  terminal  may  be  a  single  device,  two  or 
three  connected  devices,  or  a  computer  system  that 
sometimes  acts  like  a  terminal. 

One  of  the  most  commonly  used  terminal  devices 
is  the  cathode  ray  tube  (CRT).  It  consists  of  a  televi- 
sion-like screen  and  a  typewriter-like  keyboard.  Data  are 
entered  through  the  keyboard  by  the  terminal  operator. 
Output  information  is  presented  in  display  form  on 
the  screen. 

Another  popular  terminal  device  is  the  printer 
terminal.  This  device  creates  hard  copy  output  instead 
of  displaying  the  output  information  on  a  screen.  It  may 
or  may  not  have  a  keyboard  for  data  entry.  When  it  is 
used  only  for  receiving  data,  it  is  called  a  receive-only 
(RO)  terminal. 

Terminals  that  do  nothing  more  than  pass  data 
back  and  forth  are  known  as  dumb  terminals.  A  ter- 
minal that  can  perform  limited  processing  functions, 
such  as  editing  data  before  sending  it  to  the  computer, 
is  known  as  an  intelligent  terminal.  In  addition  to  a 
keyboard  and  display  screen  (or  printer),  an  intelligent 
terminal  has  a  processor,  main  storage,  and  input/output 
interfaces.  The  processor  circuitry  may  be  a  micropro- 
cessor. 

Data  may  be  entered  and  sent  to  a  location(s) 
remote  from  the  central  computer  site,  if  desired. 
Through  remote  job  entry  (RJE),  data  to  be  proc- 
essed is  entered  directly  into  the  computer  via  a  re- 
mote terminal.  Processing  occurs  on  the  central 


computer  and  output  is  sent  back  to  one  or  more  re- 
mote terminals. 


DATA  COMMUNICATIONS 

The  transmission  of  data  from  one  location  to 
another  is  known  as  data  communication.  When  the 
transmission  is  over  a  long  distance  by  means  of  tele- 
phone or  telegraph  lines,  microwave,  or  radio,  it  is 
referred  to  as  telecommunications.  Teleprocessing, 
on  the  other  hand,  is  an  activity  that  involves  both 
data  processing  and  telecommunications. 

A  system  in  which  data  are  collected  from  one  or 
more  points  of  origin,  transmitted  to  a  central  location, 
and  processed  to  produce  results  that  are  distributed 
to  one  or  more  points  of  use  is  known  as  a  teleprocess- 
ing system. 

The  device  used  to  convert  electronic  signals  to  a 
form  for  transmission  over  telephone  lines  is  a  modula- 
tor-demodulator, usually  called  a  modem.  It  may  be 
a  separate  unit  or  part  of  a  communication  control  unit. 
Both  the  receiving  end  and  the  sending  end  of  the  line 
require  a  modem  to  convert  signals.  A  special  type  of 
modem  is  known  as  an  acoustic  coupler,  which  uses  a 
standard  office  telephone  and  does  not  require  special 
wiring. 

Data  are  transmitted  over  data  communication 
channels  (also  called  data  lines  or  data  links).  There  are 
several  types  of  data  communication  channels,  includ- 
ing ( 1 )  standard  telephone  lines,  (  2 )  coaxial  cable, 
(3)  microwave  transmission,  (4)  satellite 
communications,  and  (5)  fiber  optics.  The  costs  of 
these  vary;  standard  telephone  lines  are  the  least 
expensive. 


738 


Data  Management 


aged.  Computers  are  one  of  the  tools  that  can  be 
used  to  manage  information  and  other  resources 
more  effectively  and  efficiently. 


Changes  have  taken  place  rapidly  over  the  past 
few  years  and  that  trend  is  expected  to  continue. 
The  real  accomplishments  of  today  may  be  only 
commonplace  tomorrow.  We  believe  the  future 
holds  great  promise  for  the  use  of  data  processing  by 
biologists.  Exact  applications  of  this  technology  are 
hard  to  predict,  however.  The  use  of  hand-held  data 
terminals,  graphics,  video  text,  laser  printing,  and 
fiber  optics  are  just  getting  started.  Many  may  be 
commonplace  in  the  future.  Sophisticated  communi- 
cations technology  will  soon  allow  electronic  mes- 
saging and  networking  to  be  part  of  our  lives.  If 


recent  trends  continue,  hardware  prices  will  de- 
crease, speed  of  computations  will  increase,  size  will 
decrease,  and  computers  will  become  more  "user 
friendly." 

Data  processing  for  financial  and  administrative 
activities  is  simple  and  widespread  compared  with 
programming  for  wildlife  and  other  natural  resource 
fields.  It  is  often  difficult  to  explain  all  of  the  re- 
source interactions  to  the  programmer  who  is  devel- 
oping a  particular  application.  However,  biologists 
should  not  allow  these  complexities  to  intimidate 
them.  Remember,  the  computer  is  just  a  tool.  It  is  a 
powerful  tool  which  can  collect,  store,  and  analyze 
data  to  expand  knowledge  about  the  resources 
we  manage  and  make  better  wildlife  management 
decisions. 


Another  factor  affecting  cost  is  line  speed.  Data  are 
transmitted  in  bits  per  second  (bps).  A  low-speed  de- 
vice transmits  40  to  300  bps.  Although  these  devices 
usually  transmit  over  telegraph  communication  lines, 
standard  telephone  lines  used  for  voice  communication 
(called  "voice-grade"  lines)  can  be  used  (with  special 
modems)  for  full-duplex  transmission  at  speeds  up 
to  2,400  bps.  For  speeds  of  4,800  bps  and  above,  spe- 
cial communications  are  always  required. 


OPERATING  SYSTEM 

Each  computer  has  an  operating  system  that 
consists  of  a  collection  of  software  to  supervise  its  op- 
erations. The  operating  system  is  usually  provided  by 
the  computer  manufacturer.  Usually  changes  cannot  be 
made  to  it  by  the  local  computer  staff.  When  enhance- 
ments are  made  by  the  manufacturer,  a  copy  is  pro- 
vided to  the  user.  Operating  systems  vary  from 
computer  to  computer  even  among  those  of  one  manu- 
facturer. They  are  closely  designed  to  the  architecture 
of  the  computer  hardware. 


Functions  of  the  operating  system  include  super- 
vising the  execution  of  programs,  directing  input/output 
devices,  and  loading  data  into  main  storage.  Most  oper- 
ating systems  have  programs  that  sort  and  merge  data 
and  transfer  data  from  one  media  to  another. 

Operating  systems  are  often  identified  by  acro- 
nyms. For  example,  Disk  Operating  System  (DOS)  is 
a  commonly  used  system  for  microcomputers. 


TIME  SHARING 

Time  sharing  is  a  method  whereby  a  computer  is 
shared  by  several  users  for  different  purposes  at  the 
same  time.  While  in  time  sharing,  a  user  may  interact 


with  a  program  while  it  is  processing,  submit  process- 
ing to  be  performed  later,  or  issue  commands  to  deter- 
mine the  status  of  processing.  Users  appear  to  be  served 
simultaneously  due  to  rapid  electronic  speeds,  although 
service  is  actually  sequential. 


DATA  BASE 

The  term  data  base,  in  a  generic  sense,  refers  to 
all  data  elements  that  will  be  created,  used,  or  manipu- 
lated by  an  automated  data  system.  It  does  not  imply 
any  particular  organization  or  structure  of  the  data,  nor 
that  a  data  base  management  system  (DBMS)  will  be 
used.  The  definition  of  and  the  procedures  for  using  a 
data  base  may  vary  because  of  the  different  structures 
and  organizations. 


The  updating  process  for  a  data  base  provides  for 
changing  the  data  for  only  one  data  element  rather  than 
a  data  element  in  several  files.  The  value  of  the  data 
element  will  be  the  same  on  the  output  from  each 
program. 


A  data  dictionary  or  data  element  dictionary  is  a 
tool  that  lists  all  data  elements,  their  definitions,  how 
and  where  they  are  used,  and  who  is  responsible  for 
them.  The  dictionary  helps  standardize  definitions  of 
each  data  element.  It  is  up  to  user  departments  to  en- 
sure that  the  data  they  need  are  truly  represented  in  an 
organization's  data  base. 


A  major  task  for  many  organizations  is  deciding 
the  data  bases  they  need,  the  best  locations,  the  data 
that  should  be  stored  in  them,  and  the  organization  of 
the  data. 


Data  Management 


739 


LITERATURE  CITED 


BOHL,  M.  1980.  Information  processing.  Sci.  Res.  Associ- 
ates, Inc.,  Chicago,  IL.  492pp. 

.  1981.  An  end-user's  guide  to  data  base.  Prentice- 
Hall,  Inc.,  Englewood  Cliffs,  NJ.  144pp. 

DE  STEIGUER,  J.W.  and  R.H.  GILES,  Jr.  1981.  Introduction 
to  computerized  land-information  systems.  J.  Forestry 
79:734-737. 

LEOPOLD,  A.  1948.  Game  management.  Charles  Scribner's 
Sons.  New  York,  NY.  481pp. 

MAYER,  K.E.  1984.  A  review  of  selected  remote  sensing 
and  computer  technologies  applied  to  wildlife  habitat 


inventories.  California  Fish  and  Game  70:102-112. 
SHELLY,  G.B.  and  T.J.  CASHMAN.  1980.  Introducton  to 

computers  and  data  processing.  Anaheim  Publishing 

Company,  Fullerton,  CA.  337pp. 
SIEGERT,  P.P.  1972.  Systems  and  general  management — a 

rationale.  Am.  Manage.  Assoc.  181pp. 
STEENHOF,  K.  1982.  Use  of  an  automated  geographic 

information  system  by  the  Snake  River  Birds  of  Prey 

Research  Project.  Computer-Environment  Urban  Sys- 
tem 7:245-251. 
STRACKBEIN,  R.  and  D.B.  STRACKBEIN.  1983.  Computers 

and  data  processing  simplified  and  self-taught.  ARCO 

Publishing,  Inc.,  NY.  94pp. 
WINKLER,  C.  1983-  The  computer  careers  handbook. 

ARCO  Publishing,  Inc.,  NY.  142pp. 


740 


Data  Management 


37 

STATISTICAL 
ANALYSIS 

William  H.  West 


U.S.  Bureau  of  Land  Management 
Service  Center 
Denver,  CO  80225 


"Quantification  as  such  has  no  merit  except  insofar 
as  it  helps  to  solve  problems." 

— P.B.  Medawar,  Advice  to  a  Young  Scientist 


Editor's  Note:  A  book  on  inventory  and  monitoring 
of  wildlife  habitat  would  not  be  complete  without 
a  discussion  of  statistics.  Statistics  are  the  basic 
tool  that  biologists  must  use  for  analyzing  quanti- 
tative data  Although  many  specialized  techniques 
have  been  developed  for  analyzing  habitat  data,  in 
many  cases  the  traditional  statistics  are  adequate. 
One  chapter  cannot,  of  course,  substitute  for  ade- 
quate basic  training  in  statistics.  It  does,  however, 
describe  the  use  of  traditional  statistics  as  well 
as  some  of  the  neiv  techniques  such  as  multivariate 
analysis  to  evaluate  habitat. 


INTRODUCTION 

Statistics  should  be  a  standard  tool  of  the  biolo- 
gist working  with  inventory  or  monitoring.  While 
statistics  are  sometimes  confusing  and  difficult  to  un- 
derstand, they  are  procedures  a  resource  manager 
or  biologist  cannot  afford  to  be  without.  Reliable 
data  are  a  must  when  managing  natural  resources  in 
today's  society.  However,  we  simply  cannot  afford  to 
measure  every  square  inch  of  a  resource!  Statistical 
inference  allows  the  biologist  to  measure  enough  of 
a  site  to  make  informed  decisions  based  on  data 
representative  of  the  area. 

Statistics  provide  a  way  of  describing  wildlife 
habitat  as  accurately  as  possible  with  the  least 
amount  of  cost.  A  biologist  can  describe  a  habitat 
with  a  few  factors  and  assign  a  level  of  confidence 
about  these  factors.  This  provides  the  manager  with 
a  good  understanding  of  the  problems  faced  and 
allows  a  more  knowledgeable  decision  to  be  made. 
Finally,  statistics  provide  a  biologist  with  a  means  to 
detect  and  test  the  results  of  management  decisions. 
Results  that  may  be  too  small  to  be  detected  by  vis- 
ual inspection  of  a  site  or  data  can  be  displayed  by 
statistics,  and  variation  from  natural  events,  rather 
than  management  actions,  can  be  separated. 

In  all  cases,  statistics  are  merely  a  tool  to  be 
used  by  a  biologist  or  manager.  Statistics  do  not 
make  decisions;  they  only  provide  managers  with 
information  to  make  decisions.  A  manager  or  biolo- 
gist who  depends  on  statistics  to  make  a  decision 
is  abusing  the  use  and  intent  of  statistical  analysis. 
Only  the  manager  can  weigh  all  available  data,  deter- 
mine the  level  of  risk  he  or  she  is  willing  to  accept, 
and  make  a  decision  based  on  that  information. 

Statistics  are  tools,  and  there  are  special  tools  to 
be  used  for  special  jobs.  No  one  tool  can  be  used 
in  all  situations,  and  it  is  important  to  use  the  right 
tool  for  the  job  at  hand.  A  statistician  can  help  select 
and  use  the  right  tool  and  should  be  consulted  be- 
fore designing  a  study;  after  the  study  has  started 
is  too  late.  Working  with  a  statistician  before  data 
are  collected  will  help  clarify  questions,  result  in  a 
better  study  design,  and  save  time  and  money. 


Statistical  Analysis 


741 


Use  of  the  right  statistical  tool  can  help  support 
or  refute  the  perceptions  of  a  biologist.  What  a  biol- 
ogist sees,  or  thinks  he  or  she  sees,  may  not  actually 
be  occurring.  Natural  variation  can  cloud  the  situa- 
tion and  lead  a  biologist  to  a  wrong  conclusion.  If 
the  right  statistical  tool  is  used,  confusion  can  be 
reduced  or  at  least  accounted  for. 


(3)  Association  between  two  (or  more)  factors 
the  biologist  has  measured; 

(4)  The  effect  on  one  factor  from  changes  to 
other  factors;  and 

(5)  Whether  a  management  action  is  having  the 
desired  effect. 


This  chapter  is  not  intended  to  teach  statistics; 
the  reader  is  assumed  to  have  some  basic  knowl- 
edge. Rather,  the  chapter  is  intended  to  acquaint  the 
reader  with  inventory  and  monitoring  situations 
that  require  or  are  appropriate  for  statistical  analysis 
and  to  supply  the  reader  with  statistical  tools  appli- 
cable to  such  situations.  Basic  statistical  courses 
teach  the  equations,  but  seldom  teach  techniques  of 
using  them.  The  intent  of  this  paper  is  to  help  teach 
how  to  use  the  equations. 


TYPES  OF  STATISTICS 

Generally  statistics  fall  into  two  categories:  de- 
scriptive and  inferential.  Descriptive  statistics  are 
used  to  describe  a  population  or  sample.  In  statistics 
a  population  is  denned  as  the  total  items  of  interest 
that  might  occur  (Box  et  al.  1978).  The  population 
is  defined  by  a  biologist  before  any  testing  is  con- 
ducted. How  a  population  is  defined  depends  on  the 
questions  a  biologist  is  interested  in  answering.  If  a 
biologist  is  interested  in  answering  questions  about 
the  production  of  prairie  falcons  in  Colorado,  then 
the  population  would  be  defined  as  the  number  of 
falcon  eyries  in  Colorado  on  a  certain  date.  The 
numbers  describing  the  population  are  the  parame- 
ters of  the  population  (e.g.,  the  mean  number  of 
falcon  eyries). 

A  description  of  the  population  is  usually  based 
on  a  sample  from  the  population.  The  sample  is  al- 
ways a  subset  of  the  population  under  study.  The 
125  falcon  eyries  surveyed  during  the  summer  of 
1984  would  be  a  sample  of  the  population  defined 
above.  To  help  meet  the  assumptions  made  when 
using  statistical  analysis,  the  population  sample  is 
usually  taken  randomly.  A  random  sample  is  one  in 
which  each  member  of  the  population  has  an  equal 
and  independent  chance  (probability)  of  being  se- 
lected in  any  sample  (Harnett  1982). 

Inferential  statistics,  as  the  name  implies,  help  a 
biologist  infer  or  conclude  something  about  a  popu- 
lation, based  on  a  sample.  Inferential  statistics  allow 
analysis  of  the  following: 


(1)    True  population  parameters  (e.g.,  mean  and 
standard  deviation); 


Natural  variation  in  populations  and  differences  in 
samples  taken  make  it  difficult  to  determine  differ- 
ences or  patterns  just  by  looking  at  raw  data.  Infer- 
ential statistics  help  remove  or  account  for  the 
effects  of  natural  and  sampling  differences,  allowing  a 
biologist  to  look  more  at  the  population  without  all 
the  "noise." 


STEPS  TO  USING  STATISTICS 

Use  of  statistics  as  a  tool  is  most  effective  when 
there  is  a  clear  understanding  of  what  is  to  be  ac- 
complished. Examining  a  data  set  without  any  clearly 
defined  purpose  to  discover  relationships  or  hypoth- 
eses is  known  as  "data  snooping."  Data  snooping  is 
useful,  but  not  the  most  efficient  or  effective  means 
of  examining  data. 

Green  ( 1979)  outlines  10  principles  for  design- 
ing a  study.  These  principles  are  presented  as  a 
guide;  not  all  situations  may  warrant  using  all  10 
steps.  A  biologist  must  choose  the  applicable  princi- 
ples and  use  the  concepts  as  guideposts. 


(1) 
(2) 

(3) 


State  the  question  concisely. 

Take  replicate  samples  within  each  combi- 
nation of  time,  location,  and  any  other  con- 
trolled variables. 


(2)    Population  differences; 


Take  random  samples.  Putting  samples  in 
representative  or  typical  places  is  not  ran- 
dom sampling. 

(4)  Test  whether  a  condition  has  an  effect.  Col- 
lect samples  where  the  condition  is  present 
and  where  it  is  absent,  but  all  else  is  the 
same. 

(5)  Do  preliminary  sampling  to  provide  a  basis 
for  evaluating  sampling  design  and  statistical 
analysis. 

(6)  Verify  that  the  sampling  device  or  method  is 
sampling  the  population  thought  to  be  sam- 
pled. 

(7)  Break  large  areas  into  relatively  homogene- 
ous subareas  and  allocate  samples  to  each  in 
proportion  to  the  size  of  the  subarea. 

(8)  Verify  that  sample  unit  size  is  appropriate  to 
the  size,  densities,  and  spatial  distribution 
of  the  organisms  being  sampled. 


742 


Statistical  Analysis 


(9)  Test  data  to  determine  if  the  error  variation 
is  homogeneous,  normally  distributed,  and 
independent  of  the  mean. 

(10)  Having  chosen  the  best  statistical  method  to 
test  an  hypothesis,  stick  with  the  result 
(Green  1979). 

Of  these  principles,  the  most  important  are  1 
and  10,  although  the  others  should  not  be  ignored. 
Stating  clearly  the  questions  being  asked  will  save 
many  problems  during  the  course  of  the  study. 
Vague  questions  lead  to  vague  answers,  collection  of 
unnecessary  data,  wasted  time  and  money,  bad  con- 
clusions, and  use  of  inappropriate  statistical  tests. 
The  importance  of  clear  questions  cannot  be  over- 
stressed.  A  good  question  is  the  starting  point  or 
foundation  of  all  good  studies,  and  can  lead  to  a 
good  study  design  and  proper  selection  and  use  of 
test  statistics. 

If  the  questions  are  properly  asked,  a  biologist 
can  avoid  discarding  these  data  because  the  results 
were  incorrect  or  unexpected.  A  conclusion  may  be 
undesirable  but  the  results  will  still  be  valid  if  the 
question  was  stated  properly  and  the  correct  test 
chosen.  Data  and  results  should  never  be  discarded 
on  the  basis  that  the  results  and  conclusions  were 
not  what  was  wanted. 


DESCRIPTIVE  STATISTICS 

Populations  can  be  described  in  many  different 
ways.  The  intent  of  any  description  is  to  convey 
an  understanding  of  the  population.  A  population  (or 
a  sample)  can  be  described  to  a  manager  by  merely 
listing  all  data  in  the  population.  For  small  data  sets 
(<10  values),  this  may  be  a  satisfactory  method 
but  listing  becomes  impractical  when  the  number  of 
values  is  large.  Giving  a  manager  a  long  list  of  num- 
bers does  not  convey  much  information.  It  would  be 
much  more  meaningful  and  easier  to  remember  if 
only  one  or  two  values  were  listed  to  describe  the 
population.  The  population  mean  and  standard  devia- 
tion, or  sample  average  and  standard  deviation,  sup- 
ply information  in  a  more  manageable  way. 

Measures  of  Central  Tendency 

Descriptive  statistics  are  used  to  convey  more 
information  in  less  space  than  listing  all  raw  data. 
The  first  statistic  used  to  describe  a  data  set  is  some 
measure  of  the  central  tendency  of  these  data.  Three 
measures  of  central  tendency  are  mode,  median, 
and  mean.  The  mode  is  a  value  in  a  data  set  (sam- 
ple) that  occurs  most  often;  the  median  is  a  middle 
value  that  arranges  data  in  ascending  order;  and  the 
mean  is  a  sum  of  all  values  divided  by  the  total  num- 
ber of  values  (Harnett  1982). 


EXAMPLE  1 


A  biologist  has  collected  data  on  the  number  of 
sagebrush  plants  per  square  meter  (square  yard) 
on  a  single  mesa  top.  A  total  of  20-meter-square  plots 
were  sampled  and  these  following  data  were  collected: 

3,  9,  1,  5,  9,  0,  4,  7,  9,  5,  5,  1,  1,  6,  3,  8,  4,  4,  7,  5 

The  data  value  that  occurs  most  often  (mode)  is 
5.  When  these  data  are  arranged  in  ascending  order, 
the  middle  value  of  the  data  set  (median)  is  also  5 
The  mean  is  4.8  plants  per  m2. 


With  continuous  rather  than  discrete  data,  the 
mode  is  better  calculated  using  groups  or  classes  of 
data  rather  than  each  individual  data  set.  If  raw  data 
are  used,  possibly  only  one  value  would  occur  for 
each  measurement.  By  grouping  these  data,  a  more 
meaningful  mode  can  be  determined.  However,  the 
overall  usefulness  of  the  calculated  mode  depends 
on  how  the  classes  are  constructed. 

The  mode  and  median  are  not  used  as  often  to 
describe  a  data  set  as  the  mean  or  average.  However, 
there  are  situations  when  the  mean  of  a  data  set 
does  not  make  sense,  as  when  working  with  nominal 
data  (functional  groups  for  vegetation  classification, 
i.e.,  shrubs,  half  shrubs,  grasses,  etc. )  or  with  ordinal 
data  in  which  subjects  are  ranked,  i.e.,  suitability  of 
ecosystems  for  elk  habitat  (Krebs  1978).  The  mean 
of  a  population  or  the  average  of  a  sample  can  be 
thought  of  as  the  "point  of  balance  of  these  data, 
analogous  to  the  center  of  gravity"  (Harnett  1982), 
and  is  the  most  commonly  used  statistical  value. 

Measures  of  Dispersion 

The  mean,  mode,  and  median  indicate  the  cen- 
tral tendency  of  these  data  but  not  how  these  data 
are  grouped  around  the  center.  Measures  of  disper- 
sion tell  how  these  data  are  grouped  or  spread 
around  the  center.  The  range  of  these  data  is  a  meas- 
ure of  dispersion,  indicating  the  difference  between 
the  high  and  low  values.  A  common  method  of  de- 
scribing these  data  spreads  is  the  standard  deviation 
or  the  variance  of  these  data.  Each  data  set  can  be 
looked  at  for  differences  from  the  average.  However, 
this  method  of  describing  these  data  has  the  same 
problem  as  discussed  earlier;  it  conveys  little  mean- 
ing and  is  difficult  to  remember.  A  single  summary 
value  describing  the  spread  of  these  data  is  much 
more  satisfactory.  However,  simply  summing  the 
deviations  conveys  even  less  information  than  listing, 
since  the  sum  of  the  deviations  will  always  equal 
zero. 

The  average  of  all  the  squared  deviations  is  the 
variance  of  these  data.  The  square  root  of  the  vari- 
ance is  the  standard  deviation  of  these  data.  The 


Statistical  Analysis 


743 


standard  deviation  can  be  thought  of  as  the  typical 
deviation  around  the  mean  (Harnett  1982)  and  is 
more  convenient  to  work  with  than  the  variance 
because  it  is  in  the  same  units  as  these  original  data. 


EXAMPLE  2 


Using  these  same  data  described  in  the  previous 
example,  the  deviations  of  each  plot  value  from  the 
average  of  the  entire  sample  (4.8  plants  per  m2)  result 
in  the  following: 

-1.8,  4.2,  -3.8,  0.2,  4.2,  -4.8,  -0.8,  2.2,  4.2,  0.2, 
0.2,  -3.8,  -3.8,  1.2,  -1.8,  3.2,  -0.8,  -0.8,  2.2,  0.2 

The  sum  of  these  deviations  is  zero.  The  sample  var- 
iance is  7.9  and  the  standard  deviation  is  2.8. 


Through  repeated  collection  of  continuous  data, 
measurements  tend  to  cluster  around  the  mean 
measurement  in  a  very  predictable  pattern.  This  pat- 
tern is  the  normal  or  bell-shaped  distribution  (see 
Figure  1 ).  This  distribution  is  described  by  two  pa- 
rameters: the  mean  (u.),  describing  the  peak  of  the 
distribution,  and  the  standard  deviation  (cr),  describ- 
ing the  spread  of  these  data  around  the  mean. 

This  distribution  has  useful  properties  for  deal- 
ing with  continuous  measurements.  Approximately 
68%  of  all  data  lies  within  plus  or  minus  one  stand- 
ard deviation  of  the  average,  and  95%  lies  within 
plus  or  minus  two  standard  deviations  of  the  aver- 
age. In  the  above  example,  68%  of  all  data  lies  be- 
tween the  values  2.0  and  7.6.  The  standard  deviation 
and  how  it  describes  a  sample  will  be  important 
when  testing  a  hypothesis  or  making  inferences 
about  the  parent  population. 


fix) 


0. 399/o 


^  0.3413 

0.1359  / 

•-la"* 

*-lo  "• 

V    0.1359 

M-2a        M-o  V  P+o        p+2o 


Figure  1.     Normal  distribution  with  mean  and 
standard  deviation. 


Frequency  Distributions 

While  the  average  and  standard  deviations  are 
the  most  common  methods  of  describing  a  data  set, 
other  descriptors  are  also  useful  in  understanding 
these  data.  By  grouping  these  data  into  equal  inter- 
vals, dividing  the  number  of  observations  in  each 
interval  by  the  total  number  of  observations,  and 
plotting  the  calculated  frequencies,  a  frequency  dis- 
tribution can  be  derived.  The  shape  of  a  frequency 
distribution  helps  explain  data  characteristics.  Attri- 
butes of  these  data  such  as  unimodal  (one  central 
peak),  bimodal  (two  central  peaks),  or  multimodal 
(many  peaks)  can  be  easily  displayed  with  a  fre- 
quency distribution.  The  shape  of  the  distribution  is 
very  sensitive  to  the  grouping  of  these  data. 


EXAMPLE  3 


A  biologist,  working  on  the  biological  assump- 
tion that  the  size  of  a  sagebrush  plant  crown  is 
directly  related  to  the  age  of  the  plant,  collects  aver- 
age crown  diameter  (in  centimeters)  from  25  ran- 
domly selected  plants  from  a  specific  vegetation 
type. 

5,  5,  6,  7,  8,  8,  8,  9,  12,  12,  15,  16,  18,  18, 
22,  24,  25,  25,  25,  26,  26,  27,  28,  29,  29 

To  attain  the  following  frequency  distribution, 
group  these  data  into  2-cm  intervals  and  divide  the 
number  in  each  class  by  the  total  number.  The  sum 
of  the  frequencies  for  all  classes  will  be  1. 


CLASS 


NUMBER      FREQUENCY 


4.50-6.49 

(3) 

0.12 

6.50-8.49 

(4) 

0.16 

8.50-10.49 

(D 

0.04 

10.50-12.49 

(2) 

0.08 

12.50-14.49 

(0) 

0.00 

14.50-16.49 

(2) 

0.08 

16.50-18.49 

(2) 

0.08 

18.50-20.49 

(0) 

0.00 

20.50-22.49 

(D 

0.04 

22.50-24.49 

(D 

0.04 

24.50-26.49 

(5) 

0.20 

26.50-28.49 

(2) 

0.08 

28.50-30.49 

(2) 

0.08 

TOTAL: 


(25) 


1.00 


The  frequency  distribution  shows  two  peaks 
(bimodal)  in  these  data  at  6.50  to  8.49  cm  and 
again  at  24.50  to  26.49  cm.  The  mean  was  calcu- 
lated at  17.32  cm  and  the  standard  deviation  at 
8.68  cm  (Figure  2). 

The  biologist  would  describe  these  data  as 
having  a  mean  diameter  of  17.32  cm  and  a  bimodal 
distribution  with  peaks  at  6.50  to  8.49  cm  and  24.50 
to  26.49  cm.  Based  on  the  biological  assumption 
made  when  these  data  were  collected,  the  biologist 
would  conclude  the  population  has  old  and  young 
plants,  but  few  plants  of  medium  age.  This  distribu- 
tion may  lead  the  biologist  to  ask  why  there  was  this 
type  of  distribution  in  age. 


744  Statistical  Analysis 


INFERENTIAL  STATISTICS 

Inferential  statistics  use  information  from  a  sam- 
ple to  infer  something  about  the  parent  population 
or  to  test  a  theory  or  hypothesis  about  the  popula- 
tion. The  type  of  statistical  test  used  will  depend  on 
the  assumptions  (statistical  assumptions,  rather  than 
biological  assumptions)  made  about  the  population. 

Parametric  statistics  are  used  when  the  popula- 
tion can  be  assumed  to  have  a  known  distribution. 
The  normal  or  bell-shaped  distribution  is  one  that  is 
familiar  to  the  average  user.  Since  standard  deviation 
of  the  population  is  usually  unknown,  a  t-distribution 
based  on  the  sample  is  substituted  for  the  normal 
distribution.  The  t-distribution  is  similar  to  the  nor- 
mal distribution  (Huntsberger  and  Billingsley  1977); 
the  exact  shape  of  the  distribution  depends  on  the 
size  of  the  sample  taken.  The  larger  the  sample  size, 
the  closer  the  t-distribution  will  approach  a  normal 
distribution. 

An  assumption  made  about  the  sample  is  that 
the  sample  items  have  a  normal  distribution  and  are 
independent  of  each  other.  The  selection  of  an  item 
(or  subject)  to  sample  will  not  affect  the  distribution 
of  other  items  in  the  sample  (Box  et  al.  1978). 

To  ensure  the  assumptions  are  met,  it  is  impor- 
tant a  random  sample  from  the  population  be  used. 
Collecting  data  in  representative  or  typical  areas 
does  not  constitute  a  random  sample  (Green  1979). 
The  largest  sample  that  is  economically  feasible 
should  be  collected  so  sample  distribution  more 


FREQUENCY     DISTRIBUTION 

SAGEBRUSH  CROWN   DIAMETER 


FREQUENCY        ( 


5.5       7.5  9.5         11.5      13.5        15.5        17.5      19.5        21.5        23.5      25.5      27.5      29.5 

CROWN    DIAMETER       (CENTIMETERS) 

Figure  2.     Frequency  distribution  of  sagebrush 
crown  measurements. 


closely  approximates  the  population  distribution.  A 
good  discussion  on  the  consequences  of  violating 
these  assumptions  when  using  ecological  data  can  be 
found  in  Green  ( 1979). 

When  no  assumption  can  be  made  about  the 
distribution  of  the  population,  nonparametric  statis- 
tics are  employed.  Nonparametric  statistics  are  used 
to  test  the  assumption  of  a  normal  distribution  and 
to  conduct  statistical  tests  without  making  assump- 
tions about  the  form  of  the  distribution.  Collecting  a 
random  sample  has  more  influence  on  the  results 
of  any  statistical  analysis  than  the  assumption  of  a 
population  with  an  unknown  distribution  (Box  et  al. 
1978).  Nonparametric  statistics  will  be  discussed 
after  some  of  the  common  parametric  statistics. 

Confidence  Intervals 

The  first  parametric  statistic  deals  with  the  infer- 
ence of  the  true  population  mean,  based  on  the  sam- 
ple average  and  standard  deviation  and  desired  level 
of  probability.  This  type  of  statistic  is  known  as  the 
confidence  interval  and  assumes  normal  distribution 
of  population.  The  confidence  interval,  in  effect, 
indicates  the  range  of  potential  values  of  the  popula- 
tion mean,  given  the  level  of  error  a  biologist  is  will- 
ing to  accept. 


EXAMPLE  4 


A  biologist  has  defined  a  population  as  the 
standing  crop  (g/m2)  of  a  particular  pasture  during 
the  week  of  July  4,  1983.  Sixty  plots,  each  1  m2, 
were  sampled,  and  a  sample  average  of  327  g/m2 
and  sample  standard  deviation  of  1 12  g/m2  were 
calculated.  Confidence  intervals  for  the  mean  at  an 
80%  confidence  level  and  a  95%  confidence  level 
were  calculated. 


80%  C.L  on  the  mean 
95%  C.L.  on  the  mean 


308  g/m2  to  345  g/m2 
298  g/m2  to  355  g/m2 


The  biologist  would  conclude  the  population 
mean  lies  somewhere  within  the  ranges  calculated 
with  the  indicated  probability. 


The  confidence  level  expresses  the  probability 
that  a  parameter  falls  within  the  indicated  limits  but, 
more  importantly,  the  level  of  risk  a  biologist  is  will- 
ing to  accept  in  coming  to  a  wrong  conclusion. 
When  a  biologist  says  he  or  she  is  80%  confident  the 
mean  is  within  the  limits,  what  is  actually  being  said 
is  he  or  she  is  willing  to  accept  a  20%  chance  of 
being  wrong.  A  95%  confidence  means  a  5%  chance 
of  being  wrong. 

The  smaller  the  acceptable  risk  of  being  wrong, 
the  larger  the  limits.  This  concept  becomes  impor- 
tant when  testing  hypotheses  using  a  high  confi- 
dence level.  The  higher  the  confidence  level,  the 
more  difficult  to  disprove  or  reject  an  hypothesis.  It 


Statistical  Analysis 


745 


must  also  be  remembered  there  is  a  trade-off  be- 
tween the  size  of  a  sample  and  the  confidence  level 
desired.  A  statistic  calculated  from  a  small  sample 
size  with  a  high  confidence  level  may  not  tell  much 
about  the  population  parameter  being  estimated. 


EXAMPLE  5 


The  biologist  decides  to  collect  only  four  sam- 
ples from  the  population  (Example  4)  and  deter- 
mines the  sample  average  to  be  327  g/m2  and  a 
sample  standard  deviation  of  1 12  g/m  .  He  or  she  is 
interested  in  taking  little  risk  when  estimating  the 
population  mean  and  desires  a  confidence  level  of 
95%  or  99%. 

95%  C.L  on  the  mean  =  148  g/m2  to  505  g/m2 
99%  C.L.  on  the  mean  =   -0.10  g/m2  to  654  g/m2 

The  biologist  has  high  confidence  the  popula- 
tion mean  is  between  the  limits  0.0  g/m2  and  654 
g/m2.  However,  because  of  the  range  of  values, 
nothing  more  is  known  than  when  started.  The  small 
sample  size  coupled  with  the  high  confidence  level 
have  made  the  numbers  meaningless. 


Sample  Size 

To  ensure  the  statistics  will  be  meaningful  for 
the  desired  confidence  level  and  range  of  limits, 
an  adequate  sample  size  should  be  calculated.  Before 
any  sampling  is  done,  a  biologist  must  determine 
the  level  of  risk,  the  maximum  range  of  acceptable 
limits,  and  the  funds  available  for  sampling  (Harnett 
1982).  With  these  conditions  in  mind,  a  biologist 
can  then  collect  a  few  samples  and  determine  the 
sample  size  needed  to  conform  to  the  stated  specifi- 
cations, based  on  the  preliminary  samples.  (See 
Example  6.) 

When  choosing  an  adequate  sample  size,  set 
reasonable  expectations  for  the  results.  Do  not  make 
the  range  of  limits  too  small!  Knowing  population 
values  with  a  very  small  range  and  high  confidence 
level  may  be  desirable,  but  collecting  these  data  may 
be  too  expensive.  In  Example  6,  if  the  biologist  had 
set  a  range  of  10  g/m2  with  a  99%  confidence  level, 
a  total  of  5,300  samples  would  have  to  be  collected. 
Additionally,  because  the  adequate  sample  is  based 
on  a  preliminary  sample,  the  adequate  sample  size 
should  be  recalculated  periodically  throughout  the 
sampling  period  to  determine  if  any  changes  in  the 
adequate  sample  size  have  occurred  from  changes  in 
the  sample  mean  or  standard  deviation. 

There  is  a  balance  between  meaningful  results 
and  the  cost  of  sampling.  It  is  important  to  strive  for 
the  most  meaningful  results  within  the  constraints 
of  time  and  money.  If  the  commitment  of  resources 
is  not  going  to  be  made  to  obtain  meaningful  results, 
then  it  is  better  to  not  sample  at  all,  or  to  reduce 
the  population  size  to  fit  within  the  monetary  con- 


straints. Do  not  reduce  the  limits  of  the  results  be- 
low the  level  where  they  convey  any  information. 
Subjective,  qualitative  information  gathered  by  a  site 
visit,  for  example,  may  be  better  than  quantitative 
information  derived  from  an  inadequate  data  set.  No 
data  are  just  as  good  as  vague  data  and  much  less 
costly.  Example  5  is  a  situation  where  the  cost  of 
data  collection  was  minimal,  but  the  information 
gathered  was  useless. 

Hypothesis  Formulation 

The  next  area  of  inferential  statistics  involves 
testing  of  ideas  or  concepts.  Essential  to  this  process 
is  the  formulation  of  hypotheses.  Two  hypotheses 
are  always  formed  about  the  ideas  being  tested.  The 
null  hypothesis  (H())  is  assumed  to  be  correct  until 
proven  false  by  the  test.  The  alternate  hypothesis 
(Ha)  is  assumed  to  be  correct  in  the  event  the  null 
hypothesis  is  proven  false.  The  null  hypothesis  and 
alternate  hypothesis  are  complementary. 

Green  ( 1979)  cited  some  general  rules  for  for- 
mulating null  hypotheses.  First,  the  null  hypothesis 
should  be  the  simplest  one  possible  and  yet  describe 
the  concept  being  tested.  If  the  hypothesis  is  com- 
plex, the  question  asked  is  too  broad.  Second,  the 
null  hypothesis  must  be  falsifiable;  the  data  collected 
must  be  able  to  disprove  the  concept  (no  amount 
of  data  can  ever  prove  a  concept).  Finally,  the  null 
hypothesis  should  have  the  fewest  number  of  un- 
known explanatory  factors.  If  the  concept  tested  has 
many  unknown  factors  affecting  the  result,  conclu- 
sions will  be  difficult  to  form. 


EXAMPLE  6 


Before  sampling  the  population  described  in 
Example  4,  the  biologist  decides  to  determine  an 
adequate  sample  size  before  completing  the  sam- 
ple. The  decision  has  been  made  to  accept  a  10% 
chance  of  error  with  a  maximum  range  in  the  limits 
of  50  g/m2.  After  collecting  10  samples,  a  sample 
average  of  327  g/m2  and  sample  standard  deviation 
of  112  g/m2  are  calculated. 


Solving  the  equation: 


(2tsr 


w 


Where: 


t    =  the  t-variable  for  the  sample  at  the 

stated  level  of  error, 
s    =  the  standard  deviation  of  the  sample, 
w  =  the  width  of  the  desired  confidence 

interval. 


Based  on  the  preliminary  sample,  the  biologist 
determines  that  67  samples  must  be  collected  to 
meet  the  desired  specifications.  An  additional  57 
samples  are  then  collected. 


746 


Statistical  Analysis 


Hypothesis  formulation  relates  directly  to  the 
questions  a  biologist  is  asking.  If  good  questions  are 
asked,  the  formulation  of  hypotheses  will  be  easier 
than  if  poorly  formed  or  vague  questions  are  asked. 
The  formulation  of  hypotheses  should  be  made  when 
the  questions  are  asked.  This  helps  determine  what 
data  should  be  collected  and  what  tests  conducted 
to  falsify  the  null  hypothesis  and  answer  the 
questions. 

Included  in  the  formulation  of  the  null  hypothe- 
sis should  be  the  probability  level  at  which  the  null 
hypothesis  will  be  rejected  in  favor  of  the  alternate 
hypothesis.  The  alternate  hypotheses  should  also 
indicate  if  all  the  chance  for  error  is  to  be  placed  in 
one  tail  or  both  tails  of  the  probability  distribution 
of  the  sample.  The  chance  for  error  can  be  placed  in 
one  tail  if  there  is  reason  to  believe  there  is  no 
chance  of  the  error  occurring  in  the  opposite  direc- 
tion. When  testing  the  hypothesis  that  standing  crop 
on  a  salt  flat  is  zero,  there  is  no  need  to  test  if  the 
standing  crop  is  less  than  zero.  Any  chance  of  error 
is  placed  on  the  side  of  standing  crop  being  greater 
than  zero. 


t-Test 

The  most  common  statistic  employed  to  test  a 
hypothesis  or  compare  the  means  of  two  samples 
(populations)  is  the  t-test.  The  test  assumes  the  pop- 
ulation has  a  normal  distribution  and  the  sample 
has  a  t-distribution. 


EXAMPLE  7 


A  biologist  has  identified  two  sagebrush  flats  on 
either  side  of  a  small  canyon  (site  A  and  site  B). 
Both  sites  are  influenced  by  the  same  weather  pat- 
terns and  are  on  the  same  soil  formations.  The  biol- 
ogist wants  to  know  whether  the  sites  are  the  same 
and  can  be  managed  as  such.  The  biologist  also 
wants  to  know  if  site  B  has  enough  standing  crop 
(at  least  1,120  g/m2)  to  support  the  deer  herd  using 
the  site. 

The  two  populations  are  defined  as  the  total 
standing  crop  (g/m2)  at  sites  A  and  B  on  the  first 
week  in  August.  The  following  specific  questions 
and  hypothesis  are  formulated: 

(1)  Is  the  mean  standing  crop  of  population  A  the 
same  as  population  B? 

H0:  Mean  A  =  Mean  B 
Ha:  Mean  A  +  Mean  B 


EXAMPLE  7  (concluded) 


(2)  Is  the  mean  standing  crop  of  population  B  at 
least  equal  to  the  minimum  standing  crop 
needed  (1,120  g/m2)  to  support  the  deer  herd 
using  the  site? 

H0:  Mean  B  =  1,120  g/m2 
Ha:  Mean  B  =  1,120  g/m2 

The  biologist  is  willing  to  accept  a  10%  chance 
of  error  of  estimating  the  mean  incorrectly,  with  a 
range  of  60  g/m2.  For  the  first  question,  the  chance 
of  error  will  be  distributed  between  both  tails  of 
the  t-distribution.  For  the  second  question,  the  prob- 
ability will  be  placed  entirely  in  the  lower  portion  of 
the  t-distribution  since  the  biologist  is  only  interested 
in  a  value  less  than  1,120  g/m2. 

Ten  preliminary  samples  were  collected  at  both 
sites.  For  site  A,  the  preliminary  sample  had  an 
average  of  973  g/m2  and  standard  deviation  of  68  g/ 
m2;  from  this  information  an  adequate  sample  size 
of  17  plots  was  calculated.  For  site  B,  with  a  sample 
average  of  1,010  g/m2  and  standard  deviation  of 
107  g/m2,  an  adequate  sample  size  of  43  was  deter- 
mined. From  the  adequate  sample  sizes,  the  final 
average  of  site  A  is  985  g/m2  with  a  standard  devia- 
tion of  70  g/m2;  the  final  average  for  site  B  is  1 ,022 
g/m2  with  a  standard  deviation  of  1 16  g/m2. 

From  these  sample  data,  a  t-test  statistic  of 
-1.22  for  Question  1  was  calculated.  Comparing  this 
value  to  a  table  value  (1.671)  for  the  t-distribution 
of  these  sample  data,  the  biologist  found  the 
chance  of  error  was  not  greater  than  the  stated 
rejection  level.  Since  the  biologist  did  not  reject  the 
null  hypothesis  (H0),  he  or  she  concluded  the  two 
populations  were  the  same,  or  the  two  samples 
were  from  the  same  population. 

For  Question  2,  a  t-test  statistic  of  -5.5  was 
calculated  and  compared  to  a  table  value  of  1.303 
for  the  t-distribution  for  these  sample  data.  The 
calculated  t-test  statistic  was  found  to  exceed  the 
permissible  error  level  (10%)  The  null  hypothesis 
(H0)  was  then  rejected  and  the  alternate  hypothesis 
(Ha)  accepted.  The  biologist  concluded  the  standing 
crop  at  site  B  was  not  sufficient  to  supply  the  mini- 
mum needs  of  the  deer  herd  at  1,120  g/m2. 


Analysis  of  Variance 

The  t-test  is  useful  when  comparing  two  sam- 
ples or  when  comparing  a  sample  to  some  hypothe- 
sized value.  However,  when  more  than  two  samples 
(populations)  are  compared,  an  Analysis  of  Variance 
(ANOVA)  table  is  used.  The  ANOVA  table  accounts 
for  the  variation  of  different  factors  of  many  popula- 
tions simultaneously  and  is  very  useful  when  com- 
paring the  effectiveness  of  different  management 
treatments.  The  table  can  separate  variation  between 
treatments  and  within  treatments  so  a  biologist  can 
more  directly  evaluate  the  response  of  the  site  to  the 


Statistical  Analysis 


747 


management  actions.  The  ANOVA  table  provides 
information  for  the  calculation  of  statistics,  such  as 
the  F-statistic,  that  can  be  used  to  compare  popula- 
tions or  test  hypotheses. 

Construction  of  an  ANOVA  table  is  flexible  and 
can  be  built  to  account  for  environmental  conditions 
and  the  experimental  design  used.  By  careful  experi- 
mental design,  a  biologist  can  remove  (account  for) 
some  of  the  variability  in  these  data  and  examine 
only  the  factors  of  interest.  A  good  experimental  de- 
sign is  the  key  to  using  an  ANOVA  table,  and  a  statis- 
tician should  be  consulted  at  the  very  beginning  of 
the  test  to  ensure  the  design  is  adequate.  Major  fac- 
tors affecting  the  experimental  design  are  the  biolog- 
ical and  environmental  conditions  present.  A 
biologist  must  have  a  good  understanding  of  these 
factors  when  the  statistician  is  consulted.  Differences 
in  soils,  climate,  or  species  being  studied  might  all 
prompt  changes  in  the  design  of  the  experiment  and 
must  be  considered  prior  to  starting  the  study. 


EXAMPLE  8 


Two  herbicide  dealers  want  a  biologist  to  use 
their  products  in  a  proposed  spraying  of  tamarix 
plants  to  reduce  plant  cover.  Both  herbicides  are 
applied  to  the  soil  and  are  absorbed  through  the 
plant  roots.  Before  buying  either  product,  the  biolo- 
gist wishes  to  determine  which  of  the  two  sprays 
is  more  effective. 

A  small  hillside  is  used  in  a  test  of  the  two 
sprays.  The  biologist  notes  a  difference  in  the  soil 
type,  depth,  and  available  moisture  at  the  bottom, 
middle,  and  top  of  the  slope;  all  other  environmental 
factors  (to  the  best  of  the  biologist's  knowledge) 
are  the  same.  Because  of  these  differences,  the  bi- 
ologist feels  these  changes  must  be  accounted 
for  in  the  experimental  design.  The  design  is,  there- 
fore, blocked  (partitioned)  based  on  the  position  of 
each  test  plot  on  the  slope.  One-ha  plots  (selected 
at  random)  at  the  top,  middle,  and  bottom  of  the 
slope  are  sprayed  with  Herbicide  A;  the  other  similar 
plots  are  sprayed  with  Herbicide  B.  Control  plots  of 
untreated  tamarix  are  also  located  at  each  position 
on  the  slope  to  compare  with  the  treated  plots 

The  population  the  biologist  is  interested  in  is 
defined  as  the  cm2  of  ground  covered  by  the  aerial 
portions  of  the  tamarix  plants  per  square  meter  of 
ground  area,  1  month  after  spraying  has  occurred. 
The  following  specific  questions  are  to  be  answered 
by  the  study: 

(1)  Is  there  any  difference  in  the  effectiveness  of 
the  three  herbicides? 

H0:  Mean  A  =  Mean  B  =  Mean  C 
Ha:  At  least  two  means  are  not  equal 

Where:  Mean  A  =  the  plots  treated  with 
Herbicide  A. 


EXAMPLE  8  (continued) 


Mean  B  =  the  plots  treated  with 
Herbicide  B. 

Mean  C  =  the  untreated  plots. 

(2)  Is  there  any  difference  in  the  cover  of  tamarix 
at  each  position  on  the  slope  after  treatment? 

H0:  Mean  L  =  Mean  M  =  Mean  T 
Ha:  At  least  two  means  are  not  equal 

Where:  Mean  L  =  the  plots  at  the  lower  level  of  the 
slope. 

Mean  M  =  the  plots  in  the  middle  of  the 
slope. 

Mean  T  =  the  plots  near  the  top  of  the 
slope. 

The  biologist  determined  a  10%  chance  of  error 
is  acceptable  and  set  a  90%  confidence  level  for 
the  test.  One  month  after  spraying  the  following 
results  were  obtained: 


SLOPE 


The  following  ANOVA  table  was  constructed 
from  these  data: 


TREATMENT 

A 

B        C 

T 

37 

38       36 

M 

88 

76       81 

L 

51 

42       47 

SOURCE  OF 

SUM  OF 

MEAN 

F 

VARIATION 

DF* 

SQUARES 

SQUARE 

STATISTIC 

Block  (position  on 

slope) 

2 

3,313.56 

1,656.78 

139  71 

Treatment 

(herbicide) 

2 

67  56 

33  78 

2.83 

Residual 

4 

47.78 

11.94 

TOTAL: 

8 

3,428  90 

N/A 

N/A 

"DF:  Degrees  of  Freedom 

Comparing  the  F-statistic  for  treatments  calcu- 
lated from  these  data  (2.83)  to  the  table  value  (4.32) 
for  the  F-distribution,  the  null  hypothesis  (HC))  for 
the  first  question  cannot  be  rejected. 

The  biologist  would  thus  conclude  that  neither 
treatment  is  effective  in  reducing  the  cover  of  tama- 
rix plants,  since  treatment  with  Herbicide  A  or  B 
did  not  significantly  reduce  the  tamarix  cover  below 
that  of  the  untreated  plots.  The  ANOVA  table  does 
not  tell  the  biologist  why  the  herbicides  are  ineffec- 
tive, only  that  there  is  no  difference  in  the  tamarix 
cover  of  the  plots.  If  the  biologist  had  found  signifi- 
cant differences  in  the  treatments,  he  or  she  could 
then  go  on  to  test  and  examine  which  treatments 
were  different. 


748 


Statistical  Analysis 


EXAMPLE  8  (concluded) 


For  the  blocks  (position  on  slope)  the  calculated 
value  (139.71)  exceeds  the  table  value  (4.32)  and 
the  null  hypothesis  (H0)  is  rejected.  Thus,  the  biolo- 
gist would  conclude  that  there  are  significant  differ- 
ences in  cover  of  tamarix  at  different  positions  on 
the  slope.  The  conclusion  would  also  be  reached 
that  the  decision  to  block  the  experiment  on  the 
basis  of  the  position  of  the  plot  on  the  slope  was 
appropriate,  as  significant  differences  were  found  in 
the  tamarix  cover  on  the  bottom,  middle,  and  top 
of  the  slope. 


In  addition  to  determining  differences  between 
treatments,  the  ANOVA  table  can  also  evaluate  inter- 
actions of  treatments  when  applying  two  or  more 
treatments  to  plots  in  combination.  This  use  of  the 
ANOVA  table  enables  a  biologist  to  examine  each 
kind  of  treatment  individually  and  in  combination 
to  determine  differences  and  interactions  of  the 
treatments. 

If  in  Example  8  the  biologist  had  also  been  in- 
terested in  the  effect  the  type  of  application  (aerial 
versus  hand  application )  had  on  the  effectiveness  of 
the  herbicides,  the  experimental  design  could  be 
altered  to  examine  these  interactions.  The  experi- 
mental design  would  be  the  same  as  described  ex- 
cept each  herbicide  treatment  would  have  been 
applied  manually  and  aerially.  The  biologist  would 
be  able  to  answer  the  same  questions  as  before, 
but  would  also  be  able  to  answer  the  following 
questions: 


( 1 )  Is  there  a  difference  in  the  effectiveness  of 
the  mode  of  application? 

(2)  Is  there  any  interaction  between  the  mode  of 
application  and  type  of  herbicide  used? 

Regression  Analysis 

Another  situation  often  encountered  is  the  need 
to  predict  the  changes  in  one  factor  as  other  factors 
change.  This  type  of  situation  is  handled  statistically 
by  the  use  of  regression  analysis,  using  either  simple 
linear  regression  (one  dependent  variable  and  one 
independent  variable )  or  multiple  linear  regression 
(one  dependent  variable  and  many  independent 
variables).  Regression  analysis  uses  associations  be- 
tween the  dependent  and  independent  variables 
to  construct  an  equation  describing  the  dependent 
variable  based  on  the  independent  variables.  From 
this  equation,  the  biologist  can  predict  what  the 
dependent  variable  will  be  as  the  independent  varia- 
bles change.  Another  application  of  the  equation  is 
to  permit  a  biologist  to  predict  a  variable  too  expen- 
sive or  difficult  to  measure  directly  (dependent  vari- 
able) from  variables  less  costly  to  measure 
(independent  variables). 


Closely  associated  with  regression  analysis  is 
correlation  analysis.  While  regression  analysis  de- 
scribes how  one  factor  will  change  as  other  factors 
change,  correlation  analysis  determines  the  degree  of 
association  between  the  factors. 


EXAMPLE  9 


The  biologist  has  noted  an  apparent  association 
between  the  density  of  sagebrush  plants  and  the 
density  of  sage  grouse  in  a  certain  mountain  basin. 
Based  on  this  association,  the  biologist  wishes  to 
predict  the  number  of  sage  grouse  per  square  kilo- 
meter from  the  number  of  sagebrush  plants  per 
hectare,  and  has  collected  the  following  sets  of 
data: 


SAGEBRUSH 

5ET 

PLANTS/ha 

SAGE  GROUSE/km2 

1 

200 

21 

2 

600 

29 

3 

500 

30 

4 

100 

12 

5 

700 

32 

6 

700 

27 

7 

100 

17 

8 

500 

32 

9 

900 

37 

10 

200 

18 

11 

300 

25 

Based  on  these  data  the  biologist  calculates  the 
regression  equation: 


Y  =  0.0255  X  +   14.32 


r  =  0.91 


Where:  Y  =  the  number  of  sage  grouse  per 
square  kilometer. 

X  =  the  number  of  sagebrush  plants 
per  hectare 

r  =  the  correlation  coefficient  of  the  two 
variables. 

The  biologist  then  asks  what  the  population 
density  of  sage  grouse  will  be  if  sagebrush  density 
is  reduced  to  300  plants  per  hectare.  Replacing  the 
X  in  the  regression  equation,  the  biologist  estimates 
there  will  be  22  sage  grouse  per  square  kilometer  at 
a  plant  density  of  300  plants  per  ha.  The  biologist 
goes  on  to  ask  for  the  range  of  estimates  in  the 
sage  grouse  density  if  he  or  she  was  willing  to  ac- 
cept only  a  10%  chance  of  error,  and  constructs 
a  90%  confidence  interval  around  the  estimate. 

90%  C.I.:     22  sage  grouse  ±  6.37  sage  grouse 


The  correlation  coefficient  given  in  the  previous 
example  is  a  measure  of  the  degree  of  association 
between  the  independent  and  dependent  variables.  A 
correlation  coefficient  (r)  approaching  1  indicates  a 


Statistical  Analysis 


749 


high  positive  association  between  the  two  variables; 
as  one  variable  increases,  the  other  variable  also 
increases.  A  correlation  coefficient  of  -1  indicates  a 
high  negative  association,  where  one  variable  in- 
creases as  the  other  decreases.  The  closer  the  r  value 
is  to  1  or  -1,  the  stronger  the  association;  the  closer 
to  zero,  the  weaker  the  association. 

The  correlation  coefficient  is  only  an  indication 
of  how  the  two  variables  will  change  in  association 
with  each  other.  It  does  not  indicate  a  cause-and- 
effect  relationship  although  there  may  be  one.  In  Ex- 
ample 9,  the  high  r  value  of  0.9 1  does  not  necessar- 
ily mean  the  change  in  sagebrush  density  is  causing  a 
change  in  sage  grouse  density. 

Even  though  two  variables  have  a  high  correla- 
tion, one  variable  will  not  necessarily  be  a  good 
predictor  of  the  other.  The  two  variables  may  be 
changing  in  association  with  each  other,  but  the 
amount  of  change  is  so  small  the  predication  equa- 
tion does  not  provide  any  information.  This  situation 
can  be  discovered  by  testing  to  see  if  the  slope  of 
the  regression  line  is  significantly  greater  than  zero. 

The  lack  of  significance  in  Example  10  may  be 
due  more  to  the  small  sample  size  than  a  true  lack  of 
slope  in  these  data.  A  larger  sample  size  could  indi- 
cate a  slope  significantly  different  from  zero.  How- 
ever, this  example  is  typical  of  the  sample  sizes 
many  biologists  use  or  must  use. 

The  example  and  descriptions  of  uses  of  simple 
linear  regression  given  here  can  be  expanded  to 
more  than  one  independent  variable  by  the  use  of 
multiple  linear  regression.  Multiple  regression  can  be 
useful  in  choosing  which  independent  variables  of 
many  can  be  useful  in  predication,  by  testing  the 
coefficients  of  each  variable  to  see  whether  they  are 
greater  than  zero  (as  done  in  Example  10). 

Nonparametric  Statistics 

Many  of  the  statistics  discussed  so  far  have  been 
based  on  the  assumption  that  the  populations,  from 
which  the  samples  come,  have  a  normal  or  some 
other  distribution.  This  assumption  may  not  always 
be  the  case  (as  in  Example  3)  or  a  biologist  may  not 
be  willing  to  make  the  assumption.  In  such  situa- 
tions, a  distribution-free  or  nonparametric  test  can 
be  used,  analogous  to  the  parametric  tests  already 
presented.  It  is  also  possible  to  transform  these  data 
so  the  assumptions  for  parametric  tests  can  be  made; 
however,  this  is  a  topic  that  will  not  be  discussed 
here. 

Mann-Whitney  U  Test.  The  Mann-Whitney  U  test  is 
used  in  situations  appropriate  for  a  t-test,  but  only 


EXAMPLE  10 


For  the  last  9  years,  the  biologist  has  been 
measuring  the  ambient  temperature  on  April  1  and 
the  peak  standing  crop  (g/m2)  on  July  1 ,  in  a  certain 
mountain  valley,  hoping  the  temperature  may  be 
used  to  predict  the  standing  crop. 

YEAR        TEMP(°C)       STANDING  CROP(g/m2 


1 

11 

500 

2 

12 

520 

3 

12 

519 

4 

13 

530 

5 

14 

535 

6 

14 

537 

7 

9 

492 

8 

11 

498 

9 

5 

480 

The  following  regression  line  is  calculated  from 
these  data: 


Y  =  6.71X  +  437.05 


r  -  0.92 


The  high  correlation  coefficient  of  the  equation 
(0.92)  suggests  there  is  a  strong  positive  associa- 
tion between  the  two  variables.  The  biologist  wants 
to  test  whether  the  temperature  is  a  good  predictor 
of  standing  crop.  To  do  this  he  or  she  asks  the 
question  and  formulates  the  hypothesis: 

Is  the  data  slope  of  the  temperature  line  signifi- 
cantly different  from  zero7 

H0:  B  =  0 
Ha:  B  +  0 

Where:  B  =  the  slope  of  the  regression  line. 

The  biologist  determines  that  only  a  2%  chance 
of  error  is  acceptable.  A  t-test  statistic  is  calculated 
for  the  slope  of  the  line  at  2.44.  It  is  then  compared 
to  a  table  value  of  the  t-distribution  of  these  data 
of  2.998.  The  test  statistic  is  not  beyond  the  level  of 
error  set  for  the  test;  therefore,  the  null  hypothesis 
cannot  be  rejected.  The  biologist  would  conclude 
the  slope  of  the  regression  line  is  not  significantly 
different  from  zero,  and  the  ambient  temperature  on 
April  1  is  not  a  good  predictor  of  standing  crop  on 
July  1,  despite  the  high  correlation  between  the  two 
variables. 


where  no  assumption  of  a  normal  distribution  is 
made  and  samples  are  from  independent  populations 
(Harnett  1982).  (See  Example  11.) 

Contingency  Table.  Another  nonparametric  proce- 
dure is  the  use  of  the  Chi-Square  statistic  to  test  the 
independence  of  two  or  more  classifications  made 
on  the  same  site  or  subject  by  using  a  contingency 
table.  Data  used  in  this  test  are  the  frequency  of 
classification  combinations  given  to  the  study  sub- 
ject. The  question  asked  is  "Is  one  classification  of  a 


750 


Statistical  Analysis 


site  independent  of  another  classification  of  the  same 
site  or  does  one  classification  tend  to  occur  consis- 
tently with  another  classification?" 


portance  as  a  tool  of  the  ecologist.  The  intent  of  this 
section  is  to  merely  introduce  the  reader  to  these 
procedures  and  to  describe  general  areas  where 
these  techniques  could  be  applied. 


EXAMPLE  1 1 


A  biologist  has  encountered  a  similar  situation 
as  described  in  Example  7,  but  is  unwilling  or  un- 
able to  make  the  assumption  that  the  population  has 
a  normal  distribution.  Again,  the  biologist  wants  to 
know  if  the  two  samples  are  from  the  same  popula- 
tion and,  thus,  forms  hypotheses  to  test  the  relation- 
ship at  a  5%  chance  of  error. 

H0:  Population  A  =  Population  B 
Ha:  Population  A  =£  Population  B 

Ten  random  samples  are  located  in  each  site 
and  standing  crop  data  (g/m2)  are  collected  from 
each  site: 


SITE  A 


SITE  B 


900 

1,065 

1,201 

843 

857 

912 

1,013 

1,104 

923 

1,292 

951 

1,312 

1,173 

742 

1,021 

961 

914 

1,093 

777 

776 

A  Mann-Whitney  U  test  statistic  is  calculated 
from  these  data  at  46  and  compared  with  a  table 
value  of  78  (Snedecor  and  Cochran  1967).  The 
calculated  value  exceeds  the  acceptable  error  es- 
tablished for  the  test  (the  calculated  value  is  less 
than  the  table  value),  and  the  null  hypothesis  is 
rejected.  The  biologist  would  conclude  the  two  sam- 
ples are  not  from  the  same  population. 


MULTIVARIATE  ANALYSIS 

All  the  analyses  dealt  with  so  far  have  used  only 
one  variable  at  a  time,  e.g.,  standing  crop,  crown 
diameter,  plants  per  unit  area.  This  type  of  analysis  is 
known  as  univariate  analysis.  Unfortunately,  most 
habitats  and  habitat  relationships  cannot  be  ade- 
quately described  by  only  one  variable,  but  require 
the  use  of  many  variables.  Descriptions  using  more 
than  one  variable  are  the  area  of  statistics  known  as 
multivariate  analysis.  This  area  of  statistical  analysis  is 
receiving  greater  emphasis  from  the  ecological  com- 
munity (Johnson  1981 ),  despite  the  complexity  of 
calculation  and  difficulty  of  interpretation.  However, 
as  the  availability  of  computer  programs  expands  and 
as  biologists  become  more  familiar  with  these  tech- 
niques, this  area  of  statistics  will  attain  greater  im- 


EXAMPLE  12 


A  specialist  has  identified  22  separate  mountain 
valleys  that  can  be  grazed  by  either  cattle  or  elk. 
By  an  arbitrary  system,  each  valley  has  been  classi- 
fied as  good,  fair,  and  poor  elk  range  and  as  good, 
fair,  and  poor  cattle  range.  The  specialist  is  inter- 
ested in  determining  if  the  two  classifications  are  in- 
dependent of  each  other,  or  if  good  elk  range  also 
tends  to  be  good  cattle  range  To  answer  the  ques- 
tion at  the  5%  level  of  rejection,  the  hypotheses 
are  formed: 

HQ:  The  classifications  are  independent. 
Ha:  The  classifications  are  dependent 

Each  valley  classification  is  arranged  in  a  con- 
tingency table. 


CATTLE        Fa 


Good  Fair  Poor 

Where  the  values  in  the  table  are  read  as — 

OBSERVED  FREQUENCY/EXPECTED  FREQUENCY 

From  these  data,  the  calculated  Chi-Square 
statistic  is  6.14,  which  compared  with  the  table 
value  (9.49)  is  not  beyond  the  level  of  rejection  es- 
tablished for  the  test.  The  null  hypothesis  cannot  be 
rejected,  and  the  specialist  would  conclude  the 
two  classifications  are  independent  of  each  other. 
Good  elk  range  is  not  always  good  cattle  range. 


ELK 

Good 

-     2/2  64 

3/1.2 

1/2.16 

Fair 

-    4/4.84 

2/2  2 

5/3.96 

Poor 

-    5/3.52 

0/1.6 

3/2.88 

Although  not  a  true  multivariate  analysis,  the 
first  technique  that  comes  to  mind  is  multiple  regres- 
sion. Touched  on  briefly  before,  this  analysis  permits 
a  biologist  to  predict  a  variable  of  interest  (depend- 
ent) from  many  other  variables  (independent).  The 
technique  can  be  used  to  predict  a  variable  too 
costly  to  measure  directly  from  less  costly  variables; 
to  predict  changes  in  a  variable  based  on  changes 
in  other  variables;  or  when  using  stepwise  multiple 
regression,  to  select  those  independent  variables  that 
account  for  the  most  variability  in  the  dependent 
variable.  A  common  use  in  habitat  analysis  is  to  pre- 
dict animal  abundance  (dependent  variable)  from 
a  set  of  habitat  attributes  (independent  variables). 


Statistical  Analysis 


751 


Discriminant  Analysis 

Discriminant  function  analysis  is  used  to  sepa- 
rate observations  into  groups  (Johnson  1981 ).  When 
a  number  of  groups  or  sites  have  been  defined  on 
the  basis  of  certain  ecological  factors,  discriminating 
between  the  different  sites  is  sometimes  difficult 
because  of  overlapping  distributions  of  ecological 
factors  used  to  describe  the  sites.  A  biologist  cannot 
clearly  assign  a  site  to  one  group  or  another  because 
the  site  could  have  attributes  of  both  groups. 

Based  on  the  variables  used  to  describe  the 
original  groups,  discriminant  analysis  creates  new 
distributions  (hybrid  distributions  of  the  original 
variables )  that  result  in  greater  separation  or  dis- 
crimination between  the  defined  groups  (Webster 
and  Burrough  1974).  The  hybrid  distributions  make 
it  easier  to  assign  a  site  to  one  group  or  another 
because  of  greater  definition  between  the  groups. 
This  is  potentially  one  of  the  most  useful  multivar- 
iate techniques,  especially  in  the  area  of  habitat 
classification. 

The  discriminant  analysis  is  a  transformation  of 
the  original  factors  for  each  site  into  a  single  value 
(score)  distribution  for  each  site,  with  more  distinc- 
tion between  the  sites.  The  biologist  calculates  the 
discriminant  score  for  each  new  site,  plots  the  new 
score  on  the  distributions  of  the  defined  sites,  and 
determines  which  site  the  new  location  more  closely 
approximates.  (See  Example  13.) 

Principal  Component  Analysis 

The  principal  component  analysis  is  a  multivar- 
iate technique  used  to  combine  "measurements  of 
similar  nature  into  a  fewer  number  that  may  be 
more  stable"  (Johnson  1981 ).  The  process  selects  a 
few  important  variables  out  of  many,  by  taking  a 
set  of  similar  observations  (measurements)  and  com- 
bining them  into  a  single  variable.  For  example, 
measurements  of  tree  height,  diameter,  age,  and 
stand  density  may  all  be  combined  into  a  single  vari- 
able. The  first  new  component  constructed  from 
these  original  data  will  account  for  the  largest  possi- 
ble variability  of  these  original  data,  and  each  addi- 
tional component  constructed  will  account  for  the 
largest  possible  amount  of  the  remaining  variability 
left  after  construction  of  previous  components. 
(See  Example  14.) 

Principal  component  analysis  is  useful  in  identi- 
fying which  groups  of  measurements  will  account  for 
the  greatest  variability  in  these  data.  If  several  soil, 
climate,  and  vegetation  measurements  are  collected 
for  a  site,  principal  component  analysis  will  identify 
which  group  (soil,  climate,  or  vegetation)  accounts 
for  the  greatest  variability  in  these  data.  Because 
all  these  data  are  included  in  the  analysis,  some 


combinations  of  measurements  could  be  difficult  to 
interpret.  Tree  measurements  and  soil  measurements 
could  be  combined  into  a  single  principal  compo- 
nent making  it  difficult  to  interpret  exactly  what  the 
principal  component  is  telling  a  biologist. 


Canonical  Correlation 

Very  similar  to  the  principal  component  analysis 
is  the  canonical  correlation  analysis.  The  difference 
is  that  canonical  correlation  analysis  uses  two  sepa- 
rate sets  of  measurements  and  derives  a  linear  com- 
bination of  each  data  set  such  that  the  correlation 
between  the  two  derived  values  (components)  is 


EXAMPLE  13 


A  biologist  defines  two  habitat  types  on  the 
basis  of  these  available  data  and  on  past  experi- 
ence. To  define  the  types,  the  biologist  relied  on  the 
percentage  of  shrub  cover,  annual  precipitation, 
and  the  annual  forage  production.  When  examining 
new  locations,  the  biologist  had  difficulty  placing 
new  locations  into  one  of  the  two  habitat  types.  One 
factor  (e.g.,  forage  production)  may  place  the  loca- 
tion in  one  type,  whereas  the  other  factors  (e.g., 
percentage  shrub  cover  and  annual  precipitation) 
would  place  the  same  location  in  a  different  type.  To 
help  differentiate  the  new  locations  into  the  two 
habitat  types,  the  biologist  performed  a  discriminant 
analysis  on  the  two  previously  defined  types  and 
obtained  a  discriminant  equation: 

D  =  aX  +bY  +  cZ  +  C 

Where:    X     =  the  standardized  shrub  cover. 

a     =  the  standardized  discriminant  coeffi- 
cient for  shrub  cover. 

Y     =  the  standardized  annual  precipita- 
tion. 

b     =  the  standardized  discriminant  coeffi- 
cient for  precipitation. 

Z     =  the  standardized  annual  forage  pro- 
duction. 

c     =  the  standardized  discriminant  coeffi- 
cient for  forage  production. 

C     =  the  constant  for  the  discriminant 
equation. 

D     =  the  discriminant  score. 

A  mean  discriminant  score  was  calculated  for 
each  of  the  two  known  habitat  types.  As  each  new 
site  was  encountered,  a  discriminant  score  was 
calculated  and  compared  to  the  means  of  the  two 
known  groups  (habitat  sites).  Each  new  site  was 
then  classified  into  one  of  the  two  known  habitat 
types,  based  on  which  of  the  known  discriminant 
groups  had  the  highest  probability  of  containing  the 
discriminant  value  from  the  new  site. 


752 


Statistical  Analysis 


EXAMPLE  14 


In  the  previous  example  (Example  13),  the  biol- 
ogist decided  not  to  limit  these  data  to  the  three 
sets  described  and  instead  decided  to  use  19  cli- 
matic variables  (data  from  Newnham  1968): 

NUMBER     VARIABLE 


Latitude 

Elevation,  feet 

Average  daily  max.  temp. 


(°F), 


Winter 
Spring 
Summer 
Fall 


Average  daily  mm.  temp.  (°F), 


9 

10 


Winter 
Spring 
Summer 
Fall 


11 
12 
13 
14 


Average  daily  mean  temp.  (°F), 


Winter 
Spring 
Summer 
Fall 


15 

Average  precipitation  (inches) 

,  Winter 

16 

Spring 

17 

Summer 

18 

Fall 

19 

Average  frost-free  period, 

days 

After  computat 

on,  three  components 

are  de- 

rived  that  account  for  92.2%  of  the  variation  of  these 

data. 

COMPONENT 

1 

2 

3 

VARIABLE 

COEFFICIENTS 

1 

-0.63 

-0.50 

1.00 

2 

-0.85 

0.10 

-067 

3 

0.97 

0.10 

-0.41 

4 

0.02 

1.00 

-0.18 

5 

-0.11 

0.99 

-0.29 

6 

0.94 

0.13 

-0.45 

7 

1.00 

-0.02 

-0.10 

8 

0.90 

0.35 

0.50 

9 

0.84 

0.31 

0.75 

10 

099 

-0.05 

-0.01 

11 

1.00 

0.02 

-0.22 

12 

0.57 

0.85 

0.17 

13 

0.38 

0.94 

0.22 

14 

0.99 

0.03 

-0.19 

15 

0.72 

-0.67 

-0.23 

16 

0.59 

-0.73 

-0.26 

17 

0.40 

-086 

0.06 

18 

0.71 

-070 

-0.10 

19 

0.93 

0.01 

0.05 

PERCENT  OF 
VARIATION 


57.4 


29.4 


5.4 


Once  each  coefficient  was  determined  for  each 
variable,  the  biologist  multiplied  each  by  its  coeffi- 
cient and  summed  the  products  to  arrive  at  the 
value  for  each  principal  component.  The  three  com- 
ponents would  then  be  used  in  the  discriminant 
analysis  as  described  in  Example  13. 


maximized  (Warwick  1975).  A  second  canonical  cor- 
relation ( component )  is  then  constructed  from  the 
remaining  variability  of  these  original  data  so  the 
highest  correlation  is  achieved  between  the  next 
two  components. 

Whereas  the  principal  component  analysis  de- 
rives values  from  a  single  set  of  data  to  account  for 
the  greatest  variation  in  these  data,  the  canonical 
correlation  analysis  derives  values  from  two  sets  of 
measurements  to  attain  the  highest  correlation  be- 
tween the  two  data  sets.  If  one  data  set  contained 
vegetation  measurements  and  a  second  contained 
abiotic  measurements,  canonical  correlation  analysis 
would  identify  which  of  the  abiotic  measurements 
is  most  highly  associated  ( correlated )  with  the  vege- 
tation measurements. 


EXAMPLE  15 


A  biologist  has  collected  data  on  relative  hu- 
midity (V1),  precipitation  (V2),  wind  speed  (V3),  and 
climatic  temperatures  (V4).  Data  have  also  been 
collected  on  vegetation  cover  (V5),  frequency  (V6), 
production  (V7),  and  standing  crop  (V8).  The  biolo- 
gist is  curious  how  the  two  data  sets  (V1  through  V4 
and  V5  through  V8)  relate  to  each  other. 

The  canonical  analysis  selects  the  V1  and  V2 
values  from  the  first  set,  and  V7  and  V8  from  the 
second  set  to  be  used  in  the  first  canonical  variant 
(CANVAR1).  Since  the  biologist  is  only  interested 
in  the  effects  of  climate  on  vegetation,  only  the  coef- 
ficients for  these  climate  data  are  presented. 


CANVAR1 


0.527(V1)  -0.463(V2) 


The  V3  and  V4  values  from  the  first  set  and  the 
V5  and  V6  values  from  the  second  set  are  selected 
for  the  second  canonical  variant  (CANVAR2). 

CANVAR2  =  0.542(V3)  -  0.635(V4) 

The  first  canonical  variant  has  a  correlation  of 
0.889,  indicating  that  these  data  in  the  first  canoni- 
cal variant  (V1,  V2,  V7,  V8)  share  79%  (the  square 
of  the  correlation)  of  the  variance  of  these  data.  The 
second  variant  has  a  correlation  of  0.824  and 
shares  68%  of  the  variance. 

After  conducting  canonical  correlation,  the  biol- 
ogist found  precipitation-humidity  data  and  produc- 
tion-standing crop  data  had  the  highest  correlation, 
followed  by  wind  speed,  temperature,  and  cover- 
frequency  data. 


As  with  principal  component  analysis,  a  biologist 
runs  the  risk  of  obtaining  combinations  of  data  that 
are  hard  to  interpret.  This  example  was  constructed 
in  such  a  way  as  to  present  the  clearest  interpreta- 
tion; however,  this  type  of  clarity  will  seldom  be 
present  in  most  ecological  situations. 


Statistical  Analysis 


753 


TIME-SERIES  ANALYSIS 

Ecologists  have  long  recognized  that  ecological 
systems  experience  a  cyclic  behavior  over  time.  The 
use  of  time-series  analysis  allows  a  biologist  to  de- 
scribe how  a  variable  changes  through  time  and 
to  predict  how  the  variable  will  change  in  the  future. 
The  time  frame  used  with  this  analysis  can  be  any 
unit  the  biologist  wishes:  seconds,  minutes,  hours, 
years,  or  other  measures  of  time  (Chatneld  1980).  In 
most  statistical  analysis,  all  measurements  are  as- 
sumed independent;  in  time-series  analysis,  observa- 
tions taken  over  time  are  often  not  independent 
and  an  observation  may  be  influenced  by  past 
occurrences. 

Time-series  analysis  has  three  major  uses  in  the 
field  of  ecology.  First,  it  can  be  used  to  describe 
the  type  of  variation  occurring  (seasonal,  cyclic, 
trend,  or  irregular  fluctuations).  Second,  the  ob- 
served variation  in  one  time  series  can  be  used  to 
explain  the  observed  variation  in  another  time  series 
(e.g.,  the  variation  in  forage  production  has  the  same 
time  series  pattern  as  the  variation  in  deer  produc- 
tivity). Third,  time  series  can  be  used  to  predict  a 
variable  at  some  future  time. 


EXAMPLE  16 


A  biologist  has  collected  20  years  of  data  on 
rabbit  population  densities  in  a  certain  mountain 
valley.  By  plotting  these  raw  data,  the  biologist  can 
see  a  cyclical  pattern  in  the  densities;  however,  due 
to  the  "noise"  in  these  data,  the  pattern  is  unclear. 
After  performing  time-series  analysis,  the  biologist 
develops  a  logistic  curve  to  describe  these  data: 


Xt 


A 


;1  +B  exp[-Ct]) 


Where:  A,  B,  and  C  =  parameters  and  t  = 
time. 

Upon  examining  the  equation  and  the  plot  of 
the  equation,  the  biologist  can  clearly  see  the  cycles 
in  these  data,  an  overall  trend  (up  or  down)  to  these 
data,  and  whether  the  amplitude  of  the  cycles  is 
increasing  or  decreasing. 

The  equation  also  allows  the  biologist  to  predict 
the  density  of  rabbits  at  some  future  time.  By  com- 
paring the  time  series  for  the  rabbits  to  time  series 
for  factors  affecting  rabbit  density  (e.g.,  coyote 
density),  the  biologist  may  also  be  able  to  explain 
the  pattern  of  the  rabbit  time  series  by  the  pattern  of 
the  time  series  of  other  factors. 


The  use  of  time-series  analysis  holds  great  prom- 
ise in  the  area  of  habitat  management.  However,  it 
is  a  complicated  procedure  that  requires  the  help  of 
a  statistician,  and  large  data  sets  are  needed  to  ade- 
quately determine  the  pattern  of  fluctuations.  As 
with  multivariate  analysis,  the  advent  of  computer 


programs  will  make  the  use  of  time-series  analysis 
more  available  to  the  biologist. 

OPTIMAL  ESTIMATION 

Better  techniques  for  estimating  resource  meas- 
urements are  beginning  to  appear  in  the  natural  re- 
sources field.  These  techniques  provide  the  biologist 
with  an  estimation  of  the  number  of  samples  that 
must  be  made  and  how  often  (e.g.,  every  5  years) 
samples  must  be  taken  to  maintain  a  predefined  level 
of  variance  in  these  data  (Jameson  1985).  These 
techniques  hold  great  promise  for  the  efficient  plan- 
ning of  monitoring  studies  over  the  long  term  and 
can  be  applied  either  to  univariate  or  multivariate 
conditions. 

The  process  utilizes  Bayesian  statistics  to  deter- 
mine when  the  process  variance  becomes  unaccepta- 
ble, based  on  actual  sample  variation  and  estimated 
change  in  the  variance  over  time.  The  change  in  the 
variance  over  time  may  be  made  by  either  assuming 
a  constant  change  or  by  simulating  the  changes 
based  on  the  conditions  present,  using  a  simulation 
model. 

As  time  passes,  the  variance  of  the  process  will 
increase  until  the  variance  surpasses  the  maximum 
tolerable  level  set  by  the  biologist  or  manager  (Fig- 
ure 3).  To  return  the  variance  to  a  tolerable  level, 
a  sample  must  be  taken  to  reduce  the  overall  vari- 
ance of  the  process. 


o 
z 
< 

< 

> 


ALLOWABLE  VARIANCE 
VARIANCE  FOR  n  =  2 
VARIANCE  FOR  n  =  3 


VARIANCE  FOR  n  =  5 


TIME  PERIOD 


Figure  3.     Changes  in  process  variance  over  time 

for  three  levels  of  sampling  (Jameson  1985). 


754 


Statistical  Analysis 


The  biologist  has  the  option  of  collecting  a  few 
samples  at  short  time  intervals  or  collecting  many 
samples  over  longer  time  intervals.  By  applying  cost 
estimates  for  the  sampling  process,  the  biologist 
can  evaluate  the  sampling  scheme  (few  samples — 
short  time  frame;  many  samples — long  time  frame) 
that  is  most  economical. 


CONCLUSION 

Statistics  are  an  important  and  powerful  re- 
source tool.  The  techniques  can  be  used  to  describe 
habitats,  to  test  ideas  and  concepts,  to  evaluate  man- 
agement alternatives  and  decisions,  and  to  predict 


changes.  While  they  are  a  powerful  tool,  statistics  are 
only  a  tool  and  are  not  capable  of  making  decisions. 
Statistics  only  aid  the  resource  biologist  in  the  deci- 
sion process. 


To  be  properly  employed  and  not  abused,  statis- 
tics must  be  considered  from  the  very  beginning  of 
any  data  collection  or  impact  study  effort.  As  part  of 
conceptualizing  a  project,  statistics  can  help  to  clar- 
ify the  questions  being  asked,  avoid  the  collection  of 
unnecessary  data,  and  reduce  the  collection  of  mean- 
ingless data.  When  used  as  intended,  statistics  can 
provide  the  biologist  with  meaningful  and  useful 
information. 


Statistical  Analysis 


755 


LITERATURE  CITED 


BOX,  G.E.P.,  W.G.  HUNTER,  and  J.S.  HUNTER.  1978.  Statis- 
tics for  experimenters — An  introduction  to  design, 
data  analysis,  and  model  building.  John  Wiley  &  Sons, 
New  York,  NY.  653pp 

CHATFIELD,  C.  1980.  The  analysis  of  time  series:  An 
introduction.  Chapman  and  Hill,  London.  268pp. 

GREEN,  R.H.  1979.  Sampling  design  and  statistical  meth- 
ods for  environmental  biologists.  John  Wiley  &  Sons, 
New  York,  NY.  257pp. 

HARNETT,  D.L.  1982.  Statistical  methods,  3rd  ed.  Addison- 
Wesley  Publ.  Co.,  Reading,  MA.  730pp. 

HUNTSBERGER,  D.V.  and  P.  BILLINGSLEY.  1977.  Ele- 
ments of  statistical  inference,  4th  ed.  Allyn  and  Bacon, 
Inc.,  Boston,  MA.  385pp. 

JAMESON,  D.A.  1985.  Sampling  intensity  for  monitoring  of 
environmental  systems.  Applied  Mathematics  and  Com- 
putation. Elsevier  Sci.  Publ.  Co.,  Inc.,  New  York,  NY. 


JOHNSON,  D.H.  1981.  The  use  and  misuse  of  statistics 

in  wildlife  habitat  studies.  Pages  11-19  in  Capen,  D.E., 
ed.  The  Use  of  Multivariate  Statistics  in  Studies  of 
Wildlife  Habitat.  U.S.  Dep.  Agric,  For.  Serv.  Gen.  Tech. 
Rep.  RM-87.  249pp. 

KREBS,  C.J.  1978.  Ecology — the  experimental  analysis  of 
distribution  and  abundance,  2nd  ed.  Harper  &  Row, 
New  York,  NY.  678pp. 

NEWNHAM,  R.M.  1968.  A  classification  of  climate  by  prin- 
cipal component  analysis  and  its  relationship  to  tree 
species  distribution.  Forest  Science  14:254-264. 

SNEDECOR,  G.  and  W.G.  COCHRAN.  1967.  Statistical 
methods,  6th  ed.  Iowa  State  University  Press,  Ames. 
593pp. 

WARWICK,  P.V.  1975.  Canonical  correlation  analysis: 

Subprogram  CANCORR.  Pages  515-527  in  Nie,  N.H., 
C.H.  Hull,  J.G.  Jenkins,  K.  Steinbrenner,  and  D.H.  Bent, 
eds.  Statistical  Package  for  the  Social  Sciences  (SPSS), 
2nd  ed.  McGraw-Hill,  New  York,  NY.  675pp. 

WEBSTER,  R.  and  P.A.  BURROUGH.  1974.  Multiple  dis- 
criminant analysis  in  soil  survey.  J.  Soil  Science 
25:120-134. 


756  Statistical  Analysis 


38 

HABITAT 

EVALUATION 

SYSTEMS 

Allen  Y.  Cooperrider 


U.S.  Bureau  of  Land  Management 
Service  Center 
Denver,  CO  80225 


"What  Is  Game  Range?'  When  the  game  manager 
asks  himself  whether  a  given  piece  of  land  is  suitable 
for  a  given  species  of  game,  he  must  realize  that  he 
is  asking  no  simple  question,  but  rather  he  is  facing 
one  of  the  great  enigmas  of  animate  nature.  An  an- 
swer good  enough  for  practical  purposes  is  usually 
easy  to  get  by  the  simple  process  of  noting  whether 
the  species  is  there  and  ready,  or  whether  it  occurs 
as  'similar'  range  nearby.  But  let  him  not  be  cocksure 
about  what  is  'similar,'  for  this  involves  the  deeper 
questions  of  why  a  species  occurs  in  one  place  and 
not  in  another,  which  is  probably  the  same  as  why  it 
persists  at  all.  No  living  man  can  answer  that  ques- 
tion fully  in  even  one  single  instance." 

— Aldo  Leopold,  Game  Management 


Editor's  Note:  Habitat  evaluation  systems  are  the 
tools  for  analyzing  wildlife  habitat,  and  the  habi- 
tat model  is  the  basic  building  block  for  such  sys- 
tems. Therefore,  use  of  such  models  and  systems 
is  basic  to  the  habitat  inventory  and  monitoring 
process.  Biologists  should  be  familiar  with  existing 
techniques  in  this  area  Most  of  the  major  concept 
and  system  developments  have  occurred  in  the  past 
15  years  and  the  methodology  is  likely  to  continue 
to  develop  rapidly.  This  chapter  provides  an  intro- 
duction to  habitat  evaluation  systems  and  models, 
their  history,  construction,  and  use  in  the  inventory 
and  monitoring  process,  as  well  as  an  overview  of 
current  models  and  systems. 

INTRODUCTION 

Unorganized  data  are  of  little  value  to  the  biolo- 
gist or  manager.  Data  need  to  be  organized,  inte- 
grated, and  summarized  into  useful  information.  The 
biologist  can  then  use  such  information  to  make 
inferences  about  relative  conditions  of  habitats  and 
predictions  about  future  conditions  under  alternative 
management  practices.  Statistical  analysis,  as  dis- 
cussed in  the  previous  chapter,  is  one  means  of  ana- 
lyzing data. 

In  this  chapter,  I  discuss  methods  specifically 
designed  for  organizing  and  analyzing  wildlife  habitat 
data.  I  do  not  discuss  in  detail  models  and  systems 
designed  for  analyzing  a  single  species  (e.g.,  a  system 
for  analyzing  mule  deer  habitat )  or  a  particular  type 
of  habitat  (e.g.,  a  system  for  analyzing  riparian  habi- 
tat ).  Rather,  I  focus  on  models  and  systems  that  can 
be  used  for  many  different  species  (e.g.,  habitat  suit- 
ability index  [HSI]  models)  or  many  different  types 
of  habitats  or  communities  (the  system  to  be  used 
determines  the  data  to  collect).  The  reverse  process, 
in  which  a  biologist  collects  data  and  then  searches 
for  a  system  that  can  use  the  data,  is  inefficient  and 
should  be  avoided. 

The  habitat  model,  a  correlation  between  habi- 
tat components  and  some  attribute  of  animal  popula- 
tions), is  central  to  any  habitat  evaluation  system, 
although  it  may  not  be  labeled  as  such.  Therefore, 
much  of  this  chapter  focuses  on  habitat  models 
which  are  the  basic  building  blocks  of  any  such 
system. 

In  this  chapter,  I — 

( 1 )    Review  the  historic  development  of  habitat 
evaluation  systems  and  models; 

(  2 )    Discuss  the  elements  and  types  of  habitat 
models; 

( 3 )    Describe  the  procedure  a  biologist  should 
follow  to  develop  or  modify  a  model  for  use 
in  habitat  evaluation; 


Habitat  Evaluation  Systems 


757 


(4)  Discuss  the  relationship  between  habitat  eval- 
uation systems  and  models  and  the  habitat 
inventory  and  monitoring  cycle;  and 

(5)  Review  models  and  systems  currently  avail- 
able or  in  use. 


DEVELOPMENT  OF  HABITAT  INVENTORY, 
MONITORING,  AND  EVALUATION  SYSTEMS 

Inventory  and  monitoring  of  wildlife  habitat  has 
been  conducted  in  the  U.S.  for  over  50  years,  al- 
though such  efforts  were  rarely  described  by  those 
names.  Aldo  Leopold,  in  his  classic  work  on  Game 
Management,  published  in  1933,  described  methods 
for  "game  range  evaluation"  that  are  essentially  habi- 
tat inventory  and  monitoring  techniques.  Such  evalu- 
ations were  mostly  limited  to  practices  such  as 
browse  condition  and  trend  surveys  for  white-tailed 
deer  (Odocoileus  virgin ianus)  winter  ranges. 

Although  more  and  better  techniques  were  de- 
veloped, until  about  1970  most  habitat  surveys  con- 
sisted of  measuring  a  few  habitat  components  for  a 
particular  game  species.  Methods  were  not  available 
for  either  analyzing  habitat  for  nongame  species  or 
analyzing  the  capability  of  a  habitat  to  support  a 
wildlife  community  (as  opposed  to  a  single  species). 

At  the  same  time,  zoologists,  naturalists,  and 
other  biologists  were  acquiring  a  large  body  of 
knowledge  on  the  distribution  of  vertebrates  and  the 
habitats  used  by  most  vertebrates  in  North  America. 
Habitat  observations  were  typically  made  inciden- 
tally by  biologists  who  were  studying  a  particular 
species  or  surveying  a  region  for  wildlife.  Few  biolo- 
gists made  any  effort  to  systematically  collect  or 
synthesize  such  information.  Until  about  1970,  the 
best  summaries  of  such  species-habitat  relationships 
were  the  popular  field  guides  such  as  Peterson's 
Field  Guide  to  Western  Birds. 


These  guides  provided  simple  "verbal  models" 
of  habitats  used  by  individual  species.  For  example, 
the  American  dipper  (Cinclus  mexicanus)  is  charac- 
terized as  using  habitat  described  as  "Fast-flowing 
streams  in  or  near  mountains.  Lower  levels  in  win- 
ter." (Peterson  1969). 

Beginning  in  the  late  1960s,  interest  in  habitat 
evaluation  increased  dramatically  in  the  U.S.  The 
public  became  (  1  )  more  aware  of  the  value  and  im- 
portance of  all  wildlife,  not  just  the  huntable  or  eco- 
nomically important  species,  and  (2)  more 
concerned  about  the  impact  of  human  actions  on 
wildlife  habitat.  This  concern  was  eventually  ex- 
pressed in  a  series  of  federal  legislative  mandates 
that  were  used  by  agencies  to  manage  habitat  for  all 


wildlife  species  and  to  publicly  address  and  mitigate, 
where  possible,  the  impact  of  management  actions 
on  wildlife  resources. 

This  mandate  required  new  systems  and  ap- 
proaches to  habitat  inventory  and  monitoring.  Wild- 
life biologists  were  thus  faced  with  new  problems 
and  challenges. 

The  first  major  challenge  for  biologists  was  deal- 
ing with  all  species  or  at  least  all  vertebrate  species. 
Wildlife  biologists  were  not  trained  or  experienced 
in  working  with  most  nongame  species.  On  the 
other  hand,  zoologists  were  not  accustomed  to  col- 
lecting the  types  of  data  that  were  relevant  to  man- 
agement. Neither  group  was  experienced  in  working 
with  an  entire  vertebrate  community.  Habitat  re- 
quirements for  many  species  were  not  known,  and 
techniques  for  measuring  habitats  or  populations  of 
many  species  groups  were  primitive.  Costs  for  meas- 
uring habitats  and  populations  of  all  species  were 
prohibitive,  yet  systems  for  setting  priorities  for  field 
work  did  not  exist. 


The  second  major  challenge  was  to  make  pre- 
dictions about  impacts  from  human  activities.  Biolo- 
gists had  always  made  such  predictions,  but  they 
tended  to  be  informal  and  based  on  experience  and 
subjective  interpretation  of  limited  data.  Legislation 
such  as  the  National  Environmental  Policy  Act 
(NEPA)  required  impact  predictions  to  be  based  on 
data  that  were  systematically  collected  and  analyzed, 
and  that  such  predictions  be  formally  presented  and 
open  to  public  scrutiny  in  environmental  assessment 
reports  (EARs),  environmental  impact  statements 
(EISs),  and  other  decision  documents. 

More  recently  the  need  for  systems  to  valuate 
wildlife  resources  has  emerged.  Economic  analysis  of 
wildlife  resources  is  still  primitive,  but  economic 
values  can  be  quite  important  in  influencing  manage- 
ment systems.  Such  economic  analyses  require  input 
on  the  wildlife  population  responses  to  management 
actions. 


At  the  same  time  new  demands  were  being 
made  of  biologists,  new  technology  became  avail- 
able. Of  particular  importance  were  the  new  devel- 
opments in  remote  sensing  and  computer 
technology. 

Remote  sensing  has  been  used  by  wildlife  biolo- 
gists for  many  years,  primarily  aerial  photography 
for  cover  mapping  and  radiotelemetry  for  tracking 
animal  movements.  However,  the  development  of 
these  technologies,  particuarly  in  the  past  1 5  years, 
has  allowed  biologists  to  collect  better  data,  more 
efficiently  than  in  the  past. 


758 


Habitat  Evaluation  Systems 


Newly  developed  computer  hardware  and  soft- 
ware have  probably  been  the  most  important  new 
technologies  for  the  biologist  concerned  with  wild- 
life habitat  inventory  and  monitoring.  Two  factors  of 
computer  technology  have  been  particularly  valu- 
able: (  1 )  the  capability  of  efficiently  storing  and  re- 
trieving large  amounts  of  data  and  (2)  the  ability 
to  rapidly  manipulate  numbers  and  images. 

The  storage  capability  of  computers  has  allowed 
large  amounts  of  inventory  and  monitoring  data  to 
be  efficiently  stored  and  retrieved.  This  has  relieved 
much  of  the  burden  of  handling  large  volumes  of 
data  and  lessened  the  errors  associated  with  tran- 
scribing data  by  hand. 

The  data  manipulation  capability  of  computers 
has  allowed  not  only  rapid  generation  of  summary 
statistics,  but  also  more  sophisticated  approaches 
such  as  multivariate  analyses.  Such  analyses  would 
not  be  practical  without  the  aid  of  modern  com- 
puters. Similarly,  many  of  the  models  developed  for 
integrating  data  would  not  be  practical  without  both 
the  storage  and  data  manipulation  capabilities. 


HABITAT  MODELS 

The  habitat  model  forms  the  basis  for  all  habitat 
inventory,  management,  and  monitoring.  It  is  thus 
the  general  underlying  principle  of  habitat 
management. 

A  habitat  model  is  a  method  of  using  a  set  of 
habitat  components  or  attributes  to  predict  some  at- 
tribute of  a  wildlife  population  or  populations  (Fig- 
ure 1 ).  All  habitat  models  are  designed  for  this 
purpose.  Habitat  models  are  complex  because  of  the 
almost  limitless  habitat  components  and  attributes 
that  can  be  used  (Table  1 ),  the  diversity  of  popula- 
tion attributes  that  can  be  predicted  (Table  2),  and 
the  complicated  relationships  between  them. 


define  "fast-flowing"  and  "near"  in  terms  of  meters 
(feet)/seconds  and  kilometers  (miles)  respectively, 
then  we  would  have  a  simple,  quantitative  model. 

To  understand  and  categorize  a  habitat  model, 
one  needs  only  to  identify  the  habitat  components 
being  used  as  predictors,  the  population  attribute 
being  predicted,  and  the  type  of  function  being  used 
to  relate  one  to  the  other.  In  the  following  discus- 
sion I  describe  each  of  these. 


Model  Components — Habitat  Attributes 

The  habitat  attributes  listed  in  Table  1  form  the 
basic  predictors  of  terrestrial  habitat  models.  They 
are  analogous  to  the  independent  variables  in  a 
regression  equation.  They  are  discussed  in  more 
detail  below  to  help  explain  the  concept  of  a  habitat 
model. 


Geographic  Location.  The  single  most  important 
predictor  of  occurrence  is  probably  geographic  loca- 
tion. Most  wildlife  species  are  quite  restricted  in 
geographic  distribution;  therefore,  geographic  loca- 
tion, together  with  knowledge  of  a  species  distribu- 
tion, is  adequate  to  predict  species  potentially 
present  in  an  area.  However,  wildlife  species  are 
rarely  present  continuously  within  their  geographic 
ranges,  and  complete  delineations  of  all  sites  used  by 
a  species  are  usually  not  available.  Furthermore,  loca- 
tion is  of  little  help  in  predicting  more  sophisticated 
attributes  of  a  wildlife  population,  such  as  relative 
abundance  or  density.  Therefore,  geographic  location 
in  habitat  models  is  usually  used  implicitly  in  models 
as  a  boundary  condition,  i.e.,  animals  outside  their 
known  geographic  range  are  assumed  to  be  absent. 
More  accurate  prediction  of  presence  and  more  de- 
tailed predictions  about  population  attributes  ob- 
viously require  more  detailed  information  on  habitat 
components  present. 


Habitat  models  have  always  been  used  by  biolo- 
gists; the  previous  example  of  the  American  dipper 
using  habitat  consisting  of  "fast-flowing  streams  in  or 
near  mountains"  is  a  simple,  verbal  model.  If  we 


Habitat  Components  or 
Attributes 

-  Cover   Type 

-  Forage    Supply 

-  Other 


Predictive 


Equation 


Population 
Characteristics 

-  Presence 

-  Abundance 

-  Density 

-  Other 


Vegetation.  Beyond  geographic  location,  vegetation 
is  probably  the  next  most  important  habitat  compon- 
ent. Almost  all  habitat  models  use  vegetation  in  pre- 
dicting animal  population  attributes.  In  the  simplest 
models,  the  presence  and  absence  of  animal  species 
are  simply  associated  with  a  vegetation  type  such 
as  pinon-juniper  (Pinus-Juniperus  sp.)  or  sagebrush 
(Artemisia  sp. )  grassland.  More  complex  models 
consider  structural  components  of  the  vegetation 
(e.g.,  tree  density,  shrub  density,  etc.)  or  more  detail 
on  plant  species  composition  (e.g.,  percentage  cover 
of  big  sagebrush  [A  tridentata],  density  of  antelope 
bitterbrush  [Purshia  tridentata}). 


Figure  1.     Basic  components  of  a  habitat  model. 


The  relative  importance  of  vegetation  structure 
(physiognomy),  as  opposed  to  plant  species  compo- 
sition (floristics)  in  determining  suitability  of  habitat 


Habitat  Evaluation  Systems 


759 


Table  1.     Habitat  components  and  attributes  useful  in  predicting  presence,  abundance,  or  density  of 
vertebrates.* 


Geographic  location 

Vegetation — Live 

Vegetation  type 

Species  composition  (floristics) 

Presence 

Abundance 

Density 

Cover 

Frequency 

Biomass 

Diversity 
Vegetation  structure  (physiognomy) 

Presence 

Abundance 

Density 

Cover 

Frequency 

Biomass 

Diversity 

Vegetation — Dead 

Litter  or  mulch 

Dead  and  down  woody  material 

Persistent 

Non-persistent 

Snags 

Physical  features 

Landform  types 

Alluvial  fan 

Rock  pediment 

Other 
Landform  attributes 

Slope 

Aspect 

Elevation 
Soils 

Edaphic  habitats 
Geomorphic  habitat  features 

Cliffs 

Caves 

Talus 

Lava  flows 

Sand  dunes 

Other 


Animal-made  habitat  features 

Beaver  dams 

Dens 

Nest  cavities 


Man-made  features 

Roads 
Bridges 
Buildings 
Other 


Water 

Presence 
Attributes 
Depth 

Flow  (velocity) 
Temperature 
Chemistry  (DO, 
Substrate 


pH,  turbidity,  TDS,  etc.) 


Food  supply 

Vegetation 

Animal  prey  base  (vertebrates) 

Macroinvertebrates 

Other 
Presence,  absence,  or  abundance  of  competitors 
Presence,  absence,  or  abundance  of  predators 
Presence,  absence,  or  abundance  of  parasites  or 

diseases 
Presence,  absence,  or  degree  of  human  disturb- 
ance 

Noise 

Human  activity 

Traffic 

Other 
Presence,  absence,  or  intensity  of  hunting  or 
harvesting 

Weather  and  climate 


Historical  occurrence 


The  spatial  and  temporal  arrangements  of  tne  attributes  and  components  are  also  an  important  determinant  of  animal  abundance 


for  wildlife  species,  is  a  subject  of  continuing  debate. 
Biologists,  such  as  Short  (1983,  1984)  and  Short 
and  Burnham  ( 1 982 ),  have  proposed  using  habitat 
models  primarily  based  on  structural  characteristics, 
whereas  some  researchers  provide  evidence  that 
species  composition  is  more  important.  The  manage- 
ment biologist  does  not  need  to  be  concerned  with 
the  controversy.  Some  wildlife  species  greatly  de- 
pend on  particular  plant  species  (e.g.,  sage  grouse 
[Centrocercus  wophasitmns]  on  sagebrush,  ruffed 
grouse  [Bonasa  umbel  I  us]  on  quaking  aspen  [Popu- 
lus  tremulokles]).  The  biologist  working  with  such 


species  is  well-advised  to  pay  attention  to  species 
composition.  On  the  other  hand,  habitat  layers  or 
other  structural  components,  as  proposed  by  Short 
( 1984),  can  provide  a  good  initial  prediction  of  ani- 
mal species  presence  or  absence.  Chapter  3 1  in  this 
book  describes  how  both  types  of  measurements  can 
be  taken  and  used  to  make  predictions  about  animal 
populations. 


Vegetation  has  been  measured  in  many  ways  for 
many  purposes,  resulting  in  literally  thousands  of 


760 


Habitat  Evaluation  Systems 


measurement  techniques  being  described  in  the 
literature.  For  this  reason,  the  chapter  on  vegetation 
(Chapter  31 )  does  not  attempt  to  cover  all  possible 
techniques.  Rather,  it  presents  an  example  of  how 
vegetation  variables  can  be  used  to  predict  species 
occurrence  and  abundance. 

Wildlife  biologists  often  collect  large  amounts  of 
vegetation  data  without  a  clear  idea  of  how  they  are 
going  to  use  such  data.  Habitat  models  are  the  mech- 
anism for  using  vegetation  data  to  make  predictions 
about  animal  populations.  The  biologist  should, 
therefore,  have  a  model  in  mind  before  collecting 
vegetation  data.  The  same  caveat  applies  to  all  other 
types  of  habitat  variables,  but  is  emphasized  here 
since  so  much  time,  effort,  and  money  continue  to 
be  expended  in  collecting  vegetation  data  that  are 
not  used. 


Table  2.     Animal  population  characteristics 
predicted  by  habitat  relationships  models. 


Single  Species  Characteristics 

Presence  or  absence 

Probability  of  occurrence 

Abundance  or  density 

Relative  abundance  or  density  index 

Carrying-capacity  index 

Biomass 

Geographic  distribution 

Mortality  or  natality 

Animal  condition 

Population  dynamics 

Multiple-Species  Characteristics 

Species  richness 

Species  diversity 

Biomass 

Guilds  or  life-forms 

Community  suitability  index 


Dead  Vegetation.  Various  types  of  dead  vegetation, 
such  as  snags,  litter,  and  downed  logs,  are  becoming 
increasingly  more  important  to  wildlife  biologists. 
As  a  result,  measures  of  these  components  are  being 
used  increasingly  in  habitat  models.  However,  few 
formal  systems  for  quantifying  such  components  are 
available,  and  in  many  cases  the  biologist  must  de- 
vise ad  hoc  measurement  systems.  These  habitat 
variables  are  described  in  more  detail  in  Chapter  27, 
Terrestrial  Physical  Features. 

Physical  Features.  Physical  features  are  important 
habitat  components  in  both  aquatic  and  terrestrial 
systems.  Because  of  their  importance,  two  chapters 
in  this  book  have  been  devoted  to  describing  sys- 
tems for  measuring  them  (Chapter  27,  Terrestrial 
Physical  Features,  and  Chapter  28,  Aquatic  Physical 
Features).  In  addition,  because  of  its  central  import- 


ance, an  entire  chapter  has  been  devoted  to  the  sub- 
ject of  soils  (Chapter  26). 


Water.   Many  terrestrial  wildlife  species  require  free 
water.  Therefore,  presence  of  free  water  is  an  impor- 
tant habitat  component  for  these  species.  This  may 
be  expressed  as  a  distance  from  free  water,  as  a  den- 
sity of  springs,  or  as  many  other  factors.  The  proper- 
ties of  the  water  are  generally  not  important.  With 
amphibious  and  aquatic  species,  however,  properties 
of  water  such  as  turbidity,  temperature,  and  pH  be- 
come important.  Therefore,  measurements  of  theses 
properties  are  described  in  a  separate  chapter  (Chap- 
ter 30,  Water  Quality ).  In  addition,  two  other  chap- 
ters deal  with  water-related  variables.  Chapter  29, 
Hydrological  Properties,  deals  with  dynamic  water 
variables  such  as  flow  rates  that  are  important  for 
many  wildlife  and  fish  species;  Chapter  32,  Macroin- 
vertebrates,  focuses  on  using  measurements  of  ma- 
croinvertebrates  as  indicators  of  water  quality  and 
other  habitat  deterioration. 

Food  Supply.  Food  is  such  an  important  factor 
determining  presence,  absence,  or  abundance  of 
wildlife  species  that  it  is  usually  incorporated  into 
habitat  models,  either  implicitly  or  explicitly.  When 
animal  species  are  simply  associated  with  a  vegeta- 
tion or  cover  type,  or  with  a  habitat  layer  within 
a  vegetation  type,  the  model  assumes  the  vegetation 
type  provides  an  adequate  food  supply.  At  the  other 
extreme,  food  supply  may  be  so  important  to  some 
species  that  actual  measurements  or  estimates  of  the 
food  supply  must  be  made.  Schroeder  ( 1984),  for 
example,  describes  a  winter  model  for  black  brant 
(Branta  bernicla)  that  uses  two  habitat  variables, 
one  of  which  is  the  percentage  of  cover  of  the  for- 
age eelgrass  (Zostera  marina).  In  the  case  of  preda- 
tory animals,  the  density  or  abundance  of  prey 
species  may  be  the  important  habitat  variable.  Use  of 
food  supply  as  a  habitat  variable  is  probably  best 
developed  for  large  herbivores  such  as  elk  (Cervus 
elaphus)  and  deer.  With  these  animals,  some  habitat 
models  (frequently  called  carrying-capacity  models) 
predict  the  number  of  animals  that  can  be  supported 
on  a  range  by  using  food  supply  as  a  factor.  Some  of 
these  models  not  only  require  data  on  weight  of 
individual  plant  species  present  by  season,  but  also 
data  on  the  nutritional  content  of  various  plant  spe- 
cies or  species  groups.  Such  models  are  discussed  in 
more  detail  in  Chapter  25,  Ungulates. 

Presence,  Absence,  or  Abundance  of  Com- 
petitors. The  presence  of  competitors  is  rarely  con- 
sidered formally  in  habitat  models.  Yet,  the 
phenomena  of  competition  is  recognized  by  many 
biologists  as  an  important  factor  affecting  distri- 
bution or  abundance  of  animal  species.  There  is,  of 
course,  much  disagreement  among  biologists  about 
the  importance  of  competition  and  the  mechanism 
by  which  competition  occurs.  Similarly,  many  biolo- 


Habitat  Evaluation  Systems 


761 


gists  do  not  consider  presence  of  a  competitor  as  a 
"habitat  component."  However,  in  the  context  of 
habitat  models,  predators  may  be  treated  as  habitat 
components  just  like  any  other.  For  example,  many 
species  of  cavity-nesting  songbirds  are  adversely 
affected  by  the  presence  of  starlings  (Sturnus  vul- 
garis). Removal  of  starlings  can  improve  the  habitat 
for  many  of  these  species  and  allow  them  to  become 
more  abundant.  Therefore,  habitat  models  for  such 
species  must  consider  the  influence  of  the  presence 
or  abundance  of  starlings  in  the  habitat. 


Presence,  Absence,  or  Abundance  of  Predators. 

Like  competitors,  predators  are  not  usually  consid- 
ered in  formal  quantitative  habitat  models,  but  are 
often  discussed  in  verbal  or  written  models.  As  with 
competition,  biologists  disagree  about  the  role  and 
mechanism  of  predation  in  limiting  animal  popula- 
tions. In  any  case,  if  a  biologist  believes  that  a  preda- 
tor is  or  could  limit  a  population,  presence  or 
abundance  of  the  predator  should  be  measured  and 
included  as  a  habitat  factor  in  the  model.  Human 
hunting,  which  may  be  considered  a  special  case  of 
predation,  is  discussed  below. 

Presence,  Absence,  or  Abundance  of  Parasites 
or  Diseases.  Like  the  previous  two  "habitat  compo- 
nents," parasites  or  diseases  are  rarely  considered 
in  formal  models.  However,  the  role  of  parasites  or 
disease  in  limiting  animal  populations  is  well-docu- 
mented in  many  specific  cases.  As  with  competition 
and  predation,  an  impact  to  a  wildlife  population 
from  parasites  or  disease  is  often  correlated  with,  if 
not  caused  by,  a  change  in  physical  habitat  condi- 
tions. In  the  case  of  disease,  however,  very  little 
study  has  been  done  on  correlating  disease  impacts 
with  changes  in  habitat. 

Presence,  Absence,  or  Degree  of  Human  Pis- 
turbance.  Disturbance  includes  a  wide  variety  of 
factors  other  than  physical  or  vegetative  features  of 
the  environment,  such  as  noise,  highway  construc- 
tion, drilling,  etc.  In  many  cases,  these  are  very  im- 
portant factors  affecting  abundance  of  vertebrates.  In 
fact,  they  are  frequently  factors  that  biologists  are 
asked  to  quantify  in  analyses  such  as  EISs  and  EARs. 
Yet  very  few,  formally  published  habitat  models  in- 
clude disturbance  as  an  explicit  habitat  component. 
Some  models  do  consider  disturbance  as  an  implicit 
factor,  however.  For  example,  Schroeder  ( 1984) 
developed  an  HSI  model  for  black  brant  in  which 
roosting  cover  was  one  of  the  habitat  variables.  Hu- 
man disturbance  has  also  been  identified  as  an  im- 
portant factor  limiting  black  brant  populations. 
Therefore,  in  the  model,  roosting  cover  is  defined  as 
the  "percent  of  shoreline  .  . .  that  contains  sandy 
areas  that  are  isolated  from  human  disturbance" 
(Schroeder  1984).  Thus  disturbance  is  considered  as 
either  present  or  absent. 


A  more  challenging  problem  is  determining  the 
effects  of  a  given  degree  of  disturbance  on  a  wildlife 
population.  Few,  if  any,  formal  models  have  even 
attempted  to  do  this.  Yet,  such  information  is  needed 
for  biologists  to  evaluate  the  impacts  to  wildlife  from 
the  myriad  human  activities  occurring  in  wildlife 
habitats.  I  would  expect  that  in  future  years  biolo- 
gists will  be  developing  and  using  more  models  that 
incorporate  disturbance  as  a  habitat  factor. 


Presence,  Absence,  or  Intensity  of  Hunting  or 
Harvesting.   Like  other  factors  above,  hunting  is  not 
generally  considered  a  habitat  factor.  However,  in 
the  context  of  habitat  models,  it  is  useful  to  consider 
it  such,  since  it  obviously  can  limit  animal  popula- 
tions. Its  impact  is  also  greatly  interdependent  on 
physical  and  vegetative  factors,  such  as  the  amount 
or  quality  of  cover.  If  hunting  pressure  is  limiting 
wildlife  populations,  then  habitat  models  that  only 
consider  physical  or  vegetative  habitat  components 
will  be  poor  predictors  of  population  responses  to 
habitat  management  or  disturbance.  Therefore,  in 
developing  and  using  habitat  models,  biologists  must 
first  consider  if  hunting  is  limiting  the  population.  If 
hunting  is  not  a  factor,  which  is  the  case  with  vir- 
tually all  nongame  species  and  many  game  species  in 
this  country,  the  biologist  may  safely  omit  it  from 
further  consideration.  However,  if  it  is  or  appears  to 
be  a  factor,  the  biologist  must  take  it  into  account. 
Analyzing  impacts  of  habitat  alteration  on  wildlife 
populations  has  little  value  when  physical  or  vegeta- 
tive habitat  factors  are  not  limiting. 


Weather  and  Climate.  Weather  is  the  state  of  the 
atmosphere  at  a  given  time;  climate  refers  to  the 
characteristics  of  the  atmospheric  conditions  of  a  re- 
gion. Or,  as  one  wag  described  it,  "climate  is  what 
you  expect;  weather  is  what  you  get."  Whereas  cli- 
mate is  relatively  predictable,  weather  is  not,  and 
weather  strongly  influences  many  wildlife  popula- 
tions. Furthermore,  as  Bailey  (1984)  pointed  out, 
even  though  the  biologist  or  manager  has  no  control 
over  the  weather,  when  devasting  impacts  occur  to 
wildlife  populations  from  severe  weather  conditions, 
the  manager  may  be  blamed,  "for  there  is  no  satisfac- 
tion in  blaming  the  weather."  Therefore,  in  inventory 
and  monitoring,  biologists  must  incorporate  weather 
measurements  or  considerations  into  their  study 
design  whenever  dealing  with  populations  that  are 
heavily  influenced  by  weather.  Because  of  the  impor- 
tance of  weather  and  climate  in  determining  animal 
abundance,  an  entire  chapter  in  this  book  (Chapter 
35,  Weather  and  Climate)  has  been  devoted  to  the 
subject.  Weather  is  not  included  explicitly  in  most 
formal  habitat  models;  however,  it  is  generally  recog- 
nized as  a  concomitant  influence  that  affects  animal 
abundance  in  many  ways  and,  under  extreme  condi- 
tions, may  override  the  importance  of  all  habitat 
variables. 


762 


Habitat  Evaluation  Systems 


Historical  Occurrence.  Historical  occurrence  is 
not  generally  thought  of  as  a  habitat  component,  yet 
it  can  be  an  important  characteristic  of  habitat.  If 
an  animal  species  has  been  recorded  as  occurring  on 
an  area  within  historic  times,  then  such  information 
is  excellent  evidence  that  the  species  can  or  does 
occur  there  as  long  as  the  physical  habitat  has  not 
been  drastically  altered.  Of  course,  the  shorter  the 
time  since  that  occurrence  has  been  verified,  the 
more  probable  it  is  that  the  species  persists  there. 
Even  when  a  species  is  known  to  have  been  extir- 
pated, historical  occurrence  is  important  information 
when  considering  such  things  as  reintroductions. 
Introduction  of  animals,  such  as  peregrine  falcons 
(Falco  peregrinas)  and  bighorn  sheep  (Ovis  cana- 
densis) into  areas  that  were  not  historic  habitats,  has 
rarely  been  successful. 


Model  Components — Population  Attributes 

As  with  the  habitat  components  of  a  habitat  model, 
many  population  attributes  can  be  predicted  (Table 
2  ).  These  may  be  either  characteristics  of  a  single 
species  population,  multiple  species  populations,  or 
community  characteristics,  such  as  species  richness. 
However,  unlike  with  habitat  components,  most  of 
these  models  can  only  predict  one  population  or 
community  attribute.  In  other  words,  several  habitat 
components  may  be  used  as  predictors,  but  only  one 
population  attribute  is  normally  predicted. 

Most  models  have  been  developed  to  predict 
attributes  of  single  species.  In  fact,  most  wildlife 
research  and  management  has  been  single-species 
oriented.  Furthermore,  even  many  multiple  species 
models,  such  as  those  for  species  richness,  are  devel- 
oped from  a  set  of  single-species  models. 

Presence  or  Absence.  The  presence  or  absence  of 
a  species  is  the  simplest  sort  of  population  attribute 
that  can  be  predicted.  However,  for  many  purposes, 
such  a  determination  is  adequate  information  for  the 
majority  of  wildlife  species. 

Probability  of  Occurrence.  Probability  of  occur- 
rence considers  presence  or  absence  as  a  probability 
rather  than  as  an  absolute.  Since  models  produce 
predictions  rather  than  absolute  truths,  uncertainty  is 
always  associated  with  their  predictions;  use  of 
models  that  predict  occurrence  probability  are  more 
realistic  because  they  attempt  to  quantify  the  uncer- 
tainty associated  with  the  predictions.  For  example,  a 
biologist  may  need  to  predict  if  an  area  is  used  by 
prairie  falcons  for  nesting.  Yet  time  is  not  available 
to  make  an  absolute  determination  since  the  analysis 
must  be  made  in  winter  and  the  proposed  impact 
will  occur  during  the  next  breeding  season.  Using  a 
model  that  has  probability  of  occurrence  as  an  out- 
put, a  biologist  can  predict,  based  on  habitat  compo- 
nents, that  "the  probability  of  the  site  being 


occupied  by  prairie  falcons  is  90% ."  Models  for 
probability  can  be  used  for  other  purposes,  such  as 
predicting  the  probability  that  a  population  of  a  cer- 
tain density  occurs.  The  most  common  use  of  proba- 
bility measures  of  animal  populations  as  predicted 
outputs  is  in  "pattern  recognition  (PATREC) 
models,"  discussed  in  more  detail  later  in  this 
chapter. 

Abundance  or  Density.  Abundance  here  refers  to 
the  number  of  animals  in  a  population,  whereas  den- 
sity refers  to  the  number  of  animals  per  unit  area. 
Abundance  or  density  is  quite  difficult  to  predict. 
Thus,  very  few  models  attempt  to  actually  predict 
such  attributes.  In  most  situations,  biologists  neither 
can  nor  need  to  make  such  precise  estimates. 

Relative  Abundance  or  Density  Index.  Abun- 
dance or  density  indexes  are  values  that  correlate 
with  abundance  or  density,  but  for  which  the  quanti- 
tative relationship  between  the  two  is  unknown. 
For  example,  an  area  with  a  mule  deer 
(O.  hemionus)  index  of  0.5  may  be  expected  to 
have  twice  as  many  deer  as  an  area  with  an  index  of 
0.25.  However,  the  relationship  between  the  index 
and  the  actual  number  of  deer  is  unknown.  If,  for 
example,  an  index  of  0.5  was  known  to  correspond 
with  a  density  of  30  deer  per  square  mile,  then  it 
would  not  be  a  density  index,  but  rather  a  predictor 
of  density.  Abundance  or  density  indexes  in  habitat 
models  (predicted  density  indexes),  like  the  meas- 
ured abundance  or  density  indexes  described  in 
Chapter  2,  Data  Types,  are  most  useful  for  compari- 
sons. A  biologist  can  compare  different  areas  or  the 
same  area  during  different  years  or  under  varying 
habitat  impacts. 

Carrying  Capacity.  Carrying  capacity  is  considered 
by  range  managers  to  be  the  number  of  animals  an 
area  can  support.  It  is,  thus,  a  property  of  the  land 
and  the  condition  of  the  habitat,  but  is  expressed  in 
terms  of  numbers  or  density  of  animals.  The  actual 
number  of  animals  present  may  be  lower  than  the 
carrying  capacity.  Carrying  capacity  is  unaffected,  for 
example,  by  weather,  hunting,  or  other  decimating 
factors  that  are  not  habitat  related.  Although  the 
term  has  generated  much  controversy  and  discussion 
(Bailey  1984:279-303),  it  remains  a  useful  working 
concept  that  allows  a  biologist  to  quantify  the  value 
of  habitat  without  considering  short-term  fluctua- 
tions in  animal  numbers  from  weather,  hunting,  or 
predation.  Carrying  capacity  has  been  used  as  an 
output  in  many  models  (models  developed  for  ungu- 
lates in  which  carrying  capacity  is  assumed  to  be 
primarily  a  function  of  forage  supply). 

Carrying-Capacity  Index.  Carrying  capacity  in- 
dexes are  similar  to  density  indexes  except,  like  with 
carrying-capacity  estimates,  they  are  strictly  a  func- 
tion of  habitat.  The  most  common  carrying-capacity 


Habitat  Evaluation  Systems 


763 


indexes  are  the  habitat  suitability  indexes  used  in 
the  U.S.  Fish  and  Wildlife  Service  HSI  models 
(Schamberger  et  al.  1982).  These  models  use  an 
index  that  varies  from  0  to  1 ,  with  1  being  some 
maximum  carrying  capacity,  and  0  indicating  unsuit- 
able habitat. 

Biomass.   Biomass  is  rarely  used  in  habitat  models 
since  it  is  usually  not  of  interest  to  managers,  at  least 
with  terrestrial  species.  However,  it  is  often  used  in 
ecological  models  designed  for  understanding  eco- 
system processes.  Nonetheless,  some  management 
models  do  exist,  particularly  in  fisheries.  For  exam- 
ple, Binns  (1979)  describes  a  model  that  uses  nine 
stream  habitat  variables  to  predict  the  standing  crop 
of  trout. 

Geographic  Distribution.  Geographic  distribution 
is  basically  an  extension  of  presence  or  absence.  In 
presence  and  absence  models,  the  biologist  merely 
predicts  whether  a  species  will  be  present  on  a 
given  site.  By  using  habitat  attributes  at  a  series  of 
sites,  one  can  predict  the  geographic  area  that  a 
wildlife  species  can  be  expected  to  occupy.  One 
may  also  note  that  "geographic  location"  and  "his- 
toric distribution"  are  also  described  as  a  model 
input.  Thus,  according  to  the  description  of  models 
provided  here,  historic  distribution  can  be  used, 
for  example,  to  predict  geographic  distribution.  This 
may  seem  to  be  rather  trivial.  However,  this  is,  in 
fact,  a  rather  commonly  used  model,  even  though  it 
is  rarely  identified  as  such.  The  implicit  model,  used 
by  most  biologists  as  historical  records  of  a  species 
in  an  area,  is  the  best  predictor  of  future  occurrence, 
and  the  more  recent  the  historical  observation,  the 
greater  the  predictive  power.  However,  as  discussed 
under  historical  occurrence,  past  observation  is  not  a 
perfect  predictor;  conditions  often  change,  resulting 
in  local  extirpations,  and  the  biologist  needs  to  be 
alert  to  such  situations. 

Mortality  or  Natality.  Some  models  predict  the 
mortality  or  natality  rather  than  a  population  density. 
Mortality  is  usually  expressed  in  terms  like  number, 
density,  or  abundance  of  animals  killed  over  winter. 
Similarly,  natality  is  usually  expressed  as  some  meas- 
ure of  productivity,  such  as  young  fledged  per  nest, 
or  as  some  ratio  of  young  to  old  animals,  such  as 
fawns  per  doe.  When  used,  both  types  of  measure- 
ments are  considered  useful  indicators  of  the 
"health"  of  the  species  population. 

Animal  Condition.  Animal  condition,  expressed  in 
such  terms  as  weight,  kidney  fat  indexes,  etc.,  is 
also  used  as  a  model  output  and,  like  productivity,  is 
considered  a  surrogate  measure  of  the  "health"  of 
the  population. 

Population  Dynamics.   In  some  cases,  biologists 
may  be  more  interested  in  predicting  the  dynamics 
of  a  population  rather  than  a  static  or  fixed  popula- 


tion level.  All  populations  fluctuate  in  numbers  and, 
in  some  cases,  the  frequency  and  amplitude  of  such 
fluctuations  may  be  of  more  interest  than  the  "aver- 
age" population  size.  Although  numerous  population 
dynamics  models  are  available,  most  do  not  use  any 
physical  or  vegetative  habitat  features  to  predict 
population  responses.  They  use  weather,  hunting 
pressure  or,  in  some  cases,  predation.  Thus  they  are 
not  commonly  considered  habitat  models. 

Very  few  models  attempt  to  predict  changes  in 
wildlife  populations  over  time  as  a  result  of  changes 
in  habitat  factors,  although  a  few  are  available 
(Cooperrider  and  Bailey  1984;  Cooperrider  and  Beh- 
rend  1980).  Models  with  dynamic  outputs  (ex- 
pressed as  changes  over  time )  require  at  least  one 
dynamic  input  (changes  independently  of  the 
model).  Such  inputs  are  said  to  "drive"  the  model.  In 
most  dynamic  or  simulation  models,  weather  is  the 
dynamic  input;  however,  other  inputs  can  be  used. 
The  model  of  Cooperrider  and  Bailey  ( 1984),  for 
example,  can  use  forage  production  to  drive  the 
model. 


Model  Components — Multiple  Population 
or  Community  Characteristics 

Characteristics  of  several  populations  are  the 
plural  counterparts  of  the  population  characteristics 
just  described.  They  are  predicted  from  similar  habi- 
tat attributes.  However,  instead  of  making  inferences 
about  only  one  species,  some  inference  is  made 
about  ( 1 )  several  taxonomically  related  species  in  an 
area  (e.g.,  songbirds,  ungulates);  (2)  several  species 
that  use  similar  habitat  components  (e.g.,  cavity- 
nesting  birds);  or  (3)  all  species  in  a  habitat. 

Species  Richness.  Species  richness  is  the  number 
of  species  present  in  an  area.  It  is  the  plural  counter- 
part of  presence  or  absence.  Models  that  predict 
species  richness  are  usually  composed  of  a  series  of 
individual  species  presence  and  absence  models,  i.e., 
the  models  usually  not  only  predict  the  number  of 
species,  but  also  the  specific  species. 

Species  Diversity.  Diversity  or  species  diversity 
refers  to  not  only  the  number  of  species  but  also 
their  relative  abundance  in  an  area.  Species  diversity 
is  rarely  used  directly  as  an  output  from  habitat 
models,  at  least  the  type  of  habitat  models  used  in 
land  management.  This  is  not  surprising,  since  a 
model  to  predict  species  diversity  must  predict  not 
only  presence  or  absence  but  also  density  or  relative 
abundance  of  each  species  present.  Some  models, 
however,  are  based  on  the  premise  that  diversity  of 
vegetation  structure  is  correlated  with  species  diver- 
sity as  measured  by  some  species  diversity  index. 
(See  Chapter  2  for  a  discussion  of  diversity  indexes.) 
These  models  predict  an  index  of  species  diversity 


764 


Habitat  Evaluation  Systems 


without  explicitly  predicting  the  species  and  their 
relative  abundance. 


Biomass.  Biomass  of  all  species,  or  all  species 
within  a  denned  species  group  such  as  small  mam- 
mals, fishes,  or  ungulates,  is  sometimes  used  as  a 
model  output.  As  with  individual  species  biomass,  it 
is  used  most  commonly  with  fish. 


Guilds  or  Life-Forms.  Guilds  and  life-forms  are 
both  groups  of  species  that  use  similar  habitat 
components  for  feeding,  breeding,  or  other 
biological  functions.  Several  systems  have  been 
devised  that  use  guilds  or  life-forms  as  outputs,  as 
described  in  the  next  section.  The  advantage  to  the 
biologist  is  that  he  or  she  can  deal  with  10  or  20 
guilds,  rather  than  400  or  500  vertebrate  species  in 
an  area.  Predictions  can  be  made  and  verified  in 
terms  of  impacts  on  whole  guilds  rather  than  on 
individual  species.  Life-forms,  as  discussed  later  in 
the  chapter,  may  be  considered  a  special  case  of 
guilds. 


Community  Suitability  Index.  In  some  cases,  a 
rating  of  the  relative  value  of  a  community  to  sup- 
port a  wildlife  community  is  the  output  from  a 
model.  Such  indexes  are  either  implicitly  or  explic- 
itly thought  to  be  correlated  with  a  population  meas- 
urement, such  as  species  richness  or  species 
diversity.  For  example,  Short  (1984)  describes  a 
model  for  generating  a  suitability  index  for  a  habitat 
based  on  the  layers  of  habitat  present.  It  is  further 
based  on  the  assumption  that  structurally  more  com- 
plex habitats  have  more  wildlife  guilds  and  greater 
species  richness. 


Model  Components — Correlation  Functions 

The  precision,  accuracy,  and  usefulness  of  habi- 
tat models  depend  on  the  selection  of  the  appropri- 
ate habitat  components  for  use.  Equally  important, 
however,  is  the  process  of  linking  these  variables 
together  in  a  meaningful  way  so  the  best  prediction 
can  be  made.  A  thorough  description  and  set  of 
guidelines  on  model  development  and  construction 
is  beyond  the  scope  of  this  publication.  However, 
a  few  critical  concepts  and  principles  are  discussed 
under  the  Developing  Habitat  Models  section,  fol- 
lowing. For  a  more  thorough  treatment  of  model 
building,  see  the  U.S.  Fish  and  Wildlife  Service  Eco- 
logical Services  Manual  on  "Standards  for  the  De- 
velopment of  Habitat  Suitability  Index  Models  ( HSI )" 
(U.S.  Department  of  the  Interior,  Fish  and  Wildlife 
Service  1981 ),  which  contains  a  good  introduction 
to  the  subject.  That  discussion  focuses  on  one  partic- 
ular type  of  model;  however,  the  procedures  can 
generally  be  applied  to  other  classes  of  models. 


DEVELOPING  HABITAT  MODELS 

The  following  basic  steps  can  be  used  in  devel- 
oping a  model: 


( 1 )  Determine  model  objectives; 

(2)  Select  and  quantify  habitat  variables; 

(3)  Determine  the  correlation  or  prediction  func- 
tion; 

(4)  Document  and  verify  the  model. 


These  steps  correspond  roughly  to  the  phases  of 
construction  described  in  the  U.S.  Department  of 
the  Interior,  Fish  and  Wildlife  Service  Ecological 
Services  Manual  (  1981 ). 

Determining  Model  Objectives 

Model  objectives  should  be  based  on  the  objec- 
tives of  the  inventory  and  monitoring  effort,  as  de- 
scribed in  more  detail  in  Chapter  1,  Inventory  and 
Monitoring  Process.  The  process  includes  determin- 
ing the  desired  output  (e.g.,  species  density,  relative 
density,  species  richness)  as  well  as  the  geographic 
area,  seasons,  and  other  factors  the  model  should 
predict. 

Selecting  and  Quantifying  Habitat  Variables 

Determining  habitat  variables  that  should  be 
incorporated  into  a  model  is  critical.  A  biologist 
should  use  the  least  number  possible,  since  habitat 
measurements  are  expensive  to  obtain.  These  varia- 
bles should  be  the  "limiting  factors"  as  described 
in  basic  ecology  texts  (Odum  1969)  or  surrogates 
for  them. 

The  concept  of  a  surrogate  variable  is  important 
in  model  construction.  A  surrogate  variable  is  one 
that  closely  relates  to  a  habitat  variable  of  interest 
but  which,  for  one  reason  or  another,  is  easier  to 
measure  or  verify.  For  example,  a  Lewis'  woodpecker 
(Melanerpes  leivis)  is  reported  to  need  forests  hav- 
ing tree  canopy  closures  less  than  75%  and  does 
best  in  areas  with  0  to  30%  closure  (Sousa  1983). 
However,  within  the  range  of  0  to  75%  canopy  clo- 
sures, the  basal  area  of  trees  highly  correlates  with 
canopy  closure  in  many  areas.  Furthermore,  basal 
area  is  usually  easier  to  measure  than  canopy  closure 
or  may  be  available  from  sources  such  as  timber 
inventories.  Therefore,  the  biologist  may  prefer  to 
substitute  basal  area  as  a  surrogate  for  canopy 
closure. 

If  a  model  is  being  used  to  predict  the  impacts 
of  a  particular  management  action,  then  at  least 
some  of  the  habitat  variables  selected  should  be 


Habitat  Evaluation  Systems 


765 


those  affected  by  the  management  action  in  some 
predictable  way. 

The  habitat  variables  then  must  be  carefully 
defined.  To  convert  a  conceptual  or  verbal  model 
into  a  written  or  even  quantitative  model  generally 
requires  carefully  defining  the  variables.  For  exam- 
ple, basic  habitat  factors  required  by  bighorn  sheep 
for  survival  have  been  described  as  food,  water,  and 
escape  terrain  (Hansen  1980).  Escape  terrain  has 
been  further  described  by  various  authors  as  open, 
steep,  rocky,  and  rugged  terrain.  The  biologist  must 
define  or  quantify  the  terms  if  the  model  is  to  be 
used  by  other  biologists  or  if  habitat  measurements 
are  to  be  repeated.  "Open"  can  be  defined  as  the 
percentage  of  canopy  cover  of  trees  and  shrubs  or  as 
some  measure  of  visual  obstruction.  Similarly, 
"steep"  can  be  defined  as  percentage  of  slope.  The 
evolution  of  models  from  concepts  to  quantitative 
descriptions  requires  carefully  defined  and  eventually 
quantified  terms  (Figure  2). 


Determining  the  Correlation  or  Prediction 
Function 

Once  habitat  variables  have  been  selected,  de- 
fined, and  quantified,  the  biologist  must  determine 
the  relative  importance  of  each  to  the  others.  In  the 
simplest  case,  a  biologist  can  determine  that  all  vari- 
ables are  equally  important;  these  variables  can  then 
be  added  to  provide  a  simple  additive  model  (Figure 
3a).  On  the  other  hand,  some  habitat  factors  may 
be  considered  more  important  than  others.  For  ex- 
ample, the  biologist  may  decide  that  food  supply 
is  twice  as  important  as  escape  terrain.  By  attaching 
a  number  to  the  variable,  the  biologist  can  weigh  the 
model  so  the  index  produced  is  influenced  more  by 
food  supply  than  other  variables,  resulting  in  a 
weighted  additive  model  (Figure  3b). 

The  relationship  between  variables  may  be 
much  more  complicated,  however,  and  the  biologist 
should  try  to  model  it.  For  example,  the  additive 


American  Dipper  Habitat 


a.  Conceptual  Model 


b.  Verbal  Model 


c.  Quantitative  Model 


"Fast  flowing  streams  in  or 
near  mountains." 


"Streams  of  over  5  cfs  above 
5,000-ft  elevation  within  6 
miles  of  mountains." 


Figure  2.     Progression  of  models  from  conceptual  to  quantitative. 
766  Habitat  Evaluation  Systems 


relationship  shown  for  desert  bighorn  sheep  in  Fig- 
ure 3a  may  not  be  realistic.  A  habitat  could  have  no 
water  at  all  and  yet  have  other  important  compo- 
nents. The  index  value  derived  would  still  be  0.66 
for  equation  (a)  or  0.75  for  equation  (b),  which  are 
both  fairly  high  ratings.  Yet  such  habitats  do  not 
support  bighorn  sheep.  For  these  reasons,  a  multipli- 
cative or  limiting  factor  model  (Figures  3c  and  3d) 
may  be  more  appropriate.  With  such  a  model,  if  any 
attribute  has  a  suitability  rating  of  0,  the  overall  in- 
dex will  become  0. 


Habitat  components  can  be  combined  in  unlim- 
ited ways  using  various  weighting  factors  and  mathe- 
matical functions.  Similarly,  seasonal  or  functional 
requirements  can  be  combined  to  form  compound 
models.  For  example,  summer  habitat  requirements 
can  be  combined  with  winter  habitat  requirements, 
or  feeding  habitat  requirements  with  breeding  habi- 
tat requirements. 

One  of  the  most  difficult  problems  of  combining 
habitat  attributes  is  taking  into  account  spatial  ar- 
rangement of  habitat  components.  Biologists  know 
that  interspersion  and  juxtaposition  of  habitat  fea- 
tures can  be  very  important  to  animals.  However, 
quantifying  these  relationships  in  a  meaningful  way  is 
difficult.  The  use  of  geographic  information  systems 
(GISs)  to  analyze  these  relationships  may  assist 
greatly  in  the  future. 


Numerous  statistical  and  numerical  techniques 
can  assist  in  determining  both  the  form  of  the  func- 
tion and  the  appropriate  weighting.  Techniques  such 
as  linear  regression,  multivariate  analysis,  or  least- 
squares  curve  fitting  may  all  be  useful  when  data  are 
available.  However,  in  many  cases,  the  biologist  must 
rely  on  literature,  expert  opinion,  and  existing 
sources  of  information,  rather  than  raw  data. 

Documenting  and  Verifying  the  Model 

Documenting  a  model  is  important.  In  many 
cases,  models  are  developed  by  using  a  group  of  spe- 
cies experts.  Even  though  every  function  or  assump- 
tion of  the  model  is  not  backed  up  by  extensive 
research,  a  model  that  represents  an  expert  concen- 
sus is  easy  to  defend.  If  the  model  is  being  con- 
structed for  some  sort  of  impact  analysis,  then 
getting  interested  parties  to  design  or  evaluate  the 
model(s)  can  be  most  helpful  in  obtaining  project 
impact  concensus.  Documentation  is  also  important 
when  made  available  to  other  biologists  with  similar 
problems. 

A  model,  once  constructed,  is  basically  no  more 
than  an  untested  or  working  hypothesis.  Ideally,  all 
models  would  be  thoroughly  tested  and  refined  be- 
fore use.  In  practice,  most  decisions  need  to  be 
made  with  incomplete  information.  However,  moni- 
toring can  be  used  to  verify  models,  as  discussed 
in  the  next  section. 


Type  of  Model 


a.  Simple  Additive  Model--- -HSI  =     Vl  +  ^2  +  Va 

b.  Weighted  Additive HSI  =    2Vl  +  ^  +  V3 

Model  4 

c.  Multiplicative  Model HSI  =  (V,   x  V2    x  V3)1 

d.  Limiting  Factor  Model HSI  =  Min  (V,,  V2,  V3) 


All  four  models  calculate  a  habitat  suitability  index  (HSI) 
from  0  to  1  for  desert  bighorn  sheep  based  on  three 
habitat  factors — a  forage  index  (V,),  an  escape  terrain 
index  (V2),  and  a  water  availability  index  (V3);  the  latter 
indexes  also  range  from  0  to  1. 


* 


Without  water,  other  important  habitat  components  for 
desert  bighorn  sheep  are  not  sufficient. 


Figure  3.     Types  of  quantitative  models. 


Habitat  Evaluation  Systems 


767 


THE  HABITAT  INVENTORY  AND 
MONITORING  CYCLE  AND  HABITAT 
MODELS 

The  habitat  inventory  and  monitoring  cycle  has 
been  explained  in  detail  in  Chapter  2.  It  is  reviewed 
here  to  clarify  the  central  role  of  the  habitat  model 
in  both  the  cycle  and  the  habitat  evaluation  systems. 
For  further  discussion  of  the  roles  of  models  in  mon- 
itoring, see  Salwasser  et  al.  ( 1983),  Cole  and  Smith 
(1983),  and  O'Neil  and  Schamberger  ( 1983). 

The  habitat  inventory  and  monitoring  cycle  may 
be  viewed  as  consisting  of  five  steps: 


(1)  Scoping, 

(2)  Data  collection  and  analysis, 

( 3 )  Prediction, 

(4)  Decision/action,  and 

(5)  Monitoring  (Figure  4). 


These  steps  correspond  roughly  to  the  stages  out- 
lined in  Chapter  1 ,  but  emphasize  model  develop- 
ment, use,  or  refinement.  Habitat  models  and 
evaluation  systems  are  used  in  all  steps,  but  most 
importantly  in  step  3,  where  they  are  the  basic  tools 
for  predictions.  Similarly,  monitoring  verifies  predic- 
tions and  can  also  be  used  to  refine  model(s). 

Scoping 

The  driving  force  for  a  monitoring  program  is  a 
"problem"  or  a  proposed  action  of  some  sort.  This 
may  be  a  very  small  action,  such  as  a  new  fence,  or  a 
major  action,  such  as  a  power  plant.  It  may  be  a 
very  loosely  defined  issue,  such  as  a  "need  to  reverse 
habitat  deterioration  which  is  causing  perceived 
decline  of  deer  in  an  area,"  or  it  may  be  a  very  spe- 
cific proposal  with  specified  alternatives.  It  may  also 
be  a  wildlife  problem,  such  as  the  deer  decline,  or 
an  activity,  such  as  a  coal  lease  that  is  expected  to 
adversely  affect  the  wildlife  resource.  A  typical  prob- 
lem is  generally  the  need  to  know  the  resources 
present  on  an  area  for  land-use  planning,  etc.  This  is 
the  driving  force  for  a  basic  inventory.  All  these 
problems,  big  and  small,  well-defined  or  vague, 
should  result  in  the  same  procedure.  An  important 
principle  is  that  the  magnitude  and  intensity  of  the 
data  collection  and  analysis  should  correlate  with  the 
perceived  impact  of  the  problem. 

The  next  step  is  to  clearly  define  the  problem. 
This  will  generally  require  some  assessment  of  the 
wildlife  resources  (animals  or  habitat)  present, 
which  are  likely  to  be  affected.  This  assessment  fre- 
quently is  based  on  some  rapid  examination  or  anal- 
ysis of  available  data.  A  critical  initial  decision  is  to 


Problem 


Scoping 


Data    Collection 
and   Analysis 

1  ' 

Monitoring 

Prediction 
of  Impacts 

n 

Decision  /  Action 

-* 

Figure  4.     The  habitat  inventory  and  monitoring 
cycle. 


determine  the  geographical  area  of  concern  and 
"bound  the  study  area."  This  process  of  taking  a 
vague  problem  or  proposed  action,  clarifying  it,  and 
making  an  initial  assessment  of  potential  resource 
impacts  is  commonly  termed  "scoping." 


Scoping  may  significantly  alter  the  direction  of  a 
project.  By  identifying  actions  particularly  detrimen- 
tal to  wildlife  at  an  early  stage,  a  biologist  may  be 
able  to  direct  planners  to  less  detrimental  alterna- 
tives. This  effort  may  result  in  much  greater  benefits 
to  wildlife  resources  than  a  thorough  analysis  at  a 
later  stage  of  an  action. 

At  this  stage,  the  biologist  uses  habitat  models, 
possibly  very  vaguely  defined  conceptually  or  quali- 
tatively. For  example,  the  biologist  may  know  that 
forage  supply  is  a  limiting  habitat  factor  for  a  big- 
horn sheep  population.  Therefore,  he  or  she  can 
generally  predict  that  a  proposal  to  expand  a  grazing 
lease  into  a  critical  summer  bighorn  range  will  be 
detrimental  to  the  bighorn.  On  the  other  hand,  the 
biologist  may  only  have  a  very  vague  idea  of  how 
detrimental  the  impact  will  be. 

If  the  decision  is  made  to  further  consider  the 
proposal,  then  the  biologist  must  plan  the  data 
needed.  At  this  point,  more  quantitative  habitat 
models  should  be  developed.  In  the  bighorn  exam- 
ple discussed  above,  the  biologist  might  need  to 
locate  or  develop  a  model  that  related  forage  supply 
to  numbers  of  bighorn.  Forage  supply  would  then 
be  defined  more  specifically  and  quantitatively  as,  for 


768 


Habitat  Evaluation  Systems 


example,  total  pounds  of  grasses  and  perennial  forbs 
on  summer  range.  This  habitat  variable  ( and  possibly 
others  identified  in  the  model )  would  then  be  the 
ones  for  which  the  biologist  would  need  to  gather 
data.  Depending  on  the  situation,  the  biologist  might 
also  need  to  collect  data  on  the  bighorn  sheep  popu- 
lation, such  as  numbers,  movements,  or  lamb 
survival. 

Data  Collection  and  Analysis 

Data  collection  and  analysis  may  suggest  refine- 
ment of  a  preliminary  model.  For  example,  after 
intensive  field  work,  the  biologist  may  determine 
that  water  supply  is  also  limiting.  The  biologist  then 
may  need  to  modify  the  model  to  incorporate  water 
as  a  habitat  factor,  and  additional  data  on  number 
of  waters,  supplies,  etc.  may  need  to  be  collected. 

Prediction 

The  next  step  is  to  use  data  to  predict  the  effect 
of  alternative  management  actions.  This  step  requires 
using  a  model  since  the  data  collected  must  be  used 
to  predict  a  future  state.  The  models  may  be  simple 
or  complex.  Returning  to  the  previous  example, 
the  biologist  might  use  the  model  to  predict  that 
expanding  the  grazing  lease  would  remove  50%  of 
the  available  forage,  reducing  the  bighorn  population 
by  30% .  In  this  case,  the  biologist  has  used  a  quanti- 
tative carrying-capacity  model.  In  other  situations,  a 
biologists  may  be  forced  to  use  a  simple  conceptual 
model.  The  trend,  however,  is  to  use  more  formal 
and  quantitative  models,  since  these  allow  more 
precision  and  can  be  clearly  documented. 

In  the  past,  biologists  have  concentrated  on 
predicting  biological  consequences  of  actions.  Re- 
cently, biologists  and  economists  have  been  asked  to 
go  one  step  further  and  assess  economic  conse- 
quences of  these  actions.  This  also  requires  using 
models  to  attach  values  to  wildlife  resources  with 
and  without  the  action. 

Decision/ Action 

The  next  step  is  the  familiar  one  of  choosing  an 
alternative  and  taking  some  action.  Except  in  the 
case  of  a  wildlife-related  action  (such  as  a  habitat  im- 
provement project),  the  biologist  usually  has  a  minor 
role  in  this  process.  The  manager  must  weigh  poten- 
tial impacts  to  other  resources;  economics;  and  other 
social,  political,  and  legal  factors.  However,  the 
model  becomes  the  basic  tool  for  summarizing  the 
collected  data  and  making  predictions.  The  model 
becomes  the  tool  of  the  manager,  even  though  the 
latter  may  be  primarily  concerned  with  the  model 
outputs.  The  model  has,  in  effect,  synthesized  the 
assumptions  about  limiting  habitat  factors  and  their 
effects  on  wildlife  populations. 


Monitoring 

Once  an  action  has  been  taken,  a  biologist  needs 
to  monitor  the  biological  effects  of  the  action.  This 
monitoring  can  serve  two  purposes.  The  biologist 
can  determine  empirically  whether  the  impact  to 
wildlife  resources  is  as  predicted  or  within  tolerable 
limits.  If  the  action  is  not  achieving  the  desired  goal 
or  if  the  impacts  are  outside  tolerable  limits,  then 
the  manager  may  be  able  to  stop  or  modify  the  ac- 
tion being  taken.  The  biologist  can  also  determine 
the  quality  of  the  model  in  predicting  impacts  and 
modify  it  as  necessary. 

Many  biologists  have  advocated  adding  monitor- 
ing as  a  key  element  in  a  resource  management  pro- 
gram and  termed  it,  among  other  things,  "adaptive 
environmental  assessment"  (Holling  1978),  "adapt- 
ive management"  (Barrett  and  Salwasser  1982), 
"cyclic  incrementalism,"  or  more  simply  "muddling 
through"  (Bailey  1982)  or  "common  sense  manage- 
ment" (Barrett  and  Salwasser  1982).  These  ap- 
proaches vary  in  details,  but  are  all  based  on  the 
following  assumptions: 


( 1 )  Many  features  of  biological  systems  ( such  as 
weather)  are  unpredictable; 

(2)  The  tools  for  both  measuring  biological  re- 
sources and  predicting  future  states  are  crude 
and  the  time,  money,  and  personnel  for  such 
efforts  are  limited;  and 

(3)  Continuing  selective  remeasurement  (moni- 
toring )  can  be  effective  in  both  correcting  or 
improving  management  actions  and  also  refin- 
ing the  predictive  tools. 

The  need  to  monitor  may  seem  obvious;  however, 
the  list  of  projects  and  programs  that  have  failed 
because  monitoring  was  not  included  (or  was  given 
the  lowest  priority )  is  long  and  growing  fast. 

Monitoring  without  an  underlying  model  may 
be  useful  in  determining  whether  management  ob- 
jectives are  being  met,  but  it  is  unlikely  to  lead  to  a 
better  understanding  of  the  system  being  managed. 

Clearly,  the  five  steps  of  the  habitat  and  moni- 
toring cycle  are  similar  and  sometimes  overlap.  Scop- 
ing, for  instance,  is  a  form  of  low-level  or  "quick- 
and-dirty"  inventory  and  prediction  phase.  Although 
monitoring  may  involve  similar  data  collection,  it 
is  done  in  the  inventor}'  phase  and  for  different  pur- 
poses. 

In  the  next  sections,  I  describe  the  systems, 
models,  and  other  programs  that  can  be  used  when 
performing  one  or  more  of  the  tasks  identified  in  this 
section. 


Habitat  Evaluation  Systems 


769 


CURRENT  HABITAT  MODELS  AND 
EVALUATION  SYSTEMS 


example,  habitat  suitability  for  the  marten  (Martes 
americana)  is  a  function  of  four  variables: 


Models,  as  considered  here,  are  any  formal 
method  for  correlating  habitat  variables  with  popula- 
tion attributes,  whether  they  are  for  single  species 
or  for  multiple  species  or  communities,  regardless  of 
the  purpose  for  which  they  were  designed  and  used. 

Simple  Correlation  Models 

Single-species  models  range  from  very  simple 
correlations  to  rather  elaborate  correlations  of  spe- 
cies presence  with  many  habitat  factors.  In  the  sim- 
plest case,  presence  of  a  species  is  simply  associated 
with  presence  of  a  vegetation  type.  More  elaborate 
variations  take  into  consideration  other  habitat  fea- 
tures within  that  habitat  type  and  other  animal  popu- 
lation attributes.  These  include  season  of  use; 
successional  stage  of  vegetation;  special  habitat  re- 
quirements; biological  functions  for  which  they  are 
used  (breeding,  feeding,  resting  [cover],  special); 
relative  importance  or  value  of  the  habitat  feature 
(critical,  optimum,  acceptable,  marginal,  etc. );  and, 
in  some  cases,  some  qualitative  measure  of  relative 
density  or  abundance  (abundant,  common,  rare, 
etc.).  The  approach  is  exemplified  in  the  models  de- 
veloped and  reported  by  Verner  and  Boss  ( 1980; 
Figure  5). 

Habitat  Suitability  Index  (HSI)  Models 

HSI  models  have  been  developed  by  the  U.S. 
Fish  and  Wildlife  Service  and  others  for  use  in  their 
Habitat  Evaluation  Procedures  (HEP;  U.S.  Department 
of  the  Interior,  Fish  and  Wildlife  Service  1981 ).  In 
developing  the  HSI  models,  the  biologist  uses  exist- 
ing data,  literature,  and  expert  opinion  to  develop  an 
equation  or  algorithm  to  use  a  small  number  of  se- 
lected habitat  variables  in  predicting  the  suitability 
of  habitat  for  a  wildlife  species.  Suitability  is  indi- 
cated by  an  index  ranging  from  0  to  1,  with  0  indi- 
cating unsuitable  habitat  and  1,  optimal  habitat.  For 


( 1 )    Percentage  of  tree  canopy  closure, 

(  2 )    Percentage  of  overstory  canopy  containing  fir 
or  spruce, 

(  3  )    Successional  stage  of  stand,  and 

(4)    Percentage  of  ground  surface  covered  by 
downfall  over  3  in.  in  diameter  (Allen  1982; 
Figure  6). 

Since  habitat  suitability  cannot  be  directly  measured, 
validation  of  HSI  models  is  difficult  if  not  impossible. 
HSI  models  are  usually  developed  for  a  single  spe- 
cies; however,  sometimes  they  are  developed  for 
a  group  of  species  or  guilds,  such  as  dabbling  ducks, 
which  respond  similarly  to  the  same  set  of  habitat 
variables.  Approximately  300  HSI  models  are  in 
some  stage  of  development  in  North  America,  of 
which,  about  half  are  for  aquatic  species. 

Pattern  Recognition  (PATREC)  Models 

PATREC  models  are  similar  to  HSI  models  in 
that  they  use  a  set  of  habitat  variables  to  predict  the 
expected  capability  of  the  habitat  to  support  an  ani- 
mal species.  They  differ  from  HSI  models  in  that  they 
use  a  formal  statistical  procedure  to  predict  habitat 
capability,  expressed  as  a  probability.  Usually  the 
probability  is  of  a  certain  population  density  or 
classes  of  animals.  For  example,  Williams  et  al. 
( 1977)  describe  a  PATREC  model  for  pronghorn 
antelope  that  predicts  the  probability  that  a  habitat 
with  a  given  set  of  characteristics  will  support  a  high 
or  low  density  antelope  population.  By  providing 
definitions  for  high  density  and  low  density,  the 
model  can  also  be  used  to  calculate  an  expected 
population  density.  PATREC  models  are  generally 
developed  from  field  data;  however,  they  can  use 
data  from  expert  opinion. 


§ 


Opti  mum 

Suitable 

Marginal 


Jl  ac  k-headed 
Grosbeak     R 


12  3  3  4  4 
A  B  A  B 


1  2  3  3  3  4  4  4 
A  B  C  A  B  C 


rfrrHri  Wr+rrr 


rr 


12  3  3  3  4  4  4 

A  BCABC 


rrrrrrr 


1    2  3  3  4  4 
ARAB 


Ett 


H 


12  3  3  3  4  4  4 
A  B  C  A  B  C 


Wririhrt\ 


12  3  3  3  4  4  4 

A  B  C  A  BC 


12  3  3  3  4  4  4 
A  B  C  A  B  C 


12  3  3  3  4  4  4 
A  BC  A  B  C 


FUNCTION 

B-Breeding 

F-Feed  i  ng 

R-Resli  ng 

S-Season    ot    Occurrence 


SUCCESSIONAL  STAGES 
1-Grass/Forb 
2-Shrub/Seeding/Saplinc 
3-Pole /Med  i  urn    Tree 
4-Large    Tree 


TREE  CANOPY  COVER  CLASSES 

A    0-39% 

B    4  0  -  6  9  % 

C    70%or  Greater 


Figure  5.     An  elaborate  correlation  model  (from  Verner  and  Boss  1980). 
770  Habitat  Evaluation  Systems 


r8 

r| 

rO 

f8 

y1 

.to 

■  o 

-U1 

r-~ 

i^ 

l^~ 

<* 

<>e 

<* 

o 

LO 

in 

-CD 

.  o 

c 

c\j 

c 

CM 

»  -< 

in 

'  CM 

D         00 

ID 

•<t 

C\J 

3         00 

to 

■<t 

CM 

O          CO         CD 

1 

CJ 

O        CO       to 

Tt        CM 

O 

O 

O 

O 

O 

o 

O 

O 

^        ci        C) 

d 

o 

--       o      d 

d      d 

xapu|  Aim 

qejjr 

S 

X3pu|  Aii|iqeii 

IS 

xepu|  Aimqe^n 

S 

X9pu| 

Allliqe^ns 

TJ 

B 

tO 

^-^ 

u 

to 

XI 

5 

01 

>> 

U-l 

5 

M 

a.  *u 

o 

0 

rn 

3 

o  o 

rl 

0)    r-l 

^^ 

(0 

a 

01 

60 

U    H 

o 

CO  T3 

60   60 

10     tO 

0 

r-4 

u    0> 

to  a 

4-1   t-l     60 

T> 

<4-l    IH 

o 

o 

CO 

r-l 

s  g 

J»tH 

CO   r-l     C 

o 

VO 

>, 

t-l    rl 

cu 

T3    t-I 

to    O 

• 

(X 

o  a 

u 

H    01  H 

rl 

T> 

r** 

O 

c 

U    0 
CO    o 

2 

1    tl    pi 

a  to  to 

o 

•o 

AI 

rl 

OI 

tO 

rl     O 

a 

O     1     to 

01 

3  XI 

4-1 

o 

a) 

w 

•rl    XI 

60     rl 

O 

(0 

CU 

01 

>    01 

o   u 

m 

to   3   o> 

CO     El    r-t 

a  3 

3     4J 

U    T) 
60    01 

•H 

g 

01 

3 

o 

1)  £    O 

riH 

rl 

x: 

•H 

M 
4J 

M-l     CO 

o   o 

u 

una 
u 

cm    a 

O      r 

5 

TJ 

l-H 

•H 

3    ^-N^ 

rr,  /-v 

o  x; 

5 

CU 

iM! 

&•«  y 

M-l 

co  <:  « 

O   Q 

fr<    y 

3 

r-l 
XI 

*~\ 

r"N 

/^ 

>^v 

nj 

N 

^w 

>* 

•H 

t>" 

> 

> 

h 

•—• 

N-> 

N-^ 

Ste*' 

CO 

> 

a 

CU 

•^ 

^ 

^ 

4J 
rl 
CO 

Q 

0 

,Si 

4J 

CO 

0 

V 

^ 

4-1 

a 

■H 

to 

a 

a 

OI 

o> 

3      • 

14 

4J 

r-l     9 

o 

(0 

CO    o 

>    r-l 

*2 

01 

<u 
a 

OI 

TS     CO 

> 
o 

0 

•h 

r-l 

a  o> 

y 

0- 

i 

CO 
r-l 

1-4 

CO 
<4-l 

S 

rl 
01 

u 

■H    3 

r-l 
(►.    CO 
4J     > 

OI 

3 

01 
4J 

t-i 

-o 

CU 

T-l 

r-l 

a 

4) 

CO 

O 

0 

r-l     CU 

CO 

•H 

"8 

•H 

u 

•o 

CO 

•H 

■H     10 
J3      01 

> 

a 

S3 

5 

>. 

•o 

CO  J3 

M 

)-■ 

0< 

X> 

4J     4J 

a* 

o 

^•\ 

y,    o> 

a 

■H 

•a 

<4-l 

M 

-a- 

3    o 

TS 

TJ 

•H 

3      CD 

a 

CO 

S2 

a 

SI 

CU 

CO     4J 

M 

x 

II 

tn 

t>0 

CO 

t-l 

U 

^-\ 

■H 

OI 

a 

0) 

r-l      P. 

IJ 

a 

01 

• 

H  15 

fr. 

T3 

* 

u 

u   m 

en 

o 

> 

a 

4J 

(s.. 

(3                CM 

M 

3 

u 

o 

•H 

4-1 

•H 

Q» 

t^ 

Q 

•H                  ~^. 

01 

o 

(0 

>,  u 

M-l 

60 

y 

TJ 

I-l 

^0 

r-l 

T3 

r-4 

o 

CL    O 

o   eo 

co 

*J     C 

T-l 

^ 

, 

>s 

r-N 

a 

r-4 

iH 

o 

a 

TI 

o> 

^^* 

J=     CO 

X) 

% 

• 

"N 

4-1 

*4f 

M 

O 

y 

/■>        c    u 

/~\ 

4)    T-l    00 

i— 1 

y 

00 

<0 

v. 

•H 

> 

<4-t 

O             CO    T-l 

o 

eor-i  c 

O 

co 

0 

■n    r-~ 

4J 

r-l 

S-. 

>»0             CJ    CM 

o 

10  t5  t-i 

CM 

y 

r4     > 

•H 

■H 

X 

4-1 

0> 

CXi-l 

r-l 

U     0)    r-l 

U 

u 

r"N 

3 

XI 

•H 

-C 

o 

A|      |"S 

CO    01    o. 

o 

3 

\D  O 

4J   X! 

CO 

CO 

c*> 

I-l 

4J 

a 

A| 

CO    CO 

CO 

•  o 

CO     60 

4-> 

> 

•H 

CO 

r-l      |       CO 

01 

1^   r-l 

3 

•H       • 

JO 

rl 

CJ 

4-1   TJ 

CO    Xl      1 

60 

tl 

■o 

to   o 

3    co 

X 

to 

O 

i-l          I)     tl 

CM 

O    3    Oi 

a 

3 

a 

42  a| 

r=      r4 

CO   i-4 

u 

<4-< 

« 

>               rl      CO 

> 

O     U    1-4 

3 

l-l 

3 

O.  J3 

4) 

CM 

•H 

01 

01     "H 

■H    J3    O 

O 

3 

o 

-* 

CO     4J 

4-1   T3 

> 

3 

CO 

u 

>      rl 

CO     CO     CL 

:►. 

rl 

-S  t>, 

u 

CO    O 

CO 

01 

u 

A        °   g- 

aI 

co 

60 

4iA 

60    r-4 

4J     Q 

>< 

3 

1               0 

01     ^>  /-s 

^-s 

> 

•rl 

u 

r-l 

<M 

O          <4-l     o 

o 

y  <  co 

u 

O 

M-l 

^ 

0 

X>      60 

r-4 

CO 

(0 

o 

^         O    CJ 

v^ 

y 

o 

o  o 

co  a 

> 

4J 

> 

M 

M 

3 
CO 

K 

l-l     4-1 
14-1 

^S 

s^ 

J3 
CO 

S 

^ 

/-^ 

^v 

^^ 

CU    c 

CO    O 
i— 1 

n 

n 

o 

r-l 

CM 

CI 

«* 

c    o 

01    i-H 

M 

T) 

> 

> 

> 

> 

•H      O. 

1-4          CM 

ro        <f 

4J      0 

to 

0> 

> — 

V"' 

v-^ 

s-^ 

0    CO 
Li     01 

>          > 

>         > 

co  u-i 

r-l 

OS 

4J 

01    U 

3    01 

T-l 

4J    U 

y  -c 

W 

01     O 

rH     4J 

* 

Q   y 

CO 

O    60 

*-\ 

1-V 

a 

^    TH 

rH 

CM 

n   co 

>-• 

^  3 

X 

c 

V 


■a 

tC 

o 
S 

C 


u 

c 

'a, 
t- 
c2 

"v 
■o 
o 
E 

53 
X 
c 
< 

U 

3 
6C 


Habitat  Evaluation  Systems 


771 


Habitat  Capability  Models 

Habitat  capability  models,  as  used  here,  are  simi- 
lar to  HSI  models,  except  the  habitat  variables  are 
used  to  predict  an  animal  density.  Thus,  these 
models  could  include  the  traditional  carrying-capac- 
ity models  used  for  many  ungulates.  Clearly,  some 
PATREC  models  fall  into  this  category  when  the 
population  variable  is  some  measure  of  population 
density.  The  U.S.  Forest  Service  uses  the  term  habitat 
capability  models  to  include  PATREC  models  and 
HSI  models;  however,  the  U.S.  Forest  Service  con- 
verts the  models  so  the  output  is  some  measure  of 
population  density. 

Matrix  Models 

The  simplest  form  of  multi-species  models  is  the 
species-habitat  matrix,  in  which  the  presence  of  each 
species  in  a  cover  type  is  indicated  by  an  X  or  other 
similar  notation.  Many  variations  on  this  approach 
exist.  A  relative  density  (e.g.,  H  =  high,  M  =  me- 
dium, L  =  low)  or  abundance  (e.g.,  A  =  abundant, 
C  =  common,  U  =  unusual,  R  =  rare,  etc. )  may  be 
indicated.  Similarly,  the  season  of  use  (e.g.,  W  = 
winter,  Sp  =  Spring,  etc.)  or  breeding  status  (e.g.,  B 
=  breeds  in  habitat,  N  =  does  not  breed  in  habitat ) 
may  be  indicated.  Such  information  is  commonly 
provided  by  computer-based  systems  such  as  RUN- 
WILD  (Patton  1979)  and  "Procedures"  (U.S.  Depart- 
ment of  the  Interior,  Fish  and  Wildlife  Service 
1980a). 

Such  models  have  commonly  been  used  in  doc- 
uments such  as  EISs  to  describe  current  wildlife 
resources  that  may  be  affected  by  a  land-use  deci- 
sion; however,  they  are  not  particularly  useful  for 
predicting  which  species  will  actually  be  affected  or 
how  it  will  be  affected.  Such  prediction  requires  that 
species  presence  or  density  be  related  to  compo- 
nents of  the  habitat,  such  as  vegetation  structure  or 
physical  features.  Several  systems,  including  life- 
forms  and  guilding,  do  this  to  some  extent. 

The  Life-Form  Approach 

The  life-form  approach  was  developed  in  Ore- 
gon by  Thomas  ( 1979),  based  on  earlier  work  by 
Haapanen  (1965)  on  birds  in  Finnish  forests.  Thomas 
and  others  expanded  the  concept  to  include  all  ter- 
restrial vertebrates.  A  life-form  is  defined  as  "a  group 
of  wildlife  species  whose  requirements  for  habitat 
are  satisfied  by  similar  successional  stages  within 
given  plant  communities"  (Thomas  1979:-i<S0) 

For  the  Blue  Mountains  of  Oregon,  Thomas 
(1979)  aggregated  over  300  vertebrate  species  into 
16  life-forms  (Figure  7).  These  life-forms  were  then 
correlated  with  successional  stages  of  each  plant 
community.  Thus,  if  a  biologist  can  predict  the  effect 


of  management  on  succession,  the  life-forms  and 
species  that  will  be  present  can  also  be  predicted. 

The  life-form  approach  was  developed  for  forest 
lands  in  response  to  timber  management  programs, 
intended  to  provide  a  forest  manager  with  informa- 
tion on  impacts  to  wildlife  from  management  prac- 
tices. A  major  impact  of  logging  is  to  set  back  forest 
succession,  and  succession  in  forest  lands  is  rela- 
tively predictable.  This  approach  is  more  suited  to 
forest  lands  than  rangelands,  where  livestock  grazing 
is  often  the  principal  commercial  use  and  plant 
succession  is  less  predictable.  Thomas  (1979)  also 
recognized  the  need  to  supplement  the  life-form 
approach  with  consideration  of  special  habitats  (e.g., 
riparian  zones);  unique  habitats  (e.g.,  cliffs,  caves); 
and  separate  models  for  featured  species  (e.g.,  mule 
deer  and  elk). 

Guilding  Approach 

Use  of  a  community  guild  model  or  "guilding"  is 
similar  to  the  life-form  approach.  A  guild  was  origi- 
nally defined  as  a  "group  of  species  that  exploit  the 
same  class  of  environmental  resources  in  a  similar 
way"  (Root  1967).  Application  of  the  guild  concept 
appears  to  have  many  potential  uses  in  wildlife  man- 
agement; however,  the  way  guilds  are  defined  and 
used  varies  greatly  (Verner  1984).  Short  (1983,  1984) 
and  Short  and  Burnham  (1982)  have  developed  for- 
mal procedures  for  applying  the  guild  concept  to 
wildlife  habitat  evaluation  and  have  termed  it  the 
"community  guild  model."  Development  of  the  tech- 
nique was  supported  in  part  by  the  U.S.  Bureau  of 
Land  Management.  The  most  recent  version  of  the 
model  is  described  here. 

Guilds  are  defined  according  to  the  habitat  lay- 
ers a  species  uses  for  breeding  and  feeding.  Species 
are  then  assigned  to  appropriate  cells  based  on  exist- 
ing data  and  literature.  For  example,  a  species,  such 
as  the  black-headed  grosbeak  that  breeds  in  shrubs 
and  feeds  in  trees,  shrubs,  and  terrestrial  (ground) 
surface  (Short  1983:186),  would  be  assigned  to  rows 
8,  6,  and  5  of  column  7  in  Figure  8.  ( In  a  life-form 
analysis,  this  bird  would  be  assigned  to  life-form  1 1 ; 
Thomas  1979:323) 

Species  occupying  the  same  combination  of 
cells  are  then  lumped  into  guilds  much  as  done  with 
life-forms.  Using  such  a  model,  a  biologist  can  pre- 
dict the  guilds  affecting  various  layers  of  habitat  that 
will  be  lost  or  adversely  affected  by  management 
actions. 

Habitat  Evaluation  Procedures  (HEP) 

The  HEP  system  was  formally  developed  and 
used  by  the  U.S.  Fish  and  Wildlife  Service  and  some 
other  federal  agencies,  particularly  the  U.S.  Bureau  of 
Reclamation  and  the  U.S.  Army  Corps  of  Engineers. 


772 


Habitat  Evaluation  Systems 


M      /       /              0       i 

A*/          ■?            /           <?           A/            ■? 

/ \M>  /                                      Q                                          /                                        <Z>                                        / /                                            <& 

1 

in  water 

in  water 

1 

bullfrog 

2 

in  water 

on  the  ground,  in  bushes, 
and/or  in  trees 

9 

long-toed  salamander,  western 
toad,  Pacific  treefrog 

3 

on  the  ground  around  water 

on  the  ground,  and  in  bushes, 
trees,  and  water 

45 

common  garter  snake,  killdeer, 
western  jumping  mouse 

4 

in  cliffs,  caves,  rimrock,  and/or 
talus 

on  the  ground  or  in  the  air 

32 

side-blotched  lizard,  common  raven, 
pika 

5 

on  the  ground  without  specific 
water,  cliff,  rimrock,  or  talus 
association 

on  the  ground 

48 

western  fence  lizard,  dark-eyed 
junco,  elk 

6 

on  the  ground 

in  bushes,  trees,  or  the  air 

7 

common  nighthawk,  Lincoln's 
sparrow,  porcupine 

7 

in  bushes 

on  the  ground,  in  water,  or  the 
air 

30 

American  robin,  Swainson's  thrush, 
chipping  sparrow 

8 

in  bushes 

in  trees,  bushes,  or  the  air 

6 

dusky  flycatcher,  yellow-breasted 
chat,  American  goldfinch 

9 

primarily  in  deciduous  trees 

in  trees,  bushes,  or  the  air 

4 

cedar  waxwing,  northern  oriole, 
house  finch 

10 

primarily  in  conifers 

in  trees,  bushes,  or  the  air 

14 

golden-crowned  kinglet,  yellow- 
rumped  warbler,  red  squirrel 

11 

in  conifers  or  deciduous  trees 

in  trees,  in  bushes,  on  the 
ground,  or  in  the  air 

24 

goshawk,  evening  grosbeak,  hoary 
bat 

12 

on  very  thick  branches 

on  the  ground  or  in  water 

7 

great  blue  heron,  red-tailed  hawk, 
great  horned  owl 

13 

in  own  hole  excavated  in  tree 

in  trees,  in  bushes,  on  the 
ground,  or  in  the  air 

13 

common  flicker,  pileated 
woodpecker,  red-breasted  nuthatch 

14 

in  a  hole  made  by  another 
species  or  in  a  natural  hole 

on  the  ground,  in  water,  or  the 
air 

37 

wood  duck,  American  kestrel, 
northern  flying  squirrel 

15 

in  a  burrow  underground 

on  the  ground  or  under  it 

40 

rubber  boa,  burrowing  owl, 
Columbian  ground  squirrel 

16 

in  a  burrow  underground 

in  the  air  or  in  the  water 

10 

bank  swallow,  muskrat,  river  otter 

Total: 

327 

'Species  assignment  to  life  form  is  based  on  predominant  habitat-use  patterns. 


Figure  7.     Life-forms  for  the  Blue  Mountains  of  Oregon  (from  Thomas  et  al.  1979). 


Habitat  Evaluation  Systems 


773 


o 
o 


D5 

C 

01 


10.  Feeds  Elsewhere 

9.  Air 

8.  Tree  Canopy 

• 

7.    Tree  Bole 

6.  Shrub  Layer 

• 

5.  Terrestrial 
Surface 

• 

4.  Terrestrial 
Subsurface 

3.  Water  Surface 

2.  Water  Column 

1.   Bottom  Water  Column 

CO 

CD 
O 

ZS 

o 

CO 

B 

CO 

b 

CL 

E 

CD 

I— 

c 
E 

o 
O 

§ 

"o 

E 
o 

b 

CD 

c 
E 

o 
O 

CD 

CD 

o 

m 

ZJ 
CO 

b 

CD 

o 

ZJ 
CO 

ID 
Z! 

CO 

to 

CD 

b 

I— 

CD 

o 

ZJ 
CO 

TO 
CD 

b 

I— 

b 
>- 

CO 

_l 

JD 

ZJ 

_c 

CO 

CD 
O 

CD 

CD 
CD 

h1 

CL 

o 

c 

CCS 

o 

CD 
CD 

H 

CD 

b 

JZ 

g 

CD 
CO 

UJ 
CO 

"O 
CD 
CD 

CO 

1. 

2. 

3. 

4. 

5 

6 

7 

8. 

9. 

10. 

Breeding  Loci 


Figure  8.     Habitat  layers  used  to  define  guilds  for  use  in  the  community  guild  model  (from  Short  1983). 


774 


Habitat  Evaluation  Systems 


The  basic  building  blocks  of  the  system  are  the  HSI 
models  previously  described.  One  or  more  HSI 
models  are  used  to  represent  components  of  habitat 
potentially  affected  by  a  project.  The  index  values 
are  multiplied  by  the  acres  of  habitat  to  produce 
"habitat  units"  present  before  and  after  the  project. 
These  habitat  units  are  then  used  to  determine  equi- 
table mitigation.  An  increase  in  habitat  units  from 
mitigation  is  calculated  in  the  same  way  to  produce 
the  number  of  units  before  and  after  a  mitigation 
action. 

HEP  is  used  mostly  by  the  U.S.  Fish  and  Wildlife 
Service  and  development  agencies  to  analyze  water 
projects  that  have  legal  requirements  for  equitable 
mitigation.  It  is  used  much  less  on  projects  that  do 
not  have  these  requirements. 

An  important  feature  and  one  of  the  major 
strengths  of  the  HEP  system  is  the  emphasis  on  get- 
ting all  interested  parties  together  early  in  the  evalu- 
ation to  agree  on  the  species  and  precise  HSI  models 
to  be  used  in  the  analysis.  This  helps  prevent  later 
disagreement  regarding  the  results  of  the  evaluation. 

In  practice,  HEP  is  used  with  very  little  follow- 
up  monitoring. 

The  HEP  system  is  thoroughly  described  by  the 
U.S.  Department  of  the  Interior,  Fish  and  Wildlife 
Service  (  1980b)  with  various  applications  discussed 
in  Schamberger  and  Farmer  ( 1978)  and  Urich  and 
Graham  (1983). 

Wildlife  and  Fish  Habitat  Relationships 
(WFHR)  System 

The  U.S.  Forest  Service  has  developed  a  system 
of  habitat  evaluation  called  the  Wildlife  and  Fish 
Habitat  Relationships  (WFHR)  system.  This  is  not  a 
formal  system  like  HEP,  but  rather  a  set  of  standards 
and  collection  of  evaluation  tools  to  be  used  by 
agency  biologists  (Nelson  and  Salwasser  1982).  Biol- 
ogists may  do  evaluations  by  using  the  life-form  ap- 
proach or  HSI,  PATREC,  or  habitat  capability  models 
as  appropriate  (Sheppard  et  al.  1982).  In  addition,  on 
some  projects,  the  U.S.  Forest  Service  makes  use  of 
dynamic  simulation  models  such  as  DYNAST  to  pre- 
dict forest  succession  after  timber  harvesting  or 
other  disturbance.  Static  models  such  as  HSI  models 
are  then  used  to  predict  the  wildlife  species  or  abun- 
dance associated  with  future  successional  stages 
(Barrett  and  Salwasser  1982).  The  U.S.  Forest  Service 
also  emphasizes  and,  thus,  has  developed  methodol- 
ogy to  analyze  cumulative  impacts  (Salwasser  et  al. 
1983)  and  viable  populations  (Salwasser  and  Samson 
1985).  Methodology  for  such  analyses  is  still  primi- 
tive but  an  important  subject  that  will  likely  be  used 
more  and  more  in  the  future. 


SUMMARY  AND  CONCLUSIONS 


The  field  of  habitat  evaluation  has  advanced 
rapidly  in  the  past  1 5  years.  The  concept  of  a  habitat 
model  as  a  basis  for  habitat  evaluation  has  been  de- 
veloped and  used  in  a  variety  of  ways.  Although  such 
models  provide  prototypes  for  biologists,  the  biolo- 
gist will  need  to  specifically  develop  models  for  the 
area  and  project  under  consideration.  Such  model 
building  is  basically  nothing  more  than  a  more  for- 
mal and  quantifiable  approach  to  the  conceptual  and 
qualitative  models  that  biologists  have  used  in  the 
past. 

Habitat  models  are  clearly  an  integral  part  of  the 
inventory  and  monitoring  process,  although  they 
may  not  be  labeled  as  such.  Although  monitoring  can 
be  done  without  any  sort  of  an  explicit  model,  such 
monitoring  is  not  likely  to  increase  the  understand- 
ing of  wildlife  resources  being  managed,  nor  lead 
in  the  long  run  to  better  management.  Human  dis- 
turbance factors  are  incorporated  into  very  few 
models,  yet  they  are  among  the  most  important  im- 
pacts to  wildlife.  Future  habitat  models  will  need 
to  be  developed  to  incorporate  these  factors. 

The  U.S.  Fish  and  Wildlife  Service  HEP  proce- 
dures, by  emphasizing  the  process  of  gaining  concen- 
sus on  models  to  be  used  in  an  analysis,  have 
demonstrated  a  useful  way  to  alleviate  confrontation 
in  the  habitat  evaluation  process.  Disagreements 
focus  on  analysis  tools  that  are  easier  to  evaluate  ob- 
jectively, rather  than  on  analysis  results.  This  ap- 
proach promises  to  be  useful  in  many  situations, 
regardless  of  whether  the  biologist  is  using  other  for- 
mal aspects  of  HEP. 

The  U.S.  Forest  Service  has  developed  and  ap- 
plied many  new  approaches.  The  linking  of  dynamic 
simulation  models  of  succession  with  static  habitat 
models,  which  the  U.S.  Forest  Service  pioneered, 
should  be  useful  wherever  succession  is  reasonably 
predictable.  Similarly,  the  efforts  of  U.S.  Forest  Ser- 
vice personnel  and  collaborators  to  develop  and  use 
cumulative  effects  models  and  viable  population 
analysis  models  are  pioneering  efforts  that  are  both 
necessary  and  promising.  Future  habitat  modeling 
will  likely  emphasize  these  two  areas. 

As  with  other  areas  of  inventory  and  monitoring, 
use  of  habitat  models  and  evaluation  systems  re- 
quires the  biologist  to  use  sound,  professional  judg- 
ment and  to  keep  aware  of  current  developments. 
The  techniques  will  probably  continue  to  develop 
rapidly,  and  the  biologist  cannot  afford  to  be  content 
with  "cookbook"  procedures  or  outdated 
technology. 


Habitat  Evaluation  Systems 


775 


LITERATURE  CITED 


ALLEN,  AW.  1982.  Habitat  suitability  index  models:  Mar- 
ten. U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  FWS/OBS- 
82-10.11.  9pp. 
BAILEY,  J.A.  1982.  Implications  of  "muddling  through"  for 
wildlife  management.  Wildl.  Soc.  Bull.  10:363-369. 

.  1984.  Principles  of  wildlife  management.  John 

Wiley  and  Sons,  New  York,  NY.  373pp 
BARRETT,  R.H.  and  H.  SALWASSER.  1982.  Adaptive  man- 
agement of  timber  and  wildlife  habitat  using  dynast 
and  wildlife-habitat  relationship  models.  Paper  pre- 
sented at  Western  Assoc,  of  Fish  and  Wildlife  Agen- 
cies, Las  Vegas,  NV,  July  21,  1982.  16pp.  (mimeo) 
BINNS,  N.A.  1979.  A  habitat  quality  index  for  Wyoming 
trout  streams.  Wyoming  Game  and  Fish  Dep.,  Fish. 
Res.  Rep.  2.  75pp. 
COLE,  C.A.  and  R.L.  SMITH.  1983.  Habitat  suitability  in- 
dices for  monitoring  wildlife  populations.  Trans. 
North  Am.  Wildl.  and  Nat.  Resour.  Conf.  48:367-375. 
COOPERRIDER,  AY.  and  J.A.  BAILEY.  1984.  A  simulation 
approach  to  forage  allocation.  Pages  525-559  in  De- 
veloping Strategies  for  Rangeland  Management.  Rep. 
prepared  by  the  Committee  on  Developing  Strategies 
for  Rangeland  Manage.,  Natl.  Res.  Council/Natl.  Acad. 
Sciences,  Westview  Press,  Boulder,  CO.  2022pp. 
and  D.F.  BEHREND.  1980.  Simulation  of  the  inter- 
action of  deer  with  northern  forest  vegetation.  New 
York  Fish  and  Game  J.  27:142-155. 
HAAPANEN,  A.  1965.  Bird  fauna  of  the  Finnish  forests  in 
relation  to  forest  succession.  I.  Ann.  Zool.  Fenn. 
3(3):153-196. 
HANSEN,  C.G.  1980.  Habitat.  Pages  64-79  in  Monson,  G. 
and  L.  Summer,  eds.  The  Desert  Bighorn.  Univ.  Ari- 
zona Press,  Tucson. 
HOLLLNG,  C.S.,  ed.  1978.  Adaptive  environmental  assess- 
ment and  management.  Vol.  3,  International  Series  on 
Applied  Systems  Analysis.  John  Wiley  and  Sons,  New 
York,  NY.  377pp. 
NELSON,  R.D.  and  H.  SALWASSER.  1982.  The  Forest  Ser- 
vice wildlife  and  fish  habitat  relationships  program. 
Trans.  North  Am.  Wildl.  and  Nat.  Resour.  Conf. 
47:174-183. 
ODUM,  E.P.  1969.  Fundamentals  of  ecology.  W.B.  Saunders 

Company,  Philadelphia,  PA.  546pp. 
O'NEIL,  L.J.  and  M.L.  SCHAMBERGER.  1983.  Habitat 

models  as  a  monitoring  tool.  Pages  424-427  in  Bell, 
J.F.  and  T.  Atterbury,  eds.  Renewable  Resource  Inven- 
tories for  Monitoring  Changes  and  Trends.  SAF83- 
14,  Oregon  State  Univ.,  Corvallis. 
PATTON,  DR.  1979.  RUNWILD  II:  A  storage  and  retrieval 
system  for  wildlife  data.  Trans.  North  Am.  Wildl.  and 
Nat.  Resour.  Conf.  44:425-430. 
PETERSON,  R.T.  1969.  A  field  guide  to  western  birds: 
Field  marks  of  all  species  found  in  North  America 
west  of  the  100th  Meridian,  with  a  section  on  the 
birds  of  the  Hawaiian  Islands.  2nd  ed.  (revised  and 
enlarged).  Houghton  Mifflin  Co.,  Boston,  MA.  336pp. 
ROOT,  R.B.  1967.  The  niche  exploration  pattern  of  the 

blue-gray  gnatcatcher.  Ecol.  Monogr.  37:317-350. 
SALWASSER,  II.,  C.K.  HAMILTON,  W.B.  KROIIN,  J.F.  LIP- 
SCOMB, and  C.H.  THOMAS.  198.3.  Monitoring  wildlife 


and  fish:  Mandates  and  their  implications.  Trans. 
North  Am.  Wildl.  and  Nat.  Resour.  Conf.  48:297-307. 

and  F.B.  SAMSON.  1985.  Cumulative  effects  analy- 
sis: An  advance  in  wildlife  planning  and  management. 
Trans.  North  Am.  Wildl.  and  Nat.  Resour.  Conf. 
50:313-321. 

SCHAMBERGER,  M.A.  and  AH.  FARMER.  1978.  The  habi- 
tat evaluation  procedures:  Their  application  in  project 
planning  and  impact  evaluation.  Trans.  North  Am. 
Wildl.  and  Nat.  Resour.  Conf.  43:274-283- 

, ,  and  J.W.  TERRELL.  1982.  Habitat  suitability 

index  models:  Introduction.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.  FWS/OBS-82- 10.  2pp. 

SCHROEDER,  R.L.  1984.  Habitat  suitability  index  models: 
Black  brant.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv. 
FWS/OBS-82/ 10.63.  11pp. 

SHEPPARD,  J.L.,  D.L.  WILLS,  andJ.L  SIMONSON.  1982. 
Project  applications  of  the  Forest  Service  Rocky 
Mountain  Region  wildlife  and  fish  habitat  relationships 
system.  Trans.  North  Am.  Wildl.  and  Nat.  Resour. 
Conf.  47:128-141. 

SHORT,  H.L.  1983.  Wildlife  guilds  in  Arizona  desert  habi- 
tats. U.S.  Dep.  Inter.,  Bur.  Land  Manage.,  Tech.  Note 
362.  258pp. 

.  1984.  Habitat  suitability  index  models:  The  Ari- 
zona guild  and  layers  of  habitat  models.  U.S.  Dep. 
Inter.,  Fish  and  Wildl.  Serv.  FWS/OBS-82/ 10.70.  37pp. 

and  K.P.  BURNHAM.  1982.  Techniques  for  structur- 
ing wildlife  guilds  to  evaluate  impacts  on  wildlife 
communities.  U.S.  Dep.  Inter.,  Spec.  Sci.  Rep. — Wildl. 
244.  34pp. 

SOUSA,  P.J.  1983  Habitat  suitability  index  models:  Lewis' 
woodpecker.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv. 
FWS/OBS-82/ 10.32.  14pp. 

THOMAS,  J.W.,  ed.  1979.  Wildlife  habitats  in  managed 

forests — the  Blue  Mountains  of  Oregon  and  Washing- 
ton. U.S.  Dep.  Agric,  For.  Serv.,  Agric.  Handbook 
553.  511pp. 

URICH,  D.L  and  J.P.  GRAHAM.  1983.  Applying  habitat 

evaluation  procedures  ( HEP )  to  wildlife  area  planning 
in  Missouri.  Wildl.  Soc.  Bull.  11:215-222. 

U.S.  DEPARTMENT  OF  THE  INTERIOR,  FISH  and  WILD 
LIFE  SERVICE.  1980a.  An  evaluation  of  "A  procedure 
for  describing  fish  and  wildlife  in  Pennsylvania."  U.S. 
Dep.  Inter.,  Fish  and  Wildl.  Serv.  FWS/OBS-79/19-A. 
15pp. 

.  1980b.  Habitat  evaluation  procedures  (HEPs)  102 

ESM.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv.  (unnum- 
bered) 

.  1981. 


Standards  for  the  development  of  habitat 
suitability  index  models.  103  ESM.  U.S.  Dep.  Inter., 
Fish  and  Wildl.  Serv.  (unnumbered) 

VERNER,  J.  1984.  The  guild  concept  applied  to  manage- 
ment of  bird  populations.  Environmental  Manage. 
8(1):1-13. 

and  A.S.  BOSS,  eds.  1980.  California  wildlife  and 

their  habitats:  Western  Sierra  Nevada.  U.S.  Dep.  Agric, 
For.  Serv.,  Gen.  Tech.  Rep.  PSW-37.  439pp. 

WILLIAMS,  G.L.,  K.R.  RUSSELL,  and  W  K.  SEITZ.  1978. 

Pattern  recognition  as  a  tool  in  the  ecological  analysis 
of  habitat.  Pages  521-531  in  Marmelstein,  A.,  ed. 
Classification,  Inventory'  and  Analysis  of  Fish  and 
Wildlife  Habitat.  U.S.  Dep.  Inter.,  Fish  and  Wildl.  Serv. 
FWS/OBS-78/76.  604pp. 


776 


Habitat  Evaluation  Systems 


39 

EVALUATION  AND 
INTERPRETATION 


Allen  Y.  Cooperrider 

U.S.  Bureau  of  Land  Management 
Service  Center 
Denver,  CO  80225 


"Statistical  analysis  is  not  a  substitute  for  thinking." 
— Alan  Speigel,  Speigel's  Laws  of  Management 


Editor's  Note:  The  best  data  sets  and  most  elegant 
quantitative  analyses  will  not  likely  affect  manage- 
ment unless  they  are  interpreted  and  evaluated  for 
those  who  make  the  decisions.  This  process  requires 
the  biologist  to  creatively  bring  together  informa- 
tion from  various  sources.  These  sources  include 
not  only  data  collection  and  statistical  analyses  of 
them,  but  also  field  observations,  experience  from 
other  areas  or  projects,  and  opinions  from  others 
knowledgeable  about  the  subject  or  area.  Few  biol- 
ogists are  effective  in  influencing  decisions,  unless 
they  have  mastered  the  art  of  evaluation  and  inter- 
pretation. This  chapter  discusses  the  process  and 
some  factors  a  biologist  should  consider  at  this 
stage  of  an  inventory  or  monitoring  effort. 


INTRODUCTION 

Interpretation  and  evaluation  of  data  are  critical 
steps  in  the  inventory  and  monitoring  process.  At 
this  stage,  the  biologist  must  explain  what  all  the 
data,  information,  and  analyses  mean.  This  is  not  a 
formal  process;  if  it  were,  it  would  be  included  with 
other  analytical  methods.  Since  the  process  is  essen- 
tially subjective,  very  little  has  been  written  about  it. 
Why,  then,  include  a  chapter  on  such  a  nebulous 
subject?  Because  biologists  often  do  a  poor  job  of 
interpretation  and  evaluation  or  tend  to  skip  this 
step. 


ciAk.., 


Evaluation  and  Interpretation  777 


Biologists  frequently  present  a  manager  with 
lists  of  species  present  in  habitats,  correlation  coeffi- 
cients, t-tests,  or  other  test  statistics  and  expect  a 
manager  to  understand  how  these  relate  to  the  land 
management  decisions  to  be  made.  The  manager 
does  not  understand;  therefore,  the  biologists  are  not 
performing  their  jobs  as  thoroughly  and  profession- 
ally as  they  could.  A  biologist  is  tasked  with  explain- 
ing the  meaning  of  data;  suggesting  implications  from 
the  findings;  cautioning  about  known  or  potential 
weakness  in  the  data;  and  explaining  the  significance 
of  the  anticipated  resource  impacts.  The  biologist 
who  fails  to  do  this  has  not  completed  the  job  and  is 
not  fully  using  the  information.  Collecting  and  ana- 
lyzing data  are  frequently  expensive,  and  the  biolo- 
gist should  not  waste  information  derived  from  such 
data. 

The  art  of  interpretation  and  evaluation  cannot 
be  taught,  any  more  than  one  can  be  taught  to  think 
or  to  use  good  judgment.  The  aims  of  this  chapter 
are  more  modest.  In  it,  I  explain  what  interpretation 
and  evaluation  are  and  list  some  of  the  many  possi- 
ble aspects  that  must  be  considered  during  the  pro- 
cess. Biologists  hoping  to  improve  their  abilities  to 
interpret  and  evaluate  may  find  some  useful  guid- 
ance. However,  the  process  cannot  be  reduced  to  a 
cookbook  procedure. 

Interpretation  and  evaluation  is  thus  one  of  the 
most  important  and  challenging  tasks  of  the  profes- 
sional, yet  one  which  is  frequently  done  poorly  and 
with  very  little  guidance.  The  reasons  biologists 
struggle  with  these  tasks  usually  relate  to  their  train- 
ing and  background.  Many  wildlife  biologists  are 
thoroughly  trained  in  the  scientific  methods  and 
have  experience  in  research.  This  scientific  training 
is  quite  valuable  and  important.  However,  it  tends  to 
make  biologists  cautious.  Scientists  are  taught  to — 


( 1 )    Collect  data  carefully  and  systematically, 

(  2  )    Concentrate  on  what  the  data  demonstrates 
conclusively, 


(3)  Avoid  predicting  trends,  and 

(4)  Leave  out  value  considerations. 


A  management  biologist  should  also  collect  data 
carefully  and  systematically.  However,  the  manage- 
ment biologist  must  operate  quite  differently  with 
the  other  three  factors  since  these  are  opposite  from 
what  a  biologist  is  asked  to  do  in  interpretation  and 
evaluation.  Typically  a  manager  wants  predictions 
about  the  future.  Rarely  does  a  biologist  have 
enough  data  to  predict  in  a  scientific  way  what  will 
happen  on  an  area;  instead  he  or  she  has  many  di- 
verse sources  of  data  and  information  upon  which  to 


base  some  future  states.  This  is  needed  and  wanted 
by  management  to  make  an  informed  decision.  A 
scientist  can  afford  to  wait  for  conclusive  results;  a 
resource  manager  cannot. 

Similarly,  biological  scientists  do  not  deal  much 
with  values  other  than  "scientific  value"  and  that 
elusive  "ecological  value;"  most  biologists  are  taught 
that  consideration  of  values  is  the  domain  of  philoso- 
phy, economics,  sociology,  psychology,  or  religion. 
Yet,  a  management  biologist  must  constantly  deal 
with  values.  Biologists  are  paid  to  manage  wildlife 
and  wildlife  habitat  because  people  place  some  value 
on  those  resources.  But,  those  values  are  quite  di- 
verse. A  professional  biologist  must  be  aware  of 
these  values  and  explain  them  to  the  manager. 

Ideally,  a  scientific  manager  would  have  "hard" 
data  and  conclusive  evidence  on  which  to  base  all  of 
his  or  her  decisions.  In  real  life,  in  resource  manage- 
ment as  well  as  in  agriculture,  business,  economics, 
politics,  and  other  fields,  decisions  are  made  on 
"soft"  data.  A  decisionmaker  may  have  good  data  on 
one  decision  factor  and  only  an  opinion  or  limited 
data  on  other  factors.  A  manager  would  be  foolish  to 
make  decisions  based  only  on  factors  for  which  good 
data  were  available,  if  he  or  she  knew  other  factors 
were  also  important. 


INTERPRETATION 

To  interpret  has  been  defined  as  "to  explain  or 
tell  the  meaning  of:  present  in  understandable  terms" 
(Merriam- Webster  1983)-  Once  data  are  collected, 
organized,  put  on  the  computer,  processed  and  mas- 
saged in  a  variety  of  ways,  and  the  results  printed, 
the  information  needs  to  be  explained.  Casual  obser- 
vations while  out  in  the  field  and  talks  with  other 
people  in  the  field;  specialists  in  other  disciplines; 
and  other  biologists  familiar  with  the  area  or  the 
problem  being  addressed,  are  other  sources  of  infor- 
mation. The  experience  of  the  biologist  with  similar 
problems  or  areas  is  also  a  source  of  information. 
All  these  sources  can  be  used  to  explain  to  the  pub- 
lic, the  manager,  or  the  decisionmaker  what  the  data 
mean.  They  can  also  be  used  to  predict  a  future  with 
or  without  a  proposed  action.  Some  of  the  following 
questions  should  be  considered. 

What  do  the  data  mean?  A  biologist  may  have 
been  studying  the  behavior  and  foraging  habits  of  elk 
{Cervns  elaphus)  to  determine  if  expansion  of  a 
grazing  lease  into  its  summer  range  will  adversely 
affect  the  herd.  The  biologist  has  measured  elk  and 
cattle  use  in  a  series  of  adjacent  pastures.  After  ana- 
lyzing the  data,  he  or  she  may  have  concluded  with 
the  aid  of  a  statistician  that  "use  of  pastures  by  elk  is 
significantly  and  negatively  correlated  with  use  by 
cattle  (R"  =  0.99)."  More  obscurely,  "Spearman's 


778 


Evaluation  and  Interpretation 


rank  correlation  coefficient  between  elk  use  days  and 
cattle  use  days  is  -0.94."  These  statements  need  to 
be  translated  into  a  language  understood  by  the  nat- 
ural resource  manager,  i.e.,  "elk  are  avoiding  areas 
used  by  cattle."  (The  pure  logician,  blissfully  igno- 
rant of  western  cattle  operations,  could  of  course 
argue  that  cattle  are  avoiding  elk,  but  managers  are 
not  too  interested  in  this  type  of  argument. ) 

What  is  the  cause?  Biologists  and  managers 
must  be  concerned  with  causes.  Knowing  that  some- 
thing happens  is  useful,  but  knowing  why  is  even 
more  useful.  Considering  the  previous  example,  the 
manager  may  want  to  know  why  elk  avoid  areas 
used  by  cattle.  One  possibility  is  that  elk  simply  pre- 
fer different  types  of  pastures  than  those  normally 
used  by  livestock  operators  for  cattle  grazing.  How- 
ever, the  biologist  may  have  observed  that  elk  use 
pastures  until  cattle  move  onto  them,  after  which 
they  avoid  them.  This  is  useful  information  in  inter- 
preting data,  even  though  it  has  not  been  collected 
with  a  well-designed  study.  The  biologist  may  be- 
lieve from  his  or  her  observations  that  elk  avoid 
pastures  because  forage  conditions  are  much  worse 
after  cattle  have  been  "highgrading"  current  forage 
production.  Furthermore,  the  biologist  may  be  able 
to  cite  studies  in  similar  situations  that  indicate  such 
avoidance  is  due  to  forage  conditions.  Thus  informa- 
tion from  a  variety  of  sources  has  been  used  to  ex- 
plain the  observed  phenomena. 

A  common  problem  in  determining  causality  is 
that  the  results  observed  are  due  to  some  unex- 
pected but  obvious  cause,  other  than  the  one  a  biol- 
ogist was  trying  to  isolate.  Weather  is  often  such  a 
factor.  For  example,  a  biologist  may  have  been  moni- 
toring changes  in  vegetation  cover  in  two  similar 
streams:  one  grazed  by  cattle;  the  other  ungrazed. 
During  the  fifth  year,  a  100-year  storm  occurs  in  the 
drainage  of  the  ungrazed  stream,  causing  severe 
flooding  and  scouring  of  the  entire  streambed.  The 
data  after  5  years  would  suggest  that  grazing  im- 
proves the  condition  of  the  streamside  vegetation, 
whereas  just  the  opposite  is  true.  A  biologist  would 
be  foolish,  not  to  mention  unprofessional,  to  me- 
chanically interpret  the  data  without  recognizing  the 
more  obvious,  but  largely  undocumented,  causes. 
In  Chapter  35,  Weather  and  Climate,  ways  of  dealing 
with  variations  caused  by  weather  are  described. 
These  methods  will  help  in  most  cases,  but  excep- 
tions will  occur.  Record  events,  such  as  the  100-year 
storm,  the  record  low  frost,  the  wettest  spring  on 
record,  are  occurring  somewhere  in  the  world  every 
day. 

What  are  the  expected  impacts  of  the  pro- 
posed management?  This  is,  of  course,  the  reason 
data  have  been  collected  and  the  crucial  question  to 
be  answered.  Consider  the  previous  example  again. 
If  elk  avoid  areas  used  by  cattle  because  forage  con- 
ditions have  been  degraded,  what  are  the  expected 


impacts  of  expanding  grazing  leases?  Initially,  one 
may  expect  that  the  expansion  will  change  the  pat- 
tern of  elk  use  on  the  area.  This  may  be  a  concern  to 
either  livestock  or  wildlife  management.  Certain 
types  of  livestock  management,  such  as  rest  rotation 
systems,  may  be  doomed  to  fail  because  elk  will 
concentrate  in  the  pastures  that  are  supposed  to  be 
"rested."  Or,  such  management  may  force  elk  to 
move  to  other  areas  where  they  may  be  more  vul- 
nerable to  poaching  or  where  they  may  damage  agri- 
cultural crops. 

The  biologist  is  thus  interpreting  and  extrapolat- 
ing the  data  that  have  been  collected  to  make  infer- 
ences that  are  useful  in  making  management 
decisions.  In  following  this  process,  he  or  she  must 
also  be  evaluating  the  adequacy  of  the  information. 
This  leads  to  another  series  of  questions. 

How  good  are  the  data?  A  biologist  rarely  can 
collect  the  quantity  or  quality  of  data  required  for 
risk-free  conclusions.  For  one  thing,  budgets  for  such 
efforts  are  usually  rather  limited.  Furthermore,  well- 
designed  studies  for  evaluating  year-to-year  variation 
typically  take  3  to  5  years  of  data  collection;  yet, 
managers  cannot  wait  5  years  to  make  decisions. 
Thus,  the  best  study  designs  can  rarely  be  used,  even 
if  a  biologist  or  agency  could  afford  them. 

Even  the  data  collected  are  often  inadequate  or 
questionable.  Erroneous  data  are  collected  by  inex- 
perienced, trained,  or  unmotivated  technicians. 
Equipment  fails  or  malfunctions  occur  resulting  in 
data  gaps  or  data  of  questionable  accuracy.  Weather, 
travel  difficulties,  or  budget  problems  prevent  cer- 
tain data  from  being  collected.  Management  actions 
or  "acts  of  God"  destroy  or  invalidate  study  sites 
or  plots.  The  list  goes  on,  but  these  are  normal 
working  conditions  for  the  field  biologist  and  should 
not  be  confused  with  problems  caused  by  poor 
study  design.  The  latter  can  be  prevented;  the  for- 
mer cannot.  An  experienced  biologist  can  lessen 
these  problems  by  planning  for  contingencies,  train- 
ing personnel  thoroughly,  carefully  testing  equip- 
ment, and  taking  other  precautions.  The  problems 
can  never  entirely  be  prevented,  however. 

This  explains  why  most  data  are  "soft."  Hope- 
fully, the  data  obtained  are  better  than  having  no 
data  at  all.  However,  the  biologist  must  determine 
the  quality  of  the  data  and  convey  this  information 
to  others  who  use  or  make  decisions  based  on  such 
data.  Consider  the  previous  elk  and  cattle  example.  If 
the  technician  who  collected  the  habitat-use  data 
could  not  distinguish  elk  from  cattle  tracks,  one  may 
not  want  to  stake  his  or  her  reputation  on  the  study- 
results. 

Study  design  may  also  cause  problems.  Design- 
ing studies  requires  making  a  series  of  reasonable 
assumptions  so  certain  statistical  procedures  can  be 


Evaluation  and  Interpretation 


779 


used.  Typically,  assumptions  that  seemed  reasonable 
at  the  beginning  of  a  study  may  become  very  ques- 
tionable after  more  is  learned  about  the  habitat  or 
population  being  studied.  For  example,  one  may 
assume  that  a  drainage  contains  a  discrete  herd  of 
mule  deer  (Odocoileus  hemionns)  but,  after  a  field 
study,  discovers  two  separate  subherds:  one  that  is 
resident  year-round  and  the  other  migratory  and 
found  on  the  area  only  in  winter.  A  study  designed 
to  determine  the  amount  of  forage  to  be  allocated 
for  deer  might  produce  misleading  results  if  based 
on  the  former  assumption. 

Thus,  the  biologist  may  have  only  poor  data,  for 
reasons  beyond  his  or  her  control.  Nevertheless, 
the  manager  should  be  informed  of  the  reliability  of 
the  data.  Poor  data  can  cause  enough  problems,  and 
problems  should  not  be  compounded  by  pretending 
the  data  are  good. 

Are  the  data  adequate?  I  have  alluded  to  the 
fact  that  biologists  rarely  consider  data  adequate;  yet, 
most  of  the  time,  they  must  live  with  such  data. 
Sometimes,  however,  collecting  more  data  is  both 
necessary  and  feasible.  The  biologist  is  responsible 
for  advising  the  manager  of  such  contingencies.  As 
with  initial  planning  decisions,  such  mid-course  deci- 
sions must  be  made  by  weighing  the  time  and  costs 
involved  against  other  priority  projects. 

To  this  point,  I  have  talked  about  the  quality  of 
the  data,  the  meaning  of  the  data,  and  predicted 
impacts.  A  closely  related  question  is  "What  is  the 
significance  of  such  impacts?"  Put  another  way, 
"What  values  will  be  enhanced  or  destroyed  or  are 
at  risk?"  This  leads  to  the  subject  of  evaluation. 


EVALUATION 

Evaluation  is  the  process  of  determining  the 
value  of  something.  Used  here,  evaluation  is  the  pro- 
cess of  determining  the  value  of  the  wildlife 
resources. 

After  data  have  been  carefully  collected,  ana- 
lyzed, and  interpreted,  the  value  of  the  wildlife  re- 
sources that  may  be  affected  needs  to  be  explained. 
When  questions  such  as  "Who  cares  if  the  elk  herd 
decreases  by  50%  ?"  or  "What  is  300  acres  of  riparian 
habitat  worth?"  are  being  asked,  the  biologist  needs 
to  place  a  value  on  the  resource  or  at  least  explain 
the  resource  values. 

Wildlife  biologists  have  traditionally  concen- 
trated on  enhancing  or  ensuring  the  production  or 
conservation  of  wildlife  resources  in  the  best  manner 
possible.  The  criteria  they  use  are  primarily  biologi- 
cal or  ecological  and  include  factors  such  as  "ecolog- 
ical diversity"  or  "ecosystem  stability"  (Krutilla 


1974).  Most  wildlife  biologists  place  high  value  on 
wildlife  and  have  not  felt  a  need  to  seriously  exam- 
ine wildlife  values  (Bailey  1984).  Furthermore,  many 
biologists  dislike  the  concept  of  placing  monetary 
values  on  wildlife,  believing  the  real  values  of  wild- 
life are  aesthetic,  biological  and,  most  importantly, 
unmeasurable.  However,  if  biologists  argue  a  case  for 
wildlife  resources  against  actions  that  destroy  wild- 
life habitat,  they  must  be  thoroughly  familiar  with 
all  wildlife  values  and  some  valuation  methods 
(Steinhoff  1971). 


Economists  have  traditionally  placed  monetary 
values  on  economically  commercial  resources  (those 
resources  that  can  be  bought  and  sold)  so  the  re- 
sources can  be  analyzed.  Amenity  resources  (aes- 
thetic, ecological,  and  other  non-commercial 
resources),  however,  are  more  difficult  to  place  val- 
ues on  and,  therefore,  more  difficult  to  analyze  for- 
mally (Krutilla  1974).  Since  many  wildlife  values  fall 
in  the  amenity  category,  economic  analysis  of  wild- 
life values  is  still  in  its  infancy.  However,  the  field 
is  developing  rapidly  and  is  of  such  importance  that 
a  chapter  on  the  subject  is  included  in  this  book. 
Whereas  in  this  chapter  I  discuss  some  types  of  wild- 
life values  for  the  biologist  to  consider  subjectively, 
Chapter  40,  Economic  Analysis,  presents  the  state-of- 
the-art  for  formal  quantitative  analysis  methods.  For 
further  reading  on  wildlife  values,  I  recommend  Bai- 
ley (1984:34-50),  King  (1966),  Steinhoff  (  1971 ), 
Shaw  and  Zube  (  1980),  or  Kellert  ( 1980a,  b). 


King  ( 1 966 )  comprehensively  categorized  wild- 
life values.  He  divided  these  values,  which  have  sub- 
sequently been  used  by  several  other  authors,  into 
seven  categories: 


( 1 )  Commercial; 

(2)  Recreational; 

(3)  Biotic; 

(4)  Scientific,  philosophical ,  and  educational; 
(  5  )  Aesthetic; 

(6)  Social;  and 

( 7 )  Negative. 


Of  these  values,  only  the  commercial  value  can  be 
easily  translated  into  dollars  for  conventional  eco- 
nomic analyses.  Techniques  for  valuation  of  recrea- 
tion are  available  and  more  are  under  development 
(Sorg  and  Loomis  1985).  Numbers  3  through  6  are, 
for  the  most  part,  unquantifiable  and  will  remain  that 


780 


Evaluation  and  Interpretation 


way  for  a  long  time,  at  least  in  terms  of  placing  a 
dollar  value  on  them.  Negative  values  can  occur  in 
all  categories.  Damage  to  orchards  by  elk  is  relatively 
easy  to  translate  into  dollar  values;  on  the  other 
hand,  the  negative  value  of  Canada  goose  droppings 
on  putting  greens  is  harder  to  quantify. 

Values,  however,  can  be  looked  at  and  measured 
in  different  ways.  Kellert  ( 1980a)  measured  and 
described  wildlife  values  in  terms  of  10  attitudes 
toward  animals  (Table  1 ).  Some  of  these,  such  as  the 
ecologistic  and  scientistic  attitudes,  correspond 
neatly  with  King's  values,  whereas  others,  moralistic 
and  dominionistic,  do  not  have  any  obvious  counter- 
parts. Kellert  (  1980a)  surveyed  American  public 
attitudes  toward  animals  and  more  specifically  wild- 
life (Table  2)  as  one  measure  of  values.  The  attitudes 
are,  of  course,  much  more  complex  than  can  be 
represented  in  a  few  simple  tables,  but  biologists 
need  to  be  aware  of  public  attitudes  (values).  Many 
biologists,  who  were  trained  and  spent  years  in  the 
profession  when  wildlife  management  was  primarily 
game  management,  were  slow  to  recognize  changing 
public  attitudes  toward  wildlife  and  the  changes 


they  wanted  from  wildlife  managers.  U.S.  Bureau  of 
Land  Management  biologists,  for  example,  differ  sig- 
nificantly from  the  general  public  in  the  same  tests 
(Peyton  and  Langenau  1985).  They  score  signifi- 
cantly lower  on  utilitarian  and  negativistic  scales  and 
higher  on  ecological,  scientistic,  dominionistic,  and 
naturalistic  scales.  Biologists  must  be  aware  of  not 
only  their  own  values  but  also  those  of  the  general 
public  and  be  able  to  articulate  these  values.  This  is 
a  difficult,  sensitive,  but  necessary  task,  for  which 
there  are  no  easy  solutions  or  cookbook  approaches. 
Kellert  (1980)  summarized  the  importance  of  the 
task: 

"The  wildlife  management  field  appears  to  be 
confronted  by  a  major  change  in  the  public  it 
serves,  with  many  new  and  atypical  groups  be- 
coming appropriate  recipients  of  professional 
attention.  This  expanded  constituency  must 
inevitably  constitute  a  threat  as  much  as  a  chal- 
lenge to  a  field  that  has  historically  defined  itself 
in  far  narrower  terms.  Nevertheless,  the  chal- 
lenge represents  a  rare  change,  and  it  would  be 
a  disservice  to  the  profession,  let  alone  to  an 


Table  1.     Classification  of  attitudes  toward  animals  (from  Kellert  1980a). 


Attitude 


Description 


Naturalistic 


Ecologistic 


Humanistic 


Moralistic  • 

Aesthetic  • 
Scientistic 

Utilitarian  • 


Dominionistic 


Negativistic 
Neutralists  • 


The  primary  characteristic  is  a  strong  interest  and  affection  for  outdoors  and  wildlife.  Obser- 
vation and  personal  involvement  with  wildlife  are  the  keys  to  the  naturalistic  interest  in  the 
outdoors.  Animals  provide  the  context  and  meaning  for  active  participation  in  natural 
settings. 

Directed  at  a  conceptual  understanding  of  the  interrelationships  of  species  in  the  context  of 
ecosystems,  with  major  concern  for  dependencies  between  animals  and  their  natural  habi- 
tats. Wild  animals  are  valued  less  as  sources  of  affection  or  amusement,  and  more  as  de- 
vices for  comprehending  the  broader  functionings  of  natural  systems. 

Emphasizes  feelings  of  strong  affection  and  attachment  to  individual  animals,  typically  pets. 
The  animal  is  the  recipient  of  feelings  and  emotional  projections  somewhat  analogous  to 
those  expressed  toward  other  people. 

Concern  for  the  ethically  appropriate  humane  treatment  of  animals.  A  basic  moralistic  princi- 
ple is  the  fundamental  similarity  of  all  animals,  each  endowed  with  equivalent  rights  to 
existence. 

Emphasizes  the  attractiveness  or  symbolic  significance  of  animals. 

Concern  for  the  biological  and  physical  characteristics  of  animals.  Wildlife  are  primarily 
interesting  as  problem-solving  objects,  not  as  sources  of  companionship  or  wilderness 
recreation. 

The  relevance  of  animals  is  largely  derived  from  their  usefulness  to  people.  Animals  should 
serve  some  human  purpose  and,  thus,  be  functional  as  sources  of  personal  gain.  Animals 
are  desirable  insofar  as  they  facilitate  some  type  of  tangible  advantage  or  reward. 

Satisfactions  derived  from  the  mastery  and  control  of  animals,  typically  in  a  sporting  context 
(e.g.,  rodeos,  trophy  hunting,  bullfighting).  Animals  provide  opportunities  for  expressions  of 
prowess,  skill,  strength,  and  masculinity. 

Active  dislike  or  fear  of  animals.  This  domain  includes  neutralism — a  more  passively  oriented 
avoidance  of  animals  for  reasons  of  indifference. 

Primary  orientation  is  a  passive  avoidance  of  animals  caused  by  indifference. 


Evaluation  and  Interpretation 


781 


Table  2.     American  attitudes  toward  wildlife  (from  Kellert  1980b). 


Attitude 

Estimated  %  of 
American  Popula- 
tion Strongly 
Oriented  Toward 
the  Attitude* 

Common  Behavioral  Expressions 

Most  Related 
Values/Benefits 

Naturalistic 

10 

Outdoor  wildlife-related  recreation;  back- 
country  use,  nature  birding,  and  nature 

hunting. 

Outdoor  recreation 

Ecologistic 

7 

Conservation  support,  activism  and  mem- 
bership, ecological  study. 

Ecological 

Humanistic 

35 

Pets,  wildlife  tourism,  casual  zoo  visitation. 

Companionship, 
affection 

Moralistic 

20 

Animal  welfare  support/membership,  kind- 
ness to  animals. 

Ethical,  existence 

Aesthetic 

15 

Nature  appreciation,  art,  wildlife  tourism. 

Aesthetic 

Scientistic 

1 

Scientific  study/hobbies,  collecting. 

Scientific 

Utilitarian 

20 

Consumption  of  furs,  raising  meat,  boun- 
ties, meat  hunting. 

Consumptive, 
utilitarian 

Dominionistic 

3 

Animal  spectator  sports,  trophy  hunting. 

Sporting 

Negativistic 

2 

Cruelty,  overt  fear  behavior. 

Little  or  negative 

Neutralists 

35 

Avoidance  of  animal  behavior. 

Little  or  negative 

'Totals  more  than  100%  since  persons  can  be  strongly  oriented  toward  more  than  one  attitude. 


American  public  and  wildlife  resource  in  need, 
if  the  professional  reaction  was  more  to  avoid 
an  alien  reality  than  a  creative  and  bold 
response  to  an  evolutionary  opportunity." 


Clearly  the  days  of  simply  managing  ducks  for 
duck  hunters  and  deer  for  deer  hunters  are  gone. 
The  biologist  must  consider  and  explain  wildlife 
values  to  managers  and  others  when  presenting  in- 
formation from  inventory  and  monitoring  studies. 


SUMMARY 

Interpretation  and  evaluation  are  an  important 
part  of  the  inventory  and  monitoring  process.  Biolo- 
gists must  explain  to  the  manager  and  others  what 
the  information  means.  Explanation,  at  this  stage, 
should  be  based  on  all  sources  of  information,  not 
merely  on  a  narrow  statistical  conclusion.  Biologists 
must  also  explain  the  values  of  the  wildlife  resources 
under  consideration.  To  do  this,  biologists  should 
be  aware  of  the  values  the  public  places  on  wildlife 
resources,  since  these  may  differ  significantly  from 
their  personal  values. 


782 


Evaluation  and  Interpretation 


LITERATURE  CITED 


BAILEY,  J. A.  1984.  Principles  of  wildlife  management.  John 
Wiley  and  Sons,  New  York,  NY.  373pp. 

KELLERT,  S.R.  1980a.  Americans'  attitudes  and  knowledge 
of  animals.  Trans.  North  Am.  Wildl.  Nat.  Resour.  Conf. 
45:1 1 1-124. 

.  1980b.  Contemporary  values  of  wildlife  in  Ameri- 
can society.  Pages  31-60  in  Shaw,  WW.  and  EH. 
Zube,  eds.  Wildlife  Values.  Center  for  Assessment  of 
Noncommodity  Natural  Resource  Values,  School  of 
Renewable  Natural  Resources,  Univ.  Arizona,  Institu- 
tional Series  Rep.  1.  Tucson.  117pp. 

KING,  R.T.  1966.  Wildlife  and  man.  New  York  Conserva- 
tionist 20:8-1 1. 

KRUTILLA,  J.V.  1974.  Methods  for  estimating  the  value  of 
wildlife  resources.  Pages  125-136  in  Bailey,  J.  A.,  W. 
Elder,  and  T.D.  McKinney,  eds.  Readings  in  Wildlife 


Conservation.  The  Wildlife  Society,  Washington,  DC. 
722pp. 

MERRIAM-WEBSTER,  A.  1983.  Webster's  ninth  new  Col- 
legiate dictionary.  Merriam-Webster  Inc.,  Publishers, 
Springfield,  MA. 

PEYTON,  R.B.  and  E.E.  LANGENAU,  Jr.  1985.  A  compari- 
son of  attitudes  held  by  BLM  biologists  and  the  gen- 
eral public  towards  animals.  Wildl.  Soc.  Bull.  13:117- 
120. 

SHAW,  WW.  and  EH.  ZUBE,  eds.  1980.  Wildlife  values. 
Center  for  Assessment  of  Noncommodity  Natural 
Resource  Values,  School  of  Renewable  Natural  Re- 
sources, Univ.  Arizona,  Institutional  Series.  Rep.  1. 
Tucson.  1 17pp. 

SORG,  C.F.  and  J.  LOOMIS.  1985.  An  introduction  to  wild- 
life valuation  techniques.  Wildl.  Soc.  Bull.  1338-46. 

STEINHOFF,  H.W.  1971.  Communicating  complete  wild- 
life values  of  Kenai.  Trans.  North  Am.  Wildl.  and  Nat. 
Resour.  Conf.  36:428-438. 


Evaluation  and  Interpretation 


783 


40 

ECONOMIC 
ANALYSIS 

John  B.  1 .00 ni is 


U.S.  Fish  and  Wildlife  Service 

Western  Energy  and  land  Use  Team 

Creekside  One  Building 

2627  Redwing  Road 

Fort  Collins,  CO  80526-2899 


Editor's  Note:  Many  biologists  are  just  beginning  to 
recognize  the  need  for  economic  analyses  of  wild- 
life values.  In  many  cases  the  economic  evaluation 
of  the  effects  of  a  project  will  affect  decisions  more 
than  the  biological  evaluation.  Biologists  need  to 
be  aware  of  bio -economic  techniques  and  theory  so 
that  they  can  plan  biological  assessments  to  pro- 
vide economists  and  managers  with  the  informa- 
tion they  need  on  economic  values  of  wildlife. 


INTRODUCTION 

Several  new  federal  directives  requiring  evalua- 
tion of  the  dollar  benefits  of  wildlife  have  increased 
the  need  for  biologists  and  economists  to  understand 
the  available  bio-economic  techniques.  Improved 
techniques  for  valuation  of  amenity  (non-marketed 
natural  resources),  such  as  wildlife,  have  largely  dis- 
placed older,  more  subjective  (and  often  erroneous) 
approaches.  Any  reader  unfamiliar  with  the  major 
advances  in  the  economics  of  non-marketed  natural 
resources  will  be  pleasantly  surprised  with  improve- 
ments made  since  the  1970s. 

The  purposes  of  this  chapter  are  to  describe 
concepts  and  measurement  of  economic  values  of 
wildlife,  to  provide  an  overview  of  bio-economic 
analysis  techniques,  and  to  discuss  some  of  the  issues 
and  common  misconceptions  about  economic  anal- 
yses of  wildlife  values.  More  detailed  information  on 
actual  bio-economic  analysis  techniques  can  be 
found  in  the  appendix. 

Most  federal  agencies  are  required  to  translate 
the  biological  effects  of  their  actions  into  economic 
values.  For  example,  the  U.S.  Bureau  of  Land  Manage- 
ment's (BLM's)  Final  Rangeland  Improvement  Policy 
(U.S.  Department  of  the  Interior,  Bureau  of  Land 
Management  1982)  indicates  that  willingness  to  pay 
values  for  wildlife  need  to  be  developed  so  they 
can  be  used  in  rangeland  investment  (SageRam)  pro- 
grams that  rank  alternative  livestock  management 
actions.  The  U.S.  Forest  Service  planning  procedures 
require  estimates  of  present  net  value  of  all  re- 
sources having  an  established  market  value  or  an 
assigned  value  (U.S.  Department  of  Agriculture,  For- 
est Service  1982).  Within  these  resources,  wildlife 
recreation  has  a  dollar  value  set  by  the  Chief  of  the 
U.S.  Forest  Service;  this  value  is  initially  developed 
from  existing  research  on  the  economic  value  of 
wildlife  recreation.  The  present  net  value  (in  dollar 
terms )  has  become  one  of  the  U.S.  Forest  Service's 
key  criteria  in  comparing  planning  alternatives  and 
in  determining  what  represents  maximum  net  public 
benefits  (Peterson  1983). 


Current  Address:  Division  of  Environmental  Studies,  University  of 
California  at  Davis,  Davis,  CA  95616. 


Economic  Analysis 


785 


WILDLIFE  VALUES  AND  THE  PUBLIC  TRUST 
DOCTRINE 

Much  confusion  exists  about  the  economic 
value  of  wildlife.  In  this  section,  the  types  of  wildlife 
values  and  their  relationship  to  the  public  trust  doc- 
trine are  discussed. 

Economic  value  is  often  confused  with  financial 
value.  For  a  good  or  service  to  have  economic  value, 
it  must  meet  the  following  two  criteria: 


( 1 )  It  must  provide  some  consumers  satisfaction 
or  enjoyment,  and 

(2)  The  good  or  service  must  be  scarce  in  the 
sense  that  consumers  want  more  than  is  avail- 
able at  no  cost. 

Wildlife  certainly  meets  both  of  these  criteria. 

Economic  values  include  commercial,  recrea- 
tional, option,  existence,  and  bequest  values  (Figure 
1 ).  Besides  the  economic  values  of  on-site  recreation 
(both  consumptive  and  non-consumptive)  and  com- 
mercial uses  of  wildlife,  there  are  many  off-site  user 
values.  These  include  option,  existence,  and  bequest 
value.  Option  value  refers  to  an  individual's  willing- 
ness to  pay  to  maintain  wildlife  recreation  opportun- 
ities. Option  value  can  be  thought  of  as  an  insurance 
premium  people  would  pay  to  ensure  wildlife  recre- 
ation opportunities  in  the  future.  Existence  value  is 
the  economic  benefit  received  from  simply  knowing 
wildlife  exists.  Bequest  value  is  the  willingness  to 
pay  for  providing  wildlife  resources  to  future 
generations. 


Figure  1.     Economic  values  of  wildlife. 
786  Economic  Analysis 


The  first  pitfall  one  faces  in  identifying  and 
quantifying  wildlife  values  concerns  the  difference 
between  economic  values  and  financial  values.  Finan- 
cial values  are  defined  as  actual  revenue  or  sales 
received  by  firms  or  public  agencies  (i.e.,  cash 
changing  hands).  Financial  values  ignore  social  bene- 
fits or  costs  that  cannot  be  captured  as  revenue  by 
firms.  At  best,  financial  values  are  a  subset  of  eco- 
nomic values  (see  Figure  1 )  and,  at  worst,  may  be  a 
serious  distortion. 

The  disparity  between  economic  and  financial 
values  for  wildlife  is  confusing  to  resource  managers 
and  local  officials  accustomed  to  dealing  only  with 
marketed  resources  such  as  coal  or  timber.  For  these 
resources,  it  is  customary  to  assume  the  economic 
and  financial  values  of  additional  units  of  outputs  are 
almost  synonymous. 

Financial  values  (sales  revenue,  profit)  are  useful 
in  answering  questions  about  profitability  of  guide 
services  or  retail  outlets  for  recreational  equipment. 
Financial  feasibility  of  business  is  often  important 
to  county  and  state  officials  from  the  standpoint  of 
job  creation  or  property  taxes.  These  legitimate  con- 
cerns have  little  to  do  with  the  economic  value  of 
wildlife,  however. 

In  the  U.S.,  laws  assign  ownership  of  the  nation's 
wildlife  resources  to  the  state  or  federal  government. 
The  governments  thus  have  the  responsibility  to 
manage  these  resources  as  trustees  for  the  benefit  of 
the  public.  Public  ownership  implicitly  recognizes 
that  inefficient  resource  allocation  would  result  with- 
out cooperative  intervention.  The  public  agent  is 
expected  to  incorporate  all  benefits  and  costs  into 
decisions  and  not  merely  the  financial  costs  and  re- 
turns seen  by  private  firms.  This  is  called  the  Public 
Trust  Doctrine. 

The  Public  Trust  Doctrine  recognizes  that  all 
benefits  and  costs  of  wildlife  to  people  are  not  cap- 
tured in  the  commodity  values  of  private  markets. 
Therefore,  the  public  agencies  are  responsible  for 
protecting  the  interests  of  birders,  hunters,  and  citi- 
zens who  enjoy  knowing  wildlife  exists. 

Although  public  agencies  are  charged  with  pro- 
tecting all  wildlife  values,  current  methodology  for 
quantifying  such  values  is  often  limited  to  recrea- 
tional and  commercial  values. 


MEASUREMENT  OF  ECONOMIC  VALUES  OF 
WILDLIFE 

As  discussed  earlier,  the  economic  values  of 
wildlife  and  wildlife  recreation  differ  from  financial 
values.  Even  in  determining  economic  values  of  wild- 
life, some  confusion  still  exists.  Before  techniques 


for  measuring  economic  value  are  discussed,  it  is 
necessary  to  have  an  operational  definition  of  just 
what  needs  to  be  measured. 

In  the  marketplace  framework  of  supply  and 
demand,  the  supply  curve  defines  the  quantities  of  a 
good  that  producers  are  willing  to  market  in  return 
for  various  payments.  A  supply  curve  has  a  positive 
slope  because,  at  a  higher  price  per  unit,  producers 
are  willing  to  supply  more  goods.  The  demand  curve 
is  generated  by  the  quantities  of  a  good  that  con- 
sumers are  willing  to  buy  when  faced  with  various 
prices.  Higher  prices  will  lead  to  a  lower  quantity 
demand,  causing  a  negatively  sloped  demand  curve. 
The  intersection  of  the  supply  and  demand  curves 
indicates  equilibrium — which  is  reached  at  a  price 
where  the  quantity  that  producers  are  willing  to 
supply  equals  the  quantity  consumers  are  willing  to 
buy  (Figure  2). 


60t 


QUANTITY  OF  X 


Figure  2.     Supply  and  demand  for  good  X. 

In  the  hypothetical  example  in  Figure  2,  equilib- 
rium is  reached  at  a  price  of  $30  and  supply  of  good 
X  equals  demand.  The  demand  curve  shows  that  a 
consumer  is  willing  and  able  to  pay  $60  for  the  first 
unit  of  good  X.  When  faced  with  a  price  of  $30,  a 
consumer  surplus  of  $30  results  on  the  first  unit. 
Alternatively,  consider  the  situation  where  X  is  free. 
In  this  case,  eight  units  of  X  would  be  demanded 
and  the  entire  area  under  the  demand  curve  would 
be  a  measure  of  consumer  surplus  with  no  revenue 
collected.  From  this  hypothetical  example,  one  can 
see  that  the  price  charged  is  a  measure  of  expendi- 
tures or  financial  value  while  the  difference  between 
what  the  individual  is  willing  to  pay  and  expendi- 
tures is  a  measure  of  net  benefits  (consumer  surplus 
or  net  willingness  to  pay). 


Although  the  same  concepts  of  supply  and  de- 
mand and  consumer  surplus  apply  to  all  goods 
whether  they  are  marketed  or  not  (as  wildlife),  em- 
pirical estimation  may  be  difficult  for  unmarketed 
goods.  Supply  of  a  species  is  regulated  by  wildlife 
managers  and  the  carrying  capacity  of  the  land. 
When  carrying  capacity  has  not  been  reached,  the 
management  costs  of  supplying  various  quantities  of 
a  species  can  be  used  to  generate  a  supply  curve. 
Estimating  a  demand  curve  is  less  straightforward.  In 
the  case  of  a  market  good,  a  demand  curve  is  gener- 
ated by  monitoring  purchasing  behavior  of  individu- 
als at  various  prices  and  mapping  a  demand  curve. 
However,  in  the  case  of  wildlife-related  recreation,  a 
demand  curve  cannot  be  derived  easily  because 
wildlife  ordinarily  is  considered  a  publicly  owned 
good  that  one  does  not  have  to  pay  to  hunt  or  view. 
Hunting  and  fishing  licenses  are  not  market-deter- 
mined prices  but  administratively  determined  fees. 
As  a  result,  prices  cannot  be  associated  with  various 
quantities,  and  a  demand  curve  cannot  be  plotted 
directly. 

For  wildlife  values,  the  demand  curve  is  used  as 
a  basis  for  estimating  consumers'  (or  in  this  situa- 
tion, recreationists' )  willingness  to  pay  for  increases 
in  wildlife  recreation  opportunities.  (Harberger 
1971;  Sassone  and  Schaffer  1978).  A  hypothetical 
demand  curve  (Figure  3)  shows  the  number  of  trips 
a  recreationist  would  take  at  different  travel  costs. 
The  total  area  under  the  curve  is  the  gross  willing- 
ness to  pay.  The  net  willingness  to  pay  is  the  area 
above  the  actual  cost  but  below  the  demand  curve. 


60   - 

Eb       50  - 

DC 
1- 

OC 

UJ         40  - 

a. 

(/) 
OC 

<         30   - 

_l 

-1 

o 

Q 

20   - 

Actual  cost    @- 

yyyyyyyyy/yyyyyfa^QmaV[^  curve 

IN. 

1        Nl 

1             ^\ 

1                      \ 

0            1            2            3          (V)         5            6 

TRIPS   PER   YEAR 

Figure  3.     Hypothetical  demand  curve  for  wildlife 
recreation. 


Economic  Analysis 


787 


If  the  actual  cost  was  $10  then  the  shaded  area  in 
Figure  3  is  the  net  willingness  to  pay  which  is  $105. 
The  gross  willingness  to  pay  would  be  $105  plus 
the  $40  (4  trips  x  $10  per  trip)  which  comes  to 
$145. 

Economists  term  consumers'  net  willingness  to 
pay  as  "consumer  surplus."  Consumer  surplus  repre- 
sents the  consumers'  additional  willingness  to  pay 
for  the  opportunity  to  hunt  or  fish  at  some  site.  It  is 
a  net  or  additional  willingness  to  pay  because  it  is 
added  to  current  expenditures. 

Consumer  surplus  sometimes  stands  as  a  con- 
ceptual "stumbling  block"  for  those  involved  in  eco- 
nomic valuation.  Consumer  surplus  is  often  not  seen 
as  a  real  or  tangible  economic  benefit  because  it 
represents  money  that  has  not  actually  been  col- 
lected by  a  business  or  government  agency.  If  indi- 
viduals could  be  feasibly  charged  a  price  equal  to 
their  maximum  willingness  to  pay  for  each  unit,  the 
estimate  of  consumer  surplus  could  be  verified.  Such 
pricing  schemes  exist,  but  the  government  does  not 
use  them  to  capture  the  full  willingness  to  pay  for 
each  unit  (the  consumer  surplus)  as  revenue. 

Several  methods  are  used  to  derive  demand 
curves  indirectly  and  provide  measures  of  net  will- 
ingness to  pay.  Those  most  often  used  include  the 
travel  cost  method  and  the  contingent  value  method. 

Travel  Cost  Method  (TCM) 

This  approach  was  developed  by  Clawson 
(1959)  to  empirically  estimate  recreation  benefits. 
The  method  is  based  on  the  belief  that  travel  cost 
can  be  used  as  a  proxy  for  price  in  deriving  a  de- 
mand curve  for  a  recreation  site.  The  first  step  of 
zonal  TCM  analysis  involves  dividing  the  area  around 
the  site  into  zones  or  counties  of  origin  of  recrea- 
tionists.  One  assumes  that  the  costs  from  a  particular 
zone  to  a  recreation  site  are  the  same  for  all  individ- 
uals in  that  zone.  Based  on  origin  data,  a  visitation 
rate  (trips  per  capita)  is  calculated  for  each  zone. 
Regression  analysis  can  be  used  to  estimate  a  func- 
tion for  visitation  rates  based  on  travel  cost  and  socio- 
economic data.  Such  a  regression  line  is  shown  in 
Figure  4. 

The  different  travel  costs  for  making  a  trip  from 
each  of  four  origins  (counties)  surrounding  a  site  are 
plotted  against  the  number  of  trips  per  capita  from 
each  county  to  the  site.  These  different  combinations 
of  travel  cost  and  trips  per  capita  represent  price- 
quantity  points  that  trace  a  demand  curve.  From  this 
demand  curve,  the  consumer  surplus  or  net  willing- 
ness to  pay  for  recreation  at  a  particular  site  can 
be  calculated  by  taking  the  area  under  the  demand 
curve  but  above  the  travel  cost  for  each  zone  of 
origin.  For  details  on  use  of  the  technique,  see 
Rosenthal  et  al.  (1984). 


NUMBER  OF  TRIPS  PER  CAPITA  TO  SITE  1 

S1  represents  supply  or  travel  costs 


Figure  4.     Regression  of  travel  costs  versus  number 
of  visits  per  capita. 

The  TCM  requires  data  on  wildlife  recreation- 
ists'  travel  cost  to  or  distance  from  a  recreation  site. 
If  surveys  providing  travel  cost  (or  distance)  and 
city  or  county  of  residence  are  available,  they  should 
be  used.  However,  one  advantage  of  the  TCM  is  that 
existing  information  from  hunting  licenses,  game 
tags,  or  even  license  plates  can  be  used  to  determine 
the  wildlife  recreationist's  city  or  county  of  resi- 
dence. If  one  knows  the  county  or  city  of  residence, 
the  round-trip  distance  to  the  fishing  or  hunting 
site  can  be  calculated  from  maps.  The  distance  can 
be  converted  to  a  travel  cost  by  using  the  govern- 
ment publication  "Cost  of  Owning  and  Operating  a 
Motor  Vehicle"  (U.S.  Department  of  Transportation 
1985). 

The  information  on  visits  needs  to  be  grouped 
by  county  so  the  trips  per  capita  can  be  calculated 
by  dividing  trips  by  county  population.  This  be- 
comes a  dependent  variable  in  the  regression  analy- 
sis used  to  statistically  estimate  the  demand  curve. 
Thus,  it  is  generally  important  to  know  the  county 
or  city  (or  zip  code)  of  wildlife  users.  Knowing  the 
number  of  wildlife  recreationists  per  vehicle  and 
number  of  days  per  trip  is  also  useful. 

Data  sources  are  often  available  for  calculating 
travel  costs.  For  example,  several  data  sources  exist 
for  big  game  hunting.  State  wildlife  management 
agencies  often  record  the  zip  code  or  city  of  resi- 
dence from  hunting  licenses  or  post-season  harvest 
surveys  for  the  general  big  game  seasons.  Wildlife 
agencies  also  record  the  herd  unit  from  which  an  an- 
imal is  harvested.  These  provide  the  two  basic  types 
of  data  for  TCM:  location  of  residence  and  herd  unit. 


788 


Economic  Analysis 


Matching  these  two  together  allows  calculation  of 
round-trip  distance  and  estimation  of  round-trip 
travel  costs.  For  details  on  these  procedures  see 
Loomis(1982). 

In  many  states,  license  plate  numbers  are  keyed 
to  county  of  residence.  By  simply  visiting  the  recrea- 
tion site  of  interest  and  recording  license  plate  num- 
bers (in  some  States,  the  county  name  is  even 
printed  on  the  plate),  one  can  get  an  approximate 
idea  of  the  origin  of  wildlife  recreationists.  This  is 
useful  in  collecting  data  on  nonconsumptive  users  of 
wildlife,  such  as  birding  or  photography. 

The  TCM  can  be  used  to  measure  the  economic 
value  of  nonconsumptive  wildlife  recreation  such  as 
birdwatching  or  wildlife  photography.  Data  on  dis- 
tance traveled  can  be  obtained  on  site  at  areas 
where  a  significant  amount  of  nonconsumptive  use 
of  wildlife  takes  place. 

A  useful  characteristic  of  TCM  is  that  it  can  be 
used  for  valuation  and  use  estimation  simultaneously. 
A  simple  single  site  TCM  demand  curve  can  be  used 
to  estimate  hunter  demand  over  time.  This  is  done 
by  multiplying  each  county's  current  trip  per  capita 
rate  times  future  county  population. 

Thus,  if  a  demand  estimate  is  needed  for  the 
year  1990  for  example,  all  one  must  do 


( 1 )  Use  the  equation  to  estimate  current  trips  per 
capita  at  the  current  travel  cost  for  each 
county,  and 

(2)  Multiply  this  current  trip  per  capita  by  each 
county's  forecasted  population  in  1990. 

When  using  a  regional  TCM  that  includes  a  qual- 
ity variable  such  as  harvest,  one  can  even  estimate  a 
future  visit  rate  per  capita  that  varies  with  future 
harvest.  To  get  a  future  demand  estimate,  the  analyst 
would  multiply  each  county's  future  trip  per  capita 
figure  by  each  county's  future  population. 

The  strengths  and  weaknesses  of  TCM  as  a  tool 
for  estimating  benefits  of  recreation  are  described  by 
Dwyer  et  al.  ( 1977)  and  Rosenthal  et  al.  ( 1984). 


Contingent  Value  Method  (CVM) 

The  Contingent  Value  Method  (CVM)  is  also 
known  as  the  "Direct  Method"  or  "Bidding  Method." 
Unlike  the  familiar  market  situation  where  people 
alter  consumption  in  response  to  price  changes, 
bidding  involves  asking  individuals  to  respond  to 
changes  in  hypothetical  prices  for  an  unmarketable 
good.  The  term  "contingent  valuation"  stems  from 
asking  individuals  how  their  behavior  would  change, 


contingent  on  a  new  hypothetical  situation.  Thus 
individual  users  are  asked  to  provide  data  to  estimate 
their  net  willingness  to  pay  (consumer  surplus). 

Estimation  (distinguished  from  application)  re- 
quires administration  of  a  carefully  constructed  sur- 
vey. The  U.S.  Water  Resources  Council  (1979)  and 
Dwyer  et  al.  (1977)  provide  good  reviews  of  the 
steps  in  survey  design  and  implementation.  Although 
wildlife  biologists  at  the  field  level  will  rarely  con- 
struct such  a  survey,  they  should  understand  how 
the  method  works  and  what  the  information  means. 

The  CVM  relies  on  mail  and  personal  or  tele- 
phone interviews.  The  interview  begins  with  an  in- 
troduction and  a  full  explanation  of  the  purpose.  The 
introduction  is  often  aided  by  the  use  of  maps, 
graphs,  or  photographs  in  an  attempt  to  define  the 
good  being  valued.  Next,  participants  are  asked  ( 1 ) 
how  much  they  would  be  willing  to  pay  to  achieve 
an  improved  situation,  (2)  how  much  compensation 
they  would  require  to  accept  a  reduction  from  their 
current  situation,  or  (  3 )  how  much  they  would  be 
willing  to  pay  for  current  circumstances.  Depending 
on  how  the  questions  are  phrased,  the  resulting  bids 
represent  either  a  willingness-to-pay  or  willingness- 
to-sell  measure  of  consumer  surplus  (see  Brookshire 
et  al.  1980). 

The  iterative  bid  procedure  may  be  necessary  to 
obtain  a  complete  measure  of  maximum  willingness 
to  pay.  In  this  process,  a  recreationist  reacts  to  a 
starting  bid  and  is  successively  asked  if  he  or  she 
would  pay  higher  incremental  amounts  until  a  nega- 
tive response  is  obtained.  The  last  positive  response 
is  considered  the  measure  of  maximum  willingness 
to  pay.  The  bidding  process  forces  individuals  to 
analyze  their  preferences  more  fully,  providing  a 
more  accurate  measure  of  consumer  surplus. 

Values  of  trout  fishing  and  deer  hunting  in  1 1 
western  states  derived  with  CVM  are  shown  in  Table 
1 .  These  dollar  values  per  day  resulted  from  an  itera- 
tive bidding  sequence  performed  in  a  personal  sur- 
vey and  represent  net  willingness  to  pay  over  and 
above  current  costs  (in  1980  dollars). 

Other  sources  of  CVM  derived  values  include 
articles  appearing  in  such  journals  as  Land  Econom- 
ics, American  Journal  of  Agricultural  Economics, 
Western  Journal  of  Agricultural  Economics,  and 
Journal  of  Leisure  Research.  PhD  dissertations  and 
Masters  theses  prepared  in  the  economics  depart- 
ment of  the  University  of  Wyoming,  University  of 
New  Mexico,  Utah  State  University,  and  University  of 
Washington  often  empirically  estimate  dollar  values 
for  wildlife  recreation  in  western  states  by  using  the 
CVM.  The  Transactions  of  the  North  American 
Wildlife  Conference  often  contains  economic  values 
derived  by  CVM. 


Economic  Analysis 


789 


Table  1.     Value  (net  willingness  to  pay)  for  trout  fishing  and  deer  hunting  (U.S.  Department  of  the  Interior, 
Fish  and  Wildlife  Service  and  U.S.  Department  of  Commerce,  Bureau  of  Census  1982). 


Dollars  per  Day  of  Trout  Fishing 

Dollars  per  Day  of  Deer  Hunting 

Arizona 

Value2 

Mean3 

Value2 

Mean3 

19.54 

±4.19 

32.50 

±4.95 

California 

20.53 

±2.08 

37.35 

±7.87 

Colorado 

16.16 

±1.95 

23.49 

±4.18 

Idaho 

12.93 

±0.93 

28.77 

±2.63 

Montana 

16.47 

±1.88 

25.42 

±2.43 

Nevada 

12.35 

±1.48 

29.02 

±3.92 

New  Mexico 

15.70 

±1.46 

29.11 

±2.60 

Oregon 

13.49 

±1.82 

21.44 

±3.06 

Utah 

12.57 

±1.17 

25.72 

±2.69 

Washington 

14.03 

±2.54 

24.18 

±4.26 

Wyoming 

16.87 

±1.54 

36.26 

±3.26 

'Values  derived  using  Contingent  Value  Method;  iterative  bidding  was  used. 

2Values  are  expressed  in  1980  dollars. 

3Dollars  are  expressed  as  mean,  plus  or  minus  standard  error. 


Experience  with  CVM  indicates  that  it  can  reli- 
ably measure  values  if  the  survey  is  carefully  con- 
structed and  carried  out.  However,  studies  generally 
indicate  that  CVM  estimates  of  net  willingness  to  pay 
are  conservative  estimates  of  value  (Bishop  and  He- 
berlein  1979;  Brookshire  et  al.  1982).  This  conclu- 
sion is  based  on  comparisons  of  actual  market  data 
to  the  Contingent  Valuation  bids. 

The  CVM  is  particularly  well-suited  for  estimat- 
ing the  value  of  nonconsumptive  wildlife  recreation 
as  well  as  option,  existence,  and  bequest  values. 

Persons  having  little  or  no  experience  with 
CVM  should  seek  assistance  from  experienced  users 
of  the  method.  Data  from  studies  that  employed 
the  method  may  also  be  useful,  since  they  are  ex- 
pressed in  units,  such  as  dollars  per  hunter  day. 

Advantages  of  Valuation  Methodologies 

Two  techniques  for  measuring  the  net  willing- 
ness to  pay  for  wildlife  recreation  have  been  dis- 
cussed— Travel  Cost  Method  and  Contingent  Value 
Method.  Each  method  has  its  strengths  and  weak- 
nesses in  valuing  wildlife  recreation.  TCM  has  two 
advantages: 


(  1 )    It  has  the  capability  to  use  existing  data  from 
hunter  applications,  licenses,  or  other  sources 


for  statistically  estimating  a  demand  curve 
for  the  activity  at  a  particular  site. 


(2)    It  can  be  used  for  both  a  benefit  estimate  and 
use  projection  (demand  projection)  over 
time  from  the  same  model  framework. 

The  advantage  of  CVM  is  its  ability  to  look  di- 
rectly at  the  dollar  value  of  improvements  in  harvest 
or  quality  of  recreation  (for  example,  the  increase 
in  value  for  higher  success  rates  or  for  trophy 
animals ). 

There  are  several  advantages  to  empirical  tech- 
niques such  as  the  TCM  and  CVM.  One  advantage 
is  that  these  methods  provide  objective  measures  of 
the  benefits  of  wildlife  recreation  as  revealed  by 
the  recreationists  themselves.  This  is  a  significant 
improvement  over  the  subjective  methods  whereby 
specialists  assign  the  value  based  on  the  benefits  they 
think  the  recreationists  receive. 

Another  advantage  of  TCM  or  CVM  is  that  confi- 
dence intervals  around  the  benefit  estimates  can  be 
displayed  to  give  the  decisionmaker  some  indication 
of  the  precision  of  the  benefit  estimate.  In  this  way, 
sensitivity  analysis  can  be  performed  and  determina- 
tions made  as  to  whether  more  precise  estimates 
of  wildlife  values  are  critical  to  ranking  of 
alternatives. 


790 


Economic  Analysis 


BIOLOGICAL-RECREATIONAL-ECONOMIC 
RELATIONSHIPS:  OVERVIEW 

Many  methods  of  predicting  bio-economic  re- 
sponse require  estimating  how  a  change  in  manage- 
ment will  cause  change  in  population  numbers  and 
thus  in  recreation  (hunter  or  angler)  days.  Values 
from  CVM  can  then  be  used  to  attach  a  dollar 
amount  to  a  recreation  day.  Five  procedures  for  bio- 
economic  analysis  are  listed  below: 


(4)  "Short  Form  for  Bio-economic  Evaluations  of 
Wildlife  in  Washington  State"  (Oliver  et  al. 
1975). 

(5)  "Human  Use  and  Economic  Evaluation" 
(HUEE)  system  (U.S.  Department  of  the  Inte- 
rior, Fish  and  Wildlife  Service  1980a). 

These  procedures  are  described  here  briefly; 
two,  Hunter  Day  Short  Form  and  HUEE,  are  docu- 
mented in  more  detail  in  Appendixes  I  and  II. 


( 1 )  "Hunter  Day  Short  Form"  developed  by  the 
U.S.  Bureau  of  Land  Management's  Moab  Dis- 
trict Office  (U.S.  Department  of  the  Interior, 
Bureau  of  Land  Management  1983).  This 
procedure  has  been  developed  for  big  game 
hunting  only. 

(2)  Suislaw  National  Forest's  Salmonoid  Fisheries 
Model  (Kunkel  andjanik  1976). 

(3)  "Biological  Response  Approach"  developed 
by  the  U.S.  Bureau  of  Land  Management,  Ore- 
gon State  Office  (U.S.  Department  of  the  Inte- 
rior, Bureau  of  Land  Management  1979). 


All  of  these  procedures  share  certain  similarities 
(Figure  5).  In  particular,  all  five  approaches  require 
that  harvest  (or  population)  figures  be  known  or 
estimable  (Figure  5). 

In  the  State  of  Washington's  "Bio-Economic 
Short  Form"  (Oliver  et  al.  1975)  and  the  U.S.  Forest 
Service's  anadromous  fish  model  (Kunkel  andjanik 
1976),  biological  effects  must  be  estimated  in  terms 
of  effects  on  populations.  If  population  simulation 
models  are  available,  the  linkage  of  resource  deci- 
sions to  populations  may  be  possible. 

Although  HUEE  analysis  (U.S.  Department  of  the 
Interior,  Fish  and  Wildlife  Service  1980a)  and  the 


USFWS 

HUEE 

(5) 


Recreation  Total 

days  =     recreation 

Animal  days 


$ 
Day 


Total 

recreational 

value 


BLM's 

(1)  Moab  District 
Short  Form 

(2)  Oregon  State 
Office 


USFS 

(3)  Salmonid 
Fisheries 
Model 


State  of  Washington 

(4)   Bio-Economic 
Short  Form 


Figure  5.     Overview  of  bio-economic  approaches. 


Economic  Analysis 


791 


Oregon  Biological  Response  approach  (U.S.  Depart- 
ment of  the  Interior,  Bureau  of  Land  Management 
1979)  can  model  the  biological  impacts  in  terms  of 
population  or  harvest,  both  have  the  capability  to 
convert  changes  in  habitat  variables  (e.g.,  food, 
cover,  reproduction)  into  changes  in  carrying  capac- 
ity. Changes  in  carrying  capacity  can  often  be  con- 
verted to  changes  in  population  or  harvest.  Since 
changes  in  habitat  variables  due  to  management  ac- 
tions are  often  easier  to  predict,  this  habitat-based 
evaluation  may  be  useful  in  evaluating  Habitat  Man- 
agement Plans  (HMPs)  or  Allotment  Management 
Plans  (AMPs). 

The  next  step  required  in  all  five  approaches  is 
the  estimation  of  wildlife  recreation  use  levels  asso- 
ciated with  a  given  number  of  animals  (or  fish)  avail- 
able for  harvest.  Typically,  this  involves  multiplying 
hunter  days  or  angler  days  per  animal  harvested 
by  the  change  in  animals  available  for  harvest. 

Once  the  change  in  days  of  wildlife  recreation  is 
known,  these  days  are  often  multiplied  by  an  average 
dollar  value  per  day.  This  yields  an  estimate  of  the 
economic  effects  of  the  change  in  wildlife  recreation 
associated  with  some  original  biological  change,  not 
the  total  economic  effects  to  society  as  a  whole.  It 
does  not  include  option,  bequest,  or  existence  val- 


ues. These  values  can  be  quantified,  but  at  a  cost  that 
often  makes  it  impractical  for  routine  analyses  of 
HMPs,  AMPs,  etc. 

The  five  methods  provide  a  range  of  analysis 
techniques  varying  from  simple  to  complex.  Thus  it 
cannot  be  said  that  one  approach  is  better  than  the 
other,  only  that  one  approach  may  be  more  cost- 
effective  for  screening  possible  allotment  manage- 
ment plans  (AMPs)  or  habitat  management  plans 
(HMPs),  whereas  another,  better  suited  for  in-depth 
analysis  of  the  remaining  (after  screening)  candidate 
AMPs  or  HMPs. 

One  bit  of  inventory  data  is  common  to  all  of 
these  basic  approaches:  determining  the  wildlife 
recreationists'  days  dependent  on  or  produced  from 
the  wildlife  unit  under  study.  This  concern  is  impor- 
tant because  very  few  animals  may  actually  be  har- 
vested in  the  area  under  study.  For  migratory 
animals,  big  game  animals,  and  fish  species,  each 
piece  of  habitat  contributes  something  to  producing 
a  harvestable  animal.  For  such  migratory  animals, 
determining  the  equivalent  number  of  animals  de- 
pendent on  a  habitat  area  can  be  fairly  complex. 
A  step-by-step  procedure  for  doing  this  is  shown  in 
Table  2. 


Table  2.     A  step-by-step  procedure  for  calculating  equivalent  number  of  animals  harvested  that  are  dependent 
on  a  habitat  area. 


Section  1— Animals  harvested  in  the  study  area. 


Enter  animals  harvested  in  study  area 

Multiply  line  1  by  the  percentage  that  are  year-round  residents 

Number  of  animals  harvested  on  site  that  are  migrants;  subtract  line  3 
from  line  1 

Multiply  line  4  by  the  dependency  of  these  migrant  animals  on  the  study 
area 

Subtotal  of  animals  harvested  on  study  area;  add  lines  3  and  6 


line  1 

line  2 

line  3  (answer) 

line  4 


line  5 

line  6  (answer) 

line  7 


Section  II — Animals  harvested  outside  the  study  area  but  dependent  on  the  study  area. 


Enter  animals  harvested  in  state  wildlife  management  area  or  herd  unit 
adjacent  to  or  surrounding  study  area 

Enter  percentage  of  these  animals  that  spend  time  on  the  study  area  but 
are  harvested  elsewhere 

Multiply  the  percentage  in  line  2  by  line  1  and  enter  on  line  3 

Multiply  line  3  by  the  importance  of  the  study  area's  habitat  to  the  species 
(can  use  time  in  study  area) 


line  1 
line  2 

line  3 

line  4 

line  5  (answer) 


Section  III— The  total  equivalent  number  of  animals  harvested  that  are  dependent  on  the  study  area  is 
found  by  adding  line  7  from  Section  I  and  line  5  from  Section  II.  Total 


792 


Economic  Analysis 


The  first  step  is  to  determine  the  number  of 
animals  of  a  particular  species  that  are  harvested  in 
the  study  area.  This  can  often  be  determined  by 
looking  at  state  wildlife  management  agencies'  har- 
vest data.  The  population  or  harvest  in  the  wildlife 
management  unit  may  need  to  be  prorated  to  one  or 
more  specific  allotments.  Although  an  acre-per-acre 
proration  is  the  simplest,  use  of  information  on  pop- 
ulation concentration  will  allow  a  more  accurate 
proration.  Details  of  the  Moab  District  Short  Form 
and  HUEE  procedures  are  described  in  Appendixes  I 
and  II,  with  additional  references  provided  there. 


DISCUSSION 

Wildlife  biologists  and  managers  are  often  con- 
fused about  the  economic  value  of  wildlife  because 
they  generally  do  not  have  enough  knowledge  of 
basic  economic  concepts.  This  is  aggravated  by  a 
common  failure  to  identify  and  separate  different 
policy  questions  that  require  different  kinds  of  dollar 
value-related  answers.  The  information  required  by 
questions  of  national  economic  efficiency,  for  exam- 
ple, are  different  from  that  needed  to  address  con- 
cerns about  local  economic  impact. 

Three  issues  cause  confusion: 


means  the  efficiency  of  resource  use  has  been  in- 
creased by  reallocating  resources  from  lower  value 
uses  to  higher  value  uses,  which  is  economically 
efficient. 

The  Resources  Planning  Act  ( RPA ),  National 
Forest  Management  Act  (NFMA),  BLM's  Rangeland 
Investment  Policy,  and  the  U.S.  Water  Resources 
Council  Principles  and  Guidelines  all  require  re- 
source valuation  in  terms  of  net  economic  surplus  to 
the  consumers  or  producers.  This  net  surplus  is  the 
value  remaining  after  all  costs  have  been  paid,  and  is 
the  net  willingness  to  pay.  This  net  economic  value 
is  not  measured  by  the  actual  expenditures  or  costs 
of  the  consumer.  Expenditure  information  is  useful 
for  certain  kinds  of  policy  decisions  requiring  knowl- 
edge of  community  dependency,  as  opposed  to  net 
benefit  and  economic  efficiency.  It  is  also  useful  for 
analyzing  economic  impacts  where  the  purpose  is  to 
expose  the  distribution  of  costs  and  benefits. 

Use  of  actual  expenditure  information  to  meas- 
ure wildlife  benefits  for  efficiency-related  decisions  is 
incorrect  and  misleading.  The  price  actually  paid  is 
an  expenditure  or  cost  incurred  by  the  purchaser. 
This  price  or  expenditure  is  received  as  sales  reve- 
nue or  wages  by  the  provider  of  the  good  to  the 
purchaser. 


( 1 )  The  distinction  between  economic  efficiency 
values  and  expenditures; 

(2)  The  relationship  of  price  to  consumer  sur- 
plus; and 

(3)  The  difference  between  net  benefits  and  re- 
source quality. 

These  are  discussed  here;  more  detail  on  the  subject 
can  be  found  in  Loomis  et  al.  (1984). 

Economic  Efficiency  Values  Versus 
Expenditures 

Many  of  the  questions  posed  by  federal  and 
state  wildlife  programs  involve  determining  whether 
the  economic  gain  from  some  investment,  such  as 
fish  ladders  or  habitat  developments,  exceeds  the 
costs  of  such  developments.  A  similar  question  is 
asked  when  evaluating  National  Forest  plans  or  the 
cost-effectiveness  of  mitigation  plans.  To  determine 
whether  the  benefits  exceed  the  cost  of  some  re- 
source action,  the  willingness  to  pay  (consumer  and 
producer  surplus )  of  project  gainers  needs  to  be 
compared  with  the  willingness  to  pay  of  the  losers 
(Dwyer  et  al.  1977;  U.S.  Water  Resources  Council 
1979,  1983;  Walsh  1986  ).  When  the  willingness-to- 
pay  values  of  project  gainers  exceed  willingness- 
to-pay  values  of  losers,  the  present  net  value  is  posi- 
tive or  the  benefit-cost  ratio  is  greater  than  1.  This 


Costs  are  defined  as  benefits  foregone.  The  more 
it  costs  society  to  harvest  a  certain  number  of  trees, 
the  less  the  net  gain  to  society.  That  is,  the  more 
one  gives  up  to  get  something,  the  less  net  benefit 
there  is  to  having  it.  A  timber  sale  will  often  require 
building  several  miles  of  expensive  road  which  will 
result  in  thousands  of  dollars  of  expenditures. 
However,  if  the  value  of  trees  is  less  than  the  ex- 
penditures, there  has  been  a  net  loss  to  society.  Ex- 
penditures in  excess  of  economic  benefits  means  the 
cost  of  what  was  given  up  exceeded  the  benefits  of 
what  was  received. 

An  example  of  beneficial  treatment  wildlife  gets 
when  the  net  benefits  (gross  benefits  minus  expendi- 
tures) of  agricultural  development  are  compared 
with  habitat  preservation  can  be  seen  in  a  paper  on 
the  Birds  of  Prey  Conservation  Area  in  southern 
Idaho  (Hyde  et  al.  1982).  Hyde  et  al.  evaluated  a 
trade-off  between  agricultural  development  and  pres- 
ervation of  the  prey  base.  The  net  benefits  of  agricul- 
tural development  were  very  low  because  of  the 
high  costs  needed  for  farmers  to  cultivate  the  land 
and  pump  water  from  the  Snake  River.  If  one  judged 
the  economic  benefits  on  an  expenditures  basis, 
the  inefficient  agricultural  development  would  look 
valuable.  However,  given  high  costs,  there  are  the 
low  net  benefits  of  agriculture.  Hence,  the  opportu- 
nity cost  of  maintaining  the  prey  base  for  the  Birds 
of  Prey  Conservation  Area  was  low  and  the  net  bene- 
fits were  greater  for  the  habitat  preservation  option. 


Economic  Analysis 


793 


There  are  legitimate  uses  of  expenditure  infor- 
mation, particularly  measuring  the  local  income  and 
employment  effects  of  a  management  action.  Like  the 
process  of  economic  growth  itself,  changes  in  re- 
source allocations  will  displace  some  job  skills  (and 
workers  having  only  those  job  skills)  and  create 
demands  for  workers  with  different  job  skills.  The 
National  Environmental  Policy  Act  and  local  public 
officials  often  require  federal  agencies  to  display  the 
types  of  workers  and  industries  that  will  be  posi- 
tively or  negatively  affected  by  resource  actions 
taken  on  public  lands. 

Price  Versus  Consumer  Surplus 

Another  obstacle  to  correct  wildlife  valuation  is 
the  common  allegation  that  market  prices  and  con- 
sumers' surplus  cannot  be  compared.  This  problem 
stems  from  confusion  about  underlying  concepts  and 
failure  to  separate  different  questions  requiring  dif- 
ferent answers.  The  correct  measure  of  gross  value 
to  the  consumer  is  the  amount  of  monies  the  con- 
sumer is  willing  to  pay  for  the  goods  in  question. 
Net  value  to  the  consumer  is  this  sum  minus  the  ex- 
penditures required  to  obtain  it,  i.e.,  consumers' 
surplus. 

The  technical  relationship  between  price  and 
consumers'  surplus  is  shown  in  Figure  6.  Again,  the 
demand  curve  is  the  measurement  of  consumers' 
valuation  (Harberger  1971 ).  The  actual  shape  of  the 
demand  function  will  depend  on  the  nature  of  the 
market  and  the  context  of  the  question.  In  Figure  6, 
a  downward-sloping  function  represents  the  entire 
market  or  industry.  For  the  good  in  question,  let  the 
current  quantity  being  exchanged  be  100  units  at  a 
price  of  $10.  The  gross  value  of  one  additional  unit 
in  this  market  area  is  $10.  This  is  the  consumers' 
gross  willingness  to  pay  for  that  unit.  The  price  or 
cost  to  the  consumer  is  also  $10,  so  the  net  willing- 
ness to  pay  is  zero.  The  marginal  unit  has  no  con- 
sumer surplus. 

If  the  quantity  supplied  is  increased  from  100 
units  to  150  units,  the  old  price  of  $10  cannot  be 
used  as  the  consumers'  gross  willingness  to  pay  for 
the  50  additional  units.  The  gross  willingness  to  pay 
for  these  50  additional  units  is  the  area  under  the 
demand  curve  between  100  and  150  units,  in  this  in- 
stance, about  $375.  At  the  new  level  of  consump- 
tion, the  1 50  units  can  all  be  purchased  for  $5  each. 
The  gross  willingness  to  pay  for  the  last  unit  con- 
sumed is  thus  the  market  price  of  $5  and  the  net 
willingness  to  pay  for  this  last  unit  is  zero. 

The  willingness  to  pay  for  the  other  49  added 
units  is  greater  than  $5,  however,  as  shown  by  the 
downward  slope  of  the  demand  curve,  and  thus  gen- 
erates consumers'  surplus.  For  example,  the  100th 
unit  consumed  could  have  been  sold  for  $10,  but  it 
has  been  sold  for  $5.  The  net  willingness  to  pay 


50  100    101      150 

QUANTITY/YEAR 


Figure  6.     Relationship  between  price  and 
consumer  surplus. 


enjoyed  by  the  consumer  for  this  unit  is  thus  $5  in 
the  form  of  consumers'  surplus.  The  additional  con- 
sumer surplus  associated  with  the  increase  of  50 
units  is  the  triangular  area  below  the  demand  curve 
between  100  and  150  units  and  above  the  price  of 
$5,  or  about  $125.  When  the  added  consumer  sur- 
plus from  the  decline  in  price  from  $10  to  $5  on  the 
original  100  units  is  added  to  the  $125,  the  total 
increase  in  consumer  surplus  is  $625.  This  is  the  net 
economic  benefit  of  the  change  whether  the  units 
are  bicycles,  trees,  or  hunter  days.  The  differences 
arise  because  some  questions  ask  for  net  willingness 
to  pay  at  the  margin  (e.g.,  a  small  change  from  100 
to  101 ),  and  some  questions  ask  for  net  willingness 
to  pay  for  non-marginal  changes  (e.g.,  a  very  large 
change  from  100  to  150).  In  any  case,  if  there  is 
surplus  created  by  a  quantity  change,  it  should  be 
measured  and  counted. 


Hunting  and  fishing  often  take  place  in  small  or 
localized  markets.  Most  hunters  and  anglers  visit 
areas  within  200  miles  of  their  homes.  Removing 
fishing  opportunities  at  one  major  cold-water  stream 
or  lake  could  make  a  substantial  change  in  the  price 
(i.e.,  travel  cost)  of  fishing  opportunities.  In  such  a 
situation,  there  will  be  significant  consumer  surplus 
associated  with  eliminating  or  adding  an  opportunity. 
On  the  other  hand,  imagine  a  place  in  Minnesota 
where  there  are  hundreds  of  identical  uncongested 
lakes,  each  no  more  than  a  mile  from  another.  The 
loss  of  one  lake  would  have  no  measurable  effect  on 
the  price  faced  by  an  angler  and  hence  no  loss  in 
consumer  surplus. 


794 


Economic  Analysis 


Net  Benefits  Versus  Resource  Quality 

Another  source  of  confusion  relates  to  differ- 
ences in  the  value  of  resources  of  different  quality. 
The  gross  willingness  to  pay  for  steelhead  fishing,  for 
example,  is  generally  higher  than  for  trout  fishing. 
However,  because  steelhead  fishing  opportunities  are 
often  available  only  at  a  few  remote  rivers,  travel 
costs  for  steelhead  fishing  tend  to  be  higher  than  for 
trout  fishing.  The  combined  effect  is  that  net  willing- 
ness to  pay  for  steelhead  fishing  may  be  lower  than 
that  for  trout  fishing.  In  these  situations,  lower  net 
benefit  flows  from  the  higher  quality  resource  be- 
cause the  higher  price  offsets  the  quality  difference. 
On  the  other  hand,  introduction  of  a  new  steelhead 
site  that  requires  travel  costs  similar  to  those  at  trout 
fishing  sites  will  produce  more  net  benefit  than  a 
new  or  existing  trout  site. 

A  parallel  situation  exists  for  timber.  For  exam- 
ple, walnut  trees  are  much  more  valuable  than  pine 
trees,  but  a  remote  stand  of  walnut  trees  in  an  inac- 
cessible location  will  have  a  lower  bid  price  than  a 
stand  of  pine  trees  near  a  mill.  In  this  example,  the 
higher  quality  trees  are  worth  less  because  of  higher 
harvest  costs. 


SUMMARY  AND  CONCLUSIONS 

As  is  evident  from  this  chapter,  several  bio-eco- 
nomic analysis  systems  provide  the  capability  to 
translate  the  benefits  of  habitat  improvements  into 
dollar  terms.  Besides  fulfilling  legal  requirements  to 


make  such  conversions,  the  benefits  can  be  com- 
pared to  the  costs  of  making  a  given  habitat 
improvement. 

Because  the  traditional  valuation  techniques 
only  measure  recreation,  recreation  benefits  cannot 
appropriately  be  interpreted  as  the  total  value  of 
animals  or  their  habitat.  Rather,  net  benefits  (present 
net  worth  or  internal  rate  of  return)  provide  infor- 
mation about  the  willingness  to  pay  of  recreationists 
for  the  proposed  habitat  improvements.  If  the  recrea- 
tion benefits  alone  exceed  the  cost,  the  project  will 
usually  increase  national  economic  well-being.  If  the 
benefits  are  less  than  the  cost,  the  decisionmaker 
must  ask  whether  the  project  substantially  increases 
intangible  economic  efficiency  benefits  (option,  ex- 
istence, bequest  values  for  a  wildlife  species  of  high 
public  interest)  or  substantially  improves  equity. 
Thus,  the  internal  rate  of  return  or  benefit-cost  ratio 
does  not  by  itself  make  the  decision  to  implement  or 
not  implement  a  management  action.  The  internal 
rate  of  return  or  benefit-cost  ratios  tell  the  decision- 
maker and  the  public  taxpayer  what  the  economic 
efficiency  of  such  investments  are. 

There  may  be  other  legitimate  social  objectives 
that  outweigh  economic  efficiency  in  determining 
whether  the  investment  is  to  be  made.  These  other 
objectives  should  be  documented  in  the  decision 
process  regardless  of  what  the  economic  efficiency 
analysis  shows.  In  this  way,  bio-economic  analysis 
and  economic  analysis  in  general  contribute  to  im- 
proved and  informed  decisions  rather  than  binding 
the  decisionmaker's  hands. 


Economic  Analysis 


795 


LITERATURE  CITED 


ALLEN,  A.  and  M.  ARMBRUSTER.  1982.  Preliminary  evalu- 
ation of  a  habitat  suitability  for  the  pronghorn  in 
McKenzie,  J.  ed.,  Proc.  10th  Pronghorn  Antelope 
Workshop. 

BISHOP,  R.  and  T.  HEBERLEIN.  1979.  Measuring  values  of 
extra  market  goods:  Are  indirect  measures  biased? 
Am.  J.  Agric.  Economics  6l(5):926-930. 

BROOKSHIRE,  D.,  A.  RANDALL,  and  J.  STOLE  1980.  Valu- 
ing increments  and  decrements  in  natural  resource 
service  flows.  Am.  J.  Agric.  Economics  63(3)165-177. 

,  M.  THAYER,  W.  SCHULZE,  and  R.  D'ARGE.  1982. 

Valuing  public  goods:  A  comparison  of  survey  and 
hedonic  approaches.  Am.  Economic  Review 
72:165-177. 

BROWN,  W.,  C.  SORHUS,  B.  CHOU-YANG,  and  J.  RICH- 
ARDS. 1983.  Using  individual  observations  to  estimate 
recreation  demand  functions:  A  caution.  Am.  J.  Agric. 
Economics  65(  1 ):  1 55- 1 57. 

CLAWSON,  M.  1959-  Methods  of  measuring  the  demand 
for  and  value  of  outdoor  recreation.  Resources  for  the 
Future.  Washington,  DC. 

DWYER,  J.J.  KELLY,  and  M.  BOWES.  1977.  Improved 

procedures  for  valuation  of  the  contribution  of  recre- 
ation to  national  economic  development.  Res.  Rep. 
1 28,  Water  Resources  Center,  Univ.  Illinois  at  Urbana. 

HARBERGER,  A.  1971.  Three  basic  postulates  for  applied 
welfare  economics:  An  interpretive  essay.  J.  Economic 
Literature  9(3):785-797. 

HYDE,  W.,  A.  DICKERMAN,  and  D.  STONE.  1982.  Devel- 
opment versus  preservation  in  the  Snake  River  Birds 
of  Prey  Conservation  Area.  Am.  J.  Agric.  Economics 
64:756-760. 

JUST,  R„  D.  HUETH,  and  A.  SCHMITZ.  1982.  Applied  wel- 
fare economics  and  public  policy.  Prentice  Hall,  En- 
glewood  Cliffs,  NJ. 

KRUTILLA,  J.V.  1967.  Conservation  reconsidered.  Am. 
Economic  Review  47:777-786. 

KUNKEL,  C.  and  PJANIK.  1976.  An  economic  evaluation 
of  salmonid  fisheries  attributable  to  Suislaw  National 
Forest.  Suislaw  National  Forest,  Corvallis,  OR. 

LOOMIS,  J.  1982.  Use  of  travel  cost  models  for  evaluating 
lottery-rationed  recreation:  Application  to  big  game 
hunting.  J.  Leisure  Res.  1 4(  2  ):1 17-124. 

and  R.  OLSON.  1981.  Recreational  and  commercial 

activities  involving  wildlife  potentially  affected  by 
the  Lower  Gunnison  River  Project.  U.S.  Dep.  Inter., 
Fish  and  Wildl.  Serv.,  Fort  Collins,  CO.  (unpubl.  rep.) 
-,  G.  PETERSON,  and  S.  SORG  1984.  A  field  guide 


to  wildlife  economic  analysis.  Trans.  North  Am.  Wildl. 
Nat.  Resour.  Conf.  49:315-324. 
MATULICK,  S.,  J.  HANSON,  I.  LINES,  and  A.  FARMER. 

1982.  HEP  as  a  planning  tool:  An  application  to  wa- 
terfowl enhancement.  Sci.  Pap.  6163,  Agric.  Res.  Cen- 
ter, College  of  Agriculture,  Washington  State 
I  Diversity.  Pullman. 


MISHAN,  E.  1976.  Cost-benefit  analysis.  Praeger  Publishers, 
New  York,  NY. 

OLIVER,  W.,  C.  YOUNG,  and  D.  ELDRED.  1975.  A  short 
form  for  bio-economic  evaluations  of  wildlife  in 
Washington  State.  Bull.  7.  Washington  Game  Dep. 
Olympia. 

PETERSON,  M.  1983.  Response  to  Assistant  Secretary 
Crowel  regarding  land  and  resource  management 
planning.  U.S.  Dep.  Agric,  For.  Serv.  Memorandum 
dated  May  17,  1983. 

ROSENTHAL,  D.J.  LOOMIS,  and  G.  PETERSON.  1984. 

The  travel  cost  model:  Concepts  and  applications.  U.S. 
Dep.  Agric,  For.  Serv.  Gen  Tech.  Rep.  RM-109-  Fort 
Collins,  CO. 

SASSONE,  P.  and  W.  SCHAFFER.  1978.  Cost-benefit  analy- 
sis: A  handbook.  Academic  Press,  New  York,  NY. 

U.S.  DEPARTMENT  OF  AGRICULTURE,  FOREST  SER 

VICE.  1982.  National  Forest  System  land  and  resource 
management  planning  in  Federal  Register,  September 
30,  1982,  Vol.  47(190):43,026-43,052. 

U.S.  DEPARTMENT  OF  THE  INTERIOR,  BUREAU  OF 
LAND  MANAGEMENT.  1979.  Biological  response  ap- 
proach. U.S.  Dep.  Inter.,  Bur.  Land  Manage.,  Oregon 
State  Office,  (unpubl.  rep.). 

.  1982.  Final  rangeland  improvement  policy.  In- 
struction Memorandum  83-27.  October  15,  1982. 
1983  Rangeland  investment  analysis  Bull.  UT-060- 


83-B6,  Moab  District  Office,  UT. 

U.S.  DEPARTMENT  OF  THE  INTERIOR,  FISH  and  WILD- 
LIFE SERVICE.  1980a.  Human  use  and  economic 
evaluation  (HUEE).  U.S.  Dep.  Inter.,  Fish  and  Wildl. 
Serv.,  Div.  Ecol.  Serv.  Manual  104:  Washington,  DC. 

.  1980b.  Habitat  evaluation  procedures  (HEP).  U.S. 

Dep.  Inter.,  Fish  and  Wildl.  Serv.,  Div.  Ecol.  Serv. 
Manual  102.  Washington,  DC. 

.  1981.  Standards  for  the  development  of  habitat 

suitability  index  models.  U.S.  Dep.  Inter.,  Fish  and 
Wildl.  Serv.,  Div.  Ecol.  Serv.  Manual  103.  Washington, 
DC. 

.  and  U.S.  DEPARTMENT  OF  COMMERCE,  BUREAU 


OF  CENSUS.  1982.  1980  National  survey  of  fishing, 
hunting,  and  wildlife  associated  recreation.  U.S.  Govt. 
Printing  Office.  Washington,  DC.  1 56pp. 

U.S.  DEPARTMENT  OF  TRANSPORTATION.  1985.  Cost 
of  owning  and  operating  a  motor  vehicle.  Federal 
Highway  Administration.  Washington,  DC. 

U.S.  WATER  RESOURCES  COUNCIL.  1979.  Procedures 
for  evaluation  of  national  economic  development 
(NED)  benefits  and  costs  in  water  resources  planning 
in  Federal  Register,  Vol.  44  (242)  Final  Rule.  Decem- 
ber 14,  1979. 

.  1983.  Economic  and  environmental  principles  and 

guidelines  for  water  and  related  land  resources  imple- 
mentation studies.  March  10,  1983.  U.S.  Govt.  Printing 
Office.  Washington,  DC. 

WALSH.  1986.  Recreation  economic  decisions.  Dep.  Agric. 
and  Resour.  Economics.  Colo.  State  Univ., 
Fort  Collins. 


796 


Economic  Analysis 


APPENDIX  I.   BLM  Moab  District  Hunter  Day  Short  Form 


The  BLM  Moab  District  Hunter  Day  Short  Form 
(U.S.  Department  of  the  Interior,  Bureau  of  Land 
Management  1983)  was  developed  to  allow  for  a  sys- 
tematic but  rapid  evaluation  of  the  change  in  big 
game  hunter  days  associated  with  changes  in  live- 
stock use  levels.  The  first  page  of  the  form  provides 
a  simple  word  model  of  the  changes  in  life  requisites 
or  habitat  components  in  relation  to  habitat  suitabil- 
ity for  a  particular  big  game  species  (Figure  A-I-l ). 
The  word  model  keys  in  on  the  critical  factors  likely 
to  be  affected  by  changes  in  livestock  uses. 

The  current  model  structure  provides  some 
information  on  limiting  habitat  factors  but  the  impor- 
tance of  that  limiting  factor  is  not  explicitly  recog- 
nized. That  is,  the  model  structure  assumes  that  lack 
of  water  can  be  offset  by  greater  cover,  space,  or 
forage.  This  assumption  may  be  tenuous  at  times, 
and  the  information  on  the  limiting  factor  can  be 
helpful  to  biologists  in  designing  habitat  improve- 
ments. If  water  is  the  limiting  factor,  then  guzzlers 
would  provide  the  biggest  payoff,  not  more  cover. 
This  basic  model  could  easily  be  modified  to  make 
the  limiting  factor  concept  explicit.  Techniques  for 
incorporating  limiting  factor  concepts  into  indexes 
are  described  in  "Standards  for  the  Development 
of  Habitat  Suitability  Index  Models"  (U.S.  Depart- 
ment of  the  Interior,  Fish  and  Wildlife  Service  1981). 

The  second  page  of  the  Moab  District  Short 
Form  requires  calculating  hunter  days  from  popula- 
tion and  application  of  the  index  factor  (Figure  A- 
1-1).  The  hunter  day /population  estimates  provided 
at  the  top  of  the  table  are  for  southeastern  Utah. 
These  numbers  may  seem  counter-intuitive  because 
they  combine  percentage  of  harvestable  population 
and  days  needed  to  harvest  one  animal. 


The  biologist  must  estimate  the  existing  popula- 
tion level.  In  addition,  for  habitat  improvement  proj- 
ects, the  prior  stable  or  potential  population  must 
be  estimated.  If  such  estimates  are  available,  this 
approach  is  probably  preferable.  However,  existing 
population  data  and  potential  population  estimates 
(prior  stable)  may  not  be  available.  In  these  cases, 
harvest  estimates  must  be  obtained.  State  wildlife 
management  agencies  generally  have  existing  harvest 
data.  Estimating  potential  harvest  is  no  more  difficult 
than  estimating  potential  population,  given  optimum 
habitat  conditions.  In  addition,  using  harvest  allows 
use  of  hunter  days  per  animal  harvested,  a  number 
that  is  also  more  readily  available.  A  Hunter  Day 
Short  Form  for  use  with  harvest  data  is  also  available 
(Figure  A-I-2).  Page  1  of  the  form  is  identical  to  that 
shown  on  Figure  A-I-l,  but  the  calculations  on  page 
2  are  based  on  hunter  days  per  harvest  rather  than 
hunter  days  per  population.  Detailed  instructions  for 
using  the  Moab  District  Hunter  Day  Short  Form  can 
be  found  in  BLM's  Moab  District  Bulletin  UT-060-83- 
86  (U.S.  Department  of  the  Interior,  Bureau  of  Land 
Management  1983). 

In  summary,  the  Hunter  Day  Short  Form  pro- 
vides a  very  easy  to  use  and  yet  systematic  approach 
for  evaluating  biological-economic  effects  for  big 
game  animals.  By  using  some  of  the  modifications 
suggested  in  this  paper,  the  short  form  analysis  can 
be  upgraded  when  sufficient  data  exist  on  the  impor- 
tance of  one  habitat  type  as  a  limiting  factor.  The 
form  can  be  modified  to  use  harvest  data  when  pop- 
ulation data  are  unavailable.  The  basic  advantages 
of  the  Moab  District  Hunter  Day  Short  Form  ap- 
proach are  minimal  data  requirements  and  speedy 
analysis.  It  is  a  very  useful  approach  when  one  has 
dozens  of  areas  to  evaluate  in  a  short  time. 


Economic  Analysis 


797 


Figure  A-I-l.     Rangeland  investment  analysis — based  on  population. 


HUMAN  USE  AND  ECONOMIC   EVALUATION 

104   ESM  A 

Appendix  A.     Bio-Economic  Analysis  of  W 

ildlife 

Table  A-2.     Rangel< 

Hu 

X      Hu 
of 

Allotment         4"| 
Wildlife  Spcoes 

Cateoory             Rank 

Forage                       3 
Compe t  i  t  ion 

1.2 

0 

-1,-2 

-3 

Forage                        3 

1,2 

0 
-1.-2 

-3 

Cover                 1,2,3 

0 

-1.-2.-3 

Water                 1,2,3 
0 
-1.-2.-3 

Space                 1,2,3 

0 

-1.-2.-3 

Tot  al    Points  /    (numt 

10    I  (     3    » 
/  (           > 

Release  No.   1-85 

ind  investment  analysis  -  based  on  population. 

nter  Day  Estimates  without   Investment   package 
nter   Day  Estimates   with    investment   package 
Habitat   changes   through   time 

Alternative             1 

TifER                                       Wildlife  Biologist 

Criteria                                             C 

Forage  consumption  conflicts   totally  or 
nearly  el iminated. 

Forage  consumption  conflicts   reduced. 

Ho   change. 

Forage  consumption  conflicts    increased. 

Forage   consumption  conflicts    become    a 
major   problem. 

Key  browse,    forb,    and   grass    species 
Increase    in   vigor    and   trend    by  more    than 
30  percent. 

Either/or  browse,    forbs,    and  grass   increase 
In  vigor   and   trend  by  less   than  30  percent. 

No  change. 

Either/or  browse,   forbs,   and  grass   decrease 
In  vigor   and  trend  by   less   than  30  percent. 

Key  browse,   forb,    and   grass   species   decrease 
1n   vigor   and   trend  by  more   than  30  percent. 

Cover  availability  Increases. 

No  change. 

Cover  availability  decreases. 

Water  availability  increases. 

No  change. 

Water   availability  decreases. 

Spatial   conflicts   decreases. 

No  Change. 

Spatial  conflicts   Increases. 

Total  Points 

>er  of  categories   considered   «  3)    ■  Adjustment 

CJTM  3/W  IIP, 

'oints        Points 
'lanae   1   Chanae  2 

2~ 

3 

1 

2- 

2- 

10 

Factor 

August  1985 

.  3)   "   |        .fo7      |   Adjustment  Factor  Change   1 

3)    •    |                     |    Adjustment   Factor   Change   2 

104-ESM-A-ll 

798 


Economic  Analysis 


Figure  A-I-l.     Rangeland  investment  analysis — based  on  population  (concluded). 


HUMAN  USE  AND  ECONOMIC  EVALUATION 

104   ESM  A 

Appendix  A 

.     Bio-Economic  Analysis  of  Wildlife 

Table  A-2.     Continued. 

Hunter  Day  /  population   (HD/pop)   estimates 

Deer                                      2.0  HD/pop 
Elk                                          4.0  HD/pop 
Antelope                               0.2  HD/pop 
Desert  Bighorn  Sheep       0.2  HD/pop 

Base   Year: 

months   o 

n  allotme 

X 

nt   f   12  -  x 

IOO 
Z-o 

ZOO 

■  ZS~ 

Existing   P 
HD/pop 
HD 
Length   of 

opul at  ion 

stay   adjustment 

Habitat  Worsening                    | 

X 

♦ 

1 
1 

Habitat    Improvement 

Z.OO       Prior  Stable  for 
IOO          Existinq   Pop   ■ 

X 

X 

_  Months   on 

Al  lotment    .-    12   ■            x 

5o 

Base  year  HD   input 

'hange  1 

•a 

SM 

Change 

X 

♦ 

2 

Base  Year  HO 
Adjustment    factor 
HD  Loss 
Base  Year  HO 

1 

1 

1 

HO   Input 

1 

1 

1 

Years    to  Change 

General  An 

IOO 
•  fc7 

k>l 

2. 

I34HD 

.2S 

33HD 
,"50  HD 

X 
X 

X 

♦ 

Potential    Change 
Adjustment   factor 
Popul at  ion  Change 
HD   /   pop 
HD 

Time   Adjustment 
HD  Change 
Base  Year  HD 

|    HD      Input 

|    Yearj   to  Change 

1 

66    1 

1 

1 

104-E 

1 

Release  No.   1-85 

-A-12 

August  1985 

Economic  Analysis 


799 


Figure  A-I-2.     Rangeland  investment  analysis — based  on  harvest. 


HUMAN  USE  AND  ECONOMIC  EVALUATION 

104  ESM  A 

Appendix  A.  Bio-Economic  Analysis  of  Wildlife 

Table  A-3.  Range' 

Hu 

of 

Allotment     j 
Wildlife  Species 

Cateoory      Rank 

rorage         3 
Compet i t  ion 

1.2 

0 

-1.-2 

-3 

Forage         3 

1.2 

0 
-1.-2 

-3 

Cover       1,2,3 

0 

-1.-2.-3 

Water       1,2,3 
0 
-1,-2,-3 

Space       1,2,3 

0 

-1.-2.-3 

Total  Points  /  (numb 
10        1   (         5      x 

1   (       x 

Release  No.  1-85 

and  investment  analysis  -  based  on  harvest. 

nter  Day  Estimates  without  investment  package 
nter  Oay  Estimates  with  investment  package 
Habitat  changes  through  time 

Alternative      / 

OtrBR                                        Wildlife  Biologist 

CTIM  .^ITU 

Points   Points 
lanqe  1  Chanqe  2 

Criteria                   C 

Forage  consumption  conflicts  totally  or 
nearly  el lminated. 

Forage  consumption  conflicts  reduced. 

No  change. 

Forage  consumption  conflicts  increased. 

Forage  consumption  conflicts  become  a 
major  problem. 

Key  browse,  forb,  and  grass  species 
increase  in  vigor  and  trend  by  more  than 
30  percent. 

Either/or  browse,  forbs,  and  grass  increase 
In  vigor  and  trend  by  less  than  30  percent. 

No  change. 

Either/or  browse,  forbs,  and  grass  decrease 
In  vigor  and  trend  by  less  than  30  percent. 

Key  browse,  forb,  and  grass  species  decrease 
in  vigor  and  trend  by  more  than  30  percent. 

Cover  availability  increases. 
No  change. 

2. 

3 

1 

Cover  availability  decreases. 

Water  availability  increases. 
No  change. 

z 

Water  availability  decreases. 

Spatial  conflicts  decreases. 
No  Change. 

z 

Spatial  conflicts  increases. 

Total  Points 
er  of  categories  considered  x  3)  »  Adjustment 

10 

Factor 

August  1985 

3)  •  |  .lol         |  Adjustment  Factor  Change  1 

3)  •  |         |  Adjustment  Factor  Change  2 

104-ESM-A-14 

HOO 


Economic  Analysis 


Figure  A-I-2.     Rangeland  investment  analysis — based  on  harvest  (concluded). 


HUMAN  USE  AND  ECONOMIC  EVALUATION 


104  ESM  A 


Appendix  A.  Bio-Economic  Analysis  of  Wildlife 


Table  A-3.  Continued. 


Hun 

ter  Days/harvest   (HD/h 

arves 

t)  estimates 

for 

Utah 

Deer 

8.0 

HDAiarvest 

Elk 

11.3 

HD/harvest 

Antelope 

3.0 

HDAiarvest 

Desert  Bi 

ghorn 

Sheep 

13.0 

HD,"harvest 

from 

Utah  Big 

Game 

Harvest 

Book 

Base  Year: 


£5"  Existing   Harvest    (or   avg.    of 

a  .  3  ye"*) 

O  HD/harvest 


2O0         HD 


months  on  allotment  f  12  ■  x 


.2.5    Length  of  stay  adjustment 


50         Base  year  HD   input 


Habitat  Worsening 


Change  1 


Change  2 


Base  rear  HD 
Adjustment  factor 
HD  Loss 
Base  Year  HD 

~|  HD  Input 

|  Years  to  Change 


Habitat  Improvement 


50      Prior  Stable 

or 

Potential 

Long 

Run 

Harvest 

25^      ExistinqHarvest 

■    zf 

Potential   Change 

X 

.(of 
/4>.75 

X 

Adjustment   Factor 
Population  Cringe 

X 

& 

X 

HD  /harvest 

3  Months   on 

\y\ 

HD 

Al  lotment    .-    12   « 

X 

.15 

33 

X 

Time   Adjustment 
HD  Change 

♦ 

1 

1 

h 

♦ 

Base   Year  HD 

8&.I 

1 

1 

HD      Input 

10  1 

1 

1 

Years    to  Change 

Release  No.   1-85 


104-ESM-A-15 


August  1985 


Economic  Analysis 


801 


APPENDIX  II.     Human  Use  and  Economic  Evaluation  (HUEE)  System 


The  Human  Use  and  Economic  Evaluation 
( HUEE )  system  was  developed  by  the  U.S.  Fish  and 
Wildlife  Service  Western  Energy  and  Land  Use  Team 
(WELUT).  Human  Use  and  Economic  Evaluation 
(HUEE)  procedures  provide  means  for  determining 
both  the  extent  of  human  uses  of  wildlife  and  the 
dollar  values  of  these  uses.  These  procedures  were 
developed  and  are  intended  for  use  with  the  Habitat 
Evaluation  Procedures  (HEP).  The  HEP  and  HUEE 
together  with  the  Habitat  Suitability  Index  (HSI) 
models  (U.S.  Department  of  the  Interior,  Fish  and 
Wildlife  Service  1980a,  1980b,  1981 )  provide  a  com- 
plete set  of  procedures  for  field  staff  to  use  in  mak- 
ing both  biological  and  economic  assessments  of 
wildlife  resources. 


HUEE  procedures  can  be  used  in  conjunction 
with  HEP  or  with  just  population  or  harvest  data.  In 
addition,  HUEE  is  compatible  with  the  Travel  Cost 
Method  (TCM)  and  Contingent  Value  Method 
(CVM)  valuation.  HUEE  can  also  use  unit  day  values 
when  the  added  costs  of  TCM  or  CVM  cannot  be 
justified. 


The  HUEE  procedures  are  designed  for  use  by 
field  staff,  principally  biologists,  assigned  to  evaluate 
the  impacts  of  resource  development  projects.  These 
procedures  may  be  applied  in  field  studies  without 
the  assistance  of  economists  or  recreation  planners. 
However,  to  apply  advanced  methods  such  as  the 
TCM  or  CVM,  the  assistance  of  a  specialist,  such  as 
an  economist  or  recreation  planner  is  recommended 
(U.S.  Department  of  the  Interior,  Fish  and  Wildlife 
Service  1980a). 


HUEE  provides  a  way  to  evaluate  both  the  ef- 
fects on  supply  of  hunter  days  and  effects  on  the 
demand  for  hunter  days  (Figure  A-II-1 ).  The  term 
"potential  use"  in  Figure  All- 1  refers  to  the  amount 
of  use  people  wish  to  make  of  the  wildlife  resource. 
The  term  "sustainable  use"  is  the  amount  of  use  days 
that  can  be  provided  by  the  habitat.  "Planned  use" 
is  the  lesser  of  the  sustainable  and  potential  use. 
Thus,  HUEE  is  general  enough  to  account  for  both 
situations  where  demand  exceeds  supply  at  the  cur- 
rent license  cost  and  also  where  supply  of  animals 
exceeds  the  demand  for  hunting.  Figure  A-II-2  shows 
how  supply  and  demand  over  time  are  related  to 
HUEE's  concepts  of  sustainable  use,  potential  use, 
and  planned  use.  In  this  example,  the  initial  factor 
determining  hunter  days  is  human  demand.  By  year 
20,  demand  has  risen  and  supply  fallen  such  that 
hunter  days  demanded  and  supplied  arc  equal.  After 
year  20,  demand  (at  current  permit  fees)  exceeds 
supply  so  that  lottery  rationing  or  shorter  seasons 
are  likely  to  keep  hunter  pressure  in  line  with  avail- 
able populations. 


HEP 

\ 

Change  in 

Habitat   Units 

(HUs) 

ir 

Future 
Population 
Estimates 

ir 

Sustainable 
Use 

Potential 
Use 

Travel  Cost 
Method 
(TCM) 

\ 

Si 

/ 

/ 

/ 

f 

Planned 
Use 

Economic 

Value  of 

Use 

<r      jt 

% 

\ 

Economic  Value 

of  Change 

in   Use 

Contingent 
Value 
Method 
(CVM) 

Figure  A-II-1.     Human  use  and  economic 
evaluation  (HUEE)  system. 


Figure  A-II-3  illustrates  a  negative  impact  to  a 
species'  population  from  a  project.  Loss  of  habitat 
generally  does  not  change  the  demand  or  potential 
use  by  humans  but  rather  adversely  affects  the  eco- 
system's capability  to  support  days  of  hunting  or 
fishing.  Thus,  the  effects  are  modeled  as  a  reduction 
in  sustainable  use.  Alternatively,  enhancement  of 
habitat  suitability  by  management  can  increase  sus- 
tainable use. 

This  conceptual  framework  is  fairly  complete, 
but  to  be  useful,  it  must  be  easy  to  translate  into 
practice.  The  two  factors  influencing  the  actual 
amount  of  wildlife  recreation  that  can  be  realized  are 
potential  use  and  sustainable  use. 

Potential  use  can  be  estimated  in  several  ways. 
A  biologist  can  do  a  time  series  regression  of  past 
hunter/angler  day  levels  in  that  management  unit. 
This  equation  can  then  be  used  to  forecast  future 
wildlife  recreation  demand,  assuming  current  trends 
continue.  It  is  important  to  remember  that  these 
forecasted  levels  of  demand  will  not  be  translated 
into  actual  use  unless  the  wildlife  habitat  can  sup- 
port such  levels  of  use. 


S02 


Economic  Analysis 


400  T    SUSTAINABLE   USE 


POTENTIAL  USE 


10  20  30  40 

TIME(YEARS) 


Potential  use 


Sustainable  use 
^without  project 


IMPACT  OF  PROJECT  „ 


TIME 


Figure  A-II-2.     Relationships  between  sustainable 
use,  potential  use,  and  planned  use. 


Figure  A-II-3.     Negative  economic  impact  of  a 
project  from  reduction  in  sustainable  use. 


An  alternative  and  better  way  to  forecast  future 
use  is  to  use  TCM.  This  technique  estimates  the  eco- 
nomic benefits  per  day  and  allows  use  of  forecasting 
methods  in  valuing  wildlife  use. 

Once  the  number  of  wildlife  recreation  days 
demanded  (potential  use)  are  known,  one  needs  to 
develop  the  amount  of  use  the  wildlife  habitat  can 
support  or  supply  on  a  sustainable  basis  (sustainable 
use).  The  steps  in  calculating  sustainable  use  are 
shown  in  Figure  A-II-4.  The  end  result  of  these  calcu- 
lations are  ( 1 )  sustainable  use  days  of  fishing  or 
hunting  or  (2)  nonconsumptive  use  days  associated 
with  a  given  alternative  plan. 

Each  species  or  group  of  species  (e.g.,  water- 
fowl, upland  game  birds,  etc. )  must  be  calculated 
separately.  The  calculation  can  begin  at  Point  1 
(Habitat  Units  [HUs]),  Point  3  (animal  population 
size),  or  Point  5  (harvest).  The  starting  point  is  the 
biologist's  decision  and  is  generally  constrained  by 
the  data  available  on  HUs,  population  size,  or  har- 
vest. If  an  HSI  model  for  the  species  being  evaluated 
is  available  along  with  either  current  population  or 
current  harvest,  it  is  fairly  easy  to  estimate  future  use 
days. 

If  suitable  HSI  models  are  not  available,  then  the 
future  population  size  or  harvest  rate  with  and  with- 
out the  project  must  be  estimated  by  some  other 
method  in  order  to  estimate  future  use  davs. 


HUs  are  the  basic  units  of  analysis  in  the  U.S. 
Fish  and  Wildlife  Service  HEP.  HEP  and  the  use  of 
HSI  models  are  described  in  other  chapters  in  this 
book  and  in  U.S.  Department  of  the  Interior,  Fish  and 
Wildlife  Service  ( 1980b,  1981 ).  An  HSI  is  a  number 
from  0  to  1  that  serves  as  an  index  to  the  quality 
of  the  habitat  in  an  area  for  a  given  species.  An  index 
of  1  represents  "optimal"  habitat  and  an  index  of  0 
represents  totally  unsuitable  habitat.  For  terrestrial 
species,  HUs  are  the  product  of  the  HSI  and  the 
acreage  for  a  given  site. 

HSI  x  Acreage  =  Habitat  Units 

example:  0.60  x   1,000  acres  =  600  habitat  units. 

Animal  numbers  are  calculated  as  the  product  of 
HUs  and  number  of  animals  per  habitat  unit  (produc- 
tivity per  HU  in  Figure  A-II-4).  This  calculation  as- 
sumes that  animal  population  size  is  linearly 
correlated  with  both  habitat  suitability  (as  measured 
by  the  HSI)  and  acreage.  If  both  values  (HUs  and 
animals  per  HU)  are  available  then  the  calculation  is 
straightforward. 

Typically  the  latter  figure  (animals  per  HU) 
must  be  estimated  or  calculated  from  other  data.  If 
current  populations  are  known,  the  animals  per  HU 
can  be  calculated  by  dividing  current  population  (or 
the  average  population  for  the  last  3-5  years)  by 
current  HUs.  This  yields  animals  per  HU  that  can 


Economic  Analysis 


803 


Percent    Catchable 

Size 

(Fish  Only) 

Catchable 

Crop 
(Fish   Only) 

Habitat 
Units 
(HUs) 

Annual 

Harvest 

(Fish) 

Annual 

Harvest 

(Wildlife) 

— ► 

Population 
Size 

V"" 

Productivity 
per  HU 

n 

Harvest 
Rate 

A 

-fe 

=     Annual 

Sustainable 

Use  Days 

Use   Days 
per  Animal 
Harvested 

Figure  A-II-4.     Steps  for  calculating  sustainable  use. 


then  be  multiplied  by  HUs  in  future  target  years  to 
calculate  estimated  animal  numbers  under  alternative 
management. 

Next,  harvest  must  be  calculated.  For  terrestrial 
species,  the  population  size  is  multiplied  by  a  sustain- 
able or  allowable  harvest  rate.  This  is  a  harvest  rate 
that  can  be  sustained  without  detriment  to  the  popu- 
lation. For  fish,  the  calculation  involves  the  addi- 
tional steps  of  calculating  percentage  of  population 
of  catchable  size  from  which  a  catchable  crop  is 
calculated. 

The  next  step  is  to  estimate  hunter  or  angler 
days  per  harvest  so  that  sustainable  use  days  can  be 
calculated.  Hunter  days  per  harvest  can  be  calculated 
from  data  on  success  rates  and  number  of  days  in 
the  field.  Otherwise,  it  must  be  estimated  or  extrapo- 
lated from  other  areas.  A  decrease  in  the  number  of 
animals  available  for  harvest  could  cause  success 
rates  to  drop,  but  use  days  would  remain  the  same. 
That  is,  the  same  number  of  hunters  continue  to 
hunt  even  though  they  must  hunt  more  days  for 
each  animal  harvested.  If  this  is  an  accurate  descrip- 


tion of  hunter  behavior  in  the  study  area,  then  HUEE 
can  reflect  this.  This  constant  level  of  lower  and 
lower  quality  hunter  days  will  be  reflected  as  a  drop 
in  economic  benefits  because  as  harvest  per  day 
decreases,  benefits  per  day  generally  decrease.  Thus, 
one  advantage  of  HUEE  over  other  systems  is  its 
ability  to  estimate  changes  in  economic  values  even 
if  the  number  of  days  hunted  or  fished  remains  un- 
changed. 

The  sustainable  use  days  are  then  compared  to 
potential  use  and  the  lessor  of  potential  use  and 
sustainable  use  is  recorded  as  "Planned  Use"  (Figure 
A-II-2).  The  planned  use  is  then  multiplied  by  value 
per  day  to  obtain  the  economic  value  of  such  recrea- 
tion. Beyond  this  point,  HUEE  provides  a  format  and 
step-by-step  procedure  and  accounting  system  for 
projecting  these  values  into  the  future  and  making 
various  types  of  corrections  and  adjustments. 

More  information  on  the  details  of  HUEE  can  be 
found  in  the  U.S.  Fish  and  Wildlife  Service  Manual 
(U.S.  Department  of  the  Interior,  Fish  and  Wildlife 
Service  1980a). 


804 


Economic  Analysis 


41 

WRITTEN 
COMMUNICATIONS 


Donald  Zimmerman 

Department  of  Technical  Journalism 
Colorado  State  University 
Ft.  Collins,  CO  80521 


"The  greatest  problem  in  communication  is  the 
illusion  that  it  has  been  achieved." 


-George  Bernard  Shaw 


Editor's  Note:  No  inventory  or  monitoring  study  is 
complete  until  results  are  communicated  to  the 
people  who  need  to  see  them.  Many  of  us  allot  little 
time  to  uniting  reports  or  making  verbal  presenta- 
tions— but  these  activities  are  vital  if  management 
and  the  public  are  to  understand  our  work  and 
make  better  natural  resource  decisions. 

This  chapter  provides  methods  to  identify  and  ac- 
complish your  objectives  for  written  communica- 
tions. The  chapter  on  verbal  presentations  does  the 
same  for  oral  communications,  because  speaking 
to  groups  is  just  as  important  in  getting  your  mes- 
sage across. 


INTRODUCTION 

Once  you  finish  your  data  analysis  and  interpre- 
tation, you  will  write  the  needed  memos,  letters,  re- 
ports, and  other  documents.  You  must  write.  It  is 
part  of  your  job. 


\  MUST  WRITS  ? 


Written  Communication 


805 


Good  writing  becomes  important  for  two  rea- 
sons. First,  writing  that  communicates  a  project's  re- 
sults may  be  as  important  as  the  study  itself.  If  you 
do  not  communicate  your  study's  results  in  writing 
so  others  understand  the  findings  and  know  the 
needed  actions  and  management  decisions,  the  re- 
source may  not  be  managed  properly.  By  writing, 
you  document  and  explain  the  action  needed. 

Second,  good  writing  becomes  important  for 
you  because  your  career  depends  on  it.  If  you  do 
not  communicate  with  your  peers,  supervisors,  and 
administrators,  they  will  not  know  your  abilities  and 
skills.  Successful  resource  managers  communicate 
through  writing  and  they  become  known  for  making 
themselves  understood.  Good  writing  advances  your 
career. 


When  you  think  about  writing,  what  ideas 
emerge?  Why  do  you  write?  What  do  you  want  to 
accomplish?  Do  you  think  about  your  readers? 
Do  you  plan  ahead  or  do  you  wait  until  the  last 
minute  to  write? 


There  is  more  to  consider:  What  constitutes 
writing?  How  do  you  organize  your  thinking?  Do 
you  outline?  Do  you  rewrite?  Do  you  find  writing 
difficult  or  easy?  Do  you  avoid  writing? 


In  this  chapter  I  present  a  1 2-step  writing  pro- 
cess for  preparing  longer  reports: 

1 .  Thinking  and  planning 

2.  Organizing  your  writing 

3.  Preparing  illustrations 

4.  Drafting  the  manuscript 

5.  Writing  abstracts  and  executive  summaries 

6.  Revising  and  self-editing 

7.  Editing  for  style  and  grammar  errors 

8.  Correcting  spelling  errors 

9.  Editing  illustrations — tables  and  figures 

10.  Having  the  manuscript  retyped 

1 1 .  Copyediting  the  manuscript 

12.  Seeking  peer  reviews. 

As  you  read  through  the  chapter,  remember  that 
some  documents  do  not  have  illustrations  and  others 
do  not  require  peer  reviews.  Furthermore,  no  two 
professionals  write  in  exactly  the  same  way.  Some  in- 
dividuals may  combine  steps  whereas  others  may 
break  the  identified  steps  into  several  activities.  From 
the  1 2-step  writing  process,  adopt  and  adapt  tech- 
niques to  improve  your  writing  process. 

I  close  the  chapter  with  a  brief  discussion  of 
printing,  the  Information  Age,  and  computers'  im- 
pacts on  writing  and  communication. 


A  WRITING  STRATEGY 


Before  I  discuss  writing  further,  consider  how 
others  judge  your  writing.  They  judge  your  writing 
on  its  content,  its  appearance,  and  its  communica- 
tion effectiveness.  The  content  centers  on  how  well 
you  carried  out  your  professional  problem  solving. 
This  book  will  help  you  improve  your  content.  Ap- 
pearance centers  on  style,  mechanics,  spelling,  typ- 
ing, layout,  and  printing.  Later,  I  suggest  style  man- 
uals and  guides  and  techniques  for  checking  your 
writing's  appearance.  Communication  effectiveness 
centers  on  your  objectives — do  you  want  to  inform, 
instruct,  or  persuade  your  readers?  Will  your  readers 
understand  the  points  you  are  making?  Will  your 
readers  agree  with  you?  Will  your  readers  accurately 
perceive  your  content? 

Consciously  or  not,  many  professionals  ask 
themselves  such  questions  as  they  write.  By  asking 
yourself  such  questions,  you  will  begin  thinking 
about  your  writing  and  answering  the  questions  that 
will  help  you  communicate  more  effectively. 

To  achieve  your  objectives,  your  writing  must 
be  clear  and  concise.  Therefore,  I  suggest  a  writing 
strategy  and  include  ways  to  improve  your  writing's 
objectives. 


Thinking  and  Planning 

Writing  begins  when  you  start  thinking  about 
what  you  will  write,  your  illustrations,  and  your  con- 
tent. Effective  writing  requires  thinking  ahead — 
weeks,  months,  or  years  for  larger  projects.  When 
you  begin  thinking  about  writing,  ask:  Who  are  my 
readers?  What  do  I  want  to  communicate  to  my 
readers?  What  is  the  appropriate  written  format? 
How  long  will  my  writing  take? 

"Who  Are  My  Readers?" 

As  a  biologist,  you  have  diverse  readers.  Think 
about  who  reads  your  writing.  Reader-oriented  writ- 
ing may  change  information  levels,  behaviors,  and 
possibly  attitudes.  But  too  often  we  fail  to  keep  our 
readers  clearly  in  mind.  Within  the  Bureau  of  Land 
Management  and  other  government  agencies  you 
have  two  major  groups  of  internal  readers:  ( 1 )  tech- 
nical readers — i.e.,  fellow  scientists,  and  (2)  man- 
agers— i.e.,  the  decisionmakers. 

Depending  on  the  content  of  your  writing,  you 
will  have  different  external  readers:  ( 1 )  biologists 
and  managers  from  other  agencies;  (2)  judges,  law- 
yers, congressmen,  and  senators;  and  (3)  the  general 
public — including  special  interest  and  advocacy 
groups. 


806 


Written  Communication 


Consider  the  functions  of  your  writing.  Most 
technical  and  scientific  writing  informs,  instructs, 
persuades,  and  documents.  Informing  requires  at- 
tracting the  reader's  attention,  having  the  message 
accepted,  having  it  interpreted,  and  having  it  stored 
for  later  use  (Schramm  and  Roberts  1972).  Instruct- 
ing requires  the  same  steps  as  informing,  plus  the 
reader  must  do  something  with  the  message — i.e., 
practice  or  use  the  information  provided  (Schramm 
and  Roberts  1972).  Persuading  requires  convincing 
the  readers  of  your  position;  it  requires  that  readers 
yield  to  your  viewpoint.  Documenting  provides  a 
permanent  record.  Most  technical  reports,  notes, 
journal  articles,  memos,  and  letters  document  a  pro- 
ject, a  finding,  or  a  decision. 

Think  about  the  outcomes  you  want  from  your 
writing.  Do  you  only  want  readers  to  understand?  Or 
do  you  want  readers  to  agree  with  you?  Do  you 
want  the  readers  to  accurately  perceive  your  facts? 
Communication  may  improve  understanding,  but  de- 
crease agreement.  Your  readers  may  accurately  per- 
ceive your  facts  and  your  recommendations,  but 
they  may  not  agree  with  you. 

When  you  write  for  managers,  keep  in  mind  that 
they  like  to  have  several  options  and  the  implica- 
tions of  each  option  clearly  spelled  out.  If  a  manager 
decides  on  option  C,  what  are  the  likely  conse- 
quences of  that  decision?  What  impact  will  option  C 
have  on  the  resources?  What  impact  will  option  C 
have  on  the  agency? 


WRITING  "SHOULD" BEGltf 
W'TH  THINKING. 


'••  • 


Consider  wild  burros.  From  a  management  view- 
point, too  many  burros  may  destroy  a  range  or  re- 
duce its  carrying  capacity  for  livestock  and  wildlife. 
So  you  recommend  reducing  burro  numbers.  From  a 
budgetary  standpoint,  shooting  them  may  be  the 
least  expensive  control  technique.  But  such  a  recom- 
mendation would  soon  cause  an  uproar.  Groups  will 
clearly  understand  your  recommendation,  but  many 
will  not  agree  with  you. 

When  you  think  of  your  readers,  think  about 
how  they  will  use  the  information  you  provide.  Will 
the  information  be  for  decisionmaking?  If  so,  what 
kinds  of  decisions  will  your  readers  be  making?  How 
will  your  readers  use  the  information?  What  informa- 
tion will  you  need  to  provide  so  your  readers  make 
a  decision  based  on  sound  management  principles? 
What  facts  must  you  provide?  How  should  you  or- 
ganize those  facts? 

How  do  your  readers  read  what  you  write? 
Communication  research  consistently  shows  that 
most  people  read  only  what  interests  them  or  they 
read  only  what  they  think  will  be  of  value  to  them. 
Readers  are  selective.  And  your  readers  will  not  read 
everything  that  you  write.  To  find  out  how  people 
read  what  you  write,  produce  a  document  and  dis- 
tribute it.  Then  ask  the  people  what  they  read.  Ask 
them  what  they  learned.  Their  answers  will  help  you 
tailor  your  subsequent  writing  to  their  needs. 

When  people  read  a  document,  they  do  not  re- 
tain all  the  information.  Too  often  what  they  retain 
and  recall  are  only  the  points  and  information  that 
reinforce  their  viewpoints. 

Beyond  how  readers  read,  consider  their  frames 
of  reference — the  sum  of  their  background,  educa- 
tion, and  experience.  Too  often  we  assume  our  read- 
ers see  the  world  as  we  do.  We  assume  our  readers 
understand  the  terms  we  use.  We  assume  they  com- 
prehend what  we  are  saying.  Often  they  do  not.  At 
best,  the  commonality  of  the  frames  of  reference  of  a 
writer  and  reader  remain  imperfect,  incomplete,  and 
partial. 

"What  Information  Do  You  Communicate?" 

With  your  readers  in  mind,  what  content  will 
you  present?  How  much  information  should  you  pre- 
sent? What  information  will  you  leave  out?  Will  leav- 
ing out  some  information  create  problems  later? 
What  information  can  you  include?  Will  including 
some  information  create  problems? 

"What  is  the  Appropriate  Format?" 

Keeping  selectivity  in  mind,  choose  the  appro- 
priate format  to  quickly  give  readers  the  key  infor- 
mation. Table  1  identifies  readers  and  selects  formats 
and  then  suggests  lengths  and  purposes.  Whenever 
possible,  be  succinct.  Use  Table  1  as  a  general  guide. 


Written  Communication 


807 


Table  1.  Summary  of  target  audiences,  type  documents,  suggested  maximum  lengths,  and  purpose. 


Target 
audience 

Type 
document 

Maximum 
length  (pages) 

Purpose 

Subordinate 
biologist 

Memo 

Letter 

Manual 

Guidelines 

Procedures 

1 

1 

Varies 

Varies 

Varies 

Inform/Instruct 
Inform/Instruct 
Inform/Instruct 
Inform/Instruct 
Inform/Instruct 

Other 

biologists 

(General) 

Technical  notes 
Technical  reports 
Journal  articles 

5 

5 

Varies 

Inform/Document 
Inform/Document 
Inform/Document 

General 
public 

Letters 
Short  report 

2 
5 

Inform 
Inform 

Advocacy 
groups  or 
individuals 

Letter 

Short  report 
Technical  report 
EIS 

Executive  summary 
of  above 

2 

5 

5 

100 

5 

Document/Inform 
Document/Inform 
Document/Inform 
Document/Inform 
Document/Inform 

Supervisory 
biologists 
Supervisory 
Staff  (i.e.,  Chief 
of  Resources) 

Short  report 
Technical  report 

Memo 

Short  report 

Technical  report 

EIS 

EAR 

5 
5 

1-2 

5 

Varies 

100 

100 

Inform 
Inform 

Inform/Persuade 

Document 

Document 

Document 

Document 

Line  manager 

Memo 

Short  report  or 
Executive  summary 
of  EIS  or 
Technical  report 

1-2  with  back- 
up reports 
5 
5 

Inform/Persuade 

Inform/Persuade 
Inform/Persuade 

Congressional 
inquiry 
(Usually 
public  through 
Congressman) 

Memo 

with  accompanying 

reports,  etc.  as 

documented 

2 

Inform/Document 

Court  Judge 

Affidavit 
Deposition 

Varies 

Inform/Document 
(Persuade) 

808 


Written  Communication 


The  more  you  know  about  your  readers,  the  more 
you  can  target  and  tailor  your  writing  and  communi- 
cate more  effectively. 

"How  Long  Will  Writing  Take?" 

When  you  begin  writing,  you  need  to  know 
how  many  hours,  days,  or  weeks  it  will  take  you  to 
write  memos,  letters,  reports,  and  other  documents. 
So  keep  records.  And  when  you  plan  a  project,  bud- 
get enough  time  for  completing  the  project's  writing 
tasks. 


You  may  find  incremental  writing  useful.  Incre- 
mental writing  involves  writing  parts  of  a  manuscript 
as  a  project  progresses.  Often,  you  can  write  sections 
of  a  report  early  in  a  project  and  use  them  in  later 
manuscripts.  Such  an  approach  allows  you  to  pro- 
duce a  final  report  in  less  time  at  the  project's  end. 
Avoid,  if  possible,  putting  off  writing  until  the  pro- 
ject's end.  By  beginning  your  writing  early,  you  will 
have  time  for  rewriting  which  will  produce  a  better 
manuscript. 

Organizing  Your  Writing 

Too  often  writers  turn  to  an  old  report,  an  old 
memo,  or  other  manuscripts  as  a  guide  for  organiz- 
ing their  writing.  Before  you  do,  ask  yourself  if  regu- 
lations, policies,  or  publication  guidances  require 
that  you  use  a  specific  organization.  If  so,  follow  it.  If 
allowed,  add  a  memo,  letter,  abstract,  or  executive 


summary  giving  the  key  points  your  readers  should 
know.  By  doing  so,  you  will  help  busy  readers  recall 
the  essential  points  if  they  do  not  have  time  to  read 
the  entire  report. 

Make  sure  your  manuscripts  communicate.  Too 
often  manuscripts  tell  about  rather  than  tell.  Con- 
sider the  memo  and  comments  in  Figure  1. 

For  longer  papers,  you  will  find  outlining  makes 
your  writing  easier.  Outlining  serves  as  a  map  to 
guide  your  writing.  Outlining  saves  you  time — mak- 
ing changes  on  an  outline  takes  less  time  than  re- 
writing a  manuscript.  An  outline  helps  you  see  your 
organization  more  clearly  than  a  manuscript's 
narrative. 

Professionals  use  different  approaches  to  outlin- 
ing. In  one  of  the  few  studies  of  outlining  practices, 
McKee  (1975)  studied  technical  writers  in  the  Soci- 
ety for  Technical  Communication.  Of  the  80  writers 
who  responded  to  McKee's  survey,  60%  used  out- 
lines consisting  of  words  or  words  and  phrases.  An- 
other 30%  used  outlines  consisting  of  words, 
phrases,  and  sentences.  Only  5%  used  the  traditional 
sentence  outline,  and  5%  used  no  outline. 

Chandler  (  1978)  wrote  that  ready-made  outlines 
cheat  you  and  your  readers.  He  stated  that  ready- 
made  outlines  hamper  communication  because  the 
authors  who  use  them  bury  their  main  points  and 
force  readers  to  read  hundreds  of  words  to  learn  the 
main  points.  Instead  of  the  ready-made  outline, 
Chandler  recommended  letting  the  contents  and  the 
points  you  want  your  readers  to  know  dictate  how 
you  organize  your  manuscripts. 

Some  professionals  summarize  their  main  points 
before  they  begin  their  outlines.  Professor  James  A. 
Bailey  of  Colorado  State  University  says  he  writes  his 
conclusions  in  one  to  five  short  sentences.  Then  he 
lets  those  sentences  guide  his  outlining. 

Bailey  also  cautions  against  letting  the  order  of 
data  analysis  dictate  your  organization.  How  you  ana- 
lyze data  may  have  no  relation  to  the  data's  order  of 
importance  and  the  conclusions  you  draw  from  the 
data.  So  let  the  importance  of  the  points  dictate  the 
data's  organization  and  order  of  presentation. 

Other  writers  begin  by  making  a  rough  outline 
of  their  major  points  and  repeatedly  expand  the  out- 
line. Gradually,  an  outline  emerges  that  the  authors 
use  as  a  guide  for  drafting  their  manuscripts. 

I  used  such  an  approach  in  outlining  this  chap- 
ter. I  prepared  seven  versions  before  a  23-page  typed 
working  outline  emerged.  After  letting  the  outline 
rest,  I  discussed  it  with  an  editor  and  other  profes- 
sionals. Next  I  incorporated  selected  points  from  the 
first  13  pages  into  the  last  10  pages.  These  changes 


Written  Communication 


809 


Original  Memo 


Comments/Questions 


FROM:      District  Biologist 

SUBJECT:      Interim  Preliminary  Report 

Barstow-Las  Vegas  Motorcycle 
Race   on  wildlife   habitats 

The   initial    investigation  was   completed 
in  October   1974  and   the   follow-up 
investigation  was  completed    in  October 
1975.      The   investigation  was   the  most 
interesting  I   have  undertaken  and   it 
should   provide  us   with  practical 
information    for  assessing   potential 
damages    from   future  motorcycle    races 
and    related   off-terrain   vehicle   use. 

There   were   two   study   sites   chosen   in 
the   vicinity   of   the   start   of   the 
Barstow-Las  Vegas  Race.      Study   Sites 
I   and  II   contained   two  different 
habitat    types.      Both  pre   and   post    race 
sites  were   evaluated. 

On  Study   Sites   I   and   II    the    pre-race 
capture    success   was    low.      The    trapping 
success   rates   after   the    race   were   even 
lower.      A  marked   reduction   in   trapping 
success   occurred   on   the   area. 

The   effects  of    the   race   on   the   habitat 
types   varied   and    the    agency   can   draw 
specific   conclusions    from  the  pre   and 
post-race   data    for    subsequent    races. 

Most   data  were  of   great   practical  use 
and    some   were   quite    interesting.     .     . 
and   show  how  motorcycles   and   off-terrain 
vehicles  may   have   an   impact   on   vegeta- 
tion  types    in   the   deserts. 

It   was  good   that   we   were   able    to 
complete   the  project   as   it   provided 
excellent    information   for  management 
decisions   and   assessing   impacts   on 
the   desert    environment. 

I  think  "atta  bodys"  should  go  to  Dave 
Smith  and  Kim  Vandermiden,  area 
biologists  who  helped  with  the  project 
on  their  own  time  and  at  their  own 
expense.  They  helped  collect  the  very 
valuable  and  important  information  for 
us. 


Why  not  say  interim  report  or  preliminary  report? 
Do  you  really  need  both  words? 


Why  take  the  whole  sentence  to  add  the  dates? 
Include  them  with  the  subject  line. 

Why  was  the  investigation  the  most  interesting  that 

you  undertook? 

What's  the  point  of  the  memo?  The  State  Biologist 

doesn't  have  time  to  wade  through  a  lengthy,  general 

narrative.  Of  what  practical  value  should  the  data  be? 

How  will  the  data  help  the  agency  assess  future  data? 

Provide  specifics. 

What  were  the  sites?  What  specific  vegetation 

sites/habitat  types  were  studied?  Why? 

Do  you  mean  that  you  had  two  or  four  study  sites? 

You're  unclear.  What's  your  experimental  design? 

What's  the  methodology?  Provide  specifics. 

What  do  you  mean  low?  Provide  specifics.  What  does 
even  lower  mean?  Provide  specifics.  What's  a  marked 
reduction?  What  are  the  limitations  on  the  data? 


How  did  the  effects  vary?  Be  specific.  What 
conclusions  should  you  draw? 


What  are  the  practical  values  of  the  data?  What 
implications  do  the  data  and  your  interpretations 
have  for  subsequent  management?  Should  you  hedge 
with  "may  have  an  impact"  or  should  you  be  specific 
and  identify  the  impact? 

Why  good?  What  does  good  mean? 

What  does  "excellent  information"  mean? 


Should  the  State  Biologist  send  a  letter  of  thanks?  Or 
take  other  action?  Be  specific. 

Reorganize  the  memo  so  you  tell  the  State  Biologist 
in  the  first  paragraph  the  impacts  of  the  race  on  the 
habitat  types  and  the  management  implications. 
Provide  alternatives;  be  specific. 


Figure  1.  Memo  that  "tells  about"  more  than  it  "tells.' 


810 


Written  Communication 


gave  a  tighter  focus  on  developing  a  writing  strategy 
and  minimizing  the  discussion  of  research  on  profes- 
sional writing.  I  then  added  more  marginal  notes. 
Figure  2  shows  one  page  that  I  used  to  guide  my 
drafting  the  manuscript. 

So,  what  guidelines  emerge  from  the  foregoing 
discussion?  First,  use  some  form  of  an  outline.  The 
exact  form  does  not  matter  as  long  as  it  helps  you 
write.  You  may  find  a  detailed  approach  helpful  or 
you  may  find  a  few  key  words  and  phrases  enough  to 
guide  your  writing.  Second,  let  your  outline  rest  be- 
fore revising  it.  Time  helps  give  you  a  critical  per- 
spective and  will  help  you  improve  the  manuscript's 
organization.  Third,  when  you  revise,  evaluate  your 
logic.  Do  you  need  to  delete  points,  add  points,  or 
rearrange  points  to  build  your  rationale  and  logic? 


Preparing  Illustrations 

Ideally,  plan  your  illustrations — tables,  photo- 
graphs, graphs,  and  line  art — when  you  begin  the 
project.  You  may  encounter  problems  if  you  wait  un- 
til the  project's  end  to  shoot  photographs  and  have 
your  line  art  prepared.  Most  illustrations  benefit  from 
early  planning  in  the  project.  At  the  latest,  prepare 
the  tables  and  graphs  after  you  have  outlined  the 
manuscript. 

Most  style  manuals  and  guides  divide  illustra- 
tions into  tables  and  figures.  Any  illustration  that  is 
not  a  table  is  a  figure.  Although  tables  normally  pre- 
sent numbers,  they  can  present  short  narratives.  Fig- 
ures include  graphs,  line  art,  photographs,  and 
listings. 

Good  illustrations  save  writing,  support  the  nar- 
rative, and  show  essential  points  that  are  difficult  to 
present  in  the  narrative.  Illustrations  help  you  com- 
municate your  key  points  clearly  and  succinctly.  And 
a  good  illustration  may  eliminate  hundreds  or  thou- 
sands of  words. 

What  kinds  of  illustrations  should  you  use?  Ta- 
bles work  best  for  presenting  statistics,  data,  and 
complex  information.  Graphs — line  graphs,  bar 
graphs,  and  circle  graphs — may  give  specific  informa- 
tion, but  more  often  they  give  a  general  impression 
and  show  trends.  Newspapers  and  popular  magazines 
use  greatly  simplified  graphs  and  tables.  Line  art 
(drawings)  serve  well  for  many  readers,  but  line  art 
must  be  oriented  for  readers.  Will  your  readers  un- 
derstand the  line  art?  Maps,  for  example,  provide 
useful,  detailed  information,  but  they  need  a  north 
arrow,  legend,  and  scale.  Without  such  information, 
readers  unfamiliar  with  an  area  will  have  trouble  lo- 
cating the  site.  Photographs  provide  an  image  of  real 
animals,  objects,  vegetation,  or  an  area,  and  often 
convey  concepts  more  effectively  than  tables  (Figure 
3 ).  Photographs  work  well  for  most  readers,  but 


some  readers  may  not  see  what  you  want  them  to 
see.  So  provide  information  in  the  legend  or  caption. 

Before  preparing  any  illustrations,  study  the 
publication's  style  for  illustrations.  If  you  have  no 
style  manual  or  guide,  follow  the  current  Journal  of 
Wildlife  Management  style  and  CBE  Style  Manual 
(CBE  Style  Manual  Committee  1983). 

Tips  on  Preparing  Tables.  A  standard  table  format 
produces  uniformity  and  makes  data  presentation 
more  efficient  and  effective.  Readers  need  not  search 
for  missing  information  and  they  will  find  table  inter- 
pretation easier.  Figure  4  shows  the  standard  parts  of 
the  table  as  adapted  from  the  CBE  Style  Manual. 
These  parts  also  conform  to  the  Journal  of  Wildlife 
Management's  style. 

A  cursory  review  I  did  of  BLM  publications 
showed  inconsistencies  such  as  lack  of  a  title  above 
tables  and  inconsistent  use  of  divider  lines.  Other 
problems  involved  cramming  too  much  information 
into  a  table  and  reducing  tables  to  a  size  that  many 
readers  could  not  easily  read  the  numbers  and 
words.  When  reducing,  some  photocopying  ma- 
chines blur  letters  and  numbers. 

Another  common  problem  involves  incorrectly 
using  significant  digits.  Some  authors  use  digits  after 
the  decimal  when  whole  numbers  would  do.  Thus, 
in  a  food  habits  study,  reporting  a  particular  food  in 
5.1%  rather  than  5%  of  the  samples  illustrates  un- 
necessary detail  and  a  level  of  precision  that  does 
not  exist. 

When  tables  will  be  printed  from  typed  tables, 
use  a  sans-serif  type  face  such  as  Letter  Gothic  or 
Orator  (Figure  5).  Sans  serif  faces  have  no  cross 
strokes  at  the  letter's  end.  Serif  type  such  as  Prestige 
Elite,  Courier,  and  Modern  has  cross  strokes  at  the 
letter's  end.  Such  type  faces  make  for  easier  reading 
of  lengthy  narratives  but  blur  when  reduced  too 
small,  and  reading  becomes  difficult. 

If  possible,  use  upright  or  vertical  format  tables 
as  Figure  4  illustrates.  Avoid  using  tables  that  force 
the  reader  to  turn  the  manuscript  sideways  to  read 
the  table  and  avoid  reducing  tables  so  small  that 
readers  have  trouble  deciphering  the  numbers,  as  in 
Figure  6.  Many  readers  will  not  take  the  time  and  ef- 
fort to  turn  a  manuscript  or  decipher  small  printing. 

Avoid  multiple-page  tables  and  tables  with  ex- 
cessive columns.  In  most  instances  you  can  avoid  the 
above-mentioned  problems  by  breaking  complex  ta- 
bles into  two  or  more  tables. 

Tips  on  Figures.  When  preparing  figures,  put  the  ti- 
tles below,  not  above  the  figures.  Use  a  consistent 
format  and  style  for  titles  and  do  not  mix  typed, 
typeset,  and  hand-lettered  titles.  For  typed 


Written  Communication 


811 


VI.   A  suggested  approach  to  writing 
A.   Thinking  and  Planning 

Provide  adequate  time  for  thinking  and  planning  your  communications. 

Di?4*****!^'  u     Who  are  my  readers?  ^It^rt  f  frA+»t  ft-  ri4<+w*> 

0^  A**  -  9  ( — t—t 

xLY*      Jt7         2.  What  is  my  content?  [Ft*»»€T'*r» '  ',t%Ti 

k  r/*"   tPk             3.   What  do  I  want  to  communicate  to  my  readers?*^  rf"r ' 


»<«* 


V 


*♦** 


^EmmmumT* 

4.  What  is  the  appropriate  format  and  organization  for  presenting 
my  information?  ^;cf0A^^    u/rf+l   U*eU*    t*»ft*v\iry$.' 

Developing  a  working  outline  y" 

1.  Your  outline  may  be  a  mixture  of  sentences,  phrases,  clauses 
and  words.   It's  a  working  document  and  seldom  will  others  see  it. 

2.  Outlining  helps  you  see  your   thinking  and  putting  your  thinking 
on  paper  helps  most  writers.  kwW"-tffy/e  -rt&At*± 

3.  Letting  an  outline  rest  a  few  days  or  more  helps  you  reconsider 
your  approach. 

4.  Remember  it's  easier  to  revise  an  outline  than  it  is  to  revise 
an  entire  document. 

5.  When  you  begin  writing,  keep  your  outline  in  mind,  but  don't 
feel  constrained  by  it.   After  having  thought  about  it,  an  improved 
organization  may  emerge. 

Prepare  your  tables  and  figures  -  -  JUut+ra+i Vt-  M*T**''al'» 

1.   Tables  and  figures  should  be  based  on  your  message  and 
reader's  abilities  to  understand  and  interpret  them.   Anything 
not  a  table  is  a  figure.   Figure  XX  provides  guidelines  for  selecting 
different  kinds  of  tables  and  figures. 


INSERT  FIGURE  XX  ABOUT  HERE 


and  figures.    ,    -        •  ..-..«*•  /Aa+a 
E  rtdLvCH     VUY*ril4C 


2.   By  planning  illustrations  early,  illustrators,  artists  and 
photographers  can  meet  the  publication's  deadlines. 


Figure  2.  Sample  page  of  working  outline  developed  for  this  chapter. 


812 


Written  Communication 


Figure  X.   Recovery  of  Ramsey  Creek  following  stream  corridor  fencing  in  1974. 


Table  XX .   Comparison  of  some  vegetation  and  physical  parameters  of  Ramsey 
Creek  from  1974  to  1980  following  stream  corridor  fencing  and  natural 
revegetation. 


1974 

1976 

1978 

1980 

Vegetation 

Percent  Cover 
Bare  Soil 
Litter 
Grass 
Forbs 
Shrubs 
Trees 

8 

2 

5 

12 

26 
4 
23 
15 
30 
2 

17 
14 
9 
22 
35 
23 

5 

12 

9 

8 

30 

36 

Stems/Acre 
over  1"  dbh 

- 

- 

354 

986 

Physical  Parameters 

Average  stream 
width  (ra) 

3.1 

2.7 

2.5 

2.1 

Average  depth  (m) 

.09 

.08 

.21 

.33 

Strearabank  stability 
(7,   vegetation) 

Poor 

(0-25) 

Fair 

(26-50) 

Good 
(51-751 

Excellent 
(75-100) 

Streamside  cover 
(rating)  (dominant 
vegetation   type) 

Poor 
(Forbs) 

Fair 
(Grass/Forbs) 

Good 
(Shrubs) 

Excellent 
(Shrubs/Trees) 

jy  Simulated  Data 


Figure  3.  Comparison  of  tables  and  photographs  illustrating  the  habitat  changes  from  1976  to  1980 


Written  Communication 


813 


© 


0 


© 


Table  XX.  Average  total  mule  deer  per  census  on  4  burned 
and  seeded  and  adjacent  unburned  control  quadrats  in  1984 
and  1985.  Each  quadrat  was  censused  3  times  each  January, 


©► 


(V)»-Mule  deer  per  census 


©1     ©• 


1984 


1985 


(jT)»Quadrat 


Burn 


Control     Burn 


Control 


©- 


Apr  1980  burn 


©' 


Sagebrush 

Pinyon/Juniper 

Grassland 

Oct  1982  burn 

Grassland 


10. 3a 
15. 8b 
18.1^ 


14. 4a 


5.4 
11.8 
10.9 


5.6 


9.4 

7.7 

11.1 


6.3 


8.4 

16. la 

7.3 


5.0 


13V 


© 


ra  Greater  (  P  <  0.05  )  than  other  treatment  within  year, 
I  Mann-Whitney  U  test. 

Lb  P  <0.1,  Mann-Whitney  l)  test. 


Figure  4.  Parts  of  a  typical  table. 

©table  number  (?)  table  title  or  legend  ©rule  ©spanner  head 

(?)  rules  (?)  sub-spanner  heads  with  rules  ©  stub  head  ©  spanner 

©rule  ©field  ©subheadings  ©stub  ©rule  ©footnotes 


814 


Written  Communication 


Serif  Type  Faces 

Prestige  Elite 

Courier 


Sans-serif  Type  Faces 

Letter  Gothic 

Presentor  Orator 


Figure  5.  Typical  type  faces  illustrating  serif  and  sans-serif  type  faces. 


publications,  type  all  titles  and  insert  them  after  the 
figures  have  been  reduced  to  fit  the  page.  Make  sure 
the  type  remains  legible  and  easy  to  read.  Make  the 
titles  complete  so  the  figure  can  stand  alone  by  in- 
cluding such  information  as  the  location  of  the  study 
and  the  plant  or  animal  species  studied. 

Have  graphs  and  artwork  drawn  larger  than,  but 
proportional  to  the  final  page  size.  When  drawing  fig- 
ures, do  not  use  too  much  information  or  too  many 
lines  and  clutter  the  figure.  But  provide  enough  in- 
formation so  the  figure  is  not  confusing.  For  exam- 
ple, adding  a  regression  formula  helps  clarify  your 
data  in  a  line  graph. 

When  you  submit  figures  for  a  publication,  mark 
the  top.  When  you  check  the  galleys  or  proofs  of  the 
article,  make  sure  the  figure  was  not  inserted  upside 
down  or  backwards.  (Don't  laugh — it  happens  much 
to  the  embarrassment  of  authors  and  editors. )  If 
someone  else  inserts  the  figures,  make  sure  they 
have  inserted  them  correctly. 

If  you  need  photographs  to  illustrate  a  manu- 
script, shoot  black  and  white  negatives  during  the 
project.  Even  if  you  foresee  no  need  for  photo- 
graphs, shoot  them  anyway.  And  when  you  have  the 
prints  prepared,  request  glossy  black  and  white 
prints.  For  a  detailed  discussion  of  illustrations,  see 
Allen  (1977)  and  CBE  Style  Manual  Committee 
(1983). 

Drafting  the  Manuscript 

To  draft  your  manuscript,  use  your  outline  and 
illustrations  as  guides  and  put  your  thoughts  on  pa- 
per quickly.  Do  not  become  a  slave  to  your  out- 
line— you  may  develop  ideas  that  will  improve  your 
organization,  your  content,  or  your  points.  If  so,  use 
them.  When  you  write  your  first  draft,  do  not  worry 
about  spelling,  style,  and  grammar — correct  them 
later. 


As  you  work  through  your  outline,  use  tables, 
figures,  and  other  illustrations  to  reduce  the  length 
of  your  narrative.  Discuss  their  important  points. 
Point  out  what  you  think  the  data  and  information 
mean.  Place  illustrations  in  the  text  right  after  their 
first  mention  in  the  narrative.  So  provide  instructions 
for  inserting  the  illustrations  into  the  text.  Insert  a 
line  such  as  v 

Insert  Figure 


XX  about   here 


to  tell  the  word  processing  operator,  typist,  or 
printer  where  to  insert  the  illustration  in  the  copy  to 
be  printed  or  reproduced. 

Today,  many  professional  writers  compose  di- 
rectly on  typewriters  or  word  processors.  Recent  ad- 
vances in  computers  suggest  that  more  and  more 
professionals  from  all  fields  will  soon  be  using  com- 
puters with  word-processing  capabilities.  But  for 
now,  many  biologists  do  not  have  access  to  a  com- 
puter with  word-processing  capabilities  or  even  a 
typewriter  for  their  personal  use.  So  they  must  write 
longhand  with  pencil  and  paper. 

Draft  your  manuscript  and  then  have  it  typed 
double  spaced  with  1-  or  1-1/2-inch  margins  on  all 
sides.  Type  on  only  one  side  of  the  paper.  Once  you 
have  a  typed  copy,  make  a  photocopy  for  revising 
and  editing.  Store  the  original  in  a  safe  place  as  you 
may  need  it  later. 

Writing  Abstracts  and  Executive  Summaries 

When  preparing  lengthy  documents,  include 
either  an  abstract  or  an  executive  summary.  Often  an 
abstract  is  the  only  part  of  your  writing  that  busy 
managers,  supervisors,  and  biologists  will  read. 
Professionals  seldom  read  an  entire  report,  article,  or 
document,  but  they  do  read  the  abstracts  and  execu- 
tive summaries.  Furthermore,  the  abstract  or  execu- 
tive summary  will  often  be  separated  from  the  report 
and  circulated  to  different  readers.  Therefore,  your 
abstract  or  executive  summary'  must  stand  alone  and 


Written  Communication 


815 


I    .    ! 


oc 

c 

1 


V 

oc 

u 
3 
O 
u 


u 
oc 
OJ 
a 

u 

s: 


51      3 


55 


3^  5 


S  S  £ 


M  £ 


i!  ?.  a  .? .«  «  f  °  $  t*  • 


'  >    c        ■   S  "5. 


i S»h|| 11 


C 
O 

t 

■o 


«5 

u 

s 

03 
OC 

c 


3 
H 

OS 


e 
•s 

u 

e 

■o 

c 

oi 


OC 

u 


M 

03 

OC 

c 
'0 

3 

■a 
u 

SO 

u 

I- 


816  Written  Communication 


contain  the  essential  information  your  readers  need 
to  know. 


avoid  sentences  with  greatly  separated  subjects  and 
verbs. 


Abstracts  and  executive  summaries  should  tell 
rather  than  tell  about  (see  Figure  1 ).  Simply  provide 
the  necessary  specifics  and  come  to  your  points 
quickly.  Usually,  abstracts  are  for  narrowly-focused, 
specific  research  projects;  executive  summaries  are 
for  more  broad-based,  diverse  investigations. 

Your  abstracts  should  contain  brief  statements 
on  the  research  question  or  purpose,  methodology, 
and  conclusions.  With  tight  writing,  you  can  con- 
dense an  abstract  to  one,  two,  or  three  sentences. 
But  normally  your  abstracts  will  run  between  25  and 
200  words,  depending  upon  the  agency's  or  publica- 
tion's guidelines.  Do  not  cite  literature,  refer  to  ta- 
bles or  figures  in  the  report,  or  include  tables  or  fig- 
ures in  the  abstract. 

Executive  summaries  may  run  up  to  1,000 
words,  and  they  explain  the  project  in  more  depth 
than  an  abstract.  Furthermore,  executive  summaries 
are  usually  prepared  for  administrators  and  less  tech- 
nical readers.  Therefore,  translate  jargon  and  define 
technical  terms  to  be  sure  your  readers  understand 
your  points. 

Keep  your  writing  succinct  and  provide  only 
enough  detail  to  support  your  points.  Keep  in  mind 
that  most  readers  will  not  read  your  full  report. 
Thus,  the  executive  summary  must  be  a  tight  presen- 
tation of  the  key  points.  And,  like  writing  the  ab- 
stract, do  not  include  citations,  illustrations,  figures, 
or  tables. 

Finally,  most  professionals  find  writing  the  ab- 
stracts and  executive  summaries  easy  after  they  have 
written  the  complete  report.  Conclusions,  headings, 
and  major  titles  often  provide  guidelines  for  develop- 
ing abstracts  and  executive  summaries. 

Revising  and  Self-editing 

Clear  writing  comes  from  clear  thinking  and  re- 
writing. Rewriting  entails  self-editing — a  way  of  revis- 
ing your  writing.  By  rewriting  you  can  sharpen  your 
narrative  and  make  your  points  succinctly.  Some 
professionals  rewrite  their  manuscripts  two,  three, 
four,  or  more  times  before  they  have  a  copy  for  oth- 
ers to  critique. 

Self-editing  helps  you  tighten  your  writing.  Too 
often  we  write  long,  complex,  involved  sentences. 
Most  readers  find  short  sentences  easier  to  under- 
stand than  long  sentences.  As  a  rule  sentences 
should  average  about  20  words,  but  avoid  short, 
choppy  sentences.  Vary  sentence  length.  Avoid  sen- 
tences with  excessive  prepositional  phrases;  edit  sen- 
tences with  long,  complex,  involved  clauses;  and 


To  practice  self-editing,  plan  to  make  several 
trips  through  your  typed  manuscript  to  achieve  the 
following  goals: 

1.  Correct  obvious  typing  and  spelling  errors. 

2.  Correct  errors  in  content. 

3.  Correct  organizational  and  logic  problems. 

4.  Eliminate  wordiness. 

To  illustrate  self-editing,  assume  you  have  drafted  the 
introduction  to  a  BLM  Technical  Note  as  in  Figure  7. 
How  can  it  be  improved?  I  pose  questions  about  its 
content  and  readers,  and  then  I  suggest  how  you 
might  tighten  the  copy.  By  applying  self-editing  to 
your  writing,  you  will  clarify  your  points. 

Correct  Typing  and  Spelling  Errors.  Begin  by 
quickly  reading  through  your  typed  manuscript  and 
correcting  obvious  typographical  and  spelling  errors. 
Correct  the  errors  by  using  standard  copyediting 
symbols  as  illustrated  in  Figure  8.  The  copyediting 
symbols  show  standard  changes  for  retyping  or  type- 
setting the  manuscript  as  galleys — the  printed  copy. 
When  copyediting,  use  a  soft-lead  pencil.  Changes 
made  with  hard-lead  pencils  are  hard  to  read  and  dif- 
ficult to  erase. 

Check  for  Content  Errors.  Read  through  your 
manuscript  again  looking  for  errors  in  content.  Have 
you  accurately  presented  the  data?  Have  any  typo- 
graphical errors  changed  any  meanings?  Are  you  us- 
ing the  correct  terms?  Have  you  used  the  terms  in 
the  correct  context?  Have  you  presented  anything  or 
used  an  approach  that  will  raise  questions  about 
your  problem-solving  skills?  If  so,  have  you  explained 
why?  Critically  review  your  manuscript  as  if  it  were 
written  by  another  author. 

Errors  easily  creep  in  if  you  are  not  extremely 
cautious  and  careful.  Thus,  the  author  who  writes 


"During  the  study  we  collected  1 2  fish  species 
in  I960,  15  in  1965,  17  in  1970,  18  in  1975, 
and  20  in  1980.  The  116%  decrease  can  be  at- 
tributed to. . ." 

Here  the  author  should  have  written,  "The 
116%  increase  in  species  can  be  attributed 
to . . ."  Although  strikingly  apparent,  such  errors 
can  creep  in  when  writers  become  too  close  to 
their  writing.  To  reduce  the  chances  of  such 
errors  slipping  into  your  writing,  put  it  aside 
for  days  or  better  yet  a  week  or  two  before 
you  review  it.  Then  carefully  check  for  errors. 
And  then  have  another  professional  check  your 
manuscript. 


Written  Communication 


817 


Original 

Comments 

Suggested  rewrite 

In  compliance  with  the  National 

Is  the  first  paragraph  needed? 

To  comply  with  the  National 

Environmental  Policy  Act  (NEPA) 

Don't  the  readers  know  about 

Environmental  Policy  Act  (NEPA) 

of  1970,  the  Bureau  of  Land 

the  NEPA  and  EISs?  If  so, 

of  1970,  the  Bureau  of  Land 

Management  (BLM)  is  required 

delete  the  paragraph.  If  not, 

Management  (BLM)  must  file 

by  law  to  address  the  resources 

tighten  it. 

Environmental  Impact  Statements 

and  impacts  of  land  use  on 

(EISs)  for  uses  of  national  public 

national  public  lands  in  the  form 

lands. 

of  Environmental  Impact 

Statements  (EISs). 

Extensive  wildlife  inventories 

Is  it  necessary  to  say  the  BLM 

The  BLM  implemented  (began?) 

were  first  implemented  on  public 

implemented  the  inventories? 

extensive  inventories  .  .  . 

lands  following  the  October  1973 

What  did  the  suit  change? 

lawsuit  filed  in  the  Federal 

Should  it  be  that  specific? 

District  Court  of  the  District  of 

Should  other  conservation 

The  plaintiffs  successfully  argued 

Columbia  by  the  Natural 

groups  be  mentioned?  Who 

that  the  initial  BLM  inventories 

Resources  Defense  Council  and 

successfully  argued?  The 

did  not  adequately  address  the 

other  conservation  groups.  It  was 

Defense  Council?  Other 

impacts  of  local  grazing  and  did 

successfully  argued  that 

groups?  Can  the  sentence  be 

not  comply  with  NEPA. 

evaluation  of  livestock  grazing  on 

recast  in  active  voice? 

public  lands,  first  filed  by  the 

BLM  with  the  Council  on 

Environmental  Quality  was  too 

general  in  content  to  properly 

address  localized  impacts,  and 

therefore,  did  not  comply  with 

NEPA. 

It  is  estimated  that  310  stream 

Why  be  so  wordy?  Recast  into 

The  BLM  administers  about  310 

miles  and  over  1 1 ,800  acres  of 

active  voice  for  a  tight 

stream  miles  and  over  1 1 ,800 

riparian  woodland  are 

sentence. 

acres  of  riparian  woodland  in 

administered  by  the  BLM  in 

Arizona.  To  comply  with  the 

Arizona.  In  an  effort  to  comply 

Here's  another  wordy 

court,  this  study  provided 

with  the  court,  it  became  the 

sentence.  Can  it  be  tightened? 

aquatic  habitat  and  water  quality 

purpose  of  this  study  to  provide 

Can  the  "Bureau  of  Land 

information  to  the  BLM's  Phoenix 

aquatic  habitat  and  water  quality 

Management"  be  abbreviated? 

District  Office. 

information  to  the  Phoenix  District 

Office  of  the  Bureau  of  Land 

Management.  These  baseline 

Again,  can  the  point  be  made 

The  baseline  data  provide 

data  are  to  serve  as  an 

more  quickly?  Is  "on  public 

information  for  developing 

information  source  to  be  utilized 

lands"  necessary?  BLM  doesn't 

management  plans  for  terrestrial 

in  the  management  plans  for 

handle  other  lands,  does  it? 

and  aquatic  resources. 

terrestrial  and  aquatic  resources 

Did  the  suit  focus  on  only 

on  public  lands. 

public  lands? 

Figure  7.  Introduction  to  a  Technical  Note  with  comments  and  suggestions  for  rewriting. 


818 


Written  Communication 


Meaning 


Delete 


Circle  means  to  spell  out 


Circle  means  to  abbreviate 


For  capitals,  underline  three  times   

Draw  a  slanted  line  through  a  letter  to  change  to 
lower  case 

Delete  letter  and  close-up  using  normal  spacing  - 


Delete  letters  and  use  normal  spacing  between 
words 


Underline  to  have  words  set  in  italics 

Begin  a  new  paragraph  

Center  copy   


Set  or  type  flush  left  - 

Set  or  type  flush  right 
Insert  missing  words  - 

Insert  missing  letter  — 

Transpose  words  

Transpose  letters  


Separate  run-on  words 
Insert  period 


Insert  comma 
Insert  hyphen 


Let  original  copy  stand 
Close  up  copy 


For  grossly  misspelled  words,  mark  out  and  write 
above 


Example 

It  has  zones  of   high  and 


26/ trees  were  sampled 
The [Bureau  of  Land  Management) .  .  . 
The  department  of  the  interior  .  .  . 

The  /gency  policy  includes  .  .  . 
Its  numbers  were  redu^ced. 

.  .  .  butiti?  occurs^  in  fewer  numbers. 

Dipodomys  merriami 
J  Eight  species  were  snap  trapped. 

"}  Introduction  [_. 
References  "C 

""JTotals 

The  highest  forj^rd's  kangaroo  rat  .  . 

.  .  .  northern  grasshopper  mouse  .  .  . 
Eight  species  were  jtrappedj  snap!. 

Eight  spcfkes    .  .  . 

Eightspecieawe  re/trapped. 

Eight  species  were  trapped  £V 

Additionally.the  white-footed  mouse  . 

.  .  .  the  white-footed  mouse  .  .  . 

Simple  -mono t ypjrc.  vegetation  types  .  • 

Arizona  Upland  was  further  divided  into, 
frhrco  otondard  habitat  -oitoo  based  ony 
gradient. They  includG(f^desertscrub 


/Ttoku-i-^ 


Its  numbers  were 


crJVje&d . 


Figure  8.  Selected  copyediting  marks  for  revised  typed  manuscripts. 


Written  Communication 


819 


On  the  basis  of  your  expertise  in  wildlife,  do 
you  see  any  errors  of  logic  in  your  approach? 
Did  you  follow  acceptable  and  logical  ways  of 
solving  the  problem?  Do  your  data  support 
your  conclusions?  Do  you  limit  your  conclu- 
sions? In  what  ways  might  your  content  be  crit- 
icized? Have  you  adequately  supported  your 
position?  What  weak  points  in  logic,  problem 
solving,  and  methodologies  become  apparent 
upon  reading?  Can  you  correct  the  weaknesses? 
As  weaknesses  become  apparent,  consider  the 
potential  criticism.  Should  you  revise?  You  may 
find  yourself  revising  sentences,  paragraphs,  or 
whole  sections. 

Check  Your  Organization.  Read  through  your 
manuscript  again.  Have  you  adequately  considered 
the  reading  habits  of  your  readers?  Do  you  give  your 
readers  an  overview  of  what  is  to  come?  Will  your 
organization  help  your  readers  understand  your  con- 
tent? Do  you  progress  logically  from  point  one  to 
point  two  to  point  three?  Should  you  make  some 
points  later  in  your  manuscript?  Should  some  points 
be  made  earlier?  Have  you  adequately  denned  and 
explained  points  when  you  introduce  them?  Does 
your  organization  raise  any  questions  about  your 
logic? 

Eliminate  Wordiness.  Your  writing  will  benefit  by 
cutting  wordiness.  With  a  heavy  edit,  you  will  im- 
prove your  writing's  clarity  and  conciseness.  When 
you  write  tightly,  your  thinking  comes  through 
clearly  and  you  will  increase  your  reader's 
understanding. 

To  cut  wordiness,  eliminate  unneeded  words, 
phrases,  clauses,  sentences,  and  paragraphs.  When 
possible,  replace  a  clause  with  a  phrase,  and  a  phrase 
with  a  word.  Question  your  words.  Can  you  say 
what  you  want  more  effectively?  Can  you  make  your 
points  in  fewer  words?  Do  you  need  that  word?  That 
phrase?  That  clause?  What  can  you  eliminate? 

Developing  your  self-editing  skills  requires  rec- 
ognizing wordy  constructions  and  knowing  how  to 
improve  them.  To  improve  your  writing: 


1.  Use  active  voice. 

2.  Use  pronouns. 

3.  Replace  small,  weak  verbs. 

4.  Eliminate  unnecessary  clauses. 

5.  Eliminate  unnecessary  phrases. 

6.  Omit  needless  words. 

7.  Replace  vague,  general,  and  abstract  terms. 

8.  Avoid  jargon  and  acronyms. 


Use  Active  Voice  Active  voice  makes  clear  who 
did  what.  It  tells  the  reader  who  carried  out  the  ac- 
tion. Compare  the  active  voice  with  passive  voice 
constructions: 

Active:  The  division  manager  cancelled  the  project. 

Passive:  The  project  was  cancelled. 

Passive:  The  project  was  cancelled  by  the  division 
manager. 

In  the  first  sentence  we  know  who  did  what.  In  the 
second  sentence,  we  do  not  know  who  cancelled  the 
project.  In  the  third  sentence,  we  finally  learn  who 
did  what,  but  late  in  the  sentence.  The  active  voice 
quickly  tells  your  reader  who  carried  out  the  action. 
Use  the  active  voice  for  shorter  sentences. 

You  can  recognize  whether  a  sentence  is  active 
or  passive  by  asking,  "Did  the  subject  carry  out  the 
action?"  If  not,  you  have  a  passive  construction.  You 
can  recognize  passive  constructions  by  looking  for  is, 
was,  were,  and  other  to  be  verb  forms.  In  the  two 
passive  constructions  above,  was  signals  the  passive 
voice.  In  some  sentences,  a  prepositional  phrase  be- 
ginning with  by  signals  the  passive  construction. 

When  should  you  use  the  passive  construction? 
The  CBE  Style  Manual  (CBE  Style  Manual  Committee 
1983)  suggests  the  use  of  passive  voice  when  the 
agent  of  the  action  is  irrelevant  and  when  something 
or  someone  other  than  the  agent  is  more  important. 
Consider: 

The  National  Environmental  Policy  Act  was 
passed  in  1970. 

Here  the  passage  of  the  act  was  more  important  than 
who  passed  it.  In  other  words,  as  illustrated  below, 
what  you  want  to  emphasize  dictates  whether  you 
use  the  active  or  passive  voice: 

Many  technical  reports  are  prepared  by  con- 
sultants. 

Consultants  prepare  many  technical  reports. 

Use  Pronouns.  Correctly  used,  pronouns  pro- 
vide a  succinct  way  of  making  your  points.  The  CBE 
Style  Manual  (CBE  Style  Manual  Committee  1983) 
endorses  "I  discovered  . . ."  rather  than  "It  was  dis- 
covered . . ."  and  calls  the  second  example  the  "pas- 
sive of  modesty."  Avoid  such  passive  voice  construc- 
tions when  your  name  or  the  agency's  name  goes  on 
the  manuscript.  By  using  personal  pronouns  you 
write  in  active  voice  and  you  make  it  clear  who  did 
what  or  who  recommended  an  action. 


820 


Written  Communication 


Replace  Small,  Weak  Verbs.  Many  writers 
build  their  sentences  around  small  verbs,  is,  are,  was, 
were,  and  related  forms.  Sometimes  writers  use  trap- 
ped verbs  as  shown  in  Figure  9. 


Noun 

Verb  Form 

refusal 

refuse 

promulgation 

promulgate 

assignment 

assign 

reliance 

rely 

verification 

verify 

happening 

happen 

occurrence 

occur 

participation 

participate 

disclosure 

disclose 

Figure  9.  Common  nouns  representing  trapped 
verbs.  Adapted  from  and  used  by  permission  of 
Felker  et  al.  (1982). 


Eliminate  Unneeded  Clauses.  Long,  complex 
sentences  often  reflect  loose,  rambling  writing.  Most 
readers  find  long  sentences  difficult  to  follow.  Can 
the  following  sentences  be  written  more  succinctly? 

It  is  the  opinion  of  field  personnel  in  the  South 
Dakota  Department  of  Game,  Fish,  and  Parks 
that  one  of  the  primary  factors  influencing  pop- 
ulation levels  of  big  game  and  turkeys  in  the 
southern  Black  Hills  is  a  limited  water  supply. 

How  can  we  tighten  the  sentence?  Consider  the  first 
clause,  "It  is  the  opinion  of . . ."  Why  not  replace  the 
clause  with  "believe."  Thus 

"Field  personnel  in  the  South  Dakota  Depart- 
ment of  Game,  Fish,  and  Parks  believe  . . ." 

Looking  further,  let's  replace  ". . .  that  one  of  the  pri- 
mary factors  influencing . . ."  with  ". . .  restricts  . . ." 
Recasting  the  sentence  gives 

"Field  personnel  in  the  South  Dakota  Depart- 
ment of  Game,  Fish,  and  Parks  believe  that  a 
limited  water  supply  restricts  the  population 
levels  of  big  game  animals  and  turkeys  in  the 
southern  Black  Hills." 

I  have  trimmed  10  words,  nearly  25%,  from  the  orig- 
inal 41 -word  sentence.  Later,  when  I  discuss  vague, 
general,  and  abstract  words,  I  show  how  to  tighten 
the  passage  by  being  specific. 


To  find  trapped  verbs,  look  for  words  ending  in: 
-al,  -tion,  -ence,  and  -ure  (Felker  et  al.  1982)  or  look 
for  nouns  near  such  verb  forms  as  effected,  obtained, 
performed,  produced,  required,  carried  out,  imple- 
mented, accomplished,  occurred  (Brandner  et  al. 
1974). 

Consider  "The  area  is  a  typical  representation  of 
.  .  ."  versus  "The  area  represents . . ."  or  "Excessive 
eutrophication  of  aquatic  ecosystems  by  livestock 
defecation  results  in  pollution  and  bacterial  contami- 
nation . . ."  versus  "Livestock  defecation  pollutes  and 
produces  excessive  eutrophication  of  aquatic  ecosys- 
tems and  contaminates  them  with  bacteria . . ." 

Compare  "I  have  made  general  recommenda- 
tions for  maintaining  and  improving  each  group  of 
reptile  and  amphibian"  versus  "I  recommend  ways  of 
maintaining  and  improving  each  reptile  group."  And 
consider  "It  was  the  purpose  of  this  study  to  supple- 
ment previous  work  completed  for  the  Bill  Williams 
drainage  (Kepner  1979)  in  addition  to  describing 
the  aquatic  resources  of  the  Hassayampa  River  drain- 
age." versus  "Our  study  supplements  the  work  of 
Kepner  (1979)  on  the  Bill  Williams  drainage  and  de- 
scribes the  aquatic  resources  of  the  Hassayampa 
River  drainage." 


Eliminate  Unneeded  Phrases.  Too  many 
prepositional  phrases  in  a  sentence  clutter  your  writ- 
ing. Often  you  can  strike  out  unneeded  prepositional 
phrases.  Look  with  suspicion  on  prepositional 
phrases  ending  sentences.  Are  they  really  needed? 
Do  they  repeat  a  point  made  earlier?  If  so,  eliminate 
them.  Consider 

"We  tagged  25  mule  deer  in  this  study."  versus 
"We  tagged  25  mule  deer." 

Change  prepositional  phrases  to  modifiers.  Consider 

"The  evaluation  report  of  the  BLM  . . ."  versus 
"The  BLM  evaluation  report . . ." 


or 


"The  population  level  of  big  game  and  tur- 
keys ..."  versus  "The  big  game  and  turkey  pop- 
ulation levels ..." 

Or  replace  prepositional  phrases  with  possessives. 
Consider  "The  report  of  the  division  biologist . . ." 
versus  "The  division  biologist's  report  . . ." 


Written  Communication 


821 


Or  change  prepositional  phrases  to  adverbs. 
Consider 

"The  rattlesnake  crawled  in  a  slow  manner 
across  . . ."  versus  "The  rattlesnake  crawled 
slowly  across ..." 

Omit  Needless  Words.  Look  for  words  you  can 
omit.  Crossing  them  out  often  improves  sentences. 
Consider 


in  order  to 

in  the  same  general  area 

brown  in  color 


to 

in  the  same  area 

brown 


Figure  10,  from  the  CBE  Style  Manual  (CBE  Style 
Manual  Committee  1983),  lists  succinct  words  and 
phrases  to  replace  wordy  constructions. 

Replace  Vague,  General,  and  Abstract  Words. 

The  example  of  eliminating  unnecessary  clauses  re- 
ferred to  "field  personnel."  What  did  the  writer 
mean?  Biologists?  Technicians?  Game  protectors? 
Conservation  officers?  Replacing  "field  personnel" 
with  "biologists"  leaves  no  doubt  who  believed  what: 

"Biologists  in  the  South  Dakota  Department  of 
Game,  Fish  and  Parks  believe  that  a  limited 
water  supply  restricts  the  levels  of  big  game 
animals  and  turkeys  in  the  southern  Black 
Hills." 

Specifying  the  big  game  animals — deer  and  bear — 
would  further  clarify  the  point. 

If  you  use  vague,  general,  and  abstract  terms, 
you  leave  the  interpretation  to  the  reader,  and  the 
reader  usually  has  a  different  frame  of  reference. 
Therefore,  replace  abstract  words  with  concrete 
words.  Replace  vague  words  with  specific  terms.  Re- 
place general  words  with  precise  words. 

Figure  1 1  lists  examples.  If  you  do  not  agree, 
your  interpretation  provides  even  stronger  evidence 
for  using  specific,  precise,  concrete  terms. 

Avoid  Jargon  and  Acronyms.  Although  you 
understand  the  jargon  of  your  speciality,  your  read- 
ers may  not.  Thus,  "Palustrine  wetlands"  may  convey 
a  precise  meaning  to  a  waterfowl  ecologist,  but 
other  biologists  may  call  such  areas  "marshes"  or 
"swamps."  Furthermore,  your  readers  may  not  define 
the  terms  as  you  do.  If  you  use  jargon,  define  your 
terms.  As  you  reread  your  manuscript,  ask  yourself  if 
your  readers  know  the  terms  you  use.  If  they  know 
the  terms,  do  they  define  the  terms  as  you  do? 

Acronyms  create  problems.  One  natural  re- 
source agency  hires  CIA  agents — not  the  Central  In- 
telligence Agency  type,  but  conservation  information 


assistants.  Your  varied  readers  may  not  know  or  un- 
derstand acronyms  you  use.  If  you  use  acronyms,  de- 
fine them  following  their  first  mention  such  as 
". . .  Habitat  Management  Plan  (HMP)  . . ." 

Some  readers  will  not  be  familiar  with  the  BLM 
or  other  Government  agencies.  In  addition,  many 
readers  will  not  know  your  specialized  field  and  the 
acronyms  you  know  so  well. 

Editing  for  Style  and  Grammar  Errors 

Read  through  your  document  checking  for 
grammar  and  style  errors.  Keep  in  mind  that  differ- 
ent agencies  and  publications  have  different  style 
rules.  For  most  technical  manuscripts,  follow  the 
current  Journal  of  Wildlife  Management  style.  To  an- 
swer specific  questions,  see  the  current  CBE  Style 
Manual  (CBE  Style  Manual  Committee  1983),  the 
Chicago  Manual  of  Style  (University  of  Chicago  Press 
1983),  or  the  U.S.  Government  Printing  Office  Style 
Manual  (U.S.  Government  Printing  Office    1984). 

If  you  are  preparing  publications  for  general  au- 
diences, check  with  the  publication's  editor  to  learn 
the  publication's  style.  Many  newspapers  and  maga- 
zines follow  the  Associated  Press  Stylebook  (Angione 
1977)  or  their  own  style  manuals. 

Periodically,  style  rules  change  and  agencies  and 
publications  issue  new  style  manuals.  If  your  style 
manuals  are  more  than  5  years  old,  check  to  see  if 
new  editions  have  been  issued. 

Correcting  Spelling  Errors 

Read  through  your  manuscript  checking  for 
spelling  errors.  When  in  doubt,  look  up  the  word.  If 
you  cannot  find  the  spelling  in  a  desk  dictionary,  use 
an  unabridged  dictionary.  The  Journal  of  Wildlife 
Management  and  CBE  Style  Manual  instruct  authors 
to  follow  Webster's  Third  International  Dictionary, 
unabridged  (Gove  1966).  Other  style  manuals  and 
publications  may  follow  other  unabridged  dictionar- 
ies. As  a  matter  of  style,  dictionaries  do  spell  some 
words  differently.  Be  especially  careful  with  hyphen- 
ated words.  Because  the  language  changes,  publish- 
ers issue  new  editions  of  unabridged  dictionaries,  so 
your  office  may  need  to  buy  a  new  unabridged 
dictionary. 

Editing  Illustrations — Tables  and  Figures 

Make  one  trip  through  your  manuscript  check- 
ing tables  and  figures.  Make  sure  they  are  numbered 
sequentially  through  the  manuscript.  Tables  have 
one  sequential  numbering  and  the  figures  have  an- 
other sequential  numbering.  Make  sure  your  narra- 
tive's table  and  figure  references  are  correct  and 


822 


Written  Communication 


Wordy 


Concise 


in  the  vicinity  of near 

in  view  of  the  fact  that because 

it  is  often  the  case often 

it  is  possible  that  the  cause  is the  cause  may  be 

it  is  this  that this 

it  would  thus  appear  that apparently 

large  numbers  of many 

lenticular  in  character lenticular 

masses  are  of  large  sizes masses  are  large 

necessitates  the  inclusion  of needs,  requires 

of  such  hardness  that so  hard  that 

on  the  basis  of from,  by,  because 

oval  in  shape,  oval-shaped oval 

plants  exhibited  good  growth  plants  grew  well 

prior  to  (in  time) before 

serves  the  function  of  being is 

subsequent  to after 

the  fish  in  question this  fish 

the  tests  have  not  as  yet the  tests  have  not 

the  treatment  having  been  performed  after  treatment 

there  can  be  little  doubt  that  this  is this  probably  is 

throughout  the  entire  area throughout  the  area 

throughout  the  whole  of  the  experiment throughout  the  experiment 

two  equal  halves halves 

a  number  of few,  many,  several 

an  innumerable  number  of  tiny  veins innumerable  tiny  veins 

as  far  as  our  own  observations  are  concerned, 

they  show we  observed 

ascertain  the  location  of find 

at  the  present  moment,  at  this  point  in  time now 

bright  green  in  color bright  green 

by  means  of by,  with 

(We)  conducted  inoculation  experiments  on inoculated 

due  to  the  fact  that  because 

during  the  time  that while 

fewer  in  number fewer 

for  the  purpose  of  examining to  examine 

for  the  reason  that because,  since 

from  the  standpoint  of according  to 

goes  under  the  name  of is  called 

if  conditions  are  such  that   if 

in  all  cases always,  invariably 

in  order  to to 

in  the  course  of during 

in  the  event  that if 

in  the  near  future soon 


Figure  10.  Replace  wordy  constructions  with  concise  terms.  From  CBE  Style  Manual  (CBE  Style  Manual 
Committee  1983)  with  permission. 


Written  Communication 


823 


Vague,  General,  Abstract  Terms 


Specific,  Concrete,  Precise  Terms 


small  pond a  4-hectare  pond 

a  few  hens  in  the  flock five  hens  in  the  flock 

many  grass  species   10  grass  species 

steep  terrain 60°  slopes 

low  biomass 10  kilograms 

a  hot  day 40°  C  with  95%  humidity 

strong  winds 55-miles-per-hour  winds 

several  days   six  days 

considerable  disparity  precipitation  ranged 

in  monthly  precipitation from  0  to  20  cm  monthly 

diminutive  in  size  less  than  8  mm 

significant P  <0.05 

often  used daily,  weekly,  monthly 

mildly  acidic report  pH  level 


Figure  11.  Suggestions  for  replacing  vague,  general,  and  abstract  terms  with  specific,  concrete,  and  precise 
terms. 


where  you  want  them  inserted  is  clearly  marked.  Be 
sure  the  insertions  follow  their  first  mention  in  the 
manuscript. 

Carefully  check  the  tables  and  figures  for  con- 
tent, style,  spelling,  and  typographical  errors.  Num- 
bers are  easily  transposed,  so  recheck  every  number 
in  every  table  and  figure  in  the  final  draft.  Make  sure 
your  calculations  are  correct.  Check  row  and  col- 
umn entries  because  they  are  easily  inverted  and 
transposed.  Make  sure  no  information  has  been 
deleted. 

Having  the  Manuscript  Retyped 

Once  you  have  made  the  changes  you  want, 
have  your  manuscript  retyped.  When  you  work  with 
typists  and  word  processing  operators,  provide  the 
publication's  layout  and  format  guidelines.  They  give 
instructions  for  setting  margins,  beginning  the  first 
page,  numbering  pages,  and  placing  information  on 
later  pages.  Some  publication  guidelines  may  call  for 
not  breaking  sentences  and  paragraphs  between 
pages. 

If  your  manuscript  was  originally  entered  on  a 
word  processor,  check  with  the  operator  on  how  to 
mark  changes.  Some  operators  want  changes  marked 
in  brightly  colored  felt-tip  pens  to  speed  up  their 
work.  If  you  delete  or  add  matter,  find  out  how  the 
operator  wants  you  to  do  so.  Follow  the  operator's 
instructions  or  you  will  slow  their  work  of  correct- 
ing the  manuscript. 


Copyediting  the  Manuscript 

After  retyping  the  manuscript  or  revising  it  on 
the  word  processor,  carefully  copyedit  it.  Look  for 
spelling,  grammar,  and  style  errors.  Check  to  make 
sure  that  nothing  has  been  deleted  or  added.  Be- 
cause you  have  worked  closely  with  your  manu- 
script, you  may  have  difficulty  spotting  errors.  You 
will  read  things  in  and  you  will  read  things  out.  By 
being  familiar  with  the  manuscript,  you  may  over- 
look errors  and  details.  To  overcome  that  problem, 
have  a  competent  peer,  secretary,  or  editor  check 
your  manuscript.  But  also  read  the  manuscript  your- 
self. 

Reading  a  manuscript  aloud  and  slowly  checking 
works  well  for  some  authors.  Others  read  their  man- 
uscripts into  a  tape  recorder  and  play  the  tape  back. 
As  the  tape  plays,  they  check  spelling,  grammar,  and 
style.  If  you  find  errors,  correct  them  with  the  stan- 
dard copyediting  symbols. 

Depending  on  the  editor's  guidelines,  you  may 
submit  a  manuscript  with  a  few  copyediting  marks, 
or  you  may  be  required  to  submit  a  clean  copy. 
Manuscripts  on  word  processors  are  easily  changed. 
But  carefully  copyedit  the  revised  pages  again;  errors 
may  creep  in.  Likewise,  when  you  have  a  page  re- 
typed, check  for  new  errors  that  may  have  been 
made  in  retyping. 

Seeking  Peer  Reviews 

Most  manuscripts  benefit  from  competent  peer 


824 


Written  Communication 


reviews.  The  number  of  reviews  you  seek  depends 
upon  your  manuscript's  length,  the  content,  and  the 
manuscript's  importance  to  the  resource  and  the 
agency.  The  more  important  the  content,  the  more 
reviews  your  manuscript  will  need.  Sometimes  you 
will  need  only  three  or  four  reviews  and  in  other  in- 
stances you  may  need  more. 

When  you  seek  peer  reviews,  you  are  asking  for 
other  professionals  to  view  your  manuscript  from 
their  perspectives.  No  two  reviewers  will  have  ex- 
actly the  same  comments.  Some  comments  will  be 
superficial;  others  will  provide  detailed,  constructive 
criticism.  Some  reviewers  will  agree  with  you;  others 
will  not. 

When  you  read  the  reviews,  keep  in  mind  that 
some  reviewers'  comments  will  be  extremely  helpful 
and  fair;  others  will  not.  When  you  receive  a  superfi- 
cial, uncritical  review,  seek  another  review.  When 
you  receive  a  highly  critical,  harsh  review,  consider 
the  reviewer's  comments  carefully.  Some  reviewers 
may  disagree  with  you  or  be  threatened  by  your 
conclusions.  Such  reviewers  may  be  unfair  and  bring 
irrelevant  objections.  So  weigh  each  reviewer's  com- 
ments carefully,  but  do  not  be  intimidated  by  their 
comments. 

Keep  in  mind  your  agency's  internal  review  pol- 
icy. Such  policies  prevent  erroneous  statements  oc- 
curring in  agency  documents.  If  your  agency  has  a 
review  policy,  follow  it. 

In  some  instances,  you  will  find  yourself  having 
to  make  changes  reviewers  suggest  before  your  man- 
uscript is  approved  for  publication.  Your  editor,  and 
perhaps  your  supervisor,  will  provide  helpful 
guidelines. 

As  you  rewrite  and  revise,  work  carefully.  Incor- 
porate the  needed  revisions  and  carefully  copyedit 
your  manuscript  for  errors.  Check  every  number  in 
every  table  again.  Remember  changes  made  now  are 
easier  to  make  and  less  costly  and  time-consuming 


than  changes  made  after  the  final  typing  or  typeset- 
ting the  manuscript. 

PRINTING,  COMPUTERS,  AND 
MANUSCRIPTS 

Much  of  the  time,  your  manuscripts  will  be 
printed  in  their  typewritten  form.  But  occasionally,  a 
manuscript  may  be  typeset  and  you  may  see  it 
through  the  printing  process.  Before  you  do,  learn 
the  basic  steps  in  the  printing  process  and  when  and 
where  to  make  changes.  The  CBE  Style  Manual  (CBE 
Style  Manual  Committee  1983)  contains  three  excel- 
lent chapters  on  printing  and  publishing.  Other  help- 
ful references  include  Turnbill  and  Baird  (1980)  and 
Hill  and  Cochran  (1979). 

The  rapid  advances  of  computers  in  the  printing 
industry  in  the  1970s  and  1980s  and  personal  com- 
puters in  the  1980s  ushered  in  the  Information  Age. 
Software  packages  enable  word  processing  on  most 
personal  computers.  Other  software  packages  check 
spelling,  grammar,  and  style.  And  computers  can  be 
linked  to  phototypesetters. 

You  may  find  yourself  more  involved  with  print- 
ing than  you  envision.  Preparing  manuscripts  for 
typesetting  will  soon  require  entering  specific  com- 
mands on  your  personal  computer.  You  will  need  to 
enter  such  information  as  the  type  face,  size,  spacing, 
column  width,  and  other  typesetting  specifications. 
At  this  point  you  can  call  your  printer  and  transmit 
the  manuscript  by  telephone  to  the  phototypesetter. 

The  Information  Age  will  change  how  we  pro- 
duce documents.  The  equipment  on  which  we  work 
may  change,  but  many  of  the  steps  will  remain  the 
same.  Equipment  and  software  will  make  some  tasks 
easier,  but  you  will  still  be  writing  and  checking 
your  work.  The  Information  Age  will  encourage  the 
proliferation  of  manuscripts,  reports,  and  other  docu- 
ments. Therefore,  it  will  become  more  and  more  im- 
portant to  target  your  communications  to  specific 
readers  and  to  write  clearly  and  concisely. 


Written  Communication 


825 


LITERATURE  CITED 

ALLEN,  A.  1977.  Steps  toward  better  scientific  illustrations. 
Allen  Press.  Lawrence,  KS.  64pp. 

ANGIONE,  E.,  ed.  1977.  The  Associated  Press  stylebook 

and  libel  manual.  The  Associated  Press.  50  Rockefeller 
Plaza.  New  York,  NY  10020.  276pp. 

BRANDNER,  L.O.,  A.  BIDWELL,  and  I.  TEARE.  1974.  Let 
logic  guide  your  writing.  Contribution  1 36,  Vice- 
President  for  Agriculture  Office,  and  1431,  Depart- 
ment of  Agronomy,  Agricultural  Experiment  Station, 
Kansas  State  University.  Also  published  in  September- 
October  1974,  J.  of  Soil  and  Water  Cons.  29(5):235- 
238. 

CBE  STYLE  MANUAL  COMMITTEE.  1983.  CBE  style 

manual:  a  guide  for  authors,  editors,  and  publishers  in 
the  biological  sciences.  5th  edition,  revised  and  ex- 
panded. Council  of  Biology  Editors.  Bethesda,  MD. 
324pp. 

CHANDLER,  HE.  1978.  The  "how  to  write  what"  book. 
American  Society  for  Metals.  Metals  Park,  OH.  92pp. 

FELKER,  D.B.,  F.  PICKERING,  V.R.  CHARROW,  V.M.  HOL- 
LAND, and  J.C.  REDISH.  1982.  Guidelines  for  docu- 
ment designers.  American  Institutes  for  Research. 


1055  Thomas  Jefferson  St.  NW,  Washington,  DC 
20007.  117pp. 

GOVE,  P.B.,  ed.  1966.  Webster's  third  new  international 
dictionary,  unabridged.  G  &  C  Merriam  Co.,  Spring- 
field, MA.  2662pp. 

HILL,  M.  and  W.  COCHRAN.  1979.  Into  print:  a  practical 
guide  to  writing,  illustrating  and  publishing.  William 
Kaufman,  Inc.  One  First  Street,  Los  Altos,  CA  94022. 
175pp. 

McKEE,  B.  1975.  Types  of  outlines  used  by  technical  writ- 
ers. J.  Eng.  Teaching  Techniques.  17:30-36. 

SCHRAMM,  W.  and  D.F.  ROBERTS.  1972.  The  process  and 
effects  of  mass  communication.  Revised  edition.  Univ. 
of  Illinois  Press.  Urbana.  997pp. 

TURNBILL,  AT.  and  R.N.  BAIRD.  1980.  The  graphics  of 
communication.  4th  ed.  Holt,  Rinehart,  Winston.  New 
York,  NY.  398pp. 

UNIVERSITY  OF  CHICAGO  PRESS.  1983.  The  Chicago 
style  manual.  1 3th  edition.  Univ.  of  Chicago  Press. 
738pp. 

U.S.  GOVERNMENT  PRINTING  OFFICE.  1984.  The  U.S. 
Government  Printing  Office  style  manual.  Revised 
edition.  Superintendent  of  Documents.  U.S.  Govern- 
ment Printing  Office,  Washington,  DC.  479pp. 


826 


Written  Communication 


APPENDIX. 


Suggestions  for  a  Professional  Library 


From  time  to  time,  you  will  have  questions  about 
writing,  mechanics,  spelling,  and  style.  Professional 
writers  and  editors  regularly  have  questions  too.  To 
answer  their  questions,  they  turn  to  their  office  li- 
brary of  books  on  writing,  data  presentation,  illustra- 
tions, printing,  grammar,  style,  and  usage.  Below 
are  suggested  publications  for  your  office  library. 


Books  on  Writing 

CHANDLER,  H.  1978.  The  "how  to  write  what"  book. 

American  Society  of  Metals,  Metals  Park,  OH  44073- 

92pp. 
ROMAN,  K.  and  J.  RAPHAELSON.  1981.  Writing  that 

works.  Harper  and  Row.  New  York,  NY.  105pp. 
ZINSSER,  W.  1983.  Writing  with  a  word  processor.  Harper 

and  Row.  New  York,  NY.  1 1 7pp. 
.  1980.  On  writing  well.  Harper  and  Row.  New 

York,  NY.  187pp. 


U.S.  GOVERNMENT  PRINTING  OFFICE.  1984.  GPO  style 
manual.  Revised  edition.  U.S.  Government  Printing 
Office.  Washington,  DC.  479pp. 

Data  Presentation  and  Illustrations 

ALLEN,  A.  1977.  Steps  toward  better  scientific  illustrations. 
Allen  Press,  Inc.  Lawrence,  KS.  64pp. 

ENRICK,  N.L.  1980.  Handbook  of  effective  graphic  and 
tabular  communication.  Robert  E.  Krieger  Publishing 
Company.  Huntington,  NY.  21 4pp. 

REYNOLDS,  L.  and  D.  SIMMONDS.  1982.  Presentation  of 
data  in  science.  Martinus  Nijhoff  Publishers.  The  Ha- 
gue and  Boston.  209pp. 

Printing 

HILL,  M.  and  W.  COCHRAN.  1979.  Into  print:  a  practical 
guide  to  writing,  illustrating,  and  publishing.  William 
Kaufman,  Inc.  One  First  Street,  Los  Altos,  CA.  94022. 
175pp. 

TURNBILL,  AT.  and  R.N.  BAIRD.  1980.  The  graphics  of 
communication.  Fourth  Edition.  Holt,  Rinehart,  and 
Winston.  New  York,  NY.  398pp. 


Style  Guides 

ANGIONE,  H.,  ed.  1977.  The  Associated  Press  style  book 
and  libel  manual.  The  Associated  Press.  50  Rockefeller 
Plaza.  New  York,  NY  10020.  276pp. 

CBE  STYLE  MANUAL  COMMITTEE.  1983  CBE  style 

manual:  a  guide  for  authors,  editors  and  publishers  in 
the  biological  sciences.  5th  edition,  revised  and  ex- 
panded. Council  of  Biology  Editors,  Inc.  Bethesda, 
MD.  324pp. 

UNIVERSITY  OF  CHICAGO  PRESS.  1983.  The  Chicago 
style  manual.  1 3th  edition.  University  of  Chicago 
Press.  Chicago.  738pp. 


Grammar,  Style,  and  Usage  References 

BERNSTEIN,  T.M.  1978.  The  careful  writer:  a  guide  to 
modern  usage.  Atheneum.  New  York,  NY.  48"?pp. 

.  1971.  Miss  Thistlebottom's  hobgoblins.  Farrar, 

Straus,  and  Giroux.  New  York,  NY.  260pp. 

CREWS,  F.  1979.  The  Random  House  guide.  Random 
House.  New  York,  NY.  434pp. 

O'HAYRE,  JO.  No  Date.  Gobbledygook  has  gotta  go.  Su- 
perintendent of  Documents.  U.S.  Government  Printing 
Office.  Washington,  DC.  20402.  112pp. 

STRUNK,  W.  and  E.B.  WHITE.  1979.  The  elements  of  style. 
Third  edition.  MacMillan  Publ.  Co.  Inc.  New  York, 
NY.  85pp. 


Written  Communication 


827 


42 

VERBAL 
PRESENTATIONS 

Eugene  Decker 


Department  of  Fishery  and  Wildlife  Biology 
Colorado  State  University 
Fort  Collins,  CO  80521 


'Silence  is  the  virtue  of  fools. 


-Francis  Bacon 


Editor's  Note:  Communication  should  not  end  with 
a  written  report.  Unfortunately,  many  of  us  avoid 
speaking  to  groups  of  people.  We  must  overcome 
this  tendency  if  wildlife  is  to  be  adequately  repre- 
sented in  multiple-use  decisions  and  allocations. 

Proper  planning,  practice,  and  advance  work  will 
greatly  improve  the  presentations  you  give,  while 
increasing  understanding  and  acceptance  of  your 
results. 


INTRODUCTION 

Communication  has  been  defined  as  "the  pro- 
cess of  informing  and  being  informed;  it  implies  un- 
derstanding." In  other  words,  if  there  is  no  under- 
standing, there  is  no  communication.  Personal 
presentations  are  often  the  best  means  to  inform  an 
audience,  as  they  provide  a  better  opportunity  to 
communicate  than  do  written  communications. 

A  speaker  usually  receives  immediate  reaction 
from  the  audience  and  can  appropriately  respond.  In 
addition  to  directly  answering  questions,  the  speaker 
can  respond  to  comments,  facial  expressions,  atten- 
tion levels,  gestures,  and  other  signals  from  the  audi- 
ence during  the  presentation.  Such  responses  can 
quickly  clarify  points  not  understood  by  the 
audience. 

Another  advantage  of  personal  presentations  is 
that  a  good  speaker  can  convey  credibility,  sincerity, 
concern,  and  emotion.  This  personal  factor  can  be 
most  effective  in  getting  a  message  across  to  a  gen- 
eral public  audience. 

A  verbal  presentation  of  a  management  plan  by 
a  member  or  members  of  the  resource  staff  may 
often  determine  the  acceptance  of  the  plan  by  other 
staff,  administrators,  or  special  interest  groups.  In 
fact,  the  success  or  failure  of  many  months  of  effort 
by  team  members  may  well  focus  on  a  single 
presentation. 

Unfortunately,  I  have  seen  such  team  efforts 
frustrated  by  presentations  that  were  poorly  pre- 
pared and  communicated.  It  is  no  wonder  that  dull, 
highly  technical,  jargon -filled  presentations  illustrated 
with  unreadable  visuals  result  in  nonacceptance  of 
management  plans.  This  chapter  is  intended  to  assist 
personnel  in  preparing  and  delivering  effective  per- 
sonal presentations  which  I  hope  will  result  in  sound 
resource  management. 


SPEAKER  CREDIBILITY 

Several  factors  affect  your  image  as  a  speaker  to 


Verbal  Presentation 


829 


the  audience,  some  even  before  the  talk  begins.  Your 
appearance  and  behavior  at  a  meeting  before  the 
talk,  support  and  endorsement  by  members  of  the 
group,  and  a  concise  introduction  by  the  host  stating 
your  qualifications,  all  help  establish  you  as  a 
professional. 

Several  factors  during  your  talk  can  also  affect 
your  credibility.  First,  knowledge  of  the  subject  mat- 
ter is  of  utmost  importance.  Due  to  constant  expo- 
sure to  media,  Americans  are  a  sophisticated  audi- 
ence and  can  easily  spot  a  phoney.  A  speaker  should 
not  try  to  discuss  subjects  or  answer  questions  be- 
yond his  or  her  expertise. 

The  second  factor  is  trust,  especially  with  spe- 
cial interest  groups.  An  introduction  by  a  host  who 
endorses  you  will  help  establish  confidence  in  the 
audience.  Also,  you  should  not  avoid  a  controversial 
issue,  especially  if  you  know  your  audience  is  al- 
ready aware  of  it.  If  there  are  other  sides  to  an  issue, 
be  sure  to  acknowledge  them  before  individuals  in 
the  audience  bring  them  up;  admit  earlier  mistakes, 
never  try  to  hide  them. 

Dynamism  is  the  third  factor.  Expressing  inter- 
est, concern,  and  enthusiasm  in  a  well-delivered 
presentation  can  greatly  increase  your  effectiveness. 
However,  if  the  first  two  factors,  knowledge  and 
trust,  are  weak,  then  it  will  make  little  difference 
how  dynamic  you  are.  But,  if  your  qualifications  and 
trust  are  sound,  an  energetic,  interesting  delivery 
will  greatly  enhance  your  acceptance  by  the 
audience. 


3.  Another  problem  is  use  of  words  that  convey 
different  meanings  to  different  people.  Such 
words  as  ecosystem,  carrying  capacity,  environ- 
ment, ecology,  conservation,  habitat,  and  wilder- 
ness have  broad  meanings.  Another  confusing 
term  often  used  is  resources.  I  heard  an  agency 
person  tell  an  audience  that  ". . .  we  have  diffi- 
culties managing  your  resources  as  Congress  did 
not  give  us  the  financial  resources  for  the  re- 
sources necessary  to  do  the  job."  You  can  imag- 
ine the  audience  reaction! 


4.  The  environment  of  the  meeting  site  can  cause 
problems.  The  audience  will  not  give  you  its  full 
attention  if  the  room  temperature  is  uncomfort- 
able, the  air  stuffy  due  to  poor  circulation,  if 
there  are  distracting  materials  on  the  walls  or 
blackboard,  disturbing  noises,  or  the  chairs  are 
positioned  awkwardly  in  reference  to  the 
speaker  or  screen.  You  should  correct  as  many 
of  these  problems  as  possible  before  the  audi- 
ence arrives. 

5.  The  audience  condition  is  also  an  important  fac- 
tor. You  will  have  problems  maintaining  audi- 
ence attention  if  people  have  been  seated  too 
long  without  a  break,  if  you  have  run  into  the 
coffee  break  or  lunch  hour,  if  you  are  the  last 
speaker  of  a  4-hour  series,  or  if  it  is  very  late  in 
the  afternoon  or  evening.  Be  aware  of  these  situ- 
ations and  try  to  correct  and  adjust  by  shorten- 
ing your  talk,  giving  them  a  break,  or  trying  to 
be  more  dynamic  and  interesting. 


BARRIERS  TO  COMMUNICATION 

A  number  of  problems  can  interfere  with  your 
audience's  understanding  of  your  messages.  Below 
are  several  common  ones: 

1.  Use  of  jargon  is  a  major  problem  with  many 
professionals.  Jargon,  bureaucratic  buzz  words, 
in-terms,  and  acronyms  (HEP,  HMP,  NEPA,  AUM, 
etc.)  are  used  to  save  time  when  communicat- 
ing with  colleagues,  but  are  usually  unknown  to 
the  general  public.  Use  of  such  terms  causes 
confusion  and  often  loss  of  attention  by  an 
audience. 

2.  Technical  terms,  scientific  names,  and  other  for- 
mal terminology  are  part  of  a  professional's  vo- 
cabulary, but  these  terms  are  not  commonly 
used  by  the  general  public.  Avoid  these  terms  if 
possible,  but  if  you  have  to  use  one,  be  sure  to 
define  it.  As  an  example,  the  term  "forb"  is  regu- 
larly used  by  agency  personnel,  but  it  should  be 
defined  for  the  general  public. 


6.  The  cultural,  social,  and  educational  aspects  of 
your  audience  must  also  be  considered.  You 
should  be  aware  of  their  history,  experiences, 
education,  and  values  in  order  for  you  to  effec- 
tively convey  your  message  to  them.  It  is  impor- 
tant that  your  presentation  relate  to  some  aspect 
of  their  lives  to  be  clearly  understood.  An  urban 
audience  (service  club,  naturalist  group,  etc.) 
may  have  problems  relating  to  issues  of  eco- 
nomic survival  of  ranchers  whereas  a  rural  audi- 
ence may  not  understand  an  integrated  manage- 
ment plan  to  meet  increasing  outdoor 
recreation  needs  by  city  residents. 


PREPARING  FOR  THE  PRESENTATION 

I  offer  the  following  suggestions  to  assist  you  in 
preparing  for  personal  presentations,  particularly 
slide  talks,  to  special  interest  groups  and  the  general 
public.  They  will  be  covered  in  four  stages:  plan- 
ning, preparation,  presentation,  and  evaluation. 
Presentations  to  agency  staff  and  management  level 
personnel  are  discussed  later. 


830 


Verbal  Presentation 


Planning 

Once  you  have  been  asked  or  assigned  to  give  a 
presentation  you  should  consider  the  following: 

1 .  Decide  on  the  purpose  of  your  presentation  and 
write  it  in  a  clear,  concise  statement  (for  exam- 
ple, "to  inform  the  Red  Rock  Livestock  Associa- 
tion of  wildlife  improvements,"  or  "to  explain 
results  of  management  alternatives  on  the  ante- 
lope herd,"  etc. ).  This  will  help  give  direction  to 
your  subsequent  efforts.  In  general,  your  pur- 
pose will  be  to  produce  a  change  in  your  audi- 
ence. This  may  be  one,  or  a  combination,  of  the 
following:  ( 1 )  awareness,  (  2  )  knowledge,  (  3  )  at- 
titude about  a  situation,  or  (4)  behavior,  i.e., 
support  or  action  related  to  a  situation. 

2.  Prepare  an  audience  analysis.  Determine  as 
much  as  possible  about  your  audience  before 
the  meeting.  Such  things  as  education  levels,  ex- 
posure to  previous  speakers,  age  and  sex  com- 
position, previous  experiences,  stands  on  similar 
issues,  their  leaders,  numbers  of  persons  usually 
present,  and  contact  with  other  staff  persons. 
Your  analysis  should  give  a  good  picture  of  your 
audience  from  which  you  can  tailor  your 
presentation. 

3.  Complete  a  preliminary  rough  outline  as  soon  as 
possible.  This  is  helpful  as  a  reminder  of  needed 
facts,  graphics,  and  other  visual  aids  that  can  be 
useful  to  your  talk. 

4.  Make  a  list  of  the  visual  aids  and  equipment  you 
will  need.  Color  slides  are  especially  useful  in 
describing  field  situations  (before/after  treat- 
ments, contrasts,  impacts,  etc.),  but  may  not  be 
appropriate  for  other  occasions  where  factual  in- 
formation only  is  to  be  offered.  For  these  pres- 
entations, overhead  transparencies  or  a  flip  chart 
alone  may  be  appropriate. 

Handouts  are  useful  for  many  occasions.  A  con- 
cise outline  of  information  may  be  helpful  to 
guide  an  audience  through  a  complex  process. 
Do  not  provide  lengthy  written  materials  during 
a  presentation,  as  some  people  will  read  it  and 
not  listen  to  you.  Such  materials  should  usually 
be  given  at  the  end  of  the  program. 

5.  Be  sure  to  put  the  date  of  the  presentation  on 
your  calendar  and  allow  time  for  preparation 
and  travel.  If  it  is  a  major  presentation,  do  not 
schedule  too  much  for  that  date,  as  a  speaker 
who  has  worked  all  day,  driven  2  hours,  and 
rushed  to  a  meeting  will  be  tired,  harried,  and 
usually  a  less  effective  speaker.  Your  presenta- 
tion should  be  the  top  priority  item  for  that  day. 


Refer  to  Appendix  I,  Checklist  for  slide  show  presen- 
tations, to  help  you  with  planning  for  your  programs. 

Preparation 

Now  that  your  planning  is  underway,  you  are 
ready  to  move  on. 

1 .  Develop  a  final  outline.  I  suggest  an  outline  that 
includes  major  items  for  your  introduction, 
main  content,  and  conclusion.  The  introduction 
to  your  talk  should  relate  the  audience  to  your 
topic.  The  introduction  should  also  tell  the  audi- 
ence what  you  will  be  covering  and  why  (this 
relates  to  your  purpose  statement  in  the  plan- 
ning phase).  You  may  also  wish  to  tell  the  audi- 
ence more  about  yourself,  your  qualifications, 
and  experience  related  to  your  topic.  A  good  in- 
troduction in  your  presentation  will  increase  the 
audience  reception  to  you  and  your  ideas. 

Some  speakers  use  5-  x  7-inch  cards  for  their 
outlines  with  just  one  or  two  brief  statements 
on  each.  (Be  sure  to  number  the  cards  to  keep 
them  in  order. )  These  cards  can  be  helpful 
when  planning  a  slide  show.  Use  one  card  for 
each  slide  and  put  the  number  of  the  slide  and  a 
simple  sketch  of  the  photo  on  the  card  with  the 
statement.  You  can  lay  these  cards  out  on  a  ta- 
ble (or  on  a  frame  "story  board")  to  help  you 
organize  your  presentation.  In  this  way  you  can 
visualize  the  whole  presentation  and  add,  re- 
move, or  rearrange  as  needed. 

Other  speakers  find  an  outline  typed  on  one  or 
two  sheets  to  be  effective.  Type  in  caps  and  tri- 
ple space  for  ease  in  reading.  Use  red  underlin- 
ing to  emphasize  important  material. 

You  should  include  a  conclusion  in  your  outline 
as  it  is  usually  a  good  idea  to  summarize  your 
talk.  You  may  then  wish  to  state  what  actions 
may  be  taken  by  the  audience  or  where  addi- 
tional information  may  be  obtained.  Be  sure  to 
have  a  meaningful  closing. 

The  outline  is  recommended  instead  of  a  verba- 
tim script  so  that  you  can  quickly  refer  to  the 
outline  and  then  talk  directly  to  the  audience. 
Few  persons  can  read  a  script  well  and  maintain 
eye  contact. 

2.  Prepare  a  brief  biographic  sketch  of  yourself  for 
the  host  to  use  in  introducing  you.  This  should 
be  a  short,  one-page  statement  of  your  educa- 
tion and  experience.  (I  find  that  typing  in  all 
caps  and  triple  spacing  makes  it  easier  to  read.) 
A  good  introduction  by  the  host  is  important,  as 
the  audience  must  know  who  you  are  and  that 
you  are  qualified  to  justify  their  attention. 


Verbal  Presentation 


831 


3.  Prepare  appropriate  and  effective  visual  aids 
(see  Visual  Aids  Section  following).  Also,  prac- 
tice using  them  with  the  equipment  that  you 
will  use  at  the  talk. 

"How  many  slides  should  I  use?"  I  believe  there 
is  no  firm  rule  as  your  subject  and  schedule  will 
usually  control  the  situation.  However,  too 
many  slides  rapidly  changed  can  provide  an  un- 
pleasant overkill.  Likewise,  it  is  not  a  good  idea 
to  spend  too  much  time  on  one  slide.  In  the  lat- 
ter situation,  use  two  or  three  similar  slides  to 
provide  some  variety  when  a  discussion  is  nec- 
essary. Good  planning  and  rehearsing  can  re- 
duce this  problem  so  that  you  have  a  clear  and 
concise  presentation. 

4.  Now  that  you  are  prepared,  take  time  to  re- 
hearse. A  small  cassette  recorder  will  provide  a 
good  tape  for  your  rehearsal.  Review  it  for  tim- 
ing, organization,  technical  terms,  jargon,  and 
verbal  distractions.  Based  on  this  review,  revise 
your  outline  and  try  again.  You  will  be  pleased 
at  the  improvement. 

Some  offices  have  portable  TV  equipment.  If  you 
can  borrow  equipment,  tape  yourself  and  then 
review  it  on  a  monitor  (TV  set).  This  can  be  a 
great  learning  experience — "to  see  yourself  as 
others  see  you." 


If  you  are  scheduled  for  a  major  presentation  in- 
volving a  potentially  controversial  subject,  I  sug- 
gest an  additional  rehearsal.  Have  several  com- 
petent colleagues  serve  as  an  audience  for  you 
and  act  as  members  of  various  interest  groups  at 
the  discussion  period  after  your  talk.  Make  your 
presentation  to  them  and  then  have  a  "no  holds 
barred"  question  and  answer  session.  Because 
they  have  probably  been  exposed  to  these  situa- 
tions, they  will  be  able  to  ask  many  slanted  and 
embarrassing  questions.  This  type  of  rehearsal 
will  give  you  a  chance  to  answer  them,  then  re- 
view how  effectively  you  replied.  Such  an  activ- 
ity can  help  you  respond  better  to  the  real  audi- 
ences, as  you  will  have  been  able  to  anticipate 
some  of  the  comments.  The  participating  staff 
can  also  help  you  with  suggestions  for  appropri- 
ate replies. 

Always  run  through  your  slides  before  the  meet- 
ing to  be  sure  that  they  are  right  side  up,  not 
backwards,  and  that  you  have  the  appropriate 
slides  in  the  cartridge.  I  also  suggest  carrying 
the  slides  with  you;  do  not  put  them  in  your 
baggage  if  you  are  flying  or  give  them  to  some- 
one else  to  bring. 


Fred  forgot  to  rehearse. 
The  audience  forgot  Fred. 


I  mentioned  earlier  that  you  should  have  pre- 
pared an  outline  for  your  presentation.  This  will 
be  most  useful  in  your  rehearsal  and  for  your  in- 
troduction and  closing,  but  the  written  outline 
of  your  major  points  will  be  difficult  to  follow  in 
a  darkened  room.  With  ample  rehearsing,  you 
should  be  familiar  enough  with  your  subject  ma- 
terial that  the  slides  serve  as  your  outline  and 


832 


Verbal  Presentation 


you  will  not  need  to  refer  to  any  written  mate- 
rial. However,  there  is  usually  enough  light  re- 
flecting from  the  screen  onto  a  correctly  placed 
podium  so  that  you  will  often  be  able  to  refer  to 
some  of  your  notes  if  necessary.  A  podium  light 
is  usually  available,  but  the  light  reflecting  from 
your  note  paper  and  hitting  the  bottom  of  your 
face  may  give  an  unpleasant  appearance. 

6.  On  the  day  of  your  presentation,  be  sure  to 
check  out  the  equipment  before  you  leave  your 
office.  Arrive  early  and  check  the  facilities,  i.e., 
seating,  podium,  heating,  ventilation,  and  light- 
ing. Make  adjustments  before  the  audience  ar- 
rives. Also,  set  up  visual  aids  and  audio  equip- 
ment and  test  them  before  the  meeting  starts. 
Position  your  podium  at  the  front  of  the  room 
so  that  you  do  not  block  the  audience's  view  of 
the  screen.  Do  not  speak  from  the  projector; 
stay  front  and  center  near  the  screen. 

If  you  are  using  a  slide  or  overhead  projector, 
make  sure  there  is  an  extra  bulb  and  practice 
changing  the  bulb.  If  a  bulb  burns  out  during 
the  program,  request  the  lights  on  and  take  a  5- 
minute  break.  This  will  give  you  an  opportunity 
to  change  the  bulb  quickly,  and  get  organized  so 
that  the  show  can  continue  smoothly.  It  is  quite 
awkward  to  change  a  bulb  when  you  do  not 
know  how  and  have  to  have  the  audience  help 
you.  Also,  never  grab  the  bulb  with  bare  fingers 
as  the  only  thing  the  audience  will  remember  is 
the  smell  of  burning  flesh  and  what  you  said 
when  you  grabbed  it. 

7.  Be  sure  to  tie  down  the  cords  from  the  projec- 
tor, both  power  and  control  cords,  to  the  stand 
or  table  leg  before  continuing  on  to  the  outlet 
or  podium.  This  prevents  people  from  tripping 
over  the  cords  or  pulling  the  projector  off  the 
table. 

8.  When  setting  up  the  equipment  before  the 
meeting,  be  sure  to  have  your  slide  image  fill 
the  screen  and  keep  it  as  high  as  possible.  If  you 
have  a  number  of  vertical  slides,  you  must  ad- 
just your  image  for  the  verticals  which  means 
that  your  horizontals  will  be  lower  on  the 
screen.  However,  if  you  have  just  a  few  verti- 
cals, you  can  usually  position  the  verticals  with 
part  of  the  top  cut  off  in  order  to  have  the  hori- 
zontals as  high  and  wide  on  the  screen  as 
possible. 

9.  Practice  with  the  microphone  before  the  audi- 
ence arrives.  Have  someone  stand  in  the  back  of 
the  room  to  advise  you  on  the  correct  volume 


level.  Adjust  the  microphone  for  your  height 
and  practice  so  that  you  know  the  correct  dis- 
tance from  "mouth  to  mike"  for  best  sound. 


Presentation 

1.  Check  your  appearance.  As  mentioned  previ- 
ously, the  impression  you  give  to  an  audience 
depends  a  great  deal  on  your  appearance,  partic- 
ularly before  the  talk  begins.  A  professional 
should  dress  as  a  professional,  which  means  that 
you  should  be  neatly  groomed  and  appropriately 
dressed.  It  is  not  what  the  audience  thinks  you 
are,  it  is  what  you  convey  to  the  audience  of 
what  you  think  of  yourself.  You  should  convey 
the  impression  that  you  are  a  professional  and 
that  you  know  you  are.  The  too  casual  approach 
can  turn  an  audience  off. 

If  at  all  possible,  try  to  take  it  easy,  leave  your 
office  early  and  plan  to  get  to  the  meeting  site 
well  ahead  of  time.  If  appropriate,  freshen  up 
and  rest  in  a  motel  before  the  meeting  in  order 
to  arrive  more  relaxed  and  less  harried. 

2.  Give  the  introduction  with  the  lights  on  before 
going  to  the  slides.  This  is  important  as  the  audi- 
ence can  become  acquainted  with  you,  your 
gestures,  and  expressions  before  the  lights  go 
off.  Provide  a  verbal  transition  to  your  slides  as 

a  clue  to  the  persons  you  have  appointed  to  op- 
erate the  projector  and  the  lights  ("Now  with 
some  slides,  I  would  like  to  show  you  . . ."  or 
"May  I  have  the  projector,  please  . . .").  The  cor- 
rect sequence  is  for  the  projector  to  go  on  and 
then  the  lights  off.  This  prevents  having  the  au- 
dience in  total  darkness  with  the  projector  oper- 
ator scrambling  to  find  the  switch. 

Likewise,  it  is  important  to  have  a  transitional 
sentence  at  the  end  as  a  clue  to  the  light  opera- 
tor to  turn  the  lights  on  and  then  the  projector 
off  ("That's  my  last  slide,  but  I  would  like  to  dis- 
cuss further . . ."  or  "May  we  have  the  lights  . . ."). 
Also,  it  is  advisable  to  have  a  meaningful  closing, 
as  described  previously,  after  the  slides,  and 
with  the  lights  on. 

3-  Change  the  slides  yourself.  Be  sure  to  have  an 
extension  cord  so  that  the  control  can  be  at  the 
podium  where  you  are.  It  is  very  awkward  to 
give  verbal  or  sound  clues  (stamping  the  floor, 
snapping  your  fingers,  etc.)  to  the  projector  op- 
erator. However,  it  is  advisable  to  have  the  pro- 
jector operator  do  the  fine  focusing  of  the  slides 
on  the  screen  as  you  will  be  too  close  to  the 
screen  to  know  whether  or  not  they  are  in 
sharp  focus. 


Verbal  Presentation 


833 


4.  Meet  your  host  and  give  him  or  her  your  bio- 
graphic sketch.  Whenever  possible,  have  your 
host  introduce  you  to  others  and  talk  with  them 
before  your  talk  if  you  are  at  a  new  place  with  a 
new  audience.  This  is  helpful  when  you  get  in 
front  of  the  group  and  find  that  there  are  a  few 
friendly  faces  in  the  audience.  The  audience  will 
also  appreciate  the  fact  that  you  took  time  to 
talk  with  some  of  their  group.  One  of  the  worst 
things  that  you  can  do  is  to  arrive  at  a  presenta- 
tion, sit  in  a  back  corner,  and  not  be  involved 
with  the  group  until  you  are  introduced.  Such 
actions  convey  the  impression  that  you  care  lit- 
tle about  them  and  are  there  just  to  do  a  job. 

5.  Maintain  eye  contact  with  the  audience.  Refer  to 
your  notes,  then  return  to  the  audience.  Try  to 
look  at  everyone  during  each  minute  or  two. 
This  is  hard  for  some,  but  practice  will  help.  Eye 
contact  is  an  important  factor  in  personal  pres- 
entation, as  it  makes  the  presentation  personal. 
The  audience  wants  to  understand  what  you  are 
telling  them  and  not  have  something  read  to 
them.  Only  by  talking  to  an  audience  can  you 
reflect  your  interests,  concern,  and  credibility. 

Do  not  talk  to  the  screen  as  this  ruins  projec- 
tion of  your  voice  for  the  people  in  the  back  of 
the  room.  Look  at  the  screen,  make  sure  you  are 
talking  about  the  right  thing,  and  turn  back  to 
the  audience  to  continue  your  presentation. 
One  way  to  break  talking  to  the  screen  is  to 
point  your  feet  toward  the  audience,  not  the 
screen. 

6.  When  using  a  pointer,  stand  to  the  side  of  the 
screen  and  lightly  touch  the  screen.  This  pre- 
vents confusion  for  the  audience  as  to  what  you 
are  pointing  at.  If  you  stand  away  from  the 
screen,  the  audience  can  see  the  pointer  plus 
the  pointer  shadow  on  the  screen  which  may  be 
indicating  different  things. 

7.  Avoid  distracting  mannerisms.  You  should  have 
picked  up  on  some  of  these  during  your  re- 
hearsals with  the  tape  recorder.  Such  things  and 
repetitious  terms  ("and  uh  . . . ,  and  . . . ,  you 
know,  this  is  a  slide,  this  is  a  typical . . .")  can  be 
most  distracting,  especially  when  the  audience 
starts  counting  them.  Excessive  or  unnatural 
movements,  hands  in  pockets,  playing  with  keys 
or  pointers,  scratching,  pacing,  can  also  distract 
an  audience  from  your  message. 

8.  Maintain  enthusiasm.  If  you  come  across  with  a 
"don't  give  a  damn"  attitude,  the  audience  will 
respond  in  the  same  way.  Monotones  or  very 
fast  delivery  can  also  turn  them  off.  You  need 
not  be  a  Johnny  Carson  to  be  effective,  but  you 


should  convey  your  interest  and  sincerity  by  a 
lively  delivery.  The  audience  will  respond  like- 
wise. An  increase  in  volume  or  an  occasional 
pause  to  emphasize  a  point  are  often  good 
techniques. 

9.  Watch  the  audience  and  be  responsive  to  them. 
They  will  convey  messages  to  you  by  gestures, 
expressions,  comments,  etc.,  and  you  will  know 
when  they  do  not  understand  a  term  or  are  not 
following  your  train  of  thought.  If  heads  start 
nodding  and  some  fall  asleep,  you  know  you 
have  lost  them. 

The  use  of  humor  can  release  tension  and  pro- 
vide a  relief  during  a  serious  or  technical  pro- 
gram. However,  the  humor  should  be  appropri- 
ate, low-key,  and  not  distracting.  There  is  no 
place  for  the  unrelated,  lengthy,  or  off-color 
joke. 

10.  Keep  on  schedule.  Your  rehearsal  should  have 
given  you  an  idea  of  the  length  of  your  talk  and 
you  should  stay  within  the  allotted  time  sched- 
ule. This  is  especially  important  if  you  are  one 
of  a  series  of  speakers  with  assigned  time  pe- 
riods. Some  speakers  put  their  watches  beside 
their  outline  on  the  podium  to  remind  them  of 
the  time.  Others  have  the  host  provide  an  inob- 
trusive  signal,  i.e.,  turn  over  a  sheet  of  paper 
with  "2  Minutes"  written  on  it  when  you  have 

,    that  time  left. 

1 1.  Instruct  your  projector  operator  to  turn  the  pro- 
jector completely  off  with  the  last  slide.  Most 
projectors  give  off  a  buzzing  sound  which  is  dis- 
tracting to  the  audience.  Also,  there  is  little  dif- 
ference in  the  life  of  your  bulb  whether  the  fan 
is  left  on  or  off. 

12.  It  is  usually  a  good  idea  to  have  agreed  before- 
hand that  the  host  will  direct  the  discussion  ses- 
sion. The  host  should  not  let  one  or  two  per- 
sons dominate  and  should  dismiss  inappropriate 
or  unrelated  comments.  This  protects  you  and 
permits  you  to  stay  on  the  subject.  Avoid  argu- 
ing as  some  persons  try  to  bait  the  speaker.  Do 
not  drag  the  discussion  session  on  unnecessarily, 
and  stop  while  the  group  is  still  attentive,  not 
when  they  start  drifting  away. 

Evaluation 

Evaluation  is  necessary  for  self-improvement  and 
to  determine  if  you  have  achieved  your  purpose. 
Most  of  this  can  be  informal.  You  should  know  if  you 
got  your  message  across  to  the  audience  by  their  at- 
tentiveness  and  overall  reactions,  comments,  and 
questions. 


834 


Verbal  Presentation 


It  is  also  advisable  to  write  a  "memo  to  the  files" 
on  your  presentation  as  soon  as  possible.  Include  the 
subject,  purpose,  date,  place,  audience  analysis,  reac- 
tion, performance  by  the  host  and  other  staff  mem- 
bers, the  quality  of  the  visual  aids,  and  suggestions 
for  improvements.  These  will  provide  reminders 
when  preparing  for  the  next  presentation  and  may 
be  helpful  to  other  staff  members  who  will  be  con- 
tacting the  same  audience  later. 

You  may  also  consider  placing  a  small  cassette 
recorder  near  the  podium  during  your  talk  so  that 
you  can  review  it  later  and  make  comments  on  the 
memo,  if  necessary. 

If  you  know  someone  in  the  audience  who 
could  objectively  review  your  presentation,  you 
might  consider  asking  them  to  do  so. 


IN-HOUSE  PRESENTATION  TO  SMALL 
GROUPS 

A  situation  often  faced  by  resource  professionals 
is  explaining  their  ideas  to  personnel  of  other  disci- 
plines or  to  high-level  managers,  administrators,  and 
other  decision  makers.  Unless  these  people  can  un- 
derstand your  views  on  impacts  of  various  alterna- 
tives, value  of  a  management  practice,  or  the  unique- 
ness of  a  special  habitat,  your  input  will  have  little 
meaningful  value.  Also,  you  must  be  effective  as  this 
may  be  the  only  chance  you  will  have  to  make  your 
case. 

The  basic  principles  discussed  previously  for 
general  public  audiences  apply  also  for  presentations 
of  this  type,  but  there  are  some  modifications  that 
should  be  considered.  A  good  general  rule  is  "the 
smaller  the  audience,  the  less  formal  the  presenta- 
tion." However,  this  does  not  mean  that  you  should 
not  be  as  prepared  or  organized.  With  small  groups, 
(even  a  single  person  may  be  your  "audience")  you 
need  to  be  more  conversational,  and  you  must  pro- 
vide opportunity  for  appropriate  audience  participa- 
tion. This  can  be  done  with  ample  pauses  which  en- 
courage comments,  and  with  direct  questions.  These 
enable  good  two-way  communications  during  a 
presentation  and  are  effective  for  many  situations. 

Visual  aids  are  as  helpful  for  these  situations  as 
for  larger  audiences.  However,  color  slides  may  not 
be  appropriate,  especially  if  you  only  have  a  20-min- 
ute  appointment  with  the  district  manager  or  with 
personnel  who  know  the  area  and  program  well.  In 
these  situations,  a  neatly  prepared  series  of  overhead 
transparencies  or  pages  on  a  portable  flip  chart  may 
be  suitable  to  support  your  presentation. 


Whenever  possible,  try  to  give  your  presenta- 
tion in  a  small  room  other  than  an  office  where  you 
can  arrange  the  seating  for  the  expected  group  size 
and  set  up  your  visual  aid  equipment  before  the 
scheduled  time.  This  is  especially  relevant  if  your  au- 
dience is  high-level  management.  Their  offices  are 
usually  busy  with  telephones  and  other  interruptions 
that  greatly  detract  from  your  message. 


VISUAL  AIDS 

Appropriate  and  properly-prepared  visual  aids 
can  greatly  increase  the  effectiveness  of  your  presen- 
tation. As  the  term  implies,  these  provide  visual  assis- 
tance for  the  audience  in  understanding  your  verbal 
message. 

There  are  five  types  of  visual  aids: 

1.  Actual  objects  are  the  best  and  should  be  used 
when  appropriate.  When  discussing  big  game 
foods,  bring  along  some  clippings  of  mountain 
mahogany,  bitterbrush,  aspen,  etc.;  when  talking 
about  sex  and  aging  of  wildlife,  take  along  some 
deer  jaws  or  grouse  wings;  or,  show  them  the 
specialized  equipment  used  in  gaining  informa- 
tion about  your  project,  such  as  telemetry  gear, 
ear  tags,  big  game  collars,  immobilizing  darts, 
and  live  traps. 


Verbal  Presentation 


835 


2.  Models  can  be  used  when  the  actual  object  is 
too  large  to  bring  or  too  small  for  the  audience 
to  see.  A  miniature  deer  trap  that  really  works 
would  be  a  good  example. 

3.  Static  graphics  are  maps,  charts,  and  diagrams 
that  may  be  shown  to  the  group  or  hung  on  the 
wall.  Unless  these  are  specially  prepared,  they 
generally  are  not  effective,  as  they  are  too  com- 
plex or  the  printing  is  too  small.  Static  graphics 
are  generally  good  to  have  available  for  the  au- 
dience to  view  after  your  presentation.  They 
should  not  be  displayed  at  the  front  of  the  room 
during  the  talk  as  they  can  be  distracting. 

4.  Active  graphics  are  graphics  that  progressively 
disclose  items  you  are  discussing.  Examples  are 
flip  charts,  hook  and  loop  (velcro)  boards,  flan- 
nel boards,  or  a  chalkboard.  These  can  be  effec- 
tive in  outlining  major  points  for  the  audience 
or  a  sequence  of  events  or  elements  of  a  pro- 
gram. 

5.  Projected  visuals  include  slides,  overheads, 
television,  and  movies. 

Although  you  may  use  several  types  of  visual  aids  in 
your  presentation,  the  following  rules  should  be  fol- 
lowed in  planning  and  preparing  graphics  for  pre- 
senting facts  and  figures: 

1.  Boldness.  The  elements  (numbers  of  words) 
must  be  seen  by  the  audience  to  be  understood. 
How  many  of  you  have  heard  a  speaker  say,  "I 
know  you  can't  see  this,  but . . ."  or  have 
strained  your  eyes  trying  to  read  visuals  at  a 
professional  conference  or  staff  meeting? 

For  elements  to  be  easily  read  by  the  audience, 
they  must  be  1-1/2  inches  high  for  small  rooms 
(up  to  30  feet),  2  to  2-1/2  inches  for  medium 
rooms  (up  to  60  feet),  and  3  inches  or  larger 
for  distances  up  to  100  feet.  It  is  important  to 
be  able  to  read  the  visual  from  the  podium. 
Move  to  the  back  of  the  room  and  see  if  you 
can  see  it  from  there.  However,  if  you  see  that 
the  letters  are  only  1/2  inch  high  on  the  screen, 
you  will  know  that  most  of  the  audience  can 
never  see  them. 

2.  Simplicity.  Because  you  want  your  audience  to 
recognize  the  elements  immediately,  use  simple 
block  letters  or  numbers  as  script;  fancy  letter- 
ing delays  recognition. 

Presentation  of  complex  data  is  confusing.  You 
can  greatly  facilitate  understanding  by  rounding 
off,  condensing,  and  abbreviating  material.  For 
example,  on  a  visual  the  reader  has  difficulty 
comparing 


3. 


ANNUAL  DEER  HARVEST  IN  MONTANA 
1960-1980 


but  when  shown  as 


the  reader  can  immediately  see  that  one  is  twice 
as  large.  Because  you  are  explaining  the  graphic, 
you  do  not  have  to  include  everything  on  the 
visual  as  you  would  in  a  publication.  The 
graphic  should  illustrate  your  main  point. 

Also,  because  you  are  generally  trying  to  have 
your  audience  understand  concepts  (projec- 
tions, trends,  comparisons,  etc.),  there  may  be 
no  need  to  present  precise  data.  The  use  of  dec- 
imal points  is  seldom  necessary  and  often  con- 
veys an  impression  of  precision  that  seldom  ex- 
ists in  natural  resources.  I  suggest  rounding  off 
to  the  nearest  10,  100,  1000,  etc.,  as  appropriate 
for  fast  comprehension.  Also,  avoid  complex, 
cluttered  graphics.  It  is  better  to  have  several 
graphics  with  a  few  items  than  to  jam  them  all 
onto  one. 

Color.  Use  of  color  should  be  considered  for 
several  reasons.  Correct  use  of  color  can  attract 
attention  and  present  a  more  pleasant  visual. 
However,  be  aware  of  contrast  as  the  elements 
must  stand  out  from  the  background  if  they  are 
to  be  easily  read.  Use  dark  colors  against  bright 
ones  such  as  yellows,  tans,  golds,  and  light  blues 
and  greens. 


836 


Verbal  Presentation 


Preparing  Graphics 

Overhead  Transparencies.  Overheads  are  suitable 
for  many  presentations.  The  advantages  are  that  they 
can  be  used  with  the  room  lights  on.  The  speaker 
can  easily  maintain  eye  contact  with  the  audience  as 
a  transparency  can  be  read  correctly  on  the  projec- 
tor. Overheads  are  especially  appropriate  for  smaller 
meetings  (in-house  management)  where  you  may 
wish  to  encourage  discussion  during  the 
presentation. 

A  disadvantage  is  that  in  most  rooms  the 
speaker  blocks  the  view  of  the  screen  for  some  of 
the  audience  while  at  the  projector.  This  can  be 
avoided  by  moving  to  the  side  of  the  screen  once 
the  overhead  has  been  positioned  correctly.  Items 
can  be  identified  on  the  screen  with  a  pointer. 

Material  for  overheads  can  be  easily  made  from 
black  elements  on  a  white  sheet  of  standard  sized  pa- 
per (8  1/2  x  11  inches).  The  sheet  is  then  run 
through  the  office  copy  machine  using  a  transparent 
sheet  instead  of  paper  for  the  copy. 

Standard  type  in  office  typewriters  does  not  pro- 
vide suitable  size  elements  as  the  letters  and  num- 
bers are  too  small  to  be  seen  when  projected.  How- 
ever, the  IBM  "orator"  type  is  large  enough  for 
making  transparencies  for  use  in  a  small  room.  When 
preparing  overheads  using  the  "orator"  type,  place 
two  sheets  of  paper  in  the  typewriter  for  better 
strike,  use  all  caps,  leave  a  space  between  each  letter 
to  prevent  bleeding,  and  use  triple  spacing  between 
lines. 

The  best  transparencies  are  made  from  letters 
that  are  3/16  inch  high  or  larger.  This  produces 
the  correct  size  image  on  the  screen  when  pro- 
jected. These  can  be  made  from  dry  transfer  (press- 


on)  letters,  from  a  lettering  machine  (many  offices 
now  have  them),  from  type  prepared  by  a  commer- 
cial print  shop,  or  made  with  a  hand-lettering  stencil 
set.  The  best  size  type  is  24-point  (a  printer's  term 
for  size),  but  18-  or  20-point  is  also  acceptable. 

Make  a  rough  layout  sketch  of  your  overhead. 
Use  this  to  position  your  printing  on  a  white  sheet  of 
paper.  I  use  a  light  table  with  a  lined  sheet  under  the 
white  paper  to  keep  spacing  uniform  and  parallel. 
With  prepared  type,  you  can  cut  and  paste  the  items 
as  the  edges  will  not  show  on  the  transparency 
when  copied.  You  can  add  lines  with  a  good  inking 
or  felt-tip  pen.  Color  can  be  added  to  these  black 
and  white  transparencies  (underlining,  etc.)  with 
bright  colored  felt-tip  pens. 

Suitable  graphs,  charts,  and  diagrammatic  maps 
can  be  made  by  using  these  colored  felt-tip  pens.  Be 
sure  the  pens  are  made  for  transparencies  or  you 
will  have  a  smeary  mess.  Rough  out  your  artwork  in 
pencil,  then  place  a  clear  transparency  over  it  and 
draw  in  with  the  colored  pens. 

Another  method  for  diagrammatic  maps  and 
some  charts  is  to  do  the  lettering,  numbers,  and  any 
lines  (roads,  streams,  x/y  axes  of  graphs,  etc.)  in 
black  on  a  white  sheet  of  paper  and  make  the  trans- 
parency on  the  copy  machine.  Then,  add  color  to 
this  transparency  with  the  colored  pens.  Different 
habitat  types  can  be  cross-hatched  in  various  colors, 
or  several  colored  lines  can  be  used  on  a  line  graph. 

More  sophisticated  materials  can  be  made  with 
the  help  of  an  experienced  graphic  artist.  The  new 
generation  computers  have  programs  that  can  pre- 
pare high  quality  artwork  on  printers.  However,  re- 
member to  keep  it  simple  and  be  sure  the  elements 
are  large  enough  to  be  seen  when  projected. 


Equipment  and  materials  used  in  making  an  overhead 
presentation.  Left,  a  thermofax  machine  that  transfers 
printed  material  onto  transparencies.  Center,  transparen- 
cies that  can  be  reproduced  in  a  copying  machine.  Right, 
felt-tip  markers  for  writing  on  acetate  film,  and  an  over- 
head projector. 


Verbal  Presentation 


837 


35mm  Slides.  Color  slides  are  appropriate  visual 
aids  for  many  presentations  as  much  field  work  can 
easily  be  photographed  with  a  35  mm  camera.  Slides 
can  be  used  to  show  contrasts,  seasonal  differences, 
relationships,  or  impacts;  explain  management  prac- 
tices and  research  techniques;  and  describe  a  species 
or  management  area.  Because  slides  are  so  useful  in 
public  presentations,  I  consider  the  35  mm  single 
lens  reflex  camera  an  important  part  of  every  biolo- 
gist's personal  and  professional  equipment.  High 
quality  equipment  is  now  reasonably  priced,  and  I 
suggest  you  own  your  own  and  know  how  to  use  it. 


it  makes  your  presentation  "first  person," 
stract  "third  person." 


not  an  ab- 


Effective  slides  showing  changes  from  pioneer 
days  can  be  made  from  old  black  and  white  photos 
usually  available  from  the  local  historical  society. 
Photograph  them  with  color  slide  film. 

Use  only  sharp,  correctly  exposed  slides  and 
keep  them  clean.  Dust  can  be  removed  with  a  camel- 
hair  brush  or  a  blower.  Smears  should  be  removed 
with  film  cleaner. 


As  you  plan  a  slide  presentation,  try  to  include 
variety  whenever  possible.  A  presentation  on  deer  or 
sage  grouse  management  when  using  only  slides  of 
deer  or  sage  grouse  and  preferred  habitats  can  be 
visually  boring.  I  suggest  including  photos  of  other 
animals  in  the  habitat:  birds,  small  mammals,  snakes, 
etc.  A  few  good  wildflowers  are  almost  always  well 
received  when  describing  plants  in  the  habitat  in  ad- 
dition to  the  main  food  or  cover  items.  Also,  try  for 
variety  in  lens  angle.  When  describing  a  habitat,  start 
with  a  wide  angle  scenic  or  panoramic,  then  go  to  a 
slide  with  a  normal  lens,  move  on  to  a  smaller  area, 
and  maybe  even  closer  to  a  plant  community  or  a 
few  key  plants. 


Include  people  in  your  photos  when  appropriate 
and  yourself,  if  possible.  You  in  a  photo  can  improve 
your  credibility  and  should  be  supported  by  verbal 
descriptions  of  your  involvement.  This  is  effective  as 


In  addition  to  pictorial  slides,  suitable  graphic 
slides  can  help  clarify  items  in  your  presentation. 
You  can  prepare  suitable  artwork  easily  and  photo- 
graph it  to  make  good  graphic  slides.  One  method  is 
to  use  an  8-1/2-  x  1 1-inch  or  9-  x  12-inch  sheet  (this 
is  about  the  proportion  of  the  35  mm  slide)  of  col- 
ored paper  as  background  for  the  artwork.  I  find 
light  colored  art  paper  in  earth  tones  (golds,  or- 
anges, tans,  and  greens)  to  be  good.  Printing  can  be 
added  by  using  24-point  (or  20)  dry  transfer  letters 
and  1/8-inch  matte  press-on  tapes  in  various  colors 
for  lines.  These  tapes  are  the  right  proportion  in  re- 
lation to  the  letters  for  lines  or  bars  in  graphs.  The 
tape  is  also  available  as  contour  tape  that  can  be 
used  for  making  curved  lines.  The  paper,  letters,  and 
tapes  are  available  in  graphic  art  or  drafting  supply 
stores. 

Make  a  rough  pencil  layout  of  your  design  be- 
fore starting  on  the  colored  paper.  Use  the  edge  of  a 


838 


Verbal  Presentation 


thin  sheet  of  white  paper  as  a  line  for  placing  letters, 
not  a  penciled  line.  The  pencil  line,  when  erased, 
will  photograph  a  smudged  line.  A  piece  of  masking 
tape  will  lift  off  mistakes  with  the  dry  transfer  letters 
when  lightly  touched  to  the  letter.  Keep  all  impor- 
tant material  in  at  least  1-1/4  inches  from  each  edge 
to  permit  easy  camera  adjustment  when  photograph- 
ing the  artwork. 

Effective  diagrammatic  maps  can  be  made  by  us- 
ing different  colors  of  paper  to  show  various  owner- 
ships, habitat  types,  seasonal  ranges,  etc.  These  are 
usually  better  than  showing  a  slide  of  a  part  of  a 
complex  topographical  sheet.  Pie  charts  are  also  ef- 
fective graphics  as  they  show  comparisons  much  bet- 
ter than  numbers  on  a  chart. 

A  typewriter  can  be  used  for  making  acceptable 
slide  graphics  if  the  material  is  kept  within  a  3-  x  5- 
inch  space  on  a  4-  x  6-inch  area  of  a  light  colored 
sheet.  You  must  come  in  close  and  photograph  the 
4-  x  6-inch  space  when  using  typed  letters  in  order 
for  the  printing  to  be  large  enough  when  projected. 
This  is  the  reason  you  cannot  photograph  a  page 
with  a  chart  or  table  directly  from  a  publication  to 
make  a  slide.  The  elements  will  be  too  small.  Take 
the  material  from  the  publication,  simplify,  condense, 
round  off,  and  abbreviate  to  make  a  new  graphic,  as 
described  above. 

You  can  do  your  own  photography  of  artwork 
to  make  your  slides.  A  copy  stand  is  handy,  but  you 
can  use  a  tripod  and  shoot  material  held  with  mask- 
ing tape  to  a  wall  or  on  a  drawing  board.  Natural 
light  works,  but  flash  or  flood  lamps  can  be  useful. 
Practice  shooting  at  various  exposures  to  get  the 
best  results.  Unless  you  have  a  lot  of  graphics  to 
shoot,  it  may  be  cheaper  to  have  a  commercial  pho- 
tographer shoot  them  for  you. 


CONCLUSION 

This  chapter  is  intended  to  help  you  with  effec- 
tive verbal  presentations  regardless  of  the  audience, 
i.e.,  peers,  administrators,  advisory  groups,  or  special 
interest  publics. 

A  professional  is  one  who  knows  his  subject 
area,  is  experienced,  and  can  communicate  ideas 
well  to  other  professionals  and  lay  people.  Unfortu- 
nately, training  in  communications  receives  minor 
emphasis  in  university  preparation.  This  situation  is 
being  improved  somewhat  by  in-service  training  in 
some  states  and  by  communications  workshops  of- 
fered annually  by  the  Arizona  Chapter  and  the  West- 
ern Section  of  the  Wildlife  Society. 

In  addition  to  these  possibilities,  I  suggest  you 
consider  joining  the  local  Toastmasters  Club.  This 
group  meets  weekly  for  the  sole  purpose  of  improv- 
ing members'  communication  skills.  If  there  is  not 
one  near  you,  start  a  club  in  your  locality.  Contact 
Toastmasters  Club  International,  P.O.  Box  10400, 
Santa  Ana,  CA  92711. 

The  only  way  to  become  a  good  speaker  is 
through  practice  behind  the  podium.  Toastmasters 
provides  such  an  opportunity.  You  might  also  con- 
sider attending  speech  classes  offered  at  the  local 
university,  junior  college,  or  through  adult  education 
programs.  In  addition  to  practicing,  you  will  also  re- 
ceive objective  critiques  which  can  be  most  useful. 

Good  communication  is  essential  in  modern  nat- 
ural resource  management.  As  professionals,  your 
challenge  is  to  obtain  an  understanding  of  complex 
wildlife/habitat  interrelationships  and  communicate 
this  understanding  to  people  whose  support  you 
need  for  your  management  program.  I  hope  you  will 
accept  this  challenge. 


Verbal  Presentation 


839 


APPENDIX  I. 


APPENDIX  II. 


Checklist  for  slide  show  presentations. 


Suggested  Readings. 


A.  Planning 

1.  Prepare  statement  of  purpose. 

2.  Analyze  audience — education,  number, 
experience,  etc. 

3.  Outline  presentation. 

4.  List  visual  aids  and  related  equipment  needs. 

5.  Clear  calendar  of  conflicts,  confirm  date. 

B.  Preparation 

1.  Prepare  final  outline. 

2.  Write  biographic  sketch  for  host. 

3.  Prepare  appropriate  graphics. 

4.  Check  slides  for  relevance,  quality,  timeliness, 
dirt,  etc. 

5.  Check  out  equipment: 

•  projector  (extra  bulb) 

•  projector  stand 

•  screen 

•  extension  cords — control,  power 

•  pointer 

•  lens  cleaning  kit — brush,  lens  cleaner, 
tissue 

6.  Rehearse — check  timing,  order  of  slides.  Use 
tape  recorder. 

7.  Arrive  early  on  day  of  presentation: 

•  check  facilities — heat,  ventilation,  seating 

•  position  projector  and  screen,  tie  down 
cords 

•  focus  and  position  first  slide 

•  adjust  PA  system  and  microphone 

•  locate  light  switches 

•  minimize  distracting  light 


Of  the  many  publications  on  public  speaking  and  vis- 
ual aids,  I  suggest  several  that  are  excellent: 

Making  Effective  Slides  for  Meetings;  Fisher,  H.L.,  Leaf- 
let 21119,  Cooperative  Extension,  U.S.  Dep.  Agric, 
University  of  California,  Berkeley,  CA  94720 

The  following  are  selections  from  many  fine  publica- 
tions available  from  Eastman  Kodak  Company,  343 
State  Street,  Rochester,  NY  14650: 

Presenting  Yourself  (S-60) 
Effective  Lecture  Slides  (S-22) 
Legibility — Artwork  to  Screen  (S-24) 
Planning  and  Producing  Slide  Programs  (S-30) 
Slides  with  a  Purpose  (VI- 15) 


D. 


Presentation 

1.  Check  appearance,  grooming,  etc. 

2.  Meet  host — give  biographic  sketch,  mix  with 
group. 

3.  Appoint  projector  and  light-switch  assistants. 

4.  Give  introduction  with  lights  on. 

5.  Begin  slides.  (Transition:  projector  on,  then  lights 
off.) 

6.  Give  meaningful  closing  with  lights  on.  (Tran- 
sition from  slides:  lights  on,  then  projector  off — 
no  fan.) 

7.  Keep  on  schedule. 

8.  Have  host  handle  discussion  period. 

Evaluation 

Write  memo  to  files — include  information  and 
slides  needed,  response,  personnel  perfor- 
mance. 


840 


Verbal  Presentation 


GLOSSARY 


This  glossary  was  prepared  by  using  definitions  provided  by  the  authors  or  from  the  following  dictionaries: 

The  Dictionary'  of  the  Biological  Sciences.  Peter  Gray.  Van  Nostrand  Reinhold  Company   New  York,  NY.  1967  edition 

A  Dictionary  of  Ecology,  Evolution,  and  Systematics.  R.J.  Lincoln,  G  A.  Boxshall.  and  P.F.  Clark.  Camhridge  University  Press.  London. 

McGraw-Hill  Dictionary  of  the  Life  Sciences.  McGraw-Hill  Book  Company.  New  York,  NY.  1974  edition. 

The  Random  House  College  Dictionary,  1st  ed.  Random  House,  Inc.  New  York,  NY.  1975  edition. 

Resource  Conservation  Glossary,  3rd  ed.  Soil  Conservation  Society  of  America.  Ankeny,  IA.  1982  edition. 

Webster's  New  Collegiate  Dictionary.  G&C  Merriam  Company.  Springfield,  MA.  1979  edition 

Webster's  Third  New  International  Dictionary  (unabridged)  G&C  Merriam  Company.  Springfield,  MA.  1966  edition. 

ABIOTIC    devoid  of  life;  non-living. 

ABLATION    the  removal  of  a  surface  layer,  as  of  ice  by  melting  or  evaporation. 

ABNEY  LEVEL    a  surveying  clinometer  consisting  of  a  short  telescope,  bubble  tube,  and  graduated  vertical  arc- 
used  especially  for  measuring  tree  heights. 

ACCURACY    relates  to  the  freedom  from  bias. 

AGL  (ABOVE  GROUND  LEVEL)    used  to  define  heights  that  certain  aerial  censuses  should  be  flown. 

ALBEDO    a  measure  of  surface  reflectivity,  that  fraction  of  incident  electromagnetic  radiation  that  is  reflected 
by  a  body  or  surface,  usually  expressed  as  a  percentage. 

ALKALI    a  substance  having  marked  basic  properties  in  contrast  to  acid  properties;  a  substance  with  a  pH 
above  7. 

ALKALINITY    the  quality  or  state  of  being  alkaline  (having  a  pH  balance  above  7). 

ALTRICIAL    helpless  and  naked  when  hatched,  as  young  pigeons  are  (compare  to  precocial). 

ARBOREAL    living  in  or  adapted  for  living  in  trees. 

AREAL    any  particular  extent  of  space  or  surface,  as  a  geographical  region. 

ASYNCHRONY    lack  of  synchronism. 

AUFEIS    German  term  for  sheets  of  ice  formed  by  the  freezing  of  overflow  water. 

AUM  (ANIMAL  UNIT  MONTH)    the  quantity  of  forage  required  by  one  mature  cow  ( 1,000  lbs.)  or  the  equiv- 
alent for  one  month. 

AUTOECOLOGICAL    the  study  of  the  individual,  or  members  of  a  species  collectively,  in  relation  to  environ- 
mental conditions. 

AUTOTROPH    an  organism  capable  of  synthesizing  organic  nutrients  directly  from  simple  inorganic  sub- 
stances such  as  carbon  dioxide  and  inorganic  nitrogen. 

BENEFITS  FOREGONE    the  benefits  that  would  have  come  from  options  not  chosen  when  a  choice  between 
options  has  to  be  made. 

BENTHIC    pertaining  to  or  living  on  the  bottom  or  at  the  greatest  depths  of  a  large  body  of  water. 

BEQUEST  VALUE    the  amount  an  individual  would  be  willing  to  spend  to  ensure  that  a  resource  would  be 
present  in  the  future  for  future  generations. 

Glossary  g4l 


BIOGEOGRAPHY    the  biological  study  of  the  geographical  distribution  of  plants  and  animals. 

BIOMASS    the  total  amount  of  living  material,  plants  and/or  animals,  above  and  below  the  ground  in  a  particu- 
lar habitat  or  area. 

BIOTIC    pertaining  to  life  or  living  organisms,  caused  or  produced  by  or  comprising  living  organisms. 

BOGS    a  plant  community  that  develops  and  grows  in  areas  with  permanently  waterlogged  peat  substrates; 
also  known  as  moor  or  quagmire. 

BOREAL    pertaining  to  cool  or  cold  temperate  regions  of  the  northern  hemisphere,  the  northern  coniferous 
zone  and  taiga. 

CARNIVOROUS    flesh  eating 

CHEMOCLINE    the  transition  in  a  meromictic  lake  between  the  mixolimnion  layer  (at  the  top)  and  the  moni- 
molimnion  layer  (at  the  bottom). 

CHIONOPHILES    animals  that  thrive  in  snow-covered  habitats. 

CHIONOPHOBES    animals  intolerant  of  snow-covered  habitats. 

CIENEGAS    an  area  where  the  water  table  is  at  or  near  the  surface  of  the  ground;  standing  water  occurs  in 
depressions  in  the  area,  and  it  is  covered  with  grass  or  sometimes  with  heavy  vegetation. 

CLASSIFICATION    the  act  or  method  of  distributing  portions  of  land  and  their  attendant  biotic/abiotic  attri- 
butes into  groups  or  divisions  wherein  each  portion  within  a  group  has  similar  or  identical  characteristics; 
also,  naming  systematically. 

CLINOMETER    any  of  various  instruments  for  measuring  angles  of  elevation  or  inclination. 

COAXIAL  CABLE    a  transmission  line  that  consists  of  a  tube  of  electrically  conducting  material  surrounding  a 
central  conductor  held  in  place  by  insulators  and  that  is  used  to  transmit  telegraph,  telephone,  and  television 
signals  of  high  frequency. 

COLD-WATER  FISH    fish  that  can  stand  temperatures  ranging  from  33  degrees  Fahrenheit  to  70  degrees  Fahr- 
enheit. 

COLINEARITY    having  corresponding  parts  arranged  in  the  same  linear  order. 

CONDUCTIVITY    a  measure  of  the  ability  of  a  material  to  conduct  an  electrical  charge;  the  reciprocal  of  re- 
sistivity. 

CONSPECIFIC    of  the  same  species. 

COOL-WATER  FISH    fish  that  can  stand  temperatures  ranging  from  33  degrees  Fahrenheit  to  75  degrees  Fahr- 
enheit. 

CORM    an  underground  stem,  such  as  that  of  the  gladiolus,  similar  to  a  bulb  but  without  scales. 

CREPUSCULAR    active  in  twilight,  at  dawn  or  dusk. 

CRYOPEDOGENIC    caused  by  or  associated  with  permanently  frozen  ground  or  intensive  frost  action. 

CURSORIAL    adapted  for  running. 

DEMAND  CURVE    a  curve  that  shows  the  amount  demanded  at  each  price  level. 

DENDROGRAM    a  branching  diagrammatic  representation  of  the  interrelations  of  a  group  of  items  sharing 
some  common  factors  (as  of  natural  groups  connected  by  a  common  ancestral  form). 

842  Glossary 


DEPILATORY    a  liquid  or  cream  used  to  remove  unwanted  hair  from  the  body. 

DETRITIVORES    organisms  that  feed  on  fragmented  particulate  organic  matter  (detritus). 

DETRITUS    fragmented  particulate  organic  matter  derived  from  the  decomposition  of  plant  and  animal  re- 
mains, organic  debris;  accumulated  in  sand,  water,  or  on  mud  or  soil. 

DIATOMS    any  of  numerous  microscopic,  unicellular  (single-celled),  marine  or  freshwater  algae  having  sili- 
ceous cell  walls. 

DIMICTIC    used  to  describe  a  lake  having  two  seasonal  overturn  periods  of  free  circulation,  with  accompany- 
ing description  of  the  thermocline. 

DIURNAL    pertaining  to  those  organisms  that  are  active  during  daytime. 

DO  (DISSOLVED  OXYGEN)    refers  to  the  amount  of  dissolved  oxygen  in  water. 

DOMINANCE  TYPE    species,  that  by  their  activity,  behavior,  or  number  have  considerable  influence  or  con- 
trol upon  the  conditions  of  existence  of  associated  species;  a  species  that  controls  its  habitat  and  food  web. 

DRUMMING    to  make  a  succession  of  strokes  or  vibrations  that  produce  sounds  like  drumbeats  by  beating 
the  wings,  like  a  male  grouse. 

DYSTROPHIC    pertaining  to  an  environment  that  does  not  supply  adequate  nutrition. 

ECHOLOCATION    the  perception  of  objects  using  high  frequency  sound  waves,  used  by  some  animals  for 
navigation  and  orientation  within  the  environment. 

ECOLOGICAL  SITES    areas  of  land  described  by  polygons  on  a  map  or  photograph  that  have  potential  for 
producing  similar  potential  vegetation,  developed  from  the  ecological  site  system  of  the  U.S.  Soil  Conservation 
Service. 

ECOTONE    a  transition  line  or  strip  of  vegetation  between  two  communities,  having  characteristics  of  both 
kinds  of  neighboring  vegetation  as  well  as  characteristics  of  its  own. 

ECTOTHERMIC    poikilothermic,  having  a  body  temperature  determined  primarily  by  the  temperature  of  the 
environment;  cold-blooded. 

EDAPHIC    of  or  pertaining  to  soil,  especially  as  it  affects  living  organisms. 

EMERGENT  PLANTS    ( aquatic )  an  aquatic  plant  having  most  of  its  vegetative  parts  above  water. 

EPHEMERAL    (  1  )  relating  to  a  stream  or  a  portion  of  a  stream  that  flows  only  in  direct  response  to  precipita- 
tion and  receives  little  or  no  water  from  springs  or  no  long-continued  supply  from  snow  or  other  sources, 
and  its  channel  is  at  all  times  above  the  water  table;  (2)  lasting  for  only  a  day;  short-lived  or  transient. 

EPILIMNION    a  freshwater  zone  of  relatively  warm  water  in  which  mixing  occurs  as  a  result  of  wind  action 
and  connection  currents. 

ESTIVATE    ( a  variant  of  AESTWATE )  to  pass  the  summer,  especially  in  a  state  of  dormancy. 

ESTUARINE    of,  relating  to,  or  formed  in  an  estuary  ( water  passage )  where  the  tide  meets  a  river  current. 

EULITTORAL    a  subdivision  of  the  benthic  division  of  the  littoral  zone  of  the  marine  environment. 

EURY-    a  prefix  meaning  wide. 

EUTROPHIC    rich  in  nutrients,  characterized  by  abundant  plankton  and  high  turbidity. 

EUTROPHIC  LAKES    lakes  that  are  rich  in  nutrients  and  organic  materials,  therefore,  highly  productive  for 
plant  growth.  These  lakes  are  often  shallow  and  seasonally  deficient  in  oxygen  in  the  hypolimnion. 

Glossary  843 


EUTROPHICATION    a  means  of  aging  of  lakes  whereby  aquatic  plants  are  abundant  and  waters  are  deficient 
in  oxygen.  The  process  is  usually  accelerated  by  enrichment  of  waters  with  surface  runoff  containing  nitrogen 
and  phosphorus. 

EXISTENCE  VALUE    the  amount  of  money  an  individual  would  be  willing  to  spend  to  know  a  resource  pres- 
ently exists  and  would  continue  to  exist. 

FACULTATIVE    contingent;  assuming  a  particular  role  or  model  of  life  but  not  restricted  to  that  condition. 
Used  of  organisms  having  the  facility  to  live,  or  living,  under  atypical  conditions. 

FECUND    capable  of  producing  offspring;  fertile,  productive. 

FELL-FIELD    a  type  of  tundra  ecosystem  having  sparse,  dwarfed  vegetation  and  flat,  very  stony  soil. 

FENS    a  peaty,  moist  tract,  usually  derived  from  the  aging  or  drainage  of  a  swamp. 

FISTULA    an  abnormal  passage  leading  from  an  abscess  or  hollow  organ  to  the  body  surface  or  from  one 
hollow  organ  to  another.  Sometimes  done  surgically  to  enable  scientists  to  examine  stomach  contents. 

FLORISTICS    the  plant  life  of  a  given  region,  habitat,  or  geological  stratum. 

FOSSORIAL    adapted  for  or  used  in  burrowing  or  digging. 

GALLINACEOUS    belonging  or  pertaining  to  the  Order  Galliformes,  comprising  the  grouse,  pheasants,  tur- 
keys, partridges,  domestic  fowls,  etc. 

GEOMORPHIC    of  or  pertaining  to  the  figure  of  the  earth  or  the  forms  of  its  surface.  Of  or  like  the  earth,  its 
shape  or  surface  configuration. 

GEOMORPHOLOGY    a  science  that  deals  with  the  land  and  submarine  relief  features  of  the  earth's  surface 
and  seeks  a  genetic  interpretation  of  them. 

GLEY    a  bluish-gray  or  olive-gray  sticky  layer  of  clay  formed  under  the  surface  of  certain  waterlogged  soils. 

GREGARIOUS    tending  to  aggregate  actively  into  groups  or  clusters. 

GUILD    a  group  of  species  having  similar  ecological  resource  requirements  and  foraging  strategies  and  there- 
fore having  similar  roles  in  the  community. 

HARD  WATER    water  containing  dissolved  salts  of  calcium  and  magnesium,  especially  water  containing  more 
than  85.5  parts  per  million  of  calcium  carbonate. 

HEP  (HABITAT  EVALUATION  PROCEDURES)    a  planning  and  evaluation  technique  which  focuses  on  the 
habitat  requirements  of  fish  and  wildlife  species.  It  is  a  standardized  procedure  for  conducting  habitat  evalua- 
tions in  the  field  developed  and  used  by  the  U.S.  Fish  and  Wildlife  Service. 

HERBrVORE    an  animal  that  feeds  on  plants. 

HERBrVOROUS  (HERBrVORY)    feeding  on  plants;  phytophagous. 

HERPETOFAUNA    animals  that  are  either  amphibians  or  reptiles. 

HETEROGENEITY    differing  in  kind;  having  unlike  qualities,  possessed  of  different  characteristics. 

HETEROTROPHIC    obtaining  nourishment  from  exogenous  organic  material.  Used  of  organisms  unable  to 
synthesize  organic  compounds  from  inorganic  substrates. 

HETEROTROPHS    an  organism  that  obtains  nourishment  from  the  ingestion  and  breakdown  of  organic 
matter. 

HETEROZYGOSITY    having  two  different  alleles  at  a  given  locus  of  a  chromosome  pair. 
844  Glossary 


HIBERNACULA    the  domicile  in  which  an  animal  hihernates  or  overwinters;  winter  quarters. 

HOLARCTIC  of  or  designating  the  zoogeographic  region  that  includes  the  northern  areas  of  the  earth  and  is 
divided  into  Nearctic  and  Palearctic  regions. 

HOME  RANGE    the  area,  usually  around  the  domicile,  over  which  an  animal  normally  travels  in  search  of 
food. 

HOMEOTHERMIC    maintains  a  relatively  constant  and  warm  hody  temperature  that  is  independent  of  envi- 
ronmental temperatures;  warm-blooded. 

HORIZON    a  layer  of  soil  or  soil  material  approximately  parallel  to  the  land  surface  and  differing  from  adja- 
cent, genetically  related  layers  in  physical,  chemical,  and  biological  properties  or  characteristics. 

HQI  (HABITAT  QUALITY  INDEX)    this  index  relates  habitat  quality  to  fish  biomass  in  streams.  (Refer  to 
Binns,  N.A.  1979-  A  habitat  quality  index  for  Wyoming  trout  streams.  Fishery  Res.  Rep.  Monogr.  Series  2. 
Wyoming  Game  and  Fish  Dep.  Cheyenne.  75pp.) 

HSI  (HABITAT  SUITABILITY  INDEX)  MODEL    a  specific  type  of  habitat  model  used  by  the  U.S.  Fish  and 
Wildlife  Service  in  its  Habitat  Evaluation  Procedures  (HEPs).  A  set  of  habitat  variables  are  used  to  predict  an 
index  from  0  to  1  of  habitat  suitability  for  a  species. 

HYDRIC  SOILS    soils  that  are  characterized  by  or  thriving  in  an  abundance  of  moisture. 

HYDROGEN  IONS    the  positively  charged  ion  of  hydrogen,  H  + ,  formed  by  removal  of  the  electron  from 
atomic  hydrogen. 

HYDROLOGIC    relating  to  the  science  dealing  with  the  properties,  distribution,  and  circulation  of  water  on 
the  surface  of  the  land,  in  the  soil  and  underlying  rocks,  and  in  the  atmosphere. 

HYDROMETEOROLOGY    the  study  of  ensembles  of  liquid  in  the  atmosphere — rain,  drizzle,  snow,  sleet. 

HYDROPERIODS    the  control  of  vegetative  processes  in  plants  by  periodic  dryness;  seasonal  hydroperiodism. 

HYDROPHYTE    (  1  )  a  plant  that  grows  in  a  moist  habitat;  (2)  a  plant  requiring  large  amounts  of  water  for 
growth. 

HYPOLIMNION    the  cold  bottom  water  zone  below  the  thermociine  in  a  lake. 

INFRALITTORAL  the  depth  zone  of  a  lake  permanently  covered  with  rooted  or  adnate  macroscopic  vegeta- 
tion, often  divided  into  upper  (with  emergent  vegetation),  middle  (with  floating  vegetation),  and  lower  zones 
(with  submerged  vegetation). 

INTERFACE  the  ability  of  computing  devices  or  programs  to  communicate  information  from  one  system  to 
another. 

KROTOVTNAS  (also  CROTOVTNAS)    a  former  animal  burrow  in  one  soil  horizon  that  has  been  filled  with 
organic  matter  or  material  from  another  horizon. 

LACUSTRINE    of,  relating  to,  or  growing  in  lakes. 

LAMINAR  FLOW    streamline  flow  in  a  viscous  fluid  near  a  solid  boundary  (compare  to  turbulent  flow). 

LANDSAT  IMAGE    photographic  images  from  satellites,  observing  earth  resources.  A  satellite  program  that 
developed  the  Earth  Resources  Technology  Satellite— 1  (ERTS-1 ),  which  is  now  called  LANDSAT  1. 

LAVA  TUBES    a  tubular-like  extrusion  of  lava. 

LEK    an  assembly  area  for  communal  courtship  displays. 

LENTIC    pertaining  to  static,  calm,  or  slow-moving  aquatic  habitats. 

Glossary  845 


LIMNETIC    of,  pertaining  to,  or  inhabiting  the  pelagic  region  of  a  body  of  fresh  water. 

LIMNOLOGICAL    relating  to  lakes,  ponds,  and  other  standing  waters  and  their  associated  biota. 

LINCOLN  INDEX    an  estimate  of  population  size  obtained  after  release  and  recapture  of  marked  animals;  the 
estimated  population  size  (N)  is  calculated  from  the  number  of  marked  animals  released  (M),  the  number 
captured  in  a  sample  (n)  after  release,  and  the  number  of  marked  individuals  in  sample  (m)  using  the  formula: 

m 
This  is  also  known  as  Lincoln/Peterson  index  or  Peterson  estimator. 

LINEAR  COMPREHENSIVE  MANAGEMENT    to  measure  all  components  of  habitats  and  populations  and 
know  all  limiting  factors  (have  comprehensive,  quantitative,  and  continuous  knowledge  of  the  system  they 
manage ). 

LITTORAL    of  or  pertaining  to  the  biogeographic  zone  between  the  high-  and  low-water  marks. 

LOCI    (plural  of  LOCUS)  ( 1 )  the  position  that  a  gene  occupies  on  a  chromosome;  (2)  the  set  or  configura- 
tion of  all  points  satisfying  specified  geometric  conditions. 

LOTIC    pertaining  to  fast-running  water  habitats,  such  as  rivers  and  streams. 

MACROINVERTEBRATES    large  or  exceptionally  prominent  animals  that  lack  a  spinal  column. 

MACROPHYTES    a  large  macroscopic  plant,  used  especially  of  aquatic  forms  such  as  kelp. 

MARGIN    the  change  that  would  occur  with  the  addition  or  subtraction  of  one  unit. 

MARINE    pertaining  to  the  sea. 

MARSH    a  periodically  wet  or  continually  flooded  area  where  the  surface  is  not  deeply  submerged,  covered 
dominantly  with  sedges,  cattails,  rushes,  or  other  hydrophytic  plants. 

MAST    the  fruit  of  the  oak  and  beech  or  other  forest  trees,  used  as  food  by  hogs,  birds,  and  other  animals. 

MEROMICTIC    used  of  a  lake  that  is  permanently  stratified  due  to  the  presence  of  a  density  gradient  resulting 
from  chemical  stratification. 

MESIC    having  a  moderate  rainfall. 

MESOTROPHIC    having  intermediate  levels  of  primary  productivity,  pertaining  to  waters  having  intermediate 
levels  of  the  minerals  required  by  green  plants. 

METALIMNION    the  zone  of  steep  temperature  gradient  (thermocline)  between  the  epilimnion  and  the  hypo- 
limnion  in  a  lake 

MICROCOMPUTER    a  compact  and  inexpensive  computer  relatively  limited  in  capability  and  capacity,  con- 
sisting of  a  microprocessor  and  other  components  of  a  computer. 

MICROHABITAT    the  smallest  unit  of  a  habitat,  like  in  a  clump  of  grass  or  a  space  between  rocks. 

MICROPROCESSOR    a  miniaturized  integrated  circuit  that  performs  all  of  the  functions  of  a  central  process- 
ing unit. 

MOBBING    a  collective  attack,  by  a  group  of  animals,  on  a  predator  that  is  too  large  or  aggressive  to  be 
repelled  by  individual  effort. 

MOLLUSCS  any  invertebrate  of  the  Phylum  Mollusca  comprising  the  chitons,  snails,  bivalves,  squids,  octo- 
puses, etc.,  typically  having  a  calcareous  shell  of  one,  two,  or  more  pieces  that  wholly  or  partly  enclose  the 
soft,  unsegmented  body. 

846  Glossary 


MONOCULTURE    raising  crops  of  a  single  species  year  after  year  on  the  same  land. 

MONOMICTIC    used  of  a  lake  having  a  single  period  of  free  circulation  or  overturn  per  year,  with  consequent 
disruption  of  the  thermocline;  may  be  either  cold  monomictic  or  warm  monomictic. 

MORAINIC    of,  pertaining  to,  forming,  or  formed  by  a  moraine  (an  accumulation  of  earth  and  stones  carried 
and  finally  deposited  by  a  glacier). 

MOSAIC    an  assemblage  of  overlapping  aerial  photographs  whose  edges  have  been  matched  to  form  a  contin- 
uous photographic  representation  of  an  area. 

NALED    a  short-lived  insecticide  of  relatively  low  toxicity  to  warm-blooded  animals  that  is  used  especially  to 
control  crop  pests  and  mosquitos. 

NEONATES    newborn,  recently  hatched. 

NEOTROPICS    a  zoogeographical  region  comprising  South  America,  West  Indies,  and  Central  America,  south 
of  the  Mexican  border. 

NONPERSISTENT    an  item  that  is  not  persistent,  does  not  exist  for  a  longer  than  usual  time  or  continuously. 

OBLIGATE    essential,  necessary,  unable  to  exist  in  any  other  state,  mode,  or  relationship. 

OLIGOTROPHIC  of  a  lake,  lacking  plant  nutrients  and  usually  containing  plentiful  amounts  of  dissolved  oxy- 
gen without  marked  stratification. 

OMNIVORE    an  animal  that  feeds  on  a  mixed  diet  of  plant  and  animal  material. 

OMNrVOROUS    feeding  on  a  mixed  diet  of  plant  and  animal  material;  pantophagous. 

OPTION  VALUE  the  option  demand  is  that  demand  from  individuals  who  are  not  now  consumers  or  are  not 
now  consuming  as  much  as  they  anticipate  consuming,  and  who  therefore  would  be  willing  to  pay  to  perpetu- 
ate the  availability  of  the  commodity.  The  option  value  is  the  amount  an  individual  is  willing  to  pay. 

OXBOW    a  bend  in  a  river  that  resembles  the  U-shaped  frame  forming  a  collar  about  an  ox's  neck  and  sup- 
porting the  yoke. 

PALUSTRINE    living  or  thriving  in  a  marshy  environment;  being  or  made  up  of  marsh. 

PELAGIC    (  1  )  pertaining  to  water  of  the  open  portion  of  a  lake;  ( 2 )  pertaining  to  water  of  the  open  portion 
of  an  ocean,  above  the  abyssal  zone  and  beyond  the  outer  limits  of  the  littoral  zone. 

PERCOLATION    the  downward  movement  of  water  through  the  soil. 

PERIGLACIAL    applies  to  an  area  bordering  the  edge  of  an  ice-sheet,  to  the  climate  of  that  area,  to  physical 
processes  involving  freeze-thaw  activity  and  to  their  results. 

PERMAFROST    permanently  frozen  subsoil. 

PERMUTATIONS    the  act  of  altering  a  given  set  of  objects  in  a  group. 

PERSISTENT    refusing  to  give  up  or  let  go;  lasting  past  maturity  without  falling  off,  like  certain  leaves  or 
flowers. 

PERTLIRBATIONS    the  state  or  condition  of  being  perturbed;  agitation. 

PETERSON  ESTIMATOR    see  Lincoln  Index 

pH  a  numerical  measure  of  acidity  or  hydrogen  ion  activity.  Neutral  is  pH  7.  All  pH  values  below  7  are  acid, 
and  all  above  7  are  alkaline. 

Glossary  847 


PHENOLOGICAL    the  relationship  between  climate  and  periodic  biological  phenomena  such  as  bird  migra- 
tion or  plant  flowering. 

PHENOLOGY    the  study  of  the  temporal  aspects  of  recurrent  natural  phenomena  and  their  relationships  to 
weather  and  climate. 

PHREATOPHYTES    a  plant  that  absorbs  water  from  the  permanent  water  table. 

PHYSIOGNOMY    the  characteristic  features  or  appearance  of  a  plant  community. 

PHYTOPLANKTON    unattached  microscopic  plants  of  plankton,  subject  to  movement  by  wave  or  current 
action. 

PIEDMONT    a  district  lying  along  or  near  the  foot  of  a  mountain  range. 

PINGO    a  low  hill  or  mound  forced  up  by  hydrostatic  pressure  in  an  area  underlain  by  permafrost. 

PIONEERING    an  animal  or  plant  species  that  establishes  itself  in  a  previously  barren  environment. 

PLXEL    derived  from  picture  element;  the  smallest  unit  on  the  ground  that  can  be  detected  by  a  multispectral 
scanner.  For  the  LANDSAT  MMS,  the  pixels  are  located  on  the  ground  on  57  x  79  meter  centers. 

PLANKTON    suspended,  floating,  or  weakly  swimming  microscopic  plants  and  animals  in  the  water  that  pro- 
vide a  basis  for  the  aquatic  food  chain. 

PLAYA    the  sandy,  salty,  or  mud-caked  flat  floor  of  a  desert  basin  having  interior  drainage,  usually  occupied  by 
a  shallow  lake  during  or  after  prolonged,  heavy  rains. 

POINT-CENTER  QUADRAT    a  method  of  plotless  sampling  of  vegetation  in  which  lines  are  erected  at  right 
angles  from  the  sampling  point  to  produce  four  quarters  in  each,  from  which  a  measure  is  taken  of  the  dis- 
tance from  the  sampling  point  to  the  nearest  neighbor. 

PRECISION    relates  to  the  repeatability  of  a  result. 

PRECOCIAL    of  or  characterizing  birds  that  are  covered  with  down  and  capable  of  moving  about  when  first 
hatched. 

PROFILE  BOARD    a  scaled  board  that  when  photographed  behind  vegetation  displays  the  physiognomy  of  a 
plant  or  group  of  plants.  Sequential  photographs  can  be  used  to  display  vegetation  changes. 

PROFUNDAL    pertaining  to  the  deep  zone  of  a  lake  below  the  level  of  effective  light  penetration,  and  hence 
of  vegetation. 

QUADRATS    a  small  plot  or  sample  area,  frequently  1  square  meter  or  1  milacre  in  size. 

RAIN  SHADOW    an  area  of  light  rainfall  situated  on  the  lee  side  of  a  range  of  mountains  or  hills. 

REDD    the  spawning  ground  or  nest  of  various  fishes. 

REDOX  POTENTIAL    reduction  potential;  a  measure  of  the  tendency  of  a  given  system  to  act  as  an  oxidizing 
(electron  acceptor)  or  reducing  (electron  donor)  agent. 

RELEVES    a  random  sample  of  vegetation. 

RELICT    a  remnant  or  fragment  of  a  flora  that  remains  from  a  former  period  when  it  was  more  widely  distrib- 
uted. 

RESPIRE  CUTANEOUSLY    to  breathe  in  and  out  (inhale  and  exhale)  through  the  skin. 

REVETMENT    a  facing  of  masonry  or  the  like,  especially  for  protecting  an  embankment. 

848  Glossary 


RIPARIAN  HABITAT    relating  to  or  living  or  located  on  the  bank  of  a  natural  watercourse  (like  a  river)  or 
sometimes  of  a  lake  or  a  tidewater. 

RIVERINE    pertaining  to  a  river. 

ROSETTES    any  structure  or  marking  resembling  a  rose. 

ROTENONE    a  white,  crystalline,  water-insoluble,  poisonous  heterocyclic  compound  obtained  from  derris 
root;  used  in  certain  insecticides. 

ROTIFER    any  of  various  minute,  multicellular  aquatic  organisms  of  the  Phylum  Rotifera,  having  at  the  ante- 
rior end  a  wheel-like  ring  of  cilia. 

SALINITY    the  concentration  of  dissolved  solids  or  salt  in  water. 

SALMONID    any  of  a  Family  Salmonidae  of  elongate  soft-finned  fishes  that  have  the  last  vertebrae  upturned. 

SAPROPEL    used  of  organisms  inhabiting  mud,  rich  in  decaying  organic  matter. 

SAVANNA    the  tropical  and  subtropical  grassland  biome,  transitional  in  character  between  grassland  or  desert 
and  rain  forest,  typically  having  drought-resistant  vegetation  dominated  by  grasses  with  scattered  tall  trees. 

SCANNER    any  of  various  electronic  or  optical  devices  by  which  images  or  recorded  information  are  sensed 
for  subsequent  modification,  integration,  or  transmission. 

SERAL    pertaining  to  a  succession  of  plant  communities  in  a  given  habitat  leading  to  a  particular  climax  asso- 
ciation; a  stage  in  a  community  succession. 

SERES    the  entire  sequence  of  ecological  communities  successfully  occupying  an  area. 

SERPENTINE  PARENT  MATERIAL    a  soil  parent  material  derived  mainly  from  serpentine  rock  and  domi- 
nated by  serpentine  minerals  such  as  antigorite,  chrysolite,  fibrolite,  and  talc. 

SOFT  WATER    water  containing  little  or  no  dissolved  salts  of  calcium  or  magnesium,  especially  water  contain- 
ing less  than  85.5  parts  per  million  of  calcium  carbonate. 

SOIL  CONSISTENCE    the  feel  of  the  soil  and  ease  with  which  a  lump  can  be  crushed  by  the  fingers.  Terms 
commonly  used  to  describe  soil  consistence  are  loose,  friable,  firm,  plastic,  sticky,  hard,  soft,  or  cemented. 

SOIL  ERODIBILITY  FACTOR  (K)    a  measure  of  the  susceptibility  of  soil  particles  to  detachment  and  trans- 
port by  rainfall  and  runoff.  The  K  factor  is  used  in  the  Universal  Soil  Loss  Equation. 

SOIL  STRUCTURE    the  combination  or  arrangement  of  primary  soil  particles  into  secondary  particles,  units, 
or  peds. 

SPATIALLY    the  pattern  of  distribution  of  organisms  in  space. 

SPECIES  RICHNESS    the  absolute  number  of  species  in  an  assemblage  or  community. 

SPECIFIC  CONDUCTANCE    the  quality  of  living  matter  responsible  for  the  transmission  of  and  progressive 
reaction  to  stimuli  within  the  living  system. 

STATISTICAL  TYPE  II  ERROR    when  a  true  alternative  hypothesis  is  rejected. 

STENO-    a  prefix  meaning  narrow. 

STOCHASTIC    pertaining  to  randomness. 

STRATA    ( 1 )  the  divisions/groups  into  which  homogeneous  polygons  that  describe  a  mapped  land  area  are 
placed  so  that  all  portions  or  polygons  within  a  division  appear  to  have  similar  or  identical  attributes;  (2) 
divisions  of  a  classification  system. 

Glossary  849 


STRATIFICATION    to  become  layered,  like  layers  of  water  temperature  in  a  body  of  water. 

STRIP  TRANSECT    a  method  of  sampling  that  entails  walking  a  predetermined  line,  counting  the  animals  ob- 
served, and  recording  the  distances  at  which  they  are  seen  or  flushed.  The  average  of  the  flushing  distance  is 
calculated  and  used  to  determine  strip  width. 

SUBMERGENTS    pertaining  to  a  plant  or  plant  structure  growing  entirely  underwater. 

SUBNIVEAN    situated  or  occurring  under  the  snow. 

SUBSTRATE    ( 1 )  (biology)  the  base  of  substance  upon  which  an  organism  is  growing;  (2 )  (hydrology)  the 
bottom  material  of  a  waterway. 

SWAMPS    areas  saturated  with  water  throughout  much  of  the  year  but  with  the  surface  of  the  soil  usually  not 
deeply  submerged. 

SYMPATRIC    used  of  populations,  species,  or  taxa  occurring  together  in  the  same  geographical  area.  The  pop- 
ulations may  occupy  the  same  habitat  (biotic  sympatry)  or  different  habitats  (neighboring  sympatry)  within 
the  same  geographical  area. 

SYNCHRONY    a  synchronous  occurrence,  movement,  or  arrangement  ( occurring  at  the  same  time ). 

SYNECOLOGICAL    a  subdivision  of  ecology  that  deals  with  the  study  of  groups  of  organisms  associated  as  a 
unit. 

TAIGA    northern  coniferous  forest  biome.  The  ecosystem  adjacent  to  the  arctic  tundra,  but  used  with  varying 
scope  to  include  only  the  arctic  timberline  ecotone  through  to  the  entire  subarctic  north  temperate  forest. 

TALUS    fragments  of  rock  and  other  soil  material  accumulated  by  gravity  at  the  foot  of  cliffs  or  steep  slopes. 

TDS  (TOTAL  DISSOLVED  SOLIDS)    dissolved  solids  are  anhydrous  residues  of  the  dissolved  substances  in 
water. 

THERMOCLINE    a  horizontal  temperature  discontinuity  layer  in  a  lake  in  which  the  temperature  falls  by  at 
least  1  degree  Celsius  per  meter  depth;  thermal  layer. 

THERMOREGULATE    a  mechanism  by  which  mammals  and  birds  attempt  to  balance  heat  gain  and  heat  loss 
in  order  to  maintain  a  constant  body  temperature  when  exposed  to  variations  in  cooling  power  of  the  external 
medium. 

THERMOREGULATION    the  maintenance  or  regulation  of  temperature;  the  maintenance  of  a  particular  tem- 
perature of  the  living  body. 

TRAMMEL  NET    a  vertically  set  fishing  net  of  three  layers,  consisting  of  a  finely  meshed  net  between  two  nets 
of  coarse  mesh. 

TRANSDUCER    a  device  that  is  activated  by  power  from  one  system  and  supplies  power  in  another  form  to  a 
second  system. 

TURBIDITY    ( 1 )  the  cloudy  condition  caused  by  suspended  solids  in  a  liquid;  (2)  a  measurement  of  the  sus- 
pended solids  in  a  liquid. 

TYPING    the  act  of  delineating  homogeneous  areas  of  land  and  their  attendant  vegetation  cover  on  aerial  pho- 
tographs or  maps. 

ULTRASONIC    pertaining  to  acoustic  frequencies  above  the  range  audible  to  the  human  ear,  or  above  approx- 
imately 20,000  cycles  per  second 

UNGULATE    an  animal  having  hooves. 


850  Glossary 


USLE  (UNIVERSAL  SOIL  LOSS  EQUATION)    an  equation  used  to  design  water  erosion  control  systems:  A  = 
RKLSPC  wherein  A  is  average  annual  soil  loss  in  tons  per  acre  per  year;  R  is  the  rainfall  factor;  K  is  the  soil 
erodibility  factor;  L  is  the  length  of  slope;  S  is  the  percent  slope;  P  is  the  conservation  practice  factor;  and  C  is 
the  cropping  and  management  factor. 

UTM  (UNIVERSAL  TRANSVERSE  MERCATOR)    a  land  data  geographic  referencing  system  that  is  based  on  a 
series  of  60  zones  worldwide,  each  covering  6  degrees  of  longitude  in  a  north-south  strip. 

WARM -WATER  FISH    fish  that  can  stand  temperatures  ranging  from  33  degrees  Fahrenheit  to  80  degrees 
Fahrenheit. 

WEG  (WIND  ERODIBILITY  GROUP)    a  soil  erodibility  grouping  for  soils  based  on  the  stability  of  soil  aggre- 
gates against  breakdown  by  tillage  and  abrasion  from  wind  erosion.  The  WEG  is  used  in  the  wind  erosion 
equation  to  estimate  soil  loss  due  to  wind. 

WEIR    a  small  dam  in  a  river  or  stream  or  a  fence,  as  of  brush,  set  in  a  stream,  channel,  etc.  for  catching  fish. 

YAGI  ANTENNA    a  highly  directional  and  selective  shortwave  antenna  consisting  of  a  horizontal  conductor  of 
one  or  two  dipoles  connected  to  the  receiver  and  a  set  of  nearly  equal  insulated  dipoles  parallel  to  and  on  a 
level  with  the  horizontal  conductor. 

ZOOPLANKTON    unattached  microscopic  animals  of  plankton  having  minimal  capability  for  locomotion. 


Glossary  851 


INDEX 


Absolute  density,  353,  354,  356,  357,  361-364,  366,  448, 

463,  465,  488,  490,  497,  499,  504,  515 
Aerial  counts,  266,  363,  382,  535-538,  540,  544,  548 
Aerial  photography,  49,  54,  61,  62,  85,  99,  103,  109,  160, 

161,  164,  193,  330,  353,  360,  382,  397,  399,  405,  491, 

496,  542,  545,  551,  589,  758 
Aerial  photos,  54,  62,  65,  67,  222,  332,  729 
Aerial  tracking,  688,  689,  691 
Age  determination,  264,  266,  559,  562 
Air  temperature,  568,  714 
Alpine  tundra,  149,  150,  151,  153-156,  161,  167,  409,  547, 

548 
Analysis  of  variance,  656,  747 
Anderson,  Bertin,  169,  693 
Animal-created  structures,  140 
Animal  data  collection,  12 
Animal-made  features,  588,  598,  600 
Annual  forage  cycle,  520 
Antennas,  681-685 
Aquatic  habitat  features,  257 
Arctic  tundra,  149,  152-155,  157-159,  163-166,  351,  353, 

356,  373 
Aspect,  23,  65,  80,  131,  133,  14 1,  144,  157,  217,  248,  270, 

378,  410,  443,  480,  482,  526,  547,  548,  552,  589, 

591,  601,  614,  615,  691,  721,  760,  830 
Auditory  census,  423-425 
BIOSTORET,  634 
Bailey,  James  A.,  519,  711 
Band-tailed  pigeon,  411,  420,  426-428 
Barriers  to  communication,  830 
Bay  ducks,  375 
Beaver,  86,  89,  181,  182,  187,  191,  197,  198,  204,  240, 

372,  429,  430,  431  434,  435,  440,  444,  445,  447,  451, 

452,  485,  588,  598,  599,  710,  836 
Belt  transects,  328,  398,  399 
Bibliographies,  31,  33,  36,  38,  46,  300,  532 
Biological  limnology,  243,  246-248,  249 
Biotic  potential,  715 

Bird  species  diversity,  183,  197,  308,  310,  657,  660 
Birthing  habitat,  528 

Blue  grouse,  407,  409,  413,  414,  423,  425,  426-428,  722 
Boyd,  Raymond  J.,  519 
Breeding  bird  atlases,  297 

Breeding  bird  survey,  297,  303,  306,  309,  311,714 
Breeding  population,  306,  347,  363,  380,  381 
Broad-scale  systems,  84 
Brood  counts,  342,  381,  382,  385 
California  quail,  411,  419,  420,  722 
Call,  Mayo,  429 

Calling  surveys,  334,  339,  343,  348 
Canonical  correlation,  79,  753,  756 
Canopy  closure,  73,  82,  84,  483,  765,  770 
Canopy  volume,  82 

Capture  methods,  263,  266,  278,  284,  504,  516 
Carrying  capacity  approach,  523,  524 
Catch  per  unit  effort,  488,  490 
Cave-like  structures,  500,  501,  502 
Caves,  50,  80,  90,  99,  101,  104,  115,  124,  131,  136,  393, 

399,  486,  500,  501  502,  515,  517,  527,  533,  546,  552, 

568,  592,  593,  595,  760,  772 
Census  accuracy,  305,  399 
Chapman,  Joseph  A.,  453 
Chi-squared  test,  15,  19 


Chihuahuan  Desert,  125,  128,  130,  135,  144,  146,  281, 

290,  418,  724 
Christmas  bird  count,  298,  300,  308,  312 
Chukar,  130,  408,  409,  412,  426,  427 
Classification  systems,  50,  54,  66,  74,  77,  89,  96-99,  152, 

153,  203,  239,603 
Classifying  vegetation,  639,  640,  648,  650 
Cliffs,  50,  67,  80,  90,  104,  106,  116,  130,  131,  134,  136, 

153,  269,  292,  332,  333,  339,  373,  378,  388,  389,  393, 

399,  405,  431,  437,  499,  501,  502,  517,  527,  546- 

548,  553,  587,  588,  592-595,  740,  760,  772,  796 
Cluster  analysis,  21,  79,  540 
Coefficients  of  variation,  315,  327 
Cold-temperate  deserts,  125 
Cold-water  fishes,  258 
Colony  site  selection,  392 
Colony  size  stability,  390 
Colvin,  Bruce  A.,  679 
Common  bird  census,  300 
Community  measurements,  20,  23 
Community  suitability  index,  761,  765 
Community  surveys,  286 
Complete  census,  423 

Computer  system,  67,  450,  729,  730,  733-736,  738 
Condition  of  individual  animals,  20,  545 
Confidence  intervals,  235,  315,  329,  340,  398,  604,  745, 

790 
Conners,  Peter  G.,  351 
Contingent  value  method,  789,  790,  802 
Continuity  of  riparian  habitat,  174 
Cooperrider,  Allen  Y„  519,  587,  699,  757,  777 
Correction  factors,  395-399,  402,  703,  706 
Cottontails,  422,  423,  427,  451,  453,  460,  461,  462,  465, 

466,  467,  468,  470  471,  472,  473,  584,  722,  724 
Counts  at  roosts  and  colonies,  330 
Covariant  analysis,  71 1,  717,  718,  719 
Cover  board  technique,  644 
Cover  mapping,  75,  217,  758 

Critical  habitat  features,  79,  99,  101,  130,  183,  207,  240 
Cross,  Stephen  P.,  497 
Cuplin,  Paul,  225,  257,  603,  633 
Dabbling  ducks,  207,  220,  224,  371,  374-376,  379,  383, 

384,  770 
Data  analysis,  1,  2,  5,  6,  66,  67,  227,  236,  326,  605,  612, 

662,  689,  715,  716,  756,  805,  809 
Data  analyzing,  691 
Data  collection,  1,  2,  4,  5,  7,  9,  1 1,  12,  20,  21,  23,  25,  27, 

32,  76,  183,  243,  268,  271,  286,  315,  340,  523,  603, 

637,  696,  721,  746,  755,  768,  769,  777,  779 
Data  collection  priorities,  183 
Data  processing  cycle,  730,  731,  734 
Data  recording,  689,  737 
Data  storage,  7,  37,  64 
Dbh,  81,  82,  86,  101,  117,  176,  296,  596 
Dead  and  down  woody  material,  80-82,  85,  86,  88,  90. 

595,  596,  601,  760 
Decker,  Eugene,  829 
Density  estimates,  24,  182,  274,  327,  354,  357,  361.  363. 

366,  439,  450,  463,  465,  488,  640,  644,  646-648 
Density-frequency  relationship,  304 
Depletion  sampling,  19 
Descriptive  statistics,  742,  743 
Descriptors,  77,  203,  640,  744 


Index 


853 


Detection  of  echolocation  calls,  514 

Developing  habitat  models,  765 

Difference  methods,  701,  702 

Digestible  energy,  521,  525,  707 

Direct  effects,  712 

Direct  observation,  19,  330,  364,  379,  465,  497,  514,  526, 

680,  681,  694,  703,  704,  710 
Direct  search  methods,  274 
Discriminant  analysis,  91,  311,  752,  753,  756 
Dissolved  materials,  212,  231,  258 
Dissolved  oxygen,  135,  231,  240,  241,  244,  246,  247,  258, 

634,  636,  638 
Disturbance,  107,  124,  133,  143,  144,  152,  157,  160,  161, 
165,  167,  190,  192,  230,  330,  331,  339,  342,  348,  373, 
390,  395,  396,  400-405,  443,  485,  500,  504,  514, 
516,  528,  534,  536,  558,  583,  642,  687,  762,  775 
Diversity  index,  22,  28,  90,  304,  648,  649,  666,  674,  764 
Down  logs,  596 

Drinking  sites,  499,  500,  502,  504,  508,  512,  515 
Dropping  counts,  422,  540 
Economic  values  of  wildlife,  785,  786 
Ecotone(s),  23,  131,  133,  134,  142,  151,  153,  174,  175, 

270,  276,  348,  409,  413,  417,  581 
Edaphic  habitats,  90,  147,  584,  587,  588,  592,  601,  760 
Edge(s),  23,  28,  73,  80-82,  84,  85,  88,  90,  91,  134,  137, 
142,  147,  174,  175,  20-4,  215,  220,  221,  230,  355,  360, 
362,  364,  372,  393,  401,  408,  4 10,  41 1,  413,  414, 
416,  421,  437,  445,  450,  452,  507,  522,  527,  561,  581, 
605,  609,  618,  619,  648,  660,  682,  713,  715,  716, 
728,  837-839 
Eiders,  376-378 

Elevation,  23,  25,  65,  67,  80,  133,  141,  142,  157,  174,  177, 
181,  182,  190,  226,  231,  238,  258,  270,  361,  363, 
410,  413,  414,  417,  438,  443,  480,  482,  523,  526,  548, 
552,  581,  589,  591,  606,  614,  615,  619,  622,  623, 
636,  664,  665,  718,  721,  753,  760 
Emergents,  156,  207,  212,  213,  215,  216,  218,  219,  220, 

372,  376 
Encounter  rate  method,  354,  356,  357 
Endangered  fish  species,  257,  261 
Eng,  Robert  L.,  371,  407 
Erodibility,  575,  577,  582,  614 
Eutrophic,  220,  239,  241,  243,  372 

Evaluation,  2,  3,  5,  6,  7,  25,  28,  37,  77,  85,  86,  89-91,  108, 
121,  161,  162,  195,  201,  202,  217,  219-222,  224, 
245,  265,  266,  290,  299,  308,  312,  343,  346,  348,  353, 
361,  366,  369,  381,  385,  421,  426,  428,  476,  492, 
496,  534,  537,  539,  558-562,  564,  575,  581,  601,  606, 
612,  628-630,  632,  660,  696,  697,  698,  710,  724, 
738,  757,  758,  768,  770,  772,  775,  776,  777,  778,  780, 
782,  785,  791,  792,  796-798,  802,  818,  821,  830, 
834,  840 
Featured  species,  3,  77,  78,  85,  86,  307,  772 
Fecal  analysis,  558,  703-706,  710 
Field  sampling,  4,  6,  245,  628,  662,  663 
Field  techniques,  246,  289,  646 
Field-tracking,  686 
Fish  and  wildlife  reference,  32 
Fish  collection,  263 
Fish  preservation,  263 
Fishery  studies,  249 
Flight  line  counts,  396,  399 
Floating  platform.  393 
Flood  tide  counts.  365.  366 

Foliage  density.  173,  1  "6,  I  78,  189,640,642,644-651, 
653,655,656.659 


Foliage  height  diversity,  24,  82,  176,  178,  179,  640,  647, 

648 
Food  supply,  87,  89,  207,  231,  258,  260,  299,  344,  373, 

407,  408,  418,  476,  479,  524,  694,  699-701,  761,  766 
Forage  utilization,  699-701,  703,  704,  707,  709 
Foraging  area  approach,  523 
Forest  habitat  classification,  74 
Forest  openings,  80,  82,  420 
Frequency  distributions,  264,  744 
Frequency  of  measurements,  644 
Galton's  log-normal  model,  304 
Gambel's  quail,  113,  177,  411,  419,  420,  423,  424,  426, 

428,  658,  722 
Geese,  160,  165,  166,  206,  207,  371,  373-375,  379-382, 

384,  385,  723 
Geographic  information  system,  67,  591,  606,  728,  740, 
Geomorphic  habitat  features,  592,  760 
Goodness-of-fit  test,  19 
Gray  partridge,  408,  409,  426-428 
Grazed-plant  methods,  701,  702 
Great  Basin  Desert,  123-125,  128-130,  134,  138,  142-145, 

147 
Ground  nest  counts,  397,  398 
Ground  surveys,  264,  334,  338-340,  342,  544 
Ground-water,  170,  192,  193,  261,  613,  614,  623,  628 
Guilds,  3,  10,  20-22,  26,  28,  74,  78,  79,  85-90,  96,  108- 

112,  116,  121,  147,  188,  224,  288,  290,601,  761,  765, 

770,  772,  774,  776 
HEP  (see  Habitat  evaluation  procedures) 
HSI  (see  Habitat  suitability  index) 
Habitat  capability  models,  772,  775 
Habitat  capability  relation,  25 
Habitat  components,  3,  7,  17,  18,  21,  22,  23,  24,  25,  26, 

52,  82,  90,  124,  130,  131,  133,  136,  137,  141,  144, 

174-177,  179,  181-183,  268,  270,  271,  288,  428,  472, 

482,  491,  552,  587,  598,  600,  662,  757,  758,  759,  760, 

761,  762,  763,  764,  765,  767,  797 
Habitat  delineation,  580 

Habitat  evaluation  procedures,  25,  86,  90,  91,  121,  221, 
222,  224,  265,  266,  343,  770,  772,  775,  776,  796,  802, 
803,  830 

Habitat  layer  index,  103,  109 

Habitat  management,  1,  2,  27,  49,  54,  66,  79,  85,  124,  146, 
147,  163,  181,  183,  194-196,  199,  225,  289,  346, 
349,  408,  421,  424,  425,  524,  528,  529,  531,  557,  562, 
564,  567,  575,  578,  580,  581,  583,  662,  679,  699, 
754,  759,  762,  792,  822 

Habitat  mapping,  4,  6,  25,  49,  54,  55,  67,  166 

Habitat  models,  28,  82,  86,  96,  315,  342,  343,  757,  759, 
760,  761,  762,  763,  764,  765,  768,  770,  775,  776 

Habitat  quality  index  (HQI),  233,  265 

Habitat  requirements,  5,  21,  26,  77,  78,  80,  86,  147,  188, 
260,  292,  315,  347,  364,  374,  375,  379,  384,  419,  428, 
454,  475,  482,  491,  520,  528,  529,  533,  534,  537, 

539,  541,  542,  543,  544,  545,  546,  550,  551,  563,  564, 
661,  694,  695,  699,  715,  758,  767,  770 

Habitat  suitability,  7,  18,  25,  28,  86,  89,  90,  221,  223,  265, 
266,  343,  346,  356,  392,  426,  492,  520,  525,  534, 

540,  557,  570,  573,  580,  581,  583,  596,  757,  764,  765, 
770,  776,  796,  797,  802,  803 

Habitat  suitability  index,  7,  9,  18,  25,  28,  86,  89,  90,  96, 
116,  221,  222,  223,  265,  266,  343,  346,  426,  596,  757, 

762,  764,  765,  770,  771,  772,  775,  776,  796,  797, 
802,  803 

Habitat  type,  6,  12,  14,  16,  21,  23,  25,  36,  54,  66,  80,  274, 
286,  326,  335,  338,  408,  421,  427,  434,  443,  444,  449, 


854 


Index 


450,  475,  482,  491.  538,  583,  584,  649,  655,  686, 

691,  770,  797 
Hand  capture,  504 
Hand  tracking,  686 
Handling  techniques,  504,  505 
Hegdal,  Paul  L„  679 
Hiders,  527 
Horizonation,  568 
Horizontal  diversity,  647-649 

Horizontal  vegetation  structure,  138,  139,  268,  271 
Hydrologic  soil  group,  575,  581 
Hydrology  of  lakes,  623 
Hypothesis  formulation,  746,  747 
Impacts  on  desert  habitats,  142 
Impoundments,  185,  191,  218,  223,  224,  239,  240,  251, 

261,  372,  385 
Indicator  species,  22,  78,  79,  85-88,  288,  307,  311,  439, 

665 
Indirect  effects,  124,  71 1,  712,  713 
Indirect  observation,  703 
Inferential  statistics,  66,  742,  745,  746 
Instream  flow  needs,  227,  605 
Integrated  mapping,  64 
International  bird  census,  300,  310 
Interpretation,  1,  2,  5,  6,  21,  62,  65,  85,  99,  103,  107,  165, 

222,  232,  274,  304,  313,  343,  344,  355,  494,  514, 

554,  575,  585,  589,  601,  616,  623,  628,  662,  696,  707, 

709,  727,  751,  753,  758,  777,  778,  782,  805,  811,  822 
Jaccard's  similarity  coefficient,  21 
Jackrabbits,  453,  454,  455,  465,  467,  468,  470,  471,  472, 

473,  478,  479 
Jones,  K.  Bruce,  1,  11,  123,  267 
Jumping  mice,  433 
Kangaroo  mice,  432 

Kangaroo  rats,  104,  136,  137,  432,  572,  582 
Kerr,  Pochard  M.,  49 
Key  species  approach,  523,  524 
King  strip  census,  422 
Kochert,  Michael,  313 
Laboratory  analysis,  662,  665 
LaBounty,  James,  237 

Landforms,  491,  588,  589,  590,  591,  592,  600,  601 
I.andsat  imagery,  158,  160,  166 
Landsat  images,  62,  5-»0 
Late  nesters,  380,  390,  391 
Lava  flows,  80,  592,  593,  594,  760 
Lemmings,  432,  433,  451 
Length  of  fish,  262,  264 
Lent,  Peter  C,  149,  519 
Lentic  habitats,  134,  135 
Levels  of  data  needed,  1,4,  11 
Levels  of  inventory,  243 
Life-form  approach,  79,  772,  775 
Life  stages,  258,  261,  263,  360,  607 
Life  zones  of  a  lake,  241 
Lincoln  estimator  (see  Lincoln  index) 
Lincoln  index,  19,  306,  398,  436,  437,  444,  446,  448,  449, 

450,  467,  490 
Line  transect,  24,  219,  302,  310,  326-328,  345,  356,  358, 

366,  368,  444,  451,  473,  558 
Linear  correlation,  656,  717 
Linear  density  transects,  362,  363 
Linear  regression,  657,  749,  750,  767 
Literature  searching,  29,  31,  33,  38 
Litter,  22-24,  65,  80,  101,  115,  116,  130-132,  138,  141, 

178,  182,  267-269,  271,  276,  283,  288,  441,  443,  470, 

472,  478,  582,  584,  588,  594-596,  760,  761 


Litter  and  mulch,  595,  596 

Livestock  grazing,  2,  4,  142,  143,  171,  173,  180,  185-190, 

195,  197-199,  234,  426,  534,  595,  605,  662,  772 
Logging  and  roads,  192 
Logs,  22,  80-82,  118,  124,  131,  132,  135,  138,  182,  183, 

187,  205,  231,  267-269,  274,  275,  276,  281,  432,  437, 

440,  443,  484,  486,  503,  596,  597,  609,  761 
Loomis,  John,  785 
Lotic  water,  1 34 
Macroinvertebrates,  46,  197,  202,  216-218,  221,  227,  229- 

231,  258,  607,  611,  661-663,  665,  669-671,  761 
Mangum,  Fred,  661 
Man-made  structures,  132,  140,  141,  326,  392,  417,  419, 

433,  485,  500,  502,  514,  588,  596,  598-600 
Mann-Whitney  U-Test,  304,  750,  751 
Manning  equation,  622 
Mapping  variables,  578 
Mark  and  recapture,  15,  19,  20,  264,  281,  306,  436,  437, 

444,  446,  449,  463,  467,  530,  536,  559 
Marked  plot  strip  transect,  355,  362 
Marking  techniques,  285,  289,  507,  694 
Marmots,  153,  432,  437,  587 

Marsh  succession,  215 

Masked  bobwhite,  99,  146,  411,  418,  426,  428 

Matney,  Iris,  727 

Matrix  models,  772 

Maximum  effort  monitoring,  245 

Maximum  level  inventory,  244 

Mearn's  quail,  410,  417,  418,  426 

Measurement  of  evaporation,  615 

Measurement  of  precipitation,  615 

Measurement  of  snow,  615 

Measures  of  central  tendency,  743 

Measures  of  dispersion,  743 

Mergansers,  207,  221,  376-379 

Minimum  effort  inventory,  243 

Minimum  effort  monitoring,  245 

Mining,  2,  7,  21,  143,  190,  234,  241,  309,  345-347,  580, 
599,  611,  614,  633,  637,  638,  662 

Mist-netting,  499,  504,  508,  512,  514 

Mojave  Desert,  28,  125,  127,  130,  275,  289,  471 

Moles,  431,  440,  591,  636 

Monitoring,  1,  2,  4-9,  11,  12,  14,  15,  17-19,  24-27,  29,  30, 
32,  33,  49,  54,  61,  62,  67,  73,  74,  77-80,  84-91,  96- 
99,  123,  124,  131,  138,  141,  142,  144,  145,  152,  153, 
158-161,  163,  164,  166,  201,  202,  217,  218,  222, 
225,  231,  233,  237,  243,  245,  251-253,  257,  265,  266, 
268,  275,  286,  288,  289,  291,  292,  295,  296,  304, 
305,  306,  307.  31 1,  313,  314,  315,  328,  330,  331,  339, 
342-344,  348,  351,  356,  363,  369,  371,  379,  387, 
390,  391,  394,  395,  401,  402,  407,  430,  434,  436,  437, 
438,  439,  442,  443,  444,  447,  448,  450,  453,  476, 
491,  497,  508,  516,  517,  519,  534,  538,  543,  544,  548, 
550,  551,  554,  559,  585,  595,  598,  600,  601.  610- 
613,  628,  632,  634,  637,  638,  639,  660-663,  6^4,  681, 
695-697,  699,  707,  709,  710,  711-718,  721,  722,  724, 
741,  742,  754,  756-759,  762,  765,  767-769,  775,  776, 
777,  779,  782,  787,  805 

Monitoring  methods  for  lakes,  243,  252,  253 

Monitoring  studies,  1,  2,  4,  7,  9,  12,  19.  27,  29,  33,  54,  62. 
73,  123,  124,  144,  145,  245,  265,  268,  387,  430,  442, 
443,  450,  497,  699,  711,  754,  782 

Mourning  dove,  1 1 3,  4 1 1 ,  42 1 ,  423,  426,  427,  658,  659 

Multivariate  analysis,  28,  89,  90,  147,  311,  562,  741,  751, 
754,  767 

Muskrats,  204,  206,  207,  218-221,  223,  429,  431.  435,  436, 

445,  446,  451,  452,  484,  576 


Index 


855 


Nest  counts,  219,  304,  381,  397,  398,  400,  404 
Nest  monitoring,  296,  305,  306 
Nest  searches,  306,  337,  339,  367 
Nest  site  selection,  345,  385,  392,  405 
Nest  spot  mapping,  357 

Nesting,  3,  7,  9,  13,  24,  62,  80,  81,  85,  109,  110,  116-120, 
130,  131,  135,  136,  138-141,  146,  157,  158,  160, 
164-166,  176,  178,  179,  194,  195,  198,  204,  207,  218, 
220,  222,  224,  292,  293,  296,  306,  307,  309-312, 
314,  315,  326,  329-343,  345-349,  351-358,  364,  366, 
367,  369,  371-376,  378,  380,  381,  383-385,  387-405, 
408,  409,  411,  412,  419-421,  428,  587,  592,  594,  596, 
598,  599,  691,  694,  697,  704,  762,  763,  764 
Nesting  surveys,  326,  329,  331-334,  340,  342,  345 
Non-nesters,  391 

Nonparametric  statistics,  745,  750 
Nutritional  approach,  523-525 
Ohmart,  Robert  D.,  169,  639 
Old  World  rats,  433 
Oligotrophic,  220,  239,  243 
Oligotrophic-eutrophic  series,  239 
Open  water  counts,  366 
Optimal  estimation,  754 
Ordination  analysis  program,  153 
Overlap  indexes,  22 
PATREC  models  (see  Pattern  recognition),  25,  763,  770, 

772,  775 
Parent  material,  568,  582 

Patchiness,  24,  28,  139,  147,  152,  164,  176,  178-180,  189, 
311,  358,  361,  471,  473,  640,  641,  642,  646,  647,  648 
Pattern  recognition,  25,  28,  763,  770,  772,  775,  776 
Permafrost,  151-153,  157-160,  164,  166 
Permanent  lentic,  131,  134,  269 
Peterson,  Larry,  727 

Ph,  90,  196,  197,  199,  207,  208,  221,  231,  240,  246,  247, 
258,  308,  311,  345,  346,  368,  369,  385,  428,  451,  525, 
559,  562,  568,  573-575,  581,  582,  584,  633-638,  665, 
697,  761,  789,  824 
Phenological  characteristics,  520,  542 
Photographic  counts,  363 

Physical  features,  23,  80,  106,  258,  292,  314,  352,  353, 
372-375,  378,  409-411,  475,  482,  499,  500,  515,  523, 
587-589,  595,  598,  600,  603,  760,  761,  772 
Physical-chemical  surveys,  246 

Pikas,  104,  153,  453,  462,  463,  465,  467,  471,  472,  473 
Plant  communities,  22,  54,  73,  77,  79-82,  84,  85,  89,  91, 
121,  122,  147,  158,  159,  161,  163,  167,  169,  173,  175, 
179,  183-185,  191,  192,  195,  201,  207,  222,  223, 
242,  271,  310,  417,  453,  455,  456,  468,  475,  479,  480, 
482,  484,  486,  581,  583-584,  639,  660,  772 
Plant  community  size,  174 
Playas,  80,  115,  116,  588,  592,  593,  594 
Plot  sampling,  24 
Plotless  sampling,  24 

Pocket  gophers,  432,  486,  572,  582,  584,  591 
Pocket  mice,  136,  432,  569,  582 
Point  count  method,  304.  326 
Point  data,  23,  593 
Point-ccnter-quarter,  18,  23,  24,  145 
Pool  quality,  234,  603,  607,  608 
Pools,  13,  131,  134-136,  145,  156,  215,  224,  227,  229, 

231,  234,  269,  275,  277,  279,  360,  362,  503,  513,  607, 
609,  724 
Population  levels,  81,  183.  32S,   t24,  453.  -+54,  467,  468, 

470,  471,  496,  533.  540,  545,  ^21.  723,  821 
Population  measurements.  20,  261.  271,  284,  297,  315, 
342,  353,  357,  361,  364,  379,  394,  121,  433,  463,  487, 


497,  504,  529,  531-534,  537,  538,  540,  542,  544, 
545,  548,  551,  554,  765 

Population  structure,  16,  20,  24,  165,  195,  275,  348,  473, 
531,  549,  555,  560 

Porcupines,  433,  438,  698 

Prairie  chicken,  410,  415,  416,  426,  427,  492 

Prairie  dogs,  104,  432,  436,  437,  585,  687 

Precipitation,  12,  20,  26,  123,  125,  127-129,  131,  133, 
141-144,  190,  207,  239,  270,  272,  282,  284,  286,  291, 
326,  329,  335,  448,  467,  470,  472,  482,  523,  543, 
550-552,  568,  574,  575,  581,  603,  613-616,  623,  637, 
713,  715,  716,  720,  722,  723.  753,  824 

Preliminary  field  work,  4,  5 

Presence,  12,  14,  16,  17,  21,  23,  25,  67,  73,  80,  86,  99- 
101,  104,  106,  116,  134,  139-141,  151,  152,  157,  159, 
161,  162,  173,  175,  179,  182,  185,  203,  215-217, 
221,  258,  263,  264,  271,  274,  288,  297,  304,  314,  315, 
330-332,  341,  353,  356,  357,  360,  361,  363,  364, 
366,  379,  393,  395,  421-423,  425,  430,  431,  432,  433, 
434,  435,  436,  437,  438,  441,  443,  447,  463,  465, 
467,  487,  488,  489,  491,  497,  499,  500,  504,  512,  513, 
514,  515,  525,  526,  529,  530,  546,  548,  552,  569, 
573,  575,  576,  581,  593,  598,  600,  611,  639,  642,  643, 
648,  685,  703,  759-764,  770,  772 

Principal  component  analysis,  296,  656,  753 

Prioritizing  of  objectives,  4 

Probability  of  capture,  19,  444,  446,  448,  449,  450 

Problem  definition,  1,  2 

Productivity,  2,  6,  16,  20,  24,  26,  84,  146,  152,  160,  164- 
167,  170,  173,  177,  186,  187,  197,  207,  223,  234,  239, 
247-249,  251-253,  260,  309,  311,  313,  332,  333,  340- 
342,  344-348,  374,  383,  385,  426,  447,  448,  452, 
494,  524,  525,  531,  533,  548,  550,  554,  559,  598,  611, 
614,  637,  664,  665,  698,  711,  754,  764,  803 

Productivity  surveys,  340-342 

Properties  of  water,  258,  360,  613,  633,  634,  761 

Ptarmigan,  153,  154,  157,  162,  163,  165,  409,  414,  426, 
428 

Quality  criteria,  633,  635,  638 

RAIDS  (Riparian/Aquatic  Information  Data  Summary),  265 

Receivers,  681,  683,  685 

Recreational  activities,  173,  190 

Regression  analysis,  659,  719,  749,  788 

Relative  abundance,  11,  19,  22,  24,  66,  67,  84,  185,  198, 
245,  262-264,  274-276,  280,  281,  295,  304,  315,  326, 
331,  337,  357,  380,  421,  422,  425,  463,  465,  486,  487, 
488,  489,  492,  494,  529,  540,  759,  761,  763-765 

Relative  density,  291,  335,  337,  354,  356,  357,  361,  363- 
366,  430,  433,  434,  437,  438,  439,  443,  444,  447,  450, 
488,  489,  499,  504,  515,  529,  530,  548,  554,  648, 
765,  770,  772 

Renesting,  381,  391 

Repeatability,  233-235,  604,  645 

Riffles,  13,  134,  227,  229,  231,  607,  609 

Ring-necked  pheasant,  408,  409,  412,  426,  428,  582 

Riparian  habitat  classification,  173 

Riparian  vegetation,  54,  68,  170,  172,  175,  184,  187-189, 
192,  194-196,  225,  227,  228,  232,  236,  258,  480,  642, 
650,  655,  660 

Riparian  zones,  80,  81,  91,  180,  198,  234,  257,  336,  772 

Road  counts,  326,  327,  328,  337 

Roadside  counts,  306,  422,  423,  427,  465 

Rock,  16,  23,  64,  80,  104-106,  112,  113,  115,  116,  119, 
124,  131,  134,  136-138,  14 1,  142,  153,  157,  197,  207, 
229,  231,  234,  235,  242,  269,  314,  378,  384,  389, 
409,  414,  432,  442,  467,  501,  502,  507,  533,  552,  568, 
569,  574,  578,  579,  581,  583-584,  589,  590,  593, 


856 


Index 


594,  595,  598,  599,  604,  608,  609,  619,  760,  831 
Rock  fragments,  569,  583,  609 
Roosting  flock  counts,  363 
Roosting  sites,  361,  362,  484,  500,  501,  502,  503,  504, 

512,  514,  515 
Rope-drag  nest  mapping,  354,  358,  367 
Ruddy  ducks,  215,  375,  376,  381 
Ruffed  grouse,  410,  413,  414,  416,  421-423,  426-428,  698, 

722,  760 
Rumen  analysis,  703,  706 

Runs,  13,  134,  229,  231,  264,  326,  327,  440,  441,  616,  753 
RUNWILD,  32,  37,  90,  120,  121,  772,  776 
Ryder,  Ronald  A.,  291 
STORET,  46,  634 

Salinity,  188,  204,  208,  221,  360,  575,  581,  583,  585 
Sample  counts  of  nests,  396 
Sample  plots,  327,  362,  363,  396,  399,  422,  441 
Sample  size,  5,  12,  19,  144,  295,  326,  327,  340,  449,  628, 

648,  666,  745-747  750 
Sampling  considerations,  1 1 ,  266 
Sampling  design,  11,  12,  28,  87,  303,  313,  344,  355,  660, 

662,  675,  742,  756 
Sampling  methods,  5,  14,  23,  24,  66,  272,  274,  662 
Sampling  of  desert  wildlife,  144 
Samson,  Fred  B.,  475 
Sand  dunes,  80,  157,  486,  592-594,  760 
Scaled  quail,  146,  411,  418,  419,  423,  426-428 
Scoping,  1,  2,  768,  769 
Scoters,  376-378 
Seasonal  precipitation,  715 
Seasonal  water  table,  575 
Selecting  specific  methods,  4,  5 
Sex  and  age  ratios,  20,  524,  534,  535,  557,  559 
Shannon- Weaver  index,  22 

Sharp-tailed  grouse,  410,  413,  416,  424,  427,  428 
Short,  Henry  L,  93 

Shrews,  154,  429,  432,  440,  441,  442,  447,  487 
Sight  distance,  82,  84 
Simpson's  similarity  coefficient,  21 
Siren-elicited  vocalization,  489 
Slope,  23,  65,  80,  119,  131,  133,  141,  143,  144,  165,  166, 

181,  192,  270,  271,  352,  353,  356,  360,  361,  393,  409, 

415,  443,  453,  480,  482,  526,  548,  553,  573,  575, 

578,  579,  587,  589-591,  594,  599,  606,  609,  614,  615, 

622,  748-750,  760,  766,  787,  794 
Snags,  18,  23,  24,  80-82,  85,  86,  88,  91,  101,  134,  138, 

175,  176,  178,  179,  183,  190,  197,  310,  315,  484,  503, 

588,  595,  596,  599,  601,  760,  761 

Snow  cover,  62,  153,  154,  160-162,  164-167,  258,  330, 
409,  416,  487,  526,  542,  543,  550,  551,  562,  563,  724 

Snow  depth,  416,  470,  526,  614,  615,  616,  715-720,  722 

Snow  quality,  721 

Snowshoe  hare,  181,  427,  457,  468,  469,  470,  472,  473, 
482,  492,  723 

Social  interactions,  345,  392,  393,  487 

Soil(s),  2,  3,  5,  12,  23,  51,  52,  54,  62,  65,  66,  75,  76,  79, 
80,89,93,94,99-101,  105,  106,  115,  119,  124,  125, 
131,  132,  133,  135-138,  141,  142,  144,  146,  153,  157- 
159,  165,  166,  170,  173,  175,  177,  178,  180,  182, 
183,  186,  187,  191,  192,  199,  201,  203,  207,  213,  215, 
216,  217,  220,  222,  228,  232,  234,  235,  243,  257, 
258,  268-271,  276,  283,  284,  353,  389,  393,  39-4,  401, 
432,  434,  439,  443,  453,  455,  491,  567-585,  588, 

589,  591,  592,  594,  596,  598,  600,  604,  606,  613,  614, 
616,  652,  716,  718,  720,  747,  748,  753,  756,  760, 
761,  826 


Soil  color,  570,  572,  584 

Soil  consistence,  570,  571 

Soil  depth,  23,  137,  138,  268,  572,  576,  584 

Soil  drainage  classes,  573 

Soil  fertility,  567,  575,  576,  584,  585 

Soil  reaction,  573 

Soil  structure,  570,  571,  573 

Soil  taxonomy,  576,  578,  585 

Sonoran  Desert,  123-125,  127-130,  138-140,  142-144,  147, 

182,  198,  271,  274,  275 
Special  lake  types,  239 
Species  composition,  22,  23,  84,  128,  158,  161,  172,  173, 

177,  184,  185,  188,  190,  208,  218,  219,  263,  271,  282, 

286,  296,  328,  353,  360,  372,  373,  375,  376,  378, 

387,  395,  409-411,  413,  420,  424,  522,  523,  554,  640, 

642,  701,  759,  760 
Species  diversity,  17,  20,  22,  23,  129,  130,  139,  140,  142- 

144,  146,  147,  156,  183,  197,  222,  304,  308,  310,  515, 

657,  660,  761,  764,  765 
Specific  conductance,  212,  231,  636,  638 
Speich,  Steven  M.,  387 
Spot-mapping  census,  300 
Spowart,  Richard  A.,  475 
Spruce  grouse,  409,  412,  413,  427,  428 
Squirrels,  104,  136,  432,  437,  438,  439,  440,  444,  446, 

451,  486,  576,  582,  591,  697,  724 
Stand  size,  73,  82,  84,  86,  655 
Stone,  James  E.,  567 
Stream  channel  stability,  225,  227,  231,  232,  603,  606, 

607,  614,  628,  629-630 
Stream  classification,  225,  226 
Stream  diversity,  227,  231,  234 
Stream  gradient,  227,  230,  258,  603,  606,  611,  666 
Stream  habitat  features,  225,  227,  234 
Stream  habitat  inventory,  231,  232,  612 
Stream  order,  226,  227 
Stream  water  chemistry,  227,  230 
Stream  width  and  depth,  225,  230 
Streambank  stability,  196,  227,  231,  603,  605 
Streambed,  174,  225,  227,  229-232,  234,  603,  607,  608, 

611,  618,  663,  665,  779 
Streambed  material,  229,  607,  608 
Streamflow,  227,  233,  263,  603-605,  612,  616,  619,  622- 

626,  628,  632,  634 
Streamflow  measurement,  605,  616,  619,  623,  628 
Streamflow  pattern,  227,  603 
Strip  census,  422,  423,  427,  444,  544 
Strip  transect,  309,  355,  358,  362,  398,  399,  722 
Structural  conditions,  17,  73,  77,  82,  85,  310 
Structural  types,  173,  174,  177,  184,  185,  651,  652,  653, 

655 
Structure,  4,  12,  16,  17,  20,  21,  24,  25,  28,  62,  66,  74,  80, 

81,  90,  93,  94,  96,  97,  99-101,  103,  104,  106-110,  115, 

116,  119,  121,  123,  128,  131-142,  146,  147,  165, 

172,  179,  181,  182,  184,  188,  189,  195,  201,  206,  207, 

212,  213,  258,  268-271,  275,  286,  289,  292,  296. 

303,  304,  308-312,  314,  328-330,  348,  352,  353,  359. 

372-376,  378,  392-394,  404,  405,  419,  420,  424,  453. 

473,  476,  480,  494,  495,  499,  500,  504,  512.  515.  520. 

523,  526,  531,  534,  540,  544,  549,  551,  554,  5SS 

557,  560,  568,  570,  571,  573,  574,  588,  596,  599,  614, 

619,  640,  642,  652,  660,  665,  674,  699,  705,  732, 

736,  739,  759,  760,  764,  772,  797 
Study  objectives,  1,  4,  5,  145 

Submergents,  203,  213,  215,  216,  218-221,  372,  375 
Subsurface  chamber,  393 


Index 


857 


Succession,  25,  68,  89,  161,  173,  196,  215,  220,  224,  304, 

308,  554,  556,  644,  772,  775,  776 
Successions  stages,  73,  77,  80-82,  85,  91,  203,  221,  420, 

772,  775 
Suspended  sediment,  6 14,  628,  631,  638 
Suspended  solids  and  turbidity,  230 
Swans,  166,  207,  371-373,  384,  385 
T-test,  747,  750 

Talus,  80,  85,  90,  101,  104,  115,  116,  131,  136,  153,  157, 
269,  393,  432,  438,  453,  462,  463,  465,  483,  486,  501, 
502,  517,  568,  587,  592-596,  599,  760 
Talus  slopes,  104,  136,  157,  269,  393,  432,  453,  462,  465, 

483,  501,  594 
Tame  animals,  547,  703,  704,  706 
Tape  playback  response,  365,  366 
TDS,  (see  Total  dissolved  solids) 
Temporary  lentic,  23,  131,  135,  269 
Territory  spot  mapping,  354 
Texture,  62,  80,  159,  182,  268,  271,  500,  568-570,  572, 

574-576,  578,  579,  581-584,  592,  597 
Thomas,  Jack  Ward,  73 
Time  series  analysis,  754 
Toe-pace  transect,  24 

Total  dissolved  solids,  172,  192,  231,  233,  525,  636 
Total  population  counts,  363   ,  •":  „*,'-'  •  .. 

Transect  methods,' 24,  302,  363,'  .426:.-    ,  '       "oV- 
Transmitters,  681,  682,-683,  685;,  68'7;  689,'.691,  697,  698 
Trapping,  14;  15,  100,  145V163,,181,T85,  2^9, 273,  274, 
276,  279-284;  289; -306^  307;  4jQ8,, 4 17,^430,  434,  435, 
436,  437,  438,  439, '440,  441,  442,  443,'444,  446, 
448,  450,  449,  452,  488,  489,  490,  504; 5,1 2,  514,  536, 
556,  562,  563,  680,  686,  694,  697,  714  ' 
Travel  cost  method,  788,  790,  802 
Tree  height,  82,  753 
Tropical-subtropical  deserts,  125 
Turbidity,  131,  192,  202,  220,  230,  233,  241,  261,  269, 

603,  611,  634,  637,  638  711,  761 
Understory  volume,  82,  84 
Unmarked  strip  transect,  355,  362 
Van  Haveren,  Bruce  P.,  613 
Variable-circular,  295,  305,  306,  309,  311,  366 
Variable  width  line  transect,  356,  358,  366 
Vegetation  dynamics,  215 
Vegetation  maps,  184,  330 

Vegetation  measurements,  2,  3,  5,  6,  12,  13,  21,  23-25,  28, 
50-54,  61,  62,  65-68,  74,  76,  80-82,  84,  89,  90,  93-101, 
103,  U0,  115,  119,  121,  123-125,  128,  130-133,  135, 
137-144,  146,  147,  149,  151-153,  155,  156,  158-167, 
170-172,  174,  175,  177-197,  201-213,  215-225,  227, 
228,  232,  233,  236,  241,  242,  250,  258,  261,  266,  268- 
271,  276-278,  288,  291,  292,  296,  304,  308-311,  314, 
327,  330,  335,  336,  338,  345,  347,  352,  353,  355- 


360,  362,  364-366,  372-376,  378-380,  382,  389,  393- 
395,  397,  398,  408-411,  414,  416,  419,  420,  424,  427, 
429,  430,  431,  432,  433,  434,  435,  438,  439,  447, 
455,  457,  460,  461,  466,  469,  470,  471,  473,  475,  477, 
479,  480,  482,  483,  486,  487,  491,  503,  510,  523, 
526,  527,  533,  537,  538,  540,  543-548,  550-554,  556, 
558,  567,  572,  575,  580,  581,  583,  587-589,  591, 
592,  594-596,  598-600,  606,  609,  614,  615,  633,  639, 
640,  642,  643,  644,  646,  647,  648,  649,  650,  651,  652, 
653,  654,  655,  656,  657,  659,  660,  665,  679,  687, 
699,  700,  702,  708,  710,  743,  744,  753,  759-761,  764, 
770,  772,  776,  779,  810,  811 

Vegetative  structure,  81,  292,  296,  314,  495 

Vehicle  tracking,  683,  687,  688 

Verner,  Jared,  73 

Vertical  diversity,  647,  649, 

Vertical  habitat  availability,  268 

Vertical  vegetation  structure,  139 

Visual  aids,  831-832,  833,  835,  836,  838,  840 

Voles,  104,  429,  432,  433,  440,  451,  478,  484,  582,  585, 
698 

Wagoner,  Ora,  29 

Warm-temperate  deserts,  125 

Warm-water  fish,  242,  258,  636,  661 

Watchers,  527 

Water  chemistry,  155,  172,  207,  227,  230,  231,  233,  246, 
633,  634,  662,  664 

Water  depth,  135,  155,  203,  206,  212,  216-219,  235,  241, 
249,  262,  352,  360,  361,  366,  604,  605,  619 

Water  quality,  135,  172,  181,  187,  190,  192,  194,  196, 

208,  225,  226,  236,  244,  252,  253,  258,  485,  525,  575, 
612,  616,  628,  632-638,  661,  662,  665,  666,  674, 
761,  818 

Water  temperature,  131,  172,  230,  231,  232,  241,  258, 
269,  603,  610,  611,  664 

Watershed  stability,  61 4 

Weight  of  fish,  262,  264 

Weller,  Milton  W.,  201 

West,  William  H.,  741 

Wild  turkey,  410,  4 16,  427,  428,  578 

Wildlife  guilds,  10,  28,  90,  96,  108-112,  116,  121,  147, 
765,  776 

Willner,  Gale  R,  453 

Winter  bird  population  study,  299,  309-311 

Winter  bird  survey,  305,  31 1 

Winter  habitat,  166,  315,  330,  372,  379,  383,  384,  413, 
470,  484,  550,  551,  557,  558,  767 

Winter  surveys,  382 
Wintering  habitat,  378,  382,  383 
Woodrats,  431,  433,  437,  439,  572 
Writing  strategy,  a,  806,  811 
Zimmerman,  Donald  E.,  805 


£Be5$:i£  cenie* 


858 


Index 


Q 

W 

U  2 

H  K 

<  D 

Q  H 

U 

c 
c 

<o 

CD 

c*. 

1^. 
00 

o 

Q. 

O 

BLM  LIBRARY 
PDEONv!°RXCO0fc0047 


m       -^ 


•i  7 


U.S.  Department  of  the  Intern 
Bureau  of  Land  Management      * 


--■- 


&'& 


'Wi 


_*      ■■■■*■»< 


jr^^pP^ 


y&t 


.  ■  ■  - 


•* 
<* 


J 


JET        'jv-. 


'-  'r  «* 


; 


</  #  ■  •■■"■ 


£3&S 


i 


F*> 


i£  \       fc"  : 


it. 

v.-       ''■ 


T 


^9 


jp*»?.\i+  '■