Skip to main content

Full text of "كتب في علم الميكروبيولوجي"

See other formats


'.YMRCKIUM 






Volt 




micro-organisms 

ms- 



■ 



advances in 

geomicrobiology 



CAMBRIDG 



.Cambridge. org/9 /80521 862221 



Micro-organisms and Earth systems - 
advances in geomicrobiology 

There is growing awareness that important environmental transformations are catalysed, 
mediated and influenced by micro-organisms, and such knowledge is having an increasing 
influence on disciplines other than microbiology, such as geology and mineralogy. Geo- 
microbiology can be defined as the study of the role that microbes have played and are 
playing in processes of fundamental importance to geology. As such, it is a truly inter- 
disciplinary subject area, necessitating input from physical, chemical and biological 
sciences. The book focuses on some important microbial functions in aquatic and terrestrial 
environments and their influence on 'global' processes and includes state-of-the-art 
approaches to visualization, culture and identification, community interactions and gene 
transfer, and diversity studies in relation to key processes. Microbial involvement in key 
global biogeochemical cycles is exemplified by aquatic and terrestrial examples. All major 
groups of geochemically active microbes are represented, including cyanobacteria, bacteria, 
archaea, microalgae and fungi, in a wide range of habitats, reflecting the wealth of diversity 
in both the natural and the microbial world. This book represents environmental 
microbiology in its broadest sense and will help to promote exciting collaborations between 
microbiologists and those in complementary physical and chemical disciplines. 

Geoffrey Michael Gadd is Professor of Microbiology and Head of the Division of Environmental 
and Applied Biology in the School of Life Sciences at the University of Dundee, UK. 

Kirk T. Semple is a Reader in the Department of Environmental Science at Lancaster University, 
UK. 

Hilary M. Lappin-Scott is Professor of Environmental Microbiology in the School of Biosciences, 
University of Exeter, UK. 



Symposia of the Society for General Microbiology 

Managing Editor: Dr Melanie Scourfield, SGM # Reading, UK 
Volumes currently available: 

43 Transposition 

45 Control of virus diseases 

47 Prokaryotic structure and function - a new perspective 

51 Viruses and cancer 

52 Population genetics of bacteria 

53 Fifty years of antimicrobials: past perspectives and future trends 

54 Evolution of microbial life 

55 Molecular aspects of host-pathogen interactions 

56 Microbial responses to light and time 

57 Microbial signalling and communication 

58 Transport of molecules across microbial membranes 

59 Community structure and co-operation in biofilms 

60 New challenges to health: the threat of virus infection 

61 Signals, switches, regulons and cascades: control of bacterial gene expression 

62 Microbial subversion of host cells 

63 Microbe-vector interactions in vector-borne diseases 

64 Molecular pathogenesis of virus infections 



SIXTY-FIFTH SYMPOSIUM OF THE 
SOCIETY FOR GENERAL MICROBIOLOGY 
HELD AT KEELE UNIVERSITY SEPTEMBER 2005 

Edited by 

G. M. Gadd, K. T. Semple & H. M. Lappin-Scott 



micro-organisms 
and earth systems 

advances in 
geomicrobiology 



Published for the Society for General Microbiology 



w 







M.J 



CAMBRIDGE 

UNIVERSITY PRESS 



CAMBRIDGE UNIVERSITY PRESS 

Cambridge, New York, Melbourne, Madrid, Cape Town, Singapore, Sao Paulo 

Cambridge University Press 

The Edinburgh Building, Cambridge CB2 2RU, UK 

Published in the United States of America by Cambridge University Press, New York 

www.camb ridge, org 

Information on this title: www.cambridge.org/9780521862226 

© Society for General Microbiology 2005 



This publication is in copyright. Subject to statutory exception and to the provision of 
relevant collective licensing agreements, no reproduction of any part may take place 
without the written permission of Cambridge University Press. 

First published in print format 2005 

ISBN-13 978-0-511-14130-0 eBook (NetLibrary) 
isbn-io 0-511-14130-0 eBook (NetLibrary) 

ISBN-13 978-0-521-86222-6 hardback 
isbn-io 0-521-86222-1 hardback 



Cambridge University Press has no responsibility for the persistence or accuracy of URLs 
for external or third-party internet websites referred to in this publication, and does not 
guarantee that any content on such websites is, or will remain, accurate or appropriate. 



CONTENTS 



Contributors vii 

Editors' Preface xi 

M . Wagner and M . W. Taylor 

Isotopic-labelling methods for deciphering the function of uncultured 

micro-organisms 1 

L. A. Warren 

Biofilms and metal geochemistry: the relevance of micro-organism-induced 

geochemical transformations 1 1 

N. D. Gray and I. M. Head 

Minerals, mats, pearls and veils: themes and variations in giant sulfur bacteria 35 

D. W. Hopkins, B. Elberling, L. G. Greenfield, E. G. Gregorich, P. Novis, 
A. G. O'Donnell and A. D. Sparrow 

Soil micro-organisms in Antarctic dry valleys: resource supply and utilization 71 

V. R. Phoenix, A. A. Korenevsky, V. R. F. Matias and T. J. Beveridge 

New insights into bacterial cell-wall structure and physico-chemistry: implications 

for interactions with metal ions and minerals 85 

J. Coombs and T. Barkay 

Horizontal gene transfer of metal homeostasis genes and its role in microbial 
communities of the deep terrestrial subsurface 1 09 

L. G. Benning, V. R. Phoenix and B. W. Mountain 

Biosilicification: the role of cyanobacteria in silica sinter deposition 131 

K. H. Nealson and R. Popa 

Metabolic diversity in the microbial world: relevance to exobiology 1 51 

D. B. Nedwell 

Biogeochemical cycling in polar, temperate and tropical coastal zones: 

similarities and differences 173 

G. M. Gadd, M. Fomina and E. P. Burford 

Fungal roles and function in rock, mineral and soil transformations 201 

K. Pedersen 

The deep intraterrestrial biosphere 233 



SGM symposium 65 



vi Contents 



J. A. Raven, K. Brown, M. Mackay, J. Beardall, M. Giordano, E. Granum, 
R. C. Leegood, K. Kilminster and D. I. Walker 

Iron, nitrogen, phosphorus and zinc cycling and consequences for primary 

productivity in the oceans 247 

J. R. Lloyd 

Mechanisms and environmental impact of microbial metal reduction 273 

M. Kriiger and T. Treude 

New insights into the physiology and regulation of the anaerobic oxidation 

of methane 303 

N. Clipson, E. Landy and M. Otte 

Biogeochemical roles of fungi in marine and estuarine habitats 321 

P. C. Bennett and A. S. Engel 

Role of micro-organisms in karstification 345 

Index 365 



SGM symposium 65 



CONTRIBUTORS 



Barkay, T. 

Department of Biochemistry and Microbiology, Cook College, Rutgers University, 
76 Lipman Dr., New Brunswick, NJ 08901, USA 

Beardall, J. 

School of Biological Sciences, Monash University, Clayton, VIC 3800, Australia 

Bennett, P. C. 

Department of Geological Sciences, The University of Texas at Austin, Austin, TX 787 1 2, 
USA 

Benning, L. G. 

Earth and Biosphere Institute, School of Earth and Environment, University of Leeds, UK 

Beveridge, T. J. 

Department of Molecular and Cellular Biology, College of Biological Science, 
University of Guelph, Guelph, Ontario, Canada N1 G 2W1 

Brown, K. 

Plant Research Unit, Division of Environmental and Applied Biology, School of Life 
Sciences, University of Dundee at SCRI, Scottish Crop Research Institute, Invergowrie, 
Dundee DD2 5DA, Scotland, UK 

Burford, E. P. 

Division of Environmental and Applied Biology, Biological Sciences Institute, School of Life 
Sciences, University of Dundee, Dundee DD1 4HN, Scotland, UK 

Clipson, N. 

Department of Industrial Microbiology, University College Dublin, Belfield, Dublin 4, Ireland 

Coombs, J. 

Department of Biochemistry and Microbiology, Cook College, Rutgers University, 
76 Lipman Dr., New Brunswick, NJ 08901, USA 

Elberling, B. 

Institute of Geography, University of Copenhagen, 0ster Voldgade 10, DK-1350, 
Copenhagen K., Denmark 

Engel, A. S. 

Department of Geology and Geophysics, Louisiana State University, Baton Rouge, 
LA 70803, USA 

Fomina, M. 

Division of Environmental and Applied Biology, Biological Sciences Institute, School of Life 
Sciences, University of Dundee, Dundee DD1 4HN, Scotland, UK 

Gadd,G. M. 

Division of Environmental and Applied Biology, Biological Sciences Institute, School of Life 
Sciences, University of Dundee, Dundee DD1 4HN, Scotland, UK 



SGM symposium 65 



viii Contributors 



Giordano, M. 

Department of Marine Science, Universita Politecnica delle Marche, 60131 Ancona, 
Italy 

Granum, E. 

Department of Animal and Plant Sciences, University of Sheffield, Sheffield S1 2TN, UK 

Gray, N. D. 

School of Civil Engineering and Geosciences, Institute for Research on the Environment 
and Sustainability and Centre for Molecular Ecology, University of Newcastle, Newcastle 
uponTyneNEI 7RU, UK 

Greenfield, L. G. 

School of Biological Sciences, University of Canterbury, Private Bag 4800, Christchurch, 
New Zealand 

Gregorich, E. G. 

Agriculture Canada, Central Experimental Farm, Ottawa, Canada K1 A 0C6 

Head, I. M. 

School of Civil Engineering and Geosciences, Institute for Research on the Environment 
and Sustainability and Centre for Molecular Ecology, University of Newcastle, Newcastle 
uponTyneNEI 7RU, UK 

Hopkins, D.W. 

School of Biological and Environmental Sciences, University of Stirling, Stirling FK9 4LA, 
Scotland, UK 

Kilminster, K. 

School of Plant Biology, University of Western Australia, M090 35 Stirling Highway, Crawley, 
WA 6009, Australia 

Korenevsky, A. A. 

Department of Molecular and Cellular Biology, College of Biological Science, 
University of Guelph, Guelph, Ontario, Canada N1G 2W1 

Kruger, M. 

Federal Institute for Geosciences and Resources (BGR), Stilleweg 2, D-30655 Hannover, 
Germany, and Max-Planck-lnstitute for Marine Microbiology, Celsiusstrasse 1, D-28359 
Bremen, Germany 

Landy, E. 

School of Biomedical and Molecular Sciences, University of Surrey, Guildford GU2 7XH, 
UK 

Leegood, R. C. 

Department of Animal and Plant Sciences, University of Sheffield, Sheffield S1 2TN, UK 

Lloyd, J. R. 

The Williamson Research Centre for Molecular Environmental Studies and the School 
of Earth, Atmospheric and Environmental Sciences, University of Manchester, Manchester 
M13 9PLUK 



SGM symposium 65 



Contributors ix 



Mackay, M. 

Plant Research Unit, Division of Environmental and Applied Biology, School of Life 
Sciences, University of Dundee at SCRI, Scottish Crop Research Institute, Invergowrie, 
Dundee DD2 5DA, Scotland, UK 

Matias,V. R. F. 

Department of Molecular and Cellular Biology, College of Biological Science, 
University of Guelph, Guelph, Ontario, Canada N1 G 2W1 

Mountain, B. W. 

Institute of Geological and Nuclear Sciences, Wairakei Research Centre, Taupo, New Zealand 

Nealson, K.H. 

Department of Earth Sciences, University of Southern California, Los Angeles, 
CA 90089-0740, USA 

Nedwell, D. B. 

Department of Biological Sciences, University of Essex, Colchester C04 3SQ, UK 

Novis, P. 

Manaaki Whenua - Landcare Research, PO Box 69, Lincoln 81 52, New Zealand 

O'Donnell, A. G. 

Institute for Research on Environment and Sustainability, University of Newcastle upon Tyne, 
Newcastle upon Tyne NE1 7RU, UK 

Otte, M. 

Department of Botany, University College Dublin, Belfield, Dublin 4, Ireland 

Pedersen, K. 

Deep Biosphere Laboratory, Department of Cell & Molecular Biology, Goteborg University, 
Box 462, SE-405 30 Goteborg, Sweden 

Phoenix, V. R. 

Department of Molecular and Cellular Biology, College of Biological Science, 
University of Guelph, Guelph, Ontario, Canada N1 G 2W1 

Popa, R. 

Department of Earth Sciences, University of Southern California, Los Angeles, 
CA 90089-0740, USA 

Raven, J. A. 

Plant Research Unit, Division of Environmental and Applied Biology, School of Life 
Sciences, University of Dundee at SCRI, Scottish Crop Research Institute, Invergowrie, 
Dundee DD2 5DA, Scotland, UK 

Sparrow, A. D. 

School of Biological Sciences, University of Canterbury, Private Bag 4800, Christchurch, 
New Zealand, and Department of Natural Resources and Environmental Sciences, University 
of Nevada, 1000 Valley Rd, Reno, NV 89512, USA 



SGM symposium 65 



Contributors 



Taylor, M.W. 

Department of Microbial Ecology, University of Vienna, Althanstr. 14, A-1 090 Vienna, 
Austria 

Treude, T. 

Max-Planck-Institute for Marine Microbiology, Celsiusstrasse 1, D-28359 Bremen, Germany 

Wagner, M. 

Department of Microbial Ecology, University of Vienna, Althanstr. 14, A-1 090 Vienna, 

Austria 

Walker, D. I. 

School of Plant Biology, University of Western Australia, M090 35 Stirling Highway, Crawley, 
WA 6009, Australia 

Warren, LA. 

School of Geography and Earth Sciences, McMaster University, 1280 Main St. West, 
Hamilton, Ontario, Canada L8S4K1 



SGM symposium 65 



EDITORS' PREFACE 



The science of the environment encompasses a huge number of biological, chemical 
and physical disciplines. For several years, scientists have been interested in large-scale 
environmental processes/phenomena, such as soil formation, global warming and 
global elemental cycling. Until recently, the role and impact of micro-organisms on 
these 'global' environmental processes has been largely ignored or, at best, underesti- 
mated. However, there is growing awareness that important environmental trans- 
formations are catalysed, mediated and influenced by micro-organisms, and such 
knowledge is having an increasing influence on disciplines other than microbiology, 
such as geology and mineralogy. Geomicrobiology can be defined as the study of the 
role that microbes have played and are playing in processes of fundamental importance 
to geology As such, it is a truly interdisciplinary subject area, necessitating input from 
physical, chemical and biological sciences, in particular combining the fields of 
environmental and molecular microbiology together with significant areas of 
mineralogy, geochemistry and hydrology. As a result, geomicrobiology is probably the 
most rapidly growing area of microbiology at present. It is timely that this topic should 
be the subject of a Plenary Symposium volume of the Society for General Microbiology 
(SGM) to emphasize and define this important area of microbiological interest, and 
help to promote exciting collaborations between microbiologists and other environ- 
mental and Earth scientists. 

This Symposium arose from the Environmental Microbiology Group of the SGM and 
presents a snapshot of some key areas of geomicrobiology written by experts in their 
respective fields. The book focuses on some important microbial functions in aquatic 
and terrestrial environments and their influence on 'global' processes and includes 
state-of-the-art approaches to visualization, culture and identification, community 
interactions and gene transfer, as well as diversity studies in relation to key pathways. 
Novel approaches for the study of diversity and function of microbial communities are 
highlighted and applied to key environmental problems, such as community 
interactions in biofilms, and the microbiology of surface, sub-surface and extreme 
environments. Microbial involvement in key global biogeochemical cycles is exem- 
plified by aquatic and terrestrial examples, and includes metal and mineral trans- 
formations and development, element cycling in marine and estuarine systems, primary 
production, and anaerobic methane oxidation. All major groups of geochemically 
active microbes are represented, including cyano bacteria, bacteria, archaea, microalgae 
and fungi, in a wide range of habitats both aquatic and terrestrial, aerobic and 
anaerobic, and benign and extreme, reflecting the wealth of diversity in both the 
natural and the microbial world. It should also be appreciated that many of the natural 
processes discussed also have application in applied contexts such as agriculture and 

SGM symposium 65 



xii Editors' preface 

plant productivity, environmental exploitation and resource utilization and microbial 
treatment of pollution. This book will truly represent environmental microbiology in 
its broadest sense and we hope that it will have broad appeal, not only to environmental 
microbiologists, but also to environmental scientists, geologists, geochemists, Earth 
scientists, ecologists and environmental biotechnologists. A beneficial outcome may 
be the promotion of exciting collaborations between microbiologists and those in 
complementary physical and chemical disciplines. Only through interdisciplinary 
scientific approaches will the roles of micro-organisms in Earth systems be better 
clarified, appreciated and understood. 

We would like to thank all the authors for enthusiastically supporting this project 
despite their obvious heavy work commitments. We also wish to thank Diane Purves 
(University of Dundee) for expert assistance with manuscripts and author liaison, and 
all at the SGM office, particularly Melanie Scourfield, Josiane Dunn and Janet Hurst, 
for their efficient help and support with the Symposium organization and production of 
this volume. 

Geoffrey M. Gadd 

Kirk T. Semple 

Hilary M. Lappin-Scott 



SGM symposium 65 



Isotopic-labelling methods for 
deciphering the function of 
uncultured micro-organisms 

Michael Wagner and Michael W. Taylor 

Department of Microbial Ecology, University of Vienna, Althanstr. 1 4, A-1 090 Vienna, Austria 



INTRODUCTION 

With the benefit of hindsight, the last 20 years in microbial ecology will probably be 
referred to as the census period that dramatically changed our perception of bio- 
diversity within the three domains of life. Bacteria and archaea are no longer viewed as 
groups of peculiar and morphologically simple organisms that show relatively little 
diversification despite their long evolutionary history, but have now been recognized 
to harbour a perplexing number of novel phylogenetic lineages (Rappe & Giovannoni, 
2003). Current estimates assume that the number of prokaryotic species ranges in the 
millions and thus vastly exceeds the fewer than 10000 described prokaryotic species 
that have been isolated to date in pure culture (Curtis et ai, 2002). This dramatic para- 
digm shift was only made possible by the development of cultivation-independent 
molecular approaches for surveying microbial diversity in nature. Whilst it is now 
evident that most prokaryotes cannot be cultured easily, due to their living in complex 
communities and their intimate metabolic links with both their abiotic and biotic 
environments, the powerful arsenal of techniques at our disposal enables us to see 
beyond the 'cultured few' and gain valuable insights into the realm of uncultured micro- 
organisms (Wagner, 2004). It is now relatively straightforward to determine the species 
richness of natural microbial communities by comparative sequence analysis of 
environmentally retrieved 16S rRNA gene sequences (Olsen et al., 1986; Schloss & 
Handelsman, 2004). Furthermore, even the abundance of bacteria in environmental 
samples can be determined easily by using quantitative PCR (Skovhus et al., 2004) or 
hybridization formats such as fluorescence in situ hybridization (FISH) (Wagner et al., 
2003). Due to these technological advances, many novel and often numerically 

SGM symposium 65: Micro-organisms and Earth systems - advances in geomicrobiology. 
Editors G. M. Gadd, K. T. Semple & H. M. Lappin-Scott. Cambridge University Press. ISBN 521 86222 1 ©SGM 2005 



M. Wagner and M. W. Taylor 

important bacterial and archaeal species have been identified in various ecosystems. 
However, for most of these players, their ecophysiology and contribution to the func- 
tioning of the ecosystem remain hidden. 

Given the above, one of the biggest challenges in contemporary microbial ecology is 
to develop strategies that enable us (i) to directly investigate metabolic properties of 
defined but uncultured micro-organisms, and (ii) to identify those uncultured organ- 
isms that are responsible for defined processes within their natural environment. These 
two criteria highlight important differences in approaches that are currently taken in 
the study of microbial community function. In a microbial world dominated by un- 
cultured representatives, those investigators who approach matters from a diversity 
standpoint typically target organisms of interest (e.g. those that are highly abundant in 
a particular ecosystem) and endeavour to find out what metabolic traits these species 
possess. An alternative, equally valid approach is to recognize a process occurring in the 
environment (e.g. sulfate reduction or denitrification) and try to establish which 
organisms are responsible. Both approaches are served equally well by the methods 
outlined in this chapter. Specifically, we provide an overview of a group of recently 
developed methods that exploit the addition of isotope-labelled substrates to complex 
microbial communities in order to bridge the gap between microbial community 
structure and function. 

IDENTIFICATION OF PLAYERS THAT INCORPORATE LABELLED 
SUBSTRATES 

One of the most elegant ways to better understand the ecophysiology of uncultured 
bacteria is to expose them to isotope-labelled substrates and, subsequently, to link their 
identification with a measurement of substrate incorporation into their biomass. This 
concept was first realized by extracting phospholipid fatty acids from complex 
microbial communities after incubation with 13 C-labelled substrates, and identifying 
active groups of micro-organisms by detection of labelled signature compounds using 
isotope-ratio mass spectrometry (Boschker et aL, 1998). However, this approach is not 
suited for the identification of uncultured bacteria, because their phospholipid fatty- 
acid composition is unknown. Consequently, several methods, which are presented in 
more detail below, were developed to exploit other, more informative biomarkers or 
even to allow simultaneous organism identification and detection of substrate 
incorporation at the single-cell level. A common feature of all these methods is that it is 
possible to fine-tune the experimental setup such that very specific questions regarding 
the ecophysiology of selected micro-organisms can be answered. This is achieved by 
performing parallel experiments using modified incubation conditions (pH, tempera- 
ture, presence/absence of different electron acceptors, presence/absence of specific 
inhibitors for selected microbial groups etc.) under which the microbial biomass 

SGM symposium 65 



Function of uncultured micro-organisms 

encounters the labelled substrate. Alternatively, this modus operandi enables microbial 
ecologists to hunt for novel micro-organisms responsible for defined functions in the 
environment. However, with the exception of the so-called HetC0 2 -microautoradio- 
graphy (MAR) approach (Hesselsoe etaL, 2005), all of these methods are dependent on 
the availability of commercial isotope labelling for the compounds of interest. Further- 
more, all mentioned techniques fail to detect micro-organisms that convert labelled 
compounds without using them for the synthesis of biomass or storage compounds. 

DNA stable-isotope probing (DNA-SIP) is based on the incorporation of stable iso- 
tope-labelled substrates into the DNA of substrate-consuming micro-organisms 
(reviewed recently by Dumont 6c Murrell, 2005). To date, DNA-SIP has only been 
performed with 13 C-labelled substrates. In DNA-SIP, the heavy DNA of the substrate 
consumers is separated from the light DNA of the other community members by 
buoyant density-gradient centrifugation with added ethidium bromide. The identity of 
the active bacteria is revealed subsequently by amplification, cloning and comparative 
sequence analysis of 16S rRNA genes or selected functional genes from the heavy DNA. 
Probing the light- and heavy-DNA fractions with specific PCR primers can be applied 
to show rapidly whether a given bacterial player has incorporated the offered substrate. 
One of the major advantages of DNA-SIP is that large genomic-DNA fragments of the 
active community fraction can be isolated and subsequently analysed by the environ- 
mental-genomics approach (DeLong, 2002, 2005). On the other hand, the DNA of 
active community members will only be labelled sufficiently for DNA-SIP if the added 
compounds are highly enriched in 13 C, and if relatively high concentrations of labelled 
substrate are offered over extended time periods (up to several weeks) so that slow- 
growing micro-organisms are also able to replicate their DNA and to divide. It is 
essential to be aware that such incubation conditions might induce major shifts in the 
composition of the resident microbial community, perhaps comparable to the biases 
reported previously for enrichment cultures (Wagner et ai, 1993). Compared to the 
rapid-growing r-strategists, K-strategists that are adapted to low substrate concen- 
trations might thus be heavily underrepresented or even missing (if the applied 
substrate concentration is inhibitory to them) in the heavy-DNA fraction. Further- 
more, due to the long incubation times required for DNA-SIP, primary consumers will 
transfer label to other community members by cross-feeding. Therefore, not all 
genomes found in the heavy DNA are from bacteria capable of using the labelled 
compound (although, on a more positive note, this effect could also be exploited to 
obtain indications of metabolic interactions among different community members). 

Several of the limitations of DNA-SIP can be overcome by RNA-SIP (Manefield et al., 
2002, 2005). In microbial cells, synthesis rates of RNA (including 16S rRNA) are 
significantly higher than rates of DNA replication; consequently, if DNA is replaced by 

SGM symposium 65 



M. Wagner and M. W. Taylor 



• • • • m . 
mm*9 f • \m . 



JO • #'S • •<••*•• 

• 



j 



•■ 




« 



(a) 



i • 



^ 



f5 



Cj 



i 



#* 



4| ft * * 


*"# 


* + *#* 


1 * #*-■ 

• 


- 






.-• 


* -H 


' 






(b) # 


- 


* 



Fig. 1. Isotope-array experiment to identify butyrate-oxidizing bacteria under aerobic conditions in 
activated sludge from a nutrient-removal wastewater-treatment plant, (a) Fluorescence scan of the 
array; (b) radioactivity scan of the same array. Pictures kindly provided by A. Loy, M. Schloter and M. 
Hesselsoe. 

RNA as a biomarker, much shorter incubation times are required to obtain sufficient 
amounts of labelled material. 13 C-labelled RNA can be separated from light RNA 
via caesium trifluoroacetate (CsTFA) density-gradient centrifugation, although it has 
been observed that labelled RNA with a specific buoyant density is found over a number 
of fractions in CsTFA gradients. To solve this problem, 16S rRNA from each density 
fraction must be amplified separately by RT-PCR and subsequently analysed by a 
fingerprinting method such as denaturing-gradient gel electrophoresis (DGGE). 
Increasing band intensities in the fingerprint pattern of certain 16S rDNA fragments 
from high-density fractions throughout the duration of the labelled-substrate pulse are 
then used to demonstrate that the 16S rRNA of a particular bacterium has become 
increasingly labelled during the duration of the experiment. Subsequently, the 
respective DGGE band is cut from the gel, reamplified by PCR and sequenced in order 
to identify the substrate-consuming micro-organisms. The disadvantage of this proce- 
dure is that it is very time-consuming and dependent on the resolving power of DGGE, 
which is insufficient in highly complex microbial communities. In future studies, the 



SGM symposium 65 



Function of uncultured micro-organisms 

identification of bacteria with a specific function could be accelerated and improved in 
resolution by hybridizing heavy DNA or RNA obtained by SIP to diagnostic rRNA- 
targeted microarrays (Guschin et al., 1997; Loy et al., 2002, 2005; Wilson et al., 2002). 

RNA as a biomarker is also exploited by the so-called isotope-array approach 
(Adamczyk et al., 2003) but, in contrast to SIP, the microbial communities are incubated 
with radioactively (and not lj C) -labelled compounds prior to nucleic-acid extraction. 
To date, exclusively 14 C-labelled compounds have been used for isotope-array analysis. 
Again, as RNA is targeted, only relatively short incubation times are required. After 
community RNA extraction, the RNA is labelled covalently with a fluorescent dye and 
hybridized to a suitable rRNA-targeted microarray by using standard procedures. 
Visualization of the microarray by using a fluorescence scanner reveals the community 
structure, whilst recording the radioactivity of each microarray spot by use of a beta- 
imager shows which community member has incorporated radioactive label into its 
rRNA and thus consumed the offered substrate (Fig. 1). Due to limited resolution of 
commercially available beta-imagers, larger probe-spot sizes are required for optimal 
results compared with standard microarrays. An advantage of the isotope array 
compared with RNA-SIP is that it is PCR-independent and thus offers the potential for 
quantitative analysis. The ratio of radioactive signal to fluorescence signal, which can 
be determined easily for each hybridized probe spot, reflects the incorporation rate of 
the labelled substrate into the RNA of the detected bacterial population (s) and can thus 
be used to compare their metabolic activities within a complex microbial community. 
On the other hand, and in contrast to SIP, the nature of this format makes it impossible 
to identify novel substrate-consuming micro-organisms whose 16S rRNA sequences 
are not yet known (i.e. you can only find what you are looking for) . In the future, the iso- 
tope array should be adapted to the use of stable isotopes so as to avoid radioactive lab 
work and also render it adaptable to the analysis of microbial communities colonizing 
humans. For this purpose, beta-imaging must be replaced by detection methods such as 
time-of-flight secondary-ion mass spectrometry. 

In contrast to the aforementioned methods, structure and function of microbial 
communities can be analysed by a combination of FISH and MAR at the single-cell 
level (Lee et al., 1999; Ouverney & Fuhrman, 1999; Daims et al., 2001; Nielsen et al., 
2002, 2003b) (Fig. 2), as long as the environmental sample is suited for FISH analysis. 
After incubation of the microbial communities with radioactively labelled substrate, 
community members are identified by FISH with rRNA-targeted oligonucleotide 
probes. In parallel, the incorporation of radioactive substrate into the biomass of 
bacterial cells is visualized with a radiation-sensitive silver halide emulsion, which is 
placed on top of the radiolabeled organisms and subsequently processed by standard 
photographic procedures. Excited silver ions will precipitate as metallic silver and 

SGM symposium 65 



6 



M. Wagner and M. W. Taylor 



2-10 pm 

cryo- 
section 




Cover slip 



Silver-grain formation above cells that 
took up radioactively labelled substrate 



CLSM 



T 






Bacteria labelled by FISH 



Fig. 2. Schematic representation of the FISH-MAR approach. CLSM, Confocal laser-scanning 
microscope. 



Methanol added 




Control 




Fig. 3. Application of FISH-MAR to investigate the ecophysiology of uncultured filamentous micro- 
organisms. The filaments showed uptake of methanol in MAR experiments (a) and were identified 
simultaneously by FISH as members of the ' Gammaproteobacteria' (b). Control experiments with 
pasteurized biomass showed no unspecific uptake or binding of labelled methanol (c, d). Pictures 
kindly provided by K. Stoecker and H. Daims. 



SGM symposium 65 



Function of uncultured micro-organisms 

appear as black grains after development of the film (Fig. 3). As FISH requires perm- 
eabilization of the bacterial cell envelope, radioactive substrates that are taken up by a 
bacterial cell, but not incorporated into macromolecules, will not result in silver-grain 
formation. It should, however, be noted that MAR cannot differentiate between con- 
version of labelled organic substrates into intracellular storage products and active 
substrate metabolism. The recently developed HetC0 2 -MAR approach (Hesselsoe 
et al., 2005) allows this limitation to be overcome by utilizing the tendency of meta- 
bolically active heterotrophic bacteria to assimilate significant amounts of C0 2 . Thus, 
14 C0 2 is added as an activity marker to complex microbial communities and, after 
suitable incubation periods, active autotrophic and heterotrophic community members 
can be identified by FISH-MAR. HetCO^-MAR also avoids difficulties associated with 
obtaining many isotope-labelled substrates, as this approach requires that only the C0 2 
is labelled and other added substrates are not. 

Major advantages of FISH-MAR are that incubation times are comparatively short, 
because every labelled macromolecule of the monitored bacterial cells contributes to 
the detection of substrate incorporation. Furthermore, the amount of substrate incor- 
poration into specific uncultured bacterial cells can be quantified by counting the 
number of silver grains on top of the cells if bacteria with known specific radioactivity 
are used as an internal standard. This technique can be applied to measure in situ 
substrate affinities {K s ) of uncultured bacteria and to study physiological differences 
within naturally occurring single populations of bacteria (Nielsen et al., 2003a). In 
future studies, it should be possible to modify the FISH-MAR method such that radio- 
actively labelled substrates are replaced by stable isotope-labelled compounds and 
the stable-isotope composition of FISH-identified microbial cells is measured by 
secondary-ion mass spectrometry (Orphan et al., 2001, 2002). 

The last few years have seen the development of an impressive battery of tools that 
measure function of uncultured micro-organisms by the addition of labelled substrates. 
After initial, proof-of-principle type studies, several of these tools are now applied 
(almost) routinely in microbial ecology laboratories and they will be used increasingly 
in time-course experiments to uncover metabolic networks in multi-component 
microbial communities (Lueders et al., 2004). Together with quantitative FISH-based 
analysis of co-occurring microbial populations (Daims et al., 2005), these approaches 
will shed new light on the complex network of interactions existing in microbial 
communities. 

OUTLOOK 

We have provided here a brief overview of cutting-edge methods for the analysis of 
microbial-community function. As a field that necessarily emphasizes the importance 

SGM symposium 65 



8 M. Wagner and M. W. Taylor 

of technological advances, microbial-community ecology should continue to benefit 
significantly from the development and application of isotopic-labelling methods. 
We predict that this area should prove a productive meeting ground for process- and 
diversity-focused microbial ecologists. 

The approaches outlined above should prove especially fruitful when used in combina- 
tion with another emerging subdiscipline of microbial ecology, that of environmental 
genomics. Much has been made recently of the power of environmental genomics, 
exemplified by the massive sequencing efforts undertaken in the Sargasso Sea (Venter 
et aL, 2004). However, such studies tell us more about potential, rather than actual, 
community function because, for many of the retrieved genes, annotation provides no 
indication of a possible function of the encoded protein. Even for those genes where 
annotation is meaningful, no information on actual expression in the environment is 
apparent. Environmental proteomics offer a partial solution to this quandary (Wilmes 
& Bond, 2004), but this field is in its infancy and widespread application is surely 
several years away. The real beauty of environmental genomics, then (at least from a 
community-function perspective), may be its role in the formulation of hypotheses 
(Wagner, 2005). Strong hints about organismal metabolism can be obtained via meta- 
genome sequencing, hints that can be evaluated by using the isotopic-labelling 
techniques described in this chapter. The synergy derived from mixing these approaches 
could be profound: imagine a situation where one discovers a 'new' organism of 
presumed importance in an environment (e.g. any of a dozen or more candidate phyla 
with currently no cultured representatives), but nothing is known of its metabolism. 
FISH-MAR, SIP and isotope arrays all offer potential avenues for investigation, yet 
where would one start? Possible carbon sources alone number in the hundreds and it 
is simply not feasible to test all of these with isotopic methods. However, if one has 
genomic information for such organisms (as obtained by bacterial artificial chromo- 
some or shotgun sequencing from a mixed sample), then valuable clues as to suitable 
target substrates should be forthcoming. The need for environmental genomics is 
arguably reduced when one approaches community function from the other side, i.e. 
starting with a process and identifying the key players. Here, the critical step is in 
defining incubation conditions where only the organisms capable of the function (s) of 
interest are active. 

Since its first coupling with molecular-phylogenetic techniques in the late 1990s, 
isotopic labelling has provided important insights into the function of specific, un- 
cultured micro-organisms. Continuing technological advances and further integration 
with environmental-genomics and -proteomics approaches can only enhance the value 
of this area of microbial ecology. 

SGM symposium 65 



Function of uncultured micro-organisms 



REFERENCES 



Adamczyk, J., Hesselsoe, M., Iversen, N., Horn, M., Lehner, A., Nielsen, P. H., Schloter, 
M., Roslev, P. & Wagner, M. (2003). The isotope array, a new tool that employs 
substrate-mediated labeling of rRNA for determination of microbial community 
structure and function. Appl Environ Microbiol 69, 6875-6887. 

Boschker, H. T. S., Nold, S. C, Wellsbury, P., Bos, D., de Graaf, W. # Pel, R., Parkes, R. J. & 
Cappenberg, T. E. (1998). Direct linking of microbial populations to specific 
biogeochemical processes by 13 C-labelling of biomarkers. Nature 392, 801-805. 

Curtis, T. P., Sloan, W. T. & Scannell, J. W. (2002). Estimating prokaryotic diversity and 
its limits. Proc Natl Acad Sci USA 99, 1 0494-1 0499. 

Daims, H., Nielsen, J. L, Nielsen, P. H., Schleifer, K.-H. & Wagner, M. (2001). In situ 
characterization of Nitrospira-Wke nitrite-oxidizing bacteria active in wastewater 
treatment plants. Appl Environ Microbiol 67, 5273-5284. 

Daims, H., Lucker, S. & Wagner, M. (2005). daime, a novel image analysis program for 
microbial ecology and biofilm research. Environ Microbiol (in press). 

DeLong, E. F. (2002). Microbial population genomics and ecology. Curr Opin Microbiol 
5, 520-524. 

DeLong, E. F. (2005). Microbial community genomics in the ocean. Nat Rev Microbiol 
3, 459-469. 

Dumont, M. G. & Murrell, J. C. (2005). Stable isotope probing - linking microbial identity 
to function. Nat Rev Microbiol 3, 499-504. 

Guschin, D. Y, Mobarry, B. K., Proudnikov, D., Stahl, D. A., Rittmann, B. E. & Mirza- 
bekov, A. D. (1997). Oligonucleotide microchips as genosensors for determinative 
and environmental studies in microbiology. Appl Environ Microbiol 63, 2397-2402. 

Hesselsoe, M., Nielsen, J. L, Roslev, P. & Nielsen, P. H. (2005). Isotope labeling and 
microautoradiography of active heterotrophic bacteria on the basis of assimilation of 
1 4 C 2 . Appl Environ Microbiol 7 1 , 646-6 5 5 . 

Lee, N., Nielsen, P. H., Andreasen, K. H., Juretschko, S., Nielsen, J. L, Schleifer, K.-H. & 
Wagner, M. (1999). Combination of fluorescent in situ hybridization and micro- 
autoradiography -a new tool for structure-function analyses in microbial ecology. 
Appl Environ Microbiol 65, 1289-1297. 

Loy, A., Lehner, A., Lee, N., Adamczyk, J., Meier, H., Ernst, J., Schleifer, K.-H. & 
Wagner, M. (2002). Oligonucleotide microarray for 16S rRNA gene-based detection 
of all recognized lineages of sulfate-reducing prokaryotes in the environment. Appl 
Environ Microbiol 68, 5064-5081 . 

Loy, A., Schulz, C, Lucker, S., Schopfer-Wendels, A., Stoecker, K., Baranyi, C, Lehner, 
A. & Wagner, M. (2005). 16S rRNA gene-based oligonucleotide microarray for 
environmental monitoring of the betaproteobacterial order " Rhodocydales" . Appl 
Environ Microbiol 71,1 373-1 386. 

Lueders, T., Wagner, B., Claus, P. & Friedrich, M. W. (2004). Stable isotope probing of 
rRNA and DNA reveals a dynamic methylotroph community and trophic interactions 
with fungi and protozoa in oxic rice field soil. Environ Microbiol 6, 60-72. 

Manefield, M., Whiteley, A. S., Griffiths, R. I. & Bailey, M. J. (2002). RNA stable isotope 
probing, a novel means of linking microbial community function to phylogeny. Appl 
Environ Microbiol 68, 5367-5373. 

Manefield, M., Griffiths, R. I., Leigh, M. B., Fisher, R. & Whiteley, A. S. (2005). 
Functional and compositional comparison of two activated sludge communities 
remediating coking effluent. Environ Microbiol 7, 71 5-722. 



SGM symposium 65 



10 M. Wagner and M. W. Taylor 



Nielsen, J. L, Juretschko, S., Wagner, M. & Nielsen, P. H. (2002). Abundance and 

phylogenetic affiliation of iron reducers in activated sludge as assessed by fluores- 
cence in situ hybridization and microautoradiography. Appl Environ Microbiol 68, 

4629-4636. 
Nielsen, J. L, Wagner, M. & Nielsen, P. H. (2003a). Use of microautoradiography to study 

in situ physiology of bacteria in biofilms. Rev Environ Sci Biotechnol 2, 261-268. 
Nielsen, J. L, Christensen, D., Kloppenborg, M. & Nielsen, P. H. (2003b). Quantification 

of cell-specific substrate uptake by probe-defined bacteria under in situ conditions by 

microautoradiography and fluorescence in situ hybridization. Environ Microbiol 5, 

202-211. 
Olsen, G. J., Lane, D. J., Giovannoni, S. J., Pace, N. R. & Stahl, D. A. (1986). Microbial 

ecology and evolution: a ribosomal RNA approach. Annu Rev Microbiol 40, 337- 

365. 
Orphan, V. J., House, C. H., Hinrichs, K.-U., McKeegan, K. D. & DeLong, E. F. (2001). 

Methane-consuming archaea revealed by directly coupled isotopic and phylogenetic 

analysis. Science 293, 484^487. 
Orphan, V. J., House, C. H., Hinrichs, K. U., McKeegan, K. D. & DeLong, E. F. (2002). 

Multiple archaeal groups mediate methane oxidation in anoxic cold seep sediments. 

Proc Natl Acad Sci U 5 A 99, 7663-7668. 
Ouverney, C. C. & Fuhrman, J. A. (1999). Combined microautoradiography's rRNA 

probe technique for the determination of radioisotope uptake by specific microbial 

cell types in situ. Appl Environ Microbiol 65, 1 746-1 752. 
Rappe, M. S. & Giovannoni, S. J. (2003). The uncultured microbial majority. Annu Rev 

Microbiol 57, 369-394. 
Schloss, P. D. & Handelsman, J. (2004). Status of the microbial census. Microbiol Mol Biol 

Rev 68, 686-691. 
Skovhus, T. L, Ramsing, N. B., Holmstrom, C, Kjelleberg, S. & Dahllof, I. (2004). Real- 
time quantitative PCR for assessment of abundance of Pseudoalteromonas species in 

marine samples. Appl Environ Microbiol 70, 2373-2382. 
Venter, J. C, Remington, K., Heidelberg, J. F. & 20 other authors (2004). Environmental 

genome shotgun sequencing of the Sargasso Sea. Science 304, 66-74. 
Wagner, M. (2004). Deciphering the function of uncultured microorganisms. ASM News 

70, 63-70. 
Wagner, M. (2005). The community level: physiology and interactions of prokaryotes in the 

wilderness. Environ Microbiol 7, 483-485. 
Wagner, M., Amann, R., Lemmer, H. & Schleifer, K.-H. (1993). Probing activated sludge 

with oligonucleotides specific for proteobacteria: inadequacy of culture-dependent 

methods for describing microbial community structure. Appl Environ Microbiol 59, 

1520-1525. 
Wagner, M., Horn, M. & Daims, H. (2003). Fluorescence in situ hybridisation for the 

identification and characterisation of prokaryotes. CurrOpin Microbiol 6, 302-309. 
Wilmes, P. & Bond, P. L. (2004). The application of two-dimensional polyacrylamide gel 

electrophoresis and downstream analyses to a mixed community of prokaryotic 

microorganisms. Environ Microbiol 6, 91 1-920. 
Wilson, K. H., Wilson, W. J., Radosevich, J. L., DeSantis, T. Z., Viswanathan, V. S., 

Kuczmarski, T. A. & Andersen, G. L. (2002). High-density microarray of small- 

subunit ribosomal DNA probes. Appl Environ Microbiol 68, 2535-2541 . 



SGM symposium 65 



Biof ilms and metal geochemistry; 
the relevance of micro-organism- 
induced geochemical 
transformations 



Lesley A. Warren 



School of Geography and Earth Sciences, McMaster University, 1 280 Main St. West, Hamilton, 
Ontario, Canada L8S4K1 



INTRODUCTION 

This chapter is intended to provide a brief overview of the key concepts underlying the 
emerging area of environmental microbial metal geochemistry, rather than an exhaus- 
tive synthesis. The reader is referred to the following more comprehensive reviews 
on the biogeochemistry of metals (Warren & Haack, 2001), metal-mineral reactions 
(Brown & Parks, 2001), emerging molecular-level geochemical techniques (O'Day, 
1999; Brown & Sturchio, 2002) and a recent synthesis of how genetic expression in the 
environment can underpin geochemical reactions (Croal et al., 2004). The relevance of 
micro-organisms to metal behaviour arises from the overlap of the biosphere with the 
geosphere and the transformations that occur because of their interactions. Micro- 
organisms have evolved in intimate association with the rocks, soils and waters (i.e. 
geosphere) in which they find themselves. In order to grow and survive, they have adap- 
ted to these environments and use the inorganic components to drive their metabolic 
machinery; the myriad functional pathways by which they do so ensure that they 
influence a number of key elemental cycles in the process. As a consequence, many 
important geochemical processes are ultimately shaped by life, rather than strict 
geochemical equilibria, a fact that is increasingly recognized as strict geochemical 
principles fail to constrain observed environmental behaviour. 

Trace-metal behaviour in the environment is of increasing global concern as water and 
soil contamination with these toxic substances continues and the detrimental effects on 
ecosystems and human health emerge. While investigation into metal behaviour has 
spanned many disciplines reflecting discipline-specific foci and approaches to the topic 

SGM symposium 65: Micro-organisms and Earth systems - advances in geomicrobiology. 
Editors G. M. Gadd, K. T. Semple & H. M. Lappin-Scott. Cambridge University Press. ISBN 521 86222 1 ©SGM 2005 



12 L.A.Warren 



(e.g. biology, chemistry, geochemistry, toxicity, physics), what coalesces from this broad, 
substantial literature is that the controls on aqueous metal behaviour supersede 
individual discipline boundaries; rather, it is often dynamic, complex, biological (prin- 
cipally micro-organism driven) and geochemical linkages that drive geochemistry 
(Reysenbach & Shock, 2002; Newman & Banfield, 2002; Warren & Kauffman, 2003; 
Warren, 2004). 

Probing microbial-mineral-metal interactions has provided substantive evidence of 
microbial shaping of metal fate (e.g. Ehrlich, 2002; Holden & Adams, 2003 ; Islam et al., 
2004), and has clearly alluded to the need to understand how microbial activity and 
geochemistry interact to ultimately determine metal impact. It is increasingly evident 
that microbial activity can substantially impact aqueous geochemical behaviour in 
ways not predicted by classic geochemical models, a serious issue for contaminated 
environments where bacteria flourish such as acid mine drainage (AMD) or contamin- 
ated subsurface environments where migration of contaminants to groundwater 
supplies represents a serious health hazard. 

AQUEOUS METAL GEOCHEMISTRY: FUNDAMENTALS 

At the most fundamental level, the single most important predictor of metal behaviour, 
i.e. mobility within the geosphere, bioavailability, bioaccumulation and toxicity, is the 
solution or dissolved (operationally defined typically as less than 0-45 Jim or, more 
rigorously, <0-l ^m) concentration of the element (e.g. Martell et al., 1988; Campbell 
& Tessier, 1989; Hare, 1992; Campbell, 1995; Unz & Shuttleworth, 1996; Warren et al., 
1998; Hassler et al., 2004). Typically, the greater the solution concentration of a metal, 
the more likely are the negative impacts associated with that contaminant, reflecting 
greater mobility, bioavailability and toxicity. Thus, trace-element geochemistry has 
focused on understanding the processes by which metals partition between the solution 
and solid compartments within environmental systems and the controls on the pro- 
cesses involved (Warren & Haack, 2001; and references therein). While precipitation 
of metals within metal-bearing minerals such as sulfides, carbonates or hydroxides 
can play a role, principally in sedimentary systems where concentrations of metals can 
reach saturation with respect to a given mineral phase, it is now well established that 
metal solid-solution partitioning is often dominantly controlled by interfacial reac- 
tions occurring between charged functional groups at solid surfaces and solution metal 
species, collectively referred to as sorption reactions (Jenne, 1968; Dzombak & Morel, 
1990; Stumm & Morgan, 1996; Brown & Parks, 2001; Brown & Sturchio, 2002). In the 
majority of cases, most metal-solid interactions can be considered dynamically 
reversible, such that changes in certain geochemical conditions can lead to release of 
metals back into solution. A further, defining caveat to the framework of metal-solid 
interactions is that neither metal (s) nor solid phases can be considered collectively, 

SGM symposium 65 



Biofilms and metal geochemistry 13 

as the behaviour observed will be predicated upon which specific metal(s) and solid 
phase(s) are involved, reflecting relative affinities, differing mechanisms of seques- 
tration and controls on reactivity 

Aqueous geochemistry has served as the foundation discipline for much of the environ- 
mental metal literature spanning freshwater and marine metal behaviour, subsurface 
metal migration and the bioavailability and toxicity of metals to biota (e.g. Lee et al., 
2004; Martino et al., 2004; Basta et al., 2005). The defining controls of pH, redox status 
and ionic strength (Allard et al., 1987; Johnson, 1990; Brown & Parks, 2001; Hassler 
et al., 2004) on metal partitioning reflect their influence on: (i) formation/dissolution 
of key solid phases for sequestering trace metals (e.g. carbonates, sulfides and oxy- 
hydroxide minerals, organic matter), as well as metal speciation and associated 
behaviour; and (ii) sorption reactions. 

The nature and extent of a mineral's importance for metal sequestration in any given 
system will reflect its respective relative abundance and spatial distribution in the 
geosphere. The heterogeneous distribution of key mineral phases reflects the differing 
conditions and controls on their formation and dissolution and also provides some 
delineation of which minerals are likely to occur and play a role in metal behaviour 
across differing geochemical environments. Minerals can sequester metals through 
sorption at their surfaces as well as through direct precipitation or solid-solution 
formation within a mineral (Stumm & Morgan, 1996; Brown & Parks, 2001). 
The mechanism by which a metal becomes associated with a solid will determine its 
relative sensitivity to remobilization and the conditions under which release may occur. 
Sorption reactions can be considered reversible: metal uptake to solids is not a perm- 
anent phenomenon, as small fluctuations in either pH or solution metal(s) concen- 
tration can often lead to release of metals from a solid surface (Warren & Haack, 2001). 
Of key importance to note is that both the relative affinities of different solid phases for 
different metal ions, as well as the controls on their reactivity, differ. Thus, predicting 
metal solid-solution behaviour requires information on the types and abundances of 
solid phases present, the specific metal(s) and respective concentrations involved, as 
well as system geochemical conditions. 

Important solid phases 

In general, there are four major solid phases that can be considered most relevant 
for metal sequestration in aqueous systems: (i) carbonates, (ii) oxyhydroxide minerals 
(Fe and Mn); (iii) organic matter (live and dead) and (iv) sulfidic minerals (Tessier et al., 
1979). Many studies have shown that these typically comprise the majority of solid 
phases involved in metal uptake (Tessier 6c Campbell, 1988; Warren & Zimmerman, 
1994; Brown et al., 1999; Filgueiras et al., 2002). Another final sedimentary pool is often 

SGM symposium 65 



14 L.A.Warren 



referred to as the refractory or mineralized component (Tessier et al., 1979), reflecting 
metals held within crystalline mineral lattices. However, metals associated with this 
refractory pool are not likely to be released under normal geochemical fluctuations and 
thus its utility is usually for determining the absolute total mass of metal associated 
with a given sediment. Changes in pH, redox or ionic strength can affect metal uptake 
to the previous four fractions, and thus much research has focused on understanding 
metal associations with these specific sedimentary pools, which commonly occur as 
heterogeneous mixtures in environmental sediments and soils. 

Carbonate minerals [e.g. limestone, CaC0 3 (s), or dolostone, Ca,MgC0 3 (s)] are typi- 
cally found in circumneutral to alkaline pH environments, where their concentrations 
can effect significant metal uptake, especially for certain elements which show high 
affinity for carbonates, such as Cd and U. Carbonate minerals are susceptible to acid 
dissolution, and thus decreasing pH can lead to dissolution and subsequent release of 
carbonate-associated metals back into solution. 

In contrast, sulfidic minerals such as pyrite [FeS(s)] tend to be associated with anoxic 
sedimentary systems where the necessary reducing conditions are present for their 
formation and/or preservation. Sulfidic minerals are commonly formed as a con- 
sequence of sulfate reduction in anoxic sediments, which leads to a build-up of reduced 
S in the sediment pore waters, which, when sufficient concentrations build up to satu- 
rate with respect to various metal sulfide mineral phases, can then precipitate as a metal 
sulfide [e.g. CdS(s), CuS(s), FeS (s) ; Warren et al., 1998]. Once formed, sulfide minerals 
tend to be fairly robust, even in the presence of oxygen, to oxidative dissolution, except 
where direct microbial catalysis is involved. Sulfidic minerals are also a key component 
in mine-associated waste or tailings rock, driving the production of AMD. 

The Fe and Mn oxyhydroxides [e.g. Fe(III) oxides, FeOOH(s) and Mn(III,IV) oxides, 
MnOOH(s)] are highly metal-reactive and are typically widespread throughout most 
environmental systems, making them a dominant solid phase controlling metal 
behaviour in the environment. Oxyhydroxides of Fe and Mn are redox-sensitive, with 
the more oxidized form of both elements, Fe(III) and Mn(III,IV), forming a solid phase, 
while the reduced forms, Fe(II) and Mn(II), are soluble. Thus changes in redox 
conditions will profoundly affect the concentrations of these solids and associated 
metal behaviour. Although it should be noted that reductive dissolution of both 
oxyhydroxides proceeds much more quickly with microbial catalysis than without 
and that, while abiotic oxidation of Fe(II) to Fe(III) [and associated hydrolysis and 
formation of FeOOH(s) solid particles] occurs spontaneously above pH 3 in the 
presence of oxygen, abiotic Mn(II) oxidation is extremely slow below pH 9 (e.g. months 
to years; Morel & Hering, 1993; Stumm & Morgan, 1996). 

SGM symposium 65 



Biofilms and metal geochemistry 15 

Natural organic matter (NOM), both living (e.g. micro-organisms) and dead (both 
labile and refractory, e.g. humic and fulvic acids), is an efficient, but often dynamically 
reversible, metal sequestration pool (Gustafsson et aL, 2003; Smiejan et aL, 2003; 
Jacob & Otte, 2004; Boullemant et aL, 2004). Sequestration by NOM can be especially 
important in systems of high NOM concentration such as wetlands and eutrophic 
lakes, typically through sorption reactions (discussed subsequently) associated with 
their highly electrically charged surfaces. Decomposition of organic matter often leads 
to release of associated metals into solution and thus the ability of the NOM pool in 
any given system to hold onto metals will reflect the relative magnitude of decom- 
positional processes. 



Sorption reactions 

The profound importance of interfacial reactions in controlling the ultimate partition- 
ing of metals between the solid and solution phases is well recognized (Brown & Parks, 
2001; Warren & Haack, 2001; and references therein). These interfacial or sorption 
reactions involve attraction or bonding between a charged solution metal ion and a 
charged functional group on the surface of the particle (Fig. la) and are viewed to 
proceed analogously to complexation reactions involving charged dissolved species in 
solution (Buffle, 1988). Mineral surface reactivity is described by physical character- 
istics such as size (surface area to volume), crystallinity and nature and density of react- 
ive binding sites (Brown et aL, 1999; Martinez & McBride, 1998). Organic matter 
reactivity reflects both these same geochemical particle characteristics and, for live cells, 
metabolic pathway and level of activity, which can shape external and internal micro- 
geochemical environments, influence mineral formation and dissolution and alter 
surface charge characteristics and thus potential reactivity (e.g. Ferris et aL, 1989, 1999; 
Nelson et aL, 1995; Parmar et aL, 2000; Philip et aL, 2000; Templeton et aL, 2003a, b; 
Brown etaL, 2004). 

This particle surface charge is imparted by surface acid functional groups that variously 
deprotonate or protonate depending on their pK a values. Commonly, mineral surface 
functional groups are hydroxyl groups, which often display a spectrum of pK a values 
reflecting the specific hydroxyl group co-ordinational environment on the particle (e.g. 
edge or defect sites within the crystal structure). Micro-organisms also carry a surface 
charge associated with their cell wall, exopolysaccharide (EPS) or sheath, arising from 
constituent functional groups (e.g. carboxylic, phosphoryl, amine as well as hydroxyl; 
Cox et aL, 1999). Given the lower pK a values particularly associated with carboxylic 
groups (pK a ~4), micro-organisms almost always carry a net negative surface charge in 
most aqueous solutions. Microbial ability to sorb cationic metal ions makes them a 
highly relevant sorbent, especially in low-pH environments where important mineral 

SGM symposium 65 



16 L.A.Warren 



(a) 



Anionic sorption 



/"~NMC£ 






Net particle charge 

Surface functional gr oup H V" H 



Cationic sorption 




Particle surface 



PH 



Solution 

Interracial 
region 



ft 



(b) 



E 

£U 
-Q 

O 
t/1 



NOM / 



' Fe oxide 




8 



PH 



Fig. 1. Interfacial reactions involving particle-surface functional groups and solution ions play 
a defining role in controlling metal partitioning between solid and solution phases, ultimately 
influencing resulting metal behaviour, (a) Particle surfaces in aqueous systems carry a net overall 
charge reflecting the acid-base characteristics of their functional surface groups, dependent on 
system pH conditions. Generally, as pH decreases, greater anionic sorption occurs as particle surface 
charge becomes net positive, reflecting greater uptake of protons. In contrast, as pH increases, a 
greater net negative charge exists on particle surfaces so that cationic solution species such as 
divalent metal ions (e.g. Cd 2+ ) sorb more effectively to the negatively charged particle. The pH at 
which individual particles (e.g. different minerals, organic matter, dead or alive) become negatively 
or positively charged is dependent on the structural composition of the functional groups involved. 
Typically, most mineral functional groups are hydroxyl groups (as shown), although organic matter 
can have several types of functional groups with widely differing p/C a values (e.g. carboxylic versus 
amine groups; Coxef a/., 1999). (b) Generalized sorption edges shown for three important metal 
sorbents in natural systems. Most cationic elements show a characteristic exponential increase in 
sorption over a relatively small pH range for a given solid, reflecting pH-p/C a relationships for that solid. 
Note that the pH at which the 'edge' effect occurs for a given solid will vary depending on the metal 
involved. Mn oxyhydroxides and organic matter are typically negatively charged at much lower pH 
values than Fe oxides, reflecting the differences in the acid-base characteristics of their respective 
surface functional groups, and thus can sorb cationic species at much lower pH values than can 
Fe oxyhydroxides, which, in contrast, are often effective anionic sorbents at low pH. 



SGM symposium 65 



Biofilms and metal geochemistry 17 

sorbents (e.g. Fe oxyhydroxides) are less effective. Further, cells need be neither viable 
nor intact to act as metal sequesters (Ferris et al., 1989), as the functional groups 
associated with their cell-wall structures can still sorb metals (often with higher 
sorptive capacities) when the cells die or lyse (Urrutia, 1997). 

Most particles in solution carry a charge and, as that surface becomes more net 
negatively or positively charged (many mineral surfaces are amphoteric), their relative 
abilities to sorb cationic or anionic species, respectively, increases. Thus, the over- 
whelming importance of pH in regulating trace-metal sorption across systems has been 
widely accepted (Stumm, 1992; Stumm & Morgan, 1996). As pH increases, there is a 
greater net deprotonation of particle surfaces and associated sorption of cationic metal 
species, partitioning them to the solid, and less mobile, bioavailable and toxic, phase. In 
contrast, as pH decreases, there is greater sorption potential for anionic species 
reflecting the greater positive charge associated with particle surfaces (the exact pH at 
which this occurs is solid dependent). Further, adding to the complexity in dealing with 
heterogeneous solid and metal mixtures as is observed in the environment, 'edge' effects 
are observed for metal sorption to a given solid (Fig. lb), whereby sorption increases 
substantially over a relatively narrow pH range as the pH of the system moves above the 
pK a for dissociation of the specific solid's surface acid-base sites (Dzombak & Morel, 
1990). However, these pH edges are both solid dependent, reflecting the different pH 
values at which a surface becomes net negatively or positively charged (Fig. lb), and 
metal dependent, reflecting differential affinities of specific elements (often correlated 
to a metal's ability to hydrolyse). As is evident from Fig. 1, organic matter, e.g. micro- 
organisms, and Mn oxyhydroxides are able to sorb cationic species at much lower pH 
values than Fe oxyhydroxides (reflecting their different pH values, or pH at which 
their surfaces shift between net negative and net positive overall charge; Fig. 1). This 
makes them of key importance in low-pH systems, whilst Fe oxyhydroxides would be 
more important for anionic species under lower pH conditions. Typically, these 
sorption edges are also element specific for a given solid phase, indicating that selective 
affinity occurs. This negates the possibility of a generalized model for metal sorption 
independent of either the metal or solid involved. 

In brief, the role of ionic strength as a control in metal solid-solution partitioning 
derives from its influence on the potential electrostatic attractions that occur in 
solution, as well as on the nature of the interfacial region of the solid surface (Puis 
et al., 1991; Lores & Pennock, 1998). The most labile fraction of solid-associated 
metals is often referred to as 'exchangeable', referring to those metals that are 
electrostatically attracted to particle surfaces, but not necessarily bonded to a specific 
site on the surface or to a specific sedimentary component within a mixed sediment 

SGM symposium 65 



18 L.A.Warren 



pool. This fraction is the most easily released back into solution with any changes 
in solution conditions. In freshwater systems, increasing ionic strength increases 
both solution interactions, thereby decreasing element activity and compressing the 
electrical double layer surrounding the particle, and the associated sorptive uptake of 
metals (Drever, 1997). 



The role of redox status 

Changes in redox status can affect both solids and metals that are redox sensitive, 
influence partitioning between the solid and solution pools and thus change metal 
behaviour. Oxyhydroxide minerals are particularly sensitive to changes in redox status, 
such that reducing conditions can lead to their reductive dissolution and the associated 
release of any sorbed metals into solution. Oxidizing conditions would favour 
(especially for Fe) solid formation and associated sequestration of metals. Reductive 
dissolution of Fe and Mn oxyhydroxides is commonly driven by micro-organism Fe and 
Mn reduction coupled to organic matter degradation (see equation below; note that 
'CH 7 0' represents organic matter), where reduction of the oxidized Fe and Mn in the 
oxyhydroxide mineral leads to dissolution of the solid and release of any associated 
metals from the oxyhydroxide into solution: 



4Fe(III)OOH(s) + CH 2 + 8H + ^4Fe(II) + C0 2 +7H 2 

Typically, reductive dissolution of both oxyhydroxides, as well as oxidative dissolution 
of sulfidic minerals or organic matter, effectively proceeds only with microbial cata- 
lysis: abiogenic reaction rates are extremely slow (e.g. Morel & Hering, 1993; Stumm & 
Morgan, 1996; O'Day et al., 2004). Further, some redox-active (and often therefore 
bioactive) metals such as U, Cr and As show profoundly different behaviour depending 
on their redox status [e.g. U(VI) or U(IV), Cr(VI) or Cr(III), As(V) or As(III)]. For 
instance, Cr speciation is controlled by a shifting array of processes that include redox 
transformations, precipitation/dissolution and sorption/desorption reactions, all of 
which are intimately tied to both oxidation state and the geochemical status of a system 
(Bartlett, 1991; Richard &; Bourg, 1991; Baruthio, 1992). In its VI oxidation state, Cr 
commonly occurs as the ligand CrOj - , which is highly mobile, toxic and soluble (Felter 
& Dourson, 1997) and is generally controlled by sorption-desorption reactions at 
lower concentrations. Cr(VI), as chromate, is sorbed to minerals under lower pH 
conditions, when surface sites are positively charged. For example, Zachara et al. (1989) 
have shown that the degree of chromate sorption is greater in acidic soils and in 
subsurface media containing iron oxyhydroxides and clays. The reduced form of Cr, 
Cr(III), is insoluble in aqueous environments. Precipitation of Cr(OH) 3 or (Cr,Fe)- 
(OH) 3 compounds commonly controls its behaviour. 

SGM symposium 65 



Biofilms and metal geochemistry 19 

THE OVERLAP BETWEEN MICROBIOLOGY AND METAL 
GEOCHEMISTRY 

The role of emerging technologies 

It is only a small step across discipline divides to see the very real and often significant 
overlap between microbial activity in the environment and metal geochemistry In large 
part, the recognition that microbial activity can affect metal geochemistry has grown 
with our ability to examine the linkages at the appropriate scale, i.e. the micron level. 
Our ability to effectively probe and 'image' at the appropriate scale of resolution, using 
such techniques as X-ray absorption spectroscopy (XAS; Brown & Sturchio, 2002), 
atomic force microscopy (AFM) and scanning transmission X-ray microspectroscopy 
(STXM; O'Day, 1999), has increased our understanding of the mechanisms involved in 
solid-metal interactions. In addition, perhaps the most powerful tool which revolution- 
izes our ability to examine micro-organisms in an environmental context has been the 
development of culture-independent molecular-biological tools for phylogenetic 
characterization (e.g. Pace, 1997; Kaeberlein et al., 2002), which have permitted evalu- 
ation of community genetic diversity. 

Relevance of microbial functional metabolism 

Micro-organisms can influence metal behaviour through a number of processes, 
reflecting the direct link between types and rates of metabolic activity and geochemical 
conditions, processes and reaction rates (Fig. 2). The occurrence of micro-organisms in 
almost every environment investigated (e.g. Bennett et al., 2000; Bond et al., 2000; 
Chapelle et al., 2002; Takai et al., 2004) hints at their widespread influence on geo- 
chemical processes. Unlike higher eukaryotic organisms, micro-organisms are not 
restricted by geographical barriers (Finlay, 2002). Micro-organisms seek energy and 
carbon to survive and grow. In the environment, they select for geochemical conditions 
which are favourable and commonly catalyse geochemical processes for their growth 
(Nealson & Stahl, 1997; Nealson, 2003). While thermodynamics sets energetic 
constraints on microbial activity, i.e. organisms must live within the realm of reactions 
that are thermodynamically favourable, it is increasingly clear that the extent of this 
domain of possible reactions in the geosphere is not yet completely described (e.g. 
Spear etal., 2005). 

Beyond the simple aerobic versus anaerobic demarcation in metabolism, there are 
myriad potential metabolic pathways (i.e. electron donors and acceptors) with 
relevance for metal behaviour through which micro-organisms can selectively shape 
geochemical processes dependent on the particular redox couples that they catalyse. 
Such processes as nitrification and denitrification, methanogenesis, methanotrophy, 
sulfur/iron/manganese oxidation and reduction (Holt & Leadbetter, 1992; Nealson & 

SGM symposium 65 



20 L. A. Warren 



UJ 

U 

O 
a* 



CO 





Mechanism 



Indirect 



Fe/Mn redox 

reactions 



»* 






/ 



* 



Fe/Mn oxides 






Dissolution 
\ 



Formation 



\ 



Geochemical 

microenvironments 



V 



4 

Su rf ace 
reactions 
/ 
/ 
I 



Release of M 



z+ 



Sorption/immobilization of M 



z+ 





Increased mobility, 
bioavailability, toxicity 



Decreased metal impacts through 
decreased solution partitioning 



Fig. 2. Microbial influence on metal geochemistry can occur through both indirect and direct 
mechanisms. In both instances, microscale influence by microbial metabolic activity can scale to bulk- 
system impacts for metal behaviour, depending on the level of metabolic activity. Indirect mechanisms 
refer to general changes in the local geochemical environment associated with metabolic activity that 
then directionally affect metal solid-solution partitioning. Direct mechanisms refer to those specifically 
catalysed by microbes, e.g. Fe oxidation will lead to the formation of Fe oxyhydroxides and will likely 
lead to metal sequestration, while Fe reduction will dissolve Fe oxyhydroxides, liberating any 
associated metals into solution, increasing their mobility, bioavailability and likely impact. 



Stahl, 1997) and hydrogen-based metabolism (Chapelle et ai, 2002; Spear et al., 2005) 
can all substantively affect metal speciation and thus behaviour (Fig. 3). 

In addition to the ecological factors of sufficient nutrients, water and protection from 
predation, probably one of the most important parameters defining metabolic habitats 
is oxygen concentration and/or flux. This is also a defining geochemical parameter 
controlling the oxidative state of reactive geochemical components and thus segre- 
gating differing geochemical processes within the environment. Thus, it is clear that 
aerobic and anaerobic organisms will select for differing environments on the basis of 



SGM symposium 65 



Biofilms and metal geochemistry 21 




Metal behaviour - 
fate, toxicity, bioavailability etc. 



Metal toxicity feedback 
on biological activity f 



/ 




/ 



Solid-solution partitioning 



* 



Dynamic process determined by 

identity and concentration of 
both solid and solution players 
and reflecting system 
geochemical conditions 



Biological activity 

Dependent on extent and 
type of microbial activity; 
influences geochemical 
properties through; 

(i) PASSIVE: influencing 
micro-geochemical 
conditions through 
metabolism(s) 

(ii) ACTIVE: specific 

metabolic control - 
solids or metals 
i.e. mineral dissolution, 
metal release 




Biological— 

geochemical 
linkages 
interdependent: 
feedback occurs 



PH 

Solid-sorption interface! reactions 
L^ 4 pH leads to: 

* Solid-associated M z+ 

Redox status 

Solids: redox sensitive 

L>. Formation/dissolution 

Metals: some redox-active metals 

L* So I i d aff i n i ty red ox-state 
dependent 



CONTROL 



Biological 



Geochemical 



Will reflect the nature and magnitude 
of micro-organism activity 



Fig. 3. Understanding the controls of metal geochemistry is increasingly interpreted as a dynamic 
array reflecting the interactions between microbial activity and the important controls on metal 
solid-solution partitioning. Feedback mechanisms of local geochemical conditions on existence, 
activity and growth of micro-organisms also occur. The relative dominance of microbial or classic 
geochemical controls on metal behaviour will reflect the extent of microbial activity in a given 
system or at a given time. 



oxygen concentration, which similarly stratifies geochemically important solids and/or 
forms of elements important in metal behaviour. Thus, metal impacts will be selectively 
dependent on both the types of metabolism and levels of activity occurring in any 
micro-environment. Regardless of the specific metabolic guild involved, the common 
and widespread occurrence of microbial control on important geochemical processes is 
increasingly demonstrated (e.g. biogenic Fe oxide formation: Chafetz et al., 1998; 
Newman & Banfield, 2002; Haack & Warren, 2003). 



SGM symposium 65 



22 L. A. Warren 



As field investigations (e.g. Smiejan et al., 2003; Brown et al., 2004; Labrenz & Banfield, 
2004; Meylan et al., 2004) and associated laboratory experimentation to probe mech- 
anisms and effects (e.g. Warren & Ferris, 1998; Parmar et al., 2000; Moreau et al., 2004; 
Tani et al., 2004a, b; Templeton et al., 2003a, b) demonstrate, micro-organisms influ- 
ence trace-metal behaviour through both passive and active mechanisms (Fig. 2). 
Passive or indirect mechanisms are viewed as those associated with general metabolic 
activity, i.e. where the particular metabolic pathways and levels of activity associated 
with a given microbial community induce changes in geochemical conditions within the 
immediate micro-geochemical environment. These can then influence metal behaviour, 
e.g. increasing pH would generally favour sorption of more metals to the solid com- 
partment. Active or direct mechanisms are defined as those that reflect direct catalysis 
by micro-organisms of geochemical processes that control metal behaviour, e.g. 
reduction of Fe resulting in the dissolution of Fe oxide minerals and the associated 
release of metals into solution. Increasing evidence indicates the ability of many micro- 
organisms to directly metabolize many bioactive (redox-active) elements such as U, Cr 
and As (e.g. Lores & Pennock, 1998; Lovley & Anderson, 2000; Liu et al., 2002), which 
alters the behaviour of these metals in the process. 

Micro-organisms can also profoundly influence metal behaviour through associated 
impacts on mineral formation and dissolution (Konhauser et al., 1993; Bennett et al., 
2000; Haack & Warren, 2003; Templeton et al., 2003a; Moreau et al., 2004). However, 
it is also clear that the geochemical conditions in which micro-organisms find them- 
selves can influence their viability, activity and growth and, commonly, interactions 
between microbial behaviour and geochemistry exert feedback on each other (Warren 
& Haack, 2001; Warren & Kauffman, 2003). Thus, there is the potential for geo- 
chemical feedback on biological processes that will in turn influence the metal 
behaviour observed, which clearly indicates that the linkages between microbial activity 
and geochemical behaviour are complex and dynamic (Fig. 3). 

Biofilms: existence, structure and function 

It is increasingly clear that many micro-organisms self-organize into diverse biofilm 
communities (Davies et al., 1998; Branda et al., 2005), which can produce complex and 
dynamic internal geochemical conditions based on the particular array of metabolic 
processes occurring within the biofilm (Little et al., 1997). Biofilms form at interfaces 
between solid and solution phases, overlapping with the same dynamic reactive zone 
that controls metal partitioning. Biofilms can also form on particles suspended in 
solution (Leppard et al., 2003, 2004; Roberts et al., 2004) or at the bed-sediment surface 
(Fig. 4a; Vigneault et al., 2001; Haack & Warren, 2003). In either scenario, these 
biologically controlled layers can selectively control metal-solid interactions depending 
on the particular geochemical conditions created by the biofilm's specific metabolic 

SGM symposium 65 



Biofilms and metal geochemistry 23 

array (Fig. 4b). Biofilms must provide a selective advantage for the organisms that form 
them. Typically, biofilms are thought to provide structural protection, through the 
associated EPS matrix, from predation and/or impacts of potentially toxic con- 
taminants (Lawrence et ai, 1995, 1998; Neu & Lawrence, 1997; Davies et al., 1998). 

Mixed-community biofilms exhibit stratified metabolic functions reflecting the ener- 
getic array of reactants and products which link these microbial communities together. 
Further, minerals are commonly associated with biofilms (e.g. Douglas 6c Beveridge, 
1998; Ferris et ai, 1999; Haack 6c Warren, 2003; Templeton et ai, 2003a; Moreau et ai, 
2004), either through passive biomineralization, where the EPS and/or cells of the bio- 
film act as nucleation templates for minerals to form, or actively, where specific minerals 
are catalysed through specific functions of the biofilm. In either scenario, minerals are 
commonly associated with biofilms as either bioreactants and/or bioproducts of micro- 
bial metabolism. 

Biofilm metabolic links to metals 

Since biofilms represent a concentrated cell mass with significant metabolic activity, 
their occurrence in the environment can lead to a significant biological overlay on 
geochemically reactive processes relevant to metal behaviour that will be governed by 
the ecological controls on biofilm formation and function. While it is generally held 
that most systems are at or near equilibrium with respect to acid-base reactions (e.g. 
pH), it is very clear that, without microbial catalysis of important redox geochemical 
reactions, many would not occur over relevant timescales due to the extremely slow 
kinetics of many of these reactions in the geosphere. Another clearly important factor 
to consider is that geochemical processes that proceed abiogenically (geochemically 
controlled) often do so under different environmental conditions and controls and with 
potentially different geochemical outcomes from biogenically micro-organism- 
catalysed processes (e.g. Warren 6c Ferris, 1998; Haack 6c Warren, 2003; Tani et ai, 
2004a, b). 

Environmental biofilm metal geochemistry 

Biofilms are common in metal-contaminated and/or low-pH systems (Southam 6c 
Beveridge, 1992; Ledin & Pedersen, 1996; Nordstrom & Southam, 1997; Edwards et 
ai, 1999, 2000; Hunt et ai, 2001; Leveille et ai, 2001; Vigneault et ai, 2001; Baker et ai, 
2004; Labrenz 6c Banfield, 2004), indicating that these extreme conditions provide 
favourable conditions for adapted microbial communities to flourish. They also 
indicate that there is significant potential for biofilms in these environments to impact 
metal geochemistry. As yet, there are not many studies that specifically evaluate the 
linkage between metal behaviour and biofilm metabolic functional activity However, 
acid mine drainage (AMD) has long been recognized as one of the earliest environ- 

SGM symposium 65 



24 L. A. Warren 



SOLID: 

suspended AW 



particles 



-►-., SOLUTION 

Metal partitioning? ^ 
Biofilm r 





SOLID: bed 

sediments 



(a) 



SOLUTION 



EMR 



Metal 
L_ 



(b) 



Sequestration 



Solubilization 



BIOFILM 

Photoautotrophs 



H + +C0 2 + P + N 



**. 





(CH 2 0) n 



u 

X 

o 



Mn oxi d-fze rs Qxic-anoxic interfa ce M n 2+ ■ 



Fe oxidizers 

* 

Sibxidizers H 2 S 



Fe 2 ' 



-► S< 



u 

X 

o 



Fe/Mn reducers 
S reducers/ 



> s 2 or 



Mn 2+ ^- 
Fe 2+ <«- 



]" ,, "j--Bjomineral formation 



Mn oxides**.. 




Fe oxides 
Fe 3+ S0 4 minerals 



"Pr 2 5-*tn7Z.S° < S 2 0§" 

disproportionators """* <; 



M 



Mineral dissoluticfh 




Fig. 4. Biofilms form at solid-solution interfaces, either at bed-sediment solution interfaces or at 
suspended particle surfaces (a), thus overlapping with the reactive interface for metal solid-solution 
partitioning. These biologically controlled interracial structures can profoundly influence the nature, 
extent and types of geochemical processes that occur, particularly with respect to those that influence 
metal partitioning in a manner dependent on the particular array of metabolic functions expressed 
within a given biofilm (b). As shown in (b), in this case a biofilm with phototrophs occurring at the 
surface driving 2 fluctuations within the surface biofilm, a generalized microbial biofilm is typically 
a stratified community that will show depth-dependent metabolic pathways typically as a function 
of oxygen status, which will also strongly influence the types of geochemical processes and the 
nature of both solid and solution elements that can occur at any given depth within the biofilm 



SGM symposium 65 



Biofilms and metal geochemistry 25 

ments for evidence of microbially controlled geochemical processes (e.g. Ledin & 
Pedersen, 1996; Nordstrom & Southam, 1997), providing some evidence of the shift in 
geochemical behaviour with micro-organism activity. AMD is strongly driven by 
microbially induced weathering and oxidation of sulfidic (e.g. pyrite, pyrrhotite) 
minerals exposed during the mining of base metals and coal (Fig. 5). The oxidation of 
sulfide-containing waste rock is significantly catalysed by iron- and sulfur-oxidizing 
bacteria that are attached or in close proximity to the sulfide minerals. This oxidation 
produces large amounts of acid. Recent investigation of AMD-associated biofilm metal 
dynamics (Haack & Warren, 2003) indicates that these biological structures are 
efficient metal sequesterers (Fig. 4b), through bacterially catalysed biomineralization 
reactions that would not be predicted to occur abiogenically 

While studies are emerging that provide direct evidence of both biomineralization by 
biofilms (e.g. Haack & Warren, 2003; Templeton et al., 2003b; Labrenz & Banfield, 
2004; Moreau et al., 2004) and metal uptake by biofilms (e.g. Templeton et al., 2001, 
2003a; Vigneault et al., 2001; Haack & Warren, 2003), as yet relatively few have 
integrated metal geochemistry dynamics with microbial community dynamics. 
However, those studies that have been done demonstrate that the following two major 
processes influence biofilm metal uptake. Firstly, the process of biofilm-associated 
mineral formation, whether passive or active, often plays a key role in metal seques- 
tration (e.g. Nelson et al., 1995, 1999; Ferris et al., 1999). However, subsequent 
dissolution of biominerals, which tend to be small and amorphous, can lead to 
dynamic temporal fluctuations in biofilm metal content as mineral formation and 
dissolution are cyclically catalysed (Haack & Warren, 2003). Secondly, the organic 
fraction of the biofilm can also play a substantive role in metal sequestration through 
sorption to EPS as well as to cell walls (live and/or dead cells) and intracellular metal 
uptake, accumulation and storage (Beveridge, 1989; Schorer 6c Eisele, 1997; Liinsdorf 
et al., 1997; Vigneault et al., 2001; Haack & Warren, 2003; Brown et al., 2004; Meylan 
etal.,2004). 



('CH 2 0' represents organic carbon molecules; EMR, electromagnetic radiation). In an interlinked 
fashion, the micro-organisms dependent on the geochemical components for their growth and 
survival and the geochemical components influenced by the microbial catalysis of reactions driving 
their consumption or production, a biofilm is a geochemical reactor that is strongly driven by biology 
and overlaps directly with the reactive metal zone between solids and solution. The nature of the 
metabolic processes that occur within each functional metabolic stratum can profoundly influence 
key geochemical processes controlling metal behaviour, particularly those reactions involving the 
formation and dissolution of metal reactive minerals such as Fe and Mn oxyhydroxides (as shown in 
b). Thus, biofilm metal behaviour is likely to vary across systems, reflecting the specific interplay of 
geochemistry and microbiology, as well as for a given biofilm structure in both a depth-dependent 
and a temporal manner. If the biofilm has a phototrophic component (as exemplified in b), this will 
show shifts in respiration dominance (and therefore 2 saturation) over diel timescales. 



SGM symposium 65 



26 L. A. Warren 




FeS 2 + 3*50 2 + H 2 
2Fe 2+ + 0-5O 2 + 2H + 



FeS 2 +14Fe 3+ + 8H 2 



Fe 2+ + 2SO|-+2H + 


(D 


Biocatalysed 
reactions 


2Fe 3+ + H 2 


(2) 


DRIVE 



I 



15Fe 2+ + 2S0 2 " + 16H+ (3) 



Abiotic 
reaction 



H 2 0, 2 
micro-organisms 




H 2 0, 2 
micro-organisms 



AMD creation: 

highly acidic and 

metal-laden 

discharge 



Fig. 5. The weathering of tailings, metal-sulfide waste rock material, is catalysed through bacterial 
action (the diagram indicates that reactions 1 and 2 are biogenic and drive reaction 3, an abiogenic 
process that liberates significant acid). Abiogenic oxidation/weathering of these reduced sulfidic 
minerals, even when exposed to 2 , are exceedingly slow. The bacterium Thiobacillus ferrooxidans 
catalyses oxidation of FeS (x) minerals through direct oxidation of FeS (x) (pyrite, FeS 2 , used here) and 
minerals (equation 1 ) and through the oxidation of Fe 2+ to Fe 3+ (equation 2, increasing rates of Fe 2+ 
oxidation up to five orders of magnitude; Singer & Stumm, 1 970). This biogenically produced Fe 3+ 
then abiogenically catalyses further (more rapid) abiogenic weathering of FeS (x) (equation 3). It is 
reaction 3 that drives significant acid production associated with AMD. In the process, acid is liberated, 
metals are typically solubilized and resulting AMD is both acidic and metal laden. The diagram 
indicates that the microbe-driven processes of (1 ) and (2) (top left half of diagram) drive abiogenic 
Fe 3+ -subsequent weathering of the FeS (x) minerals. 



CONCLUSIONS 

Trace-metal behaviour is controlled by a complex and shifting array of processes that 
include redox transformations, precipitation/dissolution and sorption/desorption 
reactions, all of which are intimately tied to the geochemical status of a system and can 
be selectively influenced by microbial activity in a community-dependent manner, 
reflecting the particular functional array involved. Our understanding of how biology 
and geochemistry integrate in such a dynamic array of potential interactions is, at best, 
rudimentary However, it is clear that, in many environmental instances, an overlap 
occurs between microbial activity and metal behaviour, reflecting the physical over- 
lap of microbial biofilm habitat and reactive metal zones in the environment. Micro- 



SGM symposium 65 



Biofilms and metal geochemistry 27 

organisms influence their immediate micro-geochemical environment as well as directly 
catalyse geochemical transformations required for their survival and growth. In both 
instances they can profoundly influence the metal behaviour observed. However, micro- 
organisms are also intimately affected by their environment and thus feedback is 
exerted in both directions as dynamic and fluctuating geochemical conditions influence 
metabolic pathways, rates and times of expression. Thus, it is evident that under- 
standing and predicting geochemical behaviour requires an understanding of the often 
complex and dynamic linkages between ecological controls on where micro-organisms 
exist, how they make their living and their level of metabolic activity, and the geo- 
chemical processes they use to sustain themselves. Further, microbial plasticity ensures 
that organisms not only exploit available redox couples, but they also respond to 
microscopic variations in geochemical conditions, whilst simultaneously creating 
heterogeneity from the output of their own metabolic activity (Finlay, 2002). Thus, 
system-dependent and site-dependent microscale spatial and temporal heterogeneity 
needs to be investigated in order to constrain at the geosphere level the complex 
interactions of micro-organisms and geochemistry that result in the ultimate behaviour 
observed. 

The challenge for this emerging field is to define the relevant controls in an integrated 
ecological and geochemical space and to develop sensitive techniques for their detection 
and accurate quantification. Emerging techniques are providing new insight into 'which 
organisms are present' in different environments, a fundamental question that was, 
until recently, difficult to answer. The extension of this question to 'what are they doing' 
is the current focus and the future of this area, as we seek to address mechanistically 
how microbial metabolism or activity links directly to geochemical reactions and 
shapes outcomes of relevant processes (Croal et al., 2004; Warren, 2004). As recog- 
nition of the need to examine the types and levels of metabolic activity occurring in the 
environment (e.g. proteomics, metallomics) has grown in appreciation of the challenge 
of quantifying microbial links to geochemical reactions, increasing methodological 
development has occurred. It is still a challenge to characterize metabolic pathways 
with microbial strains only identified on the basis of 16S rRNA oligonucleotide 
sequences, as so few environmental micro-organisms are currently in culture. This 
means that, while an exponentially growing database of environmental strains is 
genetically identified, the vast majority of these novel organisms remain undescribed in 
terms of their metabolic pathways. Because of this hurdle, other techniques are focused 
on evaluation of gene expression in the environment (dependent on their character- 
ization) or on using isotopes to track community diversity, structure and function. The 
field of molecular ecology is providing exciting new studies that track both molecular 
diversity and function, with apparently promising selectivity, through both labelling 
and probing of environmental processes using stable isotopes (e.g. Engel et al., 2004; 

SGM symposium 65 



28 L. A. Warren 



Griffiths et al., 2004; Lueders et ai, 2004; Manefield et al., 2004; Pearson et ai, 2004; 
Andersen et ai, 2005; Lu et al., 2005). Until such time as gene expression (i.e. functional 
activity) for any specific organism in the environment is easily identifiable and 
quantifiable, the use of isotopes as an additional tool to probe actively occurring 
geochemical transformations linked to microbial activity appears highly promising. 

ACKNOWLEDGEMENTS 

This chapter presents a synthesis of ideas that have grown, coalesced and crystallized predomi- 
nantly from the many discussions and interactions I have had with my students over the last 
5 years. I would like especially to thank Dr Elizabeth Haack, Luc Bernier, Derek Amoves, Corrie 
Kennedy, Tara Nelson and Lisa Melymuk. I would also like to acknowledge substantial funding 
support from NSERC, CF1, OIT, McMaster University and Noranda/Falconbridge Ltd. 

REFERENCES 

Allard, B., Hakansson, K. & Karlsson, S. (1987). The importance of sorption phenomena in 

relation to trace element speciation and mobility. In Speciation of Metals in Water, 

Sediment and Soil Systems, Lecture Notes in Earth Sciences, no. 11, pp. 99-1 12. 

Edited by L. Landner. Berlin: Springer. 
Andersen, J. S., Lam, Y. W., Leung, A. K. L., Ong, S. E., Lyon, C.-E., Lamond, A. I. & 

Mann, M. (2005). Nucleolar proteome dynamics. Nature 433, 77-83. 
Baker, B. J., Lutz, M. A., Dawson, S. C, Bond, P. L & Banf ield, J. K. (2004). Metabolically 

active eukaryotic communities in extremely acidic mine drainage. Appl Environ 

Microbiol 70, 6264-627 '1. 
Bartlett, R. J. (1991). Chromium cycling in soils and water: links, gaps, and methods. 

Environ Health Perspect 92, 1 7-24. 
Baruthio, F. (1992). Toxic effects of chromium and its compounds. Biol Trace Elem Res 32, 

145-153. 
Basta, N. T., Ryan, J. A. & Chaney, R. L. (2005). Trace element chemistry in residual-treated 

soil: key concepts and metal bioavailability. J Environ Qual 34, 49-63. 
Bennett, P. C, Hiebert, F. K. & Roberts Rogers, J. (2000). Microbial control of mineral- 

groundwater equilibria: macroscale to microscale. Hydrogeol J 8, 47-62. 
Beveridge, T. J. (1989). Role of cellular design in bacterial metal accumulation and 

mineralization. Annu Rev Microbiol 43, 147-171 . 
Bond, P. L., Smriga, S. P. & Banf ield, J. F. (2000). Phylogeny of microorganisms populating 

a thick, subaerial, predominantly lithotrophic biofilm at an extreme acid mine 

drainage site. Appl Environ Microbiol 66, 3842-3849. 
Boullemant, A., Vigneault, B., Fortin, C. & Campbell, P. G. C. (2004). Uptake of neutral 

metal complexes by a green alga: influence of pH and humic substances. Aust J 

Chem 57, 931-936. 
Branda, S. S., Vik, A., Friedman, L. & Kolter, R. (2005). Biofilms: the matrix revisited. 

Trends Microbiol 1 3, 20-26. 
Brown, G. E., Jr & Parks, G. A. (2001). Sorption of trace elements on mineral surfaces: 

modern perspectives from spectroscopic studies, and comments on sorption in the 

marine environment. Int Geo! Rev 43, 963-1 073. 
Brown, G. E., Jr & Sturchio, N. C. (2002). An overview of synchrotron radiation applications 

SGM symposium 65 



Biofilms and metal geochemistry 29 



to low temperature geochemistry and environmental science. Rev Miner Geochem 

49, 1-115. 
Brown, G. E., Jr, Henrich, V. E. f Casey, W. H. & 11 other authors (1999). Metal oxide 

surfaces and their interactions with aqueous solutions and microbial organisms. 

Chem Rev 99, 77-1 74. 
Brown, G. E., Jr, Yoon, T. H., Johnson, S. B., Templeton, A. S., Trainor, T. P., Bostick, 

B. C, Kendelewicz, T., Doyle, C. S. & Spormann, A. M. (2004). The role of organic 

molecules and microbial organisms in metal ion sorption processes. Geochim 

Cosmochim Acta 68, A1 60. 
Buffle, J. (1988). Complexation Reactions in Aquatic Systems: an Analytical Approach. 

Chichester: Ellis Horwood. 
Campbell, P. G. C. (1995). Interactions between trace metals and aquatic organisms: a 

critique of the free-ion activity model. In Metal Speciation and Bioavailability in 

Aquatic Systems, IUPAC Series on Analytical and Physical Chemistry of Environmental 

Systems, pp. 45-102. Edited by A. Tessier & D. Turner. Chichester: Wiley. 
Campbell, P. G. C. & Tessier, A. (1989). Geochemistry and bioavailability of trace metals in 

sediments. In Aquatic Ecotoxicology, pp. 124-148. Edited by A. Boudou & F. Ribeyre. 

Boca Raton, FL: CRC Press. 
Chafetz, H. S., Akdim, B., Julia, R. & Reid, A. (1998). Mn- and Fe-rich black travertine 

shrubs; bacterially (and nanobacterially) induced precipitates. J Sediment Res 68, 

404-412. 
Chapelle, F. H., O'Neill, K., Bradley, P. M., Methe, B. A., Ciufo, S. A., Knobel, L L & 

Lovley, D. R. (2002). A hydrogen-based subsurface microbial community dominated 

by methanogens. Nature 41 5, 3 1 2-3 1 5. 
Cox, J. S., Smith, D. S., Warren, L. A. & Ferris, F. G. (1999). Characterizing heterogeneous 

bacterial surface functional groups using discrete affinity spectra for proton binding. 

Environ Sci Techno! 33, 45 1 4-452 1 . 
Croal, L. R., Gralnick, J. A., Malasarn, D. & Newman, D. K. (2004). The genetics of 

geochemistry. Annu Rev Genet 38, 1 75-202. 
Davies, D. G., Parsek, M. R., Pearson, J. P., Iglewski, B. H., Costerton, J. W. & 

Greenberg, E. P. (1998). The involvement of cell-to-cell signals in the development 

of a bacterial biofilm. Science 280, 295-298. 
Douglas, S. & Beveridge, T. J. (1998). Mineral formation by bacteria in natural microbial 

communities. FEMS Microbiol Ecol 26, 79-88. 
Drever, J. I. (1997). The Geochemistry of Natural Waters: Surface and Groundwater 

Environments, 3rd edn. London: Prentice-Hall. 
Dzombak, D. A. & Morel, F. M. M. (1990). Surface Complexation Modeling: Hydrous Ferric 

Oxide. New York: Wiley. 
Edwards, K. J., Gihring, T. M. & Banfield, J. F. (1999). Seasonal variations in microbial 

populations and environmental conditions in an extreme acid mine drainage 

environment. Appl Environ Microbiol 65, 3627-3632. 
Edwards, K. J., Bond, P. L, Gihring, T. M. & Banfield, J. F. (2000). An archaeal 

iron-oxidizing extreme acidophile important in acid mine drainage. Science 287, 

1796-1799. 
Ehrlich, H. L. (2002). How microbes mobilize metals in ores: a review of current under- 
standings and proposals for further research. Miner Metallurg Process 19, 220- 

224. 
Engel, A. S., Porter, M. L, Stern, L. A., Quinlan, S. & Bennett, P. C. (2004). Bacterial 

diversity and ecosystem function of filamentous microbial mats from aphotic (cave) 



SGM symposium 65 



30 L. A. Warren 



sulfidic springs dominated by chemolithoautotrophic " Epsilonproteobacteria" . FEMS 

Microbiol Ecol 51 ,31-53. 
Felter, S. P. & Dourson, M. L. (1997). Hexavalent chromium-contaminated soils: options for 

risk assessment and risk management. Regul Toxicol Pharmacol 25, 43-59. 
Ferris, F. G., Schultze-Lam, S., Witten, T. C, Fyfe, W. S. & Beveridge, T. J. (1989). Metal 

interactions with microbial biofilms in acidic and neutral pH environments. Appl 

Environ Microbiol 55, 1 249-1 257. 
Ferris, F. G., Konhauser, K. O., Lyven, B. & Pedersen, K. (1999). Accumulation of metals 

by bacteriogenic iron oxides in a subterranean environment. Geomicrobiol J 16, 

181-192. 
Filgueiras, A. V., Lavilla, I. & Bendicho, C. (2002). Chemical sequential extraction for metal 

partitioning in environmental solid samples. J Environ Monit4, 823-857. 
Finlay, B. J. (2002). Global dispersal of free-living microbial eukaryote species. Science 296, 

1061-1063. 
Griffiths, R. I., Manefield, M., Ostle, N., McNamara, N., O'Donnell, A. G., Bailey, M. J. 

& Whiteley, A. S. (2004). 13 C0 2 pulse labelling of plants in tandem with stable 

isotope probing: methodological considerations for examining microbial function in 

the rhizosphere. J Microbiol Methods 58, 1 1 9-1 29. 
Gustafsson, J. P., Pechova, P. & Berggren, D. (2003). Modelling metal binding to soils: the 

role of natural organic matter. Environ Sci Techno! 37, 2161-211 A. 
Haack, E. A. & Warren, L. A. (2003). Biofilm hydrous manganese oxyhydroxides and metal 

dynamics in acid rock drainage. Environ Sci Technol 37, 41 38^-147. 
Hare, L. (1992). Aquatic insects and trace metals: bioavailability, bioaccumulation, and 

toxicity. Crit Rev Toxicol 22, 327-369. 
Hassler, C. S., Slaveykova, V. I. & Wilkinson, K. J. (2004). Some fundamental (and often 

overlooked) considerations underlying the free ion activity and biotic ligand models. 

Environ Toxicol Chem 23, 283-291 . 
Holden, J. F. & Adams, M. W. W. (2003). Microbe-metal interactions in marine hydro- 
thermal environments. Curr Opin Chem Biol 7, 1 60-1 65. 
Holt, S. C. & Leadbetter, E. R. (1992). Structure-function relationships in prokaryotic 

cells. In Prokaryotic Structure and Function: a New Perspective (Society for 

General Microbiology Symposium no. 47), pp. 1 1-44. Edited by S. Mohan, C. Dow 

& J. A. Coles. Cambridge: Cambridge University Press. 
Hunt, A. P., Hamilton-Taylor, J. & Parry, J. D. (2001). Trace metal interactions with epilithic 

biofilms in small acidic mountain streams. Arch Hydrobiol 1 53, 1 55-1 76. 
Islam, F. S., Gault, A. G., Boothman, C, Polya, D. A., Charnock, J. M., Chatterjee, D. & 

Lloyd, J. R. (2004). Role of metal-reducing bacteria in arsenic release from Bengal 

delta sediments. Nature 430, 68-71 . 
Jacob, D. L. & Otte, M. L. (2004). Long-term effects of submergence and wetland 

vegetation on metals in a 90-year old abandoned Pb-Zn mine tailings pond. Environ 

Pollute 30, 337-345. 
Jenne, E. A. (1968). Controls on Mn, Fe, Co, Ni, Cu and Zn concentrations in soils and 

water: the significant role of hydrous Mn and Fe oxides. In 7race Inorganics in Water, 

American Chemical Society Publication no. 73, pp. 337-387. Edited by R. A. Baker. 

Washington, DC: American Chemical Society. 
Johnson, B. B. (1990). Effect of pH, temperature, and concentration on the adsorption of 

cadmium on goethite. Environ Sci Technol 24, 1 1 2-1 1 8. 
Kaeberlein, T, Lewis, K. & Epstein, S. S. (2002). Isolating "uncultivable" microorganisms 

in pure culture in a simulated natural environment. Science 296, 1 1 27-1 1 29. 



SGM symposium 65 



Biofilms and metal geochemistry 31 



Konhauser, K. O., Fyfe, W. S., Ferris, F. G. & Beveridge, T. J. (1993). Metal sorption and 

mineral precipitation by bacteria in two Amazonian river systems: Rio Solimoes 

and Rio Negro, Brazil. Geology 21, 1 103-1 106. 
Labrenz, M. & Banfield, J. K. (2004). Sulfate-reducing bacteria-dominated biofilms that 

precipitate ZnS in a subsurface circumneutral-pH mine drainage system. Microb Ecol 

47,205-217. 
Lawrence, J. R., Korber, D. R., Wolfaardt, G. M. & Caldwell, D. E. (1995). Behavioural 

strategies of surface-colonizing bacteria. Adv Microb Ecol 14, 1-75.. 
Lawrence, J. R., Swerhone, G. D. W. & Kwong, Y. T. J. (1998). Natural attenuation of 

aqueous metal contamination by an algal mat. Can J Microbiol 44, 825-832. 
Ledin, M. & Pedersen, K. (1996). The environmental impact of mine wastes - roles of 

microorganisms and their significance in treatment of mine wastes. Earth 5c/7?e\/41, 

67-108. 
Lee, D. Y, Fortin, C. & Campbell, R G. C. (2004). Influences of chloride on silver uptake by 

two green algae, Pseudokirchneriella subcapitata and Chlorella pyrenoidosa. Environ 

Toxicol Chem 23, 1012-1018. 
Leppard, G. G., Droppo, I. G., West, M. M. & Liss, S. N. (2003). Compartmentalization of 

metals within the diverse colloidal matrices comprising activated sludge microbial 

floes. J Environ Qual 32, 2 1 00-2 1 08. 
Leppard, G. G., Mavrocordatos, D. & Perret, D. (2004). Electron-optical characterization 

of nano- and macro-particles in raw and treated waters: an overview. Water Sci 

Technol 50 (12), 1-8. 
Leveille, S. A., Leduc, L. G., Ferroni, G. D., Telang, A. J. & Voordouw, G. (2001). 

Monitoring of bacteria in acid mine environments by reverse sample genome 

probing. Can J Microbiol 47, 431-442. 
Little, B. J., Wagner, P. A. & Lewandowski, Z. (1997). Spatial relationships between 

bacteria and mineral surfaces. In Geomicrobiology: Interactions between Microbes 

and Minerals, Reviews in Mineralogy no. 35, pp. 123-1 59. Edited by J. F. Banfield & 

K. H. Nealson. Washington, DC: American Mineralogical Society. 
Liu, C. X., Gorby, Y A., Zachara, J. M., Fredrickson, J. K. & Brown, C. F. (2002). 

Reduction kinetics of Fe(lll), Co(lll), U(VI), Cr(VI), and Tc(VII) in cultures of dissimilatory 

metal-reducing bacteria. Biotechnol Bioeng 80, 637-649. 
Lores, E. M. & Pennock, J. R. (1998). The effect of salinity on binding of Cd, Cr, Cu and 

Zn to dissolved organic matter. Chemosphere 37, 861-874. 
Lovley, D. R. & Anderson, R. T. (2000). Influence of dissimilatory metal reduction on fate of 

organic and metal contaminants in the subsurface. Hydrogeol 7 8, 77-88. 
Lu, Y H., Lueders, T., Friedrich, M. W. & Conrad, R. (2005). Detecting active methano- 

genic populations on rice roots using stable isotope probing. Environ Microbiol 7, 

326-336. 
Lueders, T., Pommerenke, B. & Friedrich, M. W. (2004). Stable-isotope probing of 

microorganisms thriving at thermodynamic limits: syntrophic propionate oxidation 

in flooded soil. Appl Environ Microbiol 70, 5778-5786. 
Lunsdorf, H., Brummer, l.,Timmis, K. N. & Wagner-Dobler, I. (1997). Metal selectivity of 

in situ microcolonies in biofilms of the Elbe River. J Bacteriol 179, 31-40. 
Manefield, M., Whiteley, A. S. & Bailey, M. J. (2004). What can stable isotope probing do 

for bioremediation? Int Biodeterior Biodegrad 54, 1 63-1 66. 
Martell, A. E., Motekaitis, R. J. & Smith, R. M. (1988). Structure-stability relationships of 

metal complexes and metal speciation in environmental aqueous solutions. Environ 

Toxicol Chem 7, 414-434. 



SGM symposium 65 



32 L. A. Warren 



Martinez, C. E. & McBride, M. B. (1998). Solubility of Cd 2+ , Cu 2+ , Pb 2+ , and Zn 2+ in aged 
coprecipitates with amorphous iron hydroxides. Environ Sci Technol 32, 743-748. 

Martino, M., Turner, A. & Nimmo, M. (2004). Distribution, speciation and particle-water 
interactions of nickel in the Mersey Estuary, UK. MarChem 88, 161-177. 

Meylan, S., Behra, R. & Sigg, L. (2004). Influence of metal speciation in natural freshwater 
on bioaccumulation of copper and zinc in periphyton: a microcosm study. Environ Sci 
Technol 38, 3104-3111. 

Moreau, J. W., Webb, R. I. & Banf ield, J. K. (2004). infrastructure, aggregation-state, and 
crystal growth of biogenic nanocrystalline sphalerite and wurtzite. Am Miner 89, 
950-960. 

Morel, F. M. M. & Hering, J. G. (1993). Principles and Applications of Aquatic Chemistry. 
London: Wiley Interscience. 

Nealson, K. H. (2003). Harnessing microbial appetites for remediation. Nat Biotechnol 21 , 
243-244. 

Nealson, K. H. & Stahl, D. H. (1997). Micro-organisms and biogeochemical cycles: 
what can we learn from stratified communities? In Geomicrobiology: Interactions 
between Microbes and Minerals, Reviews in Mineralogy no. 35, pp. 5-34. Edited by 
J. F. Banf ield & K. H. Nealson. Washington, DC: American Mineralogical Society. 

Nelson, Y. M., Lo, W„ Lion, L W., Shuler, M. L & Ghiorse, W. C. (1995). Lead distribution 
in a simulated aquatic environment: effects of bacterial biofilms and iron oxide. 
Water Res 29, 1934-1944. 

Nelson, Y. M., Lion, L W., Ghiorse, W. C. & Shuler, M. L (1999). Production of biogenic 
Mn oxides by Leptothrix discophora SS-1 in a chemically defined growth medium 
and evaluation of their Pb adsorption characteristics. Appl Environ Microbiol 65, 
175-180. 

Neu, T. R. & Lawrence, J. R. (1997). Development and structure of microbial biofilms in 
river water studied by confocal laser scanning microscopy. FEM5 Microbiol Ecol 24, 
11-25. 

Newman, D. K. & Banfield, J. F. (2002). Geomicrobiology: how molecular-scale inter- 
actions underpin biogeochemical systems. Science 296, 1071-1077. 

Nordstrom, D. K. & Southam, G. (1997). Geomicrobiology of sulfide mineral oxidation. 
In Geomicrobiology: Interactions between Microbes and Minerals, Reviews in Miner- 
alogy no. 35, pp. 361-390. Edited by J. F. Banfield & K. H. Nealson. Washington, DC: 
American Mineralogical Society. 

O'Day, P. A. (1999). Molecular environmental geochemistry. Rev Geophys 37, 249-274. 

O'Day, P. A., Vlassopoulos, D., Root, R. & Rivera, N. (2004). The influence of sulfur and 
iron on dissolved arsenic concentrations in the shallow subsurface under changing 
redox conditions. Proc Natl Acad Sci USA 101, 13703-13708. 

Pace, N. R. (1997). A molecular view of microbial diversity and the biosphere. Science 276, 
734-740. 

Parmar, N., Warren, L. A., Roden, E. E. & Ferris, F. G. (2000). Solid phase capture of 
strontium by the iron reducing bacteria Shewanella alga strain BrY. Chem Geol 169, 
281-288. 

Pearson, A., Sessions, A. L., Edwards, K. J. & Hayes, J. M. (2004). Phylogenetically specific 
separation of rRNA from prokaryotes for isotopic analysis. Mar Chem 92, 295- 
306. 

Philip, L., Iyengar, L. & Venkobachar, C. (2000). Site of interaction of copper on Bacillus 
polymyxa. Water Air Soil Pollut 119,11-21. 

Puis, R. W., Powell, R. M., Clark, D. & Eldred, C. J. (1991). Effects of pH, solid/solution 

SGM symposium 65 



Biofilms and metal geochemistry 33 



ratio, ionic strength, and organic acids on Pb and Cd sorption on kaolinite. Water Air 
Soil Pollut 57-58, 423-430. 

Reysenbach, A.-L. & Shock, E. (2002). Merging genomes with geochemistry in hydro- 
thermal ecosystems. Science 296, 1 077-1 082. 

Richard, F. C. & Bourg, A C. M. (1991). Aqueous geochemistry of chromium: a review. 
Water Res 25, 807-816. 

Roberts, K. A., Santschi, P. H., Leppard, G. G. & West, M. M. (2004). Characterization of 
organic-rich colloids from surface and ground waters at the actinide-contaminated 
Rocky Flats Environmental Technology Site (RFETS), Colorado, USA. Colloids Surfaces 
A Physicochem Eng Aspects 244, 1 05-1 1 1 . 

Schorer, M. & Eisele, M. (1997). Accumulation of inorganic and organic pollutants by 
biofilms in the aquatic environment. Water Air Soil Pollut 99, 651-659. 

Singer, P. C. & Stumm, W. (1970). Acid mine drainage: the rate limiting step. Science 167, 
1121-1123. 

Smiejan, A., Wilkinson, K. J. & Rossier, C. (2003). Cd bioaccumulation by a freshwater 
bacterium, Rhodospirillum rubrum. Environ Sci Technol 37, 701-706. 

Southam, G. & Beveridge, T. J. (1992). Enumeration of thiobacilli from pH-neutral and 
acidic mine tailings and their role in the development of secondary mineral soil. Appl 
Environ Microbiol 58, 1 904-1 912. 

Spear, J. R., Walker, J. J., McCollom, T. M. & Pace, N. R. (2005). Hydrogen and bio- 
energetics in the Yellowstone geothermal ecosystem. Proc Natl Acad Sci USA 102, 
2555-2560. 

Stumm, W. (1 992). Chemistry of the Solid-Water Interface: Processes at the Mineral-Water 
and Particle-Water Interface in Natural Systems. New York: Wiley. 

Stumm, W. & Morgan, J. J. (1996). Aquatic Chemistry Chemical Equilibria and Rates in 
Natural Waters, 3rd edn. New York: Wiley. 

Takai, K., Gamo, T., Tsunogai, U., Nakayama, N., Hirayama, H., Nealson, K. H. & 
Horikoshi, K. (2004). Geochemical and microbiological evidence for a hydrogen- 
based, hyperthermophilic subsurface lithoautotrophic microbial ecosystem 
(HyperSLiME) beneath an active deep-sea hydrothermal field. Extremophiles 8, 
269-282. 

Tani, Y., Ohasi, M., Miyata, N., Seyama, H., Iwahori, K. & Soma, M. (2004a). Sorption of 
Co(ll), Ni(ll), and Zn(ll) on biogenic manganese oxides produced by a Mn -oxidizing 
f u ngus, strain KR2 1-2.7 Environ Sci Health Part A Toxic Hazard Subst Environ Eng 39, 
2641-2660. 

Tani, Y., Miyata, N., Osahi, M., Ohnuki, T., Seyama, H., Iwahori, K. & Soma, M. 
(2004b). Interaction of inorganic arsenic with biogenic manganese oxide produced 
by a Mn-oxidizing fungus, strain KR21 -2. Environ Sci Technol 38, 661 8-6624. 

Templeton, A. S., Trainor, T. P., Traina, S. J., Spormann, A. M. & Brown, G. E., Jr (2001). 
Pb(ll) distributions at biofilm-metal oxide surfaces. Proc Natl Acad Sci USA 98, 
11897-11902. 

Templeton, A. S., Trainor, T. P., Spormann, A. M. & Brown, G. E., Jr (2003a). Selenium 
speciation and partitioning within Burkholderia cepacea biofilms formed on a-AI 2 3 
surfaces. Geochim Cosmochim Acta 67, 3547-3557. 

Templeton, A. S., Trainor, T. P., Spormann, A. M., Newville, M., Sutton, S. R., 
Dohnalkova, A., Gorby, Y. & Brown, G. E., Jr (2003b). Sorption versus bio- 
mineralization of Pb(ll) within Burkholderia cepacea biofilms. Environ Sci Technol 37, 
300-307. 

Tessier, A. & Campbell, P. G. C. (1988). Comments on the testing of the accuracy of an 



SGM symposium 65 



34 L. A. Warren 



extraction procedure for determining the partitioning of trace metals in sediments. 

Anal Chem 60, 1475-1476. 
Tessier, A., Campbell, P. G. C. & Bisson, M. (1979). Sequential extraction procedure for the 

speciation of particulate trace metals. Anal Chem 51, 844-851. 
Unz, R. F. & Shuttleworth, K. L. (1 996). Microbial mobilization and immobilization of heavy 

metals. Curr Opin Biotechnol 7, 307-31 0. 
Urrutia, M. M. (1997). General bacterial sorption processes. In Biosorbents for Metal Ions, 

pp. 39-66. Edited by J. Wase & C. Forster. London: Taylor and Francis. 
Vigneault, B. ; Campbell, P. G. C, Tessier, A. & De Vitre, R. (2001). Geochemical changes 

in sulfidic mine tailings stored under a shallow water cover. Water Res 35, 1 066- 

1076. 
Warren, L. A. (2004). A special issue dedicated to microbial geochemistry. Geochim 

Cosmochim Acta 68, 3 1 39. 
Warren, L. A. & Ferris, F. G. (1998). Continuum between sorption and precipitation of Fe(lll) 

on microbial surfaces. Environ Sci Tech nol 32, 2331-2337. 
Warren, L. A. & Haack, E. A. (2001). Biogeochemical controls on metal behaviour in 

freshwater environments. Earth Sci Rev 54, 261-320. 
Warren, L A. & Kauff man, M. E. (2003). Microbial geoengineers. Science 299, 1 027- 

1029. 
Warren, L. A. & Zimmerman, A. P. (1994). The influence of temperature and NaCI on 

cadmium, copper and zinc partitioning among suspended particulate and dissolved 

phases in an urban river. Water Res 28, 1921-1931. 
Warren, L. A., Tessier, A. & Hare, L. (1998). Modelling cadmium accumulation by benthic 

invertebrates in situ: the relative contributions of sediment and overlying water 

reservoirs to organism cadmium concentrations. Limnol Oceanogr 43, 1442-1454. 
Zachara, J. M., Ainsworth, C. C, Cowan, C. E. & Resch, C. T. (1989). Adsorption of 

chromate by subsurface soil horizons. Soil Sci Soc Am J 53, 418-428. 



SGM symposium 65 



Minerals, mats, pearls and veils: 
themes and variations in giant 
sulfur bacteria 



Neil D. Gray and Ian M. Head 



School of Civil Engineering and Geosciences, Institute for Research on the Environment and 
Sustainability and Centre for Molecular Ecology, University of Newcastle, Newcastle upon Tyne, 
NE17RU, UK 



MINERALS, MATS, PEARLS AND VEILS: A PANOPLY OF GIANT 
SULFUR BACTERIA 

The biogeochemical cycling of sulfur has been at the heart of microbial ecology since 
the mid-19th century This is due, at least in part, to the striking forms of many of the 
organisms involved in the transformation of reduced sulfur species. Giant sulfur 
bacteria were among the earliest micro-organisms to capture the interest of micro- 
biologists exploring the links between geochemical cycling of the elements and the 
microbiota responsible. Consequently, giant sulfur bacteria were among the first 
bacteria described. Organisms resembling Beggiatoa ('Oscillatoria alba') were des- 
cribed as early as 1803 (Vaucher, 1803), but were included in the genus Beggiatoa some 
time later (Trevisan, 1842). Thiothrix (Rabenhorst, 1865; Winogradsky, 1888), Achrom- 
atium (Schewiakoff, 1893) and Tbioploca (Lauterborn, 1907) were all described by the 
early 20th century and Winogradsky (1887, 1888) had already formulated the principles 
of lithotrophic growth based on sulfide oxidation, from his work on Beggiatoa species. 
Surprisingly for such conspicuous organisms, novel giant sulfur bacteria are still being 
described (Guerrero et ai, 1999; Schulz et ai, 1999). 

Achromatium 

Bacteria of the genus Achromatium are remarkable. Cells of up to 125 ^im in length 
have been reported (Babenzien et al., 2005; Head et al., 2000a) and, in addition to 
characteristic sulfur globules that become visible on treatment of the cells with dilute 
acid, their large oval cells are typically filled with enormous inclusions of calcium 
carbonate (Fig. 1). As with many giant sulfur bacteria, they have as yet eluded 

SGM symposium 65: Micro-organisms and Earth systems - advances in geomicrobiology. 
Editors G. M. Gadd, K. T. Semple & H. M. Lappin-Scott. Cambridge University Press. ISBN 521 86222 1 ©SGM 2005 



36 N. D. Gray and I. M. Head 




Fig. 1. DAPI (4,6-diamidino-2-phenylindole)-stained preparation of a crudely purified suspension of 
Achromatium cells. The diatom cells illustrate the large size of the Achromatium cells. 



cultivation in the laboratory. Nevertheless, the ability to physically enrich cells from 
environmental samples (De Boer et aL, 1971; Head et aL, 2000a) has permitted the 
inference of some physiological properties (Gray et aL, 1997, 1999a, b, 2000, 2004). 
Achromatium species are capable of oxidizing reduced sulfur species completely to 
sulfate (Gray et al., 1997). This is a feature of bacteria capable of conserving energy 
from sulfur oxidation, in contrast to chemoheterotrophic sulfur bacteria that appear to 
utilize sulfide oxidation as a means to detoxify sulfide or neutralize reactive oxygen 
species produced during aerobic metabolism (Burton & Morita, 1964), although 
considerable ambiguity and debate still surround the precise role of sulfide oxidation in 
these bacteria (Strohl, 2005). Recent evidence also suggests that, although oxygen 
is probably the preferred electron acceptor for Achromatium species, nitrate is also 
probably used as an alternative electron acceptor (Gray et aL, 2004). The use of nitrate 
as a terminal electron acceptor is a trait shared with several other giant sulfur bacteria. 
Although the name of only one species of Achromatium has been validly published 
{Achromatium oxaliferum; Schewiakoff, 1893), it is clear from comparative analysis of 
16S rRNA gene sequences recovered from Achromatium cells physically enriched from 
sediment samples that several distinct species exist and, typically, several species coexist 
in a single sediment (Babenzien et aL, 2005; Gray et aL, 1999b; Head et aL, 1996). All 
Achromatium species investigated to date form a distinct monophyletic group within 



SGM symposium 65 



Ecology of giant sulfur bacteria 37 



1% 




Large marine Beggiatoa, Thiomargarita and Thioploo 



d Jhioploca ingrica and other freshwater Thbploca 
Thin, cultured marine Beggiatoa 



Thin, freshwater Beggiatoa 



' Gammaproteobacteha ' 




Sulfur-oxidizing symbionts 



Sulfur-oxidizing symbionts 
^^fl Thtomkmspira 

Thiothrix 

Achromatium 




Aliochromatium/Tb ermo en romatium 
Sulfur-oxidizing symbionts 

Addithiobatillus 



Thiovulum rnajus 



SutfuKttidiang symbionts ' Eps il onproteobact eria' 



Fig. 2. Phylogenetic tree of representatives of the sulfur-oxidizing bacteria from the 

' Gammaproteobacteria' and ' Epsilonproteobacteria' , including the giant sulfur bacteria from the 

genera Achromatium, Beggiatoa, Thiomargarita, Thbploca, Thiothrix and Thiovulum. 



the 'Gammaproteobacteria' (Fig. 2) and, in most analyses, appear to be related to 
anoxygenic, photosynthetic bacteria of the family Cbromatiaceae. However, the fifth 
release of the taxonomic outline of prokaryotes from the second edition of Bergeys 
Manual of Systematic Bacteriology lists Achromatium in the order 'Thiotricbales' as 
genus II within the family 'Tbiotrichaceae' (Garrity et al., 2005). 

Beggiatoa 

In contrast to Achromatium species, which are found as large populations of free-living 
unicells, Beggiatoa species often form extensive, conspicuous mats on the surface of 
sediments or associated with sulfur springs, hydrothermal vents or cold-seep environ- 
ments (Strohl, 2005; Teske &C Nelson, 2004). The mats can be extremely thin and 
compact (hundreds of micrometres) or much more diffuse and thick (centimetres), 
depending on the prevailing hydrodynamic conditions, which dictate whether substrate 
transport to the cells is diffusion-limited (Gunderson et al., 1992; Jorgensen & 
Revsbech, 1983). Indeed, there have been reports of B^gg/^^o^-dominated mats as thick 
as 30-60 cm in some marine hydrothermal systems associated with populations of tube 



SGM symposium 65 



38 N. D. Gray and I. M. Head 

worms (McHatton et al., 1996; Nelson et al., 1989). Several species have been obtained 
in pure culture, but the largest mat-forming Beggiatoa species discovered have yet to be 
grown in pure culture. Beggiatoa species are physiologically varied: some strains have 
been shown to be obligate chemolithoautotrophs, whilst others are facultative chemo- 
lithoautotrophs or mixotrophs (Hagen & Nelson, 1996; Nelson & Jannasch, 1983). 
Strains capable of autotrophic growth have been obtained from marine environments, 
whereas evidence to date suggests that freshwater isolates are primarily heterotrophs 
and evidence for lithoautotrophic growth of freshwater Beggiatoa isolates has been 
contentious (Strohl, 2005). Nevertheless, lithoautotrophic growth has been shown for 
more recently isolated freshwater Beggiatoa strains (Grabovich et ai, 2001; Patritskaya 
et al., 2001) and evidence for biochemical control of the switch from heterotrophic 
growth to mixotrophic or lithoautotrophic growth has been provided (Eprintsev et al., 
2003, 2004; Stepanova et al., 2002). The heterotrophic freshwater strains produce 
characteristically thin filaments, as do cultivated marine strains capable of autotrophic 
growth. Neither of these thin-filament types produces large vacuoles and both have 
been cultivated in the laboratory. Beggiatoa cells with wide filaments (up to 200 ^im in 
diameter) have yet to be grown in laboratory culture and are characterized by large 
vacuoles that are believed to play a role in storing nitrate, which is used as an oxidant in 
these enormous bacteria. However, recently studied vacuolated filaments of giant 
bacteria that resemble Thiotbrix species morphologically, but are related most closely 
to Beggiatoa species, do not appear to store nitrate, suggesting that the role of large 
intracellular vacuoles may be more complex than simply providing a repository for 
large amounts of oxidant (Kalanetra et al., 2004). Most large-vacuolated Beggiatoa 
species accumulate high concentrations of nitrate (Kalanetra et al., 2004; McHatton 
et al., 1996; Mufimann et al., 2003), which is used as an oxidant either by reduction to 
ammonia or through denitrification, thus providing a link between the biogeochemical 
cycles of nitrogen and sulfur (Teske &C Nelson, 2004). Like Achromatium species, 
Beggiatoa species belong to the "Gammaproteohacteria\ where they form a mono- 
phyletic group with Thioploca and Thiomargarita species (Fig. 2; Kalanetra et al., 2004; 
Kojima et al., 2003 ; Mu&nann et al., 2003 ; Teske et al., 1996) . The uncultured vacuolate 
bacteria {Beggiatoa, Thioploca and Thiomargarita species) themselves form a distinct 
phylogenetic group within this, with freshwater and marine thin-filament Beggiatoa 
species each occupying distinct lineages (Teske 6c Nelson, 2004). It is also clear that 
there is considerable diversity within communities of Beggiatoa, which often exhibit 
filaments with different diameters (Nelson et al., 1989). 



Thioploca 

In many respects, Thioploca species resemble Beggiatoa species, especially the large- 
vacuolated Beggiatoa found in marine environments. Members of the genus Thioploca 



SGM symposium 65 



Ecology of giant sulfur bacteria 39 

are differentiated from those of the genus Beggiatoa on the basis that trichomes of 
Thioploca species are usually enclosed in a thick polysaccharide sheath. Whilst some 
Beggiatoa strains have been isolated in pure culture, no Thioploca strains have been 
grown axenically in the laboratory. Like Beggiatoa strains, they form multicellular 
filaments that can be several centimetres long. Several morphotypes can be differ- 
entiated on the basis of filament diameter (Schulz et al., 1996) and these correspond to 
different species, defined on the basis of 16S rRNA gene sequence identity (Teske et al., 
1996). Moreover, most Thioploca species described to date, in common with large 
Beggiatoa species, harbour vacuoles that are presumed to store nitrate, which is used as 
an oxidant by both Beggiatoa and Thioploca species (McHatton et al., 1996). These 
occur even in Thioploca species with relatively thin (2-5 ^m diameter) filaments 
(Zemskaya et al., 2001), whereas thin Beggiatoa species apparently do not have large 
vacuoles (Teske & Nelson, 2004). rRNA gene sequence-based analyses of Thioploca 
isolates recovered from naturally occurring mats have shown that they are related 
closely to, and form a monophyletic group with, the large-vacuolated Beggiatoa species 
and members of the genus Thiomargarita, which also exhibit a large internal vacuole 
implicated in nitrate storage (Mufimann et al., 2003; Teske et al., 1996, 1999). Given the 
similarity with Beggiatoa and the fact that, under certain circumstances, Thioploca 
trichomes leave their sheaths and are then morphologically indistinguishable from 
Beggiatoa filaments, it is likely that the classification of these two very similar genera 
will be reconciled as more phylogenetic and physiological information accumulates. 
Extensive mats of Thioploca cells have been discovered worldwide in both marine and 
freshwater environments (Teske & Nelson, 2004) and Jorgensen & Gallardo (1999) 
have suggested that the discontinuous occurrence of Thioploca mats that extend for a 
distance of over 3000 km along the west coast of South America probably represent the 
largest communities of sulfur bacteria on the planet. 



Mats of Thioploca cells differ in structure from those composed of Beggiatoa cells. 
This is principally a consequence of the presence of multiple Thioploca trichomes 
within an extensive sheath that can penetrate several centimetres below the sediment 
surface (Schulz et al., 1996). This innovation, coupled with the intracellular storage 
of large amounts of nitrate, has suggested a novel ecological strategy whereby Thio- 
ploca species exploit pools of oxidant and reductant that are separated in space. 
Thioploca trichomes exhibit a tactic response towards nitrate and, when nitrate 
concentrations are high in the overlying water, the trichomes extend into the water 
column, where it is believed that they accumulate nitrate within their extensive system 
of vacuoles. They then retreat back into their sheaths, carrying their store of oxidant to 
greater depth in the sediment, where electron donor in the form of sulfide is freely 
available either from sulfate reduction or geothermal sources (Huettel et al., 1996). 

SGM symposium 65 



40 N. D. Gray and I. M. Head 





Fig. 3. Mats and streamers of filamentous sulfur bacteria from springs on Sulfur Mountain, CA, USA. 

Thiothrix 

Conspicuous streamers of Thiothrix cells are often noticeable in flowing waters from 
sulfur springs (Fig. 3). In this respect, they differ from the sulfur bacteria considered 
above. Beggiatoa species, in particular, rely on molecular diffusion through a stagnant 
boundary layer to provide electron donor and acceptor to the cells (Schulz & Jorgensen, 
2001). By virtue of specialized holdfast cells, members of the genus Thiothrix can 
maintain themselves even in vigorously flowing waters that would wash away 
filamentous sulfur bacteria lacking this adaptation. The rapid fluid flow decreases the 



SGM symposium 65 



Ecology of giant sulfur bacteria 41 





Fig. 4. Rosettes of Thiothrix sp. from a sulfur spring, Gilsland Spa, Northumberland, UK. 

unmixed boundary layer, freeing Thiothrix cells from the limitations of molecular 
diffusion. It has long been considered that defining features of members of the genus 
Thiothrix are their ability to deposit intracellular sulfur and their growth in character- 
istic rosette structures that emanate long filaments (Fig. 4; Larkin, 1989). This has, 
however, been called into question by a number of observations. The genus Thiothrix 
includes organisms with highly variable morphology, such as Thiothrix eikelboomii 
and Thiothrix defluvii, that do not necessarily produce rosettes (Howarth et aL, 1999; 
Unz & Head, 2005). Furthermore, there is evidence that Leucothrix species, which also 



SGM symposium 65 



42 N.D.Gray and I. M. Head 

produce rosettes and were until recently considered as heterotrophs, may deposit intra- 
cellular sulfur when growing lithoheterotrophically (Grabovich et al., 2002). In addi- 
tion, recently characterized bacteria that are morphologically reminiscent of Thiothrix 
species have been shown to be related more closely to vacuolated Beggiatoa and 
Thioploca strains (Kalanetra et al., 2004). Spectacular accumulations of Thiothrix cells 
have been reported from sulfur springs (Brigmon et al., 2003; Larkin & Strohl, 1983; 
McGlannan & Makemson, 1990) and in a range of cave environments (Brigmon et al., 
1994; Engel et al., 2004), associated with methane seeps (Pimenov et al., 2000) and, 
more recently, they have been reported as an important component of novel prokaryotic 
communities with a string-of-pearls morphology (Moissl et al., 2002). 

Although not related closely to the genera Beggiatoa or Thioploca, Thiothrix species 
also belong to the ' ' Gammaproteobacteria' and this group is replete with morpho- 
logically distinctive sulfur bacteria (Fig. 2). 

Thiomargarita 

Thiomargarita cells are enormous. The bacterium produces large, spherical cells, and 
cells in excess of 700 ^m in diameter have been reported (Schulz et al., 1999). The large, 
spherical cells reside in a mucus sheath that links the otherwise unconnected cells in a 
chain. Despite the huge size of Thiomargarita cells, the bacterium was discovered only 
recently (Schulz & Jorgensen, 2005; Schulz et al., 1999). This may reflect the habitat 
occupied by the bacterium. Thiomargarita isolates were first reported from seafloor 
sediments at a depth of approximately 100 m, off the coast of south-western Africa. 
Like Thioploca-dom'mated sediments off the western seaboard of South America, the 
overlying water has low oxygen tension and relatively high nitrate concentrations, 
which are a consequence of the very high productivity in these regions that are 
characterized by upwelling of nutrient-rich deep waters. Like the large, marine 
Beggiatoa and Thioploca species, Thiomargarita cells have large, central vacuoles that 
constitute approximately 98 % of the biovolume of the cells, and accumulate nitrate to 
levels of hundreds of millimolar (Schulz et al., 1999). Thiomargarita species do not 
exhibit the gliding motility that is characteristic of Beggiatoa and Thioploca species 
and rely instead on being transported passively into the water column by processes such 
as gas venting, storms and wave action. This permits them to access their sources of 
oxidant (oxygen and nitrate), which can be used to oxidize stored elemental sulfur or 
sulfide when the cells are deposited back into the organic-rich sulfidic sediment that 
they normally inhabit. The kinds of event that transport Thiomargarita cells passively 
into the water column are innately sporadic and this may explain their huge size, which 
allows them to store large quantities of oxidant that may be required for long-term 
survival between resuspension events. In addition, they are tolerant of high oxygen 
concentrations, whereas Beggiatoa and Thioploca species are microaerophilic and 

SGM symposium 65 



Ecology of giant sulfur bacteria 43 

have a phobic response to high oxygen concentrations (Huettel et al., 1996). Empirical 
observations suggest that Thiomargarita cells may survive in the absence of nitrate for 
over 700 days (Schulz & Jorgensen, 2001). 

This large-vacuolated bacterium is phylogenetically related most closely to the vacuo- 
lated Beggiatoa and Thioploca species and, based on comparative 16S rRNA gene 
sequence analysis, forms a monophyletic group with them in the 'Gammaproteo- 
bacteria' (Fig. 2; Schulz etal., 1999). 

Thiovulum 

In 1913, Hinze described two species of Thiovulum (Hinze, 1913; Starr & Skerman, 
1965): Thiovulum majus (the type species) and 'Thiovulum minus'. These were identi- 
fied in marine environments and all isolated Thiovulum species to date come from 
marine habitats. Although perhaps not in the same realm as the microbial giants 
described above [Thiovulum cells range from 5 to 25 \im in diameter; Fenchel, 1994), 
conspicuous 'veils' of Thiovulum cells are produced, associated with gradients of 
sulfide and oxygen (Jorgensen & Revsbech, 1983). Perhaps their most remarkable 
feature is their rapid motility, which has been measured at over 600 ^m (24-60 cell 
lengths) s . This is equivalent to a 2 m fish swimming at up to >400 km h _1 (60 body 
lengths s _1 ). Their rapid swimming is facilitated by extensive flagellation over the 
surface of the cells and they use this either to propel the multicellular floating veils that 
they form into the optimum position in opposed gradients of sulfide and oxygen 
(Jorgensen & Revsbech, 1983) or to direct water flow around tethered veils to reduce 
diffusional limitation of substrate transport to the cells (Fenchel & Glud, 1998). It is 
now clear that a number of other bacteria have adopted similar strategies to enhance 
transport of substrates to the cells (Thar & Kuril, 2002). 

Apart from aspects of their chemosensory behaviour, motility and ultrastructure, there 
are few studies on the physiology of Thiovulum species and only one partial 16S rRNA 
gene sequence from authenticated Thiovulum cells has been deposited in the public 
databases. The one Thiovulum species that has been characterized phylogenetically 
to date belongs to the 'Epsilonproteobacteria' (Fig. 2; Lane et al., 1992; Romaniuk et 
al., 1987), although it seems likely that, with more extensive sampling, the picture of 
the taxonomy of the genus Thiovulum will change. 

GIANT SULFUR BACTERIA: ENERGY AND GEOCHEMICAL 
SIGNIFICANCE 

Bacteria involved in the oxidative side of the sulfur cycle couple the oxidation of 
reduced sulfur species to the reduction of dissolved oxygen, nitrate and possibly 
oxidized metals. Many generate energy from these processes and reduce C0 2 for 

SGM symposium 65 



44 N.D.Gray and I. M. Head 

biosynthesis. Some also have the capacity to fix dinitrogen gas and Beggiatoa species 
have been best studied in this respect (Nelson et ai, 1982; Polman & Larkin, 1988). 
Sulfur bacteria therefore play a pivotal role in the biogeochemical cycling of sulfur, 
carbon, nitrogen and possibly metals. Their activities also affect the precipitation and 
dissolution of minerals and can have consequences for higher trophic levels in a variety 
of ecosystems. 

The reduced sulfur that provides the energy to drive the biogeochemical activities of 
sulfur bacteria can come from a variety of sources. Reduced sulfur can be liberated 
from decaying organic matter, be produced from dissimilatory sulfate reduction or 
released as a result of volcanic activity or from other geological sources. Environments 
characterized by different sources of reduced sulfur present different geochemical 
constraints on electron-donor and -acceptor availability, which have led to the numer- 
ous ecological adaptations observed in sulfur-oxidizing bacteria. One major ecological 
adaptation has been the evolution of sulfur bacteria with a large cell size. 

The size of giant sulfur bacteria not only represents a remarkable evolutionary adap- 
tation to the lifestyles that they have adopted, but also permits them to be manipulated 
experimentally under in situ conditions and used as models to establish links between 
environment, physiology and evolutionary diversification that may be extrapolated 
more widely Some of the adaptations in the energetics and biochemistry of giant sulfur 
bacteria have consequences for ecosystem-scale processes. These are discussed in 
the context of the geochemical features of habitats dominated by conspicuous 
sulfur bacteria. 

Energetic considerations of sulfur oxidation 

The transformation of S"~to SOj - by bacteria requires the transfer of eight electrons to 
an appropriate electron acceptor. What is known or assumed about electron transport 
in giant sulfur bacteria has been gleaned both directly (Cannon et ai, 1979; Schmidt & 
DiSpirito, 1990; Strohl et ai, 1986) and by analogy with mechanisms observed in other 
sulfur bacteria. What therefore follows, by necessity, does not represent a compre- 
hensive picture of sulfur-oxidation pathways, but focuses on those areas where direct or 
circumstantial evidence has been obtained from studies of giant sulfur bacteria. 
Readers seeking a comprehensive review of sulfur-oxidation pathways are directed to 
some excellent reviews (Friedrich et ai, 2001; Kappler &C Dahl, 2001; Kelly, 1999; Kelly 
et ai, 1997). What these studies indicate is that there is no universal pathway of sulfur 
oxidation and the giant sulfur bacteria, like all sulfur oxidizers, display considerable 
variation in pathways of dissimilatory sulfur metabolism, even between closely related 
species. For instance, the oxidation of hydrogen sulfide to sulfur, the first step in sulfur 
oxidation, is linked either to the reduction of a c-type cytochrome, mediated by 

SGM symposium 65 



Ecology of giant sulfur bacteria 45 

flavocytochrome c-sulfide dehydrogenase, or reduction of a quinone, mediated by 
sulfide-quinone reductase, in phototrophic and lithotrophic bacteria (Friedrich et al., 
2001). However, very little information is available on how these processes are mediated 
in the giant sulfur bacteria. Some studies have shown that the conversion of sulfide to 
sulfur in freshwater Beggiatoa alba cells was not prevented by dibromothymoquinone, 
an inhibitor of ubiquinone reduction, whereas inhibitors of flavoproteins (e.g. thenoyl- 
trifluoroacetone) did suppress sulfide oxidation (Schmidt et al., 1987). This suggests 
that electrons from sulfide enter the electron-transport chain via a flavocytochrome in 
Beggiatoa alba (Schmidt et al., 1987). How typical this is of other giant sulfur bacteria is 
not clear and Beggiatoa alba is not considered to be capable of chemolithotrophic 
growth. This first step in the oxidation process is critical, because the sulfur produced 
can then be either oxidized further or stored by precipitation as intracellular inclusions. 
Although a characteristic of all giant sulfur bacteria is the deposition of intracellular 
sulfur, a central problem still to be resolved in sulfur oxidation by chemolithotrophs is 
the means by which zero-valent sulfur is transformed to SO'f and whether energy is 
conserved from this step (Friedrich et al., 2001; Kelly et al., 1997). It has been proposed 
that the process may involve oxidation of S° to sulfite by a sulfur oxygenase; sulfite 
generated may then react with elemental sulfur to generate thiosulfate, which is 
observed as a product in vitro (Kelly, 1999). The requirement for molecular oxygen by a 
sulfur oxygenase means that this mechanism would not allow energy conservation and 
would not be feasible during anaerobic sulfur oxidation by nitrate-reducing Tbioploca, 
Beggiatoa or Tbiomargarita species, for example. Electrons from Tbiobacillus denitrifi- 
cans growing with nitrate as an electron acceptor in the absence of oxygen have been 
shown to be linked to a respiratory chain (Kelly, 1999) and there is evidence that, even in 
aerobic sulfur oxidizers, energy is conserved during elemental sulfur oxidation and does 
not involve a non-energy-conserving oxygenase. It is clear that several mechanisms may 
be involved in elemental sulfur oxidation and the details may be different in different 
bacterial taxa (Kelly, 1999). The final step in the sulfide-oxidation process is the 
transformation of sulfite to sulfate and it is this section of the sulfur-oxidation pathway 
for which most data are available in the giant sulfur bacteria. 

Sulfite oxidation to sulfate is known to occur by three different pathways: one direct 
and two indirect (Fig. 5; Kappler & Dahl, 2001). The first of the indirect pathways is in 
essence a reversal of sulfate reduction: sulfite is oxidized to sulfate via the formation of 
adenosine 5'-phosphosulfate (APS), mediated by APS reductase, and subsequent 
decomposition by reaction with pyrophosphate to form ATP and sulfate (catalysed 
by ATP sulfurylase; Fig. 5, pathway Ha). Alternatively, APS and orthophosphate may 
be converted to ADP by ADP sulfurylase (adenylylsulfate : phosphate andenylyl- 
transf erase), with the ADP being converted to ATP by adenylate kinase (Fig. 5, pathway 
lib). Both indirect pathways therefore produce a molecule of ATP by substrate-level 

SGM symposium 65 



46 N. D. Gray and I. M. Head 



Sulfide 



2e 



A 



Sulfur 



4e- 




3 



03 
Q. 



l 



Sulfite : acceptor 

oxidoreductase 



2e- 




Sulfite 



AMP || 



y/ APS re 



reductase 



PR 
ATP 

sulfurylase 
ATP 



r* 



APS 




Sulfate 



lib 



Adenylylsulfate: 

phosphate 
ad enylyltransf erase 



Adenylate A ^-ADP 
kinase y 
ATP + AMP 



3 

a. 

a 

I 



Fig. 5. Some of the alternative pathways of sulfur oxidation thought to be predominant in giant 
sulfur bacteria. The pathways for oxidation of sulfide through sulfur to sulfite and the involvement of 
intermediates, such as thiosulfate and tetrathionate, have not been elucidated in giant sulfur bacteria 
and a minimal representation of oxidation of sulfide to sulfite is provided. This shows the number of 
electrons generated at each step with potential to contribute to oxidative phosphorylation, but does 
not represent the full range of reactions that may contribute to sulfide and sulfur oxidation. 



phosphorylation. The direct mechanism of sulfite oxidation involves the enzyme 
sulfite : acceptor oxidoreductase, which oxidizes sulfite to sulfate without the need for 
AMP (Fig. 5, pathway I). This pathway only generates ATP by oxidative phosphory- 
lation. Many sulfur bacteria carry both enzyme systems. However, whilst sulfite : 
acceptor oxidoreductase is widespread among sulfur-oxidizing lithoautotrophs, the 
APS pathway is less prevalent (Kappler & Dahl, 2001). Nevertheless, some Beggiatoa 
strains exhibit both a direct and an indirect pathway (Hagen & Nelson, 1997) and the 
APS reductase pathway occurs in sulfur-oxidizing symbionts of invertebrates and APS 
reductase genes have been detected in Achromatium species (Head et al., 2000b; Nelson 
& Hagen, 1995). In contrast, some Beggiatoa isolates do not exhibit APS reductase 
activity (Grabovich et al., 1998, 2001; Hagen 6c Nelson, 1997). 



SGM symposium 65 



Ecology of giant sulfur bacteria 47 

However, there is potential for substrate-level phosphorylation in some sulfur-oxidizing 
pathways and, indeed, in some Thiothrix species, substrate-level phosphorylation is 
believed to be the sole mechanism of energy conservation (Grabovich et ai, 1999; 
Odintsova & Dubinina, 1993). 

The thermodynamics and, hence, the energy yield of these oxidative processes are 
dependent upon the particular electron acceptor used, the chemistry of the reduced 
sulfur compound being oxidized and the reduced and oxidized products of the reaction. 
Thermodynamic and kinetic considerations of sulfur oxidation have considerable bear- 
ing on the ecology and physiology of the sulfur bacteria. For instance, the energy yield 
from the oxidation of 1 mol HS - by 2 mol 2 to sulfate ostensibly yields -732-6 kj 
(mol sulfur) -1 (Kelly, 1999). Whilst this may be important for giant sulfur bacteria, 
many that are adapted to low oxygen concentrations may use nitrate as an alternative 
electron acceptor. There is good evidence that the large-vacuolated Beggiatoa and 
Thioploca species reduce nitrate to ammonium rather than dinitrogen gas, which only 
accounted for approximately 15 % of the nitrate reduced (Otte et ai, 1999), and there 
has been some debate regarding the validity of some earlier suggestions that Beggiatoa 
species are capable of denitrification (Fossing et ai, 1995; McHatton et al., 1996; 
Sweerts et al., 1990). Based on the free energy of formation values used by Kelly (1999), 
it is possible to calculate the energy yield from the oxidation of sulfide coupled to 
nitrate reduction to N 2 according to the equation 

5HS-+8N03+3H + ^5SOf +4N 2 +4H 2 

This yields approximately the same amount of energy [-744-76 kj (mol sulfur) -1 ] as 
sulfide oxidation with oxygen [-732-6 kj (mol sulfur) -1 ]. Note that the energy yield 
from reduction of nitrate to dinitrogen gas or ammonium is considerably greater than 
that from reduction of nitrate to nitrite. It should also be noted that these values were 
calculated under standard conditions and are useful for comparative purposes, but 
probably do not reflect the energy yield under in situ conditions. 



In contrast, the reaction 

HS - +N03 + H + +H 2 O^NHj+SOj - 

yields -463-8 kj (mol sulfur) -1 . On the face of it, one might conclude that these free- 
energy values make the dissimilatory reduction of nitrate to ammonium (nitrate 
ammonification or DNRA) less favourable, but several giant sulfur bacteria adopt this 
way of life and reduction of NOT to NH4 is quantitatively important in many organic- 
rich coastal marine sediments (Sorensen, 1978). In these sediments, whilst reduction of 
nitrate to N 2 and consumption of 2 are restricted to the upper few millimetres 
of sediment, the reduction of NO3 to NH|is significant at greater depth. The enzymol- 
ogy of the two pathways of nitrate reduction is, however, quite different, as shown by 

SGM symposium 65 



48 N. D. Gray and I. M. Head 

acetylene treatment of sediment microcosms. Acetylene prevents denitrification, but 
DNRA is unaffected (Bonin et ai, 1998). Interestingly, the relative balance of these two 
pathways has a significant effect on the sedimentary nitrogen cycle. This is because 
denitrification produces nitrogen gas, which is biologically and chemically inert; in 
contrast, DNRA produces ammonium that, under suitable conditions, can be reoxi- 
dized to nitrate. Thus, biologically available nitrogen is more likely to be retained in 
sediments where the majority of nitrate is reduced to ammonium (Sorensen, 1978). 
Whilst this is a useful outcome with respect to the bacteria that utilize nitrate, it does 
not in itself explain why, given the higher energy yield, nitrate is not always reduced 
preferentially to N ? . A number of studies have indicated that, in high-sulfide environ- 
ments, DNRA coupled to sulfide oxidation is favoured because of the inhibition 
of denitrification at high sulfide concentrations (e.g. Brunet & Garcia-Gil, 1996). 
However, it is also noteworthy that, in reduced environments, because of the stoichio- 
metry of the two reactions, DNRA is probably no less thermodynamically favourable 
than denitrification. In reducing marine sediments, sulfide is unlikely to be limiting, 
whereas nitrate is likely to be less abundant. Only 1 mol nitrate is required to oxidize 
1 mol sulfide to sulfate when nitrate is reduced to ammonium (an eight-electron 
transfer), whereas 1-6 mol nitrate is required to oxidize 1 mol sulfide to sulfate when the 
nitrate is reduced to N 7 (a five-electron transfer). Despite a more favourable energy yield 
for denitrification than for DNRA when the values are expressed as kj (mol sulfur 
oxidized) -1 , the energy yield in kj (mol nitrate reduced) -1 is broadly the same for the 
two reactions, i.e. -465-5 kj (mol nitrate) -1 for denitrification and -463-8 kj (mol 
nitrate) -1 for DNRA. From this different perspective, if nitrate is limiting compared 
with the availability of electron donor, it would appear that the two processes are 
equally energetically favourable. 

However, the energy yield of electron transport is not dictated solely by thermo- 
dynamics. The biology of the system may also have significance regarding the ability to 
realize the maximal thermodynamic yield. To generate ATP by oxidative phosphory- 
lation, it is necessary to generate a proton-motive force. This depends on the ability of 
each step in the electron-transport pathway to export protons efficiently across the cell 
membrane. The efficiency is dependent on the mode of respiration employed. The 
aerobic electron-transport chain, because of its topography, is more efficient at 
converting the chemical energy of electron donors into ATP (Berks et al., 1995). This is 
because cytochrome aa 3 is not only a proton pump, but it also delivers electrons to the 
cytoplasmic side of the membrane, where they combine with oxygen and protons and 
hence create a greater electrochemical gradient. In contrast, in denitrification, at least in 
some organisms, the enzymes nitrite reductase, nitric oxide reductase and nitrous oxide 
reductase are located on the periplasmic side of the membrane, the consumption of 
protons occurs externally and the net gain in translocated protons is lower than might 

SGM symposium 65 



Ecology of giant sulfur bacteria 49 

be suggested solely by the reduction potentials of the respective electron acceptors 
(Berks et al., 1995). Thus, despite similar thermodynamics, different respiratory pro- 
cesses can have very different energetic efficiency in terms of ATP formation. In the 
process of DNRA, after the reduction of nitrate to nitrite, nitrite is reduced to 
ammonium without the release of intermediates (Simon, 2002). In this process, nitrite is 
reduced by a cytochrome c-nitrite reductase complex, which mediates the multi- 
electron and multi-proton reduction of nitrite to ammonium. It is clear that this 
process, when coupled to sulfide oxidation, produces ATP (Simon, 2002). In the small 
number of cases studied to date, nitrite reduction to ammonium occurs on the outside 
of the cell membrane where protons are consumed, reducing the effective proton- 
motive force as in denitrification (Simon, 2002). 

Geochemical significance of giant sulfur bacteria 

The metabolic diversity of giant sulfur bacteria and the huge abundances that they can 
achieve in appropriate environments mean that they play a significant role in a range of 
biogeochemical processes. A further consequence of the activity of sulfur-oxidizing 
bacteria is that they may have pronounced effects at higher trophic levels. The various 
giant sulfur bacteria oxidize reduced sulfur species, deposit elemental sulfur, fix 
inorganic carbon, utilize organic compounds as electron donors or carbon sources, 
consume oxygen and reduce nitrate and nitrite to ammonium and perhaps, to a lesser 
extent, gaseous products. Several are capable of fixing dinitrogen gas and some may 
mobilize insoluble metals and other minerals. Giant sulfur bacteria can therefore 
have important implications for the cycling of many elements. 

Giant sulfur bacteria and the sulfur cycle. The oxidation of sulfide to elemental 
sulfur and ultimately sulfate by giant sulfur bacteria clearly has considerable 
consequences for sulfur cycling, especially where extensive mats of bacteria occur. The 
sediments off the Pacific coast of South America, which harbour extensive mats of 
Thioploca cells, exhibit high rates of sulfate reduction (up to 1500 nmol cm -3 day -1 ); 
however, sulfide rarely accumulates above a few tens of micromolar (Ferdelman et al., 
1997; Thamdrup 6c Canfield, 1996). The low concentrations of sulfide can be explained 
to some extent by reaction of sulfide with reactive iron, which is considered an impor- 
tant sink for sulfide in the Pacific coast sediments (Thamdrup 6c Canfield, 1996). 
Nevertheless, it has been estimated that anywhere between 16 and 91 % of sulfide 
reoxidation can be attributed to mats of Thioploca cells (Ferdelman et al., 1997; Otte 
et al., 1999). This has the dual role of replenishing an important oxidant for anaerobic 
carbon mineralization (sulfate) and reducing the steady-state concentration of toxic 
sulfide. Sediments off the Namibian coast also exhibit low levels of dissolved sulfide 
(Briichert et al., 2003). This is particularly unexpected, because of the relatively low 
levels of reactive iron species in Namibian coastal sediments (Briichert et al., 2000; 

SGM symposium 65 



50 N. D. Gray and I. M. Head 

Morse 6c Emeis, 1990). The effective removal of sulfide in this environment is therefore 
due, in large part, to the effective reoxidation of sulfide by the mats of Tbiomargarita 
cells. In this case, the giant sulfur bacteria could account for up to 55 % of the reoxi- 
dation of sulfide and a positive correlation was found between the measured sulfide flux 
from the sediment and the size of the population of Tbiomargarita cells (Brtichert et al., 
2003). Furthermore, considering that Tbiomargarita cells may contain up to 1-7 M 
sulfur in their cytoplasm and that biomass densities as high as 176 g m -2 may occur off 
the coast of south-west Africa (Brtichert et al., 2003), they represent a major repository 
of elemental sulfur in these sediments. For example, the biovolume of Tbiomargarita 
cells in the top few centimetres of Namibian coastal sediments was approximately 
4 ^1 mm (Schulz et al., 1999). Given that only 2 % of this volume is cytoplasm, this 
gives a value of 0-08 ,ul cytoplasm per cubic millimetre of sediment, which will contain 
up to 4-35 \xg sulfur. Converting this to an areal basis, this equates to approximately 
4-35 g sulfur m -2 contributed by the Tbiomargarita cells. Interestingly, measurements of 
intracellular and extracellular elemental sulfur in Tbioploca-dom'mated sediments 
off the coast of Chile showed that extracellular sulfur was abundant and had a different 
depth distribution from intracellular sulfur, suggesting that chemical oxidation also 
played an important role in sulfide oxidation in these sediments (Zopfi et al., 2001). 
This is in agreement with the conclusions drawn based upon the relative availability of 
reactive iron in Namibian and Chilean coastal sediments (Brtichert et al., 2003). 

In freshwater sediments, the concentration of sulfate is orders of magnitude lower than 
that in marine systems and it is rarely possible to measure dissolved sulfide in these 
environments. Consequently, fewer studies of sulfur cycling have been conducted in 
freshwater sediments. In many circumstances where there are extensive accumulations 
of sulfur bacteria in freshwater environments, this may be due to locally high levels of 
sulfide associated with decaying organic matter, rather than from sulfate reduction. 
Nevertheless, even when no dissolved sulfide is detectable, large populations of giant 
sulfur bacteria may be supported (Head et al., 1996, 2000b). These may have greater 
significance for biogeochemical cycling in freshwater systems than the sulfate concen- 
tration would suggest. Achromatium species are extremely effective at reoxidizing 
reduced sulfur generated by sulfate reduction (Gray et al., 1997). This is the case even in 
environments with extremely high levels of reactive iron, which would be expected to 
effectively outcompete Achromatium species for the low levels of sulfide produced by 
sulfate reduction (Head et al., 2000b). This not only has consequences for the sulfur 
bacteria themselves, but also suggests that a greater fraction of organic carbon mineral- 
ization in Achromatium-btznng sediments is channelled through sulfate reduction 
than would be predicted from measurements of sulfate reduction in incubations 
performed under anoxic conditions (Gray et al., 1997; Head et al., 2000a, b). It has been 
suggested that, for sulfur bacteria to thrive under conditions of low sulfide and high 

SGM symposium 65 



Ecology of giant sulfur bacteria 51 

reactive-iron concentrations, they must compete effectively with the rapid chemical 
reaction between sulfide and iron (Thamdrup 6c Canfield, 1996). In the case of Thio- 
ploca cells in marine sediments, this may be achieved through close physical association 
with the sulfate-reducing bacteria that generate the sulfide (Thamdrup 6c Canfield, 
1996). However, this does not seem to be the case for Achromatium species, where other 
bacterial cells have not been found associated with Achromatium cells. Achromatium 
may therefore possess high-affinity sulfide-uptake systems that are capable of compet- 
ing effectively with chemical reactions with iron to permit the utilization of low levels 
of dissolved sulfide. The kinetics of sulfide oxidation by Achromatium species have yet 
to be studied in detail. 

On the other hand, Achromatium cells may not use dissolved sulfide directly, but 
may instead use solid-phase iron sulfides as their principal source of electron donor 
(Gray et al., 1997; Head et al., 2000a, b). If this is the case, the precipitation of intra- 
cellular calcium carbonate by these bacteria may provide a means by which they could 
dissolve solid-phase iron sulfides and thus drive reoxidation of sulfur at greater rates 
than would be predicted from very low free-sulfide concentrations. Several explanations 
for calcite deposition in Achromatium species, i.e. neutralization of acid formed by 
sulfur oxidation, regulation of cell buoyancy or maintenance of high C0 2 partial 
pressures to facilitate autotrophic growth, have been proposed (Head et al., 2000b). 
However, this physiological feature, which is unique to Achromatium species in the 
bacterial domain, suggests that it represents an adaptation to an ecophysiological 
challenge not faced by other organisms. 

Calcite precipitation linked to sulfide dissolution may therefore be an ability unique to 
members of the genus Achromatium. Protons exported during chemiosmotic energy 
generation are consumed in the mineral dissolution reaction 

FeS + H + ^Fe 2+ + HS" 

and the sulfide generated can then be oxidized to sulfate: 

HS-+20 2 -HS04 

As some of the exported protons will be consumed by their reaction with metal 
sulfides, they will be unavailable for transport across the cell membrane by ATP 
synthase. An inevitable consequence is that a surplus of hydroxyl ions will occur on the 
cytoplasmic side of the cell membrane. This is exacerbated if nitrate is used as an 
electron acceptor, as sulfide oxidation with nitrate consumes protons: 

HS-+N03 + H + +H 2 0-*NH++SOj- 

5HS-+8N03+3H + -5SOf +4N 2 + H 2 

SGM symposium 65 



52 N.D.Gray and I. M. Head 



Mineral surface 



FeS 



\ 



Periplasm 
Ca 2+ 



HCO^ 



H + «- 




Fe 2+ + HS 



Cytoplasm 



H 2 + COJ 



t 



+ HC0 3 + OH 



t 

OH 




H 2 



HS" + 20- 



. 



hso; 



Calcite inclusion 



-► Ca 2+ 

+ 



CO 



3 



1 



CaCO, 



Fig. 6. Putative mechanism for the coupling of intracellular calcite precipitation and sulfide mineral 
dissolution and oxidation in Achromatium species. 

A potential sink for excess hydroxyl ions is through buffering with bicarbonate and the 
formation and precipitation of calcite: 

OH-+HCOj-*CO^ + H 2 

Ca 2+ +C05--CaC0 3 (s) 

The coccolithophorid algae have previously been proposed to buffer increases in 
cytoplasmic pH via intracellular precipitation of calcite (Borowitzka, 1982). These 
photosynthetic organisms deposit intracellular calcite to maintain an high internal 
partial pressure of C0 2 to facilitate carbon fixation by ribulose-l,5-bisphosphate 
carboxylase/oxygenase (RuBisCO) (Borowitzka, 1982). Under conditions of low 
dissolved C0 7 and high dissolved oxygen, bicarbonate is converted to C0 1 and 
hydroxyl ions by carbonic anhydrase (Borowitzka, 1982). The subsequent fixation of 
C0 ? , however, raises cytoplasmic pH, due to the residual hydroxyl ions. To neutralize 
the cytoplasmic pH, a second molecule of bicarbonate is precipitated as calcium 
carbonate within an intracellular membrane-bound structure, the coccolith-containing 
vesicle (Fig. 6). This reaction consumes hydroxyl ions with the generation of carbonate 
and water: 



SGM symposium 65 



Ecology of giant sulfur bacteria 53 



OH-+HCOj-*CO^ + H 2 



Ca 2+ +CO|-^CaC0 3 (s 



An alternative mechanism by which Achromatium species might maintain a neutral pH 
in the cytoplasm would be to liberate protons directly from the reaction of bicarbonate 
ions with calcium. This is unlikely, as calcite is precipitated not at the cell membrane, 
but within membrane-bound intracellular inclusions (Fig. 6). 

Ca 2+ +HC03-CaC0 3 + H + 

Despite the fact that this mechanism of carbonate precipitation is encountered widely 
in the literature, the direct precipitation of calcium carbonate in this manner does not 
occur in nature. This is because calcite is ionic and can only be formed from its 
component ions. As a result, CO^must be formed before calcite formation can take 
place (Wright & Oren, 2005). 

This proposed basis for calcite precipitation in members of the genus Achromatium 
means that there should be a direct link between calcite precipitation and energy 
generation. For this reason, the energetics of the whole process need to be evaluated. 
Interestingly, the stoichiometry for iron sulfide oxidation linked to calcium carbonate 
precipitation shows that there is no net gain or loss of protons and, thus, the process 
should be chemiosmotically neutral: 

FeS + Ca 2+ +20 2 +HC03-HS04+CaC0 3 +Fe 2+ 

However, there are other factors that should be considered: for instance, Achromatium 
cells contain large amounts of calcite, even when calcium is below detection limits in its 
environment. This implies a massive accumulation of calcium against a considerable 
concentration gradient. There is thus likely to be an energetic cost of precipitating 
calcite that must be balanced against the enhanced energy generation supported by the 
consumption of intracellular hydroxyl ions. In environments with high levels of 
dissolved sulfide, it is unlikely that this mechanism to access solid-phase sulfides would 
be competitive, because of the cost of calcium carbonate precipitation. Interestingly, 
the only Achromatium species reported routinely from sulfidic marine systems is 
'Achromatium vo\utans\ which does not precipitate calcite. 

Giant sulfur bacteria and the nitrogen cycle. Some giant sulfur bacteria have an 
important role to play in the biogeochemical cycling of nitrogen. There is evidence that 
some Beggiatoa species are capable of fixing dinitrogen gas (Polman 6c Larkin, 1988), 
but more attention has been diverted toward the ability of a number of giant sulfur 
bacteria to use nitrate as a respiratory oxidant. Initially, this was presumed to be linked 
to denitrification (Fossing et ai, 1995; Sweerts et ai, 1990), but more recent evidence 

SGM symposium 65 



54 N.D.Gray and I. M. Head 

strongly supports the notion that the principal route for respiratory nitrate reduction is 
through DNRA (Otte et al., 1999). This has prompted a considerable reappraisal of the 
significance of mats of giant sulfur bacteria in the nitrogen dynamics of sedimentary 
environments. 

In many coastal marine environments, nitrogen is the most important limiting nutrient 
(Nixon, 1981; Ryther 6c Dunstan, 1971) and removal of nitrogen via denitrification can 
be an important control on eutrophication (Seitzinger, 1988). 

As organic carbon inputs to sediments increase in marine environments in particular, 
denitrification becomes less important and DNRA more so. This has been attributed to 
increased sulfate reduction and the accumulation of sulfide to concentrations that 
inhibit key enzymes in the denitrification pathway (Brunet &C Garcia-Gil, 1996), but, as 
discussed above, energetic considerations may also play a role. Nevertheless, it is clear 
that a shift in nitrogen-reduction pathways from denitrification to a conservative route 
of nitrate reduction, such as DNRA, will have important consequences for the overall 
budget of nitrogen and the productivity of a system. In this context, the ability of giant 
sulfur bacteria, such as members of the genera Thioploca and Beggiatoa, which 
accumulate high levels of nitrate, to reduce nitrate to ammonium coupled with sulfide 
oxidation (McHatton et al., 1996; Otte etal., 1999) clearly has significance. 

Studies of sediments off the coast of Chile have shown that nitrate fluxes into sediments 
with large populations of Thioploca cells are >50 % higher than in sediments with a 
low abundance of Thioploca cells (Zopfi et al., 2001). It has been suggested that, 
because Thioploca cells can extend into nitrate-containing waters overlying the 
sediments that they inhabit, they may monopolize the available nitrate before it can 
be transported to other sediment-dwelling bacteria. This would imply that other forms 
of nitrate respiration might be of minimal significance in nitrogen cycling in sediments 
harbouring large populations of Thioploca cells. Experimental evidence refutes 
this suggestion and denitrification rates in Thioploca-rich sediments were actually 
slightly higher (4-5-9 mmolrrT 2 day -1 ) than typical values for coastal sediments 
(Herbert, 1999). Thus, there was still sufficient nitrate reaching the sediment to support 
significant denitrification rates, despite considerable assimilation of water-column 
nitrate by Thioploca cells. Furthermore, nitrous oxide was also consumed rapidly, 
suggesting that complete denitrification occurred in the sediments (Zopfi et al., 2001). 
Although the sole route of nitrate reduction in these sediments is not through DNRA 
linked to sulfide oxidation and not all sulfide oxidation is mediated by Thioploca 
species, it has been estimated that if 25 % of sulfide oxidation is mediated by nitrate- 
reducing Thioploca cells [toward the lower range of reported estimates (16-91 %)], this 
would result in an increase in the net flux of ammonium from the sediments of S3 %. 



SGM symposium 65 



Ecology of giant sulfur bacteria 55 

This represents a significant increase in the amount of nitrogen retained in the system 
and is likely to have a significant effect on the overall primary productivity in such 
coastal ecosystems, which are typically nitrogen-limited. Indeed, although Graco et al. 
(2001) found that the main source of ammonium in Chilean coastal sediments was in 
fact mineralization of organic matter and only 17 % of the total recycled ammonium 
resulted from the DNRA activity of giant sulfur bacteria, the activity of the sulfur 
bacteria still made a significant contribution to the recycled nitrogen and could account 
for >10 % of the primary productivity supported by recycled nitrogen. Similar con- 
clusions have been drawn from studies of sediments in Tokyo Bay, which have also 
found that nitrate-accumulating, vacuolated sulfur bacteria have an important role in 
nitrogen retention in this coastal ecosystem. Interestingly, conditions of high 
productivity are likely to exacerbate the situation, as higher inputs of organic matter 
stimulate sulfate reduction, leading to higher sulfide concentrations that may repress 
denitriflcation and favour DNRA (Brunet & Garcia-Gil, 1996), resulting in greater 
nitrogen retention in the system with potential effects on long-term eutrophication. It is 
interesting to speculate about the wider implications of nitrogen retention on the bio- 
geochemical cycling of sulfur and carbon. Effective recycling of nitrogen within 
the system will ultimately lead to greater sulfate regeneration and, consequently, the 
importance of carbon mineralization through sulfate reduction may also be increased. 
This would be further enhanced by the greater stoichiometric efficiency of DNRA 
relative to denitriflcation. 

Giant sulfur bacteria and the carbon cycle. The vast accumulations of biomass in 
mats of giant sulfur-oxidizing bacteria represent a significant pool of organic carbon. It 
has been shown that sedimentary carbon dynamics are affected by the presence of mats 
of Thioploca and Beggiatoa cells, and sediments characterized by sulfur-oxidizing 
bacterial mats have been shown to promote higher rates of carbon burial (Graco et al., 
2001). It has been suggested that this may be a consequence of sheaths and cells of 
Thioploca and Beggiatoa containing refractory carbon and limited grazing on the giant 
bacterial cells, although no direct measurements of the resistance of sulfur bacterial 
sheaths have been made (Graco et al., 2001). 

Most studies of the effect of sulfur bacterial mats on carbon cycling have focused on 
cold-seep environments, where methane and hydrocarbons are important carbon 
sources driving the microbial ecosystem. Mats of different colour have often been 
reported from such sites and recent studies have shown that the colour of the mats, 
which is believed to reflect high cytochrome content of the orange-mat bacteria, 
correlates with the type of carbon metabolism exhibited by the sulfur bacteria 
(Nikolaus et al., 2003). However, in this case, it was indicated that the water-soluble 
pigments, which had an absorbance maximum at 390 nm, were unlike typical cyto- 

SGM symposium 65 



56 N. D. Gray and I. M. Head 

chromes (Nikolaus et al., 2003). The non-pigmented mats expressed high but variable 
levels of RuBisCO activity, suggesting that the non-pigmented sulfur bacteria had 
chemoautotrophic potential. RuBisCO activity in the pigmented filaments was several 
orders of magnitude lower, suggesting that they were more likely to be heterotrophic, 
perhaps utilizing hydrocarbons or hydrocarbon-oxidation products prevalent at the 
seep site (Nikolaus et al., 2003). 

Sulfur bacterial mats are characteristic of many methane-rich environments and it has 
even been suggested that the mats may form a relatively impermeable barrier that 
restricts the movement of methane out of the sediment (Orphan et al., 2004). This may 
help to promote anaerobic oxidation of methane by holding methane in surface 
sediments, close to the large pool of sulfate in the overlying water and also to a source of 
sulfate from reoxidation of sulfide by the sulfur bacteria themselves. Consequently, high 
levels of anaerobic methane oxidation appear to be associated with conspicuous sulfur 
bacterial mats (Joye et al., 2004; Treude et al., 2003). The high levels of anaerobic 
methane oxidation, which are typically linked to sulfate reduction in marine systems, 
appear to be supported by the effective reoxidation of sulfide by the bacterial mats (Joye 
et al., 2004). Isotopic data, obtained from a combination of fluorescent in situ hybridi- 
zation and secondary-ion mass spectrometry (FISH-SIMS; Orphan et al., 2001), 
suggest that the filamentous mat bacteria also incorporate methane-derived light 
carbon (Orphan et al., 2004). It is clear from these data and the precipitation of iso- 
topically light carbonates in seep environments that much of the carbon from active 
seep sites associated with sulfur bacterial mats is retained in the system by the 
combined activities of the anaerobic methane-oxidizing archaeal-sulfate-reducer 
consortia and mats of giant sulfur bacteria (Boetius & Suess, 2004). It is not only the 
microbial component of the ecosystem that is affected by the carbon dynamics in seep 
environments; higher trophic levels are also influenced by the presence of sulfur 
bacterial mats. One study has shown that both the density and diversity of fauna are 
greater at seep sites with extensive sulfur bacterial mats, compared with similar settings 
that lack the bacterial mats. Mats of Tbioploca cells in the Gulf of Mexico in particular 
appeared to support a more diverse foraminiferan community, compared with mats of 
Beggiatoa cells and off-mat sites (Robinson et al., 2004). 

EVOLUTIONARY AND ECOLOGICAL DIVERSITY IN THE GIANT 
SULFUR BACTERIA 

Giant sulfur bacteria have been shown by comparative 16S rRNA gene sequence 
analysis to be distributed between two classes of the phylum Proteobacteria (Teske 
et al., 1996). The majority (Beggiatoa-Thioploca-Thiomargarita, Achromatium and 
Thiothrix) form well-defined clades within the l Gammaproteobacteria\ However, the 
genus Thiovolum clusters with the 'Epsilonproteobacteria' (Fig. 2). 

SGM symposium 65 



Ecology of giant sulfur bacteria 57 

The different genera of giant sulfur bacteria have each evolved major morphological 
and physiological adaptations that allow them to exploit sulfur oxidation in different 
geochemical settings. Within each of these genera, however, there is evidence for more 
recent diversification and adaptation. It has been possible to correlate fine-scale 
patterns of genetic diversity with ecological function in giant sulfur bacteria and, by 
extrapolation, this is providing insights into the ecological mechanisms that underpin 
diversification in the microbial world more generally. 

Within-genus diversity in morphology and physiology 

On the basis of comparative 16S rRNA gene sequence analysis, the genera Beggiatoa, 
Thioploca and Tbiomargarita occupy a monophyletic group within the Proteobacteria. 
Within this, four distinct lineages are evident (Teske & Nelson, 2004). The large- 
vacuolated, marine Beggiatoa and Thioploca species, Tbiomargarita namibiensis and 
a recently described, vacuolated, rosette-forming sulfur bacterium (Kalanetra et al., 
2004) occupy one group, whilst freshwater Thioploca species form a second group of 
closely related organisms. The non-vacuolated, marine Beggiatoa species form the third 
group, with the fourth comprising freshwater Beggiatoa isolates (Teske & Nelson, 
2004). Until relatively recently, our knowledge of the diversity of these bacteria has 
largely been defined by a few recognized species, identified on the basis of filament 
diameter, the presence of nitrate-storing vacuoles and the presence or absence of sheath 
material (Teske 6c Nelson, 2004). Some of these morphological features map onto 
the 16S rRNA gene sequence-based phylogeny (e.g. filament diameter and presence of 
vacuoles), whereas others (sheath formation) do not. As more 16S rRNA gene sequence 
data become available for giant sulfur bacteria, it is becoming apparent that there 
is considerably greater genetic diversity within each of the described genera than the 
original, morphologically based descriptions would suggest. For instance, a recent 
study of the diversity and population structure of filamentous sulfur bacteria in the 
Danish Limfjorden and the German Wadden Sea showed high diversity in nitrate- 
storing, vacuolated Beggiatoa species coexisting within the same sediments (Mufonann 
et al., 2003). These organisms represented novel phylogenetic clusters distinct from 
those characterized previously. It was predicted from the wide spectrum of filament 
diameters encountered in these environments and the relatively restricted diameter of 
individual Beggiatoa species identified by using FISH probes that the entire diversity 
that was present within the genus Beggiatoa had not been surveyed completely. 
Interestingly, the coexisting Beggiatoa species identified by FISH showed depth-related 
differences in their distribution (MuEmann et al., 2003). This was attributed either to 
the ability of species with a wider average filament diameter to store nitrate in greater 
quantities and hence reside at greater depths for longer, or the ability of species with a 
smaller average diameter to scavenge more efficiently for dissolved sulfide, which is in 
limited supply in the near-surface environment. This is echoed by the findings that the 

SGM symposium 65 



58 N. D. Gray and I. M. Head 

depth distribution of Thioploca chileae, Thioploca araucae and a third Thioploca 
species, designated SCM (short-cell morphotype), in South American coastal sedi- 
ments was different, and the nitrate and sulfur content of the different Thioploca 
species correlated with the different location of the bacteria in the sediment (Zopfi 
et al., 2001). As in most other locations, nitrate-accumulating Beggiatoa cells from 
organic-rich sediments in Tokyo Bay also had varying widths of trichome (Kojima & 
Fukui, 2003). The morphotype with wider trichomes was related most closely to 
uncultured Beggiatoa species from other geographical localities, identified from 16S 
rRNA gene sequence analysis, and distinct from the Tokyo Bay morphotype with 
narrow trichomes. Among the narrower types, which formed a new branch within 
the Beggiatoa-Thopmargarita-Thioploca clade, a sample of cells from a tidal flat 
harboured a narrow Beggiatoa-like morphotype that was genetically distinct from 
those identified in samples taken in water depths of 10 and 20 m. This suggests that 
there may be some ecological differentiation between closely related Beggiatoa geno- 
types. Analysis of 16S rRNA gene sequences from Chilean Thioploca isolates revealed 
several coexisting, genetically distinct species that can be differentiated by filament 
diameter. Not only were these found in the same sediment sample, but they also 
occurred commonly within the same sheath (Teske et al., 1996). In contrast, a more 
recent study of freshwater Thioploca isolates from Lake Biwa, Japan, and Lake 
Constance, Germany, indicated that two freshwater Thioploca isolates from geo- 
graphically separated environments were barely distinguishable on the basis of 
trichome diameter and had almost identical 16S rRNA gene sequences (Kojima et al., 
2003). 

As a relatively large number of Beggiatoa species have been isolated in axenic culture, a 
great deal more is known about the physiology of members of the genus Beggiatoa than 
other large sulfur bacteria. Members of the genus Beggiatoa have a diverse carbon 
metabolism, ranging from obligate chemolithoautotrophy, facultative chemolitho- 
autotrophy and mixotrophy through to lithoheterotrophy This phenotypic diversity is 
observed at the species level, so it appears that carbon metabolism has been a key driver 
of ecological and evolutionary diversification. For instance, although it appears that the 
marine Beggiatoa species (vacuolated and unvacuolated) display predominantly 
chemolithoauto trophic nutrition (Teske & Nelson, 2004), it also appears that they can 
be either obligate or facultative autotrophs, as demonstrated by the related marine 
strains MS-81-lc and MS-81-6 (Hagen & Nelson, 1996). For instance, in the presence of 
acetate, carbon fixation by the obligately chemolithotrophic strain MS-81-lc was not 
reduced significantly (Hagen & Nelson, 1996). In addition, the use of radiolabeled 
substrates demonstrated that this organism was unable to respire acetate and that 
2-oxoglutarate reductase, an enzyme necessary for the respiration of organic substrates 
in the tricarboxylic cycle, was absent. Even in the presence of organic compounds, 80 % 

SGM symposium 65 



Ecology of giant sulfur bacteria 59 

of cell carbon was obtained from C0 9 . Strain MS-81-lc also obtained more energy 
from the oxidation of reduced sulfur species than the facultatively chemolitho- 
autotrophic marine strain MS-81-6 (Hagen & Nelson, 1997). MS-81-6, although able 
to grow chemolithoautotrophically, was found to be metabolically more versatile and 
could utilize acetate freely for energy or biosynthesis (Hagen & Nelson, 1996). 
Beggiatoa sp. strains MS-81-lc and MS-81-6 were originally isolated from the same 
environment (Great Sippewissett salt marsh, Woods Hole, MA, USA) and the physio- 
logical differences between the strains may be a factor in the maintenance of the 
diversity of Beggiatoa species in this single geographical location. 

In contrast to marine organisms, there has been considerable debate as to whether 
genetically differentiated freshwater Beggiatoa strains are capable of genuine 
autotrophic or lithoheterotrophic growth (Hagen & Nelson, 1997; Strohl, 2005; Teske 
& Nelson, 2004). Whilst mixotrophic nutrition has been claimed for a large number 
of freshwater strains (Hagen & Nelson, 1996), it has only recently been shown that a 
freshwater isolate can grow lithoautotrophically (Grabovich et al., 2001). Nevertheless, 
it is clear that the genetic differentiation between the marine and freshwater Beggiatoa 
species is correlated with fundamental differences in carbon metabolism, given the 
predominance of heterotrophic/mixotrophic metabolism in the freshwater strains and 
autotrophic metabolism in the marine species. The ecological sense of these differences 
may be explained by the availability of sulfide in marine and freshwater environments. 
The metabolic versatility in the freshwater strains is consistent with the potentially low 
and variable supply of sulfide in these environments and hence the requirement to 
supplement growth by the utilization of organic carbon. 

At present, very little is known about the genetic diversity of the genus Thiomargarita 
and 16S rRNA gene sequence data are only available for Thiomargarita namibiensis 
from Walvis Bay, Namibia, south-west Africa (Schulz et al., 1999). Unlike most com- 
munities of Achromatium, Beggiatoa and Thioploca species, samples of Thiomargarita 
namibiensis from Walvis Bay only produced a single 16S rRNA gene sequence, suggest- 
ing either that Thiomargarita namibiensis is genetically much more homogeneous than 
other giant sulfur bacteria or that any genetic differentiation of Thiomargarita species 
cannot be resolved on the basis of 16S rRNA gene sequence analysis. It may be of 
significance that the habitats where several distinct species of giant sulfur bacteria 
coexist tend to be spatially structured, relatively undisturbed sediments. The different 
genotypes observed in these systems may therefore reflect species adapted to different 
conditions in the sediment environment (Gray et al., 1999b). Members of the genus 
Thiomargarita are believed to have quite a different lifestyle from sulfur bacteria that 
exhibit genetic diversity at a single location. It is thought that Thiomargarita cells 
survive by living in an environment that is subject to periodic disturbance and mixing. 

SGM symposium 65 



60 N. D. Gray and I. M. Head 

The mixing brings sessile cells in sulfide-rich sediment into contact with water 
containing nitrate and oxygen, the oxidants that it uses for sulfide oxidation (Schulz & 
Jorgensen, 2001). It is often considered that mixed environments sustain less diversity 
than structured environments, as they provide a narrower range of environmental 
conditions that can be exploited by ecologically distinct organisms. This may in part 
explain the apparent homogeneity of communities of Thiomargarita cells. None- 
theless, chains of Thiomargarita cells also fall into distinct cell-size classes that, by 
analogy with other giant sulfur bacteria, may correlate with different species (Schulz 
et al., 1999). Alternatively, the failure to identify different Thiomargarita species may 
be due to limited sampling of the communities. 

The genus Thiothrix forms a deep branch within the ' Gammaproteobacteria' ; however, 
there is considerable physiological variation within the genus (Aruga et al., 2002; 
Howarth et ai, 1999; Rossetti et ai, 2003). Thiothrix species, like Beggiatoa species, are 
varied in their carbon metabolism and heterotrophs, mixotrophs and chemolitho- 
autotrophs are all represented by different Thiothrix species. At present, no studies have 
established how this genetic and physiological diversity relates to the distribution and 
abundance of Thiothrix species in engineered or natural ecosystems. However, a 
number of studies have shown that bacteria related closely to Thiothrix unzii, which 
can grow with organic carbon sources but has an obligate requirement for sulfide or 
thiosulfate (Unz &C Head, 2005), appear to be prevalent in some sulfidic cave environ- 
ments (Brigmon et ai, 2003; En gel et al., 2004). 

Genetic and ecological diversity in the genus Achromatium 

The tantalizing link between fine-scale patterns of genetic diversity and ecological 
differentiation in giant sulfur bacteria has been explored most thoroughly in the genus 
Achromatium. Natural communities of Achromatium cells that were originally 
thought to be genetically homogeneous in fact comprise a number of phylogenetically 
distinct subpopulations, distinguishable by FISH (Glockner et al., 1999; Gray et ai, 
1999b). The degree of identity observed in Achromatium-dtrivtd 16S rRNA gene 
sequences (<97-5%) from individual samples of freshwater sediment indicates 
that different species of Achromatium exist both in geographically separated locations 
and within a single sediment (Gray et al., 1999b). As with the marine Beggiatoa 
species described above, it has been shown that individual, coexisting Achrom- 
atium species exhibit physiological and ecological differentiation. Coexisting 
Achromatium species in sediment from the margins of Rydal Water in the English Lake 
District, like Beggiatoa and Thioploca species, fall into distinct size classes (Gray et ai, 
1999b). Furthermore, the composition of the Achromatium community was different 
in oxidizing and reducing zones within the sediment (Gray et al., 1999b). On this basis, 
it was hypothesized that genetic diversity in coexisting Achromatium communities 

SGM symposium 65 



Ecology of giant sulfur bacteria 61 

reflected exploitation of different redox-related niches. This was based on the competi- 
tive-exclusion principle, which states that if two similar species coexist in a stable 
environment, they do so as a result of niche differentiation. If, however, there is no niche 
differentiation, the species will compete and one will eliminate or exclude the other 
(Begon et aL, 1996; Gause, 1934; Hardin, 1960). Redox-related niche differentiation 
in Achromatium communities was demonstrated experimentally (Gray et aL, 2004). It 
was reasoned that, in sediment microcosms harbouring Achromatium communities, 
growth or maintenance of a particular Achromatium species exposed to particular 
redox conditions, accompanied by neutral or antagonistic effects on other coexisting 
Achromatium species, would be indicative of niche differentiation. In anoxic micro- 
cosms, Achromatium sp. RY8 decreased and Achromatium spp. RY5 and RYKS 
increased over time. Addition of increasing concentrations of nitrate, however, main- 
tained the population of Achromatium sp. RY8. When high levels of nitrate were 
maintained throughout the incubation, the composition of the Achromatium 
community remained stable over time (Fig. 7). This suggested that all of the coexisting 
Achromatium species are obligate or facultative anaerobes that can utilize nitrate as an 
electron acceptor, but it was not possible to establish whether Achromatium species 
utilized DNRA or denitrification. Achromatium sp. RY8 was clearly more sensitive to 
sediment redox conditions than the other Achromatium species. These results give a 
much clearer picture of the mechanism that supports coexistence of several 
Achromatium species adapted to different redox conditions (Gray et aL, 2004). Redox 
conditions in sediments vary with depth, and coexisting Achromatium species are 
adapted to different redox conditions and thus avoid direct competition. Clearly, then, 
the genetic diversity observed in Achromatium communities correlates to functional 
diversity. This in turn may have an impact on geochemical cycling. The inherently low 
levels of sulfate that typify sediments inhabited by freshwater Achromatium species 
suggest that regeneration of sulfate is an important process and may lead to a greater 
flow of electrons from organic carbon through sulfate reduction than would be 
suggested by the steady-state sulfate concentration. Consequently, Achromatium 
community composition may have a direct effect on the efficiency of the ecosystem 
process by allowing efficient reoxidation of sulfide under a range of different redox 
conditions. 

GENOMICS MEETS ECOLOGY: UNDERSTANDING ECOLOGICAL 
DIVERSITY BY USING METAGENOMICS 

Many giant sulfur bacteria exist as communities comprising several related species that 
are ecologically and physiologically distinct. Although techniques such as combined 
microautoradiography and FISH can provide some information on physiological 
differences between coexisting species in natural communities (Gray et aL, 2000), this 
provides at best a shuttered view of the properties that distinguish the coexisting 

SGM symposium 65 



62 N. D.Gray and I. M. Head 



(a) (i) 



(ii) 



(iii) 



1 

S3 



.1 
I 



Oh. 

£ 

— 

ai 



E 

C 

I 
s 

e 




*» 1^=1180 

1600 




0-8 
0-7 
06 
0-5 
04 
f>3 
0-2 
r>1 



e 




n= 1377 



rth» 



i 



Q 



A 



n 



4l 



Time 1 



0-5 01 0-01 







Initial nitrate concentration (mM) 



Achromatium or Thioploca species. With the advent of genomic technologies and the 
emergence of metagenomic analysis (DeLong, 2002), the potential now exists to recover 
large amounts of genomic information directly from natural microbial communities. 
Recent forays into metagenomics have helped to frame some of the limitations of 
the approach. The largest-scale metagenomic analysis conducted to date examined the 
microbial diversity of the Sargasso Sea (Venter et al., 2004). This breathtaking work 
resulted in the recovery of 1 Gb non-redundant sequence, yet still barely scratched the 
surface of the diversity present. Even with extreme brute force, the metagenome of 
marine ecosystems is a powerful adversary In contrast, a more comprehensive sampling 
with less intensive sequencing effort was obtained by focusing on an ecosystem known 
to have very limited diversity (acid minewaters from Iron Mountain, CA, USA; Tyson 
et al., 2004). It proved possible to reconstruct the genomes from the principal players 



SGM symposium 65 



Ecology of giant sulfur bacteria 63 



(b) (i) 



n=1134 



100 
60 

40 

20 




(ii) 



2 

E 
*-' 

"5 
IS 



06 

0-5 
0-4 
03 
02 



0-1 



- 



A 



n= 1041 n=883 



n=354 



n = 126S 




A: 



Time 



1 



01 



0-01 



& 







Initial nitrate concentration (mM) 



Fig. 7. The effect of nitrate availability on different Achromatium species in a freshwater sediment, (a) 
Relative abundance (i) and absolute numbers (ii) of different Achromatium species within sediment 
microcosms incubated for 7 days under anoxic conditions with different initial nitrate concentrations. 
Empty bars, Achromatium sp. RY5; shaded bars, Achromatium sp. RYKS; filled bars, Achromatium sp. 
RY8; diagonally hatched bars, Achromatium sp. RY1; horizontally hatched bars, Eub338-positive cells. 
Ammonium (□), sulfate (A) and nitrate (■) in sediment microcosms incubated for 7 days under 
different redox conditions (iii). (b) Relative abundance of different Achromatium species in sediment 
microcosms incubated for 13 days and supplied repeatedly with nitrate at different concentrations (i). 
Empty bars, Achromatium sp. RY5; shaded bars, Achromatium sp. RYKS; filled bars, Achromatium sp. 
RY8; diagonally hatched bars, Achromatium sp. RY1; horizontally hatched bars, Eub338-positive cells. 
Sulfate (A) and nitrate (■) in sediment microcosms after 1 3 days incubation (48 h after last addition of 
nitrate-containing overlying water) (ii). Time-zero analyses were conducted on replicate microcosms 
sampled sacrificially at the beginning of the experiments. Error bars on all data indicate sem of triplicate 
incubations. Where error bars are not shown, the error bars were smaller than the symbols, n, No. 
cells counted for each replicate set of microcosms. Reproduced from Gray etal. (2004) with 
permission from Blackwell Publishing. 

in the highly acidic environment and, consequently, to link the genomic information 
with the specific metabolic and ecological role of the most abundant organisms present. 
The analyses of the Sargasso Sea and the acid minewater were both essentially 
exploratory in nature. Whilst this is a valid way to proceed and leads to many novel 
and exciting discoveries, different opportunities are offered by hypothesis-driven 
research. Testing of hypotheses is only possible within the framework of a rigorous 
experimental design. The poor sampling afforded by metagenomic analysis of many 
environments precludes such an experimental design. The work of Tyson et al. (2004), 



SGM symposium 65 



64 N. D. Gray and I. M. Head 

however, illustrates that it is possible to obtain meaningful sampling of a microbial 
community that would lend itself to a defensible comparative analysis. Communities of 
giant sulfur bacteria offer an ideal system for the pursuit of hypothesis-driven 
metagenomics. Giant sulfur bacteria occur naturally as highly enriched, high-biomass 
communities. The communities typically comprise several coexisting species that can 
be distinguished genetically at the level of comparative 16S rRNA gene sequence 
analysis and the different genotypes can be shown to be physiologically and ecolo- 
gically distinct. Extraction of total nucleic acids and generation of high-coverage clone 
libraries from such communities have a high chance of reconstructing the genomes of 
the most abundant community members. Comparative genome analysis would then 
provide the opportunity to identify the genetic basis for the ecological differentiation 
observed in natural communities of giant sulfur bacteria. 

REFERENCES 

Aruga, S., Kamagata, Y, Kohno, T., Hanada, S. ; Nakamura, K. & Kanagawa, T. (2002). 

Characterization of filamentous Eikelboom type 021 N bacteria and description of 

Thiothrix disciformis sp. nov. and Thiothrix flexilis sp. nov. IntJ Syst Evol Microbiol 52, 

1309-1316. 
Babenzien, H.-D., Glockner, F.-O. & Head, I. M. (2005). Genus Achromatium. In Bergey's 

Manual of Systematic Bacteriology, 2nd edn, vol. 2, part B: Gammaproteobacteria, 

pp. 142-147. Edited by G. Garrity. New York: Springer. 
Begon, M., Harper, J. L. & Townsend, C. R. (1996). Ecology: Individuals, Populations and 

Communities, 3rd edn. Oxford: Blackwell Science. 
Berks, B. C, Ferguson, S. J., Moir, J. W. B. & Richardson, D. J. (1995). Enzymes and 

associated electron transport systems that catalyse the respiratory reduction of 

nitrogen oxides and oxyanions. Biochim Biophys Acta 1232, 97-173. 
Boetius, A. & Suess, E. (2004). Hydrate Ridge: a natural laboratory for the study of microbial 

life fueled by methane from near-surface gas hydrates. Chem Geo/ 205, 291-310. 
Bonin, P., Omnes, P. & Chalamet, A. (1998). Simultaneous occurrence of denitrification 

and nitrate ammonification in sediments of the French Mediterranean Coast. 

Hydrobiologia 389, 169-182. 
Borowitzka, M. A. (1982). Morphological and cytological aspects of algal calcification. Int 

RevCytollA, 127-162. 
Brigmon, R. L, Martin, H. W., Morris, T. L, Bitton, G. & Zam, S. G. (1994). Biogeo- 

chemical ecology of Thiothrix spp. in underwater limestone caves. Geomicrobiol J 12, 

141-159. 
Brigmon, R. L, Furlong, M. & Whitman, W. B. (2003). Identification of Thiothrix unzii in 

two distinct ecosystems. LettAppI Microbiol 36, 88-91 . 
Bruchert, V., Perez, M. E. & Lange, C. B. (2000). Coupled primary production, benthic 

foraminiferal assemblage, and sulfur diagenesis in organic-rich sediments of the 

Benguela upwelling system. MarGeol 163, 27-40. 
Bruchert, V., Jorgensen, B. B., Neumann, K., Riechmann, D., Schlosser, M. & Schulz, H. 

(2003). Regulation of bacterial sulfate reduction and hydrogen sulfide fluxes in 

the central Namibian coastal upwelling zone. Geochim Cosmochim Acta 67, 

4505-4518. 



SGM symposium 65 



Ecology of giant sulfur bacteria 65 



Brunet, R. C. & Garcia-Gil, L. J. (1996). Sulfide-induced dissimilatory nitrate reduction to 

ammonia in anaerobic freshwater sediments. FE MS Microbiol Ecol 21 , 131-138. 
Burton, S. D. & Morita, R. Y. (1964). Effect of catalase and cultural conditions on growth of 

Beggiatoa. J Bacteriol 88, 1 755-1 761 . 
Cannon, G. C, Strohl, W. R., Larkin, J. M. & Shively, J. M. (1979). Cytochromes in 

Beggiatoa alba. Curr Microbiol 2, 263-266. 
De Boer, W. E., La Riviere, J. W. M. & Schmidt, K. (1971). Some properties of 

Achromatium oxaliferum. Antonie van Leeuwenhoek 37, 553-563. 
DeLong, E. F. (2002). Microbial population genomics and ecology. Curr Opin Microbiol 

5, 520-524. 
Engel, A. S., Porter, M. L, Stern, L. A. f Quinlan, S. & Bennett, P. C. (2004). Bacterial 

diversity and ecosystem function of filamentous microbial mats from aphotic (cave) 

sulfidic springs dominated by chemolithoautotrophic " Epsilonproteobacteria" . FEM5 

Microbiol Ecol '51,31-53. 
Eprintsev, A. T. f Falaleeva, M. I., Stepanova, I. Y. & Parfenova, N. V. (2003). Purification 

and physicochemical properties of malate dehydrogenase from bacteria of the genus 

Beggiatoa. Biochemistry (Mosc) 68, 1 72-1 76. 
Eprintsev, A. T., Falaleeva, M. I., Grabovich, M. Yu., Parfenova, N. V., Kashirskaya, 

N. N. & Dubinina, G. A. (2004). The role of malate dehydrogenase isoforms in 

the regulation of anabolic and catabolic processes in the colorless sulfur bacterium 

Beggiatoa leptomitiformis D-402. Microbiology (English translation of Mikro- 

biologiia) 73, 367-371. 
Fenchel, T. (1994). Motility and chemosensory behaviour of the sulphur bacterium 

Thiovulum majus. Microbiology 140, 31 09-31 1 6. 
Fenchel, T. & Glud, R. N. (1998). Veil architecture in a sulphide-oxidizing bacterium 

enhances countercurrent flux. Nature 394, 367-369. 
Ferdelman, T. G., Lee, C, Pantoja, S., Harder, J., Bebout, B. M. & Fossing, H. (1997). 

Sulfate reduction and methanogenesis in a 777/op/oca-dominated sediment off the 

coast of Chile. Geochim Cosmochim Acta 61, 3065-3079. 
Fossing, H., Gallardo, V. A., Jorgensen, B. B. & 12 other authors (1995). Concentration 

and transport of nitrate by the mat-forming sulphur bacterium Thioploca. Nature 

374,713-715. 
Friedrich, C. G., Rother, D., Bardischewsky, F., Quentmeier, A. & Fischer, J. (2001). 

Oxidation of reduced inorganic sulfur compounds by bacteria: emergence of a 

common mechanism? Appl Environ Microbiol 67 , 2873-2882. 
Garrity, G. M., Bell, J. A. & Lilburn, T. G. (2005). Taxonomic outline of the prokaryotes. In 

Bergey's Manual of Systematic Bacteriology, 2nd edn, vol. 2. Edited by G. Garrity. 

New York: Springer (http://dx.doi.org/10.1007/bergeysoutline). 
Gause, G. F. (1934). The Struggle for Existence. Baltimore, MD: Williams & Wilkins. 
Glockner, F. O., Babenzien, H.-D., Wulf, J. & Amann, R. (1999). Phylogeny and diversity 

of Achromatium oxaliferum. Syst Appl Microbiol 22, 28-38. 
Grabovich, M. Yu., Dubinina, G. A., Lebedeva, V. Yu. & Churikova, V. V. (1998). 

Mixotrophic and lithoheterotrophic growth of the freshwater filamentous sulfur 

bacterium Beggiatoa leptomitiformis D-402. Microbiology (English translation of 

Mikrobiologiia) 67, 383-388. 
Grabovich, M. Yu., Muntyan, M. S., Lebedeva, V. Yu., Ustiyan, V. S. & Dubinina, G. A. 

(1999). Lithoheterotrophic growth and electron transfer chain components of the 

filamentous gliding bacterium Leucothrix mucor DSM 2157 during oxidation of 

sulfur compounds. FEMS Microbiol Lett 1 78, 1 55-1 61 . 



SGM symposium 65 



66 N. D. Gray and I. M. Head 



Grabovich, M. Yu., Patritskaya, V. Yu., Muntyan, M. S. & Dubinina, G. A. (2001). 

Lithoautotrophic growth of the freshwater strain Beggiatoa D-402 and energy 
conservation in a homogeneous culture under microoxic conditions. FEM5 Microbiol 
Lett 204, 341-345. 

Grabovich, M. Yu., Dul'tseva, N. M. & Dubinina, G. A. (2002). Carbon and sulfur 
metabolism in representatives of two clusters of bacteria of the genus Leucothrix: a 
comparative study. Microbiology (English translation ofMikrobiologiia) 71, 255-261 . 

Graco, M., Farias, L, Molina, V., Gutierrez, D. & Nielsen, L. P. (2001). Massive 
developments of microbial mats following phytoplankton blooms in a naturally 
eutrophic bay: implications for nitrogen cycling. Limnol Oceanogr 46, 821-832. 

Gray, N. D., Pickup, R. W., Jones, J. G. & Head, I. M. (1997). Ecophysiological evidence 
that Achromatium oxaliferum is responsible for the oxidation of reduced sulfur 
species to sulfate in a freshwater sediment. Appl Environ Microbiol 63, 1 905-1 91 0. 

Gray, N. D., Howarth, R., Pickup, R. W., Jones, J. G. & Head, I. M. (1999a). Substrate 
uptake by uncultured bacteria from the genus Achromatium determined by 
microautoradiography. Appl Environ Microbiol 65, 51 00-51 06. 

Gray, N. D., Howarth, R., Rowan, A., Pickup, R. W., Jones, J. G. & Head, I. M. (1999b). 
Natural communities of Achromatium oxaliferum comprise genetically, morpho- 
logically, and ecologically distinct subpopulations. Appl Environ Microbiol 65, 5089- 
5099. 

Gray, N. D., Howarth, R., Pickup, R. W., Jones, J. G. & Head, I. M. (2000). Use of 
combined microautoradiography and fluorescence in situ hybridization to determine 
carbon metabolism in mixed natural communities of uncultured bacteria from the 
genus Achromatium. Appl Environ Microbiol 66, 451 8-4522. 

Gray, N. D., Comaskey, D., Miskin, I. P., Pickup, R. W., Suzuki, K. & Head, I. M. (2004). 
Adaption of sympatric Achromatium spp. to different redox conditions as a 
mechanism for coexistence of functionally similar sulphur bacteria. Environ Microbiol 
6, 669-677. 

Guerrero, R., Haselton, A., Sole, M., Wier, A. & Margulis, L (1999). Titanospirillum 
velox: a huge, speedy, sulfur-storing spirillum from Ebro Delta microbial mats. Proc 
Natl Acad Sci US A 96, 11584-11588. 

Gunderson, J. K., Jorgensen, B. B., Larsen, E. & Jannasch, H. W. (1992). Mats of giant 
sulphur bacteria on deep-sea sediments due to fluctuating hydrothermal flow. 
Nature 360, 454-456. 

Hardin, G. (1960). The competitive exclusion principle. Science 131, 1292-1297. 

Hagen, K. D. & Nelson, D. C. (1996). Organic carbon utilization by obligately and 
facultatively autotrophic Beggiatoa strains in homogeneous and gradient cultures. 
Appl Environ Microbiol 62, 947-953. 

Hagen, K. D. & Nelson, D. C. (1997). Use of reduced sulfur compounds by Beggiatoa 
spp. enzymology and physiology of marine and freshwater strains in homogeneous 
and gradient cultures. Appl Environ Microbiol 63, 3957-3964. 

Head, I. M., Gray, N. D., Clarke, K. J., Pickup, R. W. & Jones, J. G. (1996). The 
phylogenetic position and ultrastructure of the uncultured bacterium Achromatium 
oxaliferum. Microbiology 142, 2341-2354. 

Head, I. M., Gray, N. D., Babenzien, H.-D. & Glockner, F. O. (2000a). Uncultured giant 
sulfur bacteria of the genus Achromatium. FEMS Microbiol Ecol 33, 171-180. 

Head, I. M., Gray, N. D., Howarth, R., Pickup, R. W., Clarke, K. J. & Jones, J. G. (2000b). 
Achromatium oxaliferum - understanding the unmistakable. Adv Microbiol Ecol 16, 
1-40. 



SGM symposium 65 



Ecology of giant sulfur bacteria 67 



Herbert, R. A. (1999). Nitrogen cycling in coastal marine ecosystems. FEMS Microbiol Rev 

23, 563-590. 
Hinze, G. (1913). Beitragezur Kenntnisderfarblosen Schwefelbakterien. Ber Dtsch Bot Ges 

31, 189-202 (in German). 
Howarth, R., Unz, R. R, Seviour, E. M., Seviour, R. J., Blackall, L L, Pickup, R. W., Jones, 

J. G., Yaguchi, J. & Head, I. M. (1999). Phylogenetic relationships of filamentous 

sulfur bacteria (Thiothrix spp. and Eikelboom type 02 1N bacteria) isolated from 

waste water- treatment plants and description of Thiothrix eikelboomii sp. nov., 

Thiothrix unziisp . nov., Thiothrix fructosivorans sp. nov. and Thiothrix def I uviisp. nov. 

Int J Syst Bacteriol 49, 1817-1827. 
Huettel, M., Forster, S., Kloser, S. & Fossing, H. (1996). Vertical migration in the sediment- 
dwelling sulfur bacteria Thioploca spp. in overcoming diffusion limitations. Appl 

Environ Microbiol S2, 1863-1872. 
Jorgensen, B. B. & Gallardo, V. A. (1999). Thioploca spp.: filamentous sulfur bacteria with 

n itrate vacuoles. FEMS Microbiol Ecol 28,301-313. 
Jorgensen, B. B. & Revsbech, N. P. (1983). Colorless sulfur bacteria, Beggiatoa spp. and 

Thiovulum spp., in 2 and H 2 S microgradients. Appl Environ Microbiol 45, 1261- 

1270. 
Joye, S. B., Boetius, A., Orcutt, B. N., Montoya, J. P., Schulz, H. N., Erickson, M. J. & 

Lugo, S. K. (2004). The anaerobic oxidation of methane and sulfate reduction in 

sediments from Gulf of Mexico cold seeps. Chem Geo/ 205, 219-238. 
Kalanetra, K. M., Huston, S. L & Nelson, D. C. (2004). Novel, attached, sulfur-oxidizing 

bacteria at shallow hydrothermal vents possess vacuoles not involved in respiratory 

nitrate accumulation. Appl Environ Microbiol 70, 7487-7496. 
Kappler, U. & Dahl, C. (2001). Enzymology and molecular biology of prokaryotic sulfite 

oxidation. FEMS Microbiol Lett 203, 1-9. 
Kelly, D. P. (1999). Thermodynamic aspects of energy conservation by chemolithotrophic 

sulfur bacteria in relation to the sulfur oxidation pathways. Arch Microbiol 171, 

219-229. 
Kelly, D. P., Shergill, J. K., Lu, W.-P. & Wood, A. P. (1997). Oxidative metabolism of 

inorganic sulfur compounds by bacteria. Antonie van Leeuwenhoek 71, 95-1 07. 
Kojima, H. & Fukui, M. (2003). Phylogenetic analysis of Beggiatoa spp. from organic rich 

sediment of Tokyo Bay, Japan. Water Res 37, 3216-3223. 
Kojima, H., Teske, A. & Fukui, M. (2003). Morphological and phylogenetic character- 
izations of freshwater Thioploca species from Lake Biwa, Japan, and Lake Constance, 

Germany. Appl Environ Microbiol 69, 390-398. 
Lane, D. J., Harrison, A. P., Jr, Stahl, D., Pace, B., Giovannoni, S. J., Olsen, G. J. & Pace, 

N. R. (1992). Evolutionary relationships among sulfur- and iron-oxidizing eubacteria. 

J Bacteriol 174, 269-278. 
Larkin, J. M. (1989). Genus II. Thiothrix Winogradsky 1888. In Bergey's Manual of 

Systematic Bacteriology ; vol. 3, pp. 2098-2101 . Edited by J. T Staley, M. P. Bryant, 

N. Pfennig & J. G. Holt. Baltimore, MD: Williams & Wilkins. 
Larkin, J. M. & Strohl, W. R. (1983). Beggiatoa, Thiothrix, and Thioploca. Annu Rev 

Microbiol 37, 341-367. 
Lauterborn, R. (1907). Eine neue Gattung der Schwefelbakterien (Thioploca schmidlei nov. 

gen. nov. spec). Ber Dtsch Bot Ges 25, 238-242 (in German). 
McGlannan, M. F. & Makemson, J. C. (1990). HCO3 fixation by naturally occurring tufts 

and pure cultures of Thiothrix nivea. Appl Environ Microbiol 56, 730-738. 
McHatton, S. C, Barry, J. P., Jannasch, H. W. & Nelson, D. C. (1996). High nitrate 



SGM symposium 65 



68 N. D. Gray and I. M. Head 



concentrations in vacuolate, autotrophic marine Beggiatoa spp. Appl Environ 

Microbiol 62, 954-958. 
Moissl, C, Rudolph, C. & Huber, R. (2002). Natural communities of novel archaea and 

bacteria with a string-of-pearls-like morphology: molecular analysis of the bacterial 

partners. Appl Environ Microbiol 68, 933-937. 
Morse, J. W. & Emeis, K. C. (1990). Controls on C/S ratios in hemipelagic sediments. Am J 

Sci 290, 1117-1135. 
MuBmann, M., Schulz, H. N., Strotmann, B., Kjaer, T. # Nielsen, L. P., Rossello-Mora, 

R. A., Amann, R. I. & Jorgensen, B. B. (2003). Phylogeny and distribution of nitrate- 
storing Beggiatoa spp. in coastal marine sediments. Environ Microbiol 5, 523-533. 
Nelson, D. C. & Hagen, K. D. (1995). Physiology and b iochemistry of symbiotic and free- 
living chemoautotrophic sulfur bacteria. Am Zoo! 35, 91-101. 
Nelson, D. C. & Jannasch, H. W. (1983). Chemoautotrophic growth of a marine Beggiatoa 

in sulfide-gradient cultures. Arch Microbiol 136, 262-269. 
Nelson, D. C, Waterbury, J. B. & Jannasch, H. W. (1982). Nitrogen fixation and nitrate 

utilization by marine and freshwater Beggiatoa. Arch Microbiol 1 33, 1 72-1 77. 
Nelson, D. C, Wirsen, C. O. & Jannasch, H. W. (1989). Characterization of large, 

autotrophic Beggiatoa spp. abundant at hydrothermal vents of the Guaymas Basin. 

Appl Environ Microbiol 55, 2909-291 7. 
Nikolaus, R., Ammerman, J. W. & MacDonald, I. R. (2003). Distinct pigmentation and 

trophic modes in Beggiatoa from hydrocarbon seeps in the Gulf of Mexico. Aguat 

MicrobEcol 32, 85-93. 
Nixon, S. W. (1981). Remineralization and nutrient cycling in coastal marine ecosystems. 

In Estuaries and Nutrients, pp. 1 1 1-138. Edited by B. J. Nelson & L. E. Cronin. 

Totowa, NJ: Humana Press. 
Odintsova, E. V. & Dubinina, G. A. (1993). Role of reduced sulfur compounds in Thiothrix 

ramosa metabolism. Microbiology (English translation of Mikrobiologiia) 62, 1 39-146. 
Orphan, V. J., House, C. H., Hinrichs, K.-U., McKeegan, K. D. & DeLong, E. F. (2001). 

Methane-consuming archaea revealed by directly coupled isotopic and phylogenetic 

analysis. Science 293, 484-487. 
Orphan, V. J., Ussier, W., Ill, Naehr, T. H., House, C. H., Hinrichs, K.-U. & Paull, C. K. 

(2004). Geological, geochemical, and microbiological heterogeneity of the seafloor 

around methane vents in the Eel River Basin, offshore California. Chem Geo! 205, 

265-289. 
Otte, S., Kuenen, J. G., Nielsen, LP.&7 other authors (1999). Nitrogen, carbon, and 

sulfur metabolism in natural Thioploca samples. Appl Environ Microbiol 65, 3148- 

3157. 
Patritskaya, V. Yu., Grabovich, M. Yu., Muntyan, M. S. & Dubinina, G. A. (2001). 

Lithoautotrophic growth of the freshwater colorless sulfur bacterium Beggiatoa 

"leptomitiformis" D-402. Microbiology (English translation of Mikrobiologiia) 70, 

145-150. 
Pimenov, N. V., Savvichev, A. S., Rusanov, I. I., Lein, A. Yu. & Ivanov, M. V. (2000). 

Microbiological processes of the carbon and sulfur cycles at cold methane seeps of 

the North Atlantic. Microbiology (English translation of Mikrobiologiia) 69, 709-720. 
Polman, J. K. & Larkin, J. M. (1988). Properties of in vivo nitrogenase activity in Beggiatoa 

alba. Arch Microbiol 1 50, 1 26-1 30. 
Rabenhorst, L. (1865). Flora Europaea Algarum aguae dulcis etsubmarinae. Sectio II: Algas 

phycochromaceas complectens, pp. 1-319. Leipzig, Germany: Eduard Kummer (in 

Latin). 



SGM symposium 65 



Ecology of giant sulfur bacteria 69 



Robinson, C. A., Bernhard, J. M. # Levin, L A., Mendoza, G. F. & Blanks, J. K. (2004). 

Surficial hydrocarbon seep infauna from the Blake Ridge (Atlantic Ocean, 21 50 m) 
and the Gulf of Mexico (690-2240 m). Mar Ecol 25, 313-336. 

Romaniuk, P. J., Zoltowska, B., Trust, T. J., Lane, D. J., Olsen, G. J., Pace, N. R. & Stahl, 
D. A. (1987). Campylobacter pylori ', the spiral bacterium associated with human 
gastritis, is not a true Campylobacter sp. J Bacteriol 1 69, 2 1 37-2 141. 

Rossetti, S., Blackall, L L, Levantesi, C, Uccelletti, D. &Tandoi, V. (2003). Phylogenetic 
and physiological characterization of a heterotrophic, chemolithoautotrophic Thio- 
thrix strain isolated from activated sludge. I nt J Syst Evol Microbiol 53, 1271-1276. 

Ryther, J. H. & Dunstan, W. M. (1971). Nitrogen, phosphorus, and eutrophication in the 
coastal marine environment. Science 171, 1008-1013. 

Schewiakoff, W. (1893). Ober einen neuen bacterienahnlichen Organismus des 
Susswassers. Habilitationsschrift, C. Winter, pp. 1-36. Heidelberg: Universitat Heidel- 
berg (in German). 

Schmidt, T. M. & DiSpirito, A. A. (1990). Spectral characterization of c-type cytochromes 
purified from Beggiatoa alba. Arch Microbiol 1 54, 453-458. 

Schmidt, T. M., Arieli, B., Cohen, Y, Padan, E. & Strohl, W. R. (1987). Sulfur metabolism 
in Beggiatoa alba. J Bacteriol 1 69, 5466-5472. 

Schulz, H. N. & Jorgensen, B. B. (2001). Big bacteria. Annu Rev Microbiol 55, 105-137. 

Schulz, H. N. & Jorgensen, B. B. (2005). Genus Thiomargarita. In Bergey's Manual of 
Systematic Bacteriology, 2nd edn, vol. 2, part B: Gammaproteobacteria, pp. 1 69— 
171 . Edited by G. Garrity. New York: Springer. 

Schulz, H. N., Jorgensen, B. B., Fossing, H. A. & Ramsing, N. B. (1996). Community 
structure of filamentous, sheath-building sulfur bacteria, Thioploca spp., off the 
coast of Chile. Appl Environ Microbiol 62, 1 855-1 862. 

Schulz, H. N., Brinkhoff, T., Ferdelman, T. G., Marine, M. H., Teske, A. & Jorgensen, 
B. B. (1999). Dense populations of a giant sulfur bacterium in Namibian shelf 
sediments. Science 16, 493-495. 

Seitzinger, S. P. (1988). Denitrification in freshwater and coastal marine ecosystems: 
ecological and geochemical significance. Li m nol Ocea nog r 33, 702-724. 

Simon, J. (2002). Enzymology and bioenergeticsof respiratory nitrite ammonification. FEMS 
Microbiol Rev 26, 285-309. 

Sorensen, J. (1978). Capacity for denitrification and reduction of nitrate to ammonia in a 
coastal marine sediment. Appl Environ Microbiol 35, 301-305. 

Starr, M. P. & Skerman, V. B. D. (1965). Bacterial diversity: the natural history of selected 
morphologically unusual bacteria. Annu Rev Microbiol 19, 407-454. 

Stepanova, I. Yu., Eprintsev, A. T., Falaleeva, M. I., Parfenova, N. V., Grabovich, M. Yu., 
Patritskaya, V. Yu. & Dubinina, G. A. (2002). Dependence of malate dehydro- 
genase structure on the type of metabolism in freshwater filamentous colorless sulfur 
bacteria of the genus Beggiatoa. Microbiology (English translation of Mikrobiologiia) 
71,377-382. 

Strohl, W. R. (2005). Genus Beggiatoa. In Bergey's Manual of Systematic Bacteriology, 2nd 
edn, vol. 2, part B: Gammaproteobacteria, pp. 148-161 . Edited by G. Garrity. New 
York: Springer. 

Strohl, W. R., Schmidt, T. M., Vinci, V. A. & Larkin, J. M. (1986). Electron transport and 
respiration in Beggiatoa and Vitreoscilla. Arch Microbiol 145, 71-75. 

Sweerts, J. -P. R. A., De Beer, D., Nielsen, L. P., Verdouw, H., Van den Heuvel, J. C, 
Cohen, Y & Cappenberg, T. E. (1990). Denitrification by sulphur oxidizing 
Beggiatoa spp. mats on freshwater sediments. Nature 344, 762-763. 



SGM symposium 65 



70 N. D. Gray and I. M. Head 



Teske, A. & Nelson, D. C. (2004). The genera Beggiatoa and Thioploca. In The Prokaryotes: 

an Evolving Electronic Resource for the Microbiological Community, 3rd edn, 

release 3.17. Edited by M. Dworkin, S. Falkow, E. Rosenberg, K.-H. Schleifer & 

E. Stackebrandt. New York: Springer (http://link.springer-ny.com/link/service/books/ 

10125/). 
Teske, A., Ramsing, N. B., Kuver, J. & Fossing, H. (1996). Phylogeny of Thioploca 

and related filamentous sulfide-oxidizing bacteria. SystAppI Microbiol 18, 517-526. 
Teske, A., Sogin, M. L, Nielsen, L P. & Jannasch, H. W. (1999). Phylogenetic relationships 

of a large marine Beggiatoa. SystAppI Microbiol 22, 39-44. 
Thamdrup, B. & Canfield, D. E. (1996). Pathways of carbon oxidation in continental 

margin sediments off central Chile. Limnol Oceanogr 41 , 1629-1650. 
Thar, R. & Kuhl, M. (2002). Conspicuous veils formed by vibrioid bacteria on sulfidic marine 

sediment. Appl Environ Microbiol 68, 631 0-6320. 
Trevisan, V. (1842). Prospetto della flora Euganea, pp. 1-68. Padova, Italy: Coi Tipi Del 

Seminario (in Italian). 
Treude, T., Boetius, A. # Knittel, K., Wallmann, K. & Jorgensen, B. B. (2003). Anaerobic 

oxidation of methane above gas hydrates at Hydrate Ridge, NE Pacific Ocean. Mar 

Ecol Prog Ser 264, 1-14. 
Tyson, G. W., Chapman, J., Hugenholtz, P. & 7 other authors (2004). Community 

structure and metabolism through reconstruction of microbial genomes from the 

environment. Nature 428, 37-43. 
Unz, R. F. & Head, I. M. (2005). Genus Thiothrix. In Bergey's Manual of Systematic 

Bacteriology, 2nd edn, vol. 2, part B: Gammaproteobacteria, pp. 131-142. Edited by 

G. Garrity. New York: Springer. 
Vaucher, J. P. (1803). Histoire des conferves d'eau douce, contenant leurs different modes 

de reproduction, et la description de leurs princi pales especes, pp. 1-285. Geneva: 

J. Paschoud (in French). 
Venter, J. C, Remington, K., Heidelberg, J. F. & 20 other authors (2004). Environmental 

genome shotgun sequencing of the Sargasso Sea. Science 304, 66-74. 
Winogradsky, S. (1887). Uber Schwefelbakterien. BotZtgAS, 489-610 (in German). 
Winogradsky, S. (1888). Beitrage zur Morphologie und Physiologie der Bacterien. Heft 1. 

Zur Morphologie und Physiologie der Schwefelbacterien, pp. 1-120. Leipzig: Arthur 

Felix (in German). 
Wright, D. & Oren, A. (2005). Nonphotosynthetic bacteria and the formation of 

carbonates and evaporites through time. Geomicrobiol J 22, 27-53. 
Zemskaya, T. I., Namsaraev, B. B., Dul'tseva, N. M., Khanaeva, T. A., Golobokova, L. P., 

Dubinina, G. A., Dulov, L. E. & Wada, E. (2001). Ecophysiological characteristics of 

the mat-forming bacterium Thioploca in bottom sediments of the Frolikha Bay, 

northern Baikal. Microbiology (English translation ofMikrobiologiia) 70, 335-341 . 
Zopfi, J., Kjaer, T., Nielsen, L. P. & Jorgensen, B. B. (2001). Ecology of Thioploca spp.: 

nitrate and sulfur storage in relation to chemical microgradients and influence of 

Thioploca spp. on the sedimentary nitrogen cycle. Appl Environ Microbiol 67, 

5530-5537. 



SGM symposium 65 



Soil micro-organisms in Antarctic 
dry valleys: resource supply and 
utilization 

D. W. Hopkins, 1 B. Elberling # 2 L G. Greenfield, 3 

E. G. Gregorich, 4 P. Novis, 5 A. G. O'Donnell 6 and 
A. D. Sparrow 3 ' 7 

1 School of Biological and Environmental Sciences, University of Stirling, Stirling FK9 4LA, 
Scotland, UK 

institute of Geography, University of Copenhagen, 0sterVoldgade 10, DK-1350, 
Copenhagen K., Denmark 

3 School of Biological Sciences, University of Canterbury, Private Bag 4800, Christchurch, New 
Zealand 

4 Agriculture Canada, Central Experimental Farm, Ottawa, Canada K1A 0C6 

5 Manaaki Whenua - Landcare Research, PO Box 69, Lincoln 81 52, New Zealand 

institute for Research on Environment and Sustainability, University of Newcastle upon Tyne, 
Newcastle upon Tyne NE1 7RU, UK 

department of Natural Resources and Environmental Sciences, University of Nevada, 
1 000 Valley Rd, Reno, NV 895 1 2, USA 

INTRODUCTION 

In 1903, the explorer Robert Scott was one of the first humans ever to see the dry valleys 
of Antarctica. He called them 'valley(s) of the dead' in which 'we have seen no sign of 
life, . . . not even a moss or lichen'. A century later, we know that the soils and rocks are 
home to many microscopic organisms that Scott could not have seen. 

The dry valleys are part of the small percentage of the land surface of the Antarctic 
continent that is ice-free, amounting to about 4000 km 2 , and thus have rock and soil 
surfaces that can be colonized by terrestrial organisms. They are an ancient polar 
desert, perhaps as much as 2 million years old, located in Victoria Land between about 
77 and 79° south (Fig. 1). The valleys are in a precipitation shadow caused by the 
Transantarctic Mountains, which rise over 4000 m. The Antarctic dry valleys are now 
recognized as one of the harshest terrestrial environments on Earth, characterized by 
summer maximum temperatures that rarely exceed 0°C and only a few tens of 
millimetres of precipitation, most of which falls as snow and is ablated by strong winds 
carrying dry air from the polar plateau - potential evaporation far exceeds precipi- 
tation (Fig. 1). The long periods of winter darkness are punctuated by a short summer, 
when 24 h daylight is reached for a few weeks either side of the summer solstice and 
when the ground surface temperature may rise to a few degrees above zero. This harsh 
environment has led to the Antarctic dry valleys being considered as analogues of 

SGM symposium 65: Micro-organisms and Earth systems - advances in geomicrobiology. 
Editors G. M. Gadd, K. T. Semple & H. M. Lappin-Scott. Cambridge University Press. ISBN 521 86222 1 ©SGM 2005 



72 D. W. Hopkins and others 



90° W 



fO 



<u 



<u 




90° E 



20 

10 



1 FMAMJ J ASOND 



g fl- 



£ -10 



-20 

-30 



-40 




Fig. 1. Map of Antarctica showing the location of the dry valleys (•) and the Transantarctic 
Mountains (dotted line), and summary temperatures for the dry valleys (filled bars) and McMurdo 
Sound (empty bars). McMurdo Sound data were measured at Scott Base on the Ross Sea coast at 
78° S; mean annual temperature, -20 °C; mean annual precipitation, 1 80 mm. Dry valley data 
were measured at Vanda Station in the Wright Valley at 78° S; mean annual temperature, -20 °C; 
mean annual precipitation, 45 mm. 

extraterrestrial habitats by the 'astrobiology/exobiology' community (Mahaney et al., 
2001 ;Onoiri etaL, 2004). 

The dry valleys are highly significant sites for studies of ecosystem processes because of 
their relative biological simplicity (low diversity and small terrestrial biomass), and for 
monitoring the effects of environmental changes because they operate under extreme 
(both coldness and dryness) climatic conditions; hence, the Taylor Valley is included in 
the US National Science Foundation network of Long-Term Ecological Research sites 
(e.g. Virginia & Wall, 1999; Wall & Virginia, 1999; Greenland & Kittel, 2002; Hobbie, 
2003; Turner et al., 2003). The terrestrial ecosystem comprises micro-organisms, mosses, 
lichens and a restricted invertebrate community. Given the absence of a conspicuous 



SGM symposium 65 



Soil micro-organisms in Antarctic dry valleys 73 

community of terrestrial autotrophs in these resource-poor ecosystems, understanding 
the carbon and energy supply to terrestrial organisms is an important geomicro- 
biological question. 

ORGANISMS IN THE DRY VALLEYS 

Although there have been no systematic surveys of the diversity of organisms in the dry 
valleys, it is known that the soils of the dry valleys support sparse moss and lichen 
communities (Bargagli et al., 1999), a low diversity of invertebrates (e.g. Treonis et al., 
1999; Stevens & Hogg, 2002) and small microbial communities, the diversity of which is 
largely uncharacterized (Wynn-Williams, 1996; Cowan &C Tow, 2004). Although there is 
little information about the soil micro-organisms, there is evidence of microbial activity 
or potential activity in the soil (Burkins et al., 2002; Treonis et al., 2002; Parsons et ai, 
2004; B. Elberling, E. G. Gregonch, D. W. Hopkins, A. D. Sparrow, P. Novis & L. G. 
Greenfield, unpublished results). Probably the most specialized terrestrial microbial 
community in the dry valleys is the endo/cryptoendolithic community of lichens, algae 
and fungi, living a few millimetres inside relatively coarse-grained rocks. The temp- 
erature and moisture regimes inside the rocks are less hostile and the organisms are 
sustained by light penetrating the translucent mineral grains (Friedmann, 1982; 
Friedmann et al., 1993; Nienow & Friedmann, 1993). However, the most researched 
group of terrestrial organisms is the invertebrate animals - the largest residents of the 
dry valleys. One hundred years after Scott proclaimed the dry valleys sterile, Wilson 
(2002) referred to the soil invertebrates as 'McMurdo's equivalent of elephants and 
tigers'. The invertebrate community includes rotifers, tardigrades, acari, collembola 
and, most notably, nematodes, which are the largest terrestrial consumers in the dry 
valleys (Treonis et al., 1999; Virginia & Wall, 1999; Courtright et al., 2001; Doran et al., 
2002; Stevens & Hogg, 2002, 2003). The most abundant nematode is usually Scottnema 
lindsayae, which feeds on bacteria and algae and typically numbers between about 500 
and 2000 (kg soil) -1 , whilst at particularly favoured sites, its number may rise to several 
thousand kg (Table 1). The enduring presence of relatively large consumers indicates 
active energy processing and nutrient cycling. 

MICROBIAL ACTIVITY IN DRY VALLEY SOILS 

The soils in the dry valleys are characterized by the absence of structure, cohesion and 
moisture, typically alkaline pH, the absence of leaching and localized salt accumu- 
lations (Campbell & Claridge, 2000). The soils contain only small reserves of organic 
matter and nitrogen (Beyer et al., 1999; Burkins et al., 2002; Moorhead et al., 2003; 
Barrett et al., 2005; B. Elberling and others, unpublished results). With the exception of 
moist soils at the lake margins and in transient stream beds, the soils generally contain 
less (typically two orders of magnitude) carbon and nitrogen than the normal range for 
temperate soils (Table 2; Barrett et al., 2002; B. Elberling and others, unpublished 

SGM symposium 65 



74 D. W. Hopkins and others 



u 

-i— < 

u 

-I— ' 

c 

< 

_CD 
"D 

o 
o 



03 

cd 



o 



— 

o 

to 



o 

_>, 

o 

Q_ 
-»— 

o 

LO 

aj 



.g 

■i— • 

03 



03 F 

" E 

Z3 ° 

o » 

-t— 03 



LO 

CD 
-i— ' 
ra 

_Q 
CD 

-I— ' 

CD 
> 



O 

to 

o 

LO 

CD 
_Q 






o 

QJ 
Q_ 

l/> 

E 

03 

"D 

< 

CD 

o3 

75 



O 



Q 




in 




+1 




O0 




c 




03 


0) 


CD 

E 


0J 

1- 


CD 


QJ 


ro 


+■» 


03 


0) 


03 
Q 


> 



o 
-a 

Yi 

CD 

+-> 
03 

i_ 

-Q 
(V 

+- 1 

i_ 

cd 

> 
c 



03 
CD 

■a 

+ 

I 

03 



■a 

03 
CD 

■a 



<D 
> 

~03 
+-» 

,o 



_CD 

"E 

s 

3 



■a 

03 
CD 

Q 



CD 

> 



JD 

03 

E 

<D 



■a 

03 
CD 

a 



(D 

> 



_CD 
03 



T3 
03 
CD 

D 



CD 

> 



CO 






^r 


CD 




r*. 


"v? 


r^ 


PNJ 


rsi 


^ — 


+ 1 


+1 


+1 


Osi 


CD 


00 


O 


CTi 





CTi 

ro 



ro 

en cd 

ro «- 

+1 +1 o 

LT» *— 

00 «- 
LO 



rsj 
CD 

on 



ro r-- 



rsj rsi *— 

+1 +1 +1 

n m oo 

T- CO 

ro «- 
ro 



i_n 

r-- cd 

+1 +1 o 

CD r- 

ro *— 
ro 



CD 

r-- o 

L/l lO N 

+1 +1 +1 

r-- o oo 

O i- 



00 

LO 

+1 O O 

CO 



CTi 

r^ en 



o 



^ 


rsi 


+1 


+1 


^ 


o 


* — 


ro 



r-. 



CD 

+l 
o 



o o 



o 

ro lti 

+1 +1 

CD CD 

o ro 

i_n 



3 

.U 

CD U 
03 b: 



O 





o? 

10 








■Q 


re 






.C 


lo 








13 




oo 
QJ 


03 

E 


03 


00 


o 




b 


3 


H — ' 

03 

E 


8 




43 

U 
CD 


QJ 


CO 


Uj 


Q^ 



+1 
LP) 



LD 

+1 

CTi 



oo 

CD 

~D 

03 

ai 

.03 



on 

a; 

"-^ 

o 



u 

"-i— < 
u 

-i— < 
C 

< 

"D 
O 
O 



CD 

CD 



LO 

'c 

CD 

Q_ 

CD 

U 

00 

"D 

C 

_03 

+- * 
C 
CD 

CD 

M— 

E 

o 

M— 
UP 

'o 

CO 



u -^ 



O, 

CD 
U 
03 



to 



O 

LO 

CD 



CD 

Q_ 

O 



"D 
CD 

-I— ' 

CJ 

_CD 

CD 

ri 

Q) 

_Q 





I 




Q. 








C " 




Q) ^. 




§"8 






.ti CT 




C ^-» 




1 ro "Z. 




O ai 




Z ^ 












c " 




<D X 




§"8 




.t o> 




C "— ' 




+ «* Z 




5 D1 




Z 3. 












C ^ 




0) }-*, 




u> — 









H V> 




+-• 




E>S? 




03 ^ 




^3 








'ut" 




-I— 1 




3 


St" 


00 


o i 


QJ 

"D 

QJ 




_C 


u O 


OO 




Iq 


'gy 


3 


o a> 


CL 


6 M. 


C 
-> 


(/) 








QJ 




_C 




■i— • 




o 




"D 




C 
03 


SjV 


CD 


<D = 


_C 




QJ 


S o» 


LU 


CD O 




■*-• tN 


cd 


03 X 


_Q 


5 D. 




^^^J 


QJ 








03 




"3" 




oo 




+1 




00 




c 




03 




QJ 




^E 




03 




-(— ' 


■ ^_ 


03 





Q 


Ul 



^D 
CO 



CD 
CO 



O 



en 


ro 
rvi 




6 


v^- 


rsi 


+i 


+1 


+1 


o 
o 


CD 


i_n 
rsi 


r^- 


^~ 


r^ 



^ 


CD 


00 


^~ 


^r 


r> 


b 


b 


CD 


+i 


+i 


+1 


t— 


r~- 


"st" 


^ 


rvi 


CD 


b 


t— 


00 



ro 




o 


CD 


CO 


CO 


ro 


^f- 


fN 


+1 


+1 


+1 


CD 


^r 


CTi 


CO 

LP) 


^ 


LO 

rsi 
ro 



o 


^~ 


o 


ro 


CD 


CD 


b 


b 


rsi 


+1 


+i 


+1 


CO 


ro 


cr\ 


ro 


CD 


oS 


< — 


t— 


rsi 



CTi 

b 


CD 

ro 


rsj 


+1 


+1 


+1 


CTi 


r^ 


r^> 


«- 


LTl 


rsj 
CD 



QJ 




Q_ 


C 


g 


O 


00 


cn 


i 


>^ 


— 


o 


31 


Q_ 



03 



QJ 
03 



SGM symposium 65 



Soil micro-organisms in Antarctic dry valleys 75 



0-3 i 



0-2 



01 



"o 



o 



s 



c 
o 
'.p 

1 

to 

2 

'5 



Hill-slope 



0-3 i Polygon 



2 



01 



03 -i Lake margin 



02 



0-1 



^^ 




Basal Water Glucose Alanine Mat 



Fig. 2. Soil respiration over 1 20 h for soils from different landscape units in the Garwood Valley, 
incubated at 5 °C in the laboratory in unamended soil (basal) and in response to addition of water, 
glucose, alanine and samples of the microbial mat from the lake margin. Each bar is the mean of 
three replicates and the bars are ± sem. 

results). Nevertheless, in situ microbial respiration is consistently measured as C0 9 
efflux at the soil surface (Parsons et at., 2004; B. Elberling and others, unpublished 
results) and Barrett et al. (2005) reported the presence of a significant amount of labile 
soil organic matter. Assuming a steady state, the estimated turnover times for organic 
matter (mass/flux) in the dry valley soils is remarkably fast: approximately 23 years in 
the Taylor Valley (Burkins et al., 2002) and 30-123 years in the Garwood Valley, depend- 
ing on soil type (B. Elberling and others, unpublished results). By contrast, turnover 
times in most temperate region soils are generally several centuries or even millennia 
(Hopkins & Gregorich, 2005). In the absence of a conspicuous community of ter- 
restrial autotrophs in the dry valleys, the rapid turnover of organic carbon suggests 
either that the soil organic carbon is not in a steady state or the presence of modern 
additions of labile organic matter (Barrett et al., 2005). The presence of labile organic 
matter is supported by in situ measurements of soil respiration (B. Elberling and others, 
unpublished results). In the Garwood Valley, soil respiration at the lake margin was 
greater than that in the drier soils from elsewhere in the valley (Fig. 2). This difference 



SGM symposium 65 



76 D. W. Hopkins and others 



16 



15 

-Q 

1 1 
< crs 



12- 



8- 



4- 



Lower moraine 






- 


+ Hill-slope 




♦ 


"■".,■* 


Delta 


Polygon 




"~ Sand dunes 


i 


i 


Upper moraine 
r i 



20 



40 



60 



80 



100 



Total organic C : total N ratio 



Fig. 3. Relationship of in situ soil respiration and organic C : N ratio for different soils in the Garwood 
Valley (B. Elberling and others, unpublished results). 

was not related solely to the water content (Fig. 2) and experimental additions of glu- 
cose and alanine and of detritus from the lake margin microbial mats (C:N ratio of 
10-12) indicated that soil respiration is limited by both carbon and nitrogen (Fig. 2). In 
the case of the addition of microbial-mat material, the soil amendment may also have 
contained viable organisms that may have contributed to the respiratory response. The 
influence of nitrogen on soil respiration is consistent with in situ measurements of soil 
C0 2 respiration rates (Fig. 3), which indicate a correlation between the ratio of the total 
organic C : total N and soil C0 2 effluxes, with the largest effluxes associated with the 
smallest ratios of organic C : total N (B. Elberling and others, unpublished results). 

SOURCES OF RESOURCES IN DRY VALLEY SOILS 

For most terrestrial habitats, there is little debate over the source of organic resources 
for soil micro-organisms. In the dry valleys, however, the sparse and discontinuous 
presence of large photoautotrophs means that alternative sources of organic resources 
need to be considered. The possible sources of fixed carbon are: (i) modern autotrophic 
activity in situ, which would include the cryptoendolithic communities, mosses, 
cyanobacteria and heterotrophic algae in the soils, and autotrophic bacteria such as 
nitrifiers; (ii) ancient in situ, or 'legacy', organic deposits from a time when the climate 
was warmer and conditions were wetter, and organic lake sediments accumulated on 
the surfaces that developed into the modern dry valleys; (iii) spatial subsidies from the 
coastal regions, where there are abundant marine and ornithogenic deposits carried 
into the dry valleys by aeolian dispersal. This would represent relatively long-range 
transport of up to about 50 km inland; and (iv) spatial subsidies from the margins of 
modern lakes, where microbial mats accumulate under favourable conditions. This 



SGM symposium 65 



Soil micro-organisms in Antarctic dry valleys 77 



c 
o 

'S3 

V 

a. 

in 







2 
to 



■10 



-15 



-20 



"? -5 






-10 



-15 



-20 







£ -5 



4© 



-10 



-15 



-20 



"En 

6! 60 Marine 



150 200 



70 










Lacustrine 



Endolitr 





51 







Bonney Riegel 

33 km from Ross Sea 



30 -25 



-20 -15 -10 
S*ac (%o) 





Marine 







66 









Lacustrine 



64 



Endolifr 



60 



58 



1P0 



62 



130 
210 



Lake Hoare 

15 km from Ross Sea 



-30 -25 -20 -15 -10 





Marine 







26 25 








25 


Lacustrine 



Endolitf 



1H 



■bu 



= 8 



100 



LakeFryxell 

12 km from Ross Sea 



-30 -25 -20 -15 

S'K (%o) 



-10 



,3 C depletion- 



Fig. 4. Isotopic signals for soils from different locations at three sites in the Taylor Valley, shown with 
boxes covering the range of isotopic signals for materials from the marine and lacustrine environments 
and for cryptoendolithic materials. The sites are each shown by a number indicating the altitude in m 
above sea level. Data extracted and redrawn from Burkins etal. (2000) with permission from The 
Ecological Society of America. 



SGM symposium 65 



78 D. W. Hopkins and others 



10 000 



o 

t! o <1> 



c 



1 000 




100 



Lake edge Polygons Hill-slope 
Landscape unit 



Fig. 5. Potential nitrification activities for soils from different landscape units in the Garwood 
Valley. The soils were incubated in a shaking incubator at 5 °C in a suspension containing 0-25 mM 
(NH 4 ) 2 S0 4 as described by Hopkins et al. (1 988). 



material may subsequently be dispersed onto the surrounding soils, representing 
relatively short-range aeolian dispersal. 

These sources are not necessarily mutually exclusive and, depending on particular site 
details, they may all contribute to the overall carbon cycling, albeit in different 
proportions. There have been few detailed investigations of the provenance of organic 
matter in the dry valley soils (Burkins et al., 2000; B. Elberling and others, unpublished 
results) and it is not therefore possible to partition soil biological activity between 
different resource drivers. Nevertheless, from the evidence available, it is possible to 
evaluate the likelihood of some of the different sources of organic resources being 
important contributors to carbon cycling. 

Burkins et al. (2000) used the carbon and nitrogen isotopic signatures of soils and 
organic materials from within and around the dry valleys to investigate the provenance 
of organic matter in the Taylor Valley (Fig. 4). These studies did not provide evidence of 
substantial ornithogenic sources of soil organic matter. They did show that, depending 
on altitude, organic matter with cryptoendolithic and lacustrine signals was present, 
with the cryptoendolithic material occurring at greater altitude, i.e. at drier and colder 
sites further from the valley floor and lakes. However, the conclusions were not clear-cut 
because, counter-intuitively, the strongest evidence for marine-derived organic matter 
occurred at the most inland site (33 km from the coast) examined, rather than at sites 
closer to the coast. 



SGM symposium 65 



Soil micro-organisms in Antarctic dry valleys 79 

The data shown in Fig. 5 provide evidence for a chemoautotrophic contribution to 
organic matter input that is supported indirectly by the soil NOT concentrations (Table 
2). However, given the low productivity (mol NH4 oxidized by autotrophic nitrifying 
bacteria) -1 (Wood, 1986) and the fact that the measurements in Fig. 5 were obtained 
under laboratory conditions where NH4 nitrogen was not limiting, it is not likely that 
nitrification is a substantial source of soil organic carbon. 

The 'legacy' model for the Taylor Valley proposes that the origin of much of the soil 
organic carbon is material laid down in ancient lake beds. It is supported indirectly by 
geomorphological data on landscape origin (e.g. Burkins et aL, 2000; Higgins et aL, 
2000), in particular the presence of the palaeolake Washburn in the Taylor Valley 
between 22 800 and 8500 years ago, and by some isotopic signatures in soils (Burkins 
et aL, 2000). However, it is difficult to reconcile the relatively rapid estimated rates of 
carbon turnover with the persistence of an ancient but active reserve of carbon in the 
soil, and Barrett et aL (2005) now discuss multiple organic matter sources. 

The 'legacy' model may be most relevant in large expanses of dry ground remote from 
lakes, in valleys with relatively small areas covered by lakes and where the geology is not 
conducive to endoliths. However, in smaller valleys with a larger proportion covered by 
lakes, at lake and stream margins and in ephemeral stream beds in both small and large 
valleys, which are recognized as hot-spots of biological activity (Greenfield, 1998; 
McKnight et aL, 1999; Moorhead et aL, 2003), modern aquatic-derived detritus may 
assume greater importance (Moorhead et aL, 2003; B. Elberling and others, unpub- 
lished results). The lakes and ponds in the dry valleys are the focus of productivity by 
carbon- and nitrogen-fixing cyanobacteria and eukaryotic algae (Olson et aL, 1998; 
Hawes & Schwarz, 1999), representing draw-down and concentration of resources 
from the atmosphere. Periodicity in lake and stream level may then cause exposure of 
resource-rich cyanobacterial and algal detritus and wash-up of nitrogen-rich foams 
originating from decomposition of algal biomass within the lake. This conceptual 
model for nutrient cycling in such dry valleys is summarized in Fig. 6. Wilson (1965), 
Parker et aL (1982) and Greenfield (1998) suggested a linkage between aquatic product- 
ivity and nitrogen fixation by cyanobacteria, with terrestrial nutrient cycling consistent 
with the spatial-subsidy model (Fig. 6). Whilst this model remains largely unparameter- 
ized, such redistribution has been noted in the field (Wilson, 1965; Parker et aL, 1982; 
Nienow & Friedmann, 1993; Moorhead et aL, 2003) and is likely to produce a gradient 
in soil carbon stocks, declining with distance from lakes, and vertical concentration 
gradients in soil profiles (B. Elberling and others, unpublished results). 

The redistribution of lacustrine microbial mat from lake shores to soils may, as 
suggested above, also lead to redistribution of viable organisms to the soils, although 

SGM symposium 65 



80 D. W. Hopkins and others 



c 
o 
♦-* 

"a. 



T3 

O 

i— * 

o 

I 

ID 



H3 



l/V 



Aquatic 

photosynthesis 
and N fixation 



Windblown 





Fig. 6. Conceptual model for resource transfer in one of the small Antarctic dry valleys. 



M 1 2 3 19 20 21 37 38 39 55 56 57 73 74 75 +ve -ve -ve M 




Primer dimer 



PCR product 



Fig. 7. Gel showing amplification of methanogenesis-specific methyl co-enzyme reductase (MCR) in 
soils from the Garwood Valley by using the ME1/ME2 primer set (Hales et a/., 1996). Lanes 37-39 and 
55-57 correspond to the hill-slope and polygon soils in Table 2, respectively. 



few studies provide direct evidence for this. However, we have preliminary indirect 
evidence from the incidence of methanogenesis and methanogens in the Garwood 
Valley We have detected substantial in situ methanogenesis as CH 4 efflux at the surface 
in soils within 1 m of lake margins in the range of 1-3-69 mgm _2 day _1 during the 
2002-2003 austral summer (E. G. Gregorich, D. W. Hopkins, B. Elberling, A. D. 
Sparrow, P. Novis & L. G. Greenfield, unpublished results). These rates are, incidentally, 
comparable to CH 4 emissions from temperate and sub-Arctic wetlands (Moore & 
Knowles, 1989, 1990). By using PCR, we have detected the gene for the methanogenesis- 
specific methyl co-enzyme reductase (MCR) using ME1/ME2 methanogenic primers 
(Hales et al., 1996) in soils between 20 and 50 m from lake margin (Fig. 7). DNA 
amplified by using this primer set suggests the presence of methanogenic bacteria in the 
dry valley soils that may have arisen from the wet, organic-rich lake margins (Table 2). 
These data provide circumstantial evidence for transfer of methanogens from the lake 



SGM symposium 65 



Soil micro-organisms in Antarctic dry valleys 81 

margin to the surrounding dry soils, where they were most unlikely to be active under 
the dry and well-aerated surface conditions. 

CONCLUDING REMARKS 

Understanding the provenance of resources for soil micro-organisms in nutrient- and 
energy-poor ecosystems, such as polar deserts, requires consideration of a wider range 
of possible sources than in most other ecosystems. When the indigenous stocks of soil 
organic carbon and nitrogen are as low as occurs in the dry valleys, even very modest 
inputs from other sources may represent a significant subsidy. Whilst there is a parallel 
in temperate ecosystems at the early stages of primary successions (Hodkinson et al., 
2002), for most ecosystems, in situ primary production is usually, and correctly, 
assumed to be the energy source for the below-ground microbial community (Hopkins 
& Gregorich, 2005). The Antarctic dry valleys may represent ecosystems dependent to a 
greater extent on spatial and/or temporal subsidies than other terrestrial ecosystems. 
Clearly, to better understand the relative contributions of resources from the different 
sources, parameterization of the models is necessary Our priority is the spatial-subsidy 
model (Fig. 6) and this will require measurements of the dispersal of lake-derived 
organic matter and organisms and investigation of the periodicity of production and 
dispersal in relation to lake level changes. 

The dry valleys may also provide an opportunity to examine whether the biological 
diversity of the soil community is causally related to function. This question has 
emerged recently as a highly topical ecological issue. However, for most soils, the 
biological diversity is so large, particularly amongst the heterotrophs, that it is difficult 
to quantify and relate to the key carbon-cycling processes. In resource-poor systems, 
where the diversity is likely to be lower, there may be a realistic opportunity to examine 
soil biodiversity and function. 

ACKNOWLEDGEMENTS 

We are grateful to Antarctica New Zealand, the UK Natural Environment Research Council, the 
Carnegie Trust for the Universities of Scotland, the Royal Society of London, the Transantarctic 
Association, the Danish Natural Science Research Council, Agriculture and Agri-Food Canada 
and the University of Canterbury (Christ church, NZ) for support for different parts of the 
research outlined here. In addition, we express our gratitude to David Wardle (Lincoln, NZ) and 
collaborators in the US Antarctic Program supported by the US National Science Foundation, 
in particular Diana Wall, Ross Virginia, Jeb Barrett and Byron Adams, for access to some of 
their data and stimulating discussions. We are grateful to Lorna English, Patrick St Georges and 
M. R. Nielsen for technical assistance, and the numerous staff of Antarctica New Zealand for 
logistic support. Finally, D. W. H. wishes to acknowledge the influential role of the late David 
Wynn-Williams for an introduction to biological research in Antarctica. 

SGM symposium 65 



82 D. W. Hopkins and others 



REFERENCES 



Bargagli, R., Smith, R. I. L, Martella, L., Monaci, R, Sanchez-Hernandez, J. C. & 

Ugolini, F. C. (1999). Solution geochemistry and behaviour of major and trace 

elements during summer in a moss community at Edmonson Point, Victoria Land, 

Antarctica. AntarctSci 11,3-12. 
Barrett, J. E., Virginia, R. A. & Wall, D. H. (2002). Trends in resin and KCI-extractable soil 

nitrogen across landscape gradients in Taylor Valley, Antarctica. Ecosystems 5, 

289-299. 
Barrett, J. E., Virginia, R. A., Parsons, A. N. & Wall, D. H. (2005). Potential soil organic 

matter turnover in Taylor Valley, Antarctica. ArctAntarctAlp Res 37, 1 08-1 1 7. 
Beyer, L, Bockheim, J. G., Campbell, I. B. & Claridge, G. G. C. (1999). Genesis, properties 

and sensitivity of Antarctic Gelisols. AntarctSci 11, 387-398. 
Burkins, M. B., Virginia, R. A., Chamberlain, C. P. & Wall, D. H. (2000). Origin and 

distribution of soil organic matter in Taylor Valley, Antarctica. Ecology 81, 2377- 

2391. 
Burkins, M. B., Virginia, R. A. & Wall, D. H. (2002). Organic carbon cycling in Taylor Valley, 

Antarctica: quantifying soil reservoirs and soil respiration. Global Change Biol 7, 

113-125. 
Campbell, I. B. & Claridge, G. G. C. (2000). Soil temperature, moisture and salinity 

patterns in Transantarctic Mountain cold desert ecosystems. In Antarctic Ecosystems: 

Models for Wider Ecological Understanding, pp. 233-240. Edited by W. Davidson, 

C. Howard-Williams & P. Broady. Christchurch, New Zealand: Caxton Press. 
Courtright, E. M., Wall, D. H. & Virginia, D. A. (2001). Determining habitat suitability for 

soil invertebrates in an extreme environment: the McMurdo Dry Valleys, Antarctica. 

AntarctSci 13, 9-17. 
Cowan, D. A. & Tow, L. A. (2004). Endangered Antarctic environments. Annu Rev 

Microbiol 58, 649-690. 
Doran, P. T., Priscu, J. C, Lyons, W. B. & 10 other authors (2002). Antarctic climate 

cooling and terrestrial ecosystem response. Nature 415, 517-520. 
Friedmann, E. I. (1982). Endolithic microorganisms in the Antarctic cold desert. Science 

215, 1045-1053. 
Friedmann, E. I., Kappen, L., Meyer, M. A. & Nienow, J. A. (1993). Long-term 

productivity in the cryptoendolithic microbial community of the Ross Desert, 

Antarctica. MicrobEcol 25, 51-69. 
Greenfield, L. G. (1998). Nitrogen in soil Nos toe mats: foams, release and implications for 

nutrient cycling in Antarctica. N ZJNat Sci 23, 101-107. 
Greenland, D. & Kittel, T. G. F. (2002). Temporal variability of climate at the US Long-Term 

Ecological Research (LTER) sites. Clim Res 19, 213-231 . 
Hales, B. A., Edwards, C, Ritchie, D. A., Hall, G., Pickup, R. W. & Saunders, J. R. (1996). 

Isolation and identification of methanogen-specific DNA from blanket bog peat by 

PCR amplification and sequence analysis. Appl Environ Microbiol 66, 668-675. 
Hawes, I. & Schwarz, A. M. (1999). Photosynthesis in an extreme shade environment: 

benthic microbial mats from Lake Hoare, a permanently ice-covered Antarctic lake. 

JPhycol 35, 448^59. 
Higgins, S. M., Denton, G. H. & Hendy, C. H. (2000). Glacial geomorphology of Bonney 

Drift, Taylor Valley, Antarctica. GeogrAnn SerA Phys Geogr 82, 365-389. 
Hobbie, J. E. (2003). Scientific accomplishments of the long term ecological research 

program: an introduction. Bioscience 53, 17-20. 



SGM symposium 65 



Soil micro-organisms in Antarctic dry valleys 83 



Hodkinson, I. D., Webb, N. R. & Coulson, S. J. (2002). Primary community assembly on 

land - missing stages: why are the heterotrophic organisms always there first? J Ecol 

90, 569-577. 
Hopkins, D. W. & Gregorich, E. G. (2005). Carbon as a substrate for soil organisms. 

In Biodiversity and Function in Soils. Edited by R. D. Bardgett, M. B. Usher & 

D. W. Hopkins. Cambridge: Cambridge University Press (in press). 
Hopkins, D. W., O'Donnell, A. G. & Shiel, R. S. (1988). The effect of fertilisation on soil 

nitrifier activity in experimental grassland plots. Biol Ferti I Soils 5, 344-349. 
Mahaney, W. C, Dohm, J. M., Baker, V. R., Newsom, H. E., Malloch, D., Hancock, 

R. G. V., Campbell, I., Sheppard, D. & Milner, M. W. (2001). Morphogenesis of 

Antarctic paleosols: Martian analogue. Icarus 154, 1 13-130. 
McKnight, D. M., Niyogi, D. K., Alger, A. S., Bomblies, A., Conovitz, P. A. & Tate, C. M. 

(1 999). Dry valley streams in Antarctica: ecosystems waiting for water. Bioscience 49, 

985-995. 
Moore, T. R. & Knowles, R. (1 989). The influence of water table levels on methane and 

carbon dioxide emissions from peatland soils. Can J Soil Sci 69, 33-38. 
Moore, T. R. & Knowles, R. (1990). Methane emissions from fen, bog and swamp 

peatlands in Quebec. Biogeochemistry 11, 45-61 . 
Moorhead, D. L, Barrett, J. E., Virginia, R. A., Wall, D. H. & Porazinska, D. (2003). 

Organic matter and soil biota of upland wetlands in Taylor Valley, Antarctica. Polar 

Biol 26,567-576. 
Nienow, J. A. & Friedmann, E. I. (1993). Terrestrial lithophytic (rock) communities. In 

Antarctic Microbiology, pp. 343^412. Edited by E. I. Friedmann. New York: Wiley. 
Olson, J. B., Steppe, T. F., Litaker, R. W. & Paerl, H. W. (1998). N 2 -fixing microbial 

consortia associated with the ice cover of Lake Bonney, Antarctica. Microb Ecol 36, 

231-238. 
Onofri, S., Selbmann, L, Zucconi, L. & Pagano, S. (2004). Antarctic microfungi as models 

for exobiology. Planet Space Sci 52, 229-237. 
Parker, B. C, Simmons, G. M., Jr, Wharton, R. A., Jr, Seaburg, K. G. & Love, F. G. (1982). 

Removal of organic and inorganic matter from Antarctic lakes by aerial escape of 

bluegreen algal mats. J Phycol 18, 72-78. 
Parsons, A. N., Barrett, J. E., Wall, D. H. & Virginia, R. A. (2004). Soil carbon dioxide flux 

in Antarctic dry valley ecosystems. Ecosystems 7, 286-295. 
Stevens, M. I. & Hogg, I. D. (2002). Expanded distributional records of Collembola and 

Acari in southern Victoria Land, Antarctica. Pedobiologia 46, 485-495. 
Stevens, M. I. & Hogg, I. D. (2003). Long-term isolation and recent range expansion from 

glacial refugia revealed for the endemic springtail Gomphiocephalus hodgsoni 

from Victoria Land, Antarctica. Mol Ecol 12, 2357-2369. 
Treonis, A. M., Wall, D. H. & Virginia, R. A. (1999). Invertebrate biodiversity in Antarctic 

dry valley soils and sediments. Ecosystems 2, 482-492. 
Treonis, A. M., Wall, D. H. & Virginia, R. A. (2002). Field and microcosm studies of 

decomposition and soil biota in a cold desert soil. Ecosystems 5, 1 59-170. 
Turner, M. G., Collins, S. L., Lugo, A. L., Magnuson, J. J., Rupp, T. S. & Swanson, F. J. 

(2003). Disturbance dynamics and ecological response: the contribution of long- 
term ecological research. Bioscience 53, 46-56. 
Virginia, R. A. & Wall, D. H. (1999). How soils structure communities in the Antarctic dry 

valleys. Bioscience 49, 973-983. 
Wall, D. H. & Virginia, R. A. (1999). Controls on soil biodiversity: insights from extreme 

environments. Appl Soil Ecol 13, 1 37-1 50. 



SGM symposium 65 



84 D. W. Hopkins and others 



Wilson, A. T. (1965). Escape of algae from frozen lakes and ponds. Ecology 46, 376. 

Wilson, E. O. (2002). The Future of Life. New York: Alfred A. Knopf. 

Wood, P. M. (1986). Nitrification as a bacterial energy source. In Nitrification (Special 

Publication of the Society for General Microbiology no. 10), pp. 39-62. Edited by 

J. I. Prosser. Oxford: IRL Press. 
Wynn-Williams, D. D. (1996). Antarctic microbial biodiversity: the basis of polar ecosystem 

processes. Biodivers Conserv 5, 1 271-1 293. 



SGM symposium 65 



New insights into bacterial 
cell-wall structure and physico 
chemistry: implications for 
interactions with metal ions 
and minerals 

V. R. Phoenix, A. A. Korenevsky, V. R. F. Matias and 
T. J. Beveridge 

Department of Molecular and Cellular Biology, College of Biological Science, 
University of Guelph, Guelph, Ontario, Canada N1G 2W1 



INTRODUCTION 

Prokaryotes are the Earth's smallest life form and, yet, have the largest surface 
area : volume ratio of all cells (Beveridge, 1988, 1989a). They are also the most ancient 
form of life and have persisted on Earth for at least 3-6 x 10 9 years, even in some of the 
most extreme environments imaginable, such as the deep subsurface. Most of these 
early primitive (and today's modern) natural environments possess reasonably high 
amounts of metal ions that are capable of precipitation under suitable pH or redox 
conditions. Deep-seated in such geochemical situations is the likelihood of suitable 
interfaces that lower the local free energy, so that interfacial metal precipitation is 
promoted. Bacteria, being minute and having highly reactive surfaces (interfaces), are 
exquisitely efficient environmental particles for metal-ion adsorption and mineral 
nucleation. Metal ions interact with available reactive groups (or ligands) on the 
bacterial surface and precipitates grow as environmental counter-ions interact with 
more and more metal at the site (Beveridge & Murray, 1976, 1980; Beveridge et al., 
1982; Ferris & Beveridge, 1986; Fortin et al., 1998). Once formed, these precipitates are 
under the influence of natural geochemical and additional microbially mediated 
conditions (Lee & Beveridge, 2001) that instigate the development of fine-grain 
minerals, usually via dehydration, so that crystalline phases are eventually developed 
(Beveridge et al., 1983). These minerals commence as so-called 'nano-mineral phases' 
and grow with time to become larger and larger. This bacterially induced mineral- 
ization is probably the natural phenomenon that so encases some cells in fine-grain 
minerals that they die and become bona fide 'microfossils' (Ferris et al., 1988). In 
ancient times, these mineral-encased prokaryotes, enduring low-temperature 

SGM symposium 65: Micro-organisms and Earth systems - advances in geomicrobiology. 
Editors G. M. Gadd, K. T. Semple & H. M. Lappin-Scott. Cambridge University Press. ISBN 521 86222 1 ©SGM 2005 



86 V. R. Phoenix and others 



metamorphic geological conditions, survived as microfossils that are still existent in 
such very old Precambrian formations as the ~ 2-0 x 10 9 years Gun Flint Chert, north of 
Lake Superior in Canada. 

It is certain that bacterial surfaces interact with environmental metal ions and can 
provide nucleation sites for mineral precipitation, but it has been extremely difficult to 
study such systems with high precision on a cell-to-cell basis, even though a wide base 
of techniques exists (Beveridge et al., 1997). This is because the cells are extremely small 
and the interactive structures even smaller, and because the reactive sites responsible for 
adsorbing metals retain their reactivity only over a certain range of pH and £ h , which is 
difficult to monitor over microscale distances. This chapter will outline our new 
advances in elucidating the structure of bacterial surfaces, as well as the determination 
of reactive sites via potentiometric examination. Additional techniques, such as zeta 
potentials and hydrophobic/hydrophilic determinations, will also be given, so as to 
provide a more general picture of how surface structure and surface reactivity affect the 
natural physico-chemical traits of bacteria. 

NEW OBSERVATIONS OF BACTERIAL SURFACES BY USING 
CRYO-TRANSMISSION ELECTRON MICROSCOPY (cryoTEM) 

By using conventional fixation, embedding and thin-section techniques, the TEM 
observation of bacteria has been a most powerful method for elucidating the internal 
cytoplasmic organization and the juxtaposition of encompassing envelope layers of 
cells (Beveridge, 1989b; Koval 6c Beveridge, 1999). Yet, there are many drawbacks to 
such conventional techniques, as the cells are fixed chemically by using harsh fixatives 
(such as glutaraldehyde and osmium tetroxide) and dehydrated before embedding in a 
plastic resin for thin-sectioning (Beveridge et al., 2005). Essential lipids are extracted, 
proteins are denatured and nucleic acids are condensed artificially ... the cells are a 
spectre of their former selves. Clearly, the images of such cells have been beneficial to 
our initial perception of the structural organization of prokaryotic cells (Beveridge, 
1989b), but hydration (which these embedded cells no longer have) is a necessary 
prerequisite for the maintenance of native structure. With proper expertise, care and 
equipment, it is now possible through the use of cryoTEM to obtain a better and more 
natural view of bacteria (Dubochet et al., 1983; Umeda et al., 1987; Beveridge & 
Graham, 1991 ; Paul et al., 1993) . 

Freeze-substitution 

One cryoTEM technique that became popular during the 1980s and 1990s was freeze- 
substitution (Hobot et al., 1984; Umeda et al., 1987; Graham & Beveridge, 1990, 1994). 
Here, cells are frozen rapidly at approximately -196 °C so as to vitrify them in amor- 
phous ice, which is not crystalline and is a kind of glass (Koval 6c Beveridge, 1999; 

SGM symposium 65 



Cell-wall structure and physico-chemistry 87 




Fig. 1. Thin-section image of a freeze-substituted wall of a Gram-positive B. subtilis cell, showing three 
distinct regions in the cell wall: the inner (1 ), middle (2) and outer (3) regions that correspond to cell- 
wall turnover and to the available reactive groups within the cell-wall network. Bar, 50 nm. 

Beveridge et al., 2005). Hence, the cells are physically 'fixed', as there is no time during 
freezing for structure to degrade; in fact, freezing occurs within milli- to microseconds 
and all molecular motion is stopped. If the bacteria are thawed, they come back to the 
living state and continue to grow and divide. This is a clear-cut measure of how well 
the cells are preserved! Once vitrification is accomplished and cellular structure is 
preserved, the temperature is raised from -196 to -80 °C and the cells are put into a 
freeze-substitution mixture. This consists of a cryogenic fluid (such as acetone) con- 
taining a chemical fixative (osmium tetroxide), a heavy-metal stain (uranyl acetate) and 
a molecular sieve (for trapping water) (Graham & Beveridge, 1990, 1994). At this 
low temperature, the cells do not melt; instead they remain vitrified and structure is 
preserved. The cellular and surrounding ice (water) sublimes and is trapped in the 
molecular sieve. Eventually, the specimen is dehydrated thoroughly and fixed in this 
freeze-substitution mixture, with the structure maintaining many of its native features. 
Now, it can be embedded in plastic resin, cured and thin-sectioned for viewing by TEM. 



Freeze-substitution and Gram-positive (Bacillus subtilis) 
cell walls 

The results of freeze-substitution are breathtaking, especially when examining the 
surface structures (Fig. 1). The fabric of Gram-positive cell walls is no longer featureless 
(as seen in conventional embeddings; Beveridge, 2000), but is a tripartite structure 
(Fig. 1). The region associated with the plasma membrane (immediately above the 
bilayer) is highly contrasted, due to the acquisition of large quantities of heavy-metal 
stain. The middle region is lightly contrasted, as the wall polymers (mainly peptido- 
glycan) are stretched almost to breaking point, so that the mass : volume ratio is much 
reduced compared with other wall regions. The outermost region consists of thin fibres 



SGM symposium 65 



88 V. R. Phoenix and others 



that extend into the external milieu. This tripartite cell-wall structure is compatible 
with current models of cell-wall turnover (Graham & Beveridge, 1994; Beveridge, 
2000). 

More important for this chapter is the fact that the polymeric structure of the wall has 
been preserved by freeze-substitution and the available reactive sites within the wall 
have been 'decorated' with the heavy-metal stain, so that we obtain a clear picture of 
where the reactive sites reside. Many sites are in the region immediately apposed to the 
membrane as, here, new wall polymers are being extruded and compacted via 
penicillin-binding proteins in the membrane. As an area of de novo wall assembly, 
where new polymers are being cross-linked into the pre-existing wall fabric, many 
reactive groups are available for decoration in this region, as the mass is great and the 
availability of reactive groups is high. However, the middle region is different. It is 
the area in the cell wall that resists turgor pressure and maintains the integrity of the 
cell. It has to be highly cross-linked to hold the cell together under such a pressure load 
and is considerably stretched. This region, then, has few reactive groups left available 
for interaction with the heavy-metal stain, as most have been used to cross-link the 
network together for strength. The outermost region is an area where the cell wall is 
being broken down by the wall's constituent autolysins. Covalent bonds are being 
broken and new reactive groups are being made. This region probably has little mass (as 
it is being shed during cell-wall turnover), but an excess of reactive sites that are 
decorated readily by the stain. 

Frozen hydrated thin sections 

A more difficult and therefore less used cryoTEM technique is the use of frozen 
hydrated sections (Dubochet et al., 1983, 1988). This technique requires skill and 
perseverance. As in freeze-substitution, cells are vitrified, but now, instead of processing 
the cells so that conventional, plastic thin sections are obtained, this frozen material is 
put immediately into a 'cryo'-ultramicrotome and thin-sectioned. The cells are 
sectioned whilst vitrified and the frozen sections are mounted immediately into a 'cryo'- 
specimen holder and inserted into the 'cryo'-chamber of a 'cryo'TEM. We emphasize 
'cryo' because the temperature must be maintained at between -196 and -140 °C during 
all manipulations, otherwise the amorphous ice embedding the cells will become 
crystalline and ruin the native structure to be observed. No chemical fixatives or heavy- 
metal stains are used during the entire process and, as the sections remain vitrified, all 
cellular macromolecules and polymers remain in a hydrous state. 

Frozen hydrated Gram-positive (6. subtilis) cell walls 

One of the advantages of the cryo-sectioning technique is that no artificial chemical 
fixatives or heavy-metal stains need to be used. However, implicit in the use of TEM is 

SGM symposium 65 



Cell-wall structure and physico-chemistry 89 

that the specimen must possess enough density to efficiently scatter high-voltage 
electrons from the electron gun of the microscope (i.e. the electron potential is typically 
approx. 100000-200000 eV). Biomaterials, once thin-sectioned, rarely have enough 
density to effectively scatter such high-powered electrons, as the thin sections are only 
~50 nm thick and the biomatter possesses only low-atomic-number elements (such as 
H, C, O, N etc.). This is the primary reason that conventional and freeze-substitution 
thin sections use stained material; the heavy-metal stains increase the density (or the 
mass : volume ratio) of the specimen so that the contrast becomes great enough for 
the cells to be visualized easily (Beveridge et al., 2005). Frozen hydrated thin sections of 
unstained bacteria do not have this luxury, as they cannot be stained once sectioned; the 
staining fluids would immediately freeze over the sample and obliterate the structure of 
the cells. 

Clearly, then, these frozen hydrated sections of bacteria are difficult to see, as their 
contrast is close to that of the surrounding vitrified ice. For this reason, we rely on the 
inherent phase function of the lenses of the cryoTEM and use phase contrast to help 
imaging by underfocusing to see the bacteria. Certain microscopes (e.g. those with 
energy filters) can also derive more additional contrast for the specimen. However, even 
then, there are additional problems in visualizing frozen sections. The energy of the 
electron beam is often high enough to locally increase the specimen's temperature, 
resulting in the formation of crystalline ice (from amorphous ice) and, eventually, in ice 
sublimation. As bacteria are excellent nucleation particles (remember how efficient they 
are at forming fine-grain minerals), the amorphous-to-crystalline phase transition of 
ice frequently occurs on the bacteria and their structure is obscured. Furthermore, as 
the specimen is kept so cold, the frozen sections act as 'cold traps' for the condensation 
of extraneous molecules within the high vacuum of the microscope column, thereby 
often contaminating the structure of the specimen. With all these associated problems, 
it is a wonder that frozen hydrated sections of bacteria can be imaged, but they can and 
they are extraordinary (Fig. 2). 

These images differ from what we see in freeze-substitution images. Remember, now we 
have no heavy-metal stains to assist contrast and must rely on the inherent density 
imparted on the cell by the constituent atoms within its molecules. Proteins will be 
discerned more readily from the surrounding ice than, say, carbohydrates, because they 
are usually larger, contain nitrogen and (sometimes) sulfur and tend to fold more 
tightly. Most importantly, all cellular constituents remain hydrated and, therefore, not 
artificially condensed because of a dehydration regimen. The ribosomes are larger and 
more robust and can be seen because of their high concentration of protein and 
rRNA (here, the phosphorus adds additional contrast) (Fig. 2). Even the bilayer of 
the membrane can be seen, because of the inherent contrast of the phosphorus in the 

SGM symposium 65 



90 V. R. Phoenix and others 




wfLJ *■■ , - -''-' ■ 



Fig. 2. Frozen hydrated thin section of a B. subtilis cell wall, showing the plasma membrane (PM), the 
inner-wall zone (IWZ) and the outer-wall zone (OWZ). Here, the IWZ corresponds to region 1 and the 
OWZ to region 2 of the freeze-substituted wall in Fig. 1 . Region 3 is not seen. This IWZ and OWZ have 
different dimensions from the regions in Fig. 1 and are visualized entirely, due to the density of the 
constituent macromolecules. Bar, 50 nm. Reproduced from Matias & Beveridge (2005) with 
permission from Blackwell Publishing. 

phospholipids. Most important for this chapter, though, is the cell wall. Here, we get a 
clear idea of the mass distribution within the Gram-positive wall of B. subtilis (Fig. 2). 
Immediately above the bilayered membrane is a low-density space with little contrast. 
Experiments have shown that this is a periplasmic space (Matias & Beveridge, 2005). 
Above this is a more densely contrasted region that represents the peptidoglycan- 
teichoic acid network of the wall. Unlike freeze-substitutions, there is not an outermost 
fibrous region (cf. Figs 1 and 2). 

Correlation of freeze-substitution and frozen hydrated 
images 

How can we reconcile the differences seen in Figs 1 and 2, remembering that the cell in 
Fig. 1 has been dehydrated and decorated with a heavy-metal stain? As it is dehydrated, 
we would expect that there would be a certain amount of contraction of the wall 
regions in freeze-substitution images (Fig. 1), because the structures are no longer 
hydrated. We would also expect reactive groups to be labelled. On the other hand, 
frozen hydrated structure would not be condensed and this is why the thickness of each 
wall region is greater in Fig. 2 than in Fig. 1. This same image shows a periplasmic 
space, whereas Fig. 1 does not. The periplasm has contracted and condensed in Fig. 1, 



SGM symposium 65 



Cell-wall structure and physico-chemistry 91 

but it has been preserved in its natural state in Fig. 2. Accordingly, the periplasm in 
Fig. 1 is more concentrated and (it seems) more reactive, as it stains strongly. In Fig. 2, 
the periplasm has not condensed and it has low density. The conclusion, then, is that the 
natural state of the periplasm in these cells is as a relatively low-density matrix of 
highly reactive biomatter occupying a definite periplasmic space, defined by the 
membrane and the middle region. Presumably, the periplasm in this space consists of 
new wall polymers, secreted proteins (and their associated chaperones) and both 
periplasmic enzymes and oligosaccharides (Matias & Beveridge, 2005). 

The region above this periplasmic space is wider in frozen sections than in freeze- 
substitutions (cf. Figs 1 and 2) and, as this is the hydrated structure, the increased 
width shows its natural state. Both figures reveal it to be of relatively low contrast; 
freeze-substitutions suggest that there are few available reactive groups and frozen 
sections suggest that there is little density (Matias & Beveridge, 2005). Therefore, this 
region, as the stress-bearing region of the wall, has been stretched taut (thereby 
reducing its mass) and most reactive groups have been utilized to ensure that the 
network is cemented firmly together. This forms a strong but elastic fabric of wall 
polymers, mainly peptidoglycan (Yao et al., 1999; Pink et al., 2000). 

The outermost fibrous region, which is highly decorated with stain in Fig. 1, is not seen 
in Fig. 2. Accordingly, this outermost region has so little mass that it cannot be seen, but 
is highly reactive. This is in accordance with cell-wall turnover, as this is a region where 
autolysins are breaking down old peptidoglycan, making it soluble. This would reduce 
its mass whilst, at the same time, generating many new reactive sites due to hydrolysis 
(Matias St Beveridge, 2005). 

How does this new interpretation of wall structure correlate 
with metal-ion interaction and mineralization? 

It is undeniable that metal ions interact strongly with bacterial surfaces and mineralize 
them (Beveridge & Murray, 1976; Beveridge & Koval, 1981; Beveridge et al., 1983; 
Beveridge & Fyfe, 1985; Beveridge, 1989c; Fein et al, 1997; Fortin et al., 1998; Fowle 
et al., 2000 ; Daughney et al., 2001 ; Martinez & Ferris, 2001 ; Yee & Fein, 2001 ; Yee et al., 
2004a). Cell walls adsorb metal ions and minerals are nucleated in Gram-positive cell 
walls because of metal-ion interaction with the peptidoglycan and secondary polymers 
(Fig. 3; Beveridge & Murray, 1980; Beveridge et al., 1982). Our new structural 
observations on Gram-positive walls now show the quantity of hydrated biomatter that 
resides in the wall for metal interaction (Fig. 2). They also reveal the positions of the 
most reactive and likely regions for metal-ion interaction and mineral growth (Fig. 1). 
The outermost fibrous region would be the most accessible reactive region and, here, 
there would be little problem of metal-ion access and of mineral-particle growth. As 

SGM symposium 65 



92 V. R. Phoenix and others 




* - 










Cytoplasm 



Fig. 3. Conventional thin section of a B. subtilis cell that has been subjected to 50 mM FeCI 3 
treatment for 1 5 min at 22 °C. The iron has begun to precipitate from solution onto the cell wal 
(arrows). Notice that most iron is associated with the wall surface and with the periplasm. No 
stains other than the iron have been used on this cell. Bar, 50 nm. 




Fibrous 







Middle 

Periplasm 
Membrane 



Fig. 4. Diagram to explain the regions of the Gram-positive cell wall where metal ions will interact 
with reactive sites and develop into fine-grained minerals. The fibrous region refers to region 3, the 
middle region to region 2 and the periplasm to region 1 of Fig. 1 . The membrane is the plasma 
membrane. The black, crystalline shapes refer to developing minerals and their size represents 
mineral abundance. 

many wall polymers are in the act of being solubilized, these polymers and their 
precipitates would be sloughed from the cells, but could continue to grow and mature 
into bona fide mineral phases in the external milieu. The middle region has less 
reactivity, as it is highly cross-linked. Because turgor pressure stretches this region 
almost to breaking point, the peptidoglycan would be in a relatively 'open' configur- 
ation (Pink et al., 2000) so that most metal ions could penetrate through. The inner 



SGM symposium 65 



Cell-wall structure and physico-chemistry 93 

region (or periplasmic space) is a highly reactive, loose gel of polymers and proteins 
that would both interact with and precipitate metal ions. Here, though, as the peri- 
plasm resides within a confined space between the plasma membrane and the middle 
wall region, mineral growth would be restricted to the accessible space. 

These correlations and interpretations allow certain predictions to be made as to where 
environmental metal ions should interact with Gram-positive bacterial surfaces. Fig. 4 
shows a model of these predictions, with most metal minerals associated with the 
outermost region and the periplasmic space. Gram-negative surfaces are structurally 
more complicated (Beveridge, 1999), but they have also been imaged via freeze- 
substitution and frozen hydrated sections (Matias et al., 2003) and are, therefore, also 
able to be correlated. Limited space for this chapter prohibits us from doing so. 

POTENTIO METRIC PROPERTIES OF CELL SURFACES 

The surface charge of any microbial cell is controlled by the density, distribution 
and protonation state of ionizable sites in the cell wall. These chemical sites are pre- 
dominantly carboxyl, phosphoryl, amino and hydroxyl groups and they deprotonate 
with increasing pH, imparting a net negative charge on the surface. The deprotonation 
of, say, a carboxylate can be described by the following equation: 



B-COOH ^ B-COCT+ FT (equation 1) 

where B-COOH is a protonated carboxyl group on the bacterial surface and B-COO~is 
its deprotonated form. The log equilibrium constant (pKJ for the reaction in equation 
1 is then defined as: 

K a = [H + ] x [B-COOI / [B-COOH] (equation 2) 

pK a = -logK a (equation 3) 

In simple terms, the pX a can be considered as the pH at which a significant number of 
those groups or ligands become negatively charged. High-resolution acid-base 
titrations (HRABT), combined with a suitable modelling approach, can be used to 
determine ligand concentrations and their corresponding pX a values (e.g. Fein et al., 
1997; Cox et al., 1999). This approach works because the adsorption and release of H + 
from such ligands causes a shift in the pH away from the expected norm during titration 
(Fig. 5). This difference, known as the charge excess, can be used to calculate ligand 
concentrations and their pK a values. These calculations can be performed by using 
a number of approaches, such as the surface complexation method with fiteql 
(Fein et al., 1997; Borrok et al., 2004a), the linear programming method (LPM; Cox 
et al., 1999), the fully optimized continuous (focus; Smith & Ferris, 2001; Martinez et 
al., 2002) method or the Donnan shell model (Plette et al., 1995; Martinez et al., 2002; 
Yeeetal., 2004b). 

SGM symposium 65 



94 V. R. Phoenix and others 



HRABT analysis of the Gram-negative bacterium Pseudomonas aeruginosa PA01, 
modelled by using the LPM approach, is shown in Fig. 6. Here, five distinct proton- 
binding sites (and their corresponding concentrations) are revealed. Based on model 
compounds, each pX a can be assigned to a most probable ligand type. Carboxyl groups 
deprotonate over the lower pH range and exhibit a range of pK a values that occur 
predominantly between pK a 2 and 6 (Perdue, 1985). Similarly, phosphoryl groups 
exhibit intermediate pK a values that are commonly between pK a 54 and 8 (Martell 
et al., 1987). Thiols (e.g. cysteine) exhibit pK a values around 8, amines generally 
between 8 and 11 and hydroxyls between 9 and 12 (Perdue, 1985; Martell et al., 1987). 
Considering this, the pK a groups in Fig. 6 can be assigned as follows: pK^ 4-7, carboxyl; 
pK a 5-9, phosphoryl (or carboxyl); pK a 6-8, phosphoryl; pK a 84, amino (or thiol); 
pX a 94, amino/hydroxyl. 

The pK a spectrum of each bacterium is controlled by the chemical composition of the 
cell wall and, thus, each detected pX a and its assigned ligand type can be attributed 
to various components of this structure. Carboxyl groups can be associated with 
proteins, peptide stems of peptidoglycan and (in Gram-negatives) lipopolysaccharides 
[LPSs; either because of ketodeoxyoctonate (Kdo) in the oligosaccharide (unless it is 
acetylated; Vinogradov et al., 2002, 2003a, 2004) or because of O side chains, depend- 
ing on O serotype]. Phosphoryl groups are associated with teichoic acids in Gram- 
positive cell walls (Beveridge et al., 1982), whereas, in Gram-negative outer membranes 
(OMs), they are found in the core oligosaccharide and lipid A fraction of the LPS and in 
the phospholipids of the membrane's inner face (Beveridge, 1999). Thiol groups, whilst 
uncommon in OM proteins (OMPs), may be found in periplasmic proteins. Amines are 
found in abundance within the peptidoglycan (peptide stems), as well as in the proteins 
of the OM and periplasm. 

Importantly, HRABT analysis provides us with both the concentration and pX a 
values of these potential metal-binding ligands (Fein et al., 1997). Additionally, 
some metal-adsorption modelling approaches utilize a pK a spectrum to underpin their 
metal-complexation models (e.g. Fein et al., 1997). 

PHYSICO-CHEMICAL STUDIES 

Cell-surface complexity and its impact on potentiometric 
properties 

The nature of the cell wall is complex and diverse, imparting individual surface 
characteristics to each bacterial species and strain. Each micro-organism displays a 
unique arrangement of ligands and thus has unique potentiometric and metal-binding 
properties (e.g. Beveridge & Fyfe, 1985; Martinez & Ferris, 2001; Borrok et al., 2004a). 

SGM symposium 65 



1 



-300 J 



Cell-wall structure and physico-chemistry 95 




Base added (ml) 



Fig. 5. Raw data from high-resolution acid-base titration (HRABT) analysis. The pH (y axis) is given in 
mV. #, Blank titration (no bacteria); O, titration with bacteria. 



E 

o 

E 



c 
O 

I 

8 

c 
o 
o 

-a 

c 

03 
D5 



0-8 -I 

07- 

0-6 

0*5- 

0-4- 

0-3- 

0-2- 

0-1- 



\ 



•f 



* 



~r 

5 



T 

6 



7 



* 



8 



9 



10 



Fig. 6. p/C, spectrum generated from HRABT analysis of P. aeruginosa PA01 , modelled by using a 
linear programming method (LPM). This spectrum was generated from three titrations. Error bars are 
SD(a = 1). 

As described earlier, novel cryo-based TEM methods have provided further insights into 
the distribution of reactive sites within this basic framework. Together with cryoTEM, 
HRABT adds another surface-analysis tool to our arsenal for deciphering the 
availability of cell-wall ligands. For example, the cyanobacterium Calothrix sp. (strain 
KC97) surrounds itself in a thick (up to 1 \im) extracellular polysaccharide (PS) sheath, 
which is above the cell wall. This is ion-permeable and ions are able to interact with 
functional groups that reside within both the sheath and cell wall. HRABT analysis of 
intact cells and isolated sheath has shown that only ~ 15 % of the available ligands are 



SGM symposium 65 



96 V. R. Phoenix and others 



located in the sheath; the remaining 85 % are located in the wall (Phoenix et al., 2002). 
Thus, the cell-surface reactivity of Calothrix sp. can be divided into a dual layer, com- 
posed of a highly reactive polar cell wall enclosed within a poorly reactive apolar 
sheath. The distribution of ionizable groups is likely to be controlled by the organism in 
a self-serving manner. For example, the motile phase of Calothrix, called hormogonia, 
does not possess a sheath and the electronegative wall is the exposed surface, making 
the cyanobacterium hydrophilic. This property seems ideally suited to a free-swimming 
phase, which must be in close contact with its aqueous milieu. However, the benthic, 
non-motile phase assembles a sheath around cells, making them more hydrophobic and 
able to stick to inanimate surfaces. The distribution of ionizable groups over the cell 
surface also has implications for metal complexation. The higher concentration of 
ionizable groups on the cell wall ensures that this structure remains the main sink for 
metals (Yee et al., 2004a). 

There are several other general types of polymeric structures besides sheaths that reside 
above cell walls (Whitfield & Valvano, 1993), such as capsules and exopolymeric subs- 
tances (frequently found in microbial biofilms). All of these structures, including 
S layers [i.e. paracrystalline arrays of (glyco)protein; Sleytr & Beveridge, 1999] aid 
in adjusting the availability of surface groups on bacteria, and such polymeric controls 
on ligand distribution can differ according to bacterial genus, species and (even) strain. 
HRABT analysis of eight different strains of Shewanella sp. demonstrated that 
those that expressed O side-chain LPS (i.e. smooth strains) commonly exhibited higher 
total ligand concentrations than those with just core oligosaccharide (i.e. rough 
strains) (V. R. Phoenix, A. A. Korenevsky, T. J. Beveridge, Y. A. Gorby & F. G. Ferris, 
unpublished data). Furthermore, comparison of individual pK a values revealed that 
Shewanella strains possessing O side chains were relatively enriched in ligands with a 
pK a of ~5, suggesting that these chains contained available carboxyl groups. Their 
existence was corroborated by structural nuclear magnetic resonance (NMR) analyses 
of selected rough and smooth LPS strains (Vinogradov et al., 2002, 2003a, b, 2004). 
From these data, one may speculate that strains that exhibit O side chains should 
display higher metal-binding capacity than their rough counterparts. 

The complex nature of cell surfaces, however, ensures that we cannot assume that all 
organisms exhibiting O side chains will contain higher ligand concentrations than 
rough strains. This may be due to more significant differences in ligand concentrations 
in other components of the cell wall, such as OMPs, or may simply be due to a low 
concentration of O side chains or O side chains that are extremely short (and therefore 
do not have many ligands on them). This is exemplified by comparing two strains of 
P. aeruginosa, one that expresses O side chains (PAOl) and one that does not (rd7513). 
Importantly, the density of O side chains on PAOl is typically quite low, with only 

SGM symposium 65 



Cell-wall structure and physico-chemistry 97 

~20 % of the LPS containing O side chains (Kropinski et al., 1987). When the total 
ligand concentrations for each strain were evaluated by using HRABT analysis, both 
PAOl and rd7513 revealed ligand concentrations that were not statistically different 
(V. R. Phoenix & T. J. Beveridge, unpublished data). This reflects the low density of O 
side chains dispersed over PAOl's surface, which was insufficient to significantly 
enhance the total ligand concentration. 

With this in mind, a combined HRABT plus NMR approach can be further exploited 
to reveal the density of certain surface polymers. As described above, although some 
LPS molecules may express O side chains, many will not and it can therefore be difficult 
to estimate total O side-chain concentration over the cell surface. We have combined 
NMR and HRABT analyses of isolated LPS to approximate O side-chain coverage on 
Shewanella alga BrY. In this example, NMR analysis of the lipid A and LPS core 
indicated a 1 :5 stoichiometric relationship between carboxyl and phosphoryl groups. 
Additionally, NMR of the O side chain revealed that this component contained one 
carboxyl group per repeating structural unit (no other ionizable groups were present). 
However, HRABT analysis of the same LPS revealed a 1 : 1 stoichiometry between 
carboxyl and phosphoryl groups. Thus, considering a 1 : 5 stoichiometry in the lipid 
A and core and the single carboxyl per repeat unit in the O side chain (from NMR 
analysis), each LPS polymer must contain, on average, an O side chain of four repeating 
structural units (V. R. Phoenix, A. A. Korenevsky, T. J. Beveridge, Y. A. Gorby & 
F. G. Ferris, unpublished data). 

When performed on whole cells, HRABT analyses probe functional groups embedded 
deep within the cell wall (although exactly how deep is uncertain). This provides 
additional information about metal-adsorption properties, as metal ions, like protons, 
will be able to migrate into the wall matrix. However, HRABT analysis may not be 
suitable for evaluating all cell-interface processes. For example, when considering 
surface charge to determine electrostatic interactions with mineral surfaces [e.g. during 
adhesion, an approach that measures the charge properties of the very outermost 
fractions of the cell surface, such as zeta-potential analysis, may be more suitable (see 
later section for more details)]. This is because changes in the composition of these 
outermost cell-wall polymers can have a marked impact on the zeta potential of a 
bacterium, and yet have a considerably smaller impact on the HRABT-determined 
properties. This is especially true if the compositions of ionizable groups embedded 
deeper within the cell wall are similar for the organisms in question. For example, in a 
recent study of several strains of Shewanella sp., the strains displayed a notable diversity 
in zero point of charge (ZPC) as determined from zeta-potential analysis, whereas 
the ZPCs from HRABT analysis were similar (V. R. Phoenix, A. A. Korenevsky, 
T. J. Beveridge, Y. A. Gorby & F. G. Ferris, unpublished data). Thus, as highlighted 

SGM symposium 65 



98 V. R. Phoenix and others 



0-8 



(a) 



CD 



o 



CD 



0-6 



0-4- 



£ 0-2- 



4 
0-5n (b) 



i 



0-4 



03- 



0-2- 



0-1 o 



5 



ft 




H^ 



5 



7 



~i 1 

9 10 



O 



o 



o 



o 



o 



5 



6 



7 
p/t 



8 



10 



Fig. 7. p/C a spectra from HRABT analysis (modelled by using an LPM approach), (a) pK a spectra of 
whole cells of 5. algae BrY (■) and 5. oneidensis MR-1 (O) (n =3, a = 1 ). (b) p/C a spectra from LPS 
isolated from 5. algae BrY (■) and 5. oneidensis MR-1 (O). 

throughout this chapter, it can be pertinent to understand not only the concentration 
and types of reactive groups within the cell wall, but also their distribution. 

This is further emphasized by comparing HRABT analyses of whole cells with those 
of polymers isolated from the outer surface of the cell wall. HRABT-determined pK a 
spectra for S. alga BrY and Shewanella oneidensis MR-1 are shown in Fig. 7(a) ; the two 
organisms clearly display quite similar pK a spectra. However, pX a spectra obtained 
from LPS isolated from MR-1 and BrY are very different (Fig. 7b). Because LPS is located 
on the outer face of the OM, it has a significant impact on the microbe's interaction 
with the surrounding environment. Considering the difference in LPS pK a spectra, it is 
unsurprising that these two strains display quite different zeta potentials and mineral- 



SGM symposium 65 



Cell-wall structure and physico-chemistry 99 

adhesion properties (A. A. Korenevsky & T. J. Beveridge, unpublished data). This 
would not have been anticipated from the pX a spectra of whole cells. Note the 
approximate 1 :1 stoichiometry between the carboxyl {pK a ~4) and phosphoryl (pK a 
6-8) groups in the LPS of BrY used, as described above, to determine the O side- 
chain density on this bacterium. Furthermore, the dominance of phosphoryl groups 
(pK a 6-S) in the LPS of MR-1 (Fig. 7b) has also been corroborated by NMR 
(Vinogradov et al, 2003a) and further emphasizes the compatibility of these two 
methods. The overall lower ligand concentrations exhibited by the LPS of MR-1 are 
also noteworthy. These are probably due to the presence of capsular polysaccharide 
(CPS) of low ligand concentration associated with the LPS (Korenevsky et al, 2002). 

Several studies have noted changes in potentiometric and metal-adsorption properties 
in Gram-positive and -negative bacteria in response to environmental conditions 
(Daughney et al, 2001; Borrok et al., 2004a; Haas, 2004). For example, B. subtilis walls 
exchange their major secondary polymer (teichoic acid) for teichuronic acid under 
phosphate starvation (Beveridge, 1989a, c). However, for the purposes of environ- 
mental modelling (to predict the transport and fate of metals in aqueous systems), it is 
desirable to describe the potentiometric and metal-adsorption properties of a wide 
range of bacteria by using a few bulk average parameters (Yee & Fein, 2001; Borrok et 
al., 2004a, b). This is viable, providing the error associated with using universal para- 
meters is smaller than other errors inherent in environmental modelling (e.g. uncertain- 
ty in cell-number distribution throughout the system). However, understanding the 
diversity of potentiometric and metal-adsorption properties displayed by different 
species is still pertinent because it (i) allows better determination of the average 
universal parameters and their associated error and (ii) provides high-resolution data 
for specific environments where more accurate modelling may be required. 

Hydrophobicity studies 

Many studies using electron microscopy of natural biofilms or planktonic cells from 
soils, sediments, mine-tailing wastes, hot springs, etc. have revealed that particulate 
mineral deposits are commonly associated with microbial cells (Fig. 8; Xue et al., 1988; 
Konhauser et al, 1993, 1994, 2004; Konhauser & Ferris, 1996; Small et al, 1999; 
Martinez et al, 2003). These minerals may arise not only because of the interaction of 
metal ions with ionized functional groups on the cell surface, but also as a result of the 
binding of pre-formed, finely dispersed minerals, such as colloidal silica, metal oxides 
and clays. Laboratory experiments on the interaction of dissimilatory metal-reducing 
bacteria, such as Shewanella putrefaciens, with nano-sized particulate iron oxides 
showed strong association between bacterial surfaces and mineral particles (Fig. 9), 
so strong that it resulted in irreversible adsorption (Caccavo et al, 1997; Glasauer et 
a/., 2001). 

SGM symposium 65 



100 V. R. Phoenix and others 




Fig. 8. Thin section of a natural freshwater-stream sediment taken from Table Mountain in Grosse 
Morne Park, Newfoundland, Canada. No electron-microscopic stains have been used and three cells 
that are surrounded by minerals have been indicated by asterisks. Bar, 250 nm. 

The binding of particulate minerals by cell surfaces is a result of the interplay of 
electrostatic (or electric double layer), van der Waals and Lewis acid-base interactions. 
The latter are responsible for such interfacial effects as hydrophobic attraction and 
hydrophilic repulsion (Van Oss et al., 2001). It has long been recognized that so-called 
'hydrophobicity' plays a significant role in the binding of particulate minerals to the 
bacterial surface, yet it is clearly a most poorly understood aspect of surface physico- 
chemistry Hydrophobicity is a relatively non-specific term that has limited utility, as it 
does not provide any scale of magnitude and is comparative; the term draws no actual 
line between 'hydrophobic' and 'hydrophilic' surfaces, as one surface is simply more 
unwettable than the other. The only reliable quantitative criterion for defining the 
relative terms 'hydrophobicity' and 'hydrophilicity' is the sign and value of free energy 
of the interfacial interaction between molecules, particles or surfaces immersed in 
water (Van Oss & Giese, 1995). 

Microbial-surface hydrophobicity is determined by the availability and number of 
apolar functional groups (e.g. -CH 3 ) on cell-wall macromolecules. Although there are 



SGM symposium 65 



Cell-wall structure and physico-chemistry 101 




Fig. 9. Conventional thin section of 5. oneidensis MR-1 that has been reacted with a fine-grained 
hydrated iron oxide, which has adsorbed to the cell surface. Bar, 200 nm. 



no set rules, proteins are considered a major factor for cell-surface hydrophobicity, 
whilst PSs determine surface hydrophilicity. The influence of these two major 
components of cell walls on surface properties can be illustrated in the Gram-negative 
OM, where phospholipids and LPSs assemble into a membrane bilayer and where, 
extending from the OM face, PSs strongly affect surface physico-chemistry and 
adhesiveness. Surface PSs (capsule or O side chains of LPS) extending from the cell 
surface screen hydrophobic OMPs. Smooth strains (expressing O side chains or 
possibly CPS) are more hydrophilic than their rough counterparts, which possess only 
core oligosaccharide on their LPS (Williams et al., 1986; Makin & Beveridge, 1996; 
Williams & Fletcher, 1996; Flemming et aL, 1998; DeFlaun et al, 1999; Faille et al, 
2002). However, this is not always true; sugar residues of microbial PS are often 
substituted with apolar methyl and acyl groups (Whitfield & Valvano, 1993), which can 
contribute to surface hydrophobicity (Makin & Beveridge, 1996). 

We have recently conducted studies on a number of different Shewanella strains in 
which a number of different techniques were used to determine cell-surface 
hydrophobicity, including contact angle measurement (CAM) and hydrophobic 
interaction chromatography (HIC) (A. A. Korenevsky & T. J. Beveridge, unpublished 
data). The first method, CAM, yielded quantitative parameters of cell-surface 
hydrophobicity and assessed the free surface energy by using the Lifshitz-van der 
Waals/Lewis acid-base approach. Here, measurements showed the overall character 
of all Shewanella surfaces to be hydrophilic and essentially monopolar, with low 
electron-acceptor but high electron-donor parameters. The second method, HIC, was 
chosen as the least perturbing cell-surface protocol, as it measures live bacteria in their 
natural, fully hydrated state (Pembrey et al, 1999). It is based on the microbial 
interaction with a hydrophobic substrate (octyl Sepharose) and appeared to be much 



SGM symposium 65 



102 V. R. Phoenix and others 



more sensitive to ultrastructural variations of the bacterial surfaces than CAM. The 
HIC values of Shewanella showed a strong relationship with known LPS/PS composi- 
tions (A. A. Korenevsky & T. J. Beveridge, unpublished data) and were in good agree- 
ment with previous studies where increased cell-surface hydrophobicity was found in 
strains expressing short O side chains or rough LPS (Hermansson et aL, 1982; Williams 
et aL, 1986; Makin & Beveridge, 1996; Williams Sc Fletcher, 1996; Flemming et aL, 
1998; Hanna et aL, 2003). The higher hydrophobicity of rough strains is considered to 
be the consequence of increased exposure of hydrophobic OMPs. 

In our studies, electrostatic interaction chromatography (ESIC) and microelectro- 
phoresis (zeta potential) were also used and provided information on cell-surface 
electronegativity (A. A. Korenevsky & T. J. Beveridge, unpublished data). Both types 
of measurements showed that the presence of smooth LPS or CPS decreased cell- 
surface electronegativity Even though encapsulated bacterial strains are often reported 
to be more electronegative than rough LPS phenotypes, the opposite was found 
with our Shewanella strains. Both methods showed the cells with PS to be significantly 
less electronegative. This tendency has been reported previously for a number of 
Gram-negative bacteria (Hermansson et aL, 1982; Makin & Beveridge, 1996; Flemming 
et aL, 1998; Razatos et aL, 1998; DeFlaun et aL, 1999). Our recent LPS/PS structural 
analyses help to explain these present electronegativity results (Vinogradov et aL, 
2002, 2003a, b, 2004). The common repeating motif of the S. alga BrY O side chain 
consists of four monosaccharides and contains 3-hydroxybutyric acid and malic acid. 
Here, the carboxyl group of malic acid is the only available ionizable group in the 
structure (Vinogradov et aL, 2004). A similar situation was seen in the CPS of 
S. oneidensis MR-4; here, the repeating unit contains only one glucuronic acid as an 
ionizable residue (E. Vinogradov, A. A. Korenevsky & T J. Beveridge, unpublished 
data). In contrast, Shewanella 's core oligosaccharides were found to be highly phos- 
phorylated (Vinogradov et aL, 2002, 2003, 2004b; Moule et aL, 2004) and thus very 
electronegative at circumneutral pH. Accordingly, Shewanella^ O side chains and CPS 
seem to be only weakly charged and are capable of screening the main surface charge 
located in the core-lipid A region of LPS. Unlike most other shewanellae, S. oneidensis 
MR-1 possesses a surprisingly low surface charge at circumneutral pH, even though it 
expresses rough LPS. The core of its LPS contains an unusual component, 8-amino- 
Kdo (Vinogradov et aL, 2003), which may account for the lowered surface charge, as the 
amino group masks the carboxylate. Furthermore, this strain possesses a microcapsule 
of 20-30 nm that, like the PSs of MR-4, could possess a low charge density (Korenevsky 
etal., 2002). 

It at first seems to be a contradiction that HIC suggests that the more polar 
(electronegative) strains are more hydrophobic than the less electronegative strains. The 

SGM symposium 65 



Cell-wall structure and physico-chemistry 103 

answer to this contradiction could reside in our traditional perception of OM structure 
and the arrangement of surface macromolecules. We too often consider this membrane 
to be a homogeneous, static assembly of macromolecules that are dispersed randomly 
over the bilayer's surface. Instead, all molecules are in high motion (usually over 
extremely short timescales) and it is probable that, at distinct time intervals, a mosaic 
of molecular patches of definite polarity and charge exists (Sokolov et al., 2001; 
Korenevsky et al., 2002; Vadillo-Rodriguez et al., 2003, 2005). Given an inanimate 
surface of consistent hydrophobicity (such as the Sepharose beads used for HIC), the 
tendency of the OM surface would be to congregate and align compatible macro- 
molecules towards the inanimate surface. Therefore, although the overall Shewanella 
cell surface was found to be hydrophilic by other analyses (e.g. CAM; A. A. Korenevsky 
& T. J. Beveridge, unpublished data), it still may be capable of hydrophobic interactions 
through such apolar congregations. This would lead to adhesion to hydrophobic sub- 
strata, such as was seen in our HIC experiments. 

It is clear that hydrophilicity and hydrophobicity have a strong bearing on the para- 
meters that affect geomicrobiology. The accessibility of metal ions to reactive sites on 
bacterial surfaces, biomineral development on such surfaces, adsorption of pre-formed 
fine-grained minerals and adhesion of bacteria to larger mineral phases all depend on 
the surfaces' polar and apolar properties. In our Shewanella study (A. A. Korenevsky & 
T. J. Beveridge, unpublished data), a strong correlation between such macroscopic 
surface parameters as surface negativity, relative hydrophobicity and bacterial adhesion 
to haematite was observed. Rough strains exhibited higher affinity and maximal sorp- 
tion capacity (by more than an order of magnitude) to haematite when compared with 
encapsulated strains. It follows, then, that hydrophobic interactions do not make a 
significant contribution to Shewanella^ adhesion to haematite. This is in accord with 
acid-base titrations, which demonstrated that hydrous ferric oxide interacts directly 
with carboxyl sites on the surface of S. putrefaciens CN32 (Smith & Ferris, 2003; 
Martinez etal., 2003). 

The assessment of the cell-surface hydrophobicity is a challenging task because 
microbial walls have high chemical and structural complexity and are degraded readily 
by native enzymes, such as autolysins (Fig. 1). Yet, a good understanding of the 
bacterial surface, together with a thorough knowledge of the techniques being used in 
the determination of surface physico-chemistry, can be illuminating and can provide 
valuable information about the interactions between microbes and the environment. In 
this chapter, we have integrated together a variety of techniques (from electron 
microscopy to HRABT to hydrophobicity/hydrophilicity studies) into an amalgam to 
help explain metal ion-bacterial surface interactions, fine-grained biomineral 
development and mineral-sorption/adhesion phenomena. Certainly, there is much more 

SGM symposium 65 



104 V. R. Phoenix and others 



to be deciphered, especially as microbes are dynamic systems that can quickly react and 
respond to changing environmental systems, such as altered redox conditions, nutrient 
limitation, pH and temperature. Microbial surfaces alter with these environmental 
fluxes and therefore can present an entirely different set of interfacial parameters. This 
dynamic cellular behaviour and the range of interdisciplinary fields that are required to 
study these animate-inanimate systems make geomicrobiology particularly 
challenging and exciting. 

ACKNOWLEDGEMENTS 

The research presented in this chapter was made possible through funding provided by a 
Canadian National Science and Engineering Research Council (NSERC) Discovery grant and 
a United States Department of Energy Natural and Accelerated Bioremediation Research 
Program (US-DOE-NAB1R) grant to T.J.B. The electron microscopy was performed in the 
Guelph Regional Integrated Imaging Facility (GRIIF), which is partially funded by an NSERC 
Major Facility Access grant to T J. B. 

REFERENCES 

Beveridge, T. J. (1988). The bacterial surface: general considerations towards design 

and function. Can J Microbiol 34, 363-372. 
Beveridge, T. J. (1989a). Role of cellular design in bacterial metal accumulation and 

mineralization. Annu Rev Microbiol 43, 147-171. 
Beveridge, T. J. (1989b). The structure of bacteria. In Bacteria in Nature: a Treatise on the 

Interaction of Bacteria and their Habitats, vol. 3, pp. 1-65. Edited by E. R. Leadbetter 

& J. S. Poindexter. New York: Plenum. 
Beveridge, T. J. (1989c). Metal ions and bacteria. In Metal Ions and Bacteria, pp. 1-29. 

Edited by T J. Beveridge & R. J. Doyle. New York: Wiley. 
Beveridge, T. J. (1 999). Structures of Gram-negative cell walls and their derived membrane 

vesicles. J Bacteriol 181, 4725^1733. 
Beveridge, T. J. (2000). Ultrastructure of Gram-positive cell walls. In Gram-Positive 

Pathogens, pp. 3-10. Edited by V. A. Fischetti, R. P. Novick, J. J. Ferretti, D. A. Portnoy 

& J. I. Rood. Washington, DC: American Society for Microbiology. 
Beveridge, T. J. & Fyfe, W. S. (1985). Metal fixation by bacterial cell walls. Can J Earth Sci 

22, 1893-1898. 
Beveridge, T. J. & Graham, L. L. (1991). Surface layers of bacteria. Microbiol Rev 55, 

684-705. 
Beveridge, T. J. & Koval, S. F. (1981). Binding of metals to cell envelopes of Escherichia coli 

K-M.AppI Environ Microbiol 42, 325-335. 
Beveridge, T. J. & Murray, R. G. E. (1976). Uptake and retention of metals by cell walls 

of Bacillus subtilis. J Bacteriol 1 27, 1 502-1 518. 
Beveridge, T. J. & Murray, R. G. E. (1980). Sites of metal deposition in the cell wall of 

Bacillus subtilis. J Bacteriol 141 , 876-887. 
Beveridge, T. J., Forsberg, C. W. & Doyle, R. J. (1982). Major sites of metal binding in 

Bacillus licheniformis walls. J Bacteriol 1 50, 1 438-1 448. 
Beveridge, T. J., Meloche, J. D., Fyfe, W. S. & Murray, R. G. E. (1983). Diagenesis 



SGM symposium 65 



Cell-wall structure and physico-chemistry 105 



of metals chemically complexed to bacteria: laboratory formation of metal 

phosphates, sulfides, and organic condensates in artificial sediments. Appl Environ 

Microbiol AS, 1094-1108. 
Beveridge, T. J., Hughes, M. N., Lee, H., Leung, K. T., Poole, R. K., Savvaidis, I., Silver, S. 

& Trevors, J. T. (1997). Metal-microbe interactions: contemporary approaches. 

Adv Microb Physiol 38, 1 77-243 . 
Beveridge, T. J., Moyles, D. & Harris, B. (2005). Electron microscopy. In Methods 

for General and Molecular Microbiology. Edited by C. A. Reddy, T. J. Beveridge, 

J. A. Breznak, L. Snyder, T. M. Schmidt & G. A. Marzluf. Washington, DC: American 

Society for Microbiology (in press). 
Borrok, D. M., Fein, J. B. & Kulpa, C. R, Jr (2004a). Cd and proton adsorption onto 

bacterial consortia grown from industrial wastes and contaminated geologic 

settings. Environ Sci Techno! 38, 5656-5664. 
Borrok, D., Fein, J. B. & Kulpa, C. F. (2004b). Proton and Cd adsorption onto natural 

bacterial consortia: testing universal adsorption behavior. Geochim Cosmochim Acta 

68,3231-3238. 
Caccavo, F., Jr, Schamberger, P. C, Keiding, K. & Nielsen, P. H. (1997). Role of hydro- 

phobicity in adhesion of the dissimilatory Fe(lll)-reducing bacterium Shewanella alga 

to amorphous Fe(lll) oxide. Appl Environ Microbiol 63, 3837-3843. 
Cox, J. S., Smith, D. S., Warren, L. A. & Ferris, F. G. (1999). Characterizing heterogeneous 

bacterial surface functional groups using discrete affinity spectra for proton binding. 

Environ Sci Techno! 33,451 4-452 1 . 
Daughney, C. J., Fowle, D. A. & Fortin, D. (2001). The effect of growth phase on proton 

and metal adsorption by Bacillus subtil is. Geochim Cosmochim Acta 65, 1025-1035. 
DeFlaun, M. F., Oppenheimer, S. R., Streger, S., Condee, C. W. & Fletcher, M. (1999). 

Alterations in adhesion, transport, and membrane characteristics in adhesion- 
deficient pseudomonad. Appl Environ Microbiol 65, 759-765. 
Dubochet, J., McDowall, A. W., Menge, B., Schmid, E. N. & Lickfeld, K. G. (1983). 

Electron microscopy of frozen-hydrated bacteria. J Bacteriol 1 55, 381-390. 
Dubochet, J., Adrian, M., Chang, J. J., Homo, J. C, Lepault, J., McDowall, A. W. & 

Schultz, P. (1988). Cryo-electron microscopy of vitrified specimens. Q Rev Biophys 

21, 129-228. 
Faille, C, Jullien, C, Fontaine, F., Bellon-Fontaine, M.-N., Slomianny, C. & Benezech, T 

(2002). Adhesion of Bacillus spores and Escherichia co// cells to inert surfaces: role of 

surface hydrophobicity. Can J Microbiol 48, 728-738. 
Fein, J. B., Daughney, C. J., Yee, N. & Davis, T. A. (1997). A chemical equilibrium model 

for metal adsorption onto bacterial surfaces. Geochim Cosmochim Acta 61, 3319- 

3328. 
Ferris, F. G. & Beveridge, T. J. (1986). Site specificity of metallic ion binding in Escherichia 

co/ZK-12 lipopolysaccharide. Can J Microbiol 32, 52-55. 
Ferris, F. G., Fyfe, W. S. & Beveridge, T. J. (1988). Metallic ion binding by Bacillus subtilis: 

implications for the fossilization of microorganisms. Geology 16, 149-152. 
Flemming, C. A., Palmer, R. J., Jr, Arrage, A. A., van der Mei, H. C. & White, D. C. 

(1998). Cell surface physicochemistry alters biofilm development of Pseudomonas 

aeruginosa lipopolysaccharide mutants. Biofouling 13, 213-231 . 
Fortin, D., Ferris, F. G. & Beveridge, T. J. (1998). Surface-mediated mineral development 

by bacteria. Rev Mineral Geochem 35, 161-180. 
Fowle, D. A., Fein, J. B. & Martin, A. M. (2000). Experimental study of uranyl adsorption 

onto Bacillus subtilis. Environ Sci Technol 34, 3737-3741 . 



SGM symposium 65 



106 V. R. Phoenix and others 



Glasauer, S. ; Langley, S. & Beveridge, T. J. (2001). Sorption of Fe (hydr)oxides to the 

surface of Shewanella putrefaciens: cell-bound fine-grained minerals are not always 

formed de novo. Appl Environ Microbiol 67, 5544-5550. 
Graham, L. L. & Beveridge, T. J. (1990). Effect of chemical fixatives on accurate 

preservation of Escherichia coli and Bacillus subtilis structure in cells prepared by 

freeze-substitution. J Bacteriol 172, 21 50-21 59. 
Graham, L. L. & Beveridge, T. J. (1994). Structural differentiation of the Bacillus subtilis 1 68 

cell walUflacfer/o/ 176, 1413-1421. 
Haas, R. J. (2004). Effects of cultivation conditions on acid-base titration properties of 

Shewanella putrefaciens. Chem Geo! 209, 67-81 . 
Hanna, A., Berg, M., Stout, V. & Razatos, A. (2003). Role of capsular colanic acid in 

adhesion of uropathogenic Escherichia coli. Appl Environ Microbiol 69, 4474-4481 . 
Hermansson, M., Kjelleberg, S., Korhonen, T. K. & Stenstrdm, T.-A. (1982). Hydro- 
phobic and electrostatic characterization of surface structures of bacteria and its 

relationship to adhesion to an air-water interface. Arch Microbiol 131, 308-31 2. 
Hobot, J. A., Carlemalm, E., Villiger, W. & Kellenberger, E. (1984). Periplasmic gel: new 

concept resulting from the reinvestigation of bacterial cell envelope ultrastructure by 

new methods. J Bacteriol 160, 143-1 52. 
Konhauser, K. O. & Ferris, F. G. (1996). Diversity of iron and silica precipitation by microbial 

mats in hydrothermal waters, Iceland: implications for Precambrian iron formations. 

Geology 24, 323-326. 
Konhauser, K. O., Fyfe, W. S., Ferris, F. G. & Beveridge, T. J. (1993). Metal sorption and 

mineral precipitation by bacteria in two Amazonian river systems: Rio Solimoes 

and Rio Negro, Brazil. Geology 21, 1 103-1 106. 
Konhauser, K. O., Schultze-Lam, S., Ferris, F. G., Fyfe, W. S., Longstaffe, F. J. & 

Beveridge, T. J. (1994). Mineral precipitation by epilithic biofilms in the Speed River, 

Ontario, Canada. Appl Environ Microbiol SO, 549-553. 
Konhauser, K. O., Jones, B., Phoenix, V. R., Ferris, G. & Renaut, R. W. (2004). The 

microbial role in hot spring silicification. Ambio 33, 552-558. 
Korenevsky, A. A., Vinogradov, E., Gorby, Y. & Beveridge, T. J. (2002). Characterization 

of the lipopolysaccharides and capsules of Shewanella spp. Appl Environ Microbiol 

68, 4653-4657. 
Koval, S. F. & Beveridge, T. J. (1999). Microscopy, electron. In Encyclopedia of Micro- 
biology, vol. 3, pp. 276-287. Edited by J. S. Lederberg. San Diego, CA: Academic 

Press. 
Kropinski, A. M. B., Lewis, V. & Berry, D. (1987). Effect of growth temperature on the 

lipids, outer membrane proteins, and lipopolysaccharides of Pseudomonas 

aeruginosa PAO. J Bacteriol 1 69, 1 960-1 966. 
Lee, J.-U. & Beveridge, T. J. (2001). Interaction between iron and Pseudomonas aeruginosa 

biofilms attached to Sepharose surfaces. Chem Geo! 180, 67-80. 
Makin, S. A. & Beveridge, T. J. (1996). The influence of A-band and B-band lipo- 

polysaccharide on the surface characteristics and adhesion of Pseudomonas 

aeruginosa to surfaces. Microbiology 142, 299-307. 
Martell, A. E., Smith, R. M. & Motekaitis, R. J. (1987). NIST critically selected stability 

constants of metal complexes database, version 4.0. College Station, TX: Texas A&M 

University. 
Martinez, R. E. & Ferris, F. G. (2001). Chemical equilibrium modeling techniques for 

the analysis of high-resolution bacterial metal sorption data. J Colloid Interface Sci 

243, 73-80. 



SGM symposium 65 



Cell-wall structure and physico-chemistry 107 



Martinez, R. E., Smith, D. S., Kulczycki, E. & Ferris, F. G. (2002). Determination of intrinsic 

bacterial surface acidity constants using a Donnan shell model and a continuous pK a 

distribution method. J Colloid Interface Sci 253, 1 30-1 39. 
Martinez, R. E., Smith, D. S., Pedersen, K. & Ferris, F. G. (2003). Surface chemical 

heterogeneity of bacteriogenic iron oxides from a subterranean environment. 

Environ Sci Technol 37 , 5671-5677. 
Matias, V. R. F. & Beveridge, T. J. (2005). Cryo-electron microscopy reveals native polymeric 

cell wall structure in Bacillus subtilis 1 68 and the existence of a periplasmic space. Mol 

Microbiol 56, 240-251. 
Matias, V. R. E, Al-Amoudi, A., Dubochet, J. & Beveridge, T. J. (2003). Cryo-transmission 

electron microscopy of frozen-hydrated sections of Escherichia coli and Pseudo- 

monas aeruginosa. J Bacteriol 185, 61 1 2-61 1 8. 
Moule, A. L, Galbraith, L, Cox, A. D. & Wilkinson, S. G. (2004). Characterization of a 

tetrasaccharide released on mild acid hydrolysis of LPS from two rough strains of 

Shewanella species representing different DNA homology groups. Carbohydr Res 

339, 1185-1188. 
Paul, T. R., Graham, L. L. & Beveridge, T. J. (1993). Freeze-substitution and conventional 

electron microscopy of medically-important bacteria. Rev Med Microbiol 4, 65-72. 
Pembrey, R. S., Marshall, K. C. & Schneider, R. P. (1999). Cell surface analysis techniques: 

what do cell preparation protocols do to cell surface properties? Appl Environ 

Microbiol 65, 2877-2894. 
Perdue, E. M. (1985). The acidic functional groups of humic substances. In Humic 

Substances in Soil, Sediment, and Water: Geochemistry, Isolation, and Character- 
ization, pp. 493-526. Edited by G. R. Aiken, D. M. McKnight, R. L. Wershaw & 

P. MacCarthy. New York: Wiley. 
Phoenix, V. R., Martinez, R. E., Konhauser, K. O. & Ferris, F. G. (2002). Characterization 

and implications of the cell surface reactivity of Calothrix sp. strain KC97. Appl 

Environ Microbiol 68, 4827-4834. 
Pink, D., Moeller, J., Quinn, B., Jericho, M. & Beveridge, T. (2000). On the architecture of 

the Gram-negative bacterial murein sacculus. J Bacteriol 182, 5925-5930. 
Plette, A. C. C, van Riemsdijk, W. H., Benedetti, M. F. & van der Wal, A. (1995). pH 

dependent charging behaviour of isolated cell walls of a Gram-positive soil 

bacterium. J Colloid Interface Sci 1 73, 354-363. 
Razatos, A., Ong, Y.-L, Sharma, M. M. & Georgiou, G. (1998). Molecular determinants 

of bacterial adhesion monitored by atomic force microscopy. Proc Natl Acad Sci U S A 

95, 11059-11064. 
Sleytr, U. B. & Beveridge, T. J. (1999). Bacterial S-layers. Trends Microbiol 7, 253-260. 
Small, T. D., Warren, L A., Roden, E. E. & Ferris, F. G. (1999). Sorption of strontium by 

bacteria, Fe(lll) oxide, and bacteria— Fe(lll) oxide composites. Environ Sci Technol 33, 

4465-4470. 
Smith, D. S. & Ferris, F. G. (2001). Proton binding by hydrous ferric oxide and aluminum 

oxide surfaces interpreted using fully optimized continuous p/C a spectra. Environ Sci 

Technol 35, 4637-4642. 
Smith, D. S. & Ferris, F. G. (2003). Specific surface chemical interactions between hydrous 

ferric oxide and iron-reducing bacteria determined using p/C a spectra. J Colloid 

Interface Sci 266, 60-67. 
Sokolov, I., Smith, D. S., Henderson, G. S., Gorby, Y. A. & Ferris, F. G. (2001). Cell surface 

electrochemical heterogeneity of the Fe(lll)-reducing bacteria Shewanella 

putrefaciens. Environ Sci Technol 35, 341-347. 



SGM symposium 65 



108 V. R. Phoenix and others 



Umeda, A., Ueki, Y. & Amako, K. (1987). Structure of the Staphylococcus aureus cell wall 
determined by thefreeze-substitution method. J Bacteriol 169, 2482-2487. 

Vadillo-Rodriguez, V., Busscher, H. J., Norde, W., de Vries, J. & van der Mei, H. C. 
(2003). On relations between microscopic and macroscopic physicochemical 
properties of bacterial cell surfaces: an AFM study on Streptococcus mitis strains. 
Langmuir 19, 2372-2377. 

Vadillo-Rodriguez, V., Busscher, H. J., van der Mei, H. C, de Vries, J. & Norde, W. 
(2005). Role of lactobacillus cell surface hydrophobicity as probed by AFM in 
adhesion to surfaces at low and high ionic strength. Colloids SurfB Biointerfaces 41 , 
33-41. 

Van Oss, C. J. & Giese, R. F. (1995). The hydrophilicity and hydrophobicity of clay minerals. 
Clays Clay Mineral 43, 474-477. 

Van Oss, C. J., Giese, R. F. & Docoslis, A. (2001). Water, treated as the continuous liquid 
in and around cells. Cell Mol Biol 47, 721-733. 

Vinogradov, E., Korenevsky, A. & Beveridge, T. J. (2002). The structure of the 
carbohydrate backbone of the LPS from Shewanella putrefaciens CN32. Carbohydr 
Res 337, 1285-1289. 

Vinogradov, E., Korenevsky, A. & Beveridge, T. J. (2003a). The structure of the rough- 
type lipopolysaccharide from Shewanella oneidensis MR-1, containing 8-amino-8- 
deoxy-Kdo and an open-chain form of 2-acetamido-2-deoxy-D-galactose. Carbohydr 
Res 338, 1991-1997. 

Vinogradov, E., Korenevsky, A. & Beveridge, T. J. (2003b). The structure of the O-specific 
polysaccharide chain of the Shewanella algae BrY lipopolysaccharide. Carbohydr 
Res 338, 385-388. 

Vinogradov, E., Korenevsky, A. A. & Beveridge, T. J. (2004). The structure of the core 
region of the lipopolysaccharide from Shewanella algae BrY, containing 8-amino- 
3,8-dideoxy-D-A77anno-oct-2-ulosonic acid. Carbohydr Res 339, 737-740. 

Whitfield, C. & Valvano, M. A. (1993). Biosynthesis and expression of cell-surface 
polysaccharides in Gram-negative bacteria. Adv Microb Physiol 35, 1 35-246. 

Williams, V. & Fletcher, M. (1996). Pseudomonas fluorescens adhesion and transport 
through porous media are affected by lipopolysaccharide composition. Appl Environ 
Microbiol 62, 1 00-1 04. 

Williams, P., Lambert, P. A., Haigh, C. G. & Brown, M. R.W. (1986). The influence of the 
and K antigens of Klebsiella aerogenes on surface hydrophobicity and susceptibility 
to phagocytosis and antimicrobial agents. J Med Microbiol 21 , 125-132. 

Xue, H.-B., Stumm, W. & Sigg, L. (1988). The binding of heavy metals to algal surfaces. 
Water Res 22, 917-926. 

Yao, X., Jericho, M., Pink, D. & Beveridge, T. (1999). Thickness and elasticity of Gram- 
negative murein sacculi measured by atomic force microscopy. J Bacteriol 181, 6865- 
6875. 

Yee, N. & Fein, J. (2001). Cd adsorption onto bacterial surfaces: a universal adsorption 
edge? Geochim Cosmochim Acta 65, 2037-2042. 

Yee, N., Benning, L. G., Phoenix, V. R. & Ferris, F. G. (2004a). Characterization of 
metal-cyanobacteria sorption reactions: a combined macroscopic and infrared 
spectroscopic investigation. Environ Sci Techno! 38, 775-782. 

Yee, N., Fowle, D. A. & Ferris, F. G. (2004b). A Donnan potential model for metal sorption 
onto Bacillus subtilis. Geochim Cosmochim Acta 68, 3657-3664. 



SGM symposium 65 



Horizontal gene transfer of metal 
homeostasis genes and its role in 
microbial communities of the 
deep terrestrial subsurface 

Jonna Coombs and Tamar Barkay 

Department of Biochemistry and Microbiology, Cook College, Rutgers University, 76 Lipman Dr., 
New Brunswick, NJ 08901, USA 



INTRODUCTION 

Both basic and applied science issues drive our interests in the microbiology of the deep 
terrestrial subsurface. As an environment that is disconnected from the Earth's surface, 
the deep subsurface is less subject to variations in temperature and light and, in un- 
saturated zones, to intense gradients across interfaces created at the microscale level. 
These characteristics dictate an average growth rate that is very slow, up to thousands of 
years per cell division (Kieft 6c Brockman, 2001), and an ecosystem where change occurs 
over very long time scales (Fredrickson 6c Onstott, 2001). Thus, the subsurface is one of 
the most extreme environments on Earth, and identifying what limits life in the sub- 
surface has value as a model for life on other planets (Chapelle et al., 2002; Nealson 6c 
Cox, 2002). The inadvertent release of contaminants from industrial processing plants 
and storage tanks, as well as the possibility of permanently depositing nuclear wastes 
deep below the Earth's surface (Pedersen, 2001), raise questions about how microbial 
activities might exacerbate or mitigate contamination problems in the subsurface. 

The terrestrial subsurface is the habitat for diverse microbial communities that, 
together with the oceanic subsurface, may be the habitat for the largest proportion of 
Earth's biomass (Whitman et al., 1998). As subsurfaces are characterized by a range 
of physical and chemical properties, from fully aerated sedimentary shallow aquifers to 
deep igneous rocks devoid of oxygen and elevated in temperatures, their microbial 
communities are equally varied (Fredrickson 6c Fletcher, 2001). Microbial life in 
the subsurface is greatly constrained by temperature, pressure, limited space and 
availability of water and scarce resources of electron donors, acceptors and micro- 

SGM symposium 65: Micro-organisms and Earth systems - advances in geomicrobiology. 
Editors G. M. Gadd, K. T. Semple & H. M. Lappin-Scott. Cambridge University Press. ISBN 521 86222 1 ©SGM 2005 



110 J. Coombs and T. Barkay 

nutrients, challenging microbial life to its limit (Colwell, 2001). Nevertheless, studies 
over the last 30 years have revealed metabolically and phylogenetically diverse microbial 
communities in the subsurface (Amy et al., 1992; Balkwill et al., 1997). 

Microbial biomass and diversity in the subsurface 

Microbial biomass has been estimated by direct and viable counts and by the quanti- 
fication of total phospholipid fatty acids (PLFA) (Kieft et al., 1997; Ringelberg et al., 
1997). Biomass estimates show variability that corresponds to the heterogeneity of the 
geological strata that were sampled. Direct counts range from 10 7 cells (g soil) -1 in 
the sediments of the Atlantic coastal plain in North America, Rainier Mesa in Nevada 
and Witwatersrand Basin in South Africa to 10 4 cells g _1 in deep sediments of the 
western USA, and groundwater samples rarely contain more the 10 4 cells ml -1 (sum- 
marized by Fredrickson 6c Onstott, 2001). Thus, the microbial biomass of both 
the soil and the aqueous subsurface is orders of magnitude lower than those of the 
corresponding surface environments. 

Representatives belonging to the major prokaryotic lineages have been detected in the 
subsurface, as revealed by culturing methods (Balkwill et al., 1997) and by molecular 
signatures of both PLFA and 16S rRNA clone libraries (Chandler et al., 1998; Feris et 
al., 2004). Interesting unique observations emerge, however, when community structure 
and diversity are examined within the context of the unique spatial properties of the 
subsurface. For example, in the vadose zone, the area located between the top soil and 
the water table, water availability is a limiting resource not so much due to desiccation 
(water potentials of >-0-l MPa, sufficient for hydration, are common), but mostly 
because water is trapped in small spaces, creating discontinuous environments limiting 
transport of microbes, nutrients and toxicants (Kieft & Brockman, 2001). This spatial 
discontinuity of microbial niches determines microbial distribution and diversity 
patterns and limits microbial interactions to microniches. For example, Takai et al. 
(2003b) demonstrated a varied distribution of methanogens in the transition from low- 
sulfate and organic- and methane-rich shale to high-sulfate and methane- and organic- 
poor sandstone, thus relating community structure to geochemical gradients and 
lithologies. Zhou et al. (2002) reported a much higher diversity in surface relative to 
subsurface soils, but both had a high degree of species evenness rather than species 
dominance, suggesting non-competitive diversity patterns. The authors proposed that, 
in soils typified by discontinuity, microbial growth is limited by lack of diffusion of 
essential substrates rather than by competition (Zhou et al., 2002, 2004). 

Microbial metabolism in the subsurface 

Diverse modes of microbial metabolism exist in the subsurface. Heterotrophic 
metabolism is supported in aquifers recharged by surface water containing soluble 

SGM symposium 65 



Horizontal gene transfer of metal homeostasis genes 111 

organic carbon, where the consumption of limited oxygen leads to anaerobic con- 
ditions and the dominance of anaerobic respiratory pathways. Organic carbon accreted 
during the slow process of sediment formation is another energy source in the sub- 
surface (Krumholz et al., 1997; Colwell, 2001), and microbial degradation of natural 
petroleum reserves is an example of heterotrophic anaerobic metabolism with far- 
reaching consequences for oil quality and quantity (Aitken et al., 2004; Head et al., 
2003). 

Dissimilatory iron reduction is a dominant respiratory pathway in anoxic aquifers 
(Lovley et al., 2004). Iron reducers representing common (Petrie et al., 2003) and novel 
(Coates et al., 2001) members of the Geobacteraceae are often detected in such 
environments. Furthermore, the novel iron reducer Rhodoferax ferrireducens, the first 
non-phototrophic species of its genus, was isolated from subsurface sediment 
(Finneran et al., 2003). Iron reducers are important targets for bioremediation efforts in 
the subsurface because they can use other electron acceptors, among them uranium 
(Lovley et al., 2004) and vanadium (Ortiz-Bernad et al., 2004), inducing the formation 
of insoluble precipitates. These activities, stimulated in situ by the injection of readily 
oxidizable substrates such as acetate (Anderson et al., 2003) or ethanol (North et al., 
2004) into contaminated aquifers, result in the precipitation of metal and radionuclide 
contaminants. Invariably, such treatments stimulate growth of Geobacteraceae in the 
treated subsurface communities (Anderson et al., 2003; North et al., 2004). 

Sulfate reduction is another common respiratory pathway in the subsurface (Wong 
et al., 2004), driven by sulfate in groundwater and possibly by the activity of pyrite- 
oxidizing bacteria (Ulrich et al., 1998). Sulfide that accumulates during sulfate 
reduction may complex with metals and radionuclides (Neal et al., 2004) and retard 
their mobility in the subsurface, thus making stimulation of sulfate-reducing bacteria 
another strategy for bioremediation in the subsurface. 

Chemolithoautotrophic metabolism is a major mode of microbial metabolism in the 
subsurface, supporting vast communities of methanogens (Chapelle et al., 2002) and 
possibly acetogens (Pedersen, 2001) in environments with very low levels of organic 
substrates, such as deep aquifers or igneous rocks. Pedersen (2001) has proposed the 
hydrogen-driven biosphere hypothesis to explain microbial life in the latter. According 
to this hypothesis, hydrogen and carbon dioxide drive methanogens and acetogens, 
which then support the activities of acetoclastic methanogens and acetate-utilizing iron 
and sulfate reducers, resulting in the formation of organic polymers that, upon their 
degradation, are converted to hydrogen and carbon dioxide. A continuous source of 
hydrogen in the subsurface is required to support this hypothesis. The issue of whether 
hydrogen can (Stevens & McKinley, 1995; Freund et al., 2002) or cannot (Anderson 

SGM symposium 65 



112 J. Coombs and T. Barkay 

et al., 1998) be produced in the subsurface by the reaction of water with minerals 
is presently undecided. The prevalence and diversity of chemolithoautotrophic meta- 
bolism among subsurface microbes is also highlighted by the isolation of a thermo- 
philic hydrogen- and sulfur-utilizing chemolithoautotroph from a thermal aquifer, 
Sulfur ihydroge nib ium subterraneum, belonging to the order Aquificales (Takai et al., 
2003a). Hydrogen-driven chemoautotrophy by aerobes may be an important, as-yet 
unexplored, process in vadose zones where little organic matter is available for hetero- 
trophic processes. This lack of information may reflect the fact that the focus of 
subsurface microbiology research has been on anaerobic processes in groundwater 
aquifers because of the higher biomass and metabolic rates relative to those in the 
vadose zone. 

In this chapter, we address the issue of the interactions of subsurface micro-organisms 
with heavy metals and how they are affected by the exchange of genetic material. Thus, 
we touch on the issues of metal homeostasis and genetic diversity in subsurface 
microbiology, two topics barely investigated to date. 

THE INTERACTIONS OF SUBSURFACE BACTERIA WITH HEAVY 
METALS 

The contamination of the deep subsurface with mixtures of radionuclides, metals and 
organic solvents that may leach into groundwater aquifers is one of the most detri- 
mental legacies of the Cold War. Immobilization of these contaminants may be the 
only feasible approach to solving this problem, and micro-organisms that convert 
inorganic contaminants to less soluble precipitates play a prominent role in in situ 
immobilization strategies for the subsurface (NABIR, 2001). As these strategies depend 
on the activity of microbes, the presence of a mixture of toxicants may result in the 
inhibition of reactions essential for immobilization. It is for that reason that strains of 
Deinococcus spp. with high levels of resistance to ionizing and gamma radiation have 
been engineered with the ability to degrade organic contaminants and withstand metal 
toxicity (Brim et al., 2000, 2003), and that the toxicity of metals and actinides to 
bacteria with a potential in bioremediation has been evaluated (Reed et al., 1999; 
Ruggiero et al., 2005). 

A study to determine the level of metal resistance among subsurface aerobic hetero- 
trophic bacteria was initiated, reasoning that these microbes play a role in facili- 
tating microbial metabolism in the subsurface (Benyehuda et al., 2003). The microbes 
tested were from the subsurface microbial culture collection (SMCC) that is maintained 
at Florida State University, USA, and included 261 isolates from the Savannah 
River Site (SRS) in South Carolina, USA (borehole P24) and 89 strains from the 
Hanford site in Washington state (borehole YB-02). The SRS isolates belonged to 

SGM symposium 65 



Horizontal gene transfer of metal homeostasis genes 113 

the a-, /5- and y-proteobacteria, as well as to the high-G+C Gram-positive group. 
The Hanford strains contained representatives of these taxonomic groups, as well 
as the low-G+C Gram-positive group. Resistance to Pb(II), Hg(II) and Cr(VI) was 
determined by disk-inhibition assays on solid growth media and by comparison with 
the response of well-characterized reference resistant and sensitive strains (Benyehuda 
et aL, 2003). Results were analysed for the relationship of metal resistance to the 
properties of the tested microbial communities and the environments from which 
they were isolated. The major findings were: 

(i) Resistance to Pb(II) and Cr(VI) was common among subsurface strains from SRS 
and Hanford sediments, while fewer, mostly Gram-positive strains, were resistant to 
Hg(H). 

(ii) With the exception of a high level of metal tolerance among Arthrobacter spp., there 
was no relationship between the phylogeny of the microbes and their metal-resistance 
patterns. This is not surprising, as metal resistance is often specified by mobile elements 
such as plasmids and transposons (Silver & Phung, 1996; Kholodii et aL, 2002). Some 
subsurface Arthrobacter isolates proved to be exceptionally resistant to Cr(VI) and 
Hg(II). Other researchers have also reported high levels of metal tolerance among soil 
Arthrobacter spp. (Roane, 1999; Megharaj et aL, 2003) and the high abundance of this 
genus in soils impacted by mixed-waste contamination (Fredrickson et aL, 2004). Thus, 
further investigation of Arthrobacter- -metal interactions is highly warranted. 

(iii) Resistances to Hg(II) and Pb(II) were more common in the SRS collection than in 
the Hanford collection (ANOVA; P<0-05) and multiple metal resistance was also 
higher for the SRS, with 33 % of all strains resistant to more than one metal in this 
group compared with 23% for the Hanford group (Fig. 1). Thus, toxic metals 
influenced the evolution of resistance more effectively in the SRS community than in the 
Hanford community. Varying geological and geochemical factors may explain this 
difference. For example, metal toxicity is mitigated by low redox potential and high clay 
and organic matter (Collins & Stotzky, 1989; Giller et aL, 1998), and the clay content of 
Hanford sediment is higher than in the more sandy SRS sediment. These results 
illustrate that, in complex environments, microbe-metal interactions are greatly 
impacted by environmental factors, most likely by controlling bioavailability and thus 
metal toxicity For more details of this study, see Benyehuda et at. (2003). 

This study, as well as others addressing the issue of survival and activities of microbes in 
mixed-waste-contaminated subsurfaces, has focused on aerobic bacteria. However, 
current in situ immobilization efforts target the activities of metal- and radionuclide- 
reducing anaerobic bacteria (Anderson et aL, 2003; Istok et aL, 2004). The few who 

SGM symposium 65 



114 J. Coombs and T. Barkay 



SRS strains Hanford strains 





Fig. 1. Multi-resistance to Pb(ll), Hg(ll) and Cr(VI) among bacteria from the subsurface. The proportion 
of strains resistant to none (white segments), one (pale grey), two (dark grey) or all three (black) of the 
test metals among strains from the SRS and the Hanford sites are shown. Redrawn with permission 
from Benyehuda etal. (2003). 



have examined the response of anaerobes to metal toxicity reported conflicting results. 
Desulfovibrio desulfuricans G20, a model organism for the immobilization of metals 
as sulfides, was susceptible to micromolar concentrations of Cu(II), Zn(II) and Pb(II) 
when a medium designed to minimize metal complexation was used (Sani et al., 2003). 
Mixed cultures of sulfate reducers were inhibited by Cr(VI) (Smith & Gadd, 2000) and 
Cu(II) and Zn(II) (Utgikar et al., 2003). Likewise, Shewanella spp., studied for their role 
in immobilizing metals and radionuclides by reducing them to insoluble forms, were 
affected by U(VI) (Wade & DiChnstina, 2000) and Cr(VI) (Viamajala et ai, 2004). 
These observations clearly suggest that susceptibility to metals and thus the acquisition 
of metal resistance by microbes in the subsurface is critical to the success of 
bioremediation in environments contaminated by mixed wastes. 

THE EVOLUTION OF METAL HOMEOSTASIS GENES BY 
HORIZONTAL GENE TRANSFER (HGT) IN SUBSURFACE 
MICROBIAL COMMUNITIES 

Gene transfer among microbes in the subsurface 
environment 

Genes encoding resistance to heavy metals are often transferred among micro- 
organisms in much the same way that antibiotic-resistance genes travel through 
microbial populations, by horizontal gene transfer (HGT) mechanisms. Like anti- 
biotics, which are frequently produced by soil organisms, heavy metals are compounds 
that subsurface organisms are likely to encounter as part of their environment. 

SGM symposium 65 



Horizontal gene transfer of metal homeostasis genes 115 

Bioavailable heavy metals are produced naturally by the geochemical weathering of 
ores and mobilized with the movement of groundwater. Anthropogenic contamination, 
however, may increase the concentrations of toxic metals in a given ecological niche 
manyfold. The presence of resistance genes on mobile genetic elements within the 
subsurface community is therefore a distinct advantage. The occurrence of HGT in 
topsoils, natural waters and in association with the internal and external surfaces 
of plants and animals is well recognized. However, HGT has barely been examined in 
the deep subsurface. Because population densities (Normander et al., 1998; Licht 
et al., 1999) and active metabolism (Smets et al., 1993) stimulate HGT, while most 
deep subsurface environments are notorious for low population densities and meta- 
bolic rates (Balkwill, 1989; Kieft & Brockman, 2001), the deep subsurface may be the 
least conducive environment for genetic exchange. To examine HGT and its role in 
the evolution of metal resistance in the subsurface, we have looked for evidence of the 
horizontal inheritance of genes encoding P IB -type ATPases in bacteria from subsurface 
sediments of the SRS (Coombs &: Barkay, 2004). 

P, B -type ATPases and their roles in metal homeostasis and 
HGT 

P IB -type ATPases are membrane-associated ion pumps that are responsible for 
maintaining metal homeostasis by mediating the transport of heavy metals using the 
energy generated by the hydrolysis of ATP. Those that are specific for monovalent 
cations [Cu(I) and/or Ag(I)] are found in the three domains of life, while those specific 
to divalent cations [Zn(II), Cd(II) and/or Pb (II)] are only found among the prokaryotes. 
These metal pumps can function in either the import of essential ions or the export of 
ions that have reached harmful levels in the cell cytoplasm, depending on the 
orientation of the protein in the membrane (Rosen, 2002). 

Several P^-type ATPases have been shown to be associated with mobile genetic 
elements, from Gram-positive organisms such as Lactococcus lactis (O'Sullivan et al., 
2001), Staphylococcus aureus (Nucifora et al., 1989) and Artbrobacter spp. (K. Jerke 
and C. Nakatsu, personal communication) and from Gram-negative organisms such as 
Ralstonia metallidurans (Borremans et al., 2001) and Stenotropbomonas maltophilia 
(Alonso et al., 2000). However, the occurrence of HGT and its effects on the evolution 
of this locus within a specific microbial community had not previously been examined. 
We have targeted the genes encoding P IB -type ATPases for a study on the role of 
HGT in the evolution of metal homeostasis among subsurface bacteria because of 
the importance of metal ion homeostasis for survival in the harsh environment of the 
metal-contaminated subsurface. Isolates from the SMCC were selected for study 
because of the large number of Pb(II)-resistant bacteria in the SRS community 
(Benyehuda et al., 2003). 

SGM symposium 65 



116 J. Coombs and T. Barkay 

Evolution of P IB -type ATPases by HGT in a subsurface microbial 
community 

We have used a retrospective approach, based on the recognition of genomic indicators 
for evolution by HGT. We reasoned that, because HGT is estimated to occur at rates of 
31 kb per million years (Lawrence & Ochman, 1997) and subsurface micro-organisms 
reproduce very slowly, prospective approaches to HGT detection, such as microcosm 
incubations, may be of little relevance to subsurface communities. However, because 
genomic data may be interpreted in different ways, calling into question the validity of 
all molecular signatures of HGT (Eisen, 2000), we employed multiple methods in 
combination while determining whether or not a gene encoding a P IB -type ATPase was 
horizontally transferred. These methods included (i) examining the congruence of the 
P IB -type ATPase phylogeny with that of the 16S rRNA gene, (ii) looking for unusual 
sequence composition (G+C content) of the P IB -type ATPase when compared with that 
of the host genome and (iii) looking for shared indels (insertion/deletion events) among 
P IB -type ATPase genes from different organisms. 

To obtain P IB -type ATPase genes (zntA/cadA/pbrA-likt genes) from the SRS aerobic 
heterotrophs, a nested PCR approach was developed. Novel primer sets for PCR 
amplification were designed by aligning conserved domains in zntAIca dAlpbrA-Y\kz 
genes that were available in databases. The first PCR targeted the phosphatase domain 
and the ATP-binding domain and the second reaction used conserved sequences in 
the transmembrane metal-binding domain and the ATP-binding domain. Using nine 
PCR primer pairs, amplification products of zntA/cadA/pbrA-like genes from 48 of 105 
Pb(II)- resistant subsurface strains were obtained and sequenced. These sequences and 
the DNA sequences of 16S rRNA coding genes of the corresponding hosts were then 
used to determine whether HGT has contributed to the evolution of zntAlcadAlpbrA- 
like genes among the subsurface bacteria. For more details about this approach, see 
Coombs 6c Barkay (2004). Phylogenetic incongruence using both parsimony (heuristic) 
and distance (neighbour-joining) methods indicated that, in four of the isolates, zntAI 
cadA/pbrA-likt genes evolved by HGT (Fig. 2), and three of these were supported by 
unusual sequence composition, i.e. G+C content and the presence of indels. All trans- 
fers were among the /5- and/or y-proteobacteria, which were the predominant groups 
among the Pb (II) -resistant subsurface bacteria from which sequence data were 
obtained. Two of these transfers were to Comamonas spp., and comparison of cluster- 
ing patterns between the zntAlcadAlpbrA and the 16S rRNA trees suggested that, in 
one case (in strain B0669), transfer could have occurred from another Comamonas sp., 
and in the other (in strain B0173), the origin could not be clearly identified (Fig. 2). A 
third HGT event, the acquisition of a Pseudomonas-like P IB -type ATPase by Ralstonia 
sp. B0665, was not supported by additional sequence features, but the phylogenetic 
evidence as indicated by the bootstrap support value was very strong. Finally, a P IB -type 

SGM symposium 65 



Horizontal gene transfer of metal homeostasis genes 117 




CD -5 E 

c E ^ 

oj o o 

It -8 

8 CU QJ 

-7-1 U +-" 

^ C fD 
"u CD u 

c= m- CD 
TO °^ 

CO .^ 

"D C U 

g .Q-D 

3 ±=! cd 

-r LO = 

o ~ 



CD 

CD 



O 
oo 
CD 
CD 



Q_ 

CD ^ 

-C O 



CD 
CD 



CD 
_Q 



CD 
TO 

00 



CD .„ • — 



on 

CD 



TO 



u to 



o 



CD 



o p 

CD — 



±L O 

00 _C 

LA) 

Cd 

in 

CD 



CD 
U 

TO 



00 

_Q 

=3 
oo 

M— 

O 

00 

CD 



E 

p 

on 

C 

TO 

-t— ' 
00 



00 

CD 
TO 

2 

CD 

C 
CD 

cri 



TO 



00 

CD 
CD 



O 



< CD 



cf 



OjLO 



TO 

t 

oo 
_Q 



CD 



O 
oo CD 

c c 

TO CD 
■*— ' Z3 

ni ^u 
C oo 

-i- CD 
TO -g 

£ O 

b _qj 

CD U 
C =5 

O C 

E cd 

TO £ 
X TO 



-= O 
oo fN 
CD 

JZ >^ 
U TO 

2 TO 

_Q CD 

T3 03 

S-g 

JZ! b 

o 
o 
~ u 

u 

CD 



en 



E 

o 



o 



-3 

CD 



CD 

^ & 
CD CT 

~D CD 

> ^ 

CD — 

CD -Q 

c -9- 

cn£ 



P c- 

o^ O 

= E 

CD OJ 

-^ 

00 ^ 

-Q -o 

-a a; 

CD ^ 



TO "D 

.y p 



"D 



Q_ 
= CD 

CD ^ 
TO lo 

oo O 

C c 
TO ~D 

cn qj 



o 



o 



" 5. 

E CL 

TO 3 

to 00 



SGM symposium 65 



118 J. Coombs and T. Barkay 



Table 1. Evidence to support HGT in the subsurface 

Supporting evidence is provided either by phylogenetic incongruence of a target gene when 
compared to the 1 6S rRNA gene phylogeny of the host organism or by the presence of mobile genetic 
elements carrying functional genes (Plasmid). Two-letter abbreviations are used for US states. 

Site of isolation Source organism(s) Gene(s) of interest Reference 

Phylogenetic incongruence 

Savannah River, SC Comamonas spp., P|B"tyP e ATPases Coombs & Barkay (2004) 

Pseudomonas spp. 

Savannah River, SC ^Proteobacteria tRNA(Leu)(UAA) Vepritskiy ef al. (2002) 

South Glens Falls, NY Gram-negative bacteria Naphthalene dioxygenase Herrickef al. (1997) 

Plasmid 

Metal-plating lagoon Pseudomonas sp. czc,ncc Smets etal. (2003) 

Savannah River, SC Heterotrophic bacteria Metal-resistance genes Fredrickson ef al. (1988) 

South Glens Falls, NY Gram-negative bacteria Naphthalene dioxygenase Ghiorse ef al. (1995) 

Savannah River, SC Sphingomonas sp. F1 99 Catabolic genes Romine ef al. (1 999) 

Various Sphingomonas spp. Dibenzo-p-dioxin, Basta ef al. (2004) 

dibenzofuran and 

naphthalene sulfonates 
degradation 

Savannah River, SC Sphingomonas sp. F199 Catechol 2,3-dioxygenase Stillwell ef al. (1995) 

Savannah River, SC Gram-negative Cryptic Brockman etal. (1989) 

Savannah River, SC Sphingomonas spp. 2,3-Dihydroxybiphenyl Kim ef al. (1996) 

1,2-dioxygenaseand 
catechol 2,3-dioxygenase 



ATPase from Acinetobacter sp. strain B0064 grouped phylogenetically within a 
/2-proteobacterial clade. The G+C content supported the phylogenetic evidence, indi- 
cating that this could have been a recent gene acquisition by the Acinetobacter strain. 

These results indicate that HGT has occurred, albeit at low frequencies, during the 
evolution of metal homeostasis genes among subsurface bacteria. Other observations 
also support these conclusions (Table 1). Plasmid-borne genes for metal resistance 
and the degradation of hydrocarbons have been obtained from subsurface isolates, and 
phylogenetic analyses have suggested transfer of hydrocarbon-degradation genes in 
a shallow aquifer contaminated with coal tar (Herrick et al., 1997). The demonstration 
of conjugation in microcosms simulating low-nutrient subsurface soils (Smets et al., 
2003) suggests that HGT can affect the evolution and genetic diversity of subsurface 
soil communities. 

Here, we used the primary DNA sequences of zntA/cadA/pbrA-likt genes to deduce the 
evolutionary pathway of an environmentally important function, metal homeostasis, 

SGM symposium 65 



Horizontal gene transfer of metal homeostasis genes 119 

among subsurface bacteria. While it is tempting to conclude from this study that HGT 
occurs in the subsurface, such a conclusion is impossible without clear evidence that the 
studied strains evolved in the subsurface. Without such evidence, the alternative 
possibility that transfer occurred prior to deposition in the subsurface cannot be ruled 
out. Evidence for evolution in situ currently exists for a collection of Arthrobacter spp. 
from the Yakima Barricade, where a coherence exists between the 16S rRNA- and recA- 
based phylogenies and the geological strata from which the strains originated, 
suggesting a long-term evolution, possibly as long as 8 million years, in the subsurface 
(van Waasbergen et al., 2000). Examining microbial communities from this and other 
similar environments may reveal HGT and other processes that affect genetic diversity 
as they have occurred in the subsurface. 

Evidence for HGT of P, B -type ATPases in complete prokaryotic 
genomes 

In order to evaluate the observed frequency of HGT among subsurface microbes, we 
analysed genes encoding P IB -type ATPases of 188 bacterial and 22 archaeal genomes. 
As a clear phylogenetic distinction between P IB -type ATPases specifying mono- and 
divalent pumps does not currently exist, our analysis encompassed all zntA and copA- 
like sequences. Only P IB -type ATPases loci that exactly matched the amino acid 
sequence of the phosphatase and the transmembrane metal-binding domains of 
enzymes with documented activity were included in the collection. The resulting 
collections of 311 P IB -type ATPases and the 16S rRNA genes of the corresponding 
genomes were subjected to phylogenetic analysis and cases of incongruence between 
the two phylogenies were identified. When incongruence was detected, supportive 
evidence was sought by examining sequence composition as described above. In 
addition, sequences proximal, i.e. within 5 kbp, were examined for the presence 
of regulatory genes that might have been co-transferred with the P IB -type ATPase genes. 
When found, these were subjected to the incongruence test as above to determine 
their phylogenetic affiliation (Coombs & Barkay, 2005). 

Twelve instances of phylogenetic incongruence were detected, six of which were 
transfers across subclasses within the Proteobacteria (Table 2). This is not surprising, 
because our collection of P IB -type ATPases was dominated by proteobacteria. However, 
the remaining transfer events were across a longer phylogenetic distance. In two cases, 
zntA loci from low-G+C Gram-positive bacteria were found in the genomes of the 
y-proteobacteria Stenotrophomonas maltopbilia and Legionella pneumophila and 
a transfer of an £-proteobacterial cop A to Ureaplasma parvum, a low-G+C Gram- 
positive, was also noted. These findings highlight an extensive involvement of 
Gram-positive bacteria in gene exchange. Evidence that zntA and cop A were 
transferred to Deinococcus radiodurans from a-proteobacteria emerged from the 

SGM symposium 65 



120 J. Coombs and T. Barkay 



CD 

~o 

_Z3 

u 

_c 

on 





— 




>» 




cu 




_Q 




E 




■a 


in 
CD 


o 

C 




CU 


E 


CD 

en 






a. 


o 


T3 




Q. 


c 


cd 

-t— > 




3 


CD 


n: 




l/) 


01 


-t— ■ 
O 




1- 


^_ 


c 




U 


03 


c 




X 


1q 

o 


CD 

JZ 






1_ 


■4— ■ 






u 


_c 






E 


"D 






CD 


CD 

-i— 1 






-i—' 


03 






CD 


C 








CD 






Q_ 


'on 






E 


CD 
~CS 






o 


on 






u 


TO 






E 


CD 






o 


_Q 








E 






M— 


3 






CD 


cz 






U 


on 






c 


Z5 






CD 
"D 


u 

_o 






"> 


CD 




* 


CD 


+- 1 




i_ 











C 


CU 




C 


o 


c 







"D 


CD 

cn 




"O 


CD 


<T 




> 


C/l 


8- 




cu 


_Q 







z__ 


l/l 

CD 


o 







C 


s 




OJ 


CD 






CT 


C7l 


N 




<D 
CU 


^C 


CD 

C 




c 


a 


O 




'— 1 


o 


C 






u 


Oj 

JZ 






"D 


■4— > 






C 
03 


o 


^ 




<^ 


E 


l/l 




4~j 


"O 


3 




5 


CD 


_o 




CD 


'o3 

■4— ' 

c 


< 

-t— < 

c 




4- J 


O 


NJ 




M— 


u 


10 




O 


CD 


C 




1— 


E 


5 




U 


o 

C 


3 




X 


CD 


O 




■ 


CD 


"6 




fM 


CD 


£ 


cu 


CD 


C 


Q 


E 






CD 


CD 


c 

CU 


H 


> 


sSi 


o 



cu 
■a 

■D 

Q) 

i_ 

-C 
l/l 



cu 

c 
cu 



_JD 

3 
Dl 

CU 

i_ 

"to 

E 

x 
o 



c 

CU 

■M 

c 
o 

u 

u 

+ 

3 
i/l 

3 
C 



O 



o 



o 



c 

N 



~C 
O 

"6 

on 

u 

u 

8 

O 
.C 

CD 

Q 



o o o o 



o 



u 

+ 

i 

_o 

on 



o o 



o 



.02 
CD 

t; 

Oj 

o 

c 

03 
& 

in 



on 

2 



o o o o 





CD 


CD 










> 


> 










_-t^ 


■4^ 










'on 


'on 










O 


O 










Q_ 


Q_ 








CD 


E 


i 

E 


03 


03 




CD 


03 


fD 


ai 


CD 


03 


H — ' 






-i— 1 


■4— ' 


V 


u 

03 


5 


ej 


u 


u 

03 


CD 
■4— ■ 


_Q 
O 

CD 


u 

+ 


+ 


_Q 

O 

CD 
■4— ' 


_Q 
O 

CD 

■4— ' 


u 

03 

_Q 
O 


O 




i 


O 


O 


cz 


Q_ 


i 


< 


Q_ 


Q_ 


Oj 


i 


o 


o 


_l 


1 


>> 


2 






QQ. 


QQ. 


u 



•5 

f 

03 

6 

to 

03 

c 
o 

E 

o 

a 

g 

o 

c 

& 



CN 

a 
o 

E 

CD 
C 
Q- 

-2 

"aj 

c 
o 

P 1 

CD 

—J 



m 

fD 

in 

o 
c 

CD 

5 

CD 
TO 

on 

0j 

c 

O 



-8 

CD 



c 

CD 

c 
u 

Oj 

-Q 

£ 

o 

N 



K 1 

& 

-Q 

03 

a 

u 
u 

8 
I 



o o 



fD 

"i— 

CD 
+-> 

u 

nj 

_Q 

O 

CD 

■4— ' 

o 

i 
on 

2 



O 



CD 

u 

JD 
O 

CD 



< 

Q. 

O 
U 



m 

.g 

-C 

c 

o 



03 



o 



on 



.02 

CD 

■4— ' 

u 

re 

_Q 

O 

CD 



2 2 

Q_ Q_ 

I l 

8 8 



on 
C 

~D 
O 

U 
U 

8 

O 

.c 

CD 

Q 



o 



.03 
CD 

■4— ' 

u 

03 

_Q 

O 

CD 

■4— ' 

o 

■1j 



5 

03 

a 

03 



in 

03 

a 

03 



on 

2 



.TO 

CD 
-i— 1 
U 
03 
_Q 
O 
CD 

O 

Q_ 

i 
2 



m 

c 

CD 
.C 

s 

& 

u 

03 

S 

o 

N 



o o 



o o o o 



o o o o 



03 
CD 

■4— ' 

u 

03 

_Q 
O 

CD 

■4— ' 

o 

Q_ 



on 



O 



on 



.03 
CD 

u 

03 

JD 

O 

CD 

■4— ' 

o 

Q_ 

i -J 

CQ. LU 



03 
CD 
03 

JZ 

u 

03 



~cD 

CD 

.§ 



03 c- 

§ I 
If 

03 CD 

03 

c 

o 

E 

-8 

CD 



.3 

u 

03 

-Q 

I 



x 

OJ 

+-> 

CD 
CD 
on 

CD 
C 
CD 
CD 

< 

C£ 

WD 
ID 



CD 



Q_ 

■*-> 

C 
CD 



CD 

C 

O 

u 



"D 

31 

03 
CD 
> 
CD 

on 

< 

■Jt 



CD 



O 
C 
CD 
CD 

CD 



O 



03 
CD 

E 

CD 



E 

o 



o 



O -43 
-t-> 03 
03 



O 

CD 
> 

_rg 

CD 

OJ 

c 

CD 
CD 

CD 



O 



03 



CD 



o 


!^_ 


CD 

c 


o 
E 


!_ 


^ 


CD 

-| P 


_Q 


on 


CD 

C 


U 


■j^ 


u 


CD 


4— ' 


*: 


CD 


T3 


CD 


on 


CD 


03 


o 


CD 


^ 


CD 



CD 



CD 

on 



— CD 

O 
u 

U 

+ 

03 
on 



X5 
CD 

-t- 1 
03 
U 

T3 

on 

CD 

C 
CD 
CD 

_rrj 
CD 

e 

(D 



o 

a 
.p 
■*— > 

03 



03 
U 
•4— ' 

CD 
C 
CD 
CD 

_C 
Q. 

CD 



< 

Q. 
O 

u 

03 



> 


O 


CD 

T3 


s 


"D 


C 

tsj 


03 
T3 


03 


C 


CT 


03 


_C 


17, 


T3 


CD 

' 


O 
U 



CD 

CD 
C 
CD 
CD 

03 

O 

^^ 

LD 



on 

-i— 1 

03 

_C 

-t- 1 

CD 
C 
CD 
CD 

03 
on 
03 

CD 

CD 
JZ 

~o 

CD 



CD 



03 CD 



X c 

O CD 

o_ ro 

4f CL 



SGM symposium 65 



Horizontal gene transfer of metal homeostasis genes 121 




] 



Archaea 



Hafobacterium NRC-1 3 
Methanobacterium thermautotrophicum 
Brucella melitensis 1 
Mesorhizobium loti 1 

Agrobacterium tumefaciens 3 
Deinococcus radiodurans 1 
Escherichia coli 2 
Yersinia pestis 1 
Vibrio cholerae 1 
Listeria monocytogenes 
Oceanobacillus iheyensis 
Baciflus hafodurans 3 

Nostoc P2 J Cyanobacterium 

Bacillus anthracis 2 ~ 

Bacillus subtilis 1 

l Thermoanaerobacter ethenogenes 1 2 Gram-positive 
Clostridium perfringens 2 
Clostridium acetobutylicum 1 
Synechococcus PCC 6803 4 Cyanobacterium 




Gamma 



Gram -positive 



'Pyrococcus abyssf 



0-1 changes 



^" Helicobacter pylori 2669 2 
Helicobacter pylori J99 2 
Escherichia coli CopA 



Epsilon 



Fig. 3. Phylogenetic evidence for the transfer of a zntA gene from a cyanobacterium to the 
euryarchaeon 'Pyrococcus abyssi' (boxed). Circles at each node indicate the level of bootstrap support 
obtained when analysed by both parsimony and distance methods: black, >80 %; grey, >50 %; 
white, supported at > 50 % by one method only. 



incongruence analysis and separate locations of these two loci in the genomes of both 
D. radiodurans and a-proteobacteria suggested that two independent transfer events 
were involved. A single transfer event of cop A between closely related euryarchaeota 
was detected in the genome of Pyrobaculum aerophilum (Table 2) and, most excitingly, 
a zntA of cyanobacterial origin was present in the genome of the archaeon 'Pyrococcus 
abyss? (Fig. 3). Supportive evidence confirming HGT was available for six of the 
transfers that were revealed by incongruent phylogenies (Table 2) and in three cases 
these included the presence of a gene with homology to regulatory elements that are 
known to control expression of zntAlcadAlpbrA loci. Phylogenetic analysis of these 
regulatory genes showed that they likely shared an origin with the zntA or cop A genes 
they accompanied. This latter criterion also confirmed the cross-domain transfer 
between archaea and cyanobacteria. 

Thus, it seems that, as we have found with the subsurface strains, the evolution of 
genes encoding P IB -type ATPases in sequenced genomes has been subjected to HGT 
but that their inheritance has mostly proceeded vertically. This is in contrast to the 
well-documented dominance of HGT in the evolution of other traits that enhance 
fitness to toxicants, such as resistance to mercury (Kholodii et ai, 2002) and antibiotics. 
As P IB -type ATPases mediate metal homeostasis, they may be considered more essential 



SGM symposium 65 



122 J. Coombs and T. Barkay 

to core metabolism than phenotypes that are exclusively involved in detoxification, thus 
enhancing stable genomic inheritance. The frequency of transfer detected among the 
microbial genomes, 12 of 311, was slightly lower than that among the subsurface 
bacteria, 4 of 48 (see above). This difference was most likely due to differences in 
the composition of the datasets. The genome study encompassed a broader phylo- 
genetic range, and therefore we were able to detect transfers across large phylogenetic 
distances, whereas the subsurface study detected transfer among more closely related 
organisms. However, it is possible that the frequency of HGT among the subsurface 
strains was underestimated. The more closely related microbes are phylogenetically, the 
more likely they are to exchange genetic material (Lawrence & Ochman, 1997), but 
the less likely are the transfer events to leave a detectable molecular footprint in the new 
host genome (Eisen, 2000). 



HGT gene microarray 

DNA and expression microarrays are a powerful tool in biological research, and 
applications to the study of microbial community structure (Small et al., 2001) 
and function (Taroncher-Oldenburg et al., 2003; Rhee et al., 2004) have been docu- 
mented. Zhou and his collaborators have identified three types of microarrays in 
microbial ecology (Zhou 6c Thompson, 2002). Phylogenetic oligonucleotide arrays 
(POA) consist of probes homologous to 16S rRNA genes and are used to study 
community composition and its response to environmental change. Functional gene 
arrays (FGA) are designed to evaluate the metabolic potential of a community by 
probing for genes that specify major biogeochemical reactions, including those 
essential for biodegradation and bioremediation. Community genome arrays (CGA) 
target genes of pure isolates from a specific environment. Our discovery of molecular 
signatures indicative of HGT in the genomes of subsurface bacteria (Coombs & 
Barkay, 2004) prompted us to develop a fourth type of microarray, possibly a variation 
on the CGA, the horizontal gene transfer array (HGT array). This array was designed 
to answer the question: 'what are the genetic elements that transfer metal-resistance 
genes among subsurface bacteria?' 

The HGT array includes 158 oligonucleotide (70-mer) probes specific for genes that 
encode replication/incompatibility (inc/rep) loci in 86 broad-host-range (BHR) 
plasmids belonging to 13 distinct plasmid groups and 100 probes for metal-resistance 
genes. The linkage of metal resistance on specific plasmids is suggested by positive 
signals obtained following hybridization of Cy3- or Cy5-labelled plasmid DNA 
extracted from subsurface isolates with the array. The emerging patterns classify 
plasmids according to their incompatibility groupings and linkage with metal- 
resistance genes. 

SGM symposium 65 



Horizontal gene transfer of metal homeostasis genes 123 




Probe F4: designed to the 
I ncP1p plasmid pEMT3 




Probe F9 designed to the 
IncPipplasmid pB4 



Probe L5 ; designed to the 

p iB'tyP e ATPase of 
Deinococcus radiodurans 



Fig. 4. HGT array hybridized with Cy3-labelled plasmid from the subsurface strain Comamonas sp. 
B01 73. Positive hybridization signals appear as green spots. 



Analysis of the four SRS subsurface isolates that were shown to have inherited Pb(II) 
resistance by HGT (Fig. 2) indicated that at least two of them carried multiple small 
plasmids. Plasmid DNA extracts of these strains and of additional metal-resistant 
isolates from contaminated subsurface sediments in Oak Ridge, TN, USA, were 
hybridized with the array following optimization and testing with 26 exact-match 
reference plasmids (J. Coombs and T. Barkay, in preparation). Of these, a plasmid 
extract from Comamonas sp. B0173, a strain that inherited its zntAI cadAlpbrA gene by 
transfer from an unknown donor (Coombs & Barkay, 2004), hybridized to two probes 
homologous to rep/inc loci of plasmids belonging to IncPl/? and to a P IB -type ATPase 
probe (Fig. 4) . This finding indicates that inheritance of Pb(II) resistance in strain B0173 
likely occurred by conjugal transfer of an IncPljSBHR Pb (II) -resistance plasmid. Array 
studies with a plasmid extract from the other SRS isolate are currently in progress. 

The three strains from the contaminated sediments in Oak Ridge, TN, are a part of a 
large collection of Gram-positive bacteria where metal-resistance patterns correlated 
well with plasmid carriage (Patty Sobecky, personal communication). Strain Bacillus sp. 
U26 hybridized to rep/inc probes from a group of characterized Bacillus loci and 
hybridized weakly to probes for arsenic-resistance genes. Interestingly, plasmid DNA 
from another Bacillus strain, strain V6, hybridized to rep/inc probes homologous to 

SGM symposium 65 



124 J. Coombs and T. Barkay 

those of plasmids that had been described previously in both Gram-positive and 
-negative bacteria. In addition, the plasmid preparation hybridized to arsenic-resistance 
probes. These results suggest either that two plasmids exist in strain V6 or that a single 
arsenic-resistance plasmid has two different origins of replication, and imply that strain 
V6 carries plasmids of a broader diversity than has previously been described in other 
Gram-positive bacteria. Although our dataset is currently very small, it emphasizes that 
interesting, previously uncharacterized metal-resistance plasmids exist in subsurface 
soil bacteria. 



CONCLUSIONS 

The studies reported here focused on two issues that are critical to the activities of 
microbial communities in the subsurface, metal homeostasis and HGT. Critical, 
because both of these are considerations that affect strategies for controlling the 
transport of metals and radionuclides in contaminated subsurface environments. 
Results showed a high frequency of resistance to divalent cations and a modest, yet 
significant, inheritance of a gene encoding metal homeostasis by HGT. This frequency 
of HGT of metal homeostasis genes was similar to that in the microbial world at large 
as suggested by the analysis of sequenced microbial genomes. While these results 
suggest that HGT may have contributed to the survival of microbes in the harsh 
subsurface environment, they could not determine whether transfer occurred in situ. 
This question could only be addressed by examining signatures of HGT in the genomes 
of subsurface microbes from communities whose evolution in the subsurface can be 
documented unequivocally, enabling the histories of microbial speciation to be related 
to geological processes. This opportunity may exist in certain ecological niches within 
the subsurface, which are permanently isolated from the surface and where microbial 
migration is restricted by the discontinuity of niches that support life. 

ACKNOWLEDGEMENTS 

The research described here was funded by the Natural and Accelerated Bioremediation 
Research (NABIR) program, Biological and Environmental Research (BER), US Department of 
Energy (grant no. DE-FG02-99ER62864). 

REFERENCES 

Aitken, C. M., Jones, D. M. & Larter, S. R. (2004). Anaerobic hydrocarbon biodegradation 
in deep subsurface oil reservoirs. Nature 431, 291-294. 

Alonso, A., Sanchez, P. & Martinez, J. L. (2000). Stenotrophomonas maltophilia D457R 
contains a cluster of genes from gram-positive bacteria involved in antibiotic and 
heavy metal resistance. Antimicrob Agents Chemother 44, 1778-1782. 

SGM symposium 65 



Horizontal gene transfer of metal homeostasis genes 125 



Amy, P. S., Haldeman, D. L, Ringelberg, D., Hall, D. H. & Russell, C. (1992). Comparison 

of identification systems for classification of bacteria isolated from water and endo- 

lithic habitats within the deep subsurface. Appl Environ Microbiol 58, 3367-3373. 
Anderson, R. T., Chapelle, F. H. & Lovley, D. R. (1998). Evidence against hydrogen-based 

microbial ecosystems in basalt aquifers. Science 281, 976-977 '. 
Anderson, R. T., Vrlonls, H. A., Ortiz-Bernad, I. & 10 other authors (2003). Stimulating 

the in situ activity of Geobacter species to remove uranium from the groundwater of 

a uranium-contaminated aquifer. Appl Environ Microbiol 69, 5884-5891 . 
Balkwill, D. L. (1989). Numbers, diversity, and morphological characteristics of aerobic, 

chemoheterotrophic bacteria in deep subsurface sediments from a site in South 

Carolina. GeomicrobiolJl, 33-52. 
Balkwill, D. L, Reeves, R. H., Drake, G. R., Reeves, J. Y, Crocker, F. H., King, M. B. & 

Boone, D. R. (1997). Phylogenetic characterization of bacteria in the subsurface 

microbial culture collection. FE MS Microbiol Rev 20, 201-216. 
Basta, T., Keck, A., Klein, J. & Stolz, A. (2004). Detection and characterization of 

conjugative degradative plasmids in xenobiotic-degrading Sphingomonas strains. 

JBacteriol 186, 3862-3872. 
Benyehuda, G., Coombs, J., Ward, P. M., Balkwill, D. & Barkay, T. (2003). Metal 

resistance among aerobic chemoheterotrophic bacteria from the deep terrestrial 

subsurface. Can J Microbiol 49, 151-156. 
Borremans, B., Hobman, J. L, Provoost, A., Brown, N. L. & van der Lelie, D. (2001). 

Cloning and functional analysis of the pbr lead resistance determinant of Ralstonia 

metallidurans CH34. J Bacteriol 183, 5651-5658. 
Brim, H., McFarlan, S. C, Fredrickson, J. K., Minton, K. W., Zhai, M., Wackett, L P. & 

Daly, M. J. (2000). Engineering Deinococcus radioduransfor metal remediation in 

radioactive mixed waste environments. Nat Biotechnol 18, 85-90. 
Brim, H., Venkateswaran, A., Kostandarithes, H. M., Fredrickson, J. K. & Daly, M. J. 

(2003). Engineering Deinococcus geothermalis for bioremediation of high- 
temperature radioactive waste environments. Appl Environ Microbiol 69, 4575- 

4582. 
Brockman, F. J., Denovan, B. A., Hicks, R. J. & Fredrickson, J. K. (1989). Isolation and 

characterization of quinoline-degrading bacteria from subsurface environments. 

Appl Environ Microbiol 55, 1029-1032. 
Chandler, D. P., Brockman, F. J., Bailey, T. J. & Fredrickson, J. K. (1998). Phylogenetic 

diversity of Archaea and Bacteria in a deep subsurface paleosol. Microb Ecol 36, 

37-50. 
Chapelle, F. H., O'Neill, K., Bradley, P. M., Methe, B. A., Ciufo, S. A., Knobel, L L & 

Lovley, D. R. (2002). A hydrogen-based subsurface microbial community dominated 

by methanogens. Nature ^S, 312-315. 
Coates, J. D., Bhupathiraju, V. K., Achenbach, L. A., McLnerney, M. J. & Lovley, D. R. 

(2001). Geobacter hydrogenophilus, Geobacter chapel lei and Geobacter grbiciae, 

three new, strictly anaerobic, dissimilatory Fe(lll)-reducers. Int J Syst Evol Microbiol 

51, 581-588. 
Collins, Y. E. & Stotzky, G. (1989). Factors affecting the toxicity of heavy metals 

to microbes. In Metal Ions and Bacteria, pp. 31-90. Edited by T. J. Beveridge & 

R.J. Doyle. New York: Wiley. 
Colwell, F. S. (2001). Constrains on the distribution of microorganisms in subsurface 

environments. In Subsurface Microbiology and Biogeochemistry, pp. 71-95. Edited 

by J. K. Fredrickson & M. Fletcher. New York: Wiley-Liss. 



SGM symposium 65 



126 J. Coombs and T. Barkay 



Coombs, J. M. & Barkay, T. (2004). Molecular evidence for the evolution of metal 

homeostasis genes by lateral gene transfer in bacteria from the deep terrestrial 

subsurface. Appl Environ Microbiol 70, 1 698-1 707. 
Coombs, J. M. & Barkay, T. (2005). New findings on the evolution of metal homeostasis 

genes: evidence from comparative genome analysis of bacteria and archaea. Appl 

Environ Microbiol (in press). 
Eisen, J. A. (2000). Horizontal gene transfer among microbial genomes: new insights from 

complete genome analysis. CurrOpin Genet Dev 10, 606-61 1 . 
Feris, K. P., Hristova, K., Gebreyesus, B., Mackay, D. & Scow, K. M. (2004). A shallow 

BTEX and MTBE contaminated aquifer supports a diverse microbial community. 

Microb Ecol 48, 589-600. 
Finneran, K. T., Johnsen, C. V. & Lovley, D. R. (2003). Rhodoferax ferrireducens sp. nov., a 

psych rotolerant, facultatively anaerobic bacterium that oxidizes acetate with the 

reduction of Fe(lll). Int J SystEvol Microbiol 53, 669-673. 
Fredrickson, J. K. & Fletcher, M. (2001). Subsurface Microbiology and Biogeochemistry. 

New York: Wiley-Liss. 
Fredrickson, J. K. & Onstott, T. C. (2001). Biogeochemical and geological significance 

of subsurface microbiology. In Subsurface Microbiology and Biogeochemistry, 

pp. 3-37. Edited by J. K. Fredrickson & M. Fletcher. New York. Wiley-Liss. 
Fredrickson, J. K., Hicks, R. J., Li, S. W. & Brockman, F. J. (1988). Plasmid incidence in 

bacteria from deep subsurface sediments. Appl Environ Microbiol 54, 291 6-2923. 
Fredrickson, J. K., Zachara, J. M., Balkwill, D. L, Kennedy, D., Li, S. M., Kostandarithes, 

H. M., Daly, M. J., Romine, M. F. & Brockman, F. J. (2004). Geomicrobiology of 

high-level nuclear waste-contaminated vadose sediments at the Hanford site, 

Washington state. Appl Environ Microbiol 70, 4230-4241 . 
Freund, F., Dickinson, J. T. & Cash, M. (2002). Hydrogen in rocks: an energy source for 

deep microbial communities. Astrobiology 2, 83-92. 
Ghiorse, W. C, Herrick, J. B., Sandoli, R. L. & Madsen, E. L. (1995). Natural selection 

of PAH-degrading bacterial guilds at coal-tar disposal sites. Environ Health Perspect 

103 (Suppl. 5), 107-111. 
Giller, K. E., Witter, E. & McGrath, S. P. (1998). Toxicity of heavy metals to microorganisms 

and microbial processes in agricultural soils: a review. Soil Biol Biochem 30, 1389-1414. 
Head, I. M., Jones, D. M. & Larter, S. R. (2003). Biological activity in the deep subsurface 

and the origin of heavy oil. Nature 426, 344-352. 
Herrick, J. B., Stuart-Keil, K. G., Ghiorse, W. C. & Madsen, E. L (1997). Natural horizontal 

transfer of a naphthalene dioxygenase gene between bacteria native to a coal tar- 
contaminated field site. Appl Environ Microbiol 63, 2330-2337. 
Istok, J. D., Senko, J. M., Krumholz, L. R., Watson, D., Bogle, M. A., Peacock, A. f 

Chang, Y. J. & White, D. C. (2004). In situ bioreduction of technetium and uranium 

in a nitrate-contaminated aquifer. Environ Sci Technol 38, 468-475. 
Kholodii, G., Gorlenko, Z., Mindlin, S., Hobman, J. & Nikiforov, V. (2002). Tn504Mike 

transposons: molecular diversity, evolutionary relationships and distribution of 

distinct variants in environmental bacteria. Microbiology 148, 3569-3582. 
Kieft, T. L. & Brockman, F. J. (2001). Vadose zone microbiology. In Subsurface Microbiology 

and Biogeochemistry, pp. 141-169. Edited by J. K. Fredrickson & M. Fletcher. New 

York: Wiley-Liss. 
Kieft, T. L, Wilch, E., O'Connor, K., Ringelberg, D. B. & White, D. C. (1997). Survival and 

phospholipid fatty acid profiles of surface and subsurface bacteria in natural 

sediment microcosms. Appl Environ Microbiol 63, 1 531-1 542. 



SGM symposium 65 



Horizontal gene transfer of metal homeostasis genes 127 



Kim, E., Aversano, P. J., Romine, M. R, Schneider, R. R & Zylstra, G. J. (1996). Homology 
between genes for aromatic hydrocarbon degradation in surface and deep- 
subsurface Sphingomonas strains. Appl Environ Microbiol 62, 1 467-1 470. 

Krumholz, L R., McKinley, J. P., Ulrich, G. A. & Suflita, J. M. (1997). Confined subsurface 
microbial communities in Cretaceous rock. Nature 386, 64-66. 

Lawrence, J. G. & Ochman, H. (1997). Amelioration of bacterial genomes: rates of change 
and exchange. J Mol Evol 44, 383-397. 

Licht, T. R. ; Christensen, B. B., Krogfelt, K. A. & Molin, S. (1999). Plasmid transfer in the 
animal intestine and other dynamic bacterial populations: the role of community 
structure and environment. Microbiology 145, 261 5-2622. 

Lovley, D. R., Holmes, D. E. & Nevin, K. P. (2004). Dissimilatory Fe(lll) and Mn(IV) reduction. 
Adv Microb Physiol 49, 2 1 9-286. 

Megharaj, M., Avudainayagam, S. & Naidu, R. (2003). Toxicity of hexavalent chromium 
and its reduction by bacteria isolated from soil contaminated with tannery waste. 
Curr Microbiol 47, 51-54. 

NABIR (2001). The NABIR strategic plan. Natural and Accelerated Bioremediation Research 
Program, Berkeley Lab, US Department of Energy. 

Neal, A. L., Amonette, J. E., Peyton, B. M. & Geesey, G. G. (2004). Uranium complexes 
formed at hematite surfaces colonized by sulfate-reducing bacteria. Environ Sci 
Techno! 38, 3019-3027. 

Nealson, K. H. & Cox, B. L. (2002). Microbial metal-ion reduction and Mars: extraterrestrial 
expectations? Curr Opin Microbiol 5, 296-300. 

Normander, B., Christensen, B. B., Molin, S. & Kroer, N. (1998). Effect of bacterial 
distribution and activity on conjugal gene transfer on the phylloplane of the bush 
bean (Phaseolus vulgaris). Appl Environ Microbiol 64, 1 902-1 909. 

North, N. N., Dollhopf, S. L, Petrie, L, Istok, J. D., Balkwill, D. L & Kostka, J. E. (2004). 
Change in bacterial community structure during in situ biostimulation of subsurface 
sediment cocontaminated with uranium and nitrate. Appl Environ Microbiol 70, 
4911-4920. 

Nucifora, G., Chu, L., Misra, T. K. & Silver, S. (1989). Cadmium resistance from Staphylo- 
coccus aureus plasmid pl258 cadA gene results from a cadmium-efflux ATPase. Proc 
Natl Acad Sci U 5 A 86, 3544-3548. 

Ortiz-Bernad, I., Anderson, R. T., Vrionis, H. A. & Lovley, D. R. (2004). Vanadium 
respiration by Geobacter metallireducens: novel strategy for in situ removal of vana- 
dium from groundwater. Appl Environ Microbiol 70, 3091-3095. 

O'Sullivan, D., Ross, R. P., Twomey, D. P., Fitzgerald, G. F., Hill, C. & Coffey, A. (2001). 
Naturally occurring lactococcal plasmid pAH90 links bacteriophage resistance and 
mobility functions to a food-grade selectable marker. Appl Environ Microbiol 67, 
929-937. 

Pedersen, K. (2001). Diversity and activity of microorganisms in deep igneous rock aquifers 
of the fennoscandian shield. In Subsurface Microbiology and Biogeochemistry, 
pp. 97-139. Edited by J. K. Fredrickson & M. Fletcher. New York: Wiley-Liss. 

Petrie, L, North, N. N., Dollhopf, S. L, Balkwill, D. L & Kostka, J. E. (2003). 
Enumeration and characterization of iron(lll)-reducing microbial communities from 
acidic subsurface sediments contaminated with uranium(VI). Appl Environ Microbiol 
69, 7467-7479. 

Reed, D. T., Vojta, Y, Quinn, J. W. & Richmann, M. K. (1999). Radiotoxicity of plutonium 
in NTA-degrading Chelatobacter heintzii cell suspensions. Biodegradation 10, 251- 
260. 



SGM symposium 65 



128 J. Coombs and T. Barkay 



Rhee, S. K. ; Liu, X., Wu, L, Chong, S. C, Wan, X. & Zhou, J. (2004). Detection of genes 

involved in biodegradation and biotransformation in microbial communities by using 

50-mer oligonucleotide microarrays. Appl Environ Microbiol 70, 4303-431 7. 
Ringelberg, D. B., Sutton, S. & White, D. C. (1997). Biomass, bioactivity and biodiversity: 

microbial ecology of the deep subsurface: analysis of ester-linked phospholipid fatty 

acids. FEMS Microbiol Rev 20,371-377. 
Roane, T. M. (1999). Lead resistance in two bacterial isolates from heavy metal- 
contaminated soils. Microb Ecol 37, 21 8-224. 
Romine, M. R, Stillwell, L. C, Wong, K. K., Thurston, S. J., Sisk, E. C, Sensen, C, 

Gaasterland, T., Fredrickson, J. K. & Saffer, J. D. (1999). Complete sequence of 

a 184-kilobase catabolic plasmid from Sphingomonas aromaticivorans F199. 

J Bacteriol 181 , 1585-1602. 
Rosen, B. P. (2002). Transport and detoxification systems for transition metals, heavy metals 

and metalloids in eukaryotic and prokaryotic microbes. Comp Biochem Physiol A 

Mol Integr Physiol 1 33, 689-693 . 
Ruggiero, C. E., Boukhalfa, H., Forsythe, J. H., Lack, J. G., Hersman, L. E. & Neu, M. P. 

(2005). Actinide and metal toxicity to prospective bioremediation bacteria. Environ 

Microbiol 7, 88-97. 
Sani, R. K., Peyton, B. M. & Jandhyala, M. (2003). Toxicity of lead in aqueous medium to 

Desulfovibrio desulfuricans G20. Environ Toxicol Chem 22, 252-260. 
Silver, S. & Phung, L. T. (1996). Bacterial heavy metal resistance: new surprises. Annu Rev 

Microbiol SO, 7 53-789. 
Small, J., Call, D. R., Brockman, F. J., Straub, T. M. & Chandler, D. P. (2001). Direct 

detection of 16S rRNA in soil extracts by using oligonucleotide microarrays. Appl 

Environ Microbiol 67, 4708-471 6. 
Smets, B. F., Rittmann, B. E. & Stahl, D. A. (1993). The specific growth rate of Pseudo- 

monas putida PAW1 influences the conjugal transfer rate of the TOL plasmid. Appl 

Environ Microbiol 59, 3430-3437. 
Smets, B. F., Morrow, J. B. & Arango Pinedo, C. (2003). Plasmid introduction in metal- 
stressed, subsurface-derived microcosms: plasmid fate and community response. 

Appl Environ Microbiol 69, 4087-4097. 
Smith, W. L. & Gadd, G. M. (2000). Reduction and precipitation of chromate by mixed 

culture sulphate-reducing bacterial biofilms. J Appl Microbiol 88, 983-991 . 
Stevens, T. O. & McKinley, J. P. (1995). Lithoautotrophic microbial ecosystems in deep 

basalt aquifers. Science 270, 450-455. 
Stillwell, L. C, Thurston, S. J., Schneider, R. P., Romine, M. F., Fredrickson, J. K. & 

Saffer, J. D. (1995). Physical mapping and characterization of a catabolic plasmid 

from the deep-subsurface bacterium Sphingomonas sp. strain F199. J Bacteriol 

177,4537-4539. 
Takai, K., Kobayashi, H., Nealson, K. H. & Horikoshi, K. (2003a). Sulfurihydrogenibium 

subterraneum gen. nov., sp. nov., from a subsurface hot aquifer. Int J Syst Evol 

Microbiol 53, 823-827. 
Takai, K., Mormile, M. R., McKinley, J. P., Brockman, F. J., Holben, W. E., Kovacik, W. P. f 

Jr & Fredrickson, J. K. (2003b). Shifts in archaeal communities associated with 

lithological and geochemical variations in subsurface Cretaceous rock. Environ 

Microbiol 5,309-320. 
Taroncher-Oldenburg, G., Griner, E. M., Francis, C. A. & Ward, B. B. (2003). 

Oligonucleotide microarray for the study of functional gene diversity in the nitrogen 

cycle in the environment. Appl Environ Microbiol 69, 1 1 59-1 171. 



SGM symposium 65 



Horizontal gene transfer of metal homeostasis genes 129 



Ulrich, G. A., Martino, D., Burger, K., Routh, J., Grossman, E. L., Ammerman, J. W. & 
Suflita, J. M. (1998). Sulfur cycling in the terrestrial subsurface: commensal inter- 
actions, spatial scales, and microbial heterogeneity. Microb Ecol 36, 141-1 51 . 

Utgikar, V. P., Tabak, H. H., Haines, J. R. & Govind, R. (2003). Quantification of toxic 
and inhibitory impact of copper and zinc on mixed cultures of sulfate-reducing 
bacteria. Biotechnol Bioeng 82, 306-312. 

van Waasbergen, L G., Balkwill, D. L, Crocker, F. H., Bjornstad, B. N. & Miller, R. V. 
(2000). Genetic diversity among Arthrobacter species collected across a hetero- 
geneous series of terrestrial deep-subsurface sediments as determined on the basis 
of 1 6S rRNA and recA gene sequences. Appl Environ Microbiol 66, 3454-3463. 

Vepritskiy, A. A., Vitol, I. A. & Nierzwicki-Bauer, S. A. (2002). Novel group I intron in the 
tRNA Leu (UAA) gene of a gamma-proteobacterium isolated from a deep subsurface 
environment. J Bacteriol 184, 1481-1487. 

Viamajala, S., Peyton, B. M., Sani, R. K., Apel, W. A. & Petersen, J. N. (2004). Toxic 
effects of chromium(VI) on anaerobic and aerobic growth of Shewanella oneidensis 
MR-1 . Biotechnol Prog 20, 87-95. 

Wade, R. & DiChristina, T. J. (2000). Isolation of U(VI) reduction-deficient mutants of 
Shewanella putrefaciens. FEMS Microbiol Lett 1 84, 1 43-1 48. 

Whitman, W. B., Coleman, D. C. & Wiebe, W. J. (1998). Prokaryotes: the unseen majority. 
Proc Natl Acad Sci U 5 A 95, 6578-6583. 

Wong, D., Suflita, J. M., McKinley, J. P. & Krumholz, L R. (2004). Impact of clay minerals 
on sulfate-reducing activity in aquifers. Microb Ecol 47, 80-86. 

Zhou, J. & Thompson, D. K. (2002). Challenges in applying microarrays to environmental 
studies. CurrOpin Biotechnol^, 204-207. 

Zhou, J., Xia, B., Treves, D. S., Wu, L Y, Marsh, T. L, O'Neill, R. V., Palumbo, A. V. & 
Tiedje, J. M. (2002). Spatial and resource factors influencing high microbial diversity 
in soil. Appl Environ Microbiol 68, 326-334. 

Zhou, J., Xia, B., Huang, H., Palumbo, A. V. & Tiedje, J. M. (2004). Microbial diversity and 
heterogeneity in sandy subsurface soils. Appl Environ Microbiol 70, 1 723-1 734. 



SGM symposium 65 



Biosilicif ication: the role of 
cyanobacteria in silica sinter 
deposition 

Liane G. Benning # 1 Vernon R. Phoenix 2 and 
Bruce W. Mountain 3 

1 Earth and Biosphere Institute, School of Earth and Environment, University of Leeds, UK 

2 Molecular and Cellular Biology, University of Guelph, Canada 

institute of Geological and Nuclear Sciences, Wairakei Research Centre, Taupo, New Zealand 



INTRODUCTION 

The contribution of micro-organisms to amorphous silica precipitation in modern 
geothermal hot-spring environments has been the topic of intense study in the last three 
to four decades. Here, we present a review on the field and laboratory studies that have 
specifically addressed bacterial silicification, with a special focus on cyanobacterial 
silicification. Studies related to the biogenic silicification processes in diatoms, 
radiolarians and sponges are not discussed, despite the fact that, in the modern oceans 
(which are undersaturated with respect to silica), the diagenetic 'ripening' of such bio- 
genic silica controls the global silica cycle (Dixit et ai, 2001). It is well-known that the 
amorphous silica in these organisms (particularly in size, shape and orientation) is 
controlled primarily by the templating functions of glycoproteins and polypeptides 
(e.g. silaffin and silicatein). For information on these issues, we refer the reader to the 
extensive reviews by Simpson & Volcani (1981), Kroger et al. {1997, 2000), Baeuerlein 
(2000), Perry & Keeling-Tucker (2000), Hildebrand & Wetherbee (2003) and Perry 
(2003). In addition, in terrestrial environments, a large pool of amorphous silica is 
cycled through higher plants (grasses and trees) that are believed to use silicification as a 
protection mechanism against pathogens and insects. Information on these processes 
can be found in the papers by Chen & Lewin (1969), Sangster & Hodson (1986) and 
Perry & Fraser (1991). 

In this review, we will focus solely on microbial silicification processes, which have 
been studied extensively in active geothermal hot-spring environments. These are 
characterized by geothermal waters supersaturated with respect to amorphous silica 

SGM symposium 65: Micro-organisms and Earth systems - advances in geomicrobiology. 
Editors G. M. Gadd, K. T. Semple & H. M. Lappin-Scott. Cambridge University Press. ISBN 521 86222 1 ©SGM 2005 



132 L. G. Benning and others 

derived from water-rock interaction at depth. The link between microbes and the 
surface manifestations of sinter formation (both carbonate- and silica-based) was first 
documented in Yellowstone National Park at the end of the 19th century (Weed, 1889). 
Since that time, active geothermal systems have been studied widely, due to their 
importance as geothermal energy sources and as a proxy to understanding the 
formation of epithermal ore deposits, which constitute the deep-seated hydrothermal 
features beneath active systems. In the last quarter of the 20th century, a multitude of 
studies have been carried out to quantify the formation of silica and carbonate terraces 
in active systems, with a view towards understanding whether micro-organisms play 
an active or passive role in their formation (Walter et al., 1972; Ferris et al., 1986; 
Schultze-Lam etal., 1995; Cady & Farmer, 1996; Konhauser & Ferris, 1996; Jones etal., 
1998, 2001; Konhauser et al., 2001; Mountain et al., 2003). Most of these studies have 
focused on the relationships between microbes and the resulting morphology and 
structure of modern siliceous sinters. They provide insights into the driving forces 
for sinter formation in contemporary deposits and are thus relevant to processes in 
Archaean and early Proterozoic settings, where microbes may have become encased and 
thus preserved as microfossils (Konhauser, 2000; Cady, 2001; Toporski et al., 2002). 

There have also been numerous experimental laboratory microbial silicification studies. 
In single-step batch experiments, it has been shown clearly that the affinity of aqueous 
silica to bind to a microbial surface is low, regardless of whether the micro-organisms 
are equilibrated with solutions supersaturated or undersaturated with respect to 
amorphous silica (Fein et al., 2002; Phoenix et al., 2003; Yee et al., 2003). Such single- 
step experiments do not reliably mimic the processes leading to the significant silica 
accumulation observed in hot springs. Other experimental studies have used high 
concentrations of either organosilicon solvents, such as tetraethylorthosilicate, or 
inorganic silica concentrations and/or a variety of temperatures and pressures to induce 
silicification in the presence of micro-organisms and demonstrated that a complex 
interplay exists between the precipitation of silica and the formed textures and 
structures (e.g. Oehler & Schopf, 1971; Leo & Barghoorn, 1976; Walters et al., 1977; 
Francis et al., 1978; Ferris et al., 1988; Westall et al., 1995; Konhauser et al., 2001; 
Toporski et al., 2002; Mountain et al., 2003). These studies offered important insights 
into the diagenetic-related fossilization processes and sinter textural development, but 
they cannot provide mechanistic data pertaining to molecular-level interactions 
between micro-organisms and silica accumulating in environments such as hot springs 
or the ancient oceans. 

Many studies of microbial silicification in active hot springs have shown that silicifi- 
cation rates are rapid, but that the silicification process is controlled by purely abiotic 
driving forces [i.e. boiling, cooling, evaporation, waves and splash; see Mountain et al. 

SGM symposium 65 



Biosilicification by cyanobacteria 133 

(2003) and references therein]. Microscopic analysis of silicified micro-organisms from 
active hot springs shows that the microbial surface may act as a nucleation site for silica 
precipitation (Schultze-Lam et aL, 1995; Konhauser & Ferris, 1996; Jones et aL, 2000; 
Phoenix et aL, 2000; Mountain et aL, 2003). Recent studies that exposed cyanobacteria 
repeatedly to freshly prepared, supersaturated, polymerizing silica solution (a pseudo- 
flow-through setting) have shown that extensive biomineralization, similar to that 
observed in hot springs, can be induced (Phoenix et aL, 2000; Benning &: Mountain, 
2004; Benning et aL, 2004a, b), with similar structures and textures to those observed in 
the field (Benning & Mountain, 2004; Fig. 1). Based on detailed microscopic and, more 
recently, spectroscopic measurements of samples from such laboratory experiments, it 
is now believed that the accumulation of amorphous silica on the surface of cyano- 
bacteria is controlled solely by silica nanoparticle aggregation, but that the contri- 
bution of the microbial sheaths or cell walls in this aggregation process is considered 
significant (see below; Phoenix et aL, 2000; Benning et aL 2004a, b). Lastly, recent 
studies by van der Meer et aL (2002) and Pancost et aL (2005) have shown that specific 
biomarker lipids can be preserved in natural modern silica sinters. Such biomarker 
studies can provide insight into the complex community structure of thermophilic and 
hyperthermophilic micro-organisms (including both archaea and bacteria) that are 
present during silica sinter formation. The knowledge of what biomolecules remain 
preserved in the rock record may provide a means to extrapolate back in time and thus 
to better understand processes in ancient rocks. 

In the following pages, we describe the current understanding of the abiotic and biotic 
processes occurring in geothermal environments through a review of (i) the chemistry 
of silica and the thermodynamic and kinetic aspects of precipitation, (ii) the role of 
specific components of the microbial cell surface and (iii) the pathways of silica-colloid 
interaction and aggregation on cell surfaces. Only such a synergistic approach can 
provide a quantitative model for the reactions that drive microbial silicification and that 
lead ultimately to sinter formation and fossil microbial preservation. 

THE CHEMISTRY OF SILICA 

Soluble silica or monomeric orthosilicic acid (H 4 Si0 4 ) is composed of a silicon atom 
coordinated tetrahedrally to four hydroxyl groups. Amorphous silica is defined as a 
non-stoichiometric, inorganic polymer made up of a mixture of Si0 2 and H 7 units in 
various ratios. Monomeric silica remains stable in solution at 25 °C, as long as its 
concentration is below the equilibrium concentration for amorphous hydrated silica [at 
25 °C, approx. 100-125 parts per million (p. p.m.); Her, 1979]. In most natural waters, 
the concentration of dissolved silica is low (between 1 and 100 ^iM; Treguer et aL, 1995) 
and, specifically in marine settings, the silica concentrations are regulated by the growth 
of diatoms and radiolarians. In contrast, in the surface expression of active geothermal 

SGM symposium 65 



134 L. G. Benning and others 





(c) 




Fig. 1. Silicified microbes from the New Zealand geothermal hot springs, (a) High-resolution field 
emission gun scanning electron micrograph showing silica nanoparticles attached to microbial cells 
from the Rotokawa Geothermal Pool; bar, 500 nm. (b) Transmission electron micrograph of silicified 
micro-organisms from the Wairakei Geothermal Field; bar, 1 \xm. (c) Transmission electron micrograph 
of fully silicified micro-organism from the Wairakei Geothermal Field; bar, 500 nm. Note the small 
(30-200 nm) silica particles that form aggregates on the surface of the bacterial sheath. 



SGM symposium 65 



Biosilicification by cyanobacteria 135 

systems, where temperatures are higher (approx. 30-100 °C), the dissolved silica 
concentration in effluent solutions often exceeds the equilibrium solubility of amor- 
phous silica. Total silica in hot-spring effluents can be as high as 1000 p.p.m. and this 
represents a level many times higher than saturation, even at 100 °C. Subsurface, 
geothermal fluids may be undersaturated with respect to amorphous silica but, upon 
reaching the surface, drastic changes in temperature and other physico-chemical 
parameters will induce the autocatalytic polycondensation/polymerization of silica 
monomers, because these changes will induce amorphous silica saturation to be 
surpassed. Field experimental determination of precipitation rates showed that the 
ratio of monomeric to polymeric silica in the effluent solution plays an important role 
in controlling silica-precipitation rates (Carroll et ai, 1998). 

In a purely inorganic system, the polycondensation process follows a series of steps that 
progress from the polymerization of initial monomers to form dimers, trimers etc. and, 
finally, to the formation of highly soluble, critical nuclei of approximately 3 nm in size, 
which correspond to approximately 800-900 silicon atoms and have an approximate 
molecular mass of around 50 kDa (Her, 1980; Perry, 2003; Icopini et al., 2005). This 
initial step occurs via the condensation of two silicic acid molecules and the expulsion 
of water: 

H 4 Si0 4 +H 4 Si0 4 - (HO) 3 Si-0-Si(OH) 3 + H 2 (equation 1) 

Once the silicic acid molecules condense and Si-O-Si siloxane bonds form, cyclic ring 
structures will grow and other monomers, dimers etc. will react preferentially with 
these nuclei via Ostwald ripening. Dove & Rimstidt (1994) showed that the surface free 
energy of such a particle, a [erg cm -2 (1 erg = 10 _7 J)], can be linked with the bulk 
precipitate AGf (bulk solid) and the particle surface area (A, cmj to give the free energy of 
the particle, AG f( ticle) . This in turn can be expressed as a function of the particle 
radius (r, cm; assuming spherical morphology) and the molar volume (V m , cm 3 mor 1 ) 
via: 

AG £ , = [-4^r 3 AG°/2 VI + [4 x 10 1( W 2 a] (equation 2) 

f(particle) m 

and, from equation 2, an expression for the solubility of a single particle can be derived 
(Dove &C Rimstidt, 1994). Alexander (1975) calculated the surface free energy for 
amorphous silica in equilibrium with a solution to be approximately 45 erg cm -2 
(4-5 x 10 -6 J cm -2 ). This number increases dramatically with particle size and ordering 
of the silica phase, reaching a value of 120 erg cm -2 (llxlO -6 ] cm -2 ) for quartz 
(Rimstidt & Cole, 1983), thus confirming that smaller and less ordered particles will 
dissolve as larger particles grow. Once formed, the critical nuclei will grow to form 

SGM symposium 65 



136 L. G. Benning and others 

either large nanoparticles (from several hundred nanometres up to a micrometre) or 
will aggregate to form three-dimensional complex structures (Her, 1979, 1980; Perry, 
2003). 

Based on the data of Gunnarsson & Arnorsson (2000), the equilibrium amorphous 
silica solubility at temperatures from 20 to 95 °C lies between 100 and approximately 
330 p. p.m. Conventionally, the equation representing the equilibrium between silica and 
water is written as: 

Si0 2 (s) + 2H 2 -> H 4 Si0 4 (aq) (equation 3) 

with the equilibrium constant K expressed as the activities (a) of the species: 

K=a(U 4 Si0 4 )/a(Si0 2 ) . a 2 (H 2 0) (equation 4) 

This reaction is valid for all thermodynamic calculations, but fails to take into account 
kinetic effects, as well as the variations in the hydration states of silica. For example, the 
ratio between Si0 2 and H 2 in the aqueous species, as well as in the solid, often differs 
from the ideal 1 :2, due to hydrogen-bonded waters of hydration in the stoichiometry. In 
addition, in equations 3 and 4, aqueous deprotonated and polynuclear species (e.g. 
H 9 Si0 4 ~, H 3 Si0 4 and H 6 Si 9 0y _ ) are not taken into account although, in some cases, 
such species may contribute to up to 40% of the dissolved silica (Aplin, 1987). 
Equation 3 is particularly important in geothermal systems where the geothermal 
solutions are supersaturated with respect to amorphous silica, and polymerization 
and precipitation are induced due to changes in physical and hydrodynamic con- 
ditions. 

From a thermodynamic point of view, the precipitation of amorphous silica is driven 
by cooling, evaporation, boiling, solution mixing and changes in pH. These factors all 
strongly affect the saturation level of amorphous silica. For a general precipitation rate, 
Rate v an equation of the type: 

Rate ppt =-d[nH 4 Si0 4 ]/dt =-A x k ppt [a{Si0 2 ) . a 2 (H 2 0)] (equation 5) 

can be written, where n is the no. moles H 4 Si0 4 , A is the interfacial area (in m") and 
k t is the pH-dependent precipitation-rate constant (Her, 1979; Rimstidt & Barnes, 
1980; Carroll et ai, 1998). In solutions that are close to saturation, a nucleation barrier 
that needs to be surpassed for the first nuclei to form inhibits the precipitation process. 
Nielsen (1959) modelled the growth of such nuclei and showed that the flux of mono- 
mers towards such nuclei is related to the collision rate, the Boltzman constant, 



SGM symposium 65 



Biosilicification by cyanobacteria 137 

temperature and the free energy of formation of a critical nucleus. Because quartz has 
a higher surface free energy than amorphous silica and its nucleation is inhibited, it 
follows that the nucleation and growth of amorphous silica in geo thermal systems are 
more favoured. 

In most geothermal systems, the amorphous silica that precipitates is composed of 
opal-A, a phase that displays varying degrees of ordering of the Si0 4 rings, as well as 
varying amounts of structural Si0 2 units and degrees of hydration. Opal-A (nominally 
Si0 7 . nH 2 0) is a poorly ordered, highly hydrated phase that displays only one weak, 
broad Bragg diffraction band. Other silica phases observed in geothermal silica- 
dominated systems are considered good indicators of a diagenetic ageing/altering 
process. During this ageing, opal-A is transformed into opal-CT, opal-C, moganite, 
cristobalite, chalcedony and ultimately quartz (Herdianita et al., 2000). The main 
factors influencing this transformation to more stable counterparts are time, re- 
equilibration with high-temperature or high-pH solutions, dehydroxylation/drying 
cycles or diagenetic recrystallization. Opal-A can contain between 1 and 13 % water 
in its structure; this water is present either as network water or as liquid water in 
interstices bound to either internal silanols or defect sites of surface silanols (Langer & 
Florke, 1974; Knauth & Epstein, 1982). During this diagenetic transformation/ripening 
process, this water is expelled gradually and this is accompanied by a gradual change in 

o 

J-spacing for the main Bragg peak from 4-12 to 4-04 A (0-412-0-404 nm). This process 
has been used to derive an indicator of structural ripening, as well as a measure of 
depth of burial and age. During the ageing and transformation process, water content 
drops and particle density increases to 2-3 gem -3 (from as low as 1-5 g cm -3 ). At the 
same time, porosity (initially between 35 and 60 %) can be reduced by more than half to 
a value below 30 % [see Herdianita et al. (2000) and references therein]. 

In an effluent solution, the saturation state and thus the precipitation rate of amor- 
phous silica are dependent on a variety of parameters that include thermal gradients, 
time, changes in pH, concentration of inorganic cations (i.e. Al and trace elements), 
organics and ionic strength. Furthermore, this rate depends on the presence of nucle- 
ation sites/surfaces, as well as hydrodynamic parameters such as evaporation, waves, 
splash etc. As a result, the precipitated, amorphous silica phases will be highly variable 
from site to site and the resulting morphology and textures will depend strongly on 
these precipitation regimes (with resulting morphologies of the precipitated silica 
varying from nanometre-sized spheroidal particles to flat sheets to bulk silica). 

The first precipitated opal-A is usually made up of nanometre-sized spheroids that 
are later filled in by silica cement to form bulk silica structures. Its formation is a 
dynamic process and even 'fresh' sinter features can appear homogeneous; thus, they 

SGM symposium 65 



138 L. G. Benning and others 

are sometimes difficult to distinguish from aged sinters. This has been a major stum- 
bling block when purely morphological and structurally preserved biosignatures 
observed in modern sinters have been used to relate and extrapolate to processes in 
ancient rocks, where subsequent diagenetic or metamorphic processes have homo- 
genized and altered the structures, mineral ordering and composition. Specifically, 
in ancient rocks, the preservation of biogenic material is hampered by the fact that, in 
most cases, the only preserved features are the mould or casings surrounding the 
microbes, whilst the cell walls or sheaths have been lost. However, recent geothermal 
sinters have revealed that some specific biomarkers (specifically, bacterial and archaeal 
lipids) can be preserved. This may be the approach to elucidate the preservation of biota 
in ancient rocks lacking unequivocal morphological indicators (van der Meer et al., 
2002; Toporski et al., 2002; Pancost et al., 2005). 

CYANOBACTERIAL SURFACE PROPERTIES AND FUNCTION 

The structure and composition of cyanobacterial cell walls display a number of 
characteristics that are atypical of Gram-negative bacteria. They exhibit a thick, highly 
cross-linked peptidoglycan layer (similar to that of Gram-positive organisms) that 
makes the cell wall notably stronger (Drews & Weckesser, 1982; Hoiczyk 6c Hansel, 
2000). Additionally, biomolecules commonly found in the cyanobacterial outer 
membrane, such as atypical fatty acids and carotenoids, are uncommon in other Gram- 
negative bacteria (Schrader et al., 1981; Resch & Gibson, 1983; Jurgens & Mantele, 
1991). 

As with all bacteria, each component of the cyanobacterial cell envelope plays a specific 
role. The inner component, the cytoplasmic membrane, behaves as a highly selective 
barrier, allowing vital nutrients into the cell and excreting waste material out of the cell. 
The outer component of the Gram-negative cell wall is the outer membrane (an 
asymmetrical bilayer composed of lipopolysaccharide and phospholipid). This bilayer 
also acts as a selective barrier, facilitating the transport of low-molecular-mass 
molecules via proteins known as porins. Housed between the two membranes are the 
peptidoglycan, which provides rigidity, strength and shape, and the periplasm, which 
contains functionally important enzymes. Significantly, these components of the cell 
envelope are highly functional, complex and sensitive. One may then ask how these 
metabolically vital, and in some cases delicate, layers would respond to encrustation in 
silica precipitates. 

For some Gram-negative bacteria, the outer membrane is the outermost component of 
the organism and it acts as the interface between the cell and the external environment. 
In other cases, the organism surrounds itself in an extracellular polysaccharide capsule 
or sheath. The structure of this extracellular layer can vary considerably, ranging from 

SGM symposium 65 



Biosilicification by cyanobacteria 139 



<-■-• 






Sheath 



ro 



U 



(impermeable 

to Si0 2 

colloids) 







♦**•*♦« 
**■#* 



PAR 



'**"*♦ UV 



o 



o 



Silica colloids 
/% Cj in external milieu 



Silica colloids accumulate 
on sheath's outer surface 



Fig. 2. Summary schematic illustrating extracellular silicification of an ensheathed cyanobacterium. 
Silica colloids accumulate on the outer surface of the sheath, due to the impermeability of the sheath 
to 'large' particles. In addition, it is shown how silica inhibits damaging UV light from reaching the 
cell, whilst the photosynthetically active radiation (PAR) can pass through the silica layer with less 
attenuation. 



diffuse to dense and fibrous. Dense, fibrous polysaccharide layers known as sheaths are 
particularly common in cyanobacteria (e.g. Rippka et al., 1979). Moreover, the cyano- 
bacterial sheath is known to be devoid of metabolically vital components and it is thus 
likely that the organism can withstand a higher degree of damage to this layer than the 
rest of the cell envelope. The exact role of the cyanobacterial sheath is not well under- 
stood but, as it encloses the more delicate components of the cell envelope, one of its 
primary objectives may be to help prevent damage to these components. It has been 
shown that a coating of extracellular polysaccharide can protect bacteria against 
dehydration (Dudman, 1977; Scott et al., 1996; Hoiczyk, 1998; Tsuneda et al., 2003) and 
predation (Dudman, 1977), or it can aid in adhesion to a solid substrate (Dudman, 
1977; Scott et al., 1996). More specifically, some cyanobacterial sheaths can contain the 
UV light-absorbing pigment scytonemin, aiding cyanobacterial resistance to solar 
radiation (Garcia-Pichel & Castenholz, 1991). Of particular relevance to this chapter is 
the ability of the cyanobacterial sheath to protect the cell from detrimental bio- 
mineralization (Phoenix et al., 2000; Benning et al., 2004a) and to aid in the aggregation 
of silica nanoparticles (Benning et al., 2004b). 

Sheathed cyanobacteria are found in abundance in hot-spring systems, where it has 
been shown that silica accumulation is restricted to the outer surface of the sheath on 



SGM symposium 65 



140 L. G. Benning and others 

living cyanobacteria (Fig. 1; Phoenix et al., 2000; Konhauser et al., 2001; Mountain 
et al., 2003). This is likely to occur because the polysaccharide meshwork of the sheath 
enables it to act as a filter against colloidal silica (Fig. 2). Permeability studies demon- 
strated that the sheath of Calothrix sp. was impermeable to particles of at least 11 nm 
in diameter, thus preventing the colloids from biomineralizing the sensitive components 
of the cell wall [Phoenix et al. (2000) and references therein]. 

Interestingly, in the Archaean oceans, which were enriched in silica (Siever, 1992) and 
inhabited by cyanobacteria, silica biomineralization was likely to have occurred 
(Cloud, 1965). This is particularly true for the shallow waters, in which intermittently 
exposed environments were inhabited by stromatolitic communities and evaporation 
may have controlled silica precipitation. It is possible that the sheath developed/evolved 
in the early oceans as a response and protection against detrimental silica accumulation 
on the cell wall. This is supported by several studies, including transmission electron 
microscope- and synchrotron-based Fourier-transform IR analysis, which have demon- 
strated that, in response to increased silica exposure, the sheath of cyanobacteria 
thickens (Phoenix et al., 2000; Benning et al., 2004a, b). Again, this indicates that the 
sheath can act as a protective layer against silicification. Naturally, when cyanobacteria 
are exposed continuously to supersaturated solutions of silica, silicification eventually 
becomes too extensive and this may be detrimental to the cyanobacteria. Phoenix et al. 
(2000) have shown that even quite thick silica crusts (approx. 5 Jim thick for a 10 ^im 
diameter cell) did not appear to be detrimental to the cells. However, whether there is 
a maximum amount of extracellular silicification that cyanobacteria can withstand has 
yet to be determined. 

One mechanism to overcome extensive silicification may be the release of transient 
motile phases (hormogonia) (Herdman & Rippka, 1988) from the ends of heavily 
encrusted filaments and this may provide a pathway for survival. Benning et al. (2004a, 
b) have followed bacterially mediated silica accumulation both in situ and in vivo via 
the changes in the IR signature induced by the increase in silica concentration on the 
organic framework of single bacterial cells. This approach allowed the quantification of 
the actual and not the apparent bacterial silica-accumulation process, and they showed 
that the role of the sheath is twofold. Initially, the cells react to exposure to a silica-rich 
solution by producing more sheath polysaccharide as protection. As this thicker sheath 
acts as a good ('sticky') substrate for further inorganically precipitated silica-colloid 
aggregation, silica precipitation is enhanced, with detrimental effects to the cell. 

This ability of cyanobacteria to survive and grow continually, despite extensive 
extracellular silicification in modern hot springs and presumably also in the ancient 
oceans, may have provided additional advantages to the microbes. This is because 

SGM symposium 65 



Biosilicification by cyanobacteria 141 

amorphous silica biomineralization has been demonstrated to act as an effective 
screen against UV radiation (Phoenix et al., 2001). This study has shown that 
damaging wavelengths of UV-B and particularly UV-C are absorbed strongly by 
amorphous silica, whilst photosynthetically active radiation (400-700 nm) will pass 
through the silica with significantly less adsorption (Fig. 2). This process enables cyano- 
bacteria to photosynthesize in environments subjected to elevated UV, a protective 
mechanism particularly relevant to the Archaean (Phoenix et al., 2001), where highly 
detrimental levels of UV irradiated the Earth's surface (Kasting, 1987). Furthermore, 
Heijnen et al. (1992) have shown that habitation of micro-niches in bentonite clays can 
protect other bacterial forms from predation by grazing protozoa and, thus, silica 
encrustation may similarly protect cyanobacteria by making them inaccessible or in- 
edible to protozoa. Biomineralization also plays a key role in the formation of siliceous 
stromatolitic communities, both modern and ancient, by increasing their structural 
integrity and thus longevity (Konhauser et al., 2001). It has also been speculated that, 
because the amorphous silica matrix is highly hydrated, it may afford the organisms 
an additional protection layer against dehydration. Interestingly, these potential 
advantages are similar to the functions of the sheath, and it thus seems that the sheath 
and enshrouding silica biominerals may work collectively to protect the organisms 
within. 

CYANOBACTERIAL BIOMINERALIZATION PATHWAYS AND 
COLLOID AGGREGATION 

Benning et al. (2004a, b) have followed the processes leading to cyanobacterial 
silicification by using in situ IR microspectroscopy and imaging and have quantified the 
complex interplay between the cyanobacterial cell components, specifically the sheath, 
and the polymerizing silica solution. The progression of nucleation, growth and 
aggregation of nanometre-sized silica particles and their effect on cyanobacterial 
feedback have been described as a three-stage progression. In the first stage, in response 
to the presence of polymerizing silica, the cyanobacteria will increase the formation of 
new polysaccharide polymers, i.e. they will thicken their sheath. Concomitantly, silica 
will form branched polymers that, upon collapse, will bind to the carbon backbone of 
the hydrated polysaccharide sheath via hydrogen bridges. This step can be expressed as: 

[>ROH] + [>ROH] -*2[>ROH] (equation 6a) 

and 

[>ROH] + [sSi-OH] - [=Si-OR<] +H 2 (equation 6b) 

where >ROH represents the surface-hydroxylated sugar polymer in the sheath and 
=Si-OH is the monomeric silica attached to a surface. Equation 6(b) implies a possible 
site-specific silica accumulation, with the silica monomers bound via hydrogen bridges 

SGM symposium 65 



142 L. G. Benning and others 

to another OH-containing radical. The sheath polysaccharides are the obvious candi- 
dates for this step. Benning et al. (2004b) used a kinetic approach to show that this 
reaction proceeds via a diffusion-limited mechanism (see below), in which polymerizing 
silica units in the supersaturated aqueous environment begin to coalesce and aggregate 
on the 'fresh' microbial-sheath surface. This process is enhanced once a silane group is 
attached to the bacterial sheath via hydrogen bonds and, thus, a further increase in Si 
load may lead to the formation of a thin, fully hydrated silica network. Subsequently, 
other silane bonds may form independently of the sheath and this process will become 
uncoupled from the formation of the silica-carbohydrate hydrogen bonds. This can be 
expressed as: 

[=Si-OR<] + [=Si-OH] - [=Si-0-Si=] + [ROH<] (equation 7) 

At this stage, the formation of additional polysaccharides will no longer compete with 
the polymerization of silica, a fact supported by the change in IR spectra, which 
become dominated by the more ionic Si-O-Si bonds. The last stage is the formation of 
inorganic silane bonds. This process has been shown to be governed by a reaction- 
limited process that leads to the growth of purely inorganic Si-O-Si bonds via the 
formation of an oxo bridge (Si-O-Si), whilst one water molecule is expelled. This is 
similar to the purely inorganic process described in equation 1. When surface 
attachment is considered, this step can be described via: 

[sSi-OH] + [OH-Sis] - [=si-0-Si=] +H 2 (equation 8) 

In this way, a silica network made of corner-sharing [Si0 4 ] tetrahedra is obtained when 
all Si— O groups have reacted and the critical silica nuclei have formed. Their further 
growth and aggregation will follow and no other connection to the polysaccharide 
sheath is needed. 

In general, for colloid or polymer growth and aggregation, two restrictive regimes have 
been defined: diffusion-limited aggregation (DLA) and reaction-limited aggregation 
(RLA) (Everett, 1988; Hunter, 1996; Jamtveit & Meakin, 1999). In the diffusion-limited 
case, the limiting step is the movement of two polymer units toward each other prior 
to encounter and formation of a cluster (or aggregate). In such reactions, monomers 
or oligomers collide and combine instantaneously, producing a relatively porous 
aggregate. For the formation of critical nuclei of silica, the DLA process has been 
confirmed experimentally (Lin etal., 1990; Martin etal., 1990; Vontoni et al., 2002). For 
polysaccharide polymers, however, such data are unavailable, although Rees (1977) 
showed that glucose polymers - specifically amylose - grow by a DLA process. On 
the other hand, in RLA, the concentrations of the encountered reactant pairs are 

SGM symposium 65 



Biosilicification by cyanobacteria 143 

maintained at equilibrium and thus condensation occurs more slowly. In addition, 
a significant repulsive barrier exists, such that the 'sticking probability' upon 
oligomer-oligomer interaction is small (Everett, 1988; Gedde, 1995). For silica, the 
RLA process results in a more compact aggregate structure during slow condensation 
(Martin, 19S7;L'metaL, 1990). 

Based on theoretical calculations for nucleation, crystallization, growth and aggre- 
gation of mineral phases and organic polymers, Hulbert (1969) and Gedde (1995) have 
derived hypothetical constants for the mechanistic constant n, which represents a 
parameter that is related to specific mechanisms and geometric shape of the final 
mineral particles or polymer. The two types of mechanisms (DLA and RLA) and 
several different shapes (needles, plates, spheres, fibres, sheaths etc.) have been 
investigated and values for n have been deduced. In general, n increases with increasing 
'dimensionality' of the resulting particle/polymer and, in heterogeneous systems, a 
change in mechanism often occurs. Hulbert (1969) and Gedde (1995) have concluded 
that a particle/polymer of low geometry (one- or two-dimensional; e.g. fibre or sheath), 
forming via a DLA process, will have values of n varying between 0-5 and 2, with the 
highest values representing two-dimensional growth. On the other hand, if a spherical 
(three-dimensional) entity grows or aggregates via a DLA mechanism, the value of n 
will vary between 1-5 and 2-5, with the higher numbers indicating a switch to an RLA 
mechanism. Lastly, if the same three-dimensional spherical entity grows or aggregates 
via a purely RLA mechanism, values of 3-4 are expected for n. 

It is well-known that the polysaccharide components of the sheath of Calothrix sp. 
(composed primarily of neutral sugars; Weckesser et al., 1988) is usually found in the 
form of amylose. Amylose is a linear polymer of glucose units joined by repeating 
covalent C— O bonds, and it normally forms complex aggregates of linear geometry 
(Rees, 1977). In the cyanobacteria silica system, a low geometric ordering of newly 
formed polymers was corroborated by Benning et al. (2004b), who have derived n values 
for the first step in polysaccharide growth of 0-8-1-1, thus confirming a one- 
dimensional DLA growth for the carbohydrate polymers. For the second step, which is 
dominated by the attachment of the formed silica nuclei to the cyanobacterial sheath, 
Benning et al. (2004b) derived a value for n of 1-8-2-2, indicating a DLA mechanism for 
three-dimensional growth. Finally, for the last step, an n value of 3-5-3-8 was derived, 
clearly indicating three-dimensional growth via an RLA mechanism. It needs to be 
noted, however, that polysaccharide and silica polymers can both form structures of 
mixed or changing geometry during growth or aggregation and that the data derived by 
Benning et al. (2004b), which were based on a kinetic approach, may not fully describe 
all steps in this complex process. This is particularly true because, in most polymers and 
colloid systems, the nucleation and aggregation behaviour in solution is affected 

SGM symposium 65 



144 L. G. Benning and others 

strongly by pH, ionic strength, temperature, organic concentration and type. However, 
for silica nucleation and growth, the aggregation steps quantified in the laboratory are 
expected to be similar to processes in modern geothermal hot springs and thus can 
be used as analogues to model processes in natural environments. 

CONCLUSIONS 

Cyanobacteria are a major group of phototrophic prokaryotes that play an important 
role in the textural development of silica sinters in modern geothermal environments. 
Processes that are analogous to those observed in modern hot springs may also have 
been active during the fossilization of microbes and the formation of siliceous 
stromatolites in the Archaean. These, in turn, can provide a proxy for the biogeo- 
chemical conditions of the early biosphere. A large number of field observations and 
experimental laboratory studies, as well as a few molecular-dynamic simulations, have 
led to the conclusion that, in active geothermal hot springs, cyanobacteria play no 
active role in the initial silica polymerization. It was shown that covalent bonds between 
silica and bacterial cell-wall or sheath components are not favoured and that the 
nucleation of silica from supersaturated aqueous solutions is driven by purely inorganic 
polycondensation reactions, which are strongly pH-, ionic strength-, temperature- and 
saturation state-dependent. Nevertheless, many field and experimental microscopic 
observations showed clear evidence of a link between silica sinter structures/textures 
and micro-organisms via the deposition of silica nanospheres onto the microbial 
surfaces. Ultimately, this process promotes the incorporation of micro-organisms into 
the sinter structure and leads to the preservation of microbial colonies as fossils. 
However, despite the large variety of studies carried out so far, the question remains 
as to the exact role of the microbial surface in the processes that lead to silica 
precipitation. We argue here that the formation of silica sinters is a multi-step process 
that is governed primarily by inorganically driven polycondensation of silica monomers 
and the formation of silica nanoparticles. This is followed by the microbially enhanced 
aggregation of the silica nanospheres into larger assemblages. In the first step, 
experimental and theoretical evidence indicates that polymerization of silica mono- 
mers leads to the formation of branching clusters that eventually collapse to form a 
spherical particle. Although the parameters and mechanisms controlling this collapse 
are unclear, some evidence indicates that it may be the dehydroxylation of silanol or 
silane clusters. When cyanobacteria and their complex surface structures are present, it 
appears that the polycondensation rates and, thus, silica nanoparticle nucleation rates, 
are not enhanced. In addition, and more importantly, it is believed that the precipitation 
of silica does not affect cyanobacterial metabolism or duplication rates. However, 
cyanobacteria do react by increasing the amount of extracellular polysaccharide that 
they produce. Once silicification is advanced and thus unavoidable, the thicker poly- 
saccharide sheath will enhance the aggregation of the inorganically nucleated silica 

SGM symposium 65 



Biosilicification by cyanobacteria 145 

nanoparticles into larger silica assemblages. This process occurs while they are alive, 
but continued silicification leads to cell death, lysis and finally fossilization. 

The surface features seen in active geothermal systems are often regarded as ideal 
model systems for study, because they can shed light onto processes occurring in the 
shallow-subsurface portions of the geothermal systems that are linked to deep-seated 
epithermal ore deposits. In the streams, pools and sinters forming in modern 
geothermal systems, a vast array of mesophilic and thermophilic organisms thrive at 
high temperatures and varied pH, as well as high toxic-metal concentrations that are 
usually detrimental to microbial growth. The knowledge gained from studying such 
communities and their interaction with the minerals precipitating from the super- 
saturated solutions can give valuable insights into processes of biomineralization. In 
addition, our understanding of processes related to the evolution of early life forms in 
the Archaean and early Proterozoic has been extrapolated from observations (both 
structural and chemical) of bacterial-mineral interactions in modern hot-spring 
environments or from morphological observations of Precambrian silicified micro- 
fossils and stromatolites. 

From field and laboratory observations, reaction pathways for the formation of silica 
sinters in modern or ancient hot springs have been determined. The abiotic versus biotic 
components of the silica biomineralization reaction, and thus the role of micro- 
organisms in this process, were defined. These studies have shown that the silicification 
process follows a series of interlinked but unavoidable steps, starting with the microbes 
reacting to highly supersaturated silica solutions by increasing their production of 
exopolymeric material. Simultaneously with this process, but driven inorganically, 
silica polycondensation is proceeding, with monomers condensing to dimers, trimers 
etc., leading to the formation of critical silica nanospheres. The newly formed, 'sticky', 
exopolymeric sugars do not affect this polymerization, but will subsequently enhance 
the aggregation of the inorganically formed nanospheres on the microbial surface. 
Finally, this will invariably lead to the full silicification of the organic microbial 
frameworks and the formation of microfossils that can thus be preserved in modern 
silica sinter environments and that provide a modern analogue to processes in the 
ancient past. 

ACKNOWLEDGEMENTS 

Support for L. G. B. from the Leverhulme Trust (ref #F/00122/F) and the Natural Environment 
Research Council (GR9/04623) is gratefully acknowledged. V.R.P. acknowledges the kind 
support of Professor Terry Beveridge. B. W. M. acknowledges funding from the Foundation for 
Research in Science and Technology (contract C05X0201) and from the NSOF Extremophiles 
Programme. 

SGM symposium 65 



146 L. G. Benning and others 



REFERENCES 



Alexander, G. B. (1975). The effect of particle size on the solubility of amorphous silica in 

water. J Phys Chem 61,1 563-1 564. 
Aplin, K. R. (1987). The diffusion of dissolved silica in dilute aqueous solution. Geochim 

Cosmochim Acta 51,21 47-2 151. 
Baeuerlein, E. (editor) (2000). Biomineralization: from Biology to Biotechnology and 

Medical Application. Weinheim, Germany: Wiley-VCH. 
Benning, L. G. & Mountain, B. W. (2004). The silicification of microorganisms: 

a comparison between in situ experiments in the field and in the laboratory. 

In Proceedings of the 11th International Symposium on Water-Rock Inter- 
actions, pp. 3-10. Edited by R. Wanty, R. Seal & A. A. Balkema. London: Taylor & 

Francis. 
Benning, L G., Phoenix, V. R., Yee, N. &Tobin, M. J. (2004a). Molecular characterization 

of cyanobacterial silicification using synchrotron infrared micro-spectroscopy. 

Geochim Cosmochim Acta 68, 729-741 . 
Benning, L. G., Phoenix, V. R., Yee, N. & Konhauser, K. O. (2004b). The dynamics of 

cyanobacterial silicification: an infrared micro-spectroscopic investigation. Geochim 

Cosmochim Acta 68, 743-757. 
Cady, S. L. (2001). Paleobiology of the Archean. Adv Appl Microbiol 50, 3-35. 
Cady, S. L. & Farmer, J. D. (1996). Fossilization processes in siliceous thermal springs: 

trends in preservation along thermal gradients. Ciba Found Symp 202, 1 50-173. 
Carroll, S., Mroczek, E., Alai, M. & Ebert, M. (1998). Amorphous silica precipitation (60 

to 120 °C): comparison of laboratory and field rates. Geochim Cosmochim Acta 62, 

1379-1396. 
Chen, C. H. & Lewin, J. (1969). Silicon as a nutrient element for Equisetum arvense. Can J 

Bot47, 125-131. 
Cloud, P. E. (1965). Significance of the Gunflint (Precambrian) microflora. Science 148, 

27-35. 
Dixit, S., Van Capellen, P. & van Bennekom, A. J. (2001). Processes controlling solubility 

of biogenic silica and pore water build-up of silicic acid in marine sediments. Mar 

Chem 73, 333-352. 
Dove, P. M. & Rimstidt, J. D. (1994). Silica-water interactions. In Silica: Physical Behaviour, 

Geochemistry and Materials Applications (Reviews in Mineralogy 29), pp. 259-308. 

Edited by P. J. Heaney, C. T Prewitt & G. V. Gibbs. Washington, DC: Mineralogical 

Society of America. 
Drews, G. & Weckesser, J. (1982). Function, structure and composition of cell walls and 

external layers. In The Biology of Cyanobacteria, pp. 333-357. Edited by N. G. Carr & 

B. A. Whitton. Oxford: Blackwell Scientific Publications. 
Dudman, W. F. (1977). The role of surface polysaccharides in natural environments. In 

Surface Carbohydrates of the Prokaryotic Cell, pp. 357-414. Edited by I. Sutherland. 

London: Academic Press. 
Everett, D. H. (1988). Basic Principles of Colloid Science. London: Royal Society of 

Chemistry. 
Fein, J. B., Scott, S. & Rivera, N. (2002). The effect of Fe on Si adsorption by Bacillus subtilis 

cell walls: insights into non-metabolic bacterial precipitation of silicate minerals. 

Chem Geo! 182, 265-273. 
Ferris, F. G., Beveridge, T. J. & Fyfe, W. S. (1986). Iron-silica crystallite nucleation by 

bacteria in a geothermal sediment. Nature 320, 609-61 1 . 



SGM symposium 65 



Biosilicification by cyanobacteria 147 



Ferris, F. G., Fyfe, W. S. & Beveridge, T. J. (1988). Metallic ion binding by Bacillus subtilis: 

implications for the fossilization of microorganisms. Geology 16, 149-152. 
Francis, S., Margulis, L. & Barghoorn, E. S. (1978). On the experimental silicification of 

microorganisms. II. On the time of appearance of eukaryotic organisms in the fossil 

record. Precambrian Res 6, 65-1 00. 
Garcia-Pichel, F. & Castenholz, R. W. (1991). Characterization and biological implications 

of scytonemin, a cyanobacterial sheath pigment. J Phycol 27, 395-409. 
Gedde, U. W. (1995). Polymer Physics. London: Chapman & Hall. 
Gunnarsson, I. & Arnorsson, S. (2000). Amorphous silica solubility and the thermodynamic 

properties of H 4 Si04 in the range of 0° to 350 °C at P sat . Geochim Cosmochim Acta 

64, 2295-2307. 
Heijnen, C. E., Hok-A-Hin, C. H. & van Veen, J. A. (1 992). Improvements to the use of 

bentonite clay as a protective agent, increasing survival levels of bacteria introduced 

into soil. Soil Biol Biochem 24, 533-538. 
Herdianita, N. R., Browne, P. R. L, Rodgers, K. A. & Campbell, K. A. (2000). 

Mineralogical and morphological changes accompanying ageing of siliceous sinter. 

Mineral Deposita 35, 48-62. 
Herdman, M. & Rippka, R. (1988). Cellular differentiation: hormogonia and baeocytes. 

Methods Enzymol 167, 232-242. 
Hildebrand, M. & Wetherbee, R. (2003). Components and control of silicification in 

diatoms. Prog Mol Subcell Biol 33, 11-57. 
Hoiczyk, E. (1998). Structural and biochemical analysis of the sheath of Phormidium 

uncinatum . J Bacteriol 180, 3923-3932. 
Hoiczyk, E. & Hansel, A. (2000). Cyanobacterial cell walls: news from an unusual 

prokaryotic envelope. J Bacteriol 182, 1 191-1 199. 
Hulbert, S. F. (1969). Models for solid-state reactions in powder compacts: a review. 

J Br Ceramic Soc 6, 1 1 -20. 
Hunter, R. J. (1996). Introduction to Modern Colloid Science. New York: Oxford University 

Press. 
Icopini, G. A., Brantley, S. L. & Heaney, P. J. (2005). Kinetics of silica oligomerization 

and nanocolloid formation as a function of pH and ionic strength at 25 °C. Geochim 

Cosmochim Acta 69, 293-303. 
Her, R. K. (1979). The Colloid Chemistry of Silica and Silicates. Ithaca, NY: Cornell University 

Press. 
Her, R. K. (1980). Isolation and characterization of particle nuclei during the polymerization 

of silicic acid to colloidal silica. J Colloid Interface Sci 75, 138-148. 
Jamtveit, B. & Meakin, P. (editors) (1999). Growth, Dissolution and Pattern Formation 

in Geosystems. Dordrecht: Kluwer. 
Jones, B., Renaut, R. W. & Rosen, M. R. (1998). Microbial biofacies in hot-spring sinters: a 

model based on Ohaaki Pool, North Island, New Zealand. J Sediment Res 68, 

413-434. 
Jones, B., Renaut, R. W. & Rosen, M. R. (2000). Stromatolites forming in acidic hot-spring 

waters, North Island, New Zealand. Palaios 15, 450-475. 
Jones, B., Renaut, R. W. & Rosen, M. R. (2001). Taphonomy of silicified filamentous 

microbes in modern geothermal sinters- implications for identification. Palaios 16, 

580-592. 
Jurgens, U. J. & Mantele, W. (1991). Orientation of carotenoids in the outer membrane 

of Synechocystis PCC 6714 (cyanobacteria). Biochim Biophys Acta 1067, 208- 

212. 



SGM symposium 65 



148 L. G. Benning and others 



Kasting, J. F. (1987). Theoretical constraints on oxygen and carbon dioxide concentrations 

in the Precambrian atmosphere. Precambrian Res 34, 205-229. 
Knauth, L. P. & Epstein, S. (1 982). The nature of water in hydrous silica. Am Mineral 67, 

510-520. 
Konhauser, K. O. (2000). Hydrothermal bacterial biomineralization: potential modern-day 

analogues for banded iron-formations. In Marine Authigenesis: From Global to 

Microbial (Society for Sedimentary Geology Special Publication no. 66), pp. 133-145. 

Edited by C. R. Glenn, J. Lucas and L. Prevot. Tulsa, OK: Society for Sedimentary Geology. 
Konhauser, K. O. & Ferris, F. G. (1996). Diversity of iron and silica precipitation by microbial 

mats in hydrothermal waters, Iceland: implications for Precambrian iron formations. 

Geology 24, 323-326. 
Konhauser, K. O., Phoenix, V. R., Bottrell, S. H., Adams, D. G. & Head, I. M. (2001). 

Microbial— silica interactions in Icelandic hot spring sinter: possible analogues for 

some Precambrian siliceous stromatolites. Sedimentology 48, 41 5-433. 
Kroger, N., Lehmann, G., Rachel, R. & Sumper, M. (1997). Characterization of a 200-kDa 

diatom protein that is specifically associated with a silica-based substructure of the 

cell wall. EurJBiochem 250, 99-105. 
Kroger, N., Deutzmann, R., Bergsdorf, C. & Sumper, M. (2000). Species-specific 

polyamines from diatoms control silica morphology. Proc Natl Acad Sci U 5 A 97, 

14133-14138. 
Langer, K. & Florke, O. W. (1974). Near infrared absorption spectra (4000-9000 cm" 1 ) of 

opals and the role of "water" in these Si0 2 . nH 2 minerals. Fortschr Miner 52, 

17-51. 
Leo, R. F. & Barghoorn, E. S. (1976). Silicification of wood. Bot Mus Leaf! Harv Univ 

25, 1-47. 
Lin, M. Y, Lindsay, H. M., Weitz, D. A., Ball, R. C, Klein, R. & Meakin, P. (1990). 

Universal reaction-limited colloid aggregation. PhysRevA 41, 2005-2020. 
Martin, J. E. (1987). Slow aggregation of colloidal silica. PhysRevA 36, 341 5-3426. 
Martin, J. E., Wilcoxon, J. P., Schaefer, D. & Odinek, J. (1990). Fast aggregation of 

colloidal silica. Phys Rev A 41, 4379-4391 . 
Mountain, B. W., Benning, L. G. & Boerema, J. A. (2003). Experimental studies on New 

Zealand hot spring sinters: rates of growth and textural development. Can J Earth Sci 

40, 1643-1667. 
Nielsen, A. E. (1959). The kinetics of crystal growth in barium sulfate precipitation. II. 

Temperature dependence and mechanism. Acta Chem Scand 13, 784-802. 
Oehler, J. H. & Schopf, J. W. (1971). Artificial microfossils: experimental studies of 

permineralization of blue-green algae in silica. Science 174, 1229-1231 . 
Pancost, R. D., Pressley, S., Coleman, J. M., Benning, L G. & Mountain, B. W. (2005). 

Lipid biomolecules in silica sinters: indicators of microbial biodiversity. Environ 

Microbiol 7 ,66-77. 
Perry, C. C. (2003). Silicification: the processes by which organisms capture and mineralize 

silica. Rev Mineral Geochem 54, 291-327. 
Perry, C. C. & Fraser, M. A. (1991). Silica deposition and ultrastructure in the cell wall of 

Equisetum arvense: the importance of cell wall structures and flow control in 

biosilicification. Philos Trans R Soc Lond B Biol Sci 334, 1 49-1 57. 
Perry, C. C. & Keeling-Tucker, T. (2000). Biosilicification: the role of the organic matrix in 

structural control. J Biol Inorg Chem 5, 537-550. 
Phoenix, V. R., Adams, D. G. & Konhauser, K. O. (2000). Cyanobacterial viability during 

hydrothermal biomineralization. Chem Geo! 169, 329-338. 



SGM symposium 65 



Biosilicification by cyanobacteria 149 



Phoenix, V. R. ; Konhauser, K. O., Adams, D. G. & Bottrell, S. H. (2001). Role of bio- 
mineralization as an ultraviolet shield: implications for Archean life. Geology 29, 

823-826. 
Phoenix, V. R., Konhauser, K. O. & Ferris, F. G. (2003). Experimental study of iron and silica 

immobilization by bacteria in mixed Fe-Si systems: implications for microbial 

silicification in hot springs. Can J Earth Sci 40, 1669-1678. 
Pontoni, D., Narayanan, T. & Rennie, A. R. (2002). Time-resolved SAXS study of 

nucleation and growth of silica colloids. Langmuir 18, 56-59. 
Rees, D. A. (1977). Polysaccharide Shapes. London: Chapman & Hall. 
Resch, C. M. & Gibson, J. (1983). Isolation of the carotenoid-containing cell wall of three 

unicellular cyanobacteria. J Bacteriol 155, 345-350. 
Rimstidt, J. D. & Barnes, H. L. (1980). The kinetics of silica-water reactions. Geochim 

Cosmochim Acta 44, 1 683-1 699. 
Rimstidt, J. D. & Cole, D. R. (1983). Geothermal mineralization. I. The mechanism of 

formation of the Beowawe, Nevada, siliceous sinter deposit. Am J Sci 283, 861-875. 
Rippka, R., Deruelles, J., Waterbury, J. B., Herdman, M. & Stanier, R. Y. (1979). Generic 

assignments, strain histories and properties of pure cultures of cyanobacteria. J Gen 

Microbiol 111 , 1-61. 
Sangster, A. G. & Hodson, M. J. (1986). Silica in higher plants. Ciba Found Symp 121, 

90-1 1 1 . 
Schrader, M., Drews, G. & Weckesser, J. (1981). Chemical analysis on cell wall constituents 

of the thermophilic cyanobacterium Synechococcus PCC671 6. FEMS Microbiol Lett 

11,37-40. 
Schultze-Lam, S., Ferris, F. G., Konhauser, K. O. & Wiese, R. G. (1995). In-situ silicification 

of an Icelandic hot spring microbial mat: implications for microfossil formation. Can J 

Earth So 32, 2021-2026. 
Scott, C, Fletcher, R. L. & Bremer, G. B. (1996). Observations of the mechanisms of 

attachment of some marine fouling blue-green algae. Biofouling 10, 161-173. 
Siever, R. (1992). The silica cycle in the Precambrian. Geochim Cosmochim Acta 56, 

3265-3272. 
Simpson, T. L. & Volcani, B. E. (editors) (1981). Silicon and Siliceous Structures in Biological 

Systems. New York: Springer. 
Toporski, J. K. W., Steele, A., Westall, F., Thomas-Keprta, K. L. & McKay, D. S. (2002). 

The simulated silicification of bacteria - new clues to the modes and timing of 

bacterial preservation and implications for the search for extraterrestrial microfossils. 

Astrobiologyl, 1-26. 
Treguer, P., Nelson, D. M., van Bennekom, A. J., DeMaster, D. J., Leynaert, A. & 

Queguiner, B. (1995). The silica balance in the world ocean - a reestimate. Science 

268, 375-379. 
Tsuneda, S., Aikawa, H., Hayashi, H., Yuasa, A. & Hirata, A. (2003). Extracellular 

polymeric substances responsible for bacterial adhesion onto solid surface. FEMS 

Microbiol Lett 223, 287-292. 
van der Meer, M. T., Schouten, S., Hanada, S., Hopmans, E. C, Damste, J. S. & Ward, 

D. M. (2002). Alkane-1,2-diol-based glycosides and fatty glycosides and wax 

esters in Roseiflexus castenholzii and hot spring microbial mats. Arch Microbiol 

178, 229-237. 
Walter, M. R., Bauld, J. & Brock, T. D. (1972). Siliceous algal and bacterial stromatolites 

in hot spring and geyser effluents of Yellowstone National Park. Science 178, 

402-405. 



SGM symposium 65 



150 L. G. Benning and others 



Walters, C. C, Margulis, L. & Barghoorn, E. S. (1977). On the experimental silicification of 
microorganisms. I. Microbial growth on organosilicon compounds. Precambrian Res 
5,241-248. 

Weckesser, J., Hofmann, K., Jurgens, U. J., Whitton, B. A. & Raffelsberger, B. (1988). 
Isolation and chemical analysis of the sheaths of the filamentous cyanobacteria 
Calothrix parietina and C scopulorum. J Gen Microbiol 134, 629-634. 

Weed, W. H. (1889). Formation of travertine and siliceous sinter by the vegetation of 
hot springs. In United States Geological Survey Ninth Annual Report (1 887-1 888), 
pp. 613-676. Washington, DC: United States Geological Survey. 

Westall, R, Boni, L. & Guerzoni, E. (1995). The experimental silicification of micro- 
organisms. Palaeontology 38, 495-528. 

Yee, N., Phoenix, V. R. ; Konhauser, K. O., Benning, L. G. & Ferris, F. G. (2003). The effect 
of cyanobacteria on silica precipitation at neutral pH: implications for bacterial 
silicification in geothermal hot springs. Chem Geol 199, 83-90. 



SGM symposium 65 



Metabolic diversity in the 
microbial world: relevance to 
exobiology 

Kenneth H. Nealson and Radu Popa 

Department of Earth Sciences, University of Southern California, Los Angeles, CA 90089-0740, 
USA 



INTRODUCTION 

Metabolic diversity is used here as a physiological or ecological concept referring to 
the metabolic repertoire available to any group of organisms: in this case, microbes. At 
least for now, metabolic diversity is conceptually distinct from genetic diversity, 
although one imagines that, as both concepts are understood in greater depth, the 
relationships between them will become clear. The metabolic repertoire encompasses, 
for the most part, the entire range of redox-related energy sources that are available on 
our planet, from photochemistry to organic and inorganic redox chemistry. Earthly 
microbes have 'learned' to harvest the energy of nearly every useful and abundant redox 
couple, revealing a nutritional versatility that to some extent could be used to describe 
what the planet has to offer energetically To turn this around, one can imagine that, if 
energy sources were defined for Earth, one might well predict what kinds of metabolism 
should have evolved to exploit them and, in fact, for the most part, this would lead 
to the correct answer. Metabolic diversity is further accentuated by various symbioses, 
syntrophisms and community interactions (intracellular, intercellular and inter- 
population), leading to the establishment of communities with seemingly new and 
unexpected abilities. The functional diversity of the prokaryotic world is thus expressed 
in terms of its redox chemistry and, with regard to geobiology, this redox chemistry/ 
metabolic connection defines a wide variety of relevant reactions, many of which 
involve phase changes of the interacting molecules (i.e. between solid, liquid and gas 
phase). In some sense, it is the world of 'accidental biominerals'; a world where 
minerals are formed, altered or dissolved as a result of the forever-starved prokaryotic 
world. In contrast, many eukaryotic microbes also exert significant effects on the 

SGM symposium 65: Micro-organisms and Earth systems - advances in geomicrobiology. 
Editors G. M. Gadd, K. T. Semple & H. M. Lappin-Scott. Cambridge University Press. ISBN 521 86222 1 ©SGM 2005 



152 K. H. Nealson and R. Popa 

geological landscape, and these work by completely different metabolic modes - 
forming minerals by directed, energy-consuming reactions, minerals that have altered 
and continue to alter the face of our planet. All in all, the way that life has adapted to 
the energy regime of our planet allows the exobiological prediction that the same 
should be true of any other abode where life resides. Understanding energy types and 
flows may well provide a critical pathway for life detection, on or off our own planet. 

MICROBIAL DIVERSITY 
Genetic diversity 

Genetic diversity is estimated by any of a number of methods, often related to DNA 
sequence similarities and differences. For the sake of this discussion, the microbial 
world has been divided into two groups, the prokaryotes and eukaryotes (Fig. 1), with 
the former group consisting of the Bacteria and the Archaea (Woese, 2004; Woese et al., 
1984, 1990) and the latter including a wide array of single-celled, microscopic 
eukaryotes, collectively called the protists or protoctista (Whittaker & Margulis, 1978). 
This grouping is based primarily on sequence comparisons of 16S rRNA genes and 
includes sequences of both cultivated and as-yet uncultivated microbes. Both the 
prokaryotic and eukaryotic microbes are underrepresented with regard to cultivated 
members, so that any discussion of metabolic diversity of these groups is of necessity 
strongly biased towards a very small subset of the whole. For example, it is estimated 
that only 0-1 to 1 % of the prokaryotes in most environments have been cultivated 
(Whitman et al., 1998). In fact, there are many phylum-level bacterial taxonomic 
groups that are represented by DNA sequences of 16S rRNA genes alone: i.e. groups 
for which no species have yet been cultivated (Pace et al., 1986). While it is obviously 
inappropriate to make ecological extrapolations based on 1 % or less of the popu- 
lation, in many cases it is the best that has been possible. 

Metabolic diversity 

Metabolic diversity, on the other hand, can be assessed by measurement of the reactants 
and products and quantified in the laboratory and the field. Just what organisms are 
responsible for a given activity is often discovered after the process has been demon- 
strated in natural samples. Thus, the connection between geochemists and micro- 
biologists has been a strong one for many years and, in some cases, the geochemists 
have observed the processes for which the causative microbes were later discovered and 
characterized. Some examples of this are the processes of anaerobic methane oxidation 
(Boetius etaL, 2000; Orphan et al., 2001b), anaerobic ammonia oxidation (Jetten et al., 
1998) and microbial iron and manganese reduction (Lovley & Phillips, 1988; Myers & 
Nealson, 1988), all of which were known in the geochemical literature prior to the 
microbiology being demonstrated. The metabolic diversity of the prokaryotes, how- 

SGM symposium 65 



Metabolic diversity and exobiology 1 53 



Prokaryotes 



f~ 



Bacteria 



Archaea 

-A. 



>r 



Eukaryotes 



BjghGC 

Gram- 
positives 



Spirochaetes 



.2 

g 



bpsiton 



Alpha 

Beta 

Gamma 




Flagellates 



7 he rm odesulfotobacterium 

Thermotogales 

Aquifex 



Fig. 1. Microbial diversity according to molecular phylogeny. This diagram shows some of the range 
of microbes as estimated by sequence similarity of 1 6S rRNA gene sequences. This genetically diverse 
assemblage is dominated by bacteria and archaea, but includes a large number of very important 
eukaryotic microbes such as carbonate- and silica-depositing algae, bacteriovores (ciliates and 

flagellates) and photosynthetic algae. 



ever, extends beyond individual abilities, as some of the truly unique things done are 
accomplished not by single microbes, but by collaborative efforts between microbes 
that specialize in the intercellular metabolite transfer called syntrophism, thus 
providing even more functional diversity than might be expected from single cells. In 
fact, many of the metabolic abilities of eukaryotes (photosynthesis, respiration, 
lithotrophy and nitrogen fixation) are carried out by symbionts acquired from the 
prokaryo tic world (Margulis, 1981). 

One way to view metabolic diversity is shown in Table 1, which illustrates the range 
of redox reactions in which microbes participate. As can be seen, the prokaryotes 
dominate this scene: only a few organics and no inorganics can be utilized by 
eukaryotes, and, with the exception of nitrate and possibly fumarate, which can be used 
by a few anaerobic fungi (Fenchel & Finlay; 1995; Tielens et al., 2002; Zumft, 1997), 
only one respiratory electron acceptor (oxygen) can be utilized by the eukaryotes. 
Purists might argue that even oxygen respiration is an acquired trait - one that was 
'invented' by prokaryotes and arose in the eukaryotes via symbiosis (Margulis, 1981), 
making respiration a hallmark trait of the prokaryotes and symbiosis, perhaps, one of 
the most important parts of eukaryotic evolution. The point is, however, that, with 



SGM symposium 65 



154 K. H. Nealson and R. Popa 



Table 1. Energy sources (fuels) and oxidants used by life 

Here we see some of the range of electron donors and electron acceptors used by all of life. The 
metabolic diversity of the prokaryotes (which can use all those listed) is dramatically expanded in 
comparison with the rather constrained eukaryotes (which can use those entries in bold only). 
Asterisks denote those molecules that form insoluble compounds (minerals) when oxidized or 
reduced, connecting this redox-driven biochemistry with the geological world. DMSO, Dimethyl 
sulfoxide; TMAO, trimethylamine A/-oxide. 



Fuels 




Oxidants 




Sunlight 




r 


Fu ma rate 




f Glucose 


Organics < 


DMSO 


Organics < 


Ethanol 

Formaldehyde 


I 

Carbon dioxide* 


TMAO 




l Methanol 


Sulfur* 




Hydrogen 




Sulfate* 




Ammonia 




Arsenate* 




Hydrogen sulfide* 


Selenite* 




Sulfur* 




Iron* 




Iron* 




Manganese* 




Manganese* 




Nitrate 




Carbon monoxide 
Arsenite* 


;* 


Oxygen 





regard to redox chemistry, the prokaryotic microbes are the experts, while the eukary- 
otes have shunned redox diversity in favour of a high energy yield - energy used to 
power their own impressive diversity of cellular structure and behaviour. The most 
pertinent features with regard to this are: (i) the use of inorganic energy sources (litho- 
trophy) is the realm of the prokaryotes; (ii) the process of anaerobic respiration (i.e. 
using electron acceptors other than oxygen) is, with few exceptions, the realm of the 
prokaryotes; (iii) the oxidation and reduction of these inorganic compounds forms 
a strong link with planetary geology, as many of the reactions either form or dissolve 
solid-phase minerals and mineraloids during the process (Table 1). 

Having extolled prokaryotic diversity, a few words of respect and admiration need to 
be inserted with regard to the anaerobic eukaryotes. Both protists and fungi are often 
abundant in many different permanently anaerobic niches (i.e. rumens, termite guts, 
anaerobic sediments), where they feed on prokaryotes and their breakdown products 
(Fenchel & Finlay, 1995). These eukaryotes are almost without exception devoid of 
mitochondria and exist either by the use of hydrogenosomes or via nitrate reduction. 
Hydrogenosomes, which are common in both anaerobic protists and fungi, are organ- 
elles believed to have arisen from mitochondria (Theissen et al., 2003) - they have 
replaced the enzymes involved in oxidative phosphorylation with those involved in 



SGM symposium 65 



Metabolic diversity and exobiology 155 



(a) 



Light 




Cell membrane 



R hod opsin => 




(b) 



Light 
H ; \ 2 

V ±_ / 2e*2H 




Q -Q, 



Chlorophyll 
PS II (P680) 




FeS Chlorophyll 
PS I (P700) 



ADP+Pi 



NADPH, 



* ^ATPr 



Fig. 2. Mechanisms for harvesting light energy, (a) Rhodopsin- mediated proton translocation 
as seen in archaea (Halobacterium halobium). It should be noted that the proton gradient that is 
established by this process can be used to power many cellular functions such as transport or flagellar 
motion or, as shown, to power the synthesis of ATP via the membrane-bound ATPase. (b) Oxygenic 
photosynthesis in chloroplasts. A and A-,, Acceptors of electrons from P700; bf, cytochrome bf 
complex; Fd, ferredoxin; Fp, flavoprotein; PC, plastocyanin; Ph, phaeophytin; Pq, plastoquinone; 
PS, photosystem; Q A and Q B , plastoquinone-binding proteins. 

fermentation, thus creating an organelle that makes ATP by substrate-level phosphory- 
lation and produces dihydrogen gas as a by-product (Martin & Muller, 1998; Muller, 
1993). 

A final point with regard to diversity relates to the use of light as an energy source. By 
a factor of about 50000, visible light dominates our planet in terms of yearly energy 
flow (Nealson & Conrad, 1999), and it is used in a number of ways. The simplest is via 
a process called rhodopsin-mediated proton pumping, in which the molecule rhodopsin 
is used to absorb light and the energy is used directly to pump a proton from the inside 
to the outside of the cell, charging the cell membrane via an osmotic and electrical 
gradient (Fig. 2a). This gradient can then be used by the cell either directly or to syn- 
thesize ATP. This type of metabolism, long known in archaea (Stoeckenius 6c Bogo- 
molni, 1982), has more recently been found to be widely distributed in the bacterial 
world (Beja et al., 2000; de la Torre et al., 2003; Venter et al., 2004). Photosynthesis, the 
more familiar use of visible light by life, was apparently 'invented' by the bacteria, 
almost certainly as an anaerobic process, using sulfur compounds or iron as electron 



SGM symposium 65 



156 K. H. Nealson and R. Popa 



Photosynthesis and the generation of oxidants 



hv 



C0 2 + r donor 



? 



c> 



(CH 2 0) + oxidized e~ donor 



ORGANISMS 



Anoxygenic photosynthesis 
Purple 

Green filamentous 
Heliobacter 
Green sulfur 

Oxygenic photosythesis 
Cyanobacteria 
Chloroplasts 
(plants and algae) 



e- donors 




H 2 



oxidized e~ donors 



> 



So, S0 4 2 ", 
Fe3 + 



c> 



Fig. 3. Generation of oxidants during photosynthesis. The general formula for photosynthesis is 
shown at the top of the diagram, with the critical difference being the molecules used as electron 
donors (and the resultant oxidized forms produced). 



Magnetite 

Vivianite (Fe.O.) 

(Fe 3 (P0 4 ) 2 8H 2 0) j 

p ttf e '&B? egg* 



Sidcritc 
(FcC0 3 ) 



Fe 



3+ 



Lactate 



Fe 



2+ 



Acetate + CO 




Fig. 4. Accidental biominerals. The redox activity of some bacteria (such as iron reducers) results in 
the formation of various minerals (vivianite, magnetite, siderite), depending primarily on the chemistry 
of the environment. 



SGM symposium 65 



Metabolic diversity and exobiology 1 57 

donor (Blankenship, 2002; Xiong et al., 2000). Such anoxygenic photosynthetic bacteria 
produce oxidized electron donors as a by-product of photosynthesis (Fig. 3) and were 
the precursors of the now widespread oxygenic photosynthesis (Blankenship, 2002). 
The more complex oxygenic photosynthesis is thought to have arisen in the cyano- 
bacteria and eventually appeared symbiotically in algae and plants (Margulis, 1981), 
one of the most important events in the evolution of complex life. 

With the above discussion of metabolic diversity, it is easy to miss an important point: 
namely that there is also an impressive unity to the metabolism of life. Virtually all 
present-day metabolism on Earth operates in a similar way, with environmental redox 
equivalents being harvested and used for the reduction of cellular electron or hydrogen 
carriers which are similar throughout all of life (Nealson, 1997; Nealson & Conrad, 
1999). These reduced carriers are then used directly for biosynthesis or for the gener- 
ation of a membrane potential (combination of a pH and electrical potential) that can 
be used for a variety of functions (including ATP formation, transport and motility). 
Thus, despite apparent rampant diversity in terms of resource allocation, the central 
energy-processing systems of all life, including those of structurally complex large 
eukaryotes, operate in a similar way. With regard to this issue, there are really only 
two metabolic means of extracting energy from the environment: chemiosmosis and 
fermentation. Almost without exception, the interactions between the living world 
and the mineral world are of the chemiosmotic type (i.e. redox chemistry). While it 
is difficult to specify exactly when the ability to respire arose in time, one imagines 
that it is indeed one of the earliest 'inventions' of life (Nealson 6c Rye, 2004) and, once 
invented, became a component of virtually all subsequent successful experiments in 
metabolic evolution. 

MINERAL FORMATION: THE GEOBIOLOGY CONNECTION 
Prokaryotic mineral formation 

As discussed above, the array of electron donors and acceptors that the prokaryotic 
world uses is, to some extent, the glue that knits microbial metabolism to geobiology. 
As shown in Table 1, many of the redox-active compounds become a solid (mineral or 
mineraloid) form - thus, oxidation or reduction of these components often results in a 
change of state of the component (i.e. between insoluble and soluble), leading to 
formation or dissolution of minerals. In this sense, the metabolism of the prokaryotes, 
designed for the purpose of harvesting energy from the environment, is inadvertently 
linked to the formation and/or dissolution of many minerals. The inadvertent nature of 
this process is shown in Fig. 4, a summary of results in which a culture of Shewanella 
putrefaciens CN-32, during the process of iron reduction, was shown to produce any of 
several mineral products depending on the environment in which the bacteria are grown 

SGM symposium 65 



158 K. H. Nealson and R. Popa 



Vesicles Organic Nucleation 

(cisternae) scales of calcite 




Growth of calcite 
into coc eoliths 



Calcite crystallites Radial ' Kxocytosis 

(heterococcoliths) organization of of coccoliths 

heterococcoliths 



Coccoliths with 
organic coating 



Fig. 5. Purposeful biominerals. Biomineralization of coccolithophores (calcareous algae) making 
crystals of calcite. This diagram is intended to demonstrate the complex and detailed interactions that 
occur in general in the formation of mineral products by eukaryotes (after Young & Henriksen, 2003). 

(Roden & Zachara, 1996). That is, while the bacteria metabolically impact the local 
mineralogy, the end products produced are controlled by the environment rather 
than by the bacteria themselves. Thus, to some extent, one can view the formation of 
minerals by some (perhaps most) of the prokaryotes as metabolic 'accidents', depend- 
ent on the metabolic chemistry, but not directed by it. As far as is known, there is 
no direct advantage to the bacterium of forming a given mineral in comparison to 
any other: this is determined in fact by the chemistry of the environment in which the 
bacteria are living. This leaves us in the interesting position of viewing prokaryotic 
minerals as a metabolic by-product - something done 'accidentally', and something, 
perhaps, that occurs as an indicator of microbial metabolism, but without the morpho- 
logical clues usually associated with biominerals or fossils. 

This being said, while the formation of a specific mineral may have no advantage for 
prokaryotes, it may well be that mineral formation in general serves a useful thermo- 
dynamic/kinetic purpose. From the point of view of the bacterium, removing one of the 
soluble products as an insoluble mineral could have a dynamic effect on the chemical 
abilities of the bacterium. This is a strategy often employed by bacteria living in sym- 
biosis with other bacteria, in which the end product of one is the substrate for another, 
and this cascade makes chemical reactions proceed at speeds far beyond those predicted 



SGM symposium 65 



Metabolic diversity and exobiology 1 59 

if the reaction products were disposed of by diffusion alone. Perhaps the most pertinent 
example is that of anaerobic methane oxidation, in which the process is really driven by 
sulfate-reducing bacteria that consume the hydrogen produced via methane oxidation, 
pushing the reaction forward (Boetius et al., 2000; Orphan et al., 2001a, b). 

Eukaryotic mineral formation 

The eukaryotes, in contrast, have perfected the art of mineral fabrication, making a 
wide array of biominerals with many different uses (Dove et al., 2003; Lowenstam, 
1981; Lowenstam & Weiner, 1989). Biomineralization is extremely common among the 
multicellular eukaryotes, many of which need structural elements to grow and function 
in three dimensions, providing advantages for predation (teeth, bones etc.) and protec- 
tion (shells, frustules, cell coverings) (Lowenstam, 1981; Lowenstam & Weiner, 1989). 
These biominerals are often ornate and recognizable structures that provide us with a 
visible fossil record, allowing the calibration of the molecular record (discussed below), 
something that is extremely difficult prior to the 'invention' of these biominerals. While 
it is less widely distributed among the eukaryotic microbes, organisms like coccolitho- 
phores and foraminifera (organisms that make carbonate minerals) and diatoms and 
radiolarians (organisms that make silica minerals) have dramatic impacts on the carbon 
and silicon cycles on the planet (Lowenstam & Weiner, 1989; Skinner & Jahren, 2004). 

It is clear from emerging mechanistic studies that the situation is dramatically different 
with regard to eukaryotic mineral formation (Dove et al., 2003). Eukaryotic bio- 
minerals are genetically directed structures, formed using protein templates that are 
used to catalyse and direct the synthesis of the specific minerals. The resulting struct- 
ures are reproducible enough that they can often be used to identify the organism that 
produced them, often to the level of genus or even species (Fig. 5). This is in marked 
contrast to prokaryotic mineral formation and dissolution, which is primarily a func- 
tion of environment rather than genetics (Fig. 3). As opposed to the prokaryotes, which 
form minerals while gaining metabolic energy, the formation of eukaryotic biominerals 
is 'costly' in the sense of requiring specific templates, energy for synthesis and often 
auxiliary systems for transport and assembly. Finally, redox chemistry is not a funda- 
mental part of eukaryotic mineral formation, as redox changes do not occur and no 
energy is gained. Given that the eukaryotes are unable to view minerals as either viable 
electron donors or acceptors, their ability to form or dissolve redox-active minerals is 
extremely limited. 

A final note with regard to the minerals produced by prokaryotes and eukaryotes: 
because protein templates and other structural aids are not involved in prokaryotic 
mineral formation, these minerals are often 'pure', sulfides, carbonates etc., while those 
of eukaryotes are 'hybrids', composed of biological material interspersed with 

SGM symposium 65 



160 K. H. Nealson and R. Popa 





* 





Fig. 6. Chains of magnetosomes in Magnetospirillum magnetotacticum AMB1 as seen in a high- 
magnification TEM section (image provided courtesy of Dr Virginia Souza-Egipsy). Bar 100 nm. 



crystalline minerals (Dove et ai, 2003). These adaptations add remarkable mechanical 
properties to the eukaryotic biominerals and make them chemically distinct from 
the prokaryotic biominerals (Dove et uL, 2003). This being said, it should also be noted 
that a wide variety of mineral and mineraloid-like inclusions (oxalates, carbonates, uric 
acid etc.) are neither species-specific nor 'hybrid' in nature. 

Biomagnetite and the death of a generalization 

The above discussion of mineral formation divides the world into prokaryotic redox 
minerals and eukaryotic structural and functional biominerals. For example, dissimi- 
latory iron-reducing bacteria can produce copious amounts of extracellular magnetite 
as a function of respiration (Lovley 6c Phillips, 1988; Roden & Zachara, 1996), while 
eukaryotic chitons can produce magnetite coatings on their teeth (Lowenstam, 1981; 
Lowenstam & Weiner, 1989). The eukaryotic magnetite is easily recognizable, but 
whether it is possible to distinguish between magnetite formed by prokaryotes and 
abiotically formed magnetite is not yet clear. 

The division between prokaryotes and eukaryotes with regard to biominerals is 
not always so clear, however. Many strains of magnetotactic bacteria (no magneto- 
tactic archaea are known) produce highly ordered, intracellular, crystalline magnetite 
inclusions called magnetosomes (Fig. 6). These single domain, mineralogically nearly 

SGM symposium 65 



Metabolic diversity and exobiology 161 

perfect, highly magnetic crystals are often arranged in chain-like arrays (Blakemore, 
1975; Bazylinski & Frankel, 2000). They provide the cells that contain them with the 
ability to align passively in a magnetic field, the response being that these highly motile 
cells move swiftly towards one magnetic pole or the other. It appears that each bacterial 
strain produces a characteristic magnetite pattern (size, shape and arrangement of the 
magnetosomes) and that specific genes (and, thus, specific proteins) are involved in 
the production of the magnetosomes (Schuler, 1999, 2004; Matsunaga & Okamura, 
2003). Furthermore, while the function(s) of the magnetosome is not yet fully proven, 
nearly everyone agrees that there must be an advantage to the cells that contain the 
magnetosomes, and most agree that it is connected with environmental sensing and 
location (Frankel etaL, 1997). 

Are there other such examples of highly structured, genetically dictated biominerals 
produced by bacteria? Recently, it was reported that highly crystalline forms of metal 
sulfides were produced by sulfate-reducing bacteria (Suzuki et al., 2002). Whether these 
are in fact structured with a 'purpose' is not known, but their highly crystalline 
structure is an intriguing finding; perhaps such preordained and structured prokaryotic 
minerals are more prevalent than was thought (Fortin, 2004). 

TIME: LINKING THE PAST WITH THE PRESENT 

The timing of the emergence of various metabolic abilities is one of the most enigmatic 
problems in the microbial world. The evidence available to us in the ancient rock record 
is, at best, akin to smudged fingerprints and, at worst, like footprints in the sand after 
the tide has come and gone. One of the great hopes of molecular phylogenetic 
approaches was that they would allow one to look back in time using sequence data: 
using these data to estimate when major metabolic 'inventions' occurred in the past. 
Such 'inventions' would include not only structural innovations visible in the fossil 
record, but metabolic inventions like respiration and photosynthesis and other 
prokaryotic specialities like nitrogen fixation, denitrification and sulfur oxidation. This 
hope has been realized to some extent but, realistically, the methods are probably good 
only as far back as there is a fossil record to support them. For example, Benner et al. 
(2002) discussed the use of various types of data to understand the evolution of 
metabolism, concluding that, within the last 50, perhaps, 100 million years, one can 
probably do this with some confidence, utilizing a combination of fossil, isotope, 
organic geochemical and molecular evolutionary clock records to infer past patterns of 
evolution (in this case, the evolution of certain sugar-fermentation abilities). Thus, to 
some extent, the fossil record is of limited use: true (recognizable) biominerals 
produced by eukaryotes are visible only ~500 million years ago, when the first sponge 
spicules and carbonate biominerals can be seen. These processes are thus geologically 
young (Li etaL, 1998; Nealson & Rye, 2004). 

SGM symposium 65 



162 K. H. Nealson and R. Popa 

Prior to the formation of these 'true' biominerals, the signatures that exist are primarily 
geochemical in nature. Thus, while prokaryotic minerals may not be readily recog- 
nizable by their morphologies or unique crystal structures, many can be judged to be 
of biological origin via the fractionation of isotopes during the 'supply' of chemical 
components for their formation. Kinetic fractionation occurs as a function of enzymic 
catalysis, leading to biological materials preferentially being composed of the 'lighter' 
isotopes. Thus light carbon is preferentially used by living organisms during carbon 
fixation and accumulates in the resulting biomass, while light sulfide is produced during 
sulfur or sulfate reduction and accumulates in sulfide minerals. These isotopic tracers 
have been of value in tracing the metabolic activities of modern organisms in both the 
laboratory and the field and provide a major tool for looking for indicators of 
metabolic activity in ancient samples. That is, it is possible, by using C and S isotopes, 
to see in the ancient rock record the appearance of processes that result in fractionation 
of these isotopes. Herein lies an important distinction that can be easily missed: while 
the isotopes may strongly suggest the existence of a process leading to fractionation, 
they cannot tell us unambiguously which process was involved and cannot be used to 
tell which organism or even group of organisms accounted for the fractionation. Unless 
one accepts this tenet, one can be easily fooled. 

Carbon isotopes have been among the most valuable and widely used of the isotopes 
with regard to life detection and definition. Carbon fractionation occurs during its 
reduction (fixation) from C0 9 to organic carbon by bacteria, algae and plants and to a 
much greater degree (i.e. light carbon) when C0 2 is fixed into methane by methano- 
genic archaea. However, even for this well-known and often-used system, deciphering 
the isotope signatures from ancient samples is difficult because (i) these pathways are 
varied and unknown, (ii) subsequent diagenetic reactions are not easy to specify and 
(iii) the signature of the source of carbon is seldom known. 

Even with these caveats, however, it seems clear from a variety of studies that bio- 
minerals can be traced to the early phases of Earth's history. Stable-isotope signatures of 
both carbon and sulfur suggest that metabolic activities were involved with the 
formation of minerals from very early times. Carbon isotope ratios ( lo C/ 12 C) have been 
used to suggest that carbon fixation may have existed as early as 3-8 Ga (billion years) 
ago (Mojzsis et al., 1996). While this number has been challenged, few would argue 
with 3-5 Ga for convincing evidence of carbon isotope signals in the ancient record. 
Similarly, sulfur isotopes ( 34 S/ 33 S) suggest that sulfur reduction of some kind was 
occurring 2-5 Ga and perhaps earlier (Canfield et ai, 2000; Shen et aL, 2001). 

Providing definitive fossil evidence from before the time of the biomineral-producing 
eukaryotes is difficult for several reasons: first, because the preservation of the materials 

SGM symposium 65 



Metabolic diversity and exobiology 163 

is often poor, making identification difficult, and second, because virtually none of the 
putative organisms seen in the samples are alive today. While they have similarities to 
other organisms, the nature of their behaviour and even their metabolism cannot be 
specified with reasonable certainty. Another discouraging development with regard 
to molecular evolution methods is the rampant appearance of examples in which it is 
now clear that the evolutionary 'clock' is neither constant (it can run at different rates) 
nor predictable (Doolittle et ai, 1996). Thus, two organisms that would be described 
as deeply branching and suspected of being of equal 'age' may be quite different 
because of differences in evolutionary clock speeds. For ancient samples, of the order 
of hundreds to thousands of millions of years and older, the situation gets even more 
uncertain. 

Another difficulty that is peculiar to the prokaryotes is that the 'fossils' used to identify 
them are reduced to either organic geochemicals (i.e. classes of chemicals peculiar to a 
certain type of cell) or isotope fractionation patterns indicative of a certain type of 
metabolism. Both analyses have their limitations. For example, while some classes 
of compounds can be identified as components of cyanobacteria in the modern world, 
there is no way of knowing with certainty that non-cyanobacterial organisms that 
contained these compounds were not present before the cyanobacteria arose. Similarly, 
when one sees isotope fractionation, such as the appearance of light sulfur, it is temp- 
ting to invoke the appearance of sulfate reduction (and sulfate-reducing bacteria) and, 
in fact, this is often done (Canfield et al., 2000). While the activity of sulfate-reducing 
bacteria would indeed explain the observed results, it is also true that sulfur can be 
fractionated by many other organisms, including sulfur-, polysulfide- and thiosulfate- 
reducers (Smock etal., 1998). 

As a footnote to this discussion, one must remember that much of the work with both 
organic biomarkers and isotope fractionations hinges on our knowledge of microbial 
physiology and the composition of cultured microbes. As noted above, it is estimated 
that less than 1 % of the microbial world has been obtained in culture (Whitman et al., 
1998), and therefore that we are often extrapolating from a very limited experimental 
base. To this end, we note in Table 2 that much progress has been made in recent years 
in uncovering novel organisms and novel metabolic abilities - microbes whose very 
existence was previously doubted or unknown. Given that methods are improving with 
regard to culturing microbes, one can hope that the situation will improve, but one 
must be cautious when trying to extrapolate backwards in time to ancient metabolisms 
based on such an incomplete database. 

Finally, we must confront the issue of lateral (horizontal) gene transfer (Doolittle, 
2002). With the advent of genome sequencing, it has become apparent that in the 

SGM symposium 65 



164 K. H. Nealson and R. Popa 



Table 2. Microbial metabolic diversity 

This table represents some of the metabolic abilities of prokaryotes that have been revealed since 
1988. 



Process 



Found in 



Year of discovery Reference(s) 



Dissimilatory Fe and Mn 
reduction 

Anaerobic methane 
oxidation consortium 

Anaerobic ammonia 
oxidation 



Bacteria and archaea 



Bacteria/archaea 



Bacteria 



1988 



2000 



1998 



Lovley & Phillips (1 988); 

Myers & Nealson (1988) 

Boetiusefa/. (2000); 
Orphan et al. (2001b) 

Jettenefa/. (1998) 



Anaerobic iron oxidation 


Photosynthetic 


bacteria 


1993 


Ehrenreich&Widdel(1994); 
Widdelefa/. (1993) 


Anaerobic iron oxidation 


Heterotrophic I 


Dacteria 


1994 


Straubefa/. (1996) 


Perchlorate reduction 


Bacteria 




1996 


Coatesefa/. (1999) 


Phosphite oxidation 


Bacteria 




2000 


Schink & Friedrich (2000) 


Dissimilatory arsenic 
reduction 


Bacteria 




1994 


Ahmannefa/. (1994) 


Dissimilatory selenate 
reduction 


Bacteria 




1994 


Oremland etal. (1994) 



past there has been a large amount of mixing of genomes, such that the so-called 
phylogenetic 'tree of life' is in fact much more like a cross-hatched bush (Doolittle, 
2002). Each of the three 'kingdoms' has obtained, and fixed into their genomes, 
ample information from the other two kingdoms, suggesting that major sharing of 
genetic information has occurred, especially among the prokaryotes, where this is still 
happening at rapid rates. Thus, one of the major issues in prokaryotic genomics is that 
of defining the gene set that characterizes a given group of microbes (as opposed to 
those genes that can apparently move in and out with non-lethal results). 

With regard to the use of molecular methods, one of the intellectual 'traps' of this 
approach should be noted - namely that, for the prokaryotes, all of the organisms on 
which the 'phylogeny' is based are alive and evolving today. That is, while they may 
contain 'ancient' traits or abilities, these are surely not as they were in the past, so that 
the phylogenetic approach allows one to look at the most likely sequence of events, but 
not to place accurate times on any of the events. Thus, one could argue that we are 
looking at the living remnants of past microbial evolution and can have a reasonably 
good idea of what preceded what (within the limitations of uncertainties introduced 
by horizontal gene transfer), but trying to ascertain when any of these processes arose, 
i.e. when a given process or organism first appeared, is very difficult (if not impossible) 
by this method. As a precaution, one might note that almost none of the organisms 
seen in the fossil record of 100 million years ago are alive today If we tried to construct 



SGM symposium 65 



Metabolic diversity and exobiology 165 

those organisms simply from molecular biology alone, it would almost certainly be a 
resounding failure. This is the dilemma we are faced with in reconstructing prokaryotic 
evolution. 

Kinetic biosignatures and layered communities 

While the above discussion has dealt with inference of the past through reading the rock 
record, much biology is concerned with measurement and manipulation of the present. 
To this end, one can look for biosignatures connected to the chemistry of the organism 
and its impact on the environment (substrate removal or alteration and product 
deposition). Such short-term interactions define, to some extent, one of the major 
differences between life and non-life: specific catalysis of reactions that would other- 
wise occur at extremely slow rates. In the absence of life, many low-temperature 
geochemical reactions (i.e. less than 100 °C) proceed at rates slower than molecular 
diffusion, so that product accumulation and gradient formation cannot occur. Enzymic 
catalysis, however, leads to the consumption of reactants at rates faster than they can 
be supplied and the production of products at rates faster than they can diffuse 
away, leading to the formation of gradients that are indicative of the life forces that 
have produced them (Nealson & Berelson, 2003). The simple argument is that, in the 
absence of life, the chemical gradients (as we call them, layered microbial communities 
or LMCs) that are so dominant in anaerobic niches on Earth would simply not exist. 
For example, in the Black Sea (Fig. 7), a series of redox zones are seen, each indicative 
of a process that occurs in a well-defined redox zone (Nealson & Berelson, 2003). At 
these interfaces or layers, the chemical profiles can be used to define the microbial 
processes that are occurring to establish the gradients. For example, as shown in Fig. 7, 
the abrupt disappearance of oxygen at -50 m is referred to as the oxygen-depletion 
zone, where catalysis by aerobic respiration occurs so quickly that oxygen is taken to 
nearly zero within a few metres. In the absence of respiration, oxygen would be nearly 
constant to the bottom. Similarly, the nitrate disappears a few metres below due to a 
process called denitrification - without this, the nitrate would almost certainly be 
uniformly distributed in the Black Sea, as the rates of the processes that might consume 
it are very slow. Below this are the zones of manganese, iron and sulfate reduction, all 
of which would not exist without life catalysing each process. 

It should be obvious that these LMCs are inferred not from biological measurements, 
but from the very existence of the chemical layers and non-linear gradients (Nealson & 
Berelson, 2003). As will be discussed below, they are more complex than is implied 
above, but the gradients themselves can be used as indicators of specific catalysis and 
thus of life. Of interest here is that, with regard to spatial scales, we see such LMCs at 
scales of micrometres in biofilms to millimetres in algal mats and centimetres in lake 
and ocean sediments - they are a universal feature of life on Earth. In environments 

SGM symposium 65 



166 K. H. Nealson and R. Popa 







Black Sea (marine basin) 
Maximum (%) 
50 



100 







50 



100 r 



Q. 

Q 



150 - 



200 - 




O 



NO: 



Mn 



2 + 



NH + 



Fig. 7. Redox-related vertical profiles in the Black Sea. This profile presents a general scheme of redox 
interfaces seen in the Black Sea (Nealson & Berelson, 2003). The gradients here are a function of the 
organisms that accumulate at the interfaces - organisms that consume reactants and create products 
at rates sufficient to account for the gradients observed. Similar chemical gradients are used as 
indicators of LMCs over scales ranging from many metres, like this one, to millimetres or less in many 
other environments. 

where rapid mixing occurs, such as the open ocean, mixed lakes and the atmosphere, 
such gradients are rapidly dissipated by convection and mixing, but the signals of the 
biological processes are nevertheless there. 

EXTREME ENVIRONMENTS: WHY DO WE GO THERE? 

'Extremophiles': one of the most popular buzzwords of the last decade, and the cause 
for great excitement among microbiologists. Finding the 'most extreme' organism with 
regard to any given variable (temperature, pH, UV radiation etc.) has been a sporting 
pleasure for many of us. However, our obsession with these physical and chemi- 
cal extremes may be, to some extent, slightly misplaced. For many physical and 
chemical variables, the prokaryotes and eukaryotes are both able to adapt reasonably 



SGM symposium 65 



Metabolic diversity and exobiology 167 



= Eukaryotes 

Physical-chemical extremes □ ■ Prokaryotes 



| -20 Temperature +121 



L° 



pH +11-5 



dw Salinity saturate^ 

Nutritional extremes 




H 2 , H 2 S -- Inorganic energy sources- CO, Fe(ll) 



Q 2 , NO3-, Mn(IV) - Oxidants - S0 4 2 ~, CQ 2 | 



Fig. 8. Types of extremophily. This diagram shows the limits of prokaryotic and eukaryotic life for a 
number of variables. The top three panels are physical/chemical variables, demonstrating that the 
structurally simpler prokaryotes are a bit tougher and more resilient than eukaryotes, while the bottom 
three panels show that some types of metabolism are 'off limits' for the eukaryotes - truly extreme, 
dw, Distilled water. 

well (Fig. 8), although the simpler and more robust prokaryotes almost always exceed 
the more complex eukaryotes. As seen in Fig. 8, however, the true extreme conditions 
might be imagined to be those of nutrition. As noted above, eukaryotes might define as 
'acceptable' only those nutrients that can be converted into glucose or pyruvate, and 
only oxygen to respire. Thus, nearly the entire inorganic world of electron donors is 
'extreme', as is the use of any electron acceptor other than oxygen. When we move into 
the anaerobic world, especially the anaerobic world at high temperature, we begin 
to see the prokaryote-dominated world that we can imagine has some relevance to the 
ancient Earth. It should not be surprising to see that many of us who want to think 
about either the ancient Earth or the possibility of life elsewhere find ourselves studying 
environments like the deep subsurface - places where nutritional extremophiles are 
abundant and, for the most part, uncharacterized. 

CONCLUSIONS 

Metabolic diversity has become a catchphrase for the wonders of the microbial world - 
wonders that may help to explain why the world is like it is through the energetic 
exploitation of the geosphere via the metabolism of the biosphere. A direct, but not 
entirely obvious, extension of this thinking is that it applies not only to understanding 
the evolution and distribution of life on our own planet, but to the possibility of life in 

SGM symposium 65 



168 K. H. Nealson and R. Popa 

abodes other than Earth: giving rise to such questions as, 'could we use this geobiolo- 
gical logic to distinguish a living from a dead planet?' A quick perusal of the metabolic 
diversity of the prokaryotic life on our planet reveals that virtually every redox-related 
energetic niche is occupied. If there is fuel (an electron donor) and something with 
which to burn the fuel (an electron acceptor), then some microbe has learned how to 
exploit this energetic niche. This exploitation, in its simplest sense, defines the meta- 
bolic diversity with which we work - microbes striving to make a living on whatever 
resources are available, and being remarkably successful in plying their trade. Given 
that such metabolic diversity is so dominant on our own planet, one imagines that we 
should expect no less of any other planet on which life has evolved. If so, then the search 
for life in all realms (the surface and subsurface of the Earth, as well as potential 
extraterrestrial sites and samples) should begin with considerations of energy types, 
levels and fluxes, always with an eye to finding things that should not be there in the 
absence of life - the true connection of microbial diversity to exobiology. We don't 
know whether life exists elsewhere than on our planet. However, given the remarkable 
ability of life to adapt to and exploit earthly energy sources, it is an easy intellectual 
leap to imagine that, in other abodes, with abundant energy and acceptable chemistry, 
some sort of life could prosper. However, assumptions about the chemistry or other 
details of any non-earthly life may lead one awry. What must be true is that life will 
evolve to take advantage of the energy sources and chemical building blocks that are 
available and that, to compete with the natural chemistry, it will need to invent catalysts 
that elevate the rates of reactions. It is this catalysis that should be the signpost of life - 
the geobiological indicator that something is amiss (reactant consumption, product 
accumulation, complex molecule generation or isotope fractionation), and the enticing 
possibility that life may have been responsible. 



REFERENCES 



Ahmann, D. f Roberts, A. L, Krumholtz, L R. & Morel, F. M. M. (1994). Microbe grows 
by reducing arsenic. Nature37 4 \, 750. 

Bazylinski, D. A. & Frankel, R. B. (2000). Biologically controlled mineralization of magnetic 
iron minerals by magnetotactic bacteria. In Environmental Microbe-Metal Inter- 
actions, pp. 109-143. Edited by D. R. Lovley. Washington, DC: American Society for 
Microbiology. 

Beja, O., Aravind, L, Koonin, E. V. & 9 other authors (2000). Bacterial rhodopsin: 
evidence for a new type of phototrophy in the sea. Science 289, 1 902-1 906. 

Benner, S. A., Caraco, M. D., Thomson, J. M. & Gaucher, E. A. (2002). Paleontological, 
geological, and molecular histories of life. Science 296, 864-868. 

Blakemore, R. (1975). Magnetotactic bacteria. Science 190, 377-379. 

Blankenship, R. E. (2002). Molecular Mechanisms of Photosynthesis. Oxford: Blackwell 
Science. 



SGM symposium 65 



Metabolic diversity and exobiology 169 



Boetius, A., Ravenschlag, K., Schubert, C. J. & 7 other authors (2000). A marine 

microbial consortium apparently mediating anaerobic oxidation of methane. Nature 

407, 623-626. 
Canfield, D. E., Habicht, K. S. & Thamdrup, B. (2000). The archean sulfur cycle and 

the early history of atmospheric oxygen. Science 288, 658-661 . 
Coates, J. D., Michaelidou, U. # Bruce, R. A., O'Connor, S. M., Crespi, J. N. & 

Achenbach, L. A. (1999). Ubiquity and diversity of dissimilatory (per)chlorate- 

reducing bacteria. Appl Environ Microbiol 65, 5234-5241 . 
de la Torre, J. R., Christianson, L. M., Beja, O., Suzuki, M. T., Karl, D. M., Heidelberg, J. 

& DeLong, E. F. (2003). Proteorhodopsin genes are distributed among divergent 

marine bacterial taxa. Proc Natl Acad Sci U 5 A 100, 12830-12835. 
Doolittle, W. F. (2002). Thinking laterally about genes (review of Lateral DNA Transfer: 

Mechanisms and Consequences, by F. Bushman). Nature 41 8, 589-590. 
Doolittle, R. F., Feng, D.-F., Tsang, S., Cho, G. & Little, E. (1996). Determining divergence 

times of the major kingdoms of living organisms with a protein clock. Science 271, 

470-477. 
Dove, P. M., De Yoreo, J. J. & Weiner, S. (editors) (2003). Biomineralization. Reviews 

in Mineralogy and Geochemistry, vol. 54. Washington, DC: Mineralogical Society 

of America. 
Ehrenreich, A. & Widdel, F. (1994). Anaerobic oxidation of ferrous iron by purple bacteria, 

a new type of phototrophic metabolism. Appl Environ Microbiol 60, 451 7-4526. 
Fenchel, T. & Finlay, B. J. (1995). Ecology and Evolution in Anoxic Worlds. Oxford: Oxford 

University Press. 
Fortin, D. (2004). What biogenic minerals tell us. Science 303, 1618-1619. 
Frankel, R. B., Bazylinski, D.A.Johnson, M.S. & Taylor, B. L (1997). Magneto-aerotaxis 

in marine coccoid bacteria. Biophys J 73, 994-1000. 
Jetten, M. S. M., Stous, M., van de Pas-Schoonen, K. T. & 7 other authors (1998). The 

anaerobic oxidation of ammonium. FEMS Microbiol Rev 22, 421-437. 
Li, C. W., Chen, J. Y. & Hua, T. E. (1 998). Precambrian sponges with cellular structures. 

Science 279, 879-882. 
Lovley, D. R. & Phillips, E.J. P. (1988). Novel mode of microbial energy metabolism: organic 

carbon oxidation coupled to dissimilatory reduction of iron or manganese. Appl 

Environ Microbiol 54, 1 472-1 480. 
Lowenstam, H. A. (1981). Minerals formed by organisms. Science 211, 1 126-1 131 . 
Lowenstam, H. A. & Weiner, S. (1989). On Biomineralization. New York: Oxford University 

Press. 
Margulis, L. (1981). Symbiosis in Cell Evolution. New York: W. H. Freeman. 
Martin, W. & Muller, M. (1998). The hydrogen hypothesis for the first eukaryote. Nature 

392,37-41. 
Matsunaga, T. & Okamura, Y (2003). Genes and proteins involved in bacterial magnetic 

particle formation. Trends Microbiol 1 1 , 536-541 . 
Mojzsis, S., Arrhenius, G., McKeegan, K. D., Harrison, T. M., Nutman, A. P. & Friend, 

C. R. L. (1996). Evidence for life on Earth before 3,800 million years ago. Nature 

384, 55-59. 
Muller, M. (1993). The hydrogenosome. J Gen Microbiol 1 39, 2879-2889. 
Myers, C. R. & Nealson, K. H. (1988). Bacterial manganese reduction and growth with 

manganese oxide as the sole electron acceptor. Science 240, 1319-1 321 . 
Nealson, K. H. (1 997). Sediment bacteria: who's there, what are they doing, and what's 

new? Annu Rev Earth Planet Sci 25, 403-434. 



SGM symposium 65 



170 K. H. Nealson and R. Popa 



Nealson, K. & Berelson, W. (2003). Layered microbial communities and the search for 

life in the universe. Geomicrobiol J 20, 451^462. 
Nealson, K. H. & Conrad, P. G. (1999). Life: past, present and future. Philos Trans R Soc 

Lond B Biol Sci 354, 1923-1939. 
Nealson, K. H. & Rye, R. (2004). Evolution of metabolism. In Treatise on Geochemistry, 

pp. 41-61. Edited by W. H. Schlesinger. Amsterdam: Elsevier. 
Oremland, R. S., Switzer Blum, J., Culbertson, C. W., Visscher, P. T., Miller, L. G., 

Dowdle, P. & Strohmaier, F. E. (1994). Isolation, growth, and metabolism of an 

obligately anaerobic, selenate-respiring bacterium, strain SES-3. Appl Environ 

Microbiol 60, 3011-3019. 
Orphan, V. J., Hinrichs, K.-U., Ussier, W., Ill, Paull, C. K., Taylor, L T., Sylva, S. P., Hayes, 

J. M. & DeLong, E. F. (2001a). Comparative analysis of methane-oxidizing archaea 

and sulfate-reducing bacteria in anoxic marine sediments. Appl Environ Microbiol 

67, 1922-1934. 
Orphan, V. J., House, C H., Hinrichs, K.-U., McKeegan, K. H. & DeLong, E. F. (2001b). 

Methane-consuming archaea revealed by directly coupled isotopic and phylogenetic 

analysis. Science 293, 484^487. 
Pace, N., Stahl, D. A., Lane, D. J. & Olsen, G. J. (1986). The analysis of natural microbial 

populations by ribosomal RNA sequences. In Advances in Microbial Ecology, 

pp. 1-55. Edited by K. C. Marshall. New York: Plenum. 
Roden, E. E. & Zachara, J. M. (1996). Microbial reduction of crystalline iron(lll) oxides: 

influence of oxide surface area and potential for cell growth. Environ Sci Technol 

30, 1618-1628. 
Schink, B. & Friedrich, M. (2000). Phosphite oxidation by sulphate reduction. Nature 

406, 37. 
Schuler, D. (1999). Formation of magnetosomes in magnetotactic bacteria. J Mol Microbiol 

Biotechnol 1,79-86. 
Schuler, D. (2004). Molecular analysis of a subcellular compartment: the magnetosome 

membrane in Magnetospirillum gryphiswaldense. Arch Microbiol 181, 1-7. 
Shen, Y. A., Buick, R. & Canfield, D. E. (2001). Isotopic evidence for microbial sulphate 

reduction in the early Archaean era. Nature 410, 77-81 . 
Skinner, H. C. W. & Jahren, A. H. (2004). Biomineralization. In Biogeochemistry, 

pp. 1 17-184. Edited by W. H. Schlesinger. Amsterdam: Elsevier. 
Smock, A. M., Bottcher, M. W. & Cypionka, H. (1998). Fractionation of sulfur isotopes 

during thiosulfate reduction by Desulfovibrio desulfuricans. Arch Microbiol 169, 

460-463. 
Stoeckenius, W. & Bogomolni, R. A. (1982). Bacteriorhodopsin and related pigments 

of halobacteria./\nnu/?e\/£/oc/?e/7? 51, 587-616. 
Straub, K. L., Benz, M., Schink, B. &Widdel, F. (1996). Anaerobic, nitrate-dependent, 

microbial oxidation of ferrous iron. Appl Environ Microbiol 62, 1458-1460. 
Suzuki, Y., Kelly, S. D., Kemner, K. M. & Banfield, J. F. (2002). Nanometre-size products 

of uranium bioreduction. Nature ^9, 134-136. 
Theissen, U., Hoffmeister, M., Grieshaber, M. & Martin, W. (2003). Single eubacterial 

origin of eukaryotic sulfide : quinone oxidoreductase, a mitochondrial enzyme 

conserved from the early evolution of eukaryotes during anoxic and sulfidic times. 

Mol Biol Evol 20, 1 564-1 574. 
Tielens, A. G. M., Rotte, C, van Hellemond, J. J. & Martin, W. (2002). Mitochondria as 

we don't know them. Trends Biochem Sci 27, 564-572. 
Venter, J. C, Remington, K., Heidelberg, J. F. & 20 other authors (2004). Environmental 



SGM symposium 65 



Metabolic diversity and exobiology 171 



genome shotgun sequencing of the Sargasso Sea. Science 304, 66-74. 
Whitman, W. B., Coleman, D. C. & Wiebe, W. J. (1998). Prokaryotes: the unseen majority. 

Proc Natl Acad Sci U 5 A 95, 6578-6583. 
Whittaker, R. H. & Margulis, L. (1978). Protist classification and the kingdoms of 

organisms. Biosystems 10, 3-18. 
Widdel, R, Schnell, S., Heising, S., Ehrenreich, A., Assmus, B. & Schink, B. (1993). 

Ferrous iron oxidation by anoxygenic phototrophic bacteria. Nature 362, 834-836. 
Woese, C. R. (2004). A new biology for a new century. Microbiol Mol Biol Rev 68, 1 73-1 86. 
Woese, C. R., Stackebrandt, E., Weisburg, W. G. & 8 other authors (1984). The 

phylogeny of purple bacteria: the alpha subdivision. SystAppI Microbiol 5, 31 5-326. 
Woese, C. R., Kandler, O. &Wheelis, M. L. (1990). Towards a natural system of organisms: 

proposal for the domains Archaea, Bacteria, and Eucarya. Proc Natl Acad Sci USA 

87, 4576-4579. 
Xiong, J., Fischer, W. M., Inoue, K., Nakahara, M. & Bauer, C. E. (2000). Molecular 

evidence for the early evolution of photosynthesis. Science 289, 1 724-1 730. 
Young, J. R. & Henriksen, K. (2003). Biomineralization within vesicles: the calcite of 

coccoliths. In Biomineralization, Reviews in Mineralogy and Geochemistry, vol. 54, 

pp. 190-215. Edited by P. M. Dove, J. J. De Yoreo & S. Weiner. Washington, DC: 

Mineralogical Society of America. 
Zumft, W. G. (1997). Cell biology and molecular basis of denitrification. Microbiol Mol Biol 

Rev 61, 533-616. 



SGM symposium 65 



Biogeochemical cycling in polar, 
temperate and tropical coastal 
zones: similarities and differences 



David B. Nedwell 



Department of Biological Sciences, University of Essex, Colchester C04 3SQ, UK 



INTRODUCTION 

This chapter will consider biogeochemical cycling in the coastal zone. This is defined as 
that area of estuarine and coastal, relatively shallow water where there is strong 
benthic-pelagic linkage and exchange between the water column and the underlying 
sediment. In deeper water this connection becomes increasingly tenuous as the 
exchange between the euphotic zone and the benthic layer declines. Longhurst et al. 
(1995) recognized the coastal boundary domain, divided into 22 provinces, as often 
bounded by a shelf-break front, and included coastal upwelling regions. The coastal 
zone generally exhibits high rates of primary production compared with the open 
ocean (Table 1), and there is the greatest impact from inputs from the land to the coastal 
sea through estuaries. Estuaries and coastal seas are highly heterotrophic systems which 
are net exporters of C0 2 to the atmosphere due to the mineralization and recycling 
of both autochthonous and allochthonous organic matter (Borges, 2005). 

PHYSICO-CHEMICAL DIFFERENCES BETWEEN LATITUDINAL 
REGIONS 

The physical-biological interactions that influence marine phytoplankton production 
have been reviewed by Daly &C Smith (1993). Because of the spherical shape of the 
Earth, more solar energy falls per unit area of surface in equatorial regions than at 
the poles (Fig. la), and the incidence of light at the equator is vertical to the surface, but 
oblique at the poles. Furthermore, the distance radiation travels through the atmos- 
phere is longer at the poles, thus reducing the irradiation incident at the poles compared 
with equatorial regions. Consequently, equatorial regions have higher energy inputs 

SGM symposium 65: Micro-organisms and Earth systems - advances in geomicrobiology. 
Editors G. M. Gadd, K. T. Semple & H. M. Lappin-Scott. Cambridge University Press. ISBN 521 86222 1 ©SGM 2005 



174 D. B. Nedwell 



Table 1. Areas and annual primary production rates of biogeochemical provinces 
based on the CZCS 1978-86 climatological data (after Longhurst et a/., 1995) 



Geographical subset 



Area (xKT 6 km 2 ) 



,-v 



Primary production rate (gCm 2 year n ) 



Coastal domain 


37-4 


Polar domain 


20-8 


Westerlies domain 


129-9 


Trades domain 


139-9 


Arctic Ocean 


17 


Atlantic Ocean 


74-0 


Pacific Ocean 


148-9 


Indian Ocean 


45-4 


Southern Ocean 


57-9 


Upwelling provinces 


8-4 



385 
310 
126 
93 
645 
199 
132 
143 
141 
398 



Table 2. Seasonal variation in characteristics of the coastal zone in polar, 
temperate and tropical coastal zones 



Characteristic 



Polar 



Temperate 



Tropical 



Insolation 
Day length 
Water temperature 
Estuarine flushing 



Extremely seasonal 

Extremely seasonal 

Constant, low 

Highly seasonal 
(summer melt) 



Moderately seasonal 
Moderately seasonal 
Seasonally variable 
Moderately seasonal 



Weakly seasonal 

Weakly seasonal 

Constant, high 

Highly seasonal 
(wet/dry season rainfall) 



and are hotter than polar regions. Rainfall and humidity are also extreme in the tropics, 
often with distinct rainy seasons which impose differences in the ecology of the coastal 
zones. During the tropical dry season, when there is little or no rainfall or river flow, 
there may be accumulation of detritus within an estuary but during the wet season 
heavy rainfall flushes the estuary, exporting a spike of material to the coastal zone, and 
're-setting' the estuary (Eyre & Balls, 1999). For example, north Queensland tropical 
estuaries exhibited flushing times varying between 2 and 30 days during the dry season 
but days in the wet, when river water essentially passed straight through the estuaries. 
In contrast, temperate Scottish estuaries exhibited much less seasonal variation in 
flushing. Pulses of nutrients (and particulate matter) leached from the catchment are 
therefore likely in tropical estuaries compared with temperate estuaries, whose rainfall 
and therefore input loads to the coastal zone are less seasonally variable. Estuaries 
are extremely important in attenuating the loads of nutrients passing from land to sea 
(see Nedwell et al., 1999); in the case of nitrogen particularly by denitrification, or of 
phosphorus by settlement of particles to which phosphate is adsorbed. Attenuation 
of nutrient fluxes from land to sea is likely to be diminished if the major loads coincide 
with periods when freshwater flushing times within the estuary are very short, as in 



SGM symposium 65 



Biogeochemical cycling in the coastal zone 175 



(a) 



N 



(b) 




20 r 




A I 1 1 I I I J I I L. 1 i J 

JFMAMJ JASOND 



Fig. 1. (a) Owing to the spherical nature of the Earth, solar energy is concentrated over a smaller area 
in equatorial regions than in polar regions, and the distance travelled through the atmosphere is also 
shorter at the equator. Consequently, energy input per unit area is greater at the equator than at the 
poles, (b) Variations in day length throughout the year in relation to latitude (reproduced with 
permission from Osborne, 2000). 

some tropical estuaries. In polar regions any terrestrial input to the coastal zone occurs 
only during the short summer, when ice and snow melt occurs, and is therefore also 
seasonally extreme. 

Day length also exhibits latitudinal changes. Because of the rotation of the Earth 
around its axis there is little difference seasonally in the day length at the equator, 
whereas at the poles there is extreme change in the day length between winter and 
summer and intermediate seasonal changes in day length at temperate latitudes 
(Fig. lb). The different environmental factors are summarized in Table 2. In essence, 
high-latitude polar coastal waters are stenothermal, with relatively constant low 
temperature (-1 to 4 °C), mid-latitude temperate waters are eurythermal, with seasonal 
temperature cycling between approximately 2 and 20 °C, and low-latitude tropical 
waters are stenothermal, with relatively constant high temperature (> 16 °C). 

PRIMARY PRODUCTION 

Biogeochemical cycling in the coastal zone is driven by autochthonous primary 
production and also by allochthonous inputs via estuaries, both of which provide the 
energy for subsequent chemo-organotrophic activity. To compare biogeochemical 
cycling in different latitudinal regions we must start by examining their primary 



SGM symposium 65 



176 D. B. Nedwell 




75 
88 
103 
120 
141 
165 
193 
226 
265 
| 310 
363 
426 
498 
583 
683 
800 



Fig. 2. Global primary production (g C rrr 2 year 1 ) modelled from satellite estimates of chlorophyll a, 
photosynthesis-light relationships and local light environment (reproduced with permission from 
Longhursteta/., 1995). 



production. Are there consistent differences in primary production in polar, temperate 
or tropical coastal zones? Longhurst et al. (1995) used satellite radiometer data to 
model primary production in the oceans (Fig. 2), based on estimates of chlorophyll by 
the CZCS radiometer on the NIMBUS satellite from 1979 to 1986, photosynthesis-light 
relationships and the light environment. It is clear that the coastal zones are in general 
areas of high production. They estimated global oceanic primary production as 44-7- 
50-2 Pg C year . The annual primary production rates for the coastal domain (Table 2) 
averaged 385 g C m - ^year -1 , for the Arctic Ocean 645 g C m - year -1 and upwelling 
provinces 398 g C m -2 year -1 . The annual production in the coastal domain was greater 
than that in all oceans except the Arctic Ocean, and the relative productivity of the 
ocean province was only 2-5-4-0 times that of the coastal provinces, despite the former's 
very much greater area. Tropical coastal provinces in the Indian and Pacific Oceans 
averaged 279 g Cm -2 year -1 , again confirming that low-temperature polar coastal 
communities are as productive as those at lower latitudes and higher temperatures. 



What is beyond doubt is that the seasonal pattern of primary production in the 
different regions varies enormously, reflecting the different seasonality imposed by 
insolation and day length, and in turn influencing the biogeochemical cycling that goes 
on. Polar regions have constantly low water temperatures (usually <4 °C) but exhibit an 
extreme seasonal pattern, with short summer ice-free periods of high insolation during 
which pelagic primary production rapidly increases, plankton biomass quickly attains 
a maximum and a pulse of organic input to the coastal zone bottom sediments 
occurs (Fig. 3) (Nedwell et al, 1993; Rysgaard et al, 1999). Rysgaard et al (1999) 
showed that there was a strong relationship in Arctic waters between the annual pelagic 



SGM symposium 65 



0) 

£ 

'c 

™ .— . 

en — 

O -D 

<D IN 

+-< I 
£ £ 

C D) 

.2 £ 

+i — ' 
i/i 

o 

Q. 
CD 
Q 

U) 

O 



Biogeochemical cycling in the coastal zone 177 



Sea ice 



4 i- 



3 - 



2 - 




AMJJASONDJFMAMJJASONDJ FMAMJJASONDJ F 
1989 1990 1991 1992 

Fig. 3. Seasonal changes in the settlement rate of organic matter from the water column to coastal 

sediments at Signy Is., South Orkney Islands, Antarctica (reproduced with permission from Nedwell 
etal., 1993). 



500 r 



03 

£ 
u 

C 

'n 

03 

L. 

C 

g 

'■e 

T3 
O 



400 - 



300 - 



200 - 



100 - 





May 




Aug Sep Oct 
1996-1997 



Feb Mar 



Fig. 4. Annual primary production (•) and grazing by ciliates (O), heterotrophic dinoflagellates (A) 
and copepods (O) in Young Sound, northeast Greenland (reproduced with permission from Rysgaard 
etal., 1999). 

primary production and the duration of the open-water period when pelagic primary 
production could occur. This is because an ice layer with snow cover strongly attenuates 
light, and it is only after ice melt that there is significant illumination of the water. The 



SGM symposium 65 



178 D. B. Nedwell 



Sea ice 



5 r- 



4 - 



£ 



5 3 

03 

E 

u 

"E 2 

03 

CJ1 
i_ 

O 



1 - 








MAM 



J J 
1991 



A 



O N D J 



F M 
1992 



A 



Fig. 5. Benthic organic matter content at Signy Is. Sediment depth: O, 0-0-5 cm, •, 1-2 cm 
(reproduced with permission from Nedwell etal., 1993). 



£ 



140 r 



.2 120 

u 

13 

TS 

CD 



100 - 



o 



80 - 



CD 
03 

-Q 



I 60 h 

03 
N 



40 - 



03 

k. 
CD 

c 
E 

03 







Sea ice 







F M A M J J 
1991 



A S O N D J 



1992 



Fig. 6. Benthic anaerobic organic matter degradation driven by sulfate reduction as a percentage 
of total mineralization indicated by benthic 2 uptake (reproduced with permission from Nedwell 
etal. ,1993). 

seasonal cycle of pelagic primary production in Young Sound, Greenland, exhibited an 
extremely short but intense period of primary production (Fig. 4) and in the coastal 
zone around Signy Island, in the South Orkney islands, at the tip of the Antarctic 
Peninsula, a similar short but intense period of summer production has been described 



SGM symposium 65 



Biogeochemical cycling in the coastal zone 179 

(Nedwell et aL, 1993). Once sea ice breaks up, the rapidly increasing light penetrates the 
clear water and primary production increases exponentially, the water column seeded 
by ice algae from the bottom surface of the melting ice. The complete seasonal cycle 
of algal production at Signy Is. occurs within 2-3 months, after which the sea ice forms 
again, and may be as short as 4-8 weeks at higher latitudes (e.g. Rysgaard et al., 1998; 
Karl et al., 1996). Lack of turbulence beneath fast ice causes rapid settlement of organic 
matter from the water column and there is, in essence, a short but intense pulse of 
organic deposition into the coastal bottom sediments. Benthic processes such as 0-, 
uptake respond quickly and increase due to the input of fresh organic matter, so that 
there is a strong seasonal signal of benthic activity, in direct contrast to tropical coastal 
systems. In the few studies carried out, annual benthic organic degradation seems 
broadly to balance net primary production in both Arctic and Antarctic coastal systems 
studied so far (Nedwell et al., 1993; Rysgaard et al., 1998), indicating stable ecosystems. 
However, while primary production occurs in a short summer burst, that pulsed input 
of organic matter sustains benthic processes for the remainder of the year, spread 
out over the period when winter ice cover minimizes further organic inputs. Thus, the 
deposited organic matter represents a food or energy reserve for the organotrophic 
benthic community which tides them over the winter period with no new inputs. For 
example, at Signy Is. the seasonal signal of deposited organic matter could be seen only 
in the top 0-5 cm of sediment (Fig. 5). As the deposited organic matter was buried into 
the sediment by very active bioturbation, the proportion of benthic organic matter 
degradation which was driven by anaerobic sulfate reduction increased through 
the winter period, while that due to aerobic metabolism declined (Fig. 6) until, with the 
input of fresh organic matter in spring onto the surface of the sediment, aerobic 
metabolism again increased dramatically and the cycle repeated. Mincks et al. (2005) 
have argued that this pulsed input to polar benthic communities represents a 'food 
bank' for polar benthic detritivors, including micro-organisms, over the winter period 
when there were no new organic inputs, and the data from Signy Is. are consistent with 
this idea. 

In the temperate coastal zone annual primary production still exhibits a strong season- 
al signal, though less intense than that of polar regions. There are significant seasonal 
changes in insolation and water temperature (e.g. approximately 3-18 °C in the 
southern North Sea) and a seasonal cycle of spring bloom primary production occurs, 
subsequently limited by depleted nutrient availability, and possibly with a smaller 
autumn bloom if summer stratification breaks down. In estuaries and inshore waters, 
where there is a continuous supply of nutrients, there may be no distinct spring bloom, 
but an increase of primary production during spring which is sustained throughout the 
summer (e.g. Kocum et al., 2002a, b). For example, in the north-central North Sea 
during 1998-99 there was a spring bloom which decreased during summer, but, in 

SGM symposium 65 



180 D. B. Nedwell 



the German Bight, where there is a continued supply of nutrients from the Rhine, 
primary production continued in a broad peak throughout the growing season (Fig. 7). 
Again, there are seasonal changes in the downward fluxes of organic matter to bottom 
sediments, albeit less extreme than in polar regions. Upton et al. (1993) estimated 
that benthic organic matter mineralization accounted for 17-45 % of net primary 
production in the various regions of the southern North Sea. 

In the tropics the seasonal insolation change is least and insolation remains at high 
values throughout the year, while water temperatures also remain relatively constant 
and high. Primary production is therefore less likely to be limited by insolation than at 
higher latitudes, but regulated by other factors such as nutrients or turbidity. The 
tropical regime typically reflects the relatively continuous formation of new organic 
matter, with little seasonal variation. In a study of coastal lagoon sediments in Fiji there 
were no statistically significant seasonal variations in either benthic oxygen uptake rates 
or rates of pelagic/benthic nutrient exchange (Sobey, 2004), reflecting a lack of seasonal 
change in inputs of organic matter. However, our knowledge of coastal zone biogeo- 
chemical processes in the tropics is based on a relatively sparse dataset compared 
with other regions, and requires much further study. Much of the tropical ocean has 
low nutrient concentrations because of the presence of a permanent thermocline 
which prevents vertical transport of new nutrients from deeper water. This conspires 
to ensure that the tropical coastal zone generally operates at lower ambient nutrient 
concentrations than at higher latitudes. While tropical coastal zone nutrient concen- 
trations may be higher than those offshore in tropical oceans, they are still usually 
lower than those in the coastal zone at higher latitudes. Tropical primary production 
therefore becomes dependent upon recycling of nutrients to sustain fresh production, 
and is likely to be nutrient limited. However, where nutrients become available from 
upwelling or from fluvial inputs in a tropical high-light environment, primary pro- 
duction may be intense (e.g. Robertson et al., 1998). 

What factors regulate primary production in the different coastal environments? 
Clearly, in polar regions it is light that is the principal seasonal regulator of primary 
production in coastal regions, at constant very low temperature, but this is not to imply 
that light is the only regulator. Primary production may be co regulated by a variety of 
interacting factors, and nutrient availability is often a significant factor despite 
the apparently relatively high nutrient concentrations. In the Southern Ocean the 
occurrence of high-nutrient, low-chlorophyll areas, where primary production is low 
despite the presence of relatively high concentrations of nitrate, has been attributed to 
lack of available iron, thus limiting photosynthesis (de Baar et al., 1995; Martin et al., 
1990). This iron limitation apparently does not happen in the Arctic, where proximity 
to land masses results in higher aeolian inputs of iron and nitrate is reduced to very low 

SGM symposium 65 



Biogeochemical cycling in the coastal zone 181 



50 r 




Fig. 7. Seasonal variation in pelagic primary production in different ICES regions of the North Sea 
(reproduced with permission from Joint & Pomroy, 1 993). 



concentrations during summer activity. Moreover, in coastal zones, in close proximity 
to land, iron supply should be higher than in the open ocean. However, low tempera- 
ture may have other influences on the ability of algae to sequester nutrients at low 
concentrations. 

My co-workers and I have shown previously (Nedwell & Rutter, 1994; Ne dwell, 1999; 
Reay et aL, 1999) that low temperature apparently reduces the affinity of bacteria and 
algae for substrates taken up by active transport, including nitrate, phosphate and 
silicate. This has been attributed to stiffening of the cell membrane at low tempera- 
tures, which reduces the efficiency of transporter proteins embedded in the membrane. 
From the data available, there is no such inhibitory effect of low temperature upon 
ammonium uptake, which occurs at least partly by passive or facilitated diffusion 
mechanisms which do not depend upon active transport. Primary production in the 
Southern Ocean seems to be very dependent upon ammonium uptake, with very low 
f-ratios (see Reay et al., 2001), and it is possible that for at least part of the summer 
growth season passive uptake of ammonium is sufficient to maintain the rates of 
primary production measured. The concentrations of dissolved inorganic nitrogen, 
predominantly nitrate, in Antarctic waters often remains comparatively high (> 10 [iM) 
even during summer, but addition of nitrate still stimulates algal growth despite 
significant concentrations of nitrate already being present, indicating continued limita- 
tion by nitrogen even at these apparently replete concentrations. 



SGM symposium 65 



182 D. B. Nedwell 

Table 3. Rates of benthic oxygen uptake reported in polar, temperate and tropical 
regions 



Site 



Rate(mmol 2 m 2 day n ) 



Source 



Polar 

E. Svalbard, 226-320 m depth 

Bering and Chukchi Seas, 1 9-25 m depth 

E. Svalbard, 1 70-240 m depth 

E. Greenland, 40 m depth 

Bering and Chukchi Seas, 30-52 m depth 

Off Newfoundland, 270 m depth 

Signy Is., Antarctica 

Temperate 

Flax Pond, Long Is. Sound 

Aarhus Bay 

North Sea sediments, 25-81 m depth 

NW Mediterranean, 60-80 m depth 

Skagerrak, 190-695 m depth 

Thames estuary sediments, UK 

Colne estuary, UK 

Tropical 

Thai mangrove 

Jamaican mangrove 

Fiji lagoon sediment 

Indus delta, Pakistan 

Hinchinbrook Is., Australia 



3-2-11-9 


Pfannkuche&Thiel(1987) 


8-7-19-2 


Henriksenefa/. (1993) 


3-9-11-2 


Hulth etal. (1994) 


17-8 


Rysgaard etal. (1996) 


7-3-25-5 


Henriksenefa/. (1993) 


8-4 


Pomeroy etal. (1991) 


15-85 


Nedwell etal. (1993) 


81-137 


Mackin&Swider(1989) 


27 


Jorgensen & Revsbech (1 985) 


5-3-27-8 


Upton et al. (1993) 


4-6-9-9 


Tahey etal. (1994) 


11-8-16-1 


Canfieldefa/. (1993) 


12-2-241 


Trimmer ef al. (2000) 


48-216 


Dong ef al. (2000) 


17-61 


Kristensenefa/. (1991) 


31-103 


Nedwell et al. (1994) 


0-24 


Sobey (2004) 


16-28 


Kristensenefa/. (1992) 


2-8-61 


Alongiefa/. (1999) 



If affinity for nitrate (or other nutrients) is decreased at low temperatures, algae may be 
nitrogen-limited despite the presence of significant concentrations of nitrate because 
the nitrate cannot be sequestered effectively. This suggestion does not necessarily 
contradict the hypothesis that iron limits Southern Ocean but not Arctic Ocean 
primary production, as iron uptake is itself probably an active transport process and its 
uptake will therefore tend to be retarded at low temperature. 

ORGANIC MATTER DEGRADATION AND BIOGEOCHEMICAL 
RECYCLING 

The strong pelagic-benthic linkage in the coastal zone ensures that the bottom 
sediments are major sites of organic matter breakdown and biogeochemical recycling. 
In comparing biogeochemical cycling in the different regions we may ask a number of 
questions about the benthic communities which bring about this recycling and the 



SGM symposium 65 



Biogeochemical cycling in the coastal zone 183 

environmental factors which control them. It is clear that primary production rates do 
not show any consistent latitudinal variation with temperature (see above), being as 
great in some polar regions as in the tropics. Similarly, is there any consistent relation- 
ship between water temperature and the biogeochemical recycling of organic matter? 
Inspection of reported rates of benthic oxygen uptake (Table 3), as a surrogate for 
benthic organic matter breakdown, again suggests that there is no consistent difference 
between high and low latitudes: the rates of benthic oxygen uptake in Factory Cove, 
Signy Is., at water temperatures near °C all year were greater than those reported for 
many tropical sediments (Nedwell et al., 1993; Glud et al., 1998). However, the overlap 
of gross rates of 2 uptake in the different regions does not imply that the microbial 
communities at high or low latitudes are operating in physiologically similar ways. 

Pomeroy and co-workers (Pomeroy & Deibel, 1986; Pomeroy et al., 1991) proposed the 
'cold water paradigm' that 'bacterial metabolism and growth are depressed to a much 
greater degree than those of phytoplankton at low (<4 °C) sea water temperatures', 
which would tend to lead to imbalance between the production and mineralization of 
organic matter. They argued that less consumption of primary production by bacteria 
at low temperature reduced the microbial loop activity and left more organic matter for 
metazoan grazers. In an investigation of bacterial growth and primary production 
along a north-south transect in the Atlantic Hoppe et al. (2002) demonstrated clear 
gradients of primary production and bacterial growth correlated with latitudinal 
temperature change. The ratio of bacterial production to primary production both 
above and below the equator was most significantly correlated with water temperature. 

In a study of growth rates of bacteria in cold oceans measured by thymidine incorp- 
oration, Rivkin et al. (1996) concluded that there was a weak relationship between 
specific growth rate and temperature, and the mean and median specific growth rates 
for bacteria from cold (<4°C) and warm (>4°C) regions were not significantly 
different, i.e. the growth rates of bacteria from cold and temperate oceans are similar at 
their respective environmental temperatures. However, they pointed out that the 
production rate is a function of both growth rate and biomass of cells in the standing 
crop, and the bacterial abundance in cold ocean regions can be 10-fold lower than in 
temperate or tropical seas. 

EFFECT OF TEMPERATURE ON ABILITY TO SEQUESTER 
SUBSTRATES BY ACTIVE TRANSPORT 

Interaction between organic substrate requirement and temperature has been reported 
for marine bacteria (Pomeroy et al., 1991; Pomeroy & Deibel, 1986; Wiebe et al., 1992, 
1993); at low temperature a higher concentration of substrate is required to support 
a given rate of growth or respiration than at a higher environmental temperature. 

SGM symposium 65 



184 D. B. Nedwell 



This temperature-substrate concentration interaction was attributed (Nedwell 6c 
Rutter, 1994; Nedwell, 1999) to decreased affinity for substrates at temperatures below 
the optimum temperature for growth (T t ). At temperatures below T micro- 
organisms become increasingly less able to sequester low concentrations of substrates 
from their environment effectively by active uptake, apparently because as the mem- 
brane stiffens as the temperature drops the transport proteins embedded in the 
membrane become less effective, and affinity for substrates declines (Nedwell, 1999). 
Because of the decreased affinity for substrates at low temperature, higher concen- 
trations of substrates are necessary in the environment to counter loss of affinity and to 
maintain a given substrate flux into the cell. This loss of affinity below T has been 
demonstrated for active uptake of both inorganic solutes such as nitrate by both algae 
and bacteria (Ogilvie et al., 1997; Reay et al., 1999) and organic substrates by chemo- 
organotrophic bacteria (Nedwell & Rutter, 1994). The paradigm also holds true 
for psychrophilic, mesophilic and thermophilic bacteria over their respective ranges of 
temperature (Nedwell, 1999). 

For a species to be selected by their environmental temperature they must first be 
physiologically capable of growing at that temperature, and the membrane of each 
species adapts to its broad temperature range by features such as the fatty acid com- 
position of the membrane lipids, which affects membrane fluidity (e.g. Russell, 1990, 
1992, 1998). The homeoviscous model of Sinensky (1974) proposed that these mem- 
brane compositional differences adapt a species to maintain membrane fluidity, and 
hence biological function, over a particular range of temperature. A certain degree of 
flexibility of this temperature range appears to be possible by variations in the 
membrane and key enzymes but in essence the biokinetic range for a particular species 
is fixed. Within the broad temperature range for a species, however, their ability 
to sequester substrates from the environment appears to decline below T . Note, 
however, that we would only expect an adverse effect by a low environmental tempera- 
ture on affinity for substrates if the environmental temperature was below the T for 
the species, so the extent to which a species population is optimally adapted to 
its environmental temperature would appear important. This raises the question of 
the extent to which species populations in different temperature environments have T t 
at or below their environmental temperature. 

In high-temperature environments such as thermal springs the optimum temperature of 
the species present conforms closely to the environmental temperature, but in low- 
temperature, polar environments it appears that this is not the case. Morita (1975) 
defined psychrophiles, mesophiles and thermophiles in relation to their cardinal 
temperatures, with obligate psychrophiles able to grow at 0°C with T <15°C. A 
category of 'psychrotolerant' types, able to grow at °C but with an optimum > 15 °C, 

SGM symposium 65 



Biogeochemical cycling in the coastal zone 185 

was subsequently added. In reality there is probably a continuum of physiological 
adaptation across the temperature range. Feller & Gerday (2003) have argued that 
the terms stenothermal and eurythermal psychrophiles are better used, respectively, for 
'obligate' psychrophiles able to grow over only a narrow temperature range or 'facul- 
tative' psychrophiles able to grow over a wide temperature range. While stenothermal 
psychrophiles are undoubtedly present in polar environments, the majority of organ- 
isms present seem to be eurythermal psychrotolerants with T for growth much higher 
than the environmental temperature (e.g. Morita, 1975; Franzmann, 1996; Upton 
& Nedwell, 1989). Additionally, many reports of biogeochemical processes such as 
sulfate reduction (Arnosti & Jorgensen, 2003 ; Nedwell, 1989) also indicate optima well 
above environmental temperature in polar regions. Li and co-workers (Li, 1980, 1985; 
Li et al., 1984) suggested that algal species at the equator have temperature optima for 
photosynthesis near to their environmental temperature, but that at higher latitudes 
and lower temperatures the environmental temperature is lower than the optimum for 
photosynthesis. If this latitudinal change is the general case then it appears that in low- 
temperature populations both algal and bacterial species may indeed be operating 
under conditions where they are less well adapted physiologically to their environ- 
mental temperature than populations at lower, warmer latitudes. 

It might be predicted that it would be adaptive for an organism to operate most 
efficiently at its environmental temperature, so it is somewhat surprising that, while 
this seems to occur in the tropics, it does not seem to be true in high-latitude popula- 
tions. This may mean that T is irrelevant to selection. A new model of the effect of 
temperature on enzyme activity, the equilibrium model (Daniel et al., 2001; Peterson 
et al., 2004), proposes that enzyme activity at any temperature is a function of an 
equilibrium between an active and an inactive form of the enzyme; it is this equilibrium 
which leads to the lowering of enzyme activity above the T t for the enzyme. It is the 
inactive form of the enzyme that is irreversibly denatured at higher temperature, and 
the reduction in activity may be due to a reversible change of conformation at the active 
site. Similarly, inactivation of enzymes can occur at low temperature. The equilibrium 
between active and inactive forms is defined by an equilibrium constant, K , and T 
is the temperature when K is 1 (i.e. when active and inactive forms are equal). Peterson 
et al. (2004) argue that adaptive evolution is more likely to operate through selection 
for T than through T for the enzyme or thermal stability. In polar, low-temperature 
populations an increase in temperature may increase the catalytic rate by the Arrhenius 
activation energy, but furthermore may change the equilibrium between active and 
inactive forms to increase the amount of active enzyme, so that the overall rate of 
reaction is enhanced in a compound manner as temperature rises. In that case the 
apparently high T for growth of polar micro-organisms may be an artefact which 
does not actually reflect their degree of adaptation to their environmental temperature. 

SGM symposium 65 



186 D. B. Nedwell 



In a review of psychrophilic enzymes Feller & Gerday (2003) argue that effective 
enzyme activity at low temperature requires highly reactive reaction centres which 
are less stable than those of mesophilic enzymes, which will therefore be heat labile. 
Psychrophilic enzymes seem to be inactivated at temperatures much lower than those 
that cause unfolding of their protein structure, unlike mesophilic or thermophilic 
enzymes. This emphasizes the severe heat-lability of the psychrophilic reaction centre 
compared with those of mesophiles or thermophiles. Feller & Gerday (2003) point out 
that 'a mobile and flexible active site binds its substrate weakly and, indeed, most 
psychrophilic enzymes have higher K m values than their mesophilic counterparts'. 
Furthermore, cold-active enzymes maintain reaction rates by increasing k c . n (the 
maximum enzyme reaction rate at a given temperature) at the expense of K m , i.e. 
affinity as defined by K m is likely to decline (K m increase) at low temperatures. Feller & 
Gerday (2003) proposed that 'psychrophilic enzymes have reached a state that is close 
to the lowest possible stability... and they cannot be less stable without losing the 
native and active conformation'. Does this provide a rationale for why physiological 
adaptation by micro-organisms to low temperature is not as complete as that attained 
at higher environmental temperatures? The assumption is extending the model for 
single enzyme reactions to cell growth, but is it too fanciful to imagine that the effect of 
temperature upon growth may mimic its effect upon the enzyme reactions that support 
growth? If true, it may also explain why, at low environmental temperatures, micro- 
organisms are more regulated by low affinity for substrates than in warmer environ- 
ments, because they are operating suboptimally in terms of temperature and/or the 
reaction centres of key psychrophilic enzymes (which may include transport proteins) 
confer a low affinity for substrates. 

The implications of this interaction between affinity for substrates and environmental 
temperature are profound. Mincks et al. (2005) argue that microbial respiration and 
mineralization of organic matter is a product of both substrate concentration (S) and a 
first-order reaction rate constant (k) and that benthic community respiration (R) is 
therefore described by 

where S. and k x are the respective rate constants and concentrations for metabolizable 
components (i) of the organic matter (Fig. 8a). As a given k { decreases due to low 
temperature, the rate can be maintained only by an increase in S t . 

Alternatively, the rate can be described (Button, 1993) using the specific affinity, a A , and 
the rate of substrate utilization is 

Roc2tf A(i) S f X 

SGM symposium 65 



Biogeochemical cycling in the coastal zone 187 



c 
o 

■M 

03 

i- 

c 

(Li 
u 

c 
o 

u 

03 



on 
-Q 

D 



8 
6 
4 H 
2 ■ 



(a) 



1-6 
14 J 

as 1-2 ■ 

a) +j 0*6 
I § °' 4 

S 8 o-2 



00 







2 4 6 8 

/C (respiration rate constant) 




(b) 



r 0035 




■ 0030 




■ 0025 


I 


• 0020 


l 

-C 


■ 0015 


o 


■ 0-010 


n 


■ 0-005 


< 

03 


• 0000 





10 15 20 

Temperature (°C) 



25 



Fig. 8. (a) Conceptual comparison between effects of substrate concentration and respiration rate 
constant for benthic organic matter in polar and temperate sediments (after S. L. Mincks and others, 
personal communication), (b) Effect of temperature on specific affinity (a A ) for glucose for an Antarctic 
eurythermal coryneform (■) and the substrate concentration required at each temperature to 
maintain a constant rate of substrate utilization of 0-001 5 umol h~ 1 (•). 



where a A ^ and S 1 are the specific affinity and concentration of substrate i, and X is the 
microbial biomass. As a A decreases because of lower temperature, the rate can only be 
maintained by either higher substrate concentrations or a higher biomass. Fig. 8(b) 
shows how a A varies with temperature for a eurythermal coryneform with a T of 
22 °C, isolated from an Antarctic lake. In order for a constant rate of substrate 
utilization of 0-0015 fimol h _1 to be maintained as temperature decreases below T 
(assuming unit biomass, X), the decreasing a A must be countered by an increased 
substrate concentration, S. This leads to the conclusion that microbial communities 
might be regarded as operating either at high affinity (a A ) but low 5, as in tropical 
environments, or at low affinity but high S, as in polar environments (Fig. 8a). This is 
analogous to fast turnover of small substrate pools in tropical systems compared with 
slow turnover of relatively large substrate pools in polar benthic environments (the 
'food reserve' concept). Holmboe et al. (2001) compared anoxic decomposition in Thai 
mangrove sediment and temperate Wadden Sea salt-marsh sediments. They reported 
high C : N and C : P ratios in both sediments, but low nutrient release and faster rates 



SGM symposium 65 



188 D. B. Nedwell 



of turnover of N and P by nutrient-deficient bacteria in the tropical sediment, which is 
consistent with the proposed model. 

Presumably, temperate environments exhibit changes between the two extremes in 
response to the seasonal temperature cycle and seasonal selection of the populations 
in the microbial community Previous studies (King & Nedwell, 1984; Sieburth, 1967) 
have demonstrated the selection of temperate bacterial communities by seasonal 
temperature change, the physiological change induced in the community lagging 2 
months behind the environmental temperature change. Selection by competition is 
not immediate in response to seasonal temperature change, and the 2-month lag is the 
period during which the competition occurs, resulting in seasonal selection of better- 
adapted populations. This presumably results in cycling between selection of low- 
affinity/high-substrate-concentration populations in the spring towards higher-affinity/ 
low-substrate-concentration populations during summer. This continual seasonal 
change in itself implies that adaptation to environmental temperature by microbial 
populations in eurythermal temperate environments with seasonally varying tempera- 
ture must be less complete than in stenothermal environments, and it is inevitable that 
physiological adaptation with respect to environmental temperature will be less 
optimized than in stenothermal environments. We have shown previously (Rutter & 
Nedwell, 1994) that, in non-steady-state temperature environments, the speed of 
response by a species to change may be more adaptive than their competitive ability 
under constant conditions. In stenothermal, constant low or high temperature environ- 
ments the speed of response to temperature change will not be an issue. 

It would seem, therefore, that physiological adaptation in polar populations might be 
suboptimal relative to environmental temperature because of the inherent limitation of 
adaptability of enzymes to very low temperature; in temperate populations it must be 
suboptimal because of the time lag between selection of populations in response to 
seasonal environmental temperature change; but in stenothermal tropical systems 
populations will be physiologically better adapted, with T near to environmental 
temperature. 



FUNCTIONAL GROUPS 

While gross rates of biogeochemical processes apparently do not differ between 
latitudinal regions, the microbial communities which catalyse them may differ. The 
process of benthic organic matter breakdown is achieved by the integrated activities of 
a variety of different functional groups of bacteria and archaea, together with meio- 
and macrofauna, including aerobes, and anaerobes such as nitrate respirers, Fe and 
Mn 4+ respirers, sulfate reducers and methanogenic archaea. Firstly, are there any 

SGM symposium 65 



Biogeochemical cycling in the coastal zone 189 

differences in the relative importance of the different functional groups of micro- 
organisms present in the different latitudinal regions? Secondly, are there any phylo- 
genetic differences in these functional groups in the different regions: i.e. are there any 
specifically low-temperature or high-temperature types? 

The evidence available suggests that there are no obvious major differences in func- 
tional groups in polar or temperate sediments. Benthic oxygen uptake rates in Arctic 
and Antarctic coastal sediments are similar (e.g. Nedwell et al., 1993; Glud et al., 1998) 
to those of temperate sediments at temperatures 20-25 °C higher. In Young Sound, 
Greenland, about half of the organic input was preserved by burial in bottom sedi- 
ments (Rysgaard et al., 1998) while, of the organic matter mineralized to C0 2 , aerobic 
respiration accounted for 38 %, sulfate reduction for 33 %, iron reduction for 25 % and 
denitrification for only 4 % (manganese reduction was unimportant in these sedi- 
ments). Reduced iron (and manganese if present) will be reoxidized by 2 in the 
surface, oxic layer of sediment, facilitated by bioturbation of the surface layers of 
sediment, but failure to measure iron and manganese reduction directly will result in 
the significance of anaerobic mineralization by iron and manganese being under- 
estimated and appearing as aerobic mineralization. Similar proportions of aerobic 
respiration and sulfate reduction have been reported for Antarctic sediments, although 
iron and nitrate respiration were not measured directly (Nedwell et al., 1993). Rates of 
sulfate reduction were similar to those reported for temperate sediments, although the 
optimum temperature for sulfate reduction (15 °C) was very much higher than ambient 
temperature. 

Methanogenesis in high-sulfate marine sediments is usually low because of competitive 
inhibition by sulfate reducers (Nedwell, 1984). While there is some evidence that low 
temperature may influence carbon flow through benthic communities, inhibiting 
hydrogenotrophic methanogenesis and shifting towards acetoclastic methanogenesis 
and acetogenesis (Nozhevnikova et al., 1997; Schulz & Conrad, 1996; Fey & Conrad, 
2000), there is no current evidence for this occurring in high-latitude coastal marine 
environments. 

As in most coastal zone sediments, nitrate respiration in polar sediments accounts for 
only a very small proportion of organic matter degradation, largely due to the relatively 
low concentrations of nitrate in bottom water. Of the denitrification in Young 
Sound, Greenland, the vast majority (93 %) was derived from coupled nitrification- 
denitrification (Rysgaard et al., 1998) within the sediment, and not from nitrate from 
the water column, where nitrate concentrations were low. Rates of denitrification 
reported from Arctic sediments (Devol et al., 1997; Glud et al., 1998; Henriksen et al., 
1993) are similar to those reported from temperate sediments (e.g. Dong et al., 2000; 

SGM symposium 65 



190 D. B. Nedwell 



Jenkins & Kemp, 1984; Lohse et al., 1996; Nielsen, 1992; Rysgaard et ai, 1995) and give 
no indication that low temperature influences nitrate respiration. 

While functional group activity in coastal zone biogeochemical processes between 
regions seems similar, is there any evidence of different phylogenetic composition of 
polar and temperate/tropical functional group communities bringing about these 
processes? Molecular studies to date have largely focused on the phylogenetically 
cohesive groups (sulfate reducers and methanogenic archaea) to which the 16S rRNA 
approach is applicable, rather than to the phylogenetically diverse aerobic and nitrate- 
reducing communities. Molecular studies of polar marine sediments (Purdy et al., 2003; 
Ravenschlag et al., 1999, 2001; Sahm et al., 1999) have shown them to have diverse 
communities of sulfate reducers and methanogenic archaea, although the diversity of 
both the archaeal and sulfate-reducer communities in Antarctic coastal sediment was 
lower than in communities in temperate coastal sediments (Purdy et al. 2003). Only five 
distinct groups of archaea were present in Signy Is. coastal sediment compared with 12 
from temperate estuarine sediments. It has been suggested previously (e.g. Colinvaux, 
1993) that community diversity is likely to be lower in physically or chemically stressed 
environments than in communities which are not physically or chemically stressed but 
where biological competition controls diversity, and this may be the case in these 
Antarctic sediments. 

Methanogenesis in a Signy Is. coastal sediment represented only 2 % of carbon flow, 
and there was an equivalently small proportion (0-1 %) of methanogenic archaeal 
rRNA in the total sedimentary rRNA. (In contrast, in Signy Is. freshwater lake sediment 
archaeal rRNA represented 34% of the prokaryotic community.) The coastal 
methanogens included members of both the Methanosarcinales, particularly Methano- 
coccoides, which metabolize C 1 compounds, and Metbanomicrobiales, which meta- 
bolize hydrogen. Inhibition of sulfate reducers immediately stimulated methanogenesis 
in the coastal sediment, which suggests that, as in temperate sediments, the methano- 
gens are outcompeted for substrates by sulfate reducers, which dominated the terminal 
steps of organic matter breakdown. Of the total prokaryotic rRNA in the Signy Is. 
coastal sediment, 15 % represented sulfate reducers, dominated by the Desulfotaleal 
Desulforhopalus group. Desulfotalea are psychrophiles first isolated from Arctic 
sediments (Knoblauch et al., 1999; Sahm et ai, 1999), and seem to be quantitatively 
important in polar sediments. They are all incomplete oxidizers of their organic 
substrates, as are Desulfovibrio, which have not been detected in either Arctic or 
Antarctic sediments (Sahm et ai, 1999); Knoblauch et al. (1999) suggested that the 
Desulfotaleal Desulforhopalus group is the ecological equivalent in polar sediments of 
Desulfovibrio. As these groups are incomplete oxidizers, what completes the minerali- 
zation of organic matter in low-temperature coastal sediments by oxidizing acetate to 

SGM symposium 65 



Biogeochemical cycling in the coastal zone 191 

C0 2 , a step which in temperate coastal sediment is brought about by sulfate reduction 
(Nedwell, 1984) and which in temperate estuarine sediments was attributed to 
Desulfobacter (Purdy et ai, 1997, 2001) ? 

The acetoclastic groups Desulfobacter and Desulfotomaculum reported in temperate 
sediments were not detected in the Antarctic sediment. Ravenschlag et al. (1999) 
reported a large number of clones in a general bacterial clone library from Svalbard 
sediment that were related to S°- or Fe° + - or Mn 4+ -reducing Desulfuromonas, which 
utilize acetate, and closely related clones were present in Signy Is. coastal sediment 
(Purdy et al., 2003). The implication is that acetate oxidation in these low-temperature 
sediments may be achieved by the acetate-oxidizing, S°-reducing Desulfuromonas (i.e. 
that S° reduction rather than sulfate reduction may be important in the terminal step 
of organic carbon mineralization at low temperature) or by acetate-utilizing Pelobacter 
or Geobacter. 

ROLE OF COASTAL WETLANDS 

Biogeochemical cycling of elements in the coastal zone may be significantly influenced 
by the presence of coastal wetlands, predominantly salt marshes in the temperate 
region and mangroves in the tropics. The two ecosystems may be regarded as ecological 
equivalents: which system dominates seems to be determined by temperature in the 
coldest month > 16 °C or in the warmest month >24 °C. Salt marshes and mangroves 
are highly productive ecosystems and one of the earliest hypotheses was that of 
'outwelling' (Teal, 1962) ; that salt marshes export energy in the form of organic matter 
to the coastal zone. The same concept was proposed for mangroves (Odum & Heald, 
1972; Robertson & Duke, 1990), and also extended to the idea of net export of 
nutrients such as nitrogen to the tropical coastal zone, in which primary production 
tends to be nutrient limited. Subsequent research indicated that net export from both 
salt marshes (Nixon, 1980) and mangroves varied enormously depending upon local 
physical conditions and the productivity of the biological community. Twilley et al. 
(1992) suggested that export from mangroves declined with increasing latitude as their 
productivity and litter production decreased; export from both mangroves and salt 
marshes has been related (e.g. Wolanski et al., 1992) to the strength of the local tidal 
influence, with high-amplitude tides and strong directional water flow in estuaries 
resulting in increased tidal export from estuarine salt marshes or mangroves. Export 
of nitrogen to the coastal zone can undoubtedly occur from some mangroves, 
predominantly as particulate or dissolved organic nitrogen (e.g. Rivera-Monroy et al., 
1995), but they are usually sinks for inorganic nitrogen. However, the zone of influence 
of the mangrove or salt marsh in the adjacent coastal zone seems to be limited. Rodelli 
et al. (1984) measured d 13 C values in mangrove tissue, algae and consumer organisms 
and could trace the signal from mangrove detritus in the adjacent coastal zone in 

SGM symposium 65 



192 D. B. Nedwell 



Malaysia only within about 2 km of the mangrove; similar restricted zones of influence 
have been reported for Australian (Alongi, 1990), African (Hemminga et al., 1994; 
Marguillier et al., 1997), Indian (Bouillon et al., 2002) and South American (Jennerjahn 
& Ittekkot, 1999, 2002) mangroves. The direct impact of mangrove-derived organic 
matter therefore seems to be of only limited extent in the adjacent coastal zone. 

While the net export of either energy or nutrients from salt marshes or mangroves to 
the coastal zone may not be significant on an annual basis, in the temperate region there 
may be significant seasonal differences in their exchange and impact on the coastal 
zone. Thus, a UK salt marsh was balanced overall with respect to its annual nitrogen 
exchange with tidal water but there was net export of dissolved and particulate organic 
nitrogen and a smaller amount of ammonium during summer, when coastal sea- 
water nitrogen concentrations are low and the coastal zone phyto plankton are nutrient- 
depleted, but consistent removal of nitrate from tidal water by the salt-marsh sediments 
(Azni & Nedwell, 1986). This ability means that salt marshes, mangroves and estuarine 
sediments are highly efficient processors of nutrients in coastal sea water and buffer the 
loads of nutrients in the coastal zone, particularly when ambient inorganic nitrogen, 
usually nitrate, concentrations are increased by anthropogenic inputs (Corredor 6c 
Morell, 1994; Nedwell, 1975; Nedwell et al., 1999). It has been estimated that estuarine 
and coastal sediments remove up to 50 % of the nitrate load through temperate 
estuaries (Nedwell, 1975; Seitzinger,1988; Nixon et al., 1996). Jickells et al. (2000) 
argued that the pristine Humber estuary, with its extensive wetland and salt-marsh 
area, was likely to have been an effective sink for both N and P, but removal of >90 % of 
these wetlands has reduced this capacity. Nonetheless, the removal of anthropogeni- 
cally increased N and P loads through estuaries by estuarine sediments, including salt 
marshes and mangroves, remains an important 'buffer' to ameliorate the impact of 
these loads on the coastal zone. 

The coastal ocean, estuaries and coastal wetlands may also be important in the global 
carbon budget. In an extensive review of ocean-atmosphere C0 9 fluxes in the coastal 
ocean Borges (2005) has argued that water-air fluxes of C0 2 and fluxes of dissolved 
inorganic carbon (DIC) from mangroves and salt marshes are important terms in their 
carbon budgets which have so far been ignored. Although the DIC fluxes do not seem to 
have significant effect on the DIC and pC0 2 values in adjacent coastal waters, their 
air-water C0 2 exchange, when scaled up, seems to have a significant influence upon 
global C0 2 budgets. If estuaries and coastal wetlands are not taken into consideration, 
the coastal ocean appears to be a sink for atmospheric C0 9 (-1-17 mol C m~ 2 year _1 ), 
and the estimated uptake of C0 2 by the global ocean is -1-93 Pg C year -1 . If estuaries 
and coastal wetlands, with their high heterotrophic activity, are included in the scale-up, 
the coastal ocean behaves as a net source of C0 2 (0-38 mol C year ) and the C0 2 

SGM symposium 65 



Biogeochemical cycling in the coastal zone 193 

uptake by the total global ocean decreases to -144 Pg year -1 . There are also interesting 
latitudinal differences. At high latitudes and in the tropics and subtropics the model 
predicts that, including estuaries and coastal wetlands, the coastal ocean is a net source 
of CO? to the atmosphere, but at temperate latitudes it remains a moderate sink for 
atmospheric C0 7 . Much work remains to be done to constrain these global estimates 
adequately, but the significance of the coastal zone processes seems clear. 

ACKNOWLEDGEMENTS 

/ would particularly like to thank my colleagues Kevin Purdy and Martin Embley for their 
productive, enjoyable, extensive and continued collaboration on the application of molecular 
techniques to coastal zone microbial ecology over an extended period and to the Natural 
Environment Research Council, UK, for their support in a number of research grants for this 
work. Thanks also to the British Antarctic Survey, particularly Cynan Ellis-Evans, for their 
collaboration and logistic support for the Antarctic research at Signy Island. I also acknowledge 
the helpful comments of Professor Michael Danson on adaptation of extremophile enzymes to 
high or low temperature. 

REFERENCES 

Alongi, D. M. (1990). Abundances of benthic microfauna in relation to outwelling of 

mangrove detritus in a tropical coastal region. Mar Ecol Prog Ser 63, 53-63. 
Alongi, D. M. ; Tirendi, R, Dixon, P., Trott, L A. & Brunskill, G. J. (1999). Mineralization 

of organic matter in intertidal sediments of a tropical semi-enclosed delta. Estuar 

Coastal Shelf Sci 48, 451-467. 
Arnosti, C. & Jorgensen, B. B. (2003). High activity and low temperature optima 

of extracellular enzymes in Arctic sediments: implications for carbon cycling by 

heterotrophic microbial communities. Mar Ecol Prog Ser 249, 1 5-24. 
Azni, S. A. &Nedwell, D. B. (1986). The nitrogen cycle of an East Coast, U.K. saltmarsh. II. 

Nitrogen fixation, nitrification, denitrification, tidal exchange. Estuar Coastal Shelf 

Sci 22, 689-704. 
Borges, A. V. (2005). Do we have enough pieces of the jigsaw to integrate C0 2 fluxes in the 

coastal ocean? Estuaries 28, 3-27. 
Bouillon, S., Raman, A. V., Dauby, P. & Dehairs, F. (2002). Carbon and nitrogen stable 

isotope ratios of subtidal benthic invertebrates in an estuarine mangrove ecosystem 

(Andhra Pradesh, India). Estuar Coastal Shelf Sci 54, 901-913. 
Button, D. K. (1993). Nutrient-limited microbial growth kinetics: overview and recent 

advances. Antonie van Leeuwenhoek 63, 225-235. 
Canfield, D. E., Jorgensen, B. B., Fossing, H. & 7 other authors (1993). Pathways of 

organic carbon oxidation in three continental margin sediments. Mar Geol 1 13, 

27-40. 
Colinvaux, P. (1993). Ecology 2. New York: Wiley. 
Corredor, J. E. & Morell, J. M. (1994). Nitrate depuration of secondary sewage effluents 

in mangrove sediments. Estuaries 17, 295-300. 
Daly, K. L. & Smith, W. O., Jr (1993). Physical-biological interactions influencing marine 

plankton production. Annu Rev Ecol Syst 24, 555-585. 

SGM symposium 65 



194 D. B. Nedwell 



Daniel, R. M. ; Danson, M. J. & Eisenthai, R. (2001). The temperature optima of enzymes: 
a new perspective on an old phenomenon. Trends Biochem Sci 26, 223-225. 

de Baar, H. J. W., de Jong, J. T. M., Bakker, D. C. E., Loscher, B. M., Veth, C, Bathmann, 
U. & Smetacek, V. (1995). Importance of iron for plankton blooms and carbon 
dioxide drawdown in the Southern Ocean. Nature 373, 412-41 5. 

Devol, A. H., Codispoti, L. A. & Christensen, J. P. (1997). Summer and winter denitrifi- 
cation rates in western Arctic shelf sediments. Continent Shelf Res 17, 1029-1033. 

Dong, L. F., Thornton, D. C. O., Nedwell, D. B. & Underwood, G. J. C. (2000). 
Denitrification in sediments of the River Colne estuary, England. Mar Ecol Prog Ser 
203, 109-122. 

Eyre, B. & Balls, P. (1999). A comparative study of nutrient behavior along the salinity 
gradient of tropical and temperate estuaries. Estuaries 22, 31 3-326. 

Feller, G. & Gerday, C. (2003). Psychrophilic enzymes: hot topics in cold adaptation. Nat Rev 
Microbiol 1, 200-208. 

Fey, A. & Conrad, R. (2000). Effect of temperature on carbon and electron flow and on the 
archaeal community in methanogenic rice field soil. Appl Environ Microbiol 66, 
4790^797. 

Franzmann, P. D. (1996). Examination of Antarctic prokaryote diversity through molecular 
comparisons. Biodivers Conserv 5, 1 295-1 305. 

Glud, R. N., Holby, O., Hoffmann, F. & Canfield, D. E. (1998). Benthic mineralization and 
exchange in Arctic sediments (Svalbard, Norway). Mar Ecol Prog Ser 173, 237-251 . 

Hemminga, M. A., Slim, F. J., Kazungu, J., Gaanssen, G. M., Nieuwenhuize, J. & Kruyt, 
N. M. (1994). Carbon outwellings from a mangrove forest with adjacent seagrass 
beds and coral reefs(Gazi Bay, Kenya). Mar Ecol Prog Ser 106, 291-301. 

Henriksen, K., Blackburn, T. H., Lomstein, B. A. & McRoy, C. P. (1993). Rates of nitrifi- 
cation, distribution of nitrifying bacteria and inorganic N fluxes in northern Bering- 
Chukchi shelf sediments. Continent Shelf Res 13, 629-651 . 

Holmboe, N., Kristensen, E. & Andersen, F. 0. (2001). Anoxic decomposition in sedi- 
ments from a tropical mangrove forest and the temperate Wadden Sea: implications 
of N and P addition experiments. Estuar Coastal Shelf Sci 53, 125-140. 

Hoppe, H.-G., Gocke, K., Koppe, R. & Begler, S. (2002). Bacterial growth and primary 
production along a north-south transect of the Atlantic Ocean. Nature 416, 1 68— 
171. 

Hulth, S., Blackburn, T. H. & Hall, P. O. J. (1994). Arctic sediments (Svalbard): consumption 
and microdistribution of oxygen. Mar Chem 46, 293-316. 

Jenkins, M. C. & Kemp, W. M. (1984). The coupling of nitrification and denitrification in 
two estuarine sediments. Limnol Oceanogr 29, 609-61 9. 

Jennerjahn, T. C. & Ittekkot, V. (1999). Changes in organic matter from surface waters to 
continental slope sediments off the Sao Francisco River, eastern Brazil. Mar Geo! 1 61 , 
129-140. 

Jennerjahn, T. C. & Ittekkot, V. (2002). Relevance of mangroves for the production and 
deposition of organic matter along tropical continental margins. Naturwissen- 
schaften 89, 23-30. 

Jickells, T., Andrews, J., Samways, G. & 7 other authors (2000). Nutrient fluxes through 
the Humber estuary -past, present and future. Ambio 29, 130-135. 

Joint, I. & Pomroy, A. (1993). Phytoplankton biomassand production in the southern North 
Sea . Mar Ecol Prog Ser 99, 1 69-1 82 . 

Jorgensen, B. B. & Revsbech, N. P. (1985). Diffusive boundary layer and the oxygen uptake 
of sediments and detritus. Limnol Oceanogr 30, 1 1 1-1 22. 



SGM symposium 65 



Biogeochemical cycling in the coastal zone 195 



Karl, D. M., Christian, J. R., Dore, J. E. & Letelier, R. M. (1996). Microbial oceanography in 

the region west of the Antarctic Peninsula: microbial dynamics, nitrogen cycle and 

carbon flux. In Foundation for Ecological Research West of the Antarctic Peninsula, 

Antarctic Research Series vol. 70, pp. 303-332. Washington, DC: American 

Geophysical Union. 
King, D. & Nedwell, D. B. (1984). Changes in the nitrate-reducing community of an 

anaerobic saltmarsh sediment in response to seasonal selection by temperature. 

J Gen Microbiol MO, 2935-2941 . 
Knoblauch, C, Sahm, K. & Jorgensen, B. B. (1999). Psychrophilic sulfate-reducing 

bacteria isolated from permanently cold Arctic marine sediments: description of 

Desulfofrigus oceanense gen. nov., sp. nov., Desulfofrigus fragile sp. nov., Desulfo- 

faba gelida gen. nov., sp. nov., Desulfotalea psychrophila gen. nov., sp. nov. and 

Desulfotalea arctica sp. nov. lntJSystBacterio/49, 1631-1643. 
Kocum, E., Underwood, G. J. C. & Nedwell, D. B. (2002a). Simultaneous measurement 

of phytoplanktonic primary production, nutrient and light availability along a 

turbid, eutrophic UK east coast estuary (the Colne Estuary). Mar Ecol Prog Ser 231 , 

1-12. 
Kocum, E., Nedwell, D. B. & Underwood, G. J. C. (2002b). Regulation of phytoplankton 

primary production along a hypernutrified estuary. Mar Ecol Prog Ser 231 , 13-22. 
Kristensen, E., Holmer, M. & Bussarawit, N. (1991). Benthic metabolism and sulphate 

reduction in a southern Asian mangrove swamp. Mar Ecol Prog Ser 73, 93-103. 
Kristensen, E., Devol, A., Ahmed, S. & Saleem, M. (1992). Preliminary study of benthic 

metabolism and sulfate reduction in a mangrove swamp of the Indus Delta, Pakistan. 

Mar Ecol Prog Ser 90, 287-297. 
Li, W. K. W. (1980). Temperature adaptation in phytoplankton: cellular and photosynthetic 

characteristics. In Primary Productivity in the Sea, Environmental Science Research 

series, vol. 19, pp. 259-279. Edited by P. G. Falkowski. New York: Plenum. 
Li, W. K. W. (1985). Photosynthetic response to temperature of marine phytoplankton along 

a latitudinal gradient (1 6° N to 74° N). Deep Sea Res 32, 1 381-1 391 . 
Li, W. K. W., Smith, J. C. & Piatt, T. (1984). Temperature responses of photosynthetic 

capacity and carboxylase activity in Arctic marine phytoplankton. Mar Ecol Prog Ser 

17,237-243. 
Lohse, L., Kloosterhuis, T., van Raaphorst, W. & Helder, W. (1996). Denitrification rates 

as measured by the isotope pairing method and by the acetylene inhibition 

technique in continental shelf sediments of the North Sea. Mar Ecol Prog Ser 132, 

169-179. 
Longhurst, A., Sathyendranath, S., Piatt, T. & Caverhill, C. (1995). An estimate of global 

primary production in the ocean from satellite radiometer data. J Plankton Res 17, 

1245-1271. 
Mackin, J. E. & Swider, K. T. (1989). Organic matter decomposition pathways and oxygen 

consumption in coastal marine sediments. J Mar Res 47, 681-716. 
Marguillier, S., van derVelde, G., Dehairs, E, Hemminga, M. A. & Rajagopal, S. (1997). 

Trophic relationships in an interlinked mangrove-seagrass ecosystem as traced by 

6 1 3 C and 6 1 5 N. Mar Ecol Prog Ser 151, 115-121. 
Martin, J. H., Gordon, R. M. & Fitzwater, S. E. (1990). Iron in Antarctic waters. Nature 

345, 156-158. 
Mincks, S. L., Smith, C. R. & DeMaster, D. J. (2005). Persistence of labile organic matter 

and microbial biomass in Antarctic shelf sediments: evidence of a sediment 'food 

bank'. Mar Ecol Prog Ser (in press). 



SGM symposium 65 



196 D. B. Nedwell 



Morita, R. Y. (1975). Psychrophilic bacteria. Bacteriol Rev 39 , 144-167. 

Nedwell, D. B. (1975). Inorganic nitrogen metabolism in a eutrophicated tropical mangrove 

estuary. Water Res 9, 221-231 . 
Nedwell, D. B. (1984). The input and mineralisation of organic carbon in anaerobic aquatic 

sediments. In Advances in Microbial Ecology, pp. 93-131 . Edited by K. C. Marshall. 

London: Plenum. 
Nedwell, D. B. (1989). Benthic microbial activity in an Antarctic coastal sediment at Signy 

Island, South Orkney Islands. Estuar Coastal Shelf Sci 28, 507-516. 
Nedwell, D. B. (1999). Effect of low temperature on microbial growth: lowered affinity for 

substrates limits growth at low temperature. FE MS Microbiol Ecol 30, 101-111. 
Nedwell, D. B. & Rutter, M. (1994). Influence of temperature on growth rate and 

competition between two psychrotolerant Antarctic bacteria: low temperature 

diminishes affinity for substrate uptake. Appl Environ Microbiol 60, 1 984-1 992. 
Nedwell, D. B., Walker, T. R., Ellis-Evans, J. C. & Clarke, A. (1993). Measurements of 

seasonal rates and annual budgets of organic carbon fluxes in an Antarctic coastal 

environment at Signy Island, South Orkney Islands, suggest a broad balance between 

production and decomposition. Appl Environ Microbiol 59, 3989-3995. 
Nedwell, D. B., Blackburn, T. H. & Wiebe, W. J. (1 994). Dynamic nature of the turnover of 

organic carbon, nitrogen and sulphur in the sediments of a Jamaican mangrove 

forest. Mar Ecol Prog Ser 1 1 0, 223-23 1 . 
Nedwell, D. B., Jickells, T., Trimmer, M. T. & Sanders, R. (1999). Nutrients in estuaries. 

In Estuaries, Advances in Ecological Research vol. 29, pp. 43-92. Edited by 

D. B. Nedwell & D. Raffaelli. London: Harcourt. 
Nielsen, L. P. (1992). Denitrification in sediment determined from nitrogen isotope pairing. 

FEMS Microbiol Ecol 86, 3 57-362 . 
Nixon, S. W. (1 980). Between coastal marshes and coastal waters - a review of twenty years 

of speculation and research on the role of saltmarshes in estuarine productivity and 

water chemistry. In Estuarine and Wetland Processes with Emphasis upon Modelling, 

pp. 437-525. Edited by P. Hamilton & K. B. Macdonald. New York: Plenum. 
Nixon, S. W., Ammerman, J. W., Atkinson, L P. & 13 other authors (1996). The fate of 

nitrogen and phosphorus at the land-sea margin of the North Atlantic Ocean. 

Biogeochemistry 35,141-180. 
Nozhevnikova, A. N., Holliger, C, Ammann, A. & Zehnder, A. J. B. (1997). Methano- 

genesis in sediments from deep lakes at different temperatures (2-70 °C). Water Sci 

Technol 36 (6-7), 57-64. 
Odum, W. E. & Heald, E. J. (1972). Trophic analyses of an estuarine mangrove community. 

Bull Mar Sci 22, 67 1-738. 
Ogilvie, B. G., Rutter, M. & Nedwell, D. B. (1997). Selection by temperature of nitrate- 
reducing bacteria from estuarine sediments: species composition and competition 

for nitrate. FEMS Microbiol Ecol 23, 1 1-22. 
Osborne, P. L. (2000). Tropical Ecosystems and Ecological Concepts. Cambridge: Cambridge 

University Press. 
Peterson, M. E., Eisenthal, R., Danson, M. J., Spence, A. & Daniel, R. M. (2004). A new 

intrinsic thermal parameter for enzymes reveals true temperature optima. J Biol 

Chem 279, 20717-20722. 
Pfannkuche, O. & Thiel, H. (1987). Meiobenthic stocks and benthic activity on the 

NE-Svalbard Shelf and in the Nansen Basin. Polar Biol 7, 253-266. 
Pomeroy, L. R. & Deibel, D. (1986). Temperature regulation of bacterial activity during 

the spring bloom in Newfoundland coastal waters. Science 233, 359-361 . 



SGM symposium 65 



Biogeochemical cycling in the coastal zone 197 



Pomeroy, L. R., Wiebe, W. J., Deibel, D., Thompson, R. J., Rowe, G. T. & Pakulski, J. D. 

(1991). Bacterial responses to temperature and substrate concentration during the 

Newfoundland spring bloom. Mar Ecol Prog Ser 75, 143-159. 
Purdy, K. J., Nedwell, D. B. ; Embley, T. M. &Takii, S. (1997). Use of 16S rRNA-targeted 

oligonucleotide probes to investigate the occurrence and selection of sulfate- 

reducing bacteria in response to nutrient addition to sediment slurry microcosms 

from a Japanese estuary. FE MS Microbiol Ecol 24, 221-234. 
Purdy, K. J., Nedwell, D. B., Embley, T. M. & Takii, S. (2001). Using 16S rRNA-targeted 

oligonucleotide probes to investigate the distribution of sulphate-reducing bacteria 

in estuarine sediments. FEMS Microbiol Ecol 36, 1 65-1 68. 
Purdy, K. J., Nedwell, D. B. & Embley, T. M. (2003). Analysis of the sulfate-reducing 

bacterial and methanogenic archaeal populations in contrasting Antarctic sediments. 

Appl Environ Microbiol 69, 3 1 8 1 -3 1 9 1 . 
Ravenschlag, K., Sahm, K., Pernthaler, J. & Amman, R. (1999). High bacterial diversity in 

permanently cold marine sediments. Appl Environ Microbiol 65, 3982-3989. 
Ravenschlag, K., Sahm, K. & Amman, R. (2001). Quantitative molecular analysis of the 

microbial community in marine Arctic sediments (Svalbard). Appl Environ Microbiol 

67, 387-395. 
Reay, D. S., Nedwell, D. B., Priddle, J. & Ellis-Evans, C. J. (1999). Temperature 

dependence of inorganic nitrogen uptake: reduced affinity for nitrate at suboptimal 

temperatures in both algae and bacteria. Appl Environ Microbiol 65, 2577-2584. 
Reay, D. S., Priddle, J., Nedwell, D. B., Whitehouse, M. J., Ellis-Evans, J. C, Deubert, C. 

& Connelly, D. (2001). Regulation by low temperature of phytoplankton growth 

and nutrient uptake in the Southern Ocean. Mar Ecol Prog Ser 219 , 51-64. 
Rivera-Monroy, H., Day, J. W., Twilley, R. W., Vera-Herrera, F. & Coronado-Molina, C. 

(1995). Flux of nitrogen and sediment in a fringe mangrove forest in Terminos 

Lagoon, Mexico. Estuar Coastal Shelf Sci 40, 139-160. 
Rivkin, R. B., Anderson, M. R. & Lajzerowicz, C. (1996). Microbial processes in cold 

oceans. I. Relationship between temperature and bacterial growth rate. Aquat 

Microb Ecol 10, 243-254. 
Robertson, A. I. & Duke, N. C. (1990). Mangrove fish communities in tropical Queensland, 

Australia: spatial and temporal patterns in densities, biomass and community 

structure. Mar Biol 104, 369-379. 
Robertson, A. I., Dixon, P. & Alongi, D. M. (1998). The influence of fluvial discharge on 

pelagic production in the Gulf of Papua, Northern Coral Sea. Estuar Coastal Shelf Sci 

46,319-331. 
Rodelli, M. R., Gearing, J. N., Gearing, P. J., Marshall, N. & Sasekumar, A. (1984). Stable 

isotope ratio as a tracer of mangrove carbon in Malaysian ecosystems. Oecologia 61, 

326-333. 
Russell, N. J. (1990). Cold adaptation of microorganisms. Philos Trans R Soc Lond B Biol Sci 

326,595-611. 
Russell, N. J. (1992). Psychrophilic bacteria. In Molecular Biology and Biotechnology of 

Extremophiles, pp. 203-224. Edited by R. A. Herbert & R. S. Sharp. Glasgow: Blackie. 
Russell, N. J. (1998). Molecular adaptations in psychrophilic bacteria: potential for 

biotechnological applications. Adv Biochem Eng Biotechnol 61, 1-21 . 
Rutter, M. & Nedwell, D. B. (1994). Influence of changing temperature on growth rate and 

competition between two psychrotolerant Antarctic bacteria: competition and 

survival in n on -steady-state temperature environments. Appl Environ Microbiol 60, 

1993-2002. 



SGM symposium 65 



198 D. B. Nedwell 



Rysgaard, S., Christensen, P. B. & Nielsen, L. P. (1995). Seasonal variation in nitrification 

and denitrification in estuarine sediment colonized by benthic microalgae and 

bioturbating infauna. Mar Ecol Prog Ser 126, 111-121. 
Rysgaard, S., Finster, K. & Dahlgaard, H. (1996). Primary production, nutrient dynamics 

and mineralisation in a northeastern Greenland fjord during the summer thaw. Polar 

Biol 16, 497-506. 
Rysgaard, S., Thamdrup, B., Risgaard-Petersen, N., Fossing, H., Berg, P., Christensen, 

P. B. & Dalsgaard, T. (1998). Seasonal carbon and nutrient mineralization in a high- 
Arctic coastal marine sediment, Young Sound, Northeast Greenland. Mar Ecol Prog 

Ser 175, 261-276. 
Rysgaard, S., Nielsen, T. G. & Hansen, B. W. (1999). Seasonal variation in nutrients, 

pelagic primary production and grazing in a high-Arctic coastal marine ecosystem, 

Young Sound, Northeast Greenland. Mar Ecol Prog Ser 17 '9, 13-25. 
Sahm, K. f Knoblauch, C. & Amann, R. (1999). Phylogenetic affiliation and quantification 

of psychrophilic sulfate-reducing isolates in marine Arctic sediments. Appl Environ 

Microbiol 65, 3976-3981 . 
Schulz, S. & Conrad, R. (1996). Influence of temperature on pathways to methane 

production in the permanently cold profundal sediment of Lake Constance. FEMS 

Microbiol Ecol 20, 1-14. 
Seitzinger, S. P. (1988). Denitrification in freshwater and coastal marine ecosystems: 

ecological and geochemical significance. Li m nol Ocea nog r 33, 702-724. 
Sieburth, J. M. (1967). Seasonal selection of estuarine bacteria by water temperature. J Exp 

Mar Biol Ecol 1 , 98-1 2 1 . 
Sinensky, M. (1974). Homeoviscous adaptation -a homeostatic process that regulates 

the viscosity of membrane lipids in Escherichia coli. Proc Natl Acad Sci USA 71, 

522-525. 
Sobey, N. S. (2004). The role of bottom sediments in nitrogen cycling in a tropical lagoon. 

PhD thesis, University of Essex, Colchester, UK. 
Tahey,T. M., Duinevald, G. C. A., Berghuis, E. M. & Helder, W. (1994). Relation between 

sediment-water fluxes of oxygen and silicate and faunal abundance at continental 

shelf, slope and deep water stations in northwest Mediterranean. Mar Ecol Prog Ser 

104, 119-130. 
Teal, J. M. (1962). Energy flow in the salt marsh ecosystem of Georgia. Ecology 43, 614- 

624. 
Trimmer, M., Nedwell, D. B., Sivyer, D. B. & Malcolm, S. J. (2000). Seasonal benthic 

organic matter mineralisation measured by oxygen uptake and denitrification along 

a transect of the inner and outer River Thames estuary, UK. Mar Ecol Prog Ser 1 97, 

103-119. 
Twilley, R. R., Chen, R. H. & Hargis, T. (1992). Carbon sinks in mangroves and their 

implications to carbon budget of tropical coastal ecosystems. Water Air Soil Poll ut 64, 

265-288. 
Upton, A. C. & Nedwell, D. B. (1989). Temperature responses of bacteria isolated from 

different Antarctic environments. In University Research in Antarctica: Proceedings of 

British Antarctic Survey Special Topic Award Scheme Symposium, pp. 97-101 . Edited 

by R. B. Heywood. Cambridge: British Antarctic Survey. 
Upton, A. C, Nedwell, D. B., Parkes, R. J. & Harvey, S. M. (1993). Seasonal benthic 

microbial activity in the southern North Sea: oxygen uptake and sulphate reduction. 

Mar Ecol Prog Ser 1 01 , 273-281 . 
Wiebe, W. J., Sheldon, W. M., Jr & Pomeroy, L R. (1992). Bacterial growth in the cold: 



SGM symposium 65 



Biogeochemical cycling in the coastal zone 199 



evidence for an enhanced substrate requirement. Appl Environ Microbiol 58, 

359-364. 
Wiebe, W. J., Sheldon, W. M., Jr & Pomeroy, L. R. (1993). Evidence for an enhanced 

substrate requirement by marine mesophilic bacterial isolates at minimal growth 

temperatures. Microb Ecol 25, 151-1 59. 
Wolanski, E., Mazda, Y. & Ridd, P. (1992). Mangrove hydrodynamics. In Tropical Mangrove 

Ecosystems, pp. 43-62. Edited by A. I. Robertson & D. M. Alongi. Washington, DC: 

American Geophysical Union. 



SGM symposium 65 



Fungal roles and function in rock f 
mineral and soil transformations 

Geoffrey M. Gadd, Marina Fomina and Euan P. Burford 

Division of Environmental and Applied Biology, Biological Sciences Institute, 
School of Life Sciences, University of Dundee, Dundee DD1 4HN, Scotland, UK 



INTRODUCTION 

The most important perceived environmental roles of fungi are as decomposer organ- 
isms, plant pathogens and symbionts (mycorrhizas, lichens), and in the maintenance of 
soil structure through their filamentous growth habit and production of exopolymers. 
However, a broader appreciation of fungi as agents of biogeochemical change is lacking 
and, apart from obvious connections with the carbon cycle, they are frequently 
neglected within broader microbiological and geochemical research contexts. While the 
profound geochemical activities of bacteria and archaea receive considerable attention, 
especially in relation to carbon-limited and/or anaerobic environments (see elsewhere 
in this volume), in aerobic environments fungi are of great importance, especially when 
considering rock surfaces, soil and the plant root-soil interface (Gadd, 2005a). For 
example, mycorrhizal fungi are associated with ~ 80 % of plant species and are involved 
in major mineral transformations and redistributions of inorganic nutrients, e.g. 
essential metals and phosphate, as well as carbon flow Free-living fungi have major 
roles in the decomposition of plant and other organic materials, including xenobiotics, 
as well as mineral solubilization (Gadd, 2004). Lichens (a symbiosis between an alga or 
cyanobacterium and a fungus) are one of the commonest members of the microbial 
consortia inhabiting exposed subaerial rock substrates, and play fundamental roles in 
early stages of rock colonization and mineral soil formation. Fungi are also major 
biodeterioration agents of stone, wood, plaster, cement and other building materials, 
and it is now realized that they are important components of rock-inhabiting microbial 
communities, with significant roles in mineral dissolution and secondary mineral 
formation (Burford et al., 2003a, b). The objective of this chapter is to outline 

SGM symposium 65: Micro-organisms and Earth systems - advances in geomicrobiology. 
Editors G. M. Gadd, K. T. Semple & H. M. Lappin-Scott. Cambridge University Press. ISBN 521 86222 1 ©SGM 2005 



202 G. M. Gadd ; M. Fomina and E. P. Burford 

important fungal roles and function in rock, mineral and soil transformations and to 
emphasize the importance of fungi as agents of geological change. 



WEATHERING PROCESSES AND THE INFLUENCE OF MICROBES 

The composition of the Earth's lithosphere, biosphere, hydrosphere and atmosphere 
is influenced by weathering processes (Ferris et al., 1994; Banfield et al, 1999; Vaughn 
et al., 2002). Rock substrates and their mineral constituents are weathered through 
physical (mechanical), chemical and biological mechanisms; the relative significance 
of each process varying widely depending on environmental and other conditions. 
Near-surface weathering of rocks and minerals, which occurs in subaerial (i.e. situated, 
formed or occurring on or immediately adjacent to the surface of the Earth) and sub- 
soil (i.e. not exposed to the open air) environments, will often involve an interaction 
between all three mechanisms (White et al., 1992), and it is the biological component of 
the overall process that provides much microbiological interest. At or near the Earth's 
surface, interaction between minerals, metals and non-metallic species in an aqueous 
fluid nearly always involves the presence of microbes or their metabolites (Banfield & 
Nealson, 1998). Mineral replacement reactions in rocks mainly occur by dissolution- 
re-precipitation processes (e.g. cation exchange, chemical weathering, leaching and 
diagenesis), where one mineral or mineral assemblage is replaced by a more stable 
assemblage (Putnis, 2002). Micro-organisms can influence this by mineral dissolution, 
biomineralization and alteration of mineral surface chemistry and reactivity (Hochella, 
2002). Mineral dissolution can also be inhibited by extracellular microbial poly- 
saccharides that passivate reactive centres on minerals (Welch & Vandevivere, 1994; 
Welch et al., 1999). In contrast, mineral dissolution may be accelerated by micro bially 
mediated pH changes and many other changes in solution chemistry. Excretion 
of organic ligands and siderophores not only enhances nutrient acquisition by micro- 
organisms but can markedly affect mineral composition and dissolution reactions 
(Grote & Krumbein, 1992; Maurice et al., 1995; Hersman et al., 1995; Stone, 1997; 
Gadd, 1999; Kraemer et al, 1999; Liermann et al., 2000; Sayer & Gadd, 2001). In 
addition, the formation of secondary minerals (biogenic crystalline precipitates) can 
occur through metabolism-independent and metabolism-dependent processes, often 
being influenced by abiotic and biotic factors such as environmental pH, microbial cell 
density and the composition of cell walls (Ferris et al., 1987; Gadd, 1990, 1993; Arnott, 
1995; Thompson & Ferris, 1990; Douglas & Beveridge, 1998; Fortin et al, 1998; 
Sterflinger, 2000; Verrecchia, 2000). The term 'biomineralization' refers to biologically 
induced mineralization where an organism modifies the local microenvironment, 
creating conditions that promote chemical precipitation of extracellular mineral phases 
(Hamilton, 2003). Secondary minerals can form by direct nucleation on cellular 
macromolecules, e.g. melanin and chitin in fungal cell walls (Gadd, 1990, 1993; Fortin 

SGM symposium 65 



Mycotransformation of minerals 203 

& Beveridge, 1997; Beveridge et al., 1997). However, indirect precipitation of secondary 
minerals also frequently occurs as a result of microbially mediated changes in solution 
conditions (Fortin &£ Beveridge, 1997; Banfield et al., 2000). Regardless of the mech- 
anism, lithospheric weathering of rocks and minerals can result in the mobilization and 
redistribution of essential nutrients (e.g. P, S) and metals (e.g. Na, K, Mg, Ca, Mn, Fe, 
Cu, Zn, Co and Ni) required for plant and microbial growth. In addition, non-essential 
metals (e.g. Cs, Al, Cd, Hg and Pb) may also be mobilized from mineral and soil pools 
(Gadd, 1993, 2001a, b; Morley et al, 1996). 

Bioweathering can be defined as the erosion, decay and decomposition of rocks 
and minerals mediated by living organisms. Micro-organisms, as well as animals and 
plants, can weather rock aggregates through biomechanical and biochemical attack on 
component minerals (Goudie, 1996; Adeyemi 6c Gadd, 2005). Filamentous micro- 
organisms, plant roots and burrowing animals can physically affect rocks and enhance 
splitting and fractionation: the disruptive (hydraulic) pressure of growing roots and 
hyphae is important here (Sterflinger, 2000; Money, 2001). However, biochemical 
actions of organisms are believed to be more significant processes than mechanical 
degradation (Sterflinger, 2000; Etienne, 2002). Microbes, e.g. bacteria, algae and fungi, 
and plants can mediate chemical weathering of rocks and minerals through excretion of 
for example H + , organic acids and other metabolites, while respiratory CO-, can lead to 
carbonic acid attack on mineral surfaces (Johnstone & Vestal, 1993; Ehrlich, 1998; 
Sterflinger, 2000; Gadd & Sayer, 2000). Biochemical weathering of rocks can result in 
changes in the microtopography of minerals through pitting and etching, mineral 
displacement reactions and even complete dissolution of mineral grains (Ehrlich, 1998; 
Kumar & Kumar, 1999; Adeyemi St Gadd, 2005). 



MICROBES IN ROCK AND MINERAL HABITATS 

Micro-organisms occur in rock and building stone in a variety of microhabitats, being 
classed as epilithic, hypolithic, endolithic, chasmolithic, cryptoendolithic and euendo- 
lithic organisms (Fig. 1) (Gerrath et al., 1995, 2000; May, 2003). Epiliths are common 
under humid conditions and occur on the surface of rocks and building stone. Hypo- 
lithic micro-organisms are often found under and attached to pebbles, particularly in 
hot and cold deserts. Endoliths inhabit the rock subsurface and may form distinct 
masses or brightly coloured layers. Endolithic micro-organisms can occur as chasmo- 
liths that grow in pre-existing cracks and fissures within rock, often being visible from 
the rock surface. Conversely, cryptoendoliths grow inside cavities and among crystal 
grains and cannot be observed from the rock surface. Euendolithic microbes are a 
specialized group of cryptoendoliths that are capable of actively penetrating (boring) 
into submerged rock (Ehrlich, 1998; Gerrath et al., 1995). 

SGM symposium 65 



204 G. M. Gadd, M. Fomina and E. P. Burford 







Fig. 1. Terminology for rock-dwelling micro-organisms. 1, Epilithic micro-organisms, occurring on 
the surface of rocks and building stone; 2, hypolithic micro-organisms, found under and attached to 
pebbles; 3-5, endolithic micro-organisms, inhabiting the rock subsurface, which include chasmoliths 
(3), which grow in pre-existing cracks and fissures within the rock, cryptoendoliths (4), which grow 
inside cavities and among crystal grains, and euendoliths (5), which are a specialized group of 
cryptoendoliths that are capable of actively penetrating (boring) into rock. Organisms shown can 
represent various microbial groups including bacteria, cyanobacteria, microalgae, fungi and lichens. 

Micro-organisms play a fundamental role in mineral transformations in the natural 
environment, most notably in the formation of mineral soils from rock and the cycling 
of elements (May, 2003; Gadd, 2005a). It is not surprising therefore that a wide variety 
of micro-organisms, including bacteria, algae and fungi, inhabit rocks and stone- 
work of buildings and historic monuments (Ehrlich, 2002; Burford et al., 2003a; 
Gleeson et al., 2005). Exposed surfaces are not necessarily conducive to microbial 
growth, as a result of moisture deficit, exposure to UV solar radiation and limited 
availability of nutrients. However, complex interactions between microbes and the 
mineral substrate are frequently observed, often to some distance into the mineral 
(May, 2003). The rock micro-environment is subject to diurnal and seasonal changes in 
for example temperature and moisture as well as in available nutrients (Gorbushina & 
Krumbein, 2000; Roldan et al., 2002). Nutrients may accumulate as a result of water 
interactions, wind-blown dust particles, animal faeces and death and degradation of 
living organisms and be utilized by microbes. Mineral grains within the host-rock may 
also serve as a source of metals essential for microbial growth. The transfer of 
biological material (e.g. fungal spores and other reproductive structures) from external 



SGM symposium 65 



Mycotransformation of minerals 205 

sources may also play a role in the colonization of subaerial environments by microbes. 
Physical properties (e.g. porosity) and elemental composition of the host rock (e.g. C, P, 
K, S and metal content) may govern initial establishment, growth and survival of 
microbial communities (Gleeson et al., 2005). Thus, colonization of rock substrates by 
microbes and the development of a microbial consortium is likely to be influenced 
by physical and chemical properties and interactions based on environmental (e.g. 
macro-/micro-climate) and biological factors resulting in and influencing ecological 
succession at the micro-scale. 

MICROBIAL PROCESSES INFLUENCED BY MINERALS 

Many important microbial processes can be influenced by minerals, including energy 
generation, nutrient acquisition, cell adhesion and biofilm formation (Hochella, 2002; 
see elsewhere in this volume). Micro-organisms can also acquire essential nutrients for 
microbial growth from mineral surfaces, which effectively concentrate these vital 
nutrients far above surrounding environmental levels, e.g. C, N, P, Fe and various 
organic compounds (Vaughn et al., 2002). Some environmental contaminants, which 
may be concentrated on mineral surfaces by various sorption reactions, can be dis- 
placed by similar microbial processes (Kraemer et al., 1999). In addition, it is likely that 
potentially toxic metals, released from minerals as a result of physico-chemical and 
biological processes, will have an effect on microbial communities (Gadd, 2005b). 
Mineral surface properties (e.g. microtopography, surface composition, surface charge 
and hydrophobicity) also play an integral role in microbial attachment and detachment. 
They are therefore critical in biofilm formation and the ecology of microbial popu- 
lations in and on mineral substrates (Wolfaardt et al., 1994; Fredrickson et al., 1995; 
Bennett et al., 1996; Rogers et al., 1998) . 

FUNGI IN THE TERRESTRIAL ENVIRONMENT 

Fungi are ubiquitous components of terrestrial microbial communities, with soil being 
regarded as their most characteristic habitat. Subaerial rock surfaces can be considered 
an inhospitable habitat for fungal (and other microbial) growth due to their high degree 
of insolation, desiccation and limited availability of nutrients (Gorbushina & Krum- 
bein, 2000). Micro-organisms that thrive under these extreme conditions have been 
termed 'poikilotrophic', i.e. able to deal with varying micro-climatic conditions such as 
light, salinity, pH and moisture. Microbial biofilms on and in rocks are believed to be 
major factors in rock decay and also in the formation of various patinas, films, 
varnishes, crusts and stromatolites in rock substrates (Gorbushina & Krumbein, 2000). 

Fungi have been reported from a wide range of rock types including limestone, 
soapstone, marble, granite, sandstone, andesite, basalt, gneiss, dolerite, amphibolite 
and quartz, from a variety of environments (Staley et al., 1982; Gorbushina et al., 1993; 

SGM symposium 65 



206 G. M. Gadd ; M. Fomina and E. P. Burford 

Sterflinger, 2000; Verrecchia, 2000; Burford et al., 2003a, b). It is likely that they are 
ubiquitous components of the microflora of all rocks and building stone, throughout 
a wide range of geographical and climatic zones. Despite the apparent inhospitality 
of the rock environment, the presence of organic and inorganic residues on mineral 
surfaces or within cracks is thought to encourage proliferation of fungi and other 
microbes. Waste products of algae and bacteria, dead cells, decaying plant material, 
dust particles, aerosols and animal faeces can all act as nutrient sources for fungi 
(Sterflinger, 2000). Some extremophilic fungi are especially evolved to survive and 
exploit microhabitats on and within mineral substrata, occurring within the lichen 
symbiosis or as free-living microcolonial fungi (Gorbushina et al., 1993; Bogomolova 
et al., 1998; Sterflinger, 2000). Microcolonial fungi include those black fungi that 
occur as spherical clusters of tightly packed thick-pigmented-walled cells or hyphae 
(Gorbushina et al., 1993; Bogomolova et al., 1998). As well as these, other filamentous 
fungi, including zygomycetes, ascomycetes and basidiomycetes, often occur on rock 
surfaces (epiliths) and in cracks, fissures and pores (endoliths). Certain fungi may also 
actively 'burrow' into rock substrates (cryptoendoliths). In addition, many deutero- 
mycetes (Fungi Imperfecti), which only exhibit asexual reproduction, are commonly 
found in mineral substrates (Kumar & Kumar, 1999; Sterflinger, 2000; Verrecchia, 2000). 

In soil, fungi generally comprise the largest pool of biomass (including other microbes 
and invertebrates) and this, combined with their filamentous growth habit, ensures that 
fungus-mineral interactions are an integral component of biogeochemical processes in 
the soil (Gadd, 1993, 1999, 2000a). They occur as free-living filamentous forms, plant 
symbionts, unicellular yeasts and animal and plant pathogens, and play an important 
role in carbon cycling and other biogeochemical cycles (Gadd & Sayer, 2000; Gadd, 
2005a). Their ability to translocate nutrients through the mycelial network also 
provides significant environmental advantages (Fomina et al., 2003; Jacobs et al., 2004). 
Mycorrhizal fungi in particular are one of the most important ecological groups of soil 
fungi in terms of mineral weathering and dissolution of insoluble minerals (Paris et al., 
1995; Jongmans et al., 1997; Lundstrom et al., 2000; Hoffland et al., 2002; Martino 
et al., 2003; Fomina et al., 2004, 2005c). 

MECHANISMS OF ROCK WEATHERING BY FUNGI 
Biomechanical deterioration 

Fungi are an important component of lithobiotic communities (an association of 
micro-organisms forming a biofilm at the mineral-microbe interface), where they 
interact with the lithic substrate, both geophysically and geochemically (de los Rios 
et al., 2002; Burford et al., 2003a, b). Biomechanical deterioration of rocks may occur 
through hyphal penetration (e.g. into decayed limestone) and by tunnelling into 

SGM symposium 65 



Mycotransformation of minerals 207 

otherwise intact mineral material (e.g. along crystal planes in calcitic and dolomitic 
rocks) (Kumar & Kumar, 1999; Sterflinger, 2000). Fungal hyphae can also exploit grain 
boundaries, cleavages and cracks to gain access to mineral surfaces (Adeyemi & Gadd, 
2005). In lichens, cleavage-bound mineral fragments as small as 5 \im in diameter can 
accumulate within the lower thallus (Banfield et aL, 1999). An important feature of 
hyphal growth is spatial exploration of the environment to locate and exploit new 
substrates (Boswell et aL, 2002, 2003; Jacobs et aL, 2002a). This is facilitated by a range 
of tropic responses that determine the direction of hyphal growth. Among the tropisms 
that may occur, thigmotropism or contact guidance is a well-known property of fungi 
that grow on and within solid substrates (Watts et aL, 1998). Hyphal growth can often 
be influenced by grooves, ridges and pores in solid substrate and is more prevalent in 
weakened mineral surfaces. Chemotropism, and other nutritional responses, may also 
be important in stress avoidance, such as that raised by toxic metals (Fomina et aL, 
2000a; G&ddetaL, 2001). 

The process of invasive hyphal growth due to turgor pressure inside hyphae allows 
fungi to acquire nutrients from many solid materials (Money, 2001). Highly pressurized 
hyphae can penetrate tougher substrates than those with lower pressures, and fungi that 
naturally invade hard materials generate extraordinarily high pressures (Money, 1999; 
Money & Howard, 1996). Melanin has also been implicated in the penetrative ability of 
plant-pathogenic fungi, as this black pigment facilitates the development of infection 
structures (Wheeler & Bell, 1988; Money & Howard, 1996). Rock-dwelling fungi are 
often melanized (Gorbushina et aL, 1993; Sterflinger, 2000). 

Biochemical deterioration 

Biochemical actions of fungi on rocks are believed to be more important than 
mechanical degradation (Kumar 6c Kumar, 1999). Fungi can solubilize minerals and 
metal compounds through several mechanisms, including acidolysis, complexolysis, 
redoxolysis and by mycelial metal accumulation (Burgstaller & Schinner, 1993). So- 
called 'heterotrophic leaching' by fungi primarily involves the first two mechanisms 
and occurs as a result of several processes, including proton efflux via the plasma 
membrane H + -ATPase and/or maintenance of charge balance during nutrient uptake, 
the production of siderophores [for Fe(III) mobilization] or as a result of respiratory 
C0 2 production. In many fungal strains, however, an important leaching mechanism 
occurs through the production of organic acids (e.g. oxalic and citric acid) (Adams 
et aL, 1992; Gadd, 1999, 2001a; Sayer & Gadd, 2001 ; Jarosz-Wilkolazka & Gadd, 2003 ; 
Fomina et aL, 2005a). In addition, fungi excrete many other metabolites with metal- 
complexing properties, e.g. amino acids and phenolic compounds (Manley & Evans, 
1986; Miiller et aL, 1995). Fungal carboxylic acids can play a significant role in the 
chemical attack of mineral surfaces, since the production of organic acids provides a 

SGM symposium 65 



208 G. M. Gadd ; M. Fomina and E. P. Burford 

source of protons for solubilization and metal-chelating anions which complex metal 
cations (Miiller et al., 1995; Gadd, 1999, 2001a; Sayer & Gadd, 2001; Fomina et al., 
2004). In one study on the effect of microscopic fungi on the mobility of copper, nickel 
and zinc compounds in polluted Al-Fe-humus podzols, it was found that the mobility 
of all studied metals increased under the impact of fungi and was predetermined mostly 
by the decomposition of soil organic matter (Bespalova et al., 2002). 

FUNGAL DETERIORATION OF ROCK AND BUILDING STONE 

Microbial attack on minerals may be specific and may depend on the groups of micro- 
organisms involved, e.g. some lichen hyphae overgrew augite and mica but avoided 
quartz (Aristovskaya, 1980). Substrate acidification by a free-living and different ecto- 
mycorrhizal species varied between species as well as in relation to the different 
minerals. Mycena galopus and Cortinarius glaucopus produced the highest acidity 
per unit biomass density, with higher substrate acidification resulting during growth 
on tricalcium phosphate (Rosling et al., 2004). 

In podzols, quartz and kaolin are usually overgrown by fungi and algae, with abundant 
fungal hyphae also being associated with apatite particles (Aristovskaya, 1980). It 
seems that alkaline (basic) rocks are generally more susceptible to fungal attack than 
acidic rocks (Eckhardt, 1985; Kumar &£ Kumar, 1999). Along with other organisms, 
fungi are believed to contribute to the weathering of silicate-bearing rocks, e.g. mica 
and orthoclase, and iron- and manganese-bearing minerals, e.g. biotite, olivine and 
pyroxene (Kumar & Kumar, 1999). Callot et al. (1987) showed that siderophore-pro- 
ducing fungi were able to pit and etch microfractures in samples of olivine and glasses 
under laboratory conditions. A polycarboxylate siderophore, rhizoferrin, showed the 
ability to bind Cr(III), Fe(III) and Al(III) (Pillichshammer et al., 1995). Fungi can also 
deteriorate natural glass and man-made antique and medieval glass (Krumbein et al., 
1991). Degradation of aluminosilicates and silicates is believed to occur as a result of 
the production of organic acids, inorganic acids, alkalis and complexing agents (Rossi 
& Ehrlich, 1990). It is also likely that C0 2 released during fungal respiration can 
enhance silicate degradation by carbonic acid attack (Sterflinger, 2000). Aspergillus 
niger can degrade olivine, dunite, serpentine, muscovite, feldspar, spodumene, kaolin 
and nepheline. Penicillium expansion can degrade basalt, while Tenicillium simplicis- 
sirnum and Scopulariopsis brevicaulis both release aluminium from aluminosilicates 
(Mehta et al., 1979; Rossi, 1979; Sterflinger, 2000). Piloderma was able to extract K and/ 
or Mg from biotite, microcline and chlorite to satisfy nutritional requirements. When 
grown under low-K conditions, the organism showed fibrillar growths, hyphal swellings 
and hyphae devoid of ornamentation, possibly indicating nutrient deficiency. 
Differences were found in growth rates, morphologies and the Mg content of hyphae 
grown with chlorite and biotite, suggesting that Mg was limiting to normal growth. 

SGM symposium 65 



Mycotransformation of minerals 209 

Energy-dispersive X-ray analysis indicated that Pilo derma extracted significantly more 
K from biotite than from microcline. The high Ca and O content of hyphal orna- 
mentation mainly resulted from calcium oxalate crystals (Glowa et al., 2003). 

In podzol E horizons under European coniferous forests, the weathering of horn- 
blendes, feldspars and granitic bedrock has been attributed to oxalic, citric, succinic, 
formic and malic acid excretion by saprotrophic and mycorrhizal hyphae. Ectomycor- 
rhizal fungi could form micropores (3-10 jxm) in weatherable minerals and hyphal tips 
could produce micro- to millimolar concentrations of these organic acids (Jongmans et 
al., 1997; van Breemen et al., 2000). In order to quantify the contribution of mineral 
tunnelling to the weathering of feldspars and ecosystem influx of Ca and K, surface 
soils of 11 podzols were studied by Smits et al. (2005). Tunnels were observed only 
in soils older than 1650 years, with the contribution of tunnelling to mineral weathering 
in the upper mineral soil being less than 1 %. Feldspar tunnelling corresponded to an 
average ecosystem influx of 04 g ha -1 year -1 for K and 0-2 g ha -1 year 1 for Ca over 5000 
years of soil development. These data indicate that the contribution of tunnelling to 
weathering is more important in older soils, but remains low (Smits et al., 2005). 

Fungal weathering of limestone, sandstone and marble is also known to occur (Kumar 
& Kumar, 1999; Ehrlich, 2002). In hot and cold deserts and semi-arid regions, clump- 
like colonies of epi- and endolithic darkly pigmented microcolonial fungi are common 
inhabitants of limestone, sandstone, marble and granite, as well as other rock types 
(Staley et al., 1982; Sterflinger, 2000; Gorbushina et al., 1993). Analysis of desert rock 
samples has shown colonies or single cells in connection with pitting and etching 
patterns, suggesting acid attack of the mineral surface, possibly a result of organic acid 
or carbonic acid production (Sterflinger, 2000). Microcolonial fungi have also been 
shown to be common inhabitants of biogenic oxalate crusts on granitic rocks (Blazquez 
etal.,1997). 

Acidolysis, complexolysis and metal accumulation were involved in solubilization of 
zinc phosphate and pyromorphite by a selection of soil fungi representing ericoid and 
ectomycorrhizal plant symbionts and an endophytic/entomopathogenic fungus, 
Beauveria caledonica. Acidolysis (protonation) was found to be the major mechanism 
of both zinc phosphate and pyromorphite dissolution for most of the fungi examined 
and, in general, the more metal-tolerant fungal strains yielded more biomass, acidified 
the medium more and dissolved more of the metal mineral than less-tolerant strains. 
However, Beauveria caledonica excreted a substantial amount of oxalic acid (up to 
0-8 raM) in the presence of pyromorphite that coincided with a dramatic increase in 
lead mobilization, providing a clear example of complexolysis (Fig. 2) (Fomina et al., 
2004,2005a). 

SGM symposium 65 



210 G. M. Gadd, M. Fomina and E. P. Burford 



25 



20 - 



15 -- 



10 -- 



5 -- 







m 
U 



U 

E 
O 



* 




Z 

en 




0.8 mM oxalic acid 



CO 






r"i 



in p. 






KD 



00 
■M 

3 



u 



Fig. 2. Importance of oxalic acid in fungal solubilization of pyromorphite. The mobilization index 
(Ml), which represents the ratio of solubilized lead to the initial amount of pyromorphite, is shown 
for several fungal strains and an abiotic control. Bars represent sem. Fungal cultures are identified as: 
DGC3, Hymenoscyphus ericae DGC3(UZ); OmCd, Oidiodendron maius Cd; Bc4, Beauveria caledonica 
4; LI8, Laccaria laccata 8; Pi23, Paxil lus involutus 23; Pi1 5, P. involutus 1 5; SI21 , Suillus luteusll; SI33, 
S. luteus 33; MG1 , Suillus bovinus MG1 ; LSt8, 5. bovinus LSt8; Tt, Telephora terrestris (adapted from 
Fomina et al, 2004). 



Calcium carbonate (CaC0 3 ) and calcium magnesium carbonate [CaMg(C0 3 ) 2 ] occur 
extensively on the Earth's surface as limestone and dolomite (Ehrlich, 2002). Near- 
surface calcretes and dolocretes cover as much as 13 % of the total land surface and are 
an important reservoir of carbon, accounting for almost 80 % of the total HCO3, 
CO? - and C0 2 in the Earth's lithosphere (Ehrlich, 2002; Goudie, 1996). Numerous 
micro-organisms, including bacteria and fungi, have been isolated from natural lime- 
stone formations. Cryptoendolithic (i.e. actively penetrating the rock matrix to several 
millimetres in depth) and chasmolithic or endolithic (i.e. living in hollows, cracks and 
fissures) fungi are known to occur in limestone. The production of organic acids is 
believed to play a major role in degradation of limestone (Ehrlich, 2002). 



The chemical basis for carbonate weathering is the instability of carbonates in acid 
solution: 



CaC0 3 + H + - Ca 2+ + HCO3 
HC03 + H + -H 2 C0 3 
H 2 C0 3 - H 2 + C0 2 



(equation 1) 
(equation 2) 
(equation 3) 



SGM symposium 65 



Mycotransformation of minerals 211 

Since Ca(HC0 3 ) 2 is very soluble compared with CaC0 3 , the CaC0 3 dissolves even in 
weakly acidic solutions. In strong acid solutions, the CaC0 3 dissolves more rapidly as 
carbonate is lost from the solution as C0 2 . Any organism capable of producing acidic 
metabolites is capable of dissolving carbonates and even the production of respiratory 
C0 2 during respiration can have the same effect: 

CO 2 + H 2 ~ H 2 C0 3 (equation 4) 

H 2 C0 3 + CaC0 3 «* Ca 2+ + 2HC0 3 (equation 5) 

Degradation of sandstone by fungi is also well documented and this has been attributed 
to the production of for example acetic, oxalic, citric, formic, fumaric, glyoxylic, 
gluconic, succinic and tartaric acids (Gomez-Alarcon et ai, 1994; Hirsch et al., 1995; 
Sterflinger, 2000). Fungi can also attack rock surfaces through redox attack of mineral 
constituents such as Mn and Fe (Timonin et ai, 1972; Grote & Krumbein, 1992; de la 
Torre & Gomez-Alarcon, 1994). 

METAL BINDING AND ACCUMULATION BY FUNGI 

Fungal biomass provides a sink for metals, by either (i) metal biosorption to biomass 
(cell walls, pigments and extracellular polysaccharides), (ii) intracellular accumulation 
and sequestration or (iii) precipitation of metal compounds onto hyphae. As well as 
immobilizing metals, the reduction in external free metal activity may shift the equili- 
brium so that more metal is released into the soil solution (Gadd, 1993, 2000b; Sterf- 
linger, 2000). Ectomycorrhizal fungi can release elements from apatite and wood ash and 
accumulate them in the mycelia. Suillus granulatus contained 3-15 times more K (3 mg 
g _1 ) and showed large calcium-rich crystals on rhizomorphs when grown with apatite, 
while Paxillus involutus had the largest amount of Ca (2-7 mg g _1 ). Wood ash addition 
increased the amounts of Ti, Mn and Pb in rhizomorphs (Wallander et ai, 2003). 

Fungi are effective biosorbents for a variety of metals, including Ni, Zn, Ag, Cu, Cd 
and Pb (Gadd, 1990, 1993). Metal binding by fungi can occur through metabolism- 
dependent or metabolism-independent binding of ions onto cell walls and other 
external surfaces and can be an important passive process in both living and dead 
fungal biomass (Gadd, 1990, 1993; Sterflinger, 2000). Metal-binding capacity can be 
influenced by external pH, with binding capacity decreasing at low pH for metals such 
as Cu, Zn and Cd (de Rome & Gadd, 1987). Cell density also effects binding capacity, 
with lower cell densities allowing a higher specific metal uptake per unit of biomass 
(Gadd, 1993). The presence of melanin in fungal cell walls also strongly influences 
biosorptive capacity (Gadd & Mowll, 1995; Manoli et ai, 1997). Metal localization 
was investigated in the lichenized ascomycete Trapelia involuta growing on a range 
of uraniferous minerals, including metazeunerite [Cu(U0 2 ) 2 (As0 4 ) 2 . 8H 2 0], meta- 

SGM symposium 65 



212 G. M. Gadd, M. Fomina and E. P. Burford 




Fig. 3. Formation of metal oxalate crystals by Beauveria caledonica. (a) Dry-mode environmental 
scanning electron microscope (ESEM) image of calcium oxalate (weddelite and whewellite) crystals 
produced on agar containing calcium carbonate (M. Fomina and G. M. Gadd, unpublished); (b) wet- 
mode ESEM image of zinc oxalate dihydrate formed by the mycelium grown on zinc phosphate and 

torbernite [Cu(U0 2 ) 2 (P0 4 ) 2 . 8H 2 0], autunite [Ca(U0 2 ) 2 (P0 4 ) 2 . 10H 2 O] and uranium- 
enriched iron oxide and hydroxide minerals. The highest U, Fe and Cu concentrations 
occurred in the outer parts of melanized apothecia, indicating that metal biosorption 
by melanin-like pigments was likely to be responsible for the observed metal fixation 
(Purvis etal y 2004). 

MYCOGENIC MINERAL FORMATION 

Fungi have been shown to precipitate a number of inorganic and organic mineral 
compounds, e.g. oxalates, carbonates and oxides (Arnott, 1995; Verrecchia, 2000; 
Gadd, 2000a; Grote 6c Krumbein, 1992). This process may be important in soil, as 
precipitation of carbonates, phosphates and hydroxides increases soil aggregation. 



SGM symposium 65 



Mycotransformation of minerals 213 




Q 



covered with a mucilaginous hyphal network; (c) light microscopy images of zinc phosphate particles 
adsorbed by the fungal mycelium after 7 days of growth at 25 °C in static liquid medium; (d) crystals 
of zinc oxalate dihydrate in the same mycelium after 1 days of growth (adapted from Fomina etal., 
2005a). Bars, 20 urn (a, c, d) and 50 urn (b). 

Cations like Si 4+ , Fe 3+ , Al 3+ and Ca 2+ (that may be released through dissolutive mech- 
anisms) stimulate precipitation of such compounds, which act as bonding agents for 
soil particles. Roots and hyphae can enmesh particles together, alter alignment and 
release organic metabolites that assist aggregate stability (Bronick & Lai, 2005). 

Oxalate precipitation 

Calcium oxalate is the most common form of oxalate associated with soils and leaf 
litter, occurring as the dihydrate (weddelite) or the more stable monohydrate (whewel- 
lite). Calcium oxalate crystals are commonly associated with free-living, pathogenic 
and plant-symbiotic fungi and are formed by the reprecipitation of solubilized calcium 
as calcium oxalate (Fig. 3) (Arnott, 1995). Fungal-derived calcium oxalate can exhibit a 



SGM symposium 65 



214 G. M. Gadd, M. Fomina and E. P. Burford 

variety of crystalline forms (tetragonal, bipyramidal, plate-like, rhombohedral or 
needles). The formation of calcium oxalate by fungi is important geo chemically, since 
it acts as a reservoir for calcium but also influences phosphate availability (Gadd, 1993, 
1999; Jacobs e**/., 2002a, b). 

Fungi can produce other metal oxalates with a variety of different metals and metal- 
bearing minerals, e.g. Cd, Co, Cu, Mn, Pb, Sr and Zn (Fig. 3) (Gadd, 2000a, b; Jarosz- 
Wilkolazka & Gadd, 2003; Burford et al., 2003b; Fomina et al., 2005a). Formation of 
metal oxalates may provide a mechanism whereby fungi can tolerate environments 
containing potentially high concentrations of toxic metals. A similar mechanism occurs 
in lichens growing on copper-sulfide-bearing rocks, where precipitation of copper 
oxalate occurs within the thallus (Arnott, 1995; Easton, 1997). 

Lichens are one of the most common members of the microbial consortia inhabiting 
exposed subaerial rock substrates and building stone. Oxalic acid excretion by lichens 
can result in the dissolution of insoluble carbonates and silicates and formation 
of water-soluble and -insoluble oxalates (Jones & Wilson, 1985; May et al., 1993; 
Edwards et al., 1991 ; Verrecchia, 2000). Oxalates can be formed with a variety of metal 
ions wherever lichens grow (Purvis, 1996). In particular, there has been concern over the 
deteriorative effects of biologically formed oxalic acid on architecturally important 
buildings, monuments and frescoes (Nimis et al., 1992). Conversely, however, several 
studies have suggested that lichen cover protects certain rock surfaces, acting as an 
'umbrella' and reducing the erosion of for example slightly soluble calcium sulfates 
(Mottershead & Lucas, 2000). 

Carbonate precipitation: calcified fungal filaments in 
limestone and calcareous soils 

Physico-chemical processes are known to play a significant role in calcrete formation 
and development, although it is now recognized that micro-organisms, particularly 
bacteria and algae, may also play a crucial and even dominant role in calcrete trans- 
formation (Goudie, 1996) . In limestone, fungi and lichens are generally considered to be 
important agents of carbonate mineral deterioration. A less studied area of fungal 
action, however, is their influence on carbonate precipitation (Goudie, 1996; Sterflinger, 
2000). It is already known that many near-surface limestones (calcretes), calcic and 
petrocalcic horizons in soils are often secondarily cemented with calcite (CaCO^) 
and whewellite (calcium oxalate monohydrate, CaC 2 4 .H 2 0) (Verrecchia, 2000). 
Although this phenomenon has partly been attributed to physico-chemical processes, 
the presence of calcified fungal filaments in limestone and calcareous soils from a range 
of localities confirms that fungi may play a prominent role in secondary calcite 
precipitation. Fungal filaments mineralized with calcite, together with whewellite, have 

SGM symposium 65 



Mycotransformation of minerals 215 

been reported in limestone and calcareous soils from a range of localities (Kahle, 1977; 
Klappa, 1979; Calvet, 1982; Callot et al, 1985a, b; Verrecchia et al, 1993; Monger St 
Adams, 1996; Bruand & Duval, 1999; Verrecchia, 2000). It has also been proposed that 
calcium oxalate can be degraded to calcium carbonate, e.g. in semi-arid environments, 
where such a process may again act in the cementation of pre-existing limestones 
(Verrecchia etal, 1990). 

Calcite formation by fungi may occur through indirect processes via the fungal 
excretion of oxalic acid and the precipitation of calcium oxalate (Verrecchia et al, 
1990; Gadd, 1999; Verrecchia, 2000). For example, oxalic acid excretion and the form- 
ation of calcium oxalate results in the dissolution of the internal pore walls of the 
limestone matrix, so that the solution becomes enriched in free carbonate. During 
passage of the solution through the pore walls, calcium carbonate recrystallizes as 
a result of a decrease in C0 2 , and this contributes to hardening of the material. 
Biodegradation of oxalate as a result of microbial activity can also lead to transform- 
ation into carbonate, resulting in precipitation of calcite in the pore interior, leading to 
closure of the pore system and hardening of the chalky parent material. During 
decomposition of fungal hyphae, calcite crystals can act as sites of further secondary 
calcite precipitation (Verrecchia, 2000). Manoli et al. (1997) demonstrated that chitin, 
the major component of fungal cell walls, is a substrate on which calcite will readily 
nucleate and subsequently grow. 

Reduction or oxidation of metals and metalloids 

Many fungi precipitate reduced forms of metals and metalloids in and around fungal 
hyphae: for example, Ag(I) reduction to elemental silver [Ag(0)], selenate [Se(VI)] and 
selenite [Se(IV)] to elemental selenium [Se(0)] and tellurite [Te(IV)] to elemental 
tellurium [Te(0)]. Reduction of Hg(II) to volatile Hg(0) can also be mediated by fungi 
(Gadd, 1993, 2000a, b). An Aspergillus sp., P37, was able to grow at arsenate con- 
centrations of 0-2 M (more than 20-fold higher than that withstood by Escherichia 
coli, Saccharomyces cerevisiae and Aspergillus nidulans), and it was suggested that 
increased arsenate reduction contributed to the hypertolerant phenotype of this fungus 
(Canovas et al., 2003a, b). 

Desert varnish, an oxidized metal layer (patina) a few millimetres thick found on rocks 
and in soils of arid and semi-arid regions, is believed to be of fungal and bacterial 
origin. Lichenothelia spp. can oxidize manganese and iron in metal-bearing minerals, 
such as siderite (FeC0 3 ) and rhodochrosite (MnC0 3 ), and precipitate them as oxides 
(Grote & Krumbein, 1992). Similar oxidation of Fe(II) and Mn(II) by fungi leads to the 
formation of dark patinas on glass surfaces (Erkhardt, 1985). A Mn-depositing fungus, 
identified as an Acremonium-like hyphomycete, was isolated from a variety of labora- 

SGM symposium 65 



216 G. M. Gadd ; M. Fomina and E. P. Burford 

tory and natural locations including Mn(III,iV)-oxide-coated stream-bed pebbles. A 
proposed role for a laccase-like multicopper oxidase was postulated, analogous to the 
Mn (II) -oxidizing factors found in certain bacteria (Miyata et al., 2004). 

Other minerals associated with fungal communities 

A wide range of minerals may be deposited under conditions that deviate strongly from 
'normal' pressure/temperature diagrams of precipitation and stability and have been 
found in association with poikilophilic communities on rock surfaces (Gorbushina 
etal.,2002). Some evaporite minerals (gypsum, CaS0 4 .2H 7 0) and iron oxides (magne- 
tite, Fe 3 4 ) also display distinctive morphologies indicative of the presence of microbial 
communities (Gorbushina et al., 2002). Another biogenic mineral (tepius) has been 
identified in association with a lichen carpet that covers high mountain ranges in 
Venezuela (Gorbushina et al., 2002). Forsterite (Mg 2 Si0 4 ), the magnesium member 
of the olivine [(Mg,Fe) 2 Si0 4 ] mineral solid solution series, is known to occur only in 
volcanic rocks, meteorites and metamorphosized carbonates (e.g. skarn deposits). The 
presence of forsterite in surficial deposits on rock surfaces can therefore be considered 
to be a possible biosignature for former or extant life (Gorbushina et al., 2002). 

FUNGAL-CLAY INTERACTIONS 

Clay mineral formation and impact on soil properties 

Silicon dioxide, when combined with oxides of Mg, Al, Ca and Fe, forms the silicate 
minerals in rocks and soil (Bergna, 1994). These high-temperature minerals are unstable 
in the biosphere and break down readily to form clays. Micro-organisms play a funda- 
mental role in the dissolution of silicate structure in rock weathering, and therefore in 
the genesis of clay minerals, and soil and sediment formation (Banfield et al., 1999). In 
fact, the presence of clay minerals can be a typical symptom of biogeochemically 
weathered rocks (Barker & Banfield, 1996, 1998; Rodriguez Navarro et al., 1997). For 
example, in lichen weathering of silicate minerals, Ca, K, Fe clay minerals and nano- 
crystalline aluminous Fe oxyhydroxides were mixed with fungal organic polymers 
(Barker & Banfield, 1998), while biotite was interpenetrated by fungal hyphae growing 
along cleavages and partially converted to vermiculite (Barker & Banfield, 1996). Other 
studies have shown that transformation of mica and chlorite to 2 : 1 expandable clays 
was predominant in the ectomycorrhizosphere compared with non-ectomycor- 
rhizosphere soils, likely to be a result of the high production of organic acids and direct 
extraction of K + and Mg 2+ by fungal hyphae (Arocena et al., 1999). 

Soil, which can be considered to be a biologically active loose mass of weathered rock 
fragments mixed with organic matter, is the ultimate product of rock weathering, 
i.e. the interaction between the biota, climate and rocks. Clay minerals are generally 

SGM symposium 65 



Mycotransformation of minerals 217 

present in soil in larger amounts than organic matter and, because of their ion- 
exchange capacity, charge and adsorption powers, they perform a significant buffering 
function in mineral soils (Ehrlich, 2002) and are important reservoirs of cations and 
organic molecules (Wild, 1993; Li & Li, 2000; Dinelli & Tateo, 2001; Dong et aL, 2001; 
Krumhansl et aL, 2001). 

Biological effects of clay minerals 

Fungi are in close contact with clay minerals in soils and sediments. Numerous studies 
have shown that interactions of micro-organisms with solid adsorbents lead to 
increases in biomass, growth rate and the production of enzymes and metabolites 
(Stotzky, 1966, 2000; Martin et aL, 1976; Fletcher, 1987; Marshall, 1988; Claus & Filip, 
1990; Lee & Stotzky, 1999; Lotareva & Prozorov, 2000; Lunsdorf et aL, 2000; Deman- 
eche et aL, 2001; Fomina & Gadd, 2003). Some clays may stimulate or inhibit fungal 
metabolism (Fomina et aL, 2000b; Fomina & Gadd, 2003). Stimulatory effects may 
arise from the abilities of different clays to serve as (i) pH buffers, (ii) a source of metal 
cationic nutrients, (iii) specific adsorbents of metabolic inhibitors, other nutrients and 
growth stimulators and (iv) modifiers of the microbial microenvironment because of 
their physico-chemical properties such as surface area and adsorptive capacity (Stotzky, 
1966; Marshall, 1988; Martin et aL, 1976; Fletcher, 1987). It has also been shown that 
clay minerals (bentonite, palygorskite and kaolinite) influence the size, shape and 
structure of mycelial pellets in liquid medium (Fig. 4) (Fomina &C Gadd, 2002). 

Fungal-clay mineral interactions in soil aggregation 

Fungal-clay mineral interactions play an important role in soil evolution. An examina- 
tion of soil clay aggregation by saprophytic {Rbizoctonia solani and Hyalodendron sp.) 
and mycorrhizal {Hymenoscyphus ericae and Hebeloma sp.) fungi supported the 
hypothesis that fungal hyphae bring mineral particles and organic materials together to 
form stable microaggregates at least <2 ^im in size and enmesh such microaggregates 
into stable aggregates of >50 [im in diameter (Tisdall et aL, 1997). Fungi not only 
entrap soil particles in their hyphae but take part in polysaccharide aggregation as well 
(Dorioz et aL, 1993; Martens 5t Frankenberger, 1992; Schlecht-Pietsch et aL, 1994; 
Puget et aL, 1999; Chantigny et aL, 1997). Only a few studies have been carried out on 
the sorption properties of mixtures of clay minerals (montmorillonite, kaolinite) and 
microbial biomass (algae, fungi): interactions between clay minerals and fungi alter the 
adsorptive properties of both clays and hyphae (Garnham et aL, 1991; Kadoshnikov 
et aL, 1995; Morley & Gadd, 1995; Fomina & Gadd, 2002, 2003). 

Clay and silicate weathering by fungi 

Fungi and bacteria play an important role in the mobilization of silica and silicates 
(Ehrlich, 2002). Their action is mainly indirect, either through the production of 

SGM symposium 65 



218 G. M. Gadd, M. Fomina and E. P. Burford 









Fig. 4. Scanning electron micrographs of Cladosporium dadosporioides grown in the presence of 
different clay minerals. (aHc) Mycelial pellets resulting from growth in medium containing 0-5 % (w/v) 
(a) and 5 % (w/v) (b) palygorskite or 0-5 % (w/v) kaolinite (c). (d)-(f) Internal structure of the central 
zone of fractured pellets grown in clay-free control medium (d) and in the presence of 0-5 % (w/v) (e) 
and 5 % (w/v) (f) bentonite. Bars, 1 00 urn (a, b, c) and 10 urn (d, e, f). Reproduced from Fomina & 
Gadd (2002). 

chelates or the production of acids (mineral or organic) or, as for certain bacteria, the 
production of ammonia or amines. Fungi isolated from weathered rock surfaces 
{Botrytis, Mucor, Tenicillium and Tricboderma spp.) were shown to be able to solu- 
bilize calcium, magnesium and zinc silicates (Webley et aL, 1963). Mobilization of 
silicate from clay minerals by Aspergillus niger was found to be a result of oxalic acid 
excretion (Henderson & Duff, 1963). The majority of fungal strains belonging to the 
genera Aspergillus, Paecilomyces, Penicillium, Scopulariopsis and Tricboderma could 
leach iron in submerged culture from a China clay sample (Mandal et aL, 2002). Large 
amounts of oxalic, citric and gluconic acids were produced by Penicillium frequentans 
in liquid culture. This caused extensive deterioration of clay silicates, as well as micas 
and feldspars, from sandstone and granite as a result of organic salts formation such as 
calcium, magnesium and ferric oxalates and calcium citrates (de la Torre et aL, 1993). 
The oxalate-excreting fungus Hysterangium crassum also weathered clay minerals in 
situ (Cromack et aL, 1979). 



SGM symposium 65 



Mycotransformation of minerals 219 




Fig. 5. Deterioration of a concrete block that was exposed to fungal weathering for 2 years in an 
experimental microcosm (M. Fomina and G. M. Gadd, unpublished). Cracking is evident, as well as an 
extensive hyphal net over the surface. Bar, 200 urn. 



MINERAL MYCOTRANSFORMATIONS AND ENVIRONMENTAL 
BIOTECHNOLOGY 

Concrete biodeterioration and radioactive waste disposal 

Fungi play an important role in the deterioration of concrete (Fig. 5) (Perfettini et al., 
1991; Nica et al., 2000), with complexolysis suggested as the main mechanism of 
calcium mobilization (Gu et al., 1998). This ability of fungi (and other microbes) to 
degrade concrete and other structural materials has implications for underground 
storage of nuclear waste. In high-level nuclear waste disposal, the bentonite buffer 
around the copper canisters is considered to be a hostile environment for most microbes 
because of the combination of radiation, heat and low water availability, but discrete 
microbial species can cope with each of these constraints (Pedersen, 1999). Endolithic, 
indigenous micro-organisms are capable of surviving gamma-irradiation doses 
simulating the near-field environment surrounding waste canisters (Pitonzo et al., 
1999). In 1997-1998, extensive fungal growth was observed on the walls and other 
building constructions in the inner part of the 'Shelter' built on the fourth unit of the 
Chernobyl nuclear power plant damaged in 1986 (Zhdanova et al., 2000). It was 
discovered that low-level gamma-radiation did not affect spore germination, but led 
to directed growth of fungal tips towards the radiation source (so-called positive 
radiotropism) (Zhdanova et al., 2001). 



SGM symposium 65 



220 G. M. Gadd ; M. Fomina and E. P. Burford 

Bioremediation 

Some of the processes detailed previously have the potential for treatment of con- 
taminated land (Gadd, 2000a, b, 2001a, b, 2002, 2004; Hochella, 2002; Fomina et aL, 
2005b). Solubilization processes provide a route for removal of metals from soil mat- 
rices, whereas immobilization processes enable metals to be transformed into insoluble, 
chemically more inert forms. Living or dead fungal biomass and fungal metabolites 
have been used to remove metal or metalloid species, compounds and particulates and 
organometal(loid) compounds from solution by biosorption. Fungi with Cr(VI)- 
reducing activity may have potential for treatment of Cr-polluted soils (Cervantes et aL, 
2001). Fungal-clay complex biomineral sorbents may combine the sorptive advantages 
of the individual counterparts, i.e. the high density of metal-binding sites per unit area 
and high sorption affinity and capacity of fungal biomass, and the high surface area per 
unit weight, mechanical strength and efficient sorption at high metal concentrations of 
the clay minerals (Fomina & Gadd, 2003). There has also been the use of extracellular 
ligands excreted by fungi, especially from Aspergillus and Penicillium spp., to leach 
metals such as Zn, Cu, Ni and Co from a variety of solid materials, including low-grade 
mineral ores (Brandl, 2001). Mycorrhizal associations may also be used with plants for 
metal clean-up in the general area of phytoremediation. Phy to extraction involves the 
use of plants to remove toxic metals from soil by accumulation in above-ground parts. 
Mycorrhiza may enhance phytoextraction directly or indirectly by increasing plant 
biomass, and some studies have shown increased plant accumulation of metals, 
especially when inoculated with mycorrhiza isolated from metalliferous environments. 
However, this is a simplistic hypothesis and many complicating factors affect successful 
exploitation (Meharg, 2003). Several other studies have shown reduced plant metal 
uptake (Tullio et aL, 2003). Arbuscular mycorrhiza can depress translocation of zinc to 
shoots of host plants in soils moderately polluted with zinc, with binding of metals in 
mycorrhizal structures and immobilization of metals in the mycorrhizosphere contri- 
buting to these effects (Christie et aL, 2004). Ectomycorrhizal fungi persistently fixed 
Cd(II) and Pb(II), and formed an efficient biological barrier that reduced movement of 
these metals in birch tissues (Krupa & Kozdroj, 2004). Such mycorrhizal metal 
immobilization around plant roots, including biomineral formation, may also assist 
soil remediation and revegetation. The development of stress-tolerant plant-mycor- 
rhizal associations may therefore be a promising new strategy for phytoremediation 
and soil amelioration (Schutzendubel & Polle, 2002). Because of the symbiosis with 
ericoid mycorrhizal fungi, ericaceous plants are able to grow in highly polluted 
environments, where metal ions can reach toxic levels in the soil substrate (Perotto et aL, 
2002; Martino et aL, 2003). Ericoid mycorrhizal fungal endophytes, and sometimes 
their plant hosts, can evolve toxic-metal resistance which enables ericoid mycorrhizal 
plants to colonize polluted soil. This seems to be a major factor in the success of ericoid 
mycorrhizal taxa in a range of harsh environments (Cairney & Meharg, 2003). 

SGM symposium 65 



Mycotransformation of minerals 221 

CONCLUSIONS 

It is clear that fungi have important biogeochemical roles in the biosphere. Symbiotic 
mycorrhizal fungi are responsible for major mineral transformations and redistribution 
of inorganic nutrients, for example essential metals and phosphate, as well as carbon 
flow, while free-living fungi have major roles in the decomposition of plant and other 
organic materials, including xenobiotics, as well as for example phosphate solubili- 
zation. Fungi are dominant members of the soil microflora, especially in acidic 
environments, and may operate over a wider pH range than most heterotrophic 
bacteria. Fungi are also major agents of biodeterioration of stone, wood, plaster, 
cement and other building materials, and are important components, including lichens, 
of rock-inhabiting microbial communities, with significant roles in mineral dissolution 
and secondary mineral formation. It is timely to draw attention to 'geomycology' 
within the umbrella of geomicrobiology and to the interdisciplinary approach that is 
necessary to further understanding of the important roles that all micro-organisms 
play in the biogeochemical cycling of elements, the chemical and biological mech- 
anisms that are involved and their environmental and biotechnological significance. 

ACKNOWLEDGEMENTS 

Some of the authors' research described within was funded by the BBSRC/BIRE programme 
(94/BRE13640) and CLRC Daresbury SRS (SRS grant 40107), which is gratefully acknowledged. 
E.RB. gratefully acknowledges receipt of an NERC post-graduate research studentship. 
Thanks are also due to Mr Martin Kierans (Centre for High Resolution Imaging and Processing, 
School of Life Sciences, University of Dundee, Scotland) for assistance with scanning electron 
microscopy. 

REFERENCES 

Adams, J. B., Palmer, F. & Staley, J. T. (1992). Rock weathering in deserts- mobilization 
and concentration of ferric iron by microorganisms. GeomicrobiolJ 10, 99-1 14. 

Adeyemi, A. O. & Gadd, G. M. (2005). Fungal degradation of calcium-, lead- and silicon- 
bearing minerals. Biometals 18, 269-281 . 

Aristovskaya, T. V. (1980). Microbiology of the Processes of Soil Formation. Leningrad: 
Nauka USSR (in Russian). 

Arnott, H. J. (1995). Calcium oxalate in fungi. In Calcium Oxalate in Biological Systems, 
pp. 73-1 1 1 . Edited by S. R. Khan. Boca Raton, FL: CRC Press. 

Arocena, J. M., Glowa, K. R., Massicotte, H. B. & Lavkulich, L. (1999). Chemical and 
mineral composition of ectomycorrhizosphere soils of subalpine fir (Abies lasiocarpa 
(Hook.) Nutt.) in the Ae horizon of a luvisol. Can J Soil Sci 79, 25-35. 

Banfield, J. F. & Nealson, K. H. (editors) (1998). Geomicrobiology: Interactions between 
Microbes and Minerals. Reviews in Mineralogy, vol. 35. Washington, DC: Mineral- 
ogical Society of America. 

Banfield, J. P., Barker, W. W., Welch, S. A. & Taunton, A. (1999). Biological impact on 

SGM symposium 65 



222 G. M. Gadd, M. Fomina and E. P. Burford 



mineral dissolution: application of the lichen model to understanding mineral 

weathering in the rhizosphere. Proc Natl Acad Sci U 5 A, 96, 3404-341 1 . 
Banfield, J. F, Welch, S. A., Zhang, H., Ebert, T. T. & Penn, R. L (2000). Aggregation- 
based crystal growth and microstructure development in natural iron oxyhydroxide 

biomineralization products. Science 289, 751-754. 
Barker, W. W. & Banfield, J. F. (1996). Biologically versus inorganically mediated 

weathering reactions: relationships between minerals and extracellular microbial 

polymers in lithobiotic communities. Chem Geol 132, 55-69. 
Barker, W. W. & Banfield, J. F. (1998). Zones of chemical and physical interaction at 

interfaces between microbial communities and minerals: a model. Geomicrobiol J 

15,223-244. 
Bennett, P. C, Hiebert, F. K. & Choi, W. J. (1996). Microbial colonization and weathering 

of silicates in petroleum-contaminated groundwater. Chem Geol 132, 45-53. 
Bergna, H. E. (1994). Colloid chemistry of silica - an overview. In The Colloid Chemistry 

of Silica, Advances in Chemistry Series vol. 234, pp. 1-47. Edited by H. E. Bergna. 

Washington, DC: American Chemical Society. 
Bespalova, A. Yu., Marfenina, O. E. & Motuzova, G. V. (2002). The effect of microscopic 

fungi on the mobility of copper, nickel, and zinc in polluted AI-Fe-humus podzols of 

the Kola Peninsula. Eurasian Soil Sci 35, 945-950. 
Beveridge, T. J., Pouwels, P. H., Sara, M. & 22 other authors (1997). Function of S-layers. 

FEMS Microbiol Rev 20, 99-1 49. 
Blazquez, F., Garcia-Vallez, M., Krumbein, W. E., Sterflinger, K. & Vendrell-Saz, M. 

(1997). Microstromatolithic deposits on granitic monuments: development and 

decay. Eur J Mineral 9, 889-901 . 
Bogomolova, E. V., Vlasov, D. Yu. & Panina, L. K. (1998). On the nature of the 

microcolonial morphology of epilithic black yeasts Phaeococcomyces de Hoog. Dokl 

Ross Akad Sci 363, 707-709. 
Boswell, G. P., Jacobs, H., Davidson, F. A., Gadd, G. M. & Ritz, K. (2002). Functional 

consequences of nutrient translocation in mycelial fungi. J Theor Biol 217, 459- 

477. 
Boswell, G. P., Jacobs, H., Davidson, F. A., Gadd, G. M. & Ritz, K. (2003). Growth and 

function of fungal mycelia in heterogeneous environments. Bull Math Biol 65, 

447-477. 
Brandl, H. (2001). Heterotrophic leaching. In Fungi in Bioremediation, pp. 383-423. Edited 

by G. M. Gadd. Cambridge: Cambridge University Press. 
Bronick, C. J. & Lai, R. (2005). Soil structure and management: a review. Geoderma 124, 

3-22. 
Bruand, A. & Duval, O. (1999). Calcified fungal filaments in the petrocalcic horizon of 

eutrochrepts in Beauce, France. Soil Sci Soc Am J 63, 1 64-1 69. 
Burford, E. P., Kierans, M. & Gadd, G. M. (2003a). Geomycology: fungal growth in 

mineral substrata. Mycologist 17 , 98-107. 
Burford, E. P., Fomina, M. & Gadd, G. M. (2003b). Fungal involvement in bioweathering 

and biotransformation of rocks and minerals. Mineral Mag 67 , 1127-1155. 
Burgstaller, W. & Schinner, F. (1993). Leaching of metals with fungi. J Biotechnol 27, 

91-116. 
Cairney, J. W. G. & Meharg, A. A. (2003). Ericoid mycorrhiza: a partnership that exploits 

harsh edaphic conditions. Eur J Soil Sci 54, 735-740. 
Callot, G., Mousain, D. & Plassard, C. (1985a). Concentrations de carbonate de calcium 

sur les parois des hyphes myceliens. Agronomie 5, 143-1 50 (in French). 



SGM symposium 65 



Mycotransformation of minerals 223 



Callot, G. f Guyon, A. & Mousain, D. (1985b). Inter-relation entre aiguilles de calcite 
et hyphes myceliens. Agronomie 5, 209-216 (in French). 

Callot, G., Maurette, M., Pottier, L. & Dubois, A. (1987). Biogenic etching of micro- 
fractures in amorphous and crystalline silicates. Nature 328, 147-149. 

Calvet, F. (1982). Constructive micrite envelope developed in vadose continental environ- 
ment in pleistocene eoliantesof Mallorca (Spain). Acta Geol Hisp 17, 169-178. 

Canovas, D., Duran, C, Rodriguez, N., Amils, R. & de Lorenzo, V. (2003a). Testing the 
limits of biological tolerance to arsenic in a fungus isolated from the River Tinto. 
Environ Microbiol 5, 1 33-1 38. 

Canovas, D., Mukhopadhyay, R., Rosen, B. P. & de Lorenzo, V. (2003b). Arsenate 
transport and reduction in the hyper-tolerant fungus Aspergillus sp. P37. Environ 
Microbiol 5, 1087-1093. 

Cervantes, C, Campos-Garcia, J., Devars, S., Gutierrez-Corona, F. G., Loza-Tavera, H., 
Torres-Guzman, J. C. & Moreno-Sanchez, R. (2001). Interactions of chromium 
with microorganisms and plants. FEMS Microbiol Rev 25, 335-347. 

Chantigny, M. H., Angers, D. A., Prevost, D., Vezina, L P. & Chalifour, F. P. (1997). Soil 
aggregation and fungal and bacterial biomass under annual and perennial cropping 
systems. Soil Sci Soc Am J S*\ , 262-267. 

Christie, P., Li, X. L. & Chen, B. D. (2004). Arbuscular mycorrhiza can depress translocation 
of zinc to shoots of host plants in soils moderately polluted with zinc. Plant Soil 261, 
209-217. 

Claus, H. & Filip, Z. (1990). Effects of clay and other solids on the activity of phenolozidases 
produced by some fungi and actinomycetes. Soil Biol Biochem 22, 483-488. 

Cromack, K., Jr, Sollins, P., Graustein, W. C, Speidel, K., Todd, A. W., Spycher, G., Li, 
C. Y. & Todd, R. L. (1979). Calcium oxalate accumulation and soil weathering 
in mats of the hypogeous fungus Hysterangium crassum. Soil Biol Biochem 1 1, 
463-468. 

de la Torre, M. A. & Gomez-Alarcon, G. (1994). Manganese and iron oxidation by fungi 
isolated from building stone. Microb Ecol 27, 1 77-1 88. 

de la Torre, M. A., Gomez-Alarcon, G., Vizcaino, C. & Garcia, M. T. (1993). Biochemical 
mechanisms of stone alteration carried out by filamentous fungi living in monu- 
ments. Biogeochemistry 19, 129-147. 

de los Rios, A., Wierzchos, J. & Ascaso, C. (2002). Microhabitats and chemical micro- 
environments under saxicolous lichens growing on granite. Microb Ecol A3, 181-188. 

Demaneche, S., Jocteur-Monrozier, L., Quiquampoix, H. & Simonet, P. (2001). 
Evaluation of biological and physical protection against nuclease degradation of clay- 
bound plasmid DNA. Appl Environ Microbiol 67, 293-299. 

de Rome, L. & Gadd, G. M. (1987). Copper adsorption by Rhizopusarrhizus, Cladosporium 
resinae, and Penicillium italicum. Appl Microbiol Biotechnol 26, 84-90. 

Dinelli, E. & Tateo, F. (2001). Sheet silicates as effective carriers of heavy metals in the 
ophiolitic mine area of Vigonzano (northern Italy). Mineral Mag 65, 121-132. 

Dong, W. M., Wang, X. K., Bian, X. Y, Wang, A. X., Du, J. Z. & Tao, Z. Y (2001). 
Comparative study on the sorption/desorption of radioeuropium on alumina, 
bentonite and red earth: effects of pH, ionic strength, fulvic acid, and iron oxide 
in red earth . Appl Radiat Isotopes 54, 603-61 0. 

Dorioz, J. M., Robert, M. & Chenu, C. (1993). The role of roots, fungi and bacteria on clay 
particles organization. An experimental approach. Geoderma 56, 179-194. 

Douglas, S. & Beveridge, T. J. (1998). Mineral formation by bacteria in natural microbial 
communities. FEMS Microbiol Ecol 26, 79-88. 



SGM symposium 65 



224 G. M. Gadd, M. Fomina and E. P. Burford 



Easton, R. M. (1997). Lichen-rock-mineral interactions: an overview. In Biological-Mineral- 

ogica I Interactions, vol. 21, pp. 209-239. Edited by J. M. Mcintosh & L. A. Groat. 

Mineralogical Association of Canada Short Course Series. Ottawa: Mineralogical 

Association of Canada. 
Eckhardt, F. E. W. (1985). Solubilisation, transport, and deposition of mineral cations by 

microorganisms-efficient rock-weathering agents. In The Chemistry of Weathering, 

NATO Advanced Study Institutes Series, Series C, vol. 149, pp 161-173. Edited by 

J. I. Drever. Dordrecht: Reidel. 
Edwards, H. G. M., Farwell, D. W. & Seaward, M. R. D. (1991). Raman spectra of oxalates 

in lichen encrustations in Renaissance frescoes. Spectrochim Acta A Mol Spectrosc 

47, 1531-1539. 
Ehrlich, H. L. (1998). Geomicrobiology: its significance for geology. Earth Sci Rev 45, 45- 

60. 
Ehrlich, H. L (2002). Geomicrobiology. New York: Marcel Dekker. 
Etienne, S. (2002). The role of biological weathering in periglacial areas: a study of 

weathering rinds in south Iceland. Geomorphology 47 , 75-86. 
Ferris, F. G., Fyfe, W. S. & Beveridge, T. J. (1987). Bacteria as nucleation sites for authigenic 

minerals in a metal-contaminated lake sediment. Chem Geo/ 63, 225-232. 
Ferris, F. G., Wiese, R. G. & Fyfe, W. S. (1994). Precipitation of carbonate minerals by 

microorganisms: implications for silicate weathering and the global carbon dioxide 

budget. Geomicrobiol J 12, 1-13. 
Fletcher, M. (1987). How do bacteria attach to solid surfaces? Microbiol Sci A, 133-136. 
Fomina, M. & Gadd, G. M. (2002). Influence of clay minerals on the morphology of fungal 

pellets. Mycol Res 1 06, 1 07-1 17. 
Fomina, M. & Gadd, G. M. (2003). Metal sorption by biomassof melanin-producing fungi 

grown in clay-containing medium. J Chem Technol Biotechnol 78, 23-34. 
Fomina, M., Ritz, K. & Gadd, G. M. (2000a). Negative fungal chemotropism to toxic 

metals. FEM5 Microbiol Lett 1 93, 207-2 1 1 . 
Fomina, M. A., Kadoshnikov, V. M., Gromosova, E. N. & Podgorsky, V. S. (2000b). The 

role of clay minerals during fungal growth under extreme conditions. In Proceedings 

of International III International Seminar 'Mineralogy and Life: Biomineral Homol- 
ogies', pp. 146-148. Syktyvkar, Russia: Geoprint. 
Fomina, M., Ritz, K. & Gadd, G. M. (2003). Nutritional influence on the ability of fungal 

mycelia to penetrate toxic metal-containing domains. Mycol Res 107, 861-871 . 
Fomina, M. A., Alexander, I. J., Hillier, S. & Gadd, G. M. (2004). Zinc phosphate 

and pyromorphite solubilization by soil plant-symbiotic fungi. Geomicrobiol J 21, 

351-366. 
Fomina, M., Hillier, S., Charnock, J. M., Melville, K., Alexander, I. J. & Gadd, G. M. 

(2005a). Role of oxalic acid overexcretion in transformations of toxic metal minerals 

by Beauveria caledonica. Appl Environ Microbiol 71,371-381. 
Fomina, ML, Burford, E. P. & Gadd, G. M. (2005b). Toxic metals and fungal communities. 

In The Fungal Community: its Organization and Role in the Ecosystem, pp. 733-758. 

Edited by J. Dighton, P. Oudemans & J. White. Boca Raton, FL: CRC Press. 
Fomina, M. A., Alexander, I. J., Colpaert, J. V. & Gadd, G. M. (2005c). Solubilization of 

toxic metal minerals and metal tolerance of mycorrhizal fungi. Soil Biol Biochem 37, 

851-866. 
Fortin, D. & Beveridge, T. J. (1997). Role of the bacterium Thiobacillus in the formation of 

silicates in acidic mine tailings. Chem Geo! 141, 235-250. 
Fortin, D., Ferris, F. G. & Scott, S. D. (1998). Formation of Fe-silicates and Fe-oxides on 



SGM symposium 65 



Mycotransformation of minerals 225 



bacterial surfaces in samples collected near hydrothermal vents of Southern Explorer 

Ridge in Northeast Pacific Ocean. Am Mineral 83, 1399-1408. 
Fredrickson, J. K., McKinley, J. P., Nierzwicki-Bauer, S. A., White, D. C, Ringelberg, 

D. B., Rawson, S. A., Shu-Mei, L, Brockman, F. J. & Bjornstad, B. N. (1995). 

Microbial community structure and biogeochemistry of Miocene subsurface 

sediments: implications for long-term microbial survival. Mol Ecol 4, 619-626. 
Gadd, G. M. (1990). Fungi and yeasts for metal accumulation. In Microbial Mineral 

Recovery, pp. 249-275. Edited by H. L. Ehrlich & C. Brierley. New York: McGraw-Hill. 
Gadd, G. M. (1993). Interactions of fungi with toxic metals. New Phytol 124, 25-60. 
Gadd, G. M. (1999). Fungal production of citric and oxalic acid: importance in metal 

speciation, physiology and biogeochemical processes. Adv Microb Physiol 41, 47- 

92. 
Gadd, G. M. (2000a). Heavy metal pollutants: environmental and biotechnological aspects. 

In Encyclopedia of Microbiology, pp. 607-617. Edited by J. Lederberg. New York: 

Academic Press. 
Gadd, G. M. (2000b). Bioremedial potential of microbial mechanisms of metal mobilization 

and immobilization. CurrOpin Biotechnol 11, 271-279. 
Gadd, G. M. (editor) (2001a). Fungi in Bioremediation. Cambridge: Cambridge University 

Press. 
Gadd, G. M. (2001b). Accumulation and transformation of metals by microorganisms. In 

Biotechnology a Multi-volume Comprehensive Treatise, vol. 10, Special Processes, 

pp. 225-264. Edited by H.-J. Rehm, G. Reed, A. Puhler & P. Stadler. Weinheim: Wiley- 

VCH. 
Gadd, G. M. (2002). Interactions between microorganisms and metals/radionuclides: 

the basis of bioremediation. In Interactions of Microorganisms with Radionuclides, 

pp. 179-203. Edited by M. J. Keith-Roach & F.R. Livens. Amsterdam: Elsevier. 
Gadd, G. M. (2004). Mycotransformation of organic and inorganic substrates. Mycologist 

18,60-70. 
Gadd, G. M. (editor) (2005a). Fungi in Biogeochemical Cycles. Cambridge: Cambridge 

University Press (in press). 
Gadd, G. M. (2005b). Microorganisms in toxic metal polluted soils. In Microorganisms in 

Soils: Roles in Genesis and Functions, pp. 325-356. Edited by F Buscot & A. Varma. 

Berlin: Springer. 
Gadd, G. M. & Mowll, J. L. (1995). Copper uptake by yeast-like cells, hyphae and 

chlamydospores of Aureobasidium pullulans. Exp Mycol 9, 230-240. 
Gadd, G. M. & Sayer, J. A. (2000). Fungal transformations of metals and metalloids. In 

Environmental Microbe-Metal Interactions, pp. 237-256. Edited by D. R. Lovley. 

Washington, DC: American Society for Microbiology. 
Gadd, G. M., Ramsay, L, Crawford, J. W. & Ritz, K. (2001). Nutritional influence on 

fungal colony growth and biomass distribution in response to toxic metals. FEMS 

Microbiol Lett 204, 311-316. 
Garnham, G. W., Codd, G. A. & Gadd, G. M. (1991). Uptake of cobalt and cesium by 

microalgal- and cyanobacterial-clay mixtures. Microb Ecol 25, 71-82. 
Gerrath, J. F., Gerrath, J. A. & Larson, D. W. (1995). A preliminary account of endolithic 

algae of limestone cliffs of the Niagara Escarpment. Can J Bot 73, 788-793. 
Gerrath, J. F., Gerrath, J. A., Matthes, U. & Larson, D. W. (2000). Endolithic algae and 

cyanobacteria from cliffs of the Niagara Escarpment, Ontario, Canada. Can J Bot 78, 

807-815. 
Gleeson, D. B., Clipson, N. J. W., Melville, K., Gadd, G. M. & McDermott, F. P. (2005). 



SGM symposium 65 



226 G. M. Gadd, M. Fomina and E. P. Burford 



Mineralogical control of fungal community structure in a weathered pegmatitic 

granite. Microb Ecol (in press). 
Glowa, K. R. # Arocena, J. M. & Massicotte, H. B. (2003). Extraction of potassium and/or 

magnesium from selected soil minerals by Piloderma. Geomicrobiol J 20, 99-1 1 1 . 
Gomez-Alarcon, G., Munoz, M. L. & Flores, M. (1994). Excretion of organic acids by 

fungal strains isolated from decayed limestone. Int Biodeterior Biodegrad 34, 169— 

180. 
Gorbushina, A. A. & Krumbein, W. E. (2000). Subaerial microbial mats and their effects 

on soil and rock. In Microbial Sediments, pp. 161-169. Edited by R. E. Riding & 

S. M. Awramik. Berlin: Springer. 
Gorbushina, A. A., Krumbein, W. E., Hamann, R., Panina, L, Soucharjevsky, S. & 

Wollenzien, U. (1993). On the role of black fungi in colour change and bio- 
deterioration of antique marbles. Geomicrobiol J 11, 205-221 . 
Gorbushina, A. A., Lyalikova, N. N., Vlasov, D. Y. & Khizhnyak, T. V. (2002). Microbial 

communities on the monuments of Moscow and St. Petersburg: biodiversity 

and trophic relations. Microbiology (English translation of Mikrobiologiia) 71, 350- 

356. 
Goudie, A. S. (1996). Organic agency in calcrete development. J Arid Environ 32, 103-1 10. 
Grote, G. & Krumbein, W. E. (1992). Microbial precipitation of manganese by bacteria 

and fungi from desert rock and rock varnish. Geomicrobiol J 10, 49-57. 
Gu, J.-D., Ford, T. E., Berke, N. S. & Mitchell, R. (1998). Biodeterioration of concrete by 

th e f u n g u s Fusarium . Int Biodeterior Biodegrad 41, 101-109. 
Hamilton, W. A. (2003). Microbially influenced corrosion as a model system for the study 

of metal microbe interactions: a unifying electron transfer hypothesis. Biofouling 19, 

65-76. 
Henderson, M. E. K. & Duff, R. B. (1 963). The release of metallic and silicate ions from 

minerals, rocks and soils by fungal activity. J Soil Sci 14, 236-246. 
Hersman, L, Lloyd, T. & Sposito, G. (1995). Siderophore-promoted dissolution of 

hematite. Geochim Cosmochim Acta 59, 3327-3330. 
Hirsch, R, Eckhardt, F. E. W. & Palmer, R. J., Jr (1995). Fungi active in weathering rock and 

stone monuments. Can J Bot 73, 1384-1390. 
Hochella, M. F. (2002). Sustaining Earth; thoughts on the present and future roles in 

mineralogy in environmental science. Mineral Mag 66, 627-652. 
Hoffland, E., Giesler, R., Jongmans, T. & van Breemen, N. (2002). Increasing feldspar 

tunnelling by fungi across a north Sweden podzol chronosequence. Ecosystems 

5, 11-22. 
Jacobs, H., Boswell, G. P., Ritz, K., Davidson, F. A. & Gadd, G. M. (2002a). Solubilization 

of calcium phosphate as a consequence of carbon translocation by Rhizoctonia 

solani. FEMS Microbiol Ecol 40, 65-71. 
Jacobs, H., Boswell, G. P., Harper, F. A., Ritz, K., Davidson, F. A. & Gadd, G. M. (2002b). 

Solubilization of metal phosphates by Rhizoctonia solani. Mycol Res 106, 1468- 

1479. 
Jacobs, H., Boswell, G. P., Scrimgeour, C. M., Davidson, F. A., Gadd, G. M. & Ritz, K. 

(2004). Translocation of carbon by Rhizoctonia solani in nutritionally-heterogeneous 

environments. Mycol Res 108, 453^62. 
Jarosz-Wilkofazka, A. & Gadd, G. M. (2003). Oxalate production by wood-rotting fungi 

growing in toxic metal-amended medium. Chemosphere 52, 541-547. 
Johnstone, C. G. & Vestal, J. R. (1993). Biogeochemistry of oxalate in the antarctic 

cryptoendolithic lichen-dominated community. Microb Ecol 25, 305-319. 



SGM symposium 65 



Mycotransformation of minerals 227 



Jones, D. & Wilson, M. J. (1985). Chemical activities of lichens on mineral surfaces - a 

review. Int Biodeterior 2 1 , 99-1 04. 
Jongmans, A. G., Van Breemen, N., Lungstrom, U. & 7 other authors (1997). Rock- 
eating fungi. Nature 389, 682-683. 
Kadoshnikov, V. M., Zlobenko, B. P., Zhdanova, N. N. & Redchitz, T. I. (1995). Studies of 

application of micromycetes and clay composition for decontamination of building 

materials. In Mixed Wastes and Environmental Restoration - Working Towards a 

Cleaner Environment, WM'95, pp. 61-63. Tucson, AZ: WM Symposia. 
Kahle, C. F. (1977). Origin of subaerial Holocene calcareous crusts: role of algae, fungi and 

sparmicristisation. Sedimentology 24, 413-435. 
Klappa, C. F. (1979). Calcified filaments in Quaternary calcretes: organo-mineral inter- 
actions in the subaerial vadose environment. J Sediment Petrol 49, 955-968. 
Kraemer, S. M., Cheah, S.-F., Zapf, R., Xu, J., Raymond, K. N. & Sposito, G. (1999). Effect 

of hydroxamate siderophores on Fe release and Pb(ll) adsorption by goethite. 

Geochim Cosmochim Acta 63, 3003-3008. 
Krumbein, W. E., Urzi, C. & Gehrmann, C. (1991). On the biocorrosion and bio- 
deterioration of antique and medieval glass. Geomicrobiol J 9, 139-160. 
Krumhansl, J. L, Brady, P. V. & Anderson, H. L. (2001). Reactive barriers for 137 Cs 

retention. J Contam Hydro! 47, 233-240. 
Krupa, P. & Kozdroj, J. (2004). Accumulation of heavy metals by ectomycorrhizal 

fungi colonizing birch trees growing in an industrial desert soil. World J Microbiol 

Biotechnol 20, 427-430. 
Kumar, R. & Kumar, A. V. (1999). Biodeterioration of Stone in Tropical Environments: 

an Overview. Los Angeles: Getty Conservation Institute. 
Lee, G.-H. & Stotzky, G. (1999). Transformation and survival of donor, recipient, 

and transformants of Bacillus subtilis in vitro and in soil. So/7 Biol Biochem 31,1 499- 

1508. 
Li, L. Y. & Li, R. S. (2000). The role of clay minerals and the effect of H + ions on removal of 

heavy metal (Pb 2+ ) from contaminated soils. Can Geotechnol J 37, 296-307. 
Liermann, L. J., Kalinowski, B. E., Brantley, S. L. & Ferry, J. G. (2000). Role of bacterial 

siderophores in dissolution of hornblende. Geochim Cosmochim Acta 64, 587- 

602. 
Lotareva, O. V. & Prozorov, A. A. (2000). Effect of the clay minerals montmorillonite and 

kaolinite on the generic transformation of competent Bacillus subtilis cells. 

Microbiology (English translation of Mi krobiologiia) 69, 571-574. 
Lundstrom, U. S., van Breemen, N. & Bain, D. (2000). The podzolization process. A 

review. Geoderma 94, 91-107. 
Lunsdorf, H., Erb, R. W., Abraham, W. R. & Timmis, K. N. (2000). 'Clay hutches': a novel 

interaction between bacteria and clay minerals. Environ Microbiol 2, 161-1 68. 
Mandal, S. K., Roy, A. & Banerjee, P. C. (2002). Iron leaching from china clay by fungal 

strains. Trans Indian Inst Metals 55, 1-7. 
Manley, E. & Evans, L. (1986). Dissolution of feldspars by low-molecular-weight aliphatic 

and aromatic acids. Soil Sci 141, 1 06-1 12. 
Manoli, F., Koutsopoulos, E. & Dalas, E. (1997). Crystallization of calcite on chitin. J Crystal 

Growth 182, 116-124. 
Marshall, K. C. (1988). Adhesion and growth of bacteria at surfaces in oligotrophic habitats. 

Can J Microbiol 34, 593-606. 
Martens, D. A. & Frankenberger, W. T. (1992). Decomposition of bacterial polymers in soil 

and their influence on soil structure. Biol FertH Soils 13, 65-73. 



SGM symposium 65 



228 G. M. Gadd, M. Fomina and E. P. Burford 



Martin, J. P., Filip, Z. & Haider, K. (1976). Effect of montmorollonite and humate 

on growth and metabolic activity of some actinomycetes. Soil Biol Biochem 8, 

409-413. 
Martino, E., Perotto, S., Parsons, R. & Gadd, G. M. (2003). Solubilization of insoluble 

inorganic zinc compounds by ericoid mycorrhizal fungi derived from heavy metal 

polluted sites. Soil Biol Biochem 35, 1 33-1 41 . 
Maurice, P. A., Hochella, M. F., Parks, G. A., Sposito, G. & Schwertmann, U. (1995). 

Evolution of hematite surface microtopography upon dissolution by simple organic 

acids. Clays Clay Miner 43, 29-38. 
May, E. (2003). Microbes on building stone for good or bad? Culture 24, 4-8. 
May, E., Lewis, F. J., Periera, S., Taylor, S., Seaward, M. R. D. & Allsopp, D. (1993). 

Microbial deterioration of building stone -a review. Biodeterior Abstrl ', 109-1 12. 
Meharg, A. A. (2003). The mechanistic basis of interactions between mycorrhizal 

associations and toxic metal cations. Mycol Res 107, 1253-1265. 
Mehta, A. P., Torma, A. E. & Murr, L. E. (1979). Effect of environmental parameters on 

the efficiency of biodegradation of basalt rock by fungi. Biotechnol Bioeng 21, 

875-885. 
Miyata, N., Tani, Y, Iwahori, K. & Soma, M. (2004). Enzymatic formation of manganese 

oxides by an >4cre/77on/i//T7-like hyphomycete fungus, strain KR21 -2. FEMS Microbiol 

fco/47, 101-109. 
Money, N. P. (1999). Biophysics: fungus punches its way in. Nature 401, 332-333. 
Money, N. P. (2001). Biomechanics of invasive hyphal growth. In The Mycota VIII, pp. 3-17. 

Edited by K. Esser. Berlin: Springer. 
Money, N. P. & Howard, R. J. (1 996). Confirmation of a link between fungal pigmentation, 

turgor pressure, and pathogenicity using a new method of turgor measurement. 

Fungal Genet Biol 20, 2 1 7-227. 
Monger, C. H. & Adams, H. P. (1996). Micromorphology of calcite-silica deposits, Yucca 

Mountain, Nevada. Soil Sci SocAm J 60, 519-530. 
Morley, G. F. & Gadd, G. M. (1 995). Sorption of toxic metals by fungi and clay minerals. 

Mycol Res 99, 1429-1438. 
Morley, G., Sayer, J., Wilkinson, S., Gharieb, M. & Gadd, G. M. (1996). Fungal 

sequestration, solubilization and transformation of toxic metals. In Fungi and 

Environmental Change, pp. 235-256. Edited by J. C. Frankland, N. Magan & 

G. M. Gadd. Cambridge: Cambridge University Press. 
Mottershead, D. & Lucas, G. (2000). The role of lichens in inhibiting erosion of a soluble 

rock. Lichenologist 32, 601-609. 
Muller, B., Burgstaller, W., Strasser, H., Zanella, A. & Schinner, F. (1995). Leaching of 

zinc from an industrial filter dust with Penicillium, Pseudomonas and Coryne- 

bacterium: citric acid is the leaching agent rather than amino acids. J Ind Microbiol 

14,208-212. 
Nica, D., Davis, J. L., Kirby, L., Zuo, G. & Roberts, D. J. (2000). Isolation and 

characterization of microorganisms involved in the biodeterioration of concrete in 

sewers. Int Biodeterior Biodegrad 46, 61-68. 
Nimis, P. L., Pinna, D. & Salvadori, O. (1992). Licheni e Conservazione del Monumenti. 

Bologna: Cooperativa Libraria Universitaria Ediatrice Bologna (in Italian). 
Paris, F., Bonnaud, P., Ranger, J. & Lapeyrie, F. (1995). In vitro weathering of phlogopite 

by ectomycorrhizal fungi. Plant Soil 177, 191-201. 
Pedersen, K. (1999). Subterranean microorganisms and radioactive waste disposal in 

Sweden. Eng Geol 52, 1 63-1 76. 



SGM symposium 65 



Mycotransformation of minerals 229 



Perfettini, J. V., Revertegat, E. & Langomazino, N. (1991). Evaluation of cement degra- 
dation by the metabolic activities of two fungal strains. Experientia 47, 527-533. 
Perotto, S., Girlanda, M. & Martino, E. (2002). Ericoid mycorrhizal fungi: some new 

perspectives on old acquaintances. Plant Soil 244, 41-53. 
Pillichshammer, M., Pumpel, T., Poder, R., Eller, K. # Klima, J. & Schinner, F. (1995). 

Biosorption of chromium to fungi. BiometalsS, 1 17-121. 
Pitonzo, B. J., Amy, P. S. & Rudin, M. (1999). Effect of gamma radiation on native endo- 

lithic microorganisms from a radioactive waste deposit site. Radiat Res 1 52, 64-70. 
Puget, P., Angers, D. A. & Chenu, C. (1999). Nature of carbohydrates associated with 

water-stable aggregates of two cultivated soils. Soil Biol Biochem 31 , 55-63. 
Purvis, O. W. (1 996). Interactions of lichens with metals. Sci Prog 79, 283-309. 
Purvis, O. W., Bailey, E. H., McLean, J., Kasama, T. & Williamson, B. J. (2004). Uranium 

biosorption by the lichen Trapelia involuta at a uranium mine. Geomicrobiol J 21, 

159-167. 
Putnis, A. (2002). Mineral replacement reactions; from macroscopic observations to 

microscopic mechanisms. Mineral Mag 66, 689-708. 
Rodriguez Navarro, C, Sebastian, E. & Rodriguez Gallego, M. (1997). An urban model 

for dolomite precipitation: authigenic dolomite on weathered building stones. 

Sediment Geo/ 109, 1-11. 
Rogers, J. R., Bennett, P. C. & Choi, W. J. (1998). Feldspars as a source of nutrients 

for microorganisms. Am Mineral 83, 1 532-1 540. 
Roldan, M., Clavero, E. & Hernandez-Marine, M. (2002). Biofilm structure of cyano- 

bacteria in catacombs. Coalition 5, 6-8. 
Rosling, A., Lindahl, B. D., Taylor, A. F. S. & Finlay, R. D. (2004). Mycelial growth and 

substrate acidification of ectomycorrhizal fungi in response to different minerals. 

FEMS Microbiol Ecol 47, 3 1 -37 . 
Rossi, G. (1979). Potassium recovery through leucite bioleaching: possibilities and 

limitations. In Metallurgical Applications of Bacterial Leaching and Related 

Phenomena, pp. 279-319. Edited by L. E.Murr,A. E.Torma&J. E. Brierley. New York: 

Academic Press. 
Rossi, G. & Ehrlich, H. L. (1990). Other bioleaching processes. In Microbial Mineral 

Recovery, pp. 149-170. Edited by H. L. Ehrlich & C. L. Brierley. New York: McGraw-Hill. 
Sayer, J. A. & Gadd, G. M. (2001). Binding of cobalt and zinc by organic acids and culture 

filtrates of Aspergillus niger grown in the absence or presence of insoluble cobalt or 

zinc phosphate. Mycol Res 105, 1261-1267. 
Schlecht-Pietsch, S., Wagner, U. & Anderson, T. H. (1 994). Changes in composition of soil 

polysaccharides and aggregate stability after carbon amendments to different 

textured soils. Appl Soil Ecol 1, 145-1 54. 
Schutzendubel, A. & Polle, A. (2002). Plant responses to abiotic stresses: heavy metal- 
induced oxidative stress and protection by mycorrhization. J Exp Bot 53, 1351-1365. 
Smits, M. M., Hoffland, E., Jongmans, A. G. & van Breemen, N. (2005). Contribution of 

mineral tunnelling to total feldspar weathering. Geoderma 125, 59-69. 
Staley, J. T., Palmer, F. & Adams, J. B. (1982). Microcolonial fungi: common inhabitants on 

desert rocks. Science 21 5, 1 093-1 095. 
Sterf linger, K. (2000). Fungi as geologic agents. Geomicrobiol J 17, 97-124. 
Stone, A. T. (1997). Reactions of extracellular organic ligands with dissolved metal ions and 

mineral surface. In Geomicrobiology: Interactions between Microbes and Minerals, 

Reviews in Mineralogy, vol. 35, pp. 309-341 . Edited by J. F. Banfield & K. H. Nealson. 

Washington, DC: Mineralogical Society of America. 



SGM symposium 65 



230 G. M. Gadd, M. Fomina and E. P. Burford 



Stotzky, G. (1966). Influence of clay minerals on microorganisms. II. Effect of various clay 

species, homoionic clays, and other particles on bacteria. Can J Microbiol 12, 

831-848. 
Stotzky, G. (2000). Persistence and biological activity in soil of insecticidal proteins from 

Bacillus thuringiensis and of bacterial DNA bound on clays and humic acids. J Environ 

Qual 29, 691-705. 
Thompson, J. B. & Ferris, F. G. (1990). Cyanobacterial precipitation of gypsum, calcite, and 

magnesite from natural alkaline lake water. Geology 18, 995-998. 
Timonin, M. I., Illman, W. I. & Hartgerink, T. (1972). Oxidation of manganous salts of 

manganese by soil fungi. Can J BotW, 793-799. 
Tisdall, J. M., Smith, S. E. & Rengasamy, P. (1997). Aggregation of soil by fungal hyphae. 

AustJSoilRes 35, 55-60. 
Tullio, M., Pierandrei, F., Salerno, A. & Rea, E. (2003). Tolerance to cadmium of vesicular 

arbuscular mycorrhizae spores isolated from a cadmium-polluted and unpolluted 

soi I . Biol Fertil Soils 37,211-214. 
van Breemen, N., Lundstrom, U. S. & Jongmans, A. G. (2000). Do plants drive 

podzolization via rock-eating mycorrhizal fungi? Geoderma 94, 163-171 . 
Vaughn, D. J., Pattrick, R. A. D. & Wogelius, R. A. (2002). Minerals, metals and mol- 
ecules; ore and environmental mineralogy in the new millennium. Mineral Mag 66, 

653-676. 
Verrecchia, E. P. (2000). Fungi and sediments. In Microbial Sediments, pp. 69-75. Edited 

by R. E. Riding & S. M. Awramik. Berlin: Springer. 
Verrecchia, E. P., Dumont, J. L. & Rolko, K. E. (1990). Do fungi building limestones exist 

in semi-arid regions? Naturwissenschaften 77, 584-586. 
Verrecchia, E. P., Dumont, J. L. & Verrecchia, K. E. (1993). Role of calcium-oxalate 

biomineralization by fungi in the formation of calcretes - a case-study from Nazareth, 

Israel. J Sediment Petrol 63, 1 000-1 006. 
Wallander, H., Mahmood, S., Hagerberg, D., Johansson, L & Pallon, J. (2003). 

Elemental composition of ectomycorrhizal mycelia identified by PCR-RFLP analysis 

and grown in contact with apatite or wood ash in forest soil. FEMS Microbiol Ecol 44, 

57-65. 
Watts, H. J., Very, A. A., Perera, T. H. S., Davies, J. M. & Gow, N. A. R. (1998). 

Thigmotropism and stretch -activated channels in the pathogenic fungus Candida 

albicans. Microbiology 144, 689-695. 
Webley, D. M., Henderson, M. E. F. & Taylor, I. F. (1963). The microbiology of rocks and 

weathered stones. J Soil Sci 14, 102-1 12. 
Welch, S. A. & Vandevivere, P. (1994). Effect of microbial and other naturally occurring 

polymers on mineral dissolution. Geomicrobiol J 12, 227-238. 
Welch, S. A., Barker, W. W. & Banfield, J. F. (1999). Microbial extracellular polysaccharides 

and plagioclase dissolution. Geochim Cosmochim Acta 63, 1405-1419. 
Wheeler, M. H. & Bell, A. A. (1988). Melanins and their importance in pathogenic fungi. 

Curr Topics Med Mycol 2, 338-385. 
White, I. D., Mottershead, D. N. & Harrison, S. J. (1992). Environmental Systems: an 

Introductory Text. London: Chapman & Hall. 
Wild, A. (1993). Soils and the Environment: an Introduction. Cambridge: Cambridge 

University Press. 
Wolfaardt, G. M., Lawrence, J. R., Robarts, R. D., Caldwell, S. J. & Caldwell, D. E. 

(1994). Multicellular organization in a degradative biofilm community. Appl Environ 

Microbiol 60, 434-446. 



SGM symposium 65 



Mycotransformation of minerals 231 



Zhdanova, N. N. f Zakharchenko, V. A. ; Vember, V. V. & Nakonechnaya, L T. (2000). 

Fungi from Chernobyl: mycobiota of the inner regions of the containment structures 
of the damaged nuclear reactor. Mycol Res 104, 1421-1426. 
Zhdanova, N., Fomina, M., Redchitz, T. & Olsson, S. (2001). Chernobyl effects: growth 
characteristics of soil fungi Cladosprorium dadosporioides (Fresen) de Vries with and 
without positive radiotropism. PolJEcol49, 309-318. 



SGM symposium 65 



The deep intraterrestrial 
biosphere 



Karsten Pedersen 



Deep Biosphere Laboratory, Department of Cell & Molecular Biology, Goteborg University, 
Box 462, SE-405 30 Goteborg, Sweden 



INTRODUCTION 

Exploration of the microbial world started slowly about 350 years ago, when van 
Leeuwenhoek and his contemporaries first focused their microscopes on extremely 
small living things. It is only during the last 20 years, however, that exploration of the 
world of intraterrestrial microbes has gathered momentum. Previously, it had generally 
been assumed that persistent life could not exist deep underground, out of reach of the 
sun and a photosynthetic ecosystem base. In the mid-1980s, scientists started to drill 
deep holes, from hundreds to a thousand metres deep, in both hard and sedimentary 
bedrock, and up came microbes in numbers equivalent to those found in many surface 
ecosystems. The world of intraterrestrial microbes had been discovered. 

Intraterrestrial ecosystems have been reviewed elsewhere and the content of those 
reports need not be repeated here (Ghiorse & Wilson, 1988; Pedersen, 1993a, 2000; 
Bachofen, 1997; Bachofen et al., 1998; Fredrickson 6c Fletcher, 2001; Amend & Teske, 
2005). Instead, this chapter will focus on characteristics that distinguish the intra- 
terrestrial from the terrestrial world. 

Most ecosystem environments have specific, distinguishing characteristics. The environ- 
ments of intraterrestrial microbial ecosystems occupy a special position, differing 
substantially in many respects from those of most surface-based ecosystems. In many 
ways, underground ecosystems must be approached quite differently from the way in 
which those on the surface would be approached. The intraterrestrial world is huge and 
diverse and this chapter can only provide a brief overview of the following matters: 

SGM symposium 65: Micro-organisms and Earth systems - advances in geomicrobiology. 
Editors G. M. Gadd, K. T. Semple & H. M. Lappin-Scott. Cambridge University Press. ISBN 521 86222 1 ©SGM 2005 



234 K. Pedersen 



Where do intraterrestrial environments begin? 

Variability of intraterrestrial environments. 

Strategies for exploring intraterrestrial environments. 

Organisms living in intraterrestrial ecosystems. 

Range of biomass in various intraterrestrial environments. 

Intraterrestrial species diversity 

Energy sources for intraterrestrial life. 

Activity of intraterrestrial life forms. 

Several different terms are used to refer to environments under the ground and seafloor 
surface. Though it is difficult for one word to encompass all such environments, this 
paper generally uses the term 'intraterrestrial'. Other common terms are 'subseafloor', 
'subsurface' and 'underground'. 

WHERE DO INTRATERRESTRIAL ENVIRONMENTS BEGIN? 

There is no consensus as to the answer to this question, and various scientists would 
probably suggest different answers. The view adopted in this chapter is that the intra- 
terrestrial world begins where contact with surface ecosystems is lost. This lies beneath 
the soil and root zones, beneath the groundwater table, beneath sediment and crust 
surfaces. A long time must have passed since the last surface contact, 'a long time' 
implying at least several decades and often hundreds of years or more. So, it is not depth 
per se that defines an intraterrestrial ecosystem, but rather the duration of isolation 
from the surface. 

VARIABILITY OF INTRATERRESTRIAL ENVIRONMENTS 

Many different minerals and rock types constitute our planet. These rock types lie 
along a continuum, ranging from very hard rocks (such as granites and basalts) through 
sedimentary rocks (such as sandstones) to fairly soft sediments not yet defined as rock. 
Just two types will be defined and discussed here: (i) those that are too tight to allow the 
presence of microbes elsewhere than in fractures, i.e. hard rocks, and (ii) those that are 
porous enough to allow microbes inside the rock, i.e. sedimentary rocks. 

When formed, hard rocks generally experience temperatures that greatly exceed the 
limit for life and are therefore sterile after formation. They cool off and fracture with 
time and the fractures can become colonized by microbes. Sedimentary rocks form 
slowly, generally on the seafloor; over geological time scales, plate tectonics may 
eventually move these rocks up from the sea to constitute land. Sedimentary rocks that 
do not exceed, when formed, the temperature limit for life (currently defined as 113 °C) 
can harbour microbes borne by the sediment particles during the sedimentation process 
as well as microbes that arrived later. 



SGM symposium 65 



The deep intraterrestrial biosphere 235 

Intraterrestrial environments can be described as basically solid but containing some 
groundwater in fractures and pores. In this environment, huge surface areas are exposed 
to groundwater; solid dissolution and precipitation processes are consequently very 
important. At the opposite extreme, aquatic environments, such as sea or lake environ- 
ments, consist of water with tiny amounts of dispersed solids; these solids are thus of 
very limited importance for sea and lake water composition. 

Mixing processes also differ between intraterrestrial and surface aquatic environments. 
In seas and lakes, the water is comparatively homogeneous and well mixed by wind and 
currents (although there certainly are layered environments such as meromictic lakes). 
Underground, in particular in fractured rocks, there are many different water types that 
mix at fracture crossroads. Each water type has it own flow-path history and has met 
with different minerals and organisms on its way to a mixing point. Understanding 
mixing is crucial when dealing with groundwater. Any groundwater can be charac- 
terized by its age relative to when it left the ground surface and also by its origin, usually 
defined as an end member of a mixed groundwater. It may seem a simple matter to 
clarify the age and origin of a groundwater sample, but it is not. Different dating 
methods, for example, analysing the amounts of tritium originating from nuclear bomb 
tests or of 14 C or 37 C1, commonly determine the same groundwater to be of different 
ages. This is because a single groundwater sample comprises a mixture of waters of 
various ages and origins. For example, rainwater penetrating deep underground in the 
Scandinavian region, mixing with water of glacial meltwater origin, would produce a 
groundwater containing two age signatures: those of recent and of 10000-year-old 
water. It would not be obvious whether any microbes detected in this groundwater were 
originally borne by the rainwater or the glacial meltwater. Understanding the origins of 
the various component waters of a groundwater is a challenge, but can be approached 
by statistical methods. By statistical analysis of extensive datasets of ground- 
water chemistry, end members have been identified for Scandinavian groundwater 
(Laaksoharju et ai, 1999). Models have been built that guide scientists as to how 
different end members contribute to the groundwater being analysed. 

The results indicate that, as well as a continuum of types of solids in the intraterrestrial 
environment, as explained above, there is also a continuum of groundwater types. 
Every sample brought up from underground is more or less different from all other 
such samples. This is because groundwater systems are non-homogeneous. Comparing 
a surface aquatic environment, such as a sea, with most intraterrestrial environments 
results in contrasts. Spatially, the sea varies little within a local area, while a ground- 
water environment can vary greatly over very short distances. However, when it 
comes to temporal variation, the reverse holds true: the sea has marked diurnal and 
seasonal cycles, while the intraterrestrial environment shows no change over days or 

SGM symposium 65 



236 K. Pedersen 



seasons. Observation of intraterrestrial environments may need to continue for many 
years before significant changes can be resolved. 

STRATEGIES FOR EXPLORING INTRATERRESTRIAL 
ENVIRONMENTS 

Sampling intraterrestrial micro-organisms requires a borehole or a tunnel. The 
preferred method of scientific drilling is the core-drilling technique, which allows 
the retrieved cores to be analysed for mineralogy and microbiology The borehole can 
later be explored in several different ways. 

To sample groundwater from water-conducting fractures or aquifers, part of the 
borehole can be packed off and sampled. Fig. 1 depicts several possible sampling 
techniques. A fairly simple sampling technique is to pump water though tubing 
(Fig. la); however, in the case of deep boreholes, approaching a kilometre or more 
in depth, it make take considerable time to pump out the water, and much of the 
microbiology and some geochemistry may change in the process. Such changes are 
induced by the pressure drop and change in temperature. A second option is to lower 
samplers and open them at the sampling depth (Fig. lb, c); in this way, the pressure 
of the sample can be kept at the pressure of the sampling depth. This strategy has 
been applied in sampling deep Fennoscandian groundwaters (Haveman et al., 1999; 
Haveman 6c Pedersen, 2002). If there is access to a tunnel or a mine beneath the surface 
of the groundwater table, groundwater can be obtained from boreholes without 
pumping (Fig. Id). Deep South African goldmines have allowed scientists to access 
the deep intraterrestrial biosphere via boreholes drilled to depths as great as 3-5 km 
(Onstott et al., 1998). Sampling of fractures that deliver groundwater to a tunnel does 
not require drilling (Fig. le). This approach was used by Ekendahl et al. (2003) when 
sampling for fungi in deep groundwater. Finally, it may be possible to work in situ by 
installing a laboratory near the microbes in a tunnel and working under in situ 
conditions. One example of an underground laboratory can be found in the Swedish 
Aspo Hard Rock Laboratory (Fig. 2). There, three water-bearing fractures have been 
intersected by drilling and then packed off. Groundwater from those fractures is 
circulated through flow cells in the laboratory and then back to the fracture again, 
maintaining the pressure, temperature and chemistry exactly at in situ levels. These 
laboratory flow cells can be regarded as artificial extensions of the fractures. 

The superdeep well SG-3 (12262 m deep) in the Pechenga-Zapolyarny area of the Kola 
Peninsula, Russia, is currently the deepest drilled hole in the world (Butler, 1994). 
Drilling a superdeep borehole becomes increasingly difficult with increasing depth, and 
success depends on the qualities of the geological formation drilled, the quality of 
equipment used and the skills of the drilling personnel. Other superdeep boreholes 

SGM symposium 65 



The deep intraterrestrial biosphere 237 




i i....^w»^!-"-^ ! . !j -.- s '-. ■'■■■'. •••. .■ ■■.•«.:':■- i — ~~ — n^ra^^rTY^?"^ 

!** i- -. ■ id i_ . ^ c \ qj ■ • ■ » 



iD 




* ,_ j J i ''III I 

».-A,-l.y'--. "W ■■ V. .'- ...v .'.l'.^, ..■.•.,.•■...' (J.. . ..i'V ^V<'- , \.- , ^^^^^^^^--v:'-i!'-. l! ^^W! 








3 
=1 
o 

CD 

> 



^CD 



if>* 



en 



2? E " :i '.: r " iu ,'. n"i 



TO 



>0 



\ -3? 
w> 




• \ 



CD 



QJ 
-i— < 
TO 



.1" "D 



TO 

QJ 
TO 



cd 

o 
u 



o 

Cn 
QJ 

y 

o 



o 

oo CD 

TO C 

cn o 

T3 O 



CD 



TO 



•— oo 



~o 


O 


M— 

o 


^ 





1/1 


(0 


~l 


> 

GJ 


O 


-i— 1 


TO 


(1) 


> 








> | 


(]> 


o 


-i— 1 


l/l 


M — 


( ) 


< 


E 


00 


^ 


GJ 


CL 


C 


GJ 


C 

4— ■ 


O 


O 
CD 


C 
O 


TO 


f ) 






I 


-»' "v 


~ 


fU 


00 




"D 


p 


!_ 


r\> 


_> 




C) 


~i 


CD 


TO 


CD 


i i 



C0~O Z5 

'^ £ 

CD TO 

T3 CD 

oo g 

CD "p- 

S c? 



TO 
U 

'c 
9^ -*= 

-Q u 

<- CD 

»— -t— > 

S-o 

£ TO 
CD s_ 

~±Z QJ 
CD 

cr c 

<f TO 

. "D 

1- 

TO c 

£ o 

TO Q- 
!> CD 

T3 -f= 



E o 

O X2 

4= o 
■a .y 

S^ E 

o^ 

u c 

-a to 

c 00 

TO TO 

-a P 1 

O ^ 

C CD 

TO "O 

<-» 00 



3 T3 

° m 

^- TO 

01 to 

4= T3 



o O 

CD -^ 

00 CD 

CD > 

-— QJ 



QJ 

Q_ 

o 



.g 

■m : 

TO CD 

-P C 

CD d 

C Z5 

CD -t- 1 

Q- CD 



Q- O 

CD -t- 1 

"O CD 

E^ 

O CD 

^— oo 

i_ TO 

CD £ 

CL U 

Q. Z5 

oo un 

O ^i 

+-- Q_ 

^ CD 



CD 



P. o 



O Td 



CD ro 



.5£2 CD 



O 



oo 
CD 



U 

QJ 

Q_ 

oo 

CD 



y cd 
-c y 

-> TO 



O 
Q_ 

E 

00 



X 

p 

Q_ 



CD- 
CD 

u 

X 

CD 

QJ 



_cy o 

CD 



O 
_Q 

M— 

o 

cn 



CD 
JZ 
-t— ' 

TO 

> 

00 

u 
u 

o 

CD 
TO 

"D 

C 

o 
cn 

~a 

c 

TO 

!ro 

QJ 
-t— ■ 
TO 



O 

QJ 
C 
O 
oo 
QJ 
-i— ' 
_TO 

O 
00 



U 



O 



u 

TO 
QJ 



QJ oo 
Q_ OJ 

E^ 
L^ E 

_CD O 

O ^ 

-c -a 

QJ QJ 

9 o 

b "° 
< S 

CQ TO 

g& 

cn_Q 



oo _ 

CL — 

E oo' 
CD 



D" TO 

4— > 00 

O QJ 
C i 

O & 
o a. 

QJ TO 
> 



TO u 
>, C 

JD to 

oo -e 

QJ = 

oo r^ 

TO Tn 

U 

C TO 

c c 

to ■;= 



QJ > 

"D > 

+- ' ^~ 

TO QJ 

4— ' 

£ TO 



y^ 



o 



o 

_Q CD 

ID £ 



— u 



o 

! 

cn 



■ ^s 



QJ 



oo 

C 

.O 

4— ' 

u 

QJ 
oo 

"to 

QJ 
> 
QJ 
oo 

O 

QJ 
C 
O 



"D 



TO -pj 

oo .y 

-Q _£= ^ 



oo 



o 


QJ 


QJ 
U 


F 


(0 


TO 


!. 


00 



^QJ O 

CD 

U 

TO 
Q_ 



cr 

TO 

O 

4— ' 

00 
00 

QJ 
u 
u 

< 



Ui 



QJ 

TO 

oo 

31 

O 

_c 

QJ 

o 

_Q 



=> QJ 

^^ 

cn _ 
qj tu 

"1 
o 9- 

4— ' 

i_ CD 

CD -C 



TO 

cd cn 



E 

o 



a. 



TO 
00 



>> 00 

oo CD 

3 y 

O '> 
CD CD 

2 cn 



E CL 

•^ E 

"O TO 

QJ °° 
oo o 

3 4-^ 

QJ "C 
_Q QJ 

O u 

4-- CD 

oo C 

oo O 
oo u 

^ TO 
CL-4- 

E^ 

o^T3 
QJ^ 
O u 

y to 

E CL 



C1J CD 

^ oo 

3 QJ 

oo n 

oo y- 

Qj QJ 

y c 

"4-J TO 

TO ->, 

4— ' -^ 

00 4— ' 

QJ TO 

-C 00 

? « 



> TO 

-o u 

Qj oo 

t; a; 

E E 

u c 

00 



_CD 

_Q 
O 

CL 



oo 

QJ 



U 

TO 



SGM symposium 65 



238 K. Pedersen 




Fig. 2. Artist's impression of the MICROBE underground laboratory at the 450 m level in the 
Aspo hard rock tunnel. The laboratory (to the right) is situated in a steel container and connected 
to three discrete fractures in the rock matrix. Tubing connects the systems in the laboratory with 
the groundwater in the aquifers under in situ pressure, chemistry and temperature. 

include several drilled as part of the German Continental Deep Drilling Program (KTB) 
into the crystalline rock of the Bavarian Black Forest (Schwarzwald) basement in 
Central Europe (Butler, 1994); the deepest of the six wells drilled reached a depth of 
9100 m, where an in situ temperature of 265 °C was recorded. One of these KTB wells 
was searched for hyperthermophiles at a depth of 4000 m; culturable micro-organisms 
could not be demonstrated. Another very deep borehole was drilled in Gravenberg, 
Sweden, in a search for deep earth gases (Gold, 1992). It reached a depth of 6800 m and 
thermophilic bacteria were successfully enriched and isolated from a depth of 5278 m, 
where the temperature was 65-75 °C (Szewzyk et al., 1994). The temperature data 
obtained from these boreholes indicate that the temperature increase with depth varies 
depending on the geographical location of the borehole. In most of Scandinavia, the 
increase is very slow, ranging between 1 and 1-5 °C per 100 m; elsewhere the gradient 
is much steeper, with the steepest temperature gradients obviously being found in 
hot-spring areas. 

ORGANISMS THAT LIVE IN INTRATERRESTRIAL 
ENVIRONMENTS 

Generally, representatives of most major groups of surface-living microbes should 
be able to live underground -with some obvious exceptions. The underground and 



SGM symposium 65 



The deep intraterrestrial biosphere 239 



Table 1. Range of total counts in various intraterrestrial environments 

Values are means of data from various reviews: Ghiorse & Wilson (1 988), Pedersen (1 993a, 2000), 
Bachofen (1997), Bachofen etal. (1998) and Fredrickson & Fletcher (2001). gdw. Gram dry weight. 



Environment 




Total counts 






Unattached (ml -1 ) 


Attached (cm -2 ) 


Attached (gdw -1 ) 


Hard rock fractures 


1 3 -1 6 


1 5 -1 7 


— 


Sedimentary rock 


1 3 -1 o 6 


— 


1 4 -1 o 8 


Subseafloor sediments 


— 


— 


1 5 -1 o 8 



subseafloor world is dark, water-filled and usually anaerobic with reducing conditions. 
The boundary between anaerobic and aerobic environments commonly follows the 
groundwater table. Consequently, strictly and facultatively anaerobic archaea, bacteria, 
phages, fungi and protozoa should be able to live an active life there, but not strict 
aerobes, pathogens, photosynthetic organisms or symbionts of plants, animals or 
insects. 

RANGE OF BIOMASS IN VARIOUS INTRATERRESTRIAL 
ENVIRONMENTS 

Many microbes live in a planktonic state, but even more tend to live in an attached 
state (Characklis & Marshall, 1990). As the intraterrestrial world is a world of solids 
with some interpenetrating water, there is always a nearby surface on which to attach. 
Experiments suggest that underground ecosystems in fractured rock are dominated 
by attached microbes making up biofilms (Pedersen, 2001), but it remains to be demon- 
strated that microbial biofilms exist on fracture surfaces in deep environments. It is 
easier to obtain a reasonably undisturbed groundwater sample from a borehole than 
it is to obtain a drill core sample of an undisturbed intersected fracture. The data 
concerning intraterrestrial organisms, in particular in fractured rock, are thus biased 
towards planktonic microbes found in groundwater samples. 

Most total counts in hard rock refer to the number of unattached microbes; in 
sediments, however, the data generally refer to the weight and include both attached 
and unattached cells. Typically, the numbers found in fractures and sediments are 
relatively equal to those found in clean surface waters and sediments, although the 
range is large (Table 1). A hundred million cells may sound like a lot, but bear in mind 
that this refers to cell numbers and not weight. In any case, the total intraterrestrial 
biomass is very impressive: the calculations of Whitman et al. (1998) suggest that 
the intraterrestrial biomass may nearly equal the surface biomass. Whether these 
calculations are off by 50 % or so is immaterial, as there will still be a tremendous 
amount of life dwelling deep under our feet. 

SGM symposium 65 



240 K. Pedersen 



INTRATERRESTRIAL SPECIES DIVERSITY 

The cloning and sequencing of the 16S/18S rRNA genes of microbes living in their 
natural environments has revealed a genetic diversity exceeding any earlier estimates 
(Pace, 1997). This methodology has been applied to microbes from various under- 
ground sites and has generally revealed great species diversity. The case of the Fenno- 
scandian hard-rock aquifers will be used to demonstrate typical results for hard rock. A 
total of 385 clones from two sites were sequenced (Ekendahl et al., 1994; Pedersen et al., 
1996a, 1997); 122 unique sequences were found, each representing a possible species 
that was not recorded in international databases at the time the analysis was conducted. 
Therefore, on average, approximately one-third of the sequenced clones represented 
unique species. These studies clearly have not exhausted the sequences, as new 
sequences were found in nearly every additional sample repetition. This molecular 
work indicates that the deep, hard-rock groundwater environments studied are 
inhabited by diverse microbial populations, consistent with the great variability of 
hydrogeochemical conditions. Similar results were obtained with DNA sequences 
of micro-organisms from the alkaline springs of Maqarin in Jordan (Pedersen et al., 
2004), natural nuclear reactors in Oklo, Gabon (Pedersen et al., 1996b), and in nuclear- 
waste buffer material (Stroes-Gascoyne et al., 1997). 

The molecular work described above has provided a good insight into the phylogenetic 
diversity of hard-rock aquifer micro-organisms but does not reveal species-specific 
information unless 100 % identity of the full 16S rRNA gene sequence of a known and 
described micro-organism is obtained. The huge diversity of the microbial world makes 
the probability of such a hit very small - none of the 122 specific sequences mentioned 
above had 100 % identities with described species. Even if a 100 % identity is obtained, 
there may yet be strain-specific differences in some characteristics which are not 
revealed by the 16S rRNA gene sequence information (Fuhrman & Campbell, 1998). If 
species information is required, time-consuming methods in systematic microbiology 
must be applied to a pure culture; known genera or species can be identified through 
these methods. 

Sequencing 16S/18S rRNA genes is a fairly blunt tool for obtaining an understanding 
of microbial activity and metabolic diversity. Therefore, our laboratory has been apply- 
ing cultivation methods for a long time, concurrent with molecular methods. The most 
probable numbers of a range of physiologically different micro-organisms have been 
analysed (Haveman et al., 1999; Haveman & Pedersen, 2002); these micro-organisms 
are iron-reducing bacteria, manganese-reducing bacteria, sulfate-reducing bacteria, 
autotrophic methanogens, heterotrophic methanogens, autotrophic acetogens, hetero- 
trophic acetogens and methanotrophic bacteria. All those groups have been found in 
hard-rock groundwater. In addition, unicellular eukaryotes, in particular fungi, have 

SGM symposium 65 



The deep intraterrestrial biosphere 241 




Fig. 3. The deep hydrogen-driven biosphere hypothesis, illustrated by its carbon cycle. Under 
relevant temperature and water-availability conditions, intraterrestrial micro-organisms are capable 
of performing a life cycle that is independent of solar-driven ecosystems. Hydrogen and carbon 
dioxide from the deep crust of the Earth are used as energy and carbon sources. 

been detected (Ekendahl et al., 2003). Phages are probably present but remain to be 
detected. 

ENERGY SOURCES FOR INTRATERRESTRIAL LIFE 

Where do all these organisms get their energy? Hydrogen may be an important electron 
and energy source and carbon dioxide an important carbon source in deep subsurface 
ecosystems. Hydrogen, methane and carbon dioxide have been found in micromolar 
concentrations at all sites tested for these gases (Sherwood Lollar et al., 1993a, b; 
Pedersen, 2001). A model has been proposed (Fig. 3) of a hydrogen-driven biosphere 
in deep igneous rock aquifers of the Fennoscandian Shield (Pedersen, 1993b, 2000). The 
organisms at the base of this ecosystem are assumed to be autotrophic acetogens 
capable of reacting hydrogen with carbon dioxide to produce acetate, autotrophic 
methanogens that produce methane from hydrogen and carbon dioxide and aceto- 
clastic methanogens that produce methane from the acetate product of the autotrophic 
acetogens. This hydrogen-driven, intraterrestrial biosphere should conceptually be 
possible anywhere underground or under the seafloor where hydrogen and carbon 
dioxide are available. The cycle produces acetate; this is a very versatile source of energy 
and carbon for anaerobic microbes, such as iron- and sulfate-reducing bacteria and 
acetoclastic methanogens. All components needed for this life cycle have been found 



SGM symposium 65 



242 K. Pedersen 



Hydrolysis 




Aerobic bacteria 



Denitrifying 
bacteria 



'Manganese- 
reducing' 
bacteria - 



'Iron-reducing' 
bacteria 



'Sulfate- 

reducing' 

bacteria 

Sulfur- 
reducing bacteria 



Methanogens 



Acetogenic 
bacteria 



Q 2 H 2 Q 



NO- 3 N 2 



Mn** Mn 2 * 




CO- 



CO- 



CO- 



Fe 3+ Fe 3+ 




CO- 



CH 



4 




H 2 + C0 2 



Acetate 



Fig. 4. The degradation of organic carbon can occur via a number of different metabolic pathways, 
characterized by the principal electron acceptor in the carbon oxidation reaction. A range of 
significant groundwater compounds are formed or consumed during this process. The degradation 
follows a typical redox ladder pattern. 

in the deep igneous rock aquifers and the expected microbial activities have been 
demonstrated (Pedersen, 2001), so the model is supported by the qualitative data 
obtained so far. Quantitative data are currently being obtained at the underground 
laboratory (Fig. 2). 

ACTIVITY OF INTRATERRESTRIAL LIFE FORMS 

How much do intraterrestrial organisms depend on their environment and to what 
extent can they influence their environment? The organisms may have little to do with 
hard rock formation but they certainly take part in the formation of less violently 
formed, sedimentary rock types. Once the rock is formed, there are still many processes 
that can be influenced by microbes and their activities. Precipitation and dissolution, 
weathering of rocks and cycling of nitrogen, carbon, phosphorus and many metals can 
be influenced by microbes to various degrees. Such processes generally require active 
microbes, and gaining an understanding of activity generally requires in situ and/or 
laboratory cultivation of the organisms of interest. Radiotracers can be used (Pedersen 
& Ekendahl, 1992a, b; Ekendahl & Pedersen, 1994), and there is also the possibility of 
analysing stable isotope fractionations using substrates added by nature (Des Marais, 
1999). 



SGM symposium 65 



The deep intraterrestrial biosphere 243 

Another approach to exploring intraterrestrial micro-organisms is to study and 
interpret geochemical data. For example, why is there no oxygen in deep ground- 
water? Oxygen is the preferred electron acceptor of most microbes and its source is 
photosynthesis, a surface-based process. Microbes tend to degrade organic carbon 
down the redox ladder (Fig. 4), and oxygen is the first electron acceptor to be utilized. 
As rainwater percolates down through the ground to become groundwater, it carries 
organic carbon and oxygen. Microbes degrade the organic substances along the ladder 
in an ongoing process (Banwart et ai, 1996). Oxygen thus does not penetrate very far 
underground (although there will be spots where this happens), so the intraterrestrial 
world is thus anaerobic. The intriguing thing about this situation is that the under- 
ground environment is obviously kept anaerobic by microbes. Life has changed the 
surface of our planet dramatically by rilling it with oxygen via photosynthesis, and it 
seems likely that life is also responsible for keeping oxygen out of the intraterrestrial 
biosphere. The intraterrestrial biosphere thus reflects times on our planet when 
photosynthesis had not yet had an effect. 

REFERENCES 

Amend, J. P. & Teske, A. (2005). Expanding frontiers in deep subsurface microbiology. 

Palaeogeogr Palaeoclimatol Palaeoecol 219,131-155. 
Bachofen, R. (editor) (1997). Proceedings of the 1996 International Symposium on 

Subsurface Microbiology (ISSM-96). Davos, Switzerland, 15-21 September 1996. 

FEM5 Microbiol Rev 20, 179-638. 
Bachofen, R., Ferloni, P. & Flynn, I. (1998). Microorganisms in the subsurface. Microbiol 

Res '\53 l 1-22. 
Banwart, S., Tullborg, E.-L, Pedersen, K., Gustafsson, E., Laaksoharju, M., Nilsson, 

A.-C, Wallin, B. & Wikberg, P. (1996). Organic carbon oxidation induced by large- 
scale shallow water intrusion into a vertical fracture zone at the Aspo Hard Rock 

Laboratory (Sweden). J Contam Hydrol2'\ l 115-125. 
Butler, R. (1994). Drilling deep holes in the continents. Geo! Today 10, 32-35. 
Characklis, W. G. & Marshall, K. C. (editors) (1990). Biofilms. New York: Wiley. 
Des Marais, D. J. (1999). Stable light isotope biogeochemistry of hydrothermal systems. 

In Evolution of Hydrothermal Ecosystems on Earth (and Mars?), pp. 83-98. Edited 

by G. R. Bock & J.A. Goode. Chichester: Wiley. 
Ekendahl, S. & Pedersen, K. (1994). Carbon transformations by attached bacterial 

populations in granitic groundwater from deep crystalline bed-rock of the Stripa 

research mine. Microbiology 140, 1 565-1 573. 
Ekendahl, S., Arlinger, J., Stahl, F. & Pedersen, K. (1994). Characterization of attached 

bacterial populations in deep granitic groundwater from the Stripa research mine 

with 16S rRNA gene sequencing technique and scanning electron microscopy. 

Microbiology 140, 1 575-1 583. 
Ekendahl, S., O'Neill, A. H., Thomsson, E. & Pedersen, K. (2003). Characterisation of 

yeasts isolated from deep igneous rock aquifers of the Fennoscandian Shield. Microb 

Ecol 46, 41 6^28. 



SGM symposium 65 



244 K. Pedersen 



Fredrickson, J. K. & Fletcher, M. (editors) (2001). Subsurface Microbiology and 

Biogeochemistry. New York: Wiley-Liss. 
Fuhrman, J. A. & Campbell, L (1998). Microbial microdiversity. Nature 393, 41 0. 
Ghiorse, W. C. & Wilson, J. T. (1988). Microbial ecology of the terrestrial subsurface. 

AdvAppI Microbiol 33, 1 07-1 72 . 
Gold, T. (1992). The deep, hot biosphere. Proc Natl Acad Sci U 5 A 89, 6045-6049. 
Haveman, S. A. & Pedersen, K. (2002). Microbially mediated redox processes in natural 

analogues for radioactive waste. J Contam Hydro! 55, 161-1 74. 
Haveman, S. A., Pedersen, K. & Routsalainen, P. (1999). Distribution and metabolic 

diversity of microorganisms in deep igneous rock aquifers of Finland. Geomicrobiol 

716,277-294. 
Laaksoharju, M., Skarman, C. & Skarman, E. (1999). Multivariate mixing and mass 

balance (M3) calculations, a new tool for decoding hydrogeochemical information. 

Appl Geochem 14, 861-871 . 
Onstott, T. C, Phelps, T. J., Colwell, F. S., Ringelberg, D., White, D. C. & Boone, D. R. 

(1998). Observations pertaining to the origin and ecology of microorganisms 

recovered from the deep subsurface of Taylorsville Basin, Virginia. Geomicrobiol J 15, 

353-383. 
Pace, N. R. (1997). A molecular view of microbial diversity and the biosphere. Science 276, 

734-740. 
Pedersen, K. (1993a). The deep subterranean biosphere. Earth Sci Rev 34, 243-260. 
Pedersen, K. (1993b). Bacterial processes in nuclear waste disposal. Microbiol Eur 1, 18- 

23. 
Pedersen, K. (2000). Exploration of deep intraterrestrial life: current perspectives. FEMS 

Microbiol Lett 1 85, 9-1 6. 
Pedersen, K. (2001). Diversity and activity of microorganisms in deep igneous rock aquifers 

of the Fennoscandian Shield. In Subsurface Microbiology and Biogeochemistry, 

pp. 97-1 39. Edited by J. K. Fredrickson & M. Fletcher. New York: Wiley-Liss Inc. 
Pedersen, K. & Ekendahl, S. (1992a). Incorporation of C0 2 and introduced organic 

compounds by bacterial populations in groundwater from the deep crystalline 

bedrock of the Stripa mine. J Gen Microbiol 1 38, 369-376. 
Pedersen, K. & Ekendahl, S. (1992b). Assimilation of C0 2 and introduced organic 

compounds by bacterial communities in ground water from Southeastern Sweden 

deep crystalline bedrock. Microb Ecol 23, 1-14. 
Pedersen, K., Arlinger, J., Ekendahl, S. & Hallbeck, L (1 996a). 1 6S rRNA gene diversity of 

attached and unattached bacteria in boreholes along the access tunnel to the Aspo 

hard rock laboratory, Sweden. FEMS Microbiol Ecol 19, 249-262. 
Pedersen, K., Arlinger, J., Hallbeck, L. & Pettersson, C. (1996b). Diversity and distribution 

of subterranean bacteria in groundwater at Oklo in Gabon, Africa, as determined by 

1 6S RNA gene sequencing. Mol Ecol 5, 427-436. 
Pedersen, K., Hallbeck, L, Arlinger, J., Erlandson, A.-C. & Jahromi, N. (1997). 

Investigation of the potential for microbial contamination of deep granitic aquifers 

during drilling using 16S rRNA gene sequencing and culturing methods. J Microbiol 

Methods 30, 179-192. 
Pedersen, K., Nilsson, E., Arlinger, J., Hallbeck, L. & O'Neill, A. (2004). Distribution, 

diversity and activity of microorganisms in the hyper-alkaline spring waters of 

Maqarin in Jordan. ExtremophilesS, 1 51-164. 
Sherwood Lollar, B., Frape, S. K., Fritz, P., Macko, S. A., Welhan, J. A., Blomqvist, R. & 

Lahermo, P. W. (1993a). Evidence for bacterially generated hydrocarbon gas in 



SGM symposium 65 



The deep intraterrestrial biosphere 245 



Canadian Shield and Fennoscandian Shield rocks. Geochim Cosmochim Acta 57, 

5073-5085. 
Sherwood Lollar, B., Frape, S. K., Weise, S. M. ; Fritz, P., Macko, S. A. & Welhan, J. A. 

(1993b). Abiogenic methanogenesis in crystalline rocks. Geochim Cosmochim Acta 

57, 5087-5097. 
Stroes-Gascoyne, S., Pedersen, K. f Haveman, S. A. & 8 other authors (1997). 

Occurrence and identification of microorganisms in compacted clay-based buffer 

material designed for use in a nuclear fuel waste disposal vault. Can J Microbiol 43, 

1133-1146. 
Szewzyk, U. ; Szewzyk, R. & Stenstrom, T. A. (1994). Thermophilic, anaerobic bacteria 

isolated from a deep borehole in granite in Sweden. Proc Natl Acad Sci USA 91, 

1810-1813. 
Whitman, W. B., Coleman, D. C. &Wiebe, W. J. (1998). Prokaryotes: the unseen majority. 

Proc Natl Acad Sci USA 95, 6578-6583. 



SGM symposium 65 



Iron, nitrogen, phosphorus and 
zinc cycling and consequences for 
primary productivity in the oceans 

John A. Raven # 1 Karen Brown # 1 Maggie Mackay, 1 

John Beardall # 2 Mario Giordano, 3 Espen Granum, 4 

Richard C. Leegood, 4 Kieryn Kilminster 5 and Diana I. Walker 5 

1 Plant Research Unit, Division of Environmental and Applied Biology, School of Life Sciences, 
University of Dundee at SCRI, Scottish Crop Research Institute, Invergowrie, Dundee DD2 5DA, 
Scotland, UK 

2 School of Biological Sciences, Monash University, Clayton, VIC 3800, Australia 

department of Marine Science, Universita Politecnica delle Marche, 60131 Ancona, Italy 

department of Animal and Plant Sciences, University of Sheffield, Sheffield 510 2TN, UK 

5 School of Plant Biology, University of Western Australia, M090 35 Stirling Highway, Crawley, 
WA 6009, Australia 

INTRODUCTION 

Primary productivity in the ocean amounts to the net assimilation of C0 2 equivalent to 
about 50 Pg (petagram, i.e. 10 15 g) C year 1 , while on land this is approximately 60 Pg C 
year 1 (Field et al., 1998). Almost all of this primary productivity involves photo- 
synthesis, and in the ocean it occurs only in the top few hundred metres, even in waters 
with the smallest light attenuation (Falkowski & Raven, 1997). About 1 Pg C of marine 
primary productivity involves benthic organisms, i.e. those growing on the substratum 
(Field et al., 1998), in the very small fraction of the ocean which is close enough to the 
surface to permit adequate photosynthetically active radiation (PAR) to allow photo- 
lithotrophic growth. This depth at which photosynthetic growth is just possible varies 
in time and space, and defines the bottom of the euphotic zone (Falkowski & Raven, 
1997). The remaining ~49 Pg C is assimilated by phytoplankton in the water column 
(Field et al., 1998). This chapter will concentrate on the planktonic realm, while 
acknowledging the importance of marine benthic primary producers and their inter- 
actions with micro-organisms (e.g. Dudley et al., 2001; Raven et al., 2002; Raven 6c 
Taylor, 2003; Cooke et al., 2004; Walker et al., 2004). 

The global net primary productivity of the oceans is less than that on land, despite 
about 70 % of the Earth being covered in ocean and primary productivity over 
considerable areas of land being limited by water supply. In considering the role of 
nutrient elements such as iron, nitrogen, phosphorus and zinc in contributing to the 
restriction of primary productivity in the ocean, we must consider them in the context 
of other limitations on primary productivity 

SGM symposium 65: Micro-organisms and Earth systems - advances in geomicrobiology. 
Editors G. M. Gadd, K. T. Semple & H. M. Lappin-Scott. Cambridge University Press. ISBN 521 86222 1 ©SGM 2005 



248 J. A. Raven and others 



These constraints are frequently divided by oceanographers into 'bottom-up' (physical 
and chemical) factors and 'top-down' (biotic) factors. This division echoes that of 
Grime (1974), who separated the constraints on terrestrial primary productivity into 
'stress' (factors which constrain the production of primary producer biomass) and 
'disturbance' (factors which remove primary producer biomass). The bottom-up 
(stress) factors include insufficient or excessive PAR, excess UV-B radiation, too high or 
too low a temperature and insufficient supply of any of the chemical elements essential 
for growth and completion of the life cycle in a form which is available to the organisms. 
The top-down (disturbance) factors are the biotic factors of grazing and parasitism 
(including viral attack) and the physical factor of removal by, in the case of phyto- 
plankton, sinking out of the euphotic zone. 

The bottom-up factors determining the growth rate of marine primary producers are 
intimately related to physical oceanography and to global biogeochemical cycles. 
Physical oceanography is significant for phytoplankton photosynthesis through the 
depth of the upper, mixed layer of the ocean in which the phytoplankton organisms 
are entrained, as modified by movement of organisms relative to the surrounding water 
as a result of the motility of flagellate organisms and upward and downward movement 
of non-flagellate organisms whose density is, respectively, lesser or greater than that of 
the surrounding sea water. When the mixing depth is greater than the so-called 'critical 
depth', i.e. the depth at which the integrated water-column photosynthesis by phyto- 
plankton equals community respiration, net growth of phytoplankton cannot occur 
(Falkowski & Raven, 1997). Physical oceanography also impacts on the supply of 
nutrients to photosynthetic organisms with respect to cycling of nutrients within 
the ocean. Nutrients are carried to depth by the sinking of live and dead biota and 
faecal pellets, generically termed 'marine snow'. Below the euphotic zone, most of the 
particulate organic material is acted on by decomposer organisms which regenerate 
inorganic nutrients that can then be used by primary producers upon return to the 
euphotic zone. This return of nutrients to the euphotic zone can occur by seasonal 
and permanent upwellings and by eddy diffusion across the thermocline at the base 
of the upper mixed layer (Falkowski & Raven, 1997). 

With this background to marine primary productivity and its determinants, we will 
address the cycles of four nutrient elements, iron, nitrogen, phosphorus and zinc, whose 
supply has been shown at times to restrict marine primary production in some parts of 
the ocean. This chapter first addresses the biogeochemical cycles of these elements and 
the evidence as to their roles in limiting primary productivity, dealing with the elements 
in the order in which they are thought to be significant in constraining primary 
productivity. Secondly, we will consider the interactions between these elements 

SGM symposium 65 



Fe, N, P and Zn cycling in the ocean 249 

in determining primary productivity. Thirdly, we consider some broader aspects of 
elemental supply and marine primary productivity in a changing world. 



BIOGEOCHEMICAL CYCLES OF NITROGEN, IRON, PHOSPHORUS 
AND ZINC IN RELATION TO MARINE PRIMARY PRODUCTIVITY 

Nitrogen 

The abundant N 7 in the atmosphere, and dissolved in the ocean, is only directly avail- 
able to diazotrophic organisms, i.e. those that use the enzyme nitrogenase to reduce N 7 
to NFL, which can then be used to produce the organic N compounds needed by the 
organism. N 2 fixation by marine organisms involves mainly cyanobacteria, of which 
the filamentous Trichodesmium predominates, although the role of unicellular 
cyano bacterial diazotrophs has recently been recognized (Capone, 2001). Other inputs 
of combined N, i.e. nitrogen sources other than N 9 , which are available to non-diazo- 
trophic primary producers, occur as atmospheric inputs resulting from lightning 
converting N 2 to NO v (NO + N0 2 ) and hence HN0 3 , and increasingly as NH 
(NHL + NHJ, NO A ., HNO3 and organic N from natural and anthropogenic terrestrial 
sources. Combined N supply also occurs as fluvial inputs of NH4, NO3 and organic 
N from terrestrial ecosystems, again with increases as a result of man's activities 
(Falkowski 6c Raven, 1997; Tyrrell, 1999). The losses of combined N from the ocean 
occur as denitrification, whereby NO3 is reduced to N 9 and N 2 in anoxic and hypoxic 
regions of the ocean, and by the incorporation of particulate organic N into marine 
sediments (Falkowski & Raven, 1997; Tyrrell, 1999). While it has been suggested that 
there is a net loss of combined N from the present ocean (Falkowski & Raven, 1997), 
consideration of the cumulative errors in the estimation of the components of the 
combined N inputs and outputs for the present ocean suggests that the data are 
also consistent with a balance of oceanic combined N inputs and outputs (Tyrrell, 
1999). Even if there is no net loss of combined N from the ocean at present, there is 
strong evidence that primary productivity in a large fraction of the world ocean 
is limited by the supply of combined N (Falkowski, 1997; Falkowski & Raven, 1997; 
Cullen, 1999; Tyrrell, 1999). This evidence is derived from enrichment experiments in 
which samples of surface ocean water are incubated for the determination of primary 
productivity, either unaltered or with the addition of particular nutrients, individually 
and in combination. The results of such experiments in many parts of the ocean 
suggest that the nutrient element limiting growth is N, with diazotrophs unable to 
supply combined N at a sufficiently rapid rate. More generally, the atomic ratio of 
inorganic combined N to P in the surface ocean is, when averaged over time and space, 
rather less than the 16:1 Redfield ratio of the average phytoplankton cell with the 
elemental composition of 106C : 16N : IP (Falkowski & Raven, 1997; Falkowski, 2000). 

SGM symposium 65 



250 J. A. Raven and others 



There is a significant difference between the two scenarios for oceanic combined N 
status (Falkowski, 1997; Tyrrell, 1999) with respect to the possible constraints on 
biological N 2 fixation in the ocean. We deal first with the possibility that there is not a 
shortfall in biological fixation, i.e. it is adequate to maintain the oceanic combined N 
pool by supplying as much combined N as is needed to replace the excess of combined 
N loss by sedimentation and denitrification over aeolian and riverine inputs (Tyrrell, 
1999). If this is the case then it may not be necessary to seek exogenous limitations 
specific to diazotrophy to account for the magnitude of N 2 fixation other than optimal 
allocation arguments that the extra machinery required by diazotrophy restricts the 
maximum specific growth rate of N^-fixing organisms relative to those using combined 
N. This would permit the diazotrophs to grow whenever there was a shortfall in 
combined N supply which constrained the growth of non-diazotrophs, despite the 
lower maximum specific growth rates, with the non-diazotrophs ousting the slower- 
growing diazotrophs when combined N is more readily available. However, it is known 
that there are additional resource costs of diazotrophic growth relative to growth on 
combined N, such as the increased energy, Fe and Mo (or V) requirement and a less 
well-defined requirement for additional P (Sanudo-Wilhelmy et al, 2001; Vitousek 
et al, 2001; Kustka et al., 2003a, b). These constraints on the growth of photosynthetic 
diazotrophs relative to that of photosynthetic organisms living on combined N could 
account for any inadequacy of N 2 fixation to maintain the oceanic combined N pool 
by not allowing the diazotrophs to fix N 9 until there was enough combined N to allow 
non-diazotrophs to out-compete diazotrophs (Falkowski, 1997). The observation is 
that Fe is commonly a limiting resource for diazotrophy in the surface ocean, although 
P and PAR are also significant constraints in some areas (Falkowski, 1997; Sanudo- 
Wilhelmy et al, 2001 ; Mills et al, 2004) . 

The combined N sources for phytoplankton growth in the ocean include nitrate, nitrite, 
ammonium and organic N (Falkowski 6c Raven, 1997). As we have noted, nitrate (and 
nitrite) enters the surface ocean by fluvial and aeolian inputs, as well as in upwellings 
and eddy diffusion from the deep ocean, where it is produced by nitrification of reduced 
N sedimented from the euphotic zone. This 'new' (to the euphotic zone) combined N 
supports what is termed 'new' primary production which, since it can be exported 
(sedimented) without altering the pre-existing combined N pool in the surface ocean, is 
also termed 'export production'. Ammonium/ammonia and organic N supply to 
the surface ocean is predominantly from excretion by higher trophic levels as well as 
from parasitism of phytoplankton, with only a minor component entering the surface 
ocean from aeolian and fluvial inputs. The primary productivity supported by the use 
of this reduced N is termed 'recycled' production, since it depends very largely on 
combined N recycled from particulate organic matter within the surface ocean. The 
reason that this 'recycled' production uses reduced combined N directly, rather than 

SGM symposium 65 



Fe, N, P and Zn cycling in the ocean 251 

after it has been converted to the thermodynamically more stable (in the oxygenated 
surface ocean) nitrate, seems to be the kinetic constraint on the exergonic conversion of 
ammonium to nitrate by nitrifying bacteria resulting from light inhibition of growth 
of the nitrifiers (Falkowski Sc Raven, 1997). These nitrifiers are chemolithotrophs which 
can assimilate about 0-19 Pg C from inorganic C per year. The energy used by these 
organisms can be regarded as the additional energy, ultimately derived from PAR, 
required for phytoplankton growth on nitrate rather than on reduced N in the euphotic 
zone. This energy is used to power inorganic C reduction deeper in the ocean (Raven, 
1996). The expected lower growth rate of photosynthetic primary producers under 
light-limiting conditions resulting from the use of nitrate rather than reduced N as the 
N source is, however, not found routinely (Raven, 1996). 

Most marine primary producers examined are able to use all of the combined N forms 
discussed above, i.e. ammonium/ammonia, one or more species of organic N, nitrite 
and nitrate (Falkowski & Raven, 1997). However, the very abundant picoplanktonic 
cyanobacterium Procblorococcus is unable to use nitrate, and some strains are also 
unable to use nitrite (Bryant, 2003; Dufresne et al., 2003; Palenik et al., 2003; Rocap 
et al., 2003). This relatively recent loss of metabolic capability during the evolution of 
Procblorococcus from within the paraphyletic genus Synecbococcus is clearly not 
a fatal impediment to Procblorococcus, which is probably the most abundant, in terms 
of number of individuals, photosynthetic organism on Earth, as well as being amongst 
the smallest (Bryant, 2003). Procblorococcus is relatively most successful in the oligo- 
trophic ocean, where it depends mainly on recycled combined N. It also typically occurs 
deeper in the water column than the metabolically more versatile Synecbococcus, where 
the energy-requiring assimilation of oxidized combined N would consume scarce 
energy in this light-limited environment (Ting et al., 2002; cf. Raven, 1987; MacFarlane 
& Raven, 1990; Falkowski &C Raven, 1997). As we shall see when considering Fe, 
the absence of nitrate assimilation capacity reduces the Fe requirement of the cells. 

A final aspect of the N nutrition of marine primary producers is the possibility of 
economizing on N by replacing a macromolecule with a high N content (usually a 
protein) with a macromolecule with a lower N content that performs a similar meta- 
bolic function (Raven et al., 2004). Exact functional substitutions are uncommon, 
making the effectiveness of this possible means of economizing on N difficult to 
evaluate. An example is in photosynthetic light harvesting, where the quantity of 
apoprotein per molecule of chromophore is at least twice as high for phycobiliproteins 
as for chlorophyll-protein complexes. However, differences in chromophore absorption 
spectra and in chromophore-specific absorption coefficients make a like-for-like 
comparison of the ecophysiological impact of such a substitution difficult to evaluate. 
Perhaps the best-examined case is that of the almost total replacement of phyco- 

SGM symposium 65 



252 J. A. Raven and others 



biliproteins by chlorophyll-protein complexes in Prochlorococcus relative to the 
ancestral Synecbococcus (Ting et al., 2002). Such substitutions can reduce the cell 
N requirement by over 10 % (Raven et al., 2004). Similar magnitudes of economy in N 
costs of growth can be achieved by replacing N-containing with N-free low relative 
molecular mass organic compounds involved in screening UV-B, scavenging and 
quenching reactive oxygen species, providing osmotic compensation as compatible 
solutes and restricting the attentions of grazers and parasites using anti-biophage 
compounds (Raven et al., 2004). Again, there are problems in establishing the 
effectiveness of the N-free compound relative to the N-containing compound in 
the organism under natural conditions. 

Iron 

Fe input to the surface ocean occurs as dissolved and particulate fluxes down rivers, 
much of which is sedimented in the coastal zone, and as aeolian inputs of dust, either 
dry or in rain (Berner & Berner, 1996; Chester, 2003; Jickells et al., 2005). There is also 
some input of Fe to the deep ocean via hydro thermal vents, while the major output of Fe 
from the ocean is in sinking particles incorporated into marine sediments (Chester, 
2003). Although there is some input of Fe into the surface ocean from the deep ocean 
by upwelling and eddy diffusion, the major inputs of Fe to the ocean surface are 
from rivers and, more importantly, from the atmosphere. The atmosphere supplies 
essentially all of the Fe inputs to the non-coastal surface ocean (Chester, 2003). 

Fe is a very abundant element on Earth, but the global oxygenation which began over 
2 billion years ago has restricted the availability of Fe to organisms by converting it into 
the insoluble ferric form. While chemical reduction of ferric to ferrous iron can occur 
in acidic atmospheric droplets, this soluble ferrous iron is converted to the ferric form 
with a half-time of about 2 min when the droplet equilibrates with sea water at a pH of 
about 8 (Chester, 2003). The requirement for large quantities of Fe in, for example, 
nitrogenase and nitrogenase reductase, photosystem I-like photosynthetic reaction 
centres, NADH dehydrogenase complexes and nitrite reductase (Raven, 1988, 1990; 
Raven et al., 1999) presumably evolved when Fe was much more readily available, more 
than 2-5 billion years ago. Squaring the large Fe requirements in these catalysts with 
the current relative unavailability of Fe is now met by a range of Fe-acquisition 
mechanisms, as well as a range of means by which the use of Fe is restricted. 

Fe acquisition from sea water involves a variety of mechanisms (Volker 6c Wolf- 
Gladrow, 1999). Cyanobacteria use siderophores which are secreted by the organism, 
and some of the molecules are taken up again after they have acquired Fe(III) from 
colloidal or ligated sources (Chester, 2003). Some eukaryotic marine primary producers 
can take up Fe(III)-loaded siderophores, although they do not produce the sidero- 

SGM symposium 65 



Fe, N, P and Zn cycling in the ocean 253 

phores (Chester, 2003). The genome of the diatom Thalassiosira pseudonana gives no 
indication of the production or uptake of siderophores, but has components of the 
mechanism which reduces external Fe(III) to Fe(II) followed by uptake of the Fe(II) by a 
process involving, paradoxically, oxidation to Fe(III) during uptake (Armbrust et al., 
2004). This mechanism is widespread in marine primary producers, using Fe(III) from a 
range of ligands (Falkowski &C Raven, 1997). A few marine planktonic primary 
producers are also phagotrophic, and so are able to take up colloidal or particulate 
Fe (Raven, 1997). 

The constraints on Fe availability to primary producers have evoked a number of 
ecological and evolutionary responses. One response is the restricted occurrence 
of diazotrophy and, in some habitats, of the use of nitrate and nitrite as N sources 
(Falkowski, 1997; Flynn & Hipkin, 1999; Sanudo-Wilhelmy et ai, 2001; Kustka et ai, 
2003a, b; Mills et al., 2004). Another response is the decrease in the Fe requirement for 
photosystem I on a cell basis in genotypes with low contents of this kind of reaction 
centre. Examples of low photosystem I contents include the cyanobacterium Prochloro- 
coccus relative to its ancestor Synecbococcus, and the open-ocean relative to the coastal 
species of the diatom Thalassiosira (Strzepek & Harrison, 2004). In at least some cases, 
the decreased content of photosystem I constrains acclimation to environments with 
varying fluxes of PAR (Strzepek 6c Harrison, 2004). These economies of N use, by 
changing the exogenous N source or by decreasing the content of photosystem I, can 
account for several-fold reduction in Fe requirement from the most costly to the least 
costly resource-acquisition mechanisms (Sunda et ai, 1991; Sanudo-Wilhelmy et ai, 
2001; Kustka et ai, 2003a, b; Strzepek & Harrison, 2004; cf. Falkowski et al, 2004). 
Much smaller economies can be effected by substituting an Fe-containing component 
which accounts for a relatively small fraction of the total cell N by an Fe-free 
component which performs essentially the same metabolic function. Examples here are 
the replacement of the Fe-containing photosynthetic cytochrome c 6 with the Cu- 
containing plastocyanin and of the Fe-containing ferredoxin with the metal-free 
flavodoxin (Raven et al., 1999). Here, near-equality of function has been established 
(Raven et al., 1999). Such replacements can be facultative (in response to Fe deficiency) 
or obligate, e.g. in the 'red' line of plastid evolution where plastocyanin is absent (Raven 
et al., 1999; Falkowski et al., 2004). Although the economy in Fe content of the 
organisms by these substitutions is small, the expression of flavodoxin in natural 
phytoplankton assemblages can be used as an indicator of actual, or incipient, Fe 
limitation of growth. 

Despite the theoretical likelihood of Fe limitation of primary productivity in at least a 
part of the surface ocean, problems with the analysis of Fe in sea water, and in carrying 
out incubations under conditions free of Fe contamination, meant that evidence 

SGM symposium 65 



254 J. A. Raven and others 



consistent with Fe limitation in significant areas of the ocean only became available 
a little over a quarter of a century ago with the insightful work of the late John Martin 
and his colleagues (Martin & Fitzwater, 1988; Martin et ai, 1990, 1991). Martin's 
conclusions were based on the chemical analysis of sea water and on the sort of incu- 
bation experiments described above for N, involving measurements of primary 
productivity of sea-water samples with and without the addition of a range of 
nutrients. This work showed that the Fe limitation of primary production occurred 
in the so-called 'high-nutrient, low-chlorophyll' (HNLC) areas of the ocean, i.e. those 
parts of the ocean that have relatively high concentrations of the major nutrients 
nitrate and phosphate in the surface waters, but a low biomass of primary producers 
as indicated by the chlorophyll content (Chester, 2003). The HNLC areas of the ocean 
include the north-east sub-Arctic Pacific, the eastern tropical Pacific and the Southern 
Ocean (Chester, 2003; Jickells et ai, 2005). There are also smaller coastal areas of 
Fe limitation in the eastern Pacific (Hutchins et ai, 1998, 2002; Hutchins & Bruland, 
1998; Jickells £?*#/., 2005). 



The incubation experiments described above did not rule out the possibility of the 
in situ occurrence of deep mixing, leading to light limitation of primary production, as 
a bottom-up factor, and/or of the impact of grazers or parasites in removing primary 
producer biomass as a top-down factor. Clarification of the situation came from meso- 
scale experiments in which Fe (as iron sulfate) was added to several square kilometres of 
the surface ocean in HNLC areas and the impact of the addition on the chemistry and 
biology of the Fe-enriched patch was followed for as long as was possible; observations 
were usually terminated by the subduction or horizontal fragmentation of the patch 
(Chester, 2003). These experiments have now been carried out more than once in each 
of the three main HNLC areas of the ocean and, in each case, a decrease in nitrate and 
phosphate and an increase in primary producer biomass (especially large diatoms) 
and in primary productivity was found, as well as a delayed increase in the biomass of 
herbivorous zooplankton (Watson et al., 2000; Boyd et ai, 2002, 2004; Boyd, 2002a, b), 
although light limitation can also contribute to the HNLC state (Boyd, 2002a; van 
Oijen et al., 2004). However, it has not yet been established whether the increased 
primary productivity is paralleled by an increased export flux of particulate organic 
matter, as might be expected from the additional primary productivity resulting from 
the exogenous input of the productivity-limiting resource (Boyd et al., 2004). It is of 
interest that the shortage of Fe does not alter the ratio of C to N in phytoplankton cells, 
so that there seems to be no preferential allocation of what Fe is available to either the 
C assimilation or the N assimilation pathways. This absence of preferential allocation 
of Fe could well be a function of homoeostasis of the C to N ratio. Readers are directed 
to Cooke et al. (2004) and Raven et al. (2004) for further information. 

SGM symposium 65 



Fe, N, P and Zn cycling in the ocean 255 

For diazotrophy in the surface ocean, we have seen in the consideration of N above 
that Fe is a common limiting resource for N 2 fixation, although there are parts of the 
ocean in which P or light energy can limit diazotrophy (Falkowski, 1997; Sanudo- 
Wilhelmy et al., 2001; Mills et al., 2004). As with phytoplankton assimilating combined 
N, the diazotroph Trichodesmium does not exhibit variations in the cellular C : N ratio 
as a function of Fe availability (see references in Cooke et al., 2004; Raven et al., 2004). 

Phosphorus 

Almost all of the biology of P occurs at the most oxidized state of phosphate. Phos- 
phate is supplied to the ocean mainly via rivers, although there is some atmospheric 
input (Berner 6c Berner, 1996; Chester, 2003). However, atmospheric input of Fe can 
sequester some of the P in surface water as complexes with ferric iron (Chester, 2003), 
recalling the decrease in Fe availability with the onset of atmospheric oxygenation more 
than 2 billion years ago (Bjerrum & Canfield, 2002). P loss from the ocean is in sinking 
particles which become incorporated into marine sediments (Bjerrum & Canfield, 
2002; Chester, 2003). While there has been much emphasis on inorganic phosphate as 
the dominant source for marine primary producers, there is an increasing appreciation 
of the role of soluble organic phosphates in the P nutrition of marine primary 
producers. Generally, organic P sources are used by uptake of inorganic phosphates 
after the action of extracellular phosphatases. Those marine planktonic primary 
producers which are also phagotrophic can acquire P by ingestion of particles, with 
absorption of digestion products across the food vacuole membrane, as well as by 
transporters in the plasmalemma (Raven, 1997). 

There are cogent geochemical arguments that the supply of P, rather than of N, is 
the long-term determinant of the productivity of Earth, since the supply of P (in the 
absence of man's activities) is restricted by the rate of weathering on land and is 
balanced by sedimentation in the ocean, whereas N supply can be maintained by 
diazotrophy in the face of sedimentation and denitrification (Falkowski 6c Raven, 
1997). However, as we have seen, there are factors in addition to the availability of P that 
restrict the extent of diazotrophy in the ocean (Falkowski, 1997; Tyrrell, 1999; Sanudo- 
Wilhelmy et al., 2001; Mills et al., 2004). Nevertheless, there are regions of the ocean 
in which there is evidence from measurements of nutrient concentrations in the surface 
ocean, and from nutrient enrichment studies on primary productivity of samples 
of surface water, consistent with P limitation of primary productivity (Karl et al., 
1993; Karl & Tien, 1997; Cullen, 1999; Wu et al., 2000; Bemtez-Nelson & Karl, 2002; 
Ridame & Guieu, 2002; Bjorkman & Karl, 2003; Heldal et al., 2003; Krom et al., 2004). 
Particular attention has been paid to P limitation in parts of the Mediterranean Sea 
(Krom etal., 2004). 

SGM symposium 65 



256 J. A. Raven and others 



The possibilities of economy in the use of P in marine primary producers by sub- 
stitution of another, more readily available, element for P while maintaining equivalent 
metabolic function are smaller for P than for N or, especially, Fe. As for reducing 
the cell quota of a P-containing component, much of the functional (not-stored) P in 
most marine primary producers is contained in rRNA. In many non-photosynthetic 
organisms, there is a positive correlation of growth rate and rRNA content even when 
growth of a given genotype is constrained by environmental factors other than P supply, 
or when growth stages of a metazoan are compared, or closely related taxa are 
compared (Sterner & Elser, 2002; Raven et al., 2004, 2005). This correlation is much less 
clear for photosynthetic organisms, including the marine primary producers (Sterner 

o 

& Elser, 2002; Elser et al., 2003; Agren, 2004; Klausmeier et al., 2004a, b; Leonardos & 
Geider, 2004; Raven et al., 2004, 2005). However, the finite rate of protein synthesis 
across each rRNA molecule means that there must be a critical rRNA cell quota below 
which growth rate is constrained by the rRNA content, although the discussion above 
suggests that the rRNA quota in adequately P-nourished marine primary producers 
is significantly above this critical level. This apparent excess of rRNA means that a 
potential P economy measure in marine primary producers has not been widely 
adopted in these organisms. 

The genome is a P-containing cell component that is not generally a large fraction of 
the non-storage cell P. However, in the very small marine primary producer Prochloro- 
coccus growing under moderate P deficiency, the genome contains over half of the cell 
P quota (Bertilsson et al., 2003). This is a function of the non-scalable nature of the 
components of the genome: the size of a gene encoding a given protein is very similar in 
very small cells and in larger cells. Thus, despite having as few as 1500 protein-coding 
genes and a high gene density in the genome, the very small size of Prochlorococcus 
means that even its miniaturized genome contains a very significant fraction of the total 
cell P. The discussion above on N sources for Prochlorococcus shows that there are 
constraints on the nutrition of this organism as a function of gene loss, and some 
strains have a restricted ability to acclimate to varying light environments. 

To a limited extent, the large fraction of cell P in the genome of Prochlorococcus can be 
attributed to the small cell quota of P relative to C and N in well-nourished cells 
(Bertilsson et al., 2003; Heldal et al., 2003). Although it is tempting to relate these high 
C:P and N:P ratios to the oligotrophic and potentially P-limited habitat of 
Prochlorococcus, it must be remembered that the low P content in comparison with the 
Redfield ratio of 106C : 16N : IP (by atoms) is not exceptional, granted the variability in 
the C :P and N : P ratios of marine phytoplankton cultured under nutrient-replete 
conditions. The Redfield ratio applies to temporal and spatial averages over the world 
ocean (Falkowski & Raven, 1997; Falkowski, 2000; Geider & La Roche, 2002). A 

SGM symposium 65 



Fe, N, P and Zn cycling in the ocean 257 

further complication in assessing the P quota of marine phytoplankton is the 
occurrence, in some organisms, of significant (relative to the intracellular content) 
surface-bound P (Sanudo-Wilhelmy et al., 2004). 

Zinc 

Zn enters the ocean by dry and, especially, wet deposition from the atmosphere, as well 
as in rivers, and leaves it by sedimentation (Chester, 2003). Evidence consistent with Zn 
limitation in parts of the ocean comes from the work of Crawford et al. (2003) and 
Franck et al. (2003) using enrichments with Fe, Zn or Fe + Zn in experiments on 
primary productivity, or components thereof, using samples of surface ocean water. 
However, no drawdown of Zn was found during the stimulation of primary product- 
ivity in HNLC regions by the addition of Fe, suggesting that Zn was not close to 
growth-limiting in these surface waters (Frew et al., 2001; Gall et al., 2001). There is 
evidence consistent with Zn limitation of phytoplankton growth in some freshwater 
habitats (Ellwood et al., 2001 ; Sterner et al., 2004) . 

The enzymes that use Zn in marine primary producers include carbonic anhydrase, 
alkaline phosphatase and, in seagrasses and in peridinin-containing dinoflagellates, 
Cu-Zn superoxide dismutase (Raven et al., 1999). Most marine primary producers have 
only the Fe/Mn superoxide dismutases (Raven et al., 1999). A possible rationale for the 
lack of Cu-Zn superoxide dismutase in most marine primary producers is the very 
widespread occurrence of inorganic carbon-concentrating mechanisms (CCMs) in 
these organisms which, in many cases, involve an accumulation of bicarbonate ions to 
concentrations in excess of those in sea water. Cu-Zn superoxide dismutases catalyse, 
in addition to the eponymous dismutation of superoxide to hydrogen peroxide and 
oxygen, a bicarbonate-mediated peroxidation activity which can damage the enzyme 
itself or, by reacting with exogenous substrates, protect the enzyme from inactivation 
(see Elam et al., 2003). Perhaps the absence of Cu-Zn superoxide dismutases in these 
organisms with CCMs is related to the increased extent of self-oxidation and damage 
of the enzyme as a function of increasing concentrations of bicarbonate. It is of interest 
that CCMs in terrestrial vascular plants, i.e. those in plants with C 4 and 'crassulacean 
acid metabolism' (CAM), and with Cu-Zn superoxide dismutases in their cytosol and 
stroma, accumulate C0 2 rather than bicarbonate (Giordano et al., 2005). 

The quantitatively most significant role of Zn in most marine primary producers is in 
carbonic anhydrases, although Co or Cd can substitute for Zn in vitro and, in some 
cases, in vivo (Giordano et al., 2005). There is also a Cd-specific carbonic anhydrase 
in the diatom Thalassiosira weissflogii (Lane & Morel, 2000b; Giordano et al., 2005). 
Carbonic anhydrases are involved, inter alia, in CCMs (Giordano et al., 2003, 2005). 
These CCMs are frequently expressed more strongly at low inorganic C concentrations 

SGM symposium 65 



258 J. A. Raven and others 



(Giordano et al., 2005), and Morel et al. (1994) and Lane & Morel (2000a) have shown 
additional possibilities of Zn limitation at low inorganic C concentrations where 
carbonic anhydrase is expressed at higher levels. Subsequently, Reinfelder et al. (2000, 
2004) and Morel et al. (2002) have offered evidence consistent with C 4 -like photo- 
synthetic metabolism in the planktonic marine diatom T. weissflogii, especially at low 
inorganic C levels, which would decrease the requirement for Zn. The argument here is 
that, if bicarbonate is the inorganic C form entering the cytosol, no carbonic anhydrase 
activity is needed in that compartment because bicarbonate is the inorganic C substrate 
for phosphoenolpyruvate carboxylase, the enzyme which generates the C 4 dicarboxylic 
acid oxaloacetate (Reinfelder et al., 2000, 2004; Morel et al., 2002). Decarboxylation of 
oxaloacetate by phosphoenolpyruvate carboxykinase in the chloroplast stroma 
generates C0 2 , the substrate for the core carboxylase of photosynthesis, ribulose 
bisphosphate carboxylase-oxygenase (Rubisco), which also occurs in the stroma 
(Reinfelder et al., 2000, 2004; Morel et al., 2002). This scheme for inorganic C 
assimilation does not require expression of carbonic anhydrase in the cytosol or the 
stroma, or on the cell surface, since inorganic C enters as bicarbonate, the predominant 
form of inorganic C in sea water. Although the evidence for C 4 metabolism as an 
obligatory component of photosynthesis in T. weissflogiiis incomplete (Johnston et al., 
2001; Granum et al., 2005; Giordano et al., 2005), the case is becoming stronger 
(Reinfelder et al., 2004). Such a mechanism does not necessitate the accumulation of 
bicarbonate in intracellular compartments, so that the argument suggested above for 
the absence of Cu-Zn superoxide dismutases from organisms which accumulate 
bicarbonate is inapplicable, as it is for higher plant C 4 and CAM photosynthesis. 

INTERACTIONS AMONG THE AVAILABILITY OF IRON # 
NITROGEN, PHOSPHORUS AND ZINC AND THAT OF OTHER 
RESOURCES IN DETERMINING MARINE PRIMARY 
PRODUCTIVITY AND ITS FATE 

Interactions among Fe, N, P and Zn 

The discussion above of the interaction of the cycles of the four elements with marine 
primary producers has perforce included discussion of some of the interactions among 
these elements, and with some other resources required for growth. An example is the 
increased Fe requirement for growth when nitrate rather than ammonium is the 
exogenous N source, with an even greater Fe requirement when diazotrophy supplies 
ammonia/ammonium from N 9 . We also saw that Fe deficiency does not specifically 
decrease the assimilation of nitrate or N 2 relative to inorganic C, despite the involve- 
ment of additional Fe-containing enzymes in the assimilation of these N sources which 
are not needed for photosynthetic growth with reduced N as the N source. Another 
example is the role of Fe in atmospheric dust deposited in the ocean on the availability 

SGM symposium 65 



Fe, N, P and Zn cycling in the ocean 259 

of Pfrom dissolved inorganic phosphate. Finally, Zn is required through its involvement 
as a cofactor in most carbonic anhydrases in the photosynthetic assimilation of 
inorganic C in most marine primary producers and through its involvement as a 
cofactor in extracellular phosphatases in the assimilation of external organic P. 

The elementary composition of marine phytoplankton grown under nutrient-sufficient 
conditions has been investigated by Quigg et al. (2003), Ho et al. (2003) and Falkowski 
et al. (2004). This work shows that the content of trace elements in the eukaryotes 
correlates with the evolutionary origin of the plastids, with the 'green' organisms 
relatively enriched in Fe, Cu and Zn, while the 'red' organisms are relatively enriched in 
Mn, Co and Cd. These trends in trace metal content cannot all be readily explained by 
differences in the use of the metals in catalysis, as indicated by the presence or absence 
of catalysis involving a given metal, and the variations in the quantity of the different 
metal-containing catalysts, among the taxa (Raven et al., 1999; Quigg et al., 2003; Ho 
etal.,2003). 

Interactions of Fe, N, P and Zn with the assimilation of 
inorganic C 

We have already discussed some interactions of Fe, N, P and Zn with photosynthesis by 
marine primary producers. Here, we consider some other interactions and comment on 
the possible impact of increasing surface ocean C0 2 concentrations resulting from 
anthropogenic C0 2 emissions to the atmosphere on these interactions. We concentrate 
particularly on the impact on the role of CCMs, since these are the means by which 
most marine primary producers have their growth rates saturated with inorganic C 
under the current environmental conditions (Beardall & Raven, 2004; Giordano et al., 
2005). 

Table 1 shows the effect of the availability of Fe, N, P and Zn on the extent of 
engagement of CCMs. The evidence as to the extent to which CCMs are involved 
in inorganic C assimilation generally comes from studies of the concentration of 
inorganic C required to achieve half of the inorganic-C-saturated rate of photo- 
synthesis, a lower half-saturation concentration suggesting a greater involvement of 
CCMs. More rarely, the evidence comes from the measured ratio of the inorganic C 
concentration in the cells to that in the sea-water medium during steady-state 
photosynthesis. Some of the data come from work on freshwater rather than marine 
organisms. 

The effect of limitation of photosynthesis by the supply of PAR, e.g. in benthic primary 
producers growing deep under water or in the shade of other organisms, or planktonic 
primary producers in a deeply mixed ocean surface layer, is to decrease the engagement 

SGM symposium 65 



260 J. A. Raven and others 



Table 1. Inorganic C affinity and expression of CCMs in cyanobacteria and algae 
with CCMs as a function of resource limitation for growth 



Growth-limiting 
resource 


Effect on inorganic 

C affinity/CCM expression 


Reference(s) 


PAR 


Decrease 


Beardall (1 991 ); Kubler & Raven (1 994, 1 995); 
Young & Beardall (2005); Giordano etal. (2005) 


Inorganic C 


Increase 


Giordano et al. (2005) 


Inorganic N 


Increase (nitrate as N source) 


Beardall etal. (1982, 1991); Young & Beardall 
(2005) 




Decrease (ammonium as 
N source) 


Giordano etal. (2003) 


Inorganic P 


Increase 


Beardall etal. (2005) 




Decrease 


Bozena etal. (2000) 


Inorganic Fe 


Increase 


Young & Beardall (2005) 


Inorganic Zn 


Decrease 


Morel etal. (1994); Buitenhuisef al. (2003) 



of CCMs in C assimilation. Limitation of photosynthesis by decreased inorganic C 
supply can occur in marine habitats, e.g. in high intertidal rock pools with high 
densities of primary producers and infrequent flushing with sea water at neap tides, and 
in blooms of plankton (Falkowski & Raven, 1997; Raven et al., 2002). Such inorganic 
C limitation increases the engagement of CCMs (Table 1). Conversely, the increase in 
sea surface water concentration of C0 2 , and to a smaller relative extent in bicarbonate 
and total inorganic C as well as decrease in pH, predicted to occur by 2100 leads to a 
decreased CCM engagement and a decreased affinity of photosynthesis for inorganic C 
(Table 1; Beardall 6c Raven, 2004). A decreased availability of combined N generally 
increases the engagement of CCMs, the exception being for an organism grown on 
ammonium rather than on nitrate, as were all of the other organisms tested (Table 1). 
There are conflicting data as to the effect of a decreased availability of P on CCM 
engagement, while the only data available on the effect of limited Fe availability indicate 
increased CCM engagement (Table 1). Finally, decreased Zn availability decreases the 
engagement of CCMs (Table 1). 

These data (Table 1) are in general agreement with predictions from theoretical 
considerations on the effect of variations in resource supply. The effects of light 
availability can be rationalized in terms of the relatively greater impact of leakage from 
the accumulated inorganic C pool as the energy supply for (re-)accumulation is 
decreased, and there is an increased 'affinity' for PAR in driving photosynthesis (higher 
rate at a given low photon flux density) paralleling the decreased CCM expression 
for organisms acclimated to low light, and the reverse for organisms acclimated to 
low inorganic C and high light (Kubler 6c Raven, 1994, 1995). The theoretical rationale 

SGM symposium 65 



Fe, N, P and Zn cycling in the ocean 261 

for the commonly observed increase in CCM expression when N availability is restrict- 
ed relates to the N costs of the additional protein components related to the CCM 
relative to the N costs of the alternative of Rubisco oxygenase and photorespiratory 
metabolism (Beardall etal., 1982). Similarly, the increased CCM engagement under low 
Fe conditions can be related to Fe requirements for thylakoid components involved in 
CCMs and in Rubisco oxygenase activity and photorespiration (Raven 6c Johnston, 
1991). The decreased CCM engagement in organisms not thought to use a C 4 -like 
metabolism when Zn availability is restricted can be related to the Zn in the increased 
levels of carbonic anhydrases attendant on increased CCM activity 

While these rationalizations are useful in interpreting the data, further mechanistic 
studies are needed, especially related to the interaction of the various resource supply 
factors that determine CCM engagement. Also required is work in which the possibility 
of long-term genetic adaptation, as well as of short-term acclimation, is addressed 
for increased CO ? supply (Collins & Bell, 2004) in relation to other resource supply 
factors. 



Interaction of Fe, N, P and Zn supply with the fate of marine 
primary producers 

So far we have concentrated on the impact of the availability of resources on the rate of 
primary production, i.e. a consideration of bottom-up or stress factors determining 
the accumulation of biomass. We now consider the implications of resource availability 
for the fate of the primary producer biomass. This includes the biotic interactions of 
grazing and parasitism, i.e. the classic top-down factors, as well as the abiotic factor 
of sinking in the case of phytoplankton; all of these would be subsumed under Grime's 
concept of disturbance. 

Dealing first with the effect of nutrient limitation on grazing, restricted availability of 
a nutrient can alter the chemical composition of primary producers and hence their 
nutritional value to herbivores (grazers). Most of the work on the impact of the 
chemical composition of aquatic primary producers on their nutritional value has 
involved fresh waters (Sterner & Elser, 2002). While there are influences on the ingestion 
rate of primary producers by grazers as a function of the resource supply conditions 
during the growth of the photosynthetic organisms, it is not clear whether these 
changes routinely alter the impact of grazers on primary producer populations (Sterner 
& Elser, 2002). The situation is made more complex by the general effect of nutrient 
deficiency not just on the ratio of major nutritional components such as proteins, 
carbohydrates and lipids, and the content of essential trace elements, but by additional 
responses of primary producers. The further responses to nutrient deficiency include 

SGM symposium 65 



262 J. A. Raven and others 



the production of additional quantities of organic chemicals which deter or poison 
grazers or render the ingested material indigestible. These responses by the primary 
producers are a matter of some controversy One debate concerns the extent to which 
the production of 'defence compounds' is selected for because of their effect in 
restricting removal of biomass when the rate of biomass production is already 
constrained by limited nutrient availability (Sterner 6c Elser, 2002). What some workers 
perceive as an alternative explanation is that the additional production of at least the 
N-free 'defence compounds' is related to the excess of photosynthate over the capacity 
of the organism to use the photosynthate in growth when nutrient deficiency constrains 
photosynthate use in growth more than it does photosynthate production (Sterner & 
Elser, 2002). This contrast may be more apparent than real, with 'excess' organic C in 
nutrient-limited primary producers used to make compounds which limit the loss to 
grazers of a population whose capacity for replacement is resource-limited. The argu- 
ments become especially complex when very small organisms, such as phytoplankton, 
are considered (Wolfe et al., 1997; Ianora et al., 2004), as the effectiveness of defence 
compounds in restricting grazing in terms of natural selection is a function of the 
ability of the grazers to distinguish among prey, and then to restrict the ingestion of 
the better-defended cells. Resolution of these problems involves not just consideration 
of the sensory and manipulative abilities of the grazers in relation to their learned or 
innate behaviour in detecting and avoiding better-defended cells, but also the nature 
of the phytoplankton populations (e.g. clonal or outbreeding) and the unit of natural 
selection in these organisms (Thornton, 2002; Raven & Waite, 2004). The implications 
of these 'defence compounds' in avoiding parasitism, including viral infection, are even 
less clear than their effects on grazers. 



Additional effects of nutrient deficiency on the palatability and nutritional value 
of marine primary producers include the impact of nutrient deficiency on the extent of 
mineralization. Tables 2 and 3 show the effects of resource deficiency on the ratio 
of mineral material to organic matter in two groups of quantitatively very important 
marine primary producers, the silicified diatoms (planktonic and benthic) and the 
calcified coccolithophores (planktonic, with benthic stages in the life cycle of a few 
coastal species). The data in Table 2 indicate that restricted availability of Fe, N, P and 
Zn, as well as of PAR and of inorganic C, increases the content of silica relative to that 
of organic matter. Indeed, every factor which increases the length of the vegetative life 
cycle of diatoms increases the time of the G2 phase of the life cycle. The G2 phase 
is when silicification occurs, so that slowing cell growth increases the relative amount 
of silica per cell (Martin-Jezequel et al., 2000). The situation is rather less clear 
for calcification in coccolithophores (Table 3). Here, restricted light availability 
decreases calcification relative to organic C production, as does restricted bicarbonate 

SGM symposium 65 



Fe, N, P and Zn cycling in the ocean 263 



Table 2. Silicification by diatoms as a function of resource limitation for growth 

Growth limitation Effect on particulate Reference(s) 
by supply of the silica/particulate 

named resource organic C 

Martin-Jezequel etal. (2000); Claquin etal. (2002) 

Milliganefa/. (2004) 

Martin-Jezequel etal. (2000); Claquin etal. (2002) 

Martin-Jezequel etal. (2000); Claquin etal. (2002) 

Hutchins & Bruland (1 998); Takeda (1 998); 
De La Rocha etal. (2000); Franckef a/. (2003); 
Leynaert etal. (2004) 

De La Rocha etal. (2000); Franck etal. (2003) 

* Increased cell Si quota at low C0 2 is a result of decreased loss of Si by efflux and dissolution rather 
than increased influx of Si(OH) 4 . 



PAR 


Increase 


Inorganic C 


Increase* 


Inorganic N 


Increase 


Inorganic P 


Increase 


Inorganic Fe 


Increase 


Inorganic Zn 


Increase 



Table 3. Calcification by coccolithophores as a function of resource limitation for 
growth 



Growth limitation 
by supply of the 
named resource 



Effect on particulate 
inorganic C/particulate 
organic C 



Reference(s) 



PAR 


Decrease 


Inorganic C 


Increase with decreased CO 




Decrease with decreased 




bicarbonate 


Inorganic N 


Increase 


Inorganic P 


Increase 


Inorganic Fe 


No effect 


Inorganic Zn 


Increase 



Paasche (1964, 1999,2001) 

Riebesellefa/. (2000) 

Sekino & Shiraiwa (1 994); Shiraiwa (2003) 

Paasche (1998, 2001) 

Paasche & Bruback (1 994); Paasche (1 998, 2001 ) 
Schulzef a/. (2004); cf. Crawford etal. (2003) 
Schulzef a/. (2004); cf. Crawford etal. (2003) 



supply, while restricted C0 2 supply increases calcification. Restricted supply of N, 
P and Zn increase calcification, while there is no significant effect of restricted supply 
of Fe. 

Does the increased mineralization as a result of nutrient deficiency restrict mortality 
due to grazing and parasitism via mechanical effects or, in the case of calcification, via 
challenges to the maintenance of low pH in any acid parts of the grazer's digestive 
system? There is little direct evidence even on how effective the 'nutrient-sufficient' 
quantities of silica and of calcite in diatoms and coccolithophores are in decreasing 
losses due to grazing and parasitism relative to notionally comparable organisms 
lacking silicification and calcification (Raven & Waite, 2004). 



SGM symposium 65 



264 J. A. Raven and others 



Increased mineralization must increase the density of the organisms, since silica and 
calcite are each two to three times as dense as the cell protoplast, thus increasing 
the sinking rate of planktonic organisms. Raven &C Waite (2004) consider the role of the 
increased rate of sinking resulting from the additional 'nutrient-deficient' quantities of 
silica and calcite in diatoms and coccolithophores in the context of movement down the 
water column to where there might be higher concentrations of nutrients. However, 
the vertical water movements in the upper mixed layer typical of habitats favoured by 
planktonic diatoms mean that such effects are essentially limited to an increased chance 
of falling out of the upper, mixed layer into the more nutrient-rich thermocline, with 
the possibility of the less-dense cells resulting from (light-limited but nutrient-suffi- 
cient) growth in the thermocline becoming reincorporated into the upper mixed layer 
(see Rodriguez et al., 2001; Ptacnik et al., 2003). The variations in density which permit 
very large-celled planktonic diatoms such as Ethmo discus and some species of 
Rhizosolenia to make vertical migrations taking days within one cell division cycle in 
oligotrophic waters, gaining photosynthate near the surface and nutrients at depth, 
must depend on changes in protoplast density, since silicification-related changes in 
density are not manifest within a cell cycle covering a range of light and nutrient-supply 
conditions (Raven & Waite, 2004). 



Aside from the effects of mineralization and its variation with resource supply in the 
life of diatoms and coccolithophores, the increased ballast per unit organic matter 
in nutrient-limited cells increases the potential for sinking of dead cells. In the case 
of diatoms, this effect is exacerbated by the absence of the mechanisms regulating 
(typically by lowering) the density of the protoplast in living cells (Raven & Waite, 
2004). This effect could be multiplicative with the effect of nutrient deficiency in 
enhancing production of extracellular polysaccharides by algal cells (Fogg & Westlake, 
1955; Hellebust, 1974; Wetz & Wheeler, 2003; but see Engel, 2002; Engel et al, 2004). 
These exopolysaccharides are important in the flocculation of particles in sea water 
(Engel et al., 2004), which increases their sinking rate for a given density by decreasing 
the surface area per unit volume (Stokes' law). While terrigenous mineral, e.g. clay, 
particles are also important as mineral ballast in the sinking of particulate organic 
matter in marine snow, the double effect of nutrient deficiency, by increasing the 
amount of biomineral ballast and the amount of floe-forming extracellular poly- 
saccharides, could enhance not only the sinking of particulate organic matter in the 
'biological pump', but also the removal of Fe, N, P and Zn from already nutrient- 
deficient surface waters. A further feedback on nutrient removal as a result of low 
nutrient availability is the capacity of the extracellular polysaccharides in the floes 
to bind Fe and Zn from sea water (Engel et al., 2004). While Passow (2004) suggests 
that the organic C fluxes may drive mineral particle fluxes rather than vice versa, it 

SGM symposium 65 



Fe, N, P and Zn cycling in the ocean 265 

seems inescapable that mineral particles increase the density, and sinking rate, of 
organic particles (see Klaas & Anchor, 2002; Boyd et al., 2004; Jickells et ai, 2005). 



CONCLUSIONS 

Fe, N and P have important roles in constraining marine primary productivity; the role 
of Zn in limiting the growth of primary producers is less clearly established. There are 
very substantial ecological, physiological and biochemical interactions among these 
nutrients, and also interactions with the supply of other resources. Decreased 
availability of these nutrient elements also potentially increases the sedimentation of 
particulate material, including these four nutrient elements; further work is needed 
to establish the significance of the potential feedbacks. 



ACKNOWLEDGEMENTS 

The work of K. B. and E. G. on inorganic C assimilation by diatoms is funded by a grant to 
R. C. L. and J.A.R. from the Natural Environment Research Council UK, and that of M. M. on 
mechanisms of inorganic C assimilation in coccolithophores is funded by the Biotechnology and 
Biological Sciences Research Council UK. J. B. s work on algal photosynthesis is funded by the 
Australian Research Council. Discussion with Professor Geoff Codd was very helpful. 



REFERENCES 



Agren, G. I. (2004). The C : N : P stoichiometry of autotrophs - theory and observations. Ecol 

Lett 7, 185-191. 
Armbrust, E. V., Berges, J. A., Bowler, C. & 42 other authors (2004). The genome of 

the diatom Thalassiosira pseudonana: ecology, evolution, and metabolism. Science 

306, 79-86. 
Beardall, J. (1991). Effects of photon flux density on the C0 2 -concentrating mechanism 

in the cya no bacterium Anabaena variabilis. J Plankton Res 1 3 (Suppl.), 1 33-1 46. 
Beardall, J. & Raven, J. A. (2004). The potential effects of global climate change on 

microalgal photosynthesis, growth and ecology. Phycologia 43, 26-40. 
Beardall, J., Griffiths, H. & Raven, J. A. (1982). Carbon isotope discrimination and 

the C0 2 accumulating mechanism in Chlorella emersonii.J Exp Bot 33, 729-737 '. 
Beardall, J., Roberts, S. & Millhouse, J. (1991). Effects of nitrogen limitation on inorganic 

carbon uptake and specific activity of ribulose-1 ,5-P 2 carboxylase in green micro- 
algae. Can J Bot 69, 1 1 46-1 1 50. 
Beardall, J., Roberts, S. & Raven, J. A. (2005). Regulation of inorganic carbon acquisition 

by phosphorus limitation in the green algae, Chlorella emersonii. Can J Bot (in press). 
Benitez-Nelson, C. R. & Karl, D. M. (2002). Phosphorus cycling in the North Pacific 

subtropical gyre using cosmogenic 32 P and 33 R Limnol Oceanogr 47 , 762-770. 
Berner, E. K. & Berner, R. A. (1996). Global Environment. Water, Air, and Geochemical 

Cycles. Upper Saddle River, NJ & London: Prentice Hall. 



SGM symposium 65 



266 J. A. Raven and others 



Bertilsson, S., Berglund, O., Karl, D. M. & Chisholm, S. W. (2003). Elemental composition 

of marine Prochlorococcus and Synechococcus: implications for ecological 

stoichiometry of the sea. Limnol Oceanogr 48, 1721-1731. 
Bjerrum, C. J. & Canfield, D. E. (2002). Ocean productivity before about 1 -9 Gyr ago 

limited by phosphorus absorption onto iron oxides. Nature 417, 159-162. 
Bjo'rkman, K. M. & Karl, D. M. (2003). Bioavailability of dissolved organic phosphorus in 

the euphotic zone at station ALOHA, North Pacific subtropical gyre. Limnol Oceanogr 

AS, 1049-1057. 
Boyd, P. W. (2002a). Environmental factors controlling phytoplankton processes in the 

Southern Ocean. J Phycol 38, 844-861 . 
Boyd, P. W. (2002b). The role of iron in the biogeochemistry of the Southern Ocean and 

equatorial Pacific: a comparison of in situ iron enrichments. Deep Sea Res II Top Stud 

Oceanogr 49 1 803-1 82 1 . 
Boyd, P. W., Jackson, G. A. & Waite, A. M. (2002). Are mesoscale perturbation 

experiments in polar waters prone to physical artefacts? Evidence from algal 

aggregation modelling studies. Geophys Res Lett 29, art. no. 1 541 . 
Boyd, P. W., Law, C. S., Wong, C. S. & 35 other authors (2004). The decline and fate of an 

iron-induced subarctic phytoplankton bloom. Nature 428, 549-553. 
Bozena, K.-S., Zielinski, P. & Maleszewski, S. (2000). Involvement of glycolate metabolism 

in acclimation of Chlorella vulgaris cultures to low phosphate supply. Plant Physiol 

Biochem 38, 727-734. 
Bryant, D. A. (2003). The beauty in small things revealed. Proc Natl Acad Sci U S A 100, 

9647-9649. 
Buitenhuis, E. I., Timmermans, K. R. & van de Baar, H. J. W. (2003). Zinc-bicarbonate 

col imitation of Emiliania huxleyii. Limnol Oceanogr AS, 1 575-1 582. 
Capone, D. G. (2001). Marine nitrogen fixation: what's the fuss? Curr Opin Microbiol A, 

341-348. 
Chester, R. (2003). Marine Geochemistry, 2nd edn. Oxford: Blackwell Science. 
Claquin, P., Martin-Jezequel, V., Kromkamp, J. C, Veldhuis, M. J. W. & Kraay, G. W. 

(2002). Uncoupling of silicon compared with carbon and nitrogen metabolisms and 

the role of the cell cycle in continuous cultures of Thalassiosira pseudonana 

(Bacillariophyceae) under light, nitrogen, and phosphorus control. J Phycol 38, 

922-930. 
Collins, S. & Bell, G. (2004). Phenotypic consequences of 1 ,000 generations of selection at 

elevated C0 2 in a green alga. Nature A3}, 566-569. 
Cooke, R. R. M., Hurd, C. L, Lord, J. M., Peake, B. M., Raven, J. A. & Rees, T. A. V. 

(2004). Iron and zinc content of Hormosira banksii in New Zealand. N Z J Mar Freshw 

Res 38, 73-85. 
Crawford, D. W., Lipsen, M. S., Purdie, D. A. & 9 other authors (2003). Influence of 

zinc and iron enrichments on phytoplankton growth in the northeastern subarctic 

Pacific. Limnol Oceanogr AS, 1 583-1 600. 
Cullen, J. J. (1999). Iron, nitrogen and phosphorus in the ocean. Nature 402, 372. 
De La Rocha, C. L, Hutchins, D. A., Brzezinski, M. A. & Zhang, Y. H. (2000). Effects of 

iron and zinc deficiency on elemental composition and silica production by diatoms. 

Mar Ecol Prog Ser 1 95, 7 1 -79. 
Dudley, B. J., Gahnstrom, A. M. E. & Walker, D. I. (2001). The role of benthic vegetation 

as a sink for elevated inputs of ammonium and nitrate in a mesotrophic estuary. 

Mar Ecol Prog Ser 2 1 9, 99-1 07 . 
Dufresne, A., Salanoubat, M., Partensky, F. & 18 other authors (2003). Genome 



SGM symposium 65 



Fe, N, P and Zn cycling in the ocean 267 



sequence of the cyanobacterium Prochlorococcus marinus SS120, a nearly minimal 

oxyphototrophic genome. Proc Natl Acad Sci USA 100, 10020-10025. 
Elam, J. S., Malek, K., Rodrigez, J. A., Doucette, P. A., Taylor, A. B., Hayward, L. J., 

Cabelli, D. E. # Valentine, J. S. & Hart, P. J. (2003). An alternative mechanism 

of bicarbonate-mediated peroxidation by copper-zinc superoxide dismutase: 

rates enhanced via proposed enzyme-associated peroxycarbonate intermediate. 

J Biol Chem 278, 2 1 032-2 1 039. 
Ellwood, M. J., Hunter, K. A. & Kim, J. P. (2001). Zinc speciation in Lakes Manapouri and 

Hayes, New Zealand. MarFreshw Res 52, 21 7-222. 
Elser, J. J., Acharya, K., Kyle, M. & 9 other authors (2003). Growth rate-stoichiometry 

couplings in diverse biota. Ecol Letts, 936-943. 
Engel, A. (2002). Direct relationship between C0 2 uptake and transparent exopolymer 

particles production in natural phytoplankton. J Plankton Res 24, 49-53. 
Engel, A., Thorns, S., Riebesell, U., Rochelle-Newall, E. & Zondervan, I. (2004). 

Polysaccharide aggregation as a potential sink of marine dissolved organic carbon. 

Nature 428, 929-932. 
Falkowski, P. G. (1997). Evolution of the nitrogen cycle and its influence on the biological 

sequestration of C0 2 in the ocean. Nature 387, 272-275. 
Falkowski, P. G. (2000). Rationalizing element ratios in unicellular algae. J Phycol 36, 3-6. 
Falkowski, P. G. & Raven, J. A. (1997). Aquatic Photosynthesis. Oxford: Blackwell Science. 
Falkowski, P. G., Katz, M. E., Knoll, A. H., Quigg, A., Raven, J. A., Schofield, O. & 

Taylor, F. J. R. (2004). The evolution of modern eukaryotic phytoplankton. Science 

305, 354-360. 
Field, C. B., Behrenfeld, M. J., Randerson, J. T. & Falkowski, P. G. (1998). Primary 

production of the biosphere: integrating terrestrial and oceanic components. Science 

281,237-240. 
Flynn, K. J. & Hipkin, C. R. (1999). Interactions between iron, light, ammonium, and 

nitrate: insights from the construction of a dynamic model of algal physiology. 

JPhycol3S, 1171-1190. 
Fogg, G. E. & Westlake, D. F. (1955). The importance of extracellular products of algae in 

freshwater. Verhein Int Verein Limnol 12, 21 9-232. 
Franck, V. M., Bruland, K. W., Hutchins, D. A. & Brzezinski, M. A. (2003). Iron and 

zinc effects on silicic acid and nitrate uptake kinetics in three high-nutrient, 

low-chlorophyll (HNLC) regions. Mar Ecol Prog Ser 252, 1 5-33. 
Frew, R., Bowie, A., Croot, P. & Pickmere, S. (2001). Macron utrient and trace-metal 

geochemistry of an in situ iron-induced Southern Ocean bloom. Deep Sea Res II 

Top Stud Oceanogr 48, 2467-2481 . 
Gall, M. P., Strzepek, R., Maldonado, M. & Boyd, P. W. (2001). Phytoplankton processes. 

Part 2. Rates of primary production and factors controlling algal growth during 

the Southern Ocean Iron REIease Experiment (SOIREE). Deep Sea Res II Top Stud 

Oceanogr AS, 2571-2590. 
Geider, R. J. & La Roche, J. (2002). Redfield revisited: variability of C : N : P in marine 

microalgae and its biochemical basis. Eur J Phycol 37, 1-17. 
Giordano, M., Norici, A., Forssen, M., Eriksson, M. & Raven, J. A. (2003). An anaplerotic 

role for mitochondrial carbonic anhydrase in Chlamydomonas reinhardtii. Plant 

Physiol 132, 2126-2134. 
Giordano, M., Beardall, J. & Raven, J. A. (2005). C0 2 concentrating mechanisms in 

algae: mechanisms, environmental modulation, and evolution. Ann u Rev Plant Biol 

56,99-131. 



SGM symposium 65 



268 J. A. Raven and others 



Granum, E., Raven, J. A. & Leegood, R. C. (2005). How do diatoms fix ten billion tonnes 

of inorganic carbon per year? CanJBot(\n press). 
Grime, J. P. (1974). Vegetation classification by reference to strategies. Nature 250, 26-31 . 
Heldal, M., Scanlan, D. J., Norland, S., Thingstad, F. & Mann, N. H. (2003). Elemental 

composition of single cells of various strains of marine Prochlorococcus and 

Synechococcus using X-ray microanalysis. Limnol Oceanogr 48, 1 732-1 743. 
Hellebust, J. A. (1974). Extracellular products. In Algal Physiology and Biochemistry, 

pp. 838-863. Edited by W. D. P. Stewart. Oxford: Blackwell Scientific. 
Ho, T.-Y, Quigg, A., Finkel, Z. V., Milligan, A. J., Wyman, K., Falkowski, P. G. & Morel, 

F. M. M. (2003). The elemental composition of some marine phytoplankton. J Phycol 

39, 1145-1159. 
Hutchins, D. A. & Bruland, K. W. (1998). Iron-limited growth and Si : N uptake ratios in a 

coastal upwelling regime. Nature393, 561-564. 
Hutchins, D. A., DiTullio, G. R., Zhang, Y. & Bruland, K. W. (1998). An iron limitation 

mosaic in the California upwelling regime. Limnol Oceanogr 43, 1037-1054. 
Hutchins, D. A., Hare, C. E., Weaver, R. S. & 12 other authors (2002). Phytoplankton iron 

limitation in the Humboldt Current and Peru Upwelling. Limnol Oceanogr 47 , 997- 

1011. 
lanora, A., Miralto, A., Poulet, S. A. & 9 other authors (2004). Aldehyde suppression 

of copepod recruitment in blooms of a ubiquitous planktonic diatom. Nature 429, 

403-407. 
Jickells, T. D., An, Z. S., Andersen, K. K. & 16 other authors (2005). Global iron 

connections between desert dust, ocean biogeochemistry, and climate. Science 

308,67-71. 
Johnston, A. M., Raven, J. A., Beardall, J. & Leegood, R. C. (2001). Photosynthesis in a 

marine diatom. Nature 41 2, 40^41 . 
Karl, D. M. & Tien, G. (1997). Temporal variability in dissolved phosphorus concentration 

in the subtropical North Pacific Ocean. Mar Chem 56, 77-96. 
Karl, D. M., Tien, G., Dore, J. & Winn, C. D. (1993). Total dissolved nitrogen and 

phosphorus concentrations at US-JGOFS Station ALOHA: Redfield reconciliation. 

Mar Chem 41, 203-208. 
Klaas, C. & Anchor, D. E. (2002). Association of sinking organic matter with various types of 

mineral ballast in the deep sea: implications for the rain ratio. Global Biogeochem 

Cyc/es 16, art. no. 116. 
Klausmeier, C. A., Litchman, E., Dufresne, T. & Levin, S. A. (2004a). Optimal nitrogen-to 

phosphorus stoichiometry of phytoplankton. Nature 429, 171-174. 
Klausmeier, C. A., Litchman, E. & Levin, S. A. (2004b). Phytoplankton growth and 

stoichiometry under multiple nutrient limitation. Limnol Oceanogr 49, 1463-1470. 
Krom, M. D., Herut, B. & Mantoura, R. F. C. (2004). Nutrient budgets for the Eastern 

Mediterranean: implications for phosphorus limitation. Limnol Oceanogr 49, 1 582- 

1592. 
Kubler, J. E. & Raven, J. A. (1994). Consequences of light-limitation for carbon acquisition 

in three rhodophytes. Mar Ecol Prog Ser 1 1 0, 203-208. 
Kubler, J. E. & Raven, J. A. (1995). Interaction between inorganic carbon acquisition 

and light supply in Palmaria palmata (Rhodophyta). J Phycol 31, 369-375. 
Kustka, A., Sahudo-Wilhelmy, S., Carpenter, E. J., Capone, D. G. & Raven, J. A. 

(2003a). A revised estimate of the iron use efficiency of nitrogen fixation, with 

special reference to the marine cyanobacterium Trichodesmium spp. (Cyanophyta). 

J Phycol 39, 12-25. 



SGM symposium 65 



Fe, N, P and Zn cycling in the ocean 269 



Kustka, A., Sanudo-Wilhelmy, S. A., Carpenter, E. J., Capone, D., Burns, J. & Sunda, W. 

G. (2003b). Iron requirements for dinitrogen- and ammonium-supported growth in 

cultures of Trichodesmium (IMS 101): comparison with nitrogen fixation rates and 

iron : carbon ratios of field populations. Limnol Oceanogr AS, 1869-1884. 
Lane, T. W. & Morel, F. M. M. (2000a). Regulation of carbonic anhydrase expression 

by zinc, cobalt, and carbon dioxide in the marine diatom Thalassiosira weissflogii. 

Plant Physiol \23, 345-352. 
Lane, T. W. & Morel, F. M. M. (2000b). A biological function for cadmium in the marine 

diatoms. Proc Natl Acad Sci U 5 A 97, 4627-4631 . 
Leonardos, N. & Geider, R. J. (2004). Response of elemental and biochemical composition 

of Chaetoceros muelleri to growth under varying light and nitrate : phosphate 

supply ratios and their influence on critical N : P. Limnol Oceanogr 49, 2105-21 14. 
Leynaert, A., Bucciarelli, E., Claquin, P., Dugdale, R. C, Martin-Jezequel, V., 

Pondaven, P. & Ragueneau, O. (2004). Effect of iron deficiency on diatom size 

and silicic acid uptake kinetics. Limnol Oceanogr 49, 1 134-1 143. 
MacFarlane, J. J. & Raven, J. A. (1990). C, N and P nutrition in Lemanea mammilosa 

Kutz. (Batrachospermales, Rhodophyta) in the Dighty Burn, Angus, Scotland. Plant 

Cell Environ 13, 1-13. 
Martin, J. H. & Fitzwater, S. E. (1988). Iron deficiency limits phytoplankton growth in the 

north-east Pacific subarctic. Nature 331, 341-343. 
Martin, J. H., Fitzwater, S. E. & Gordon, R. M. (1 990). Iron deficiency limits phytoplankton 

growth in Antarctic waters. Global Biogeochem Cycles A, 5-13. 
Martin, J. H., Gordon, R. E. & Fitzwater, S. E. (1991). The case for iron. Limnol Oceanogr 

36, 1793-1802. 
Martin-Jezequel, V., Hildebrand, M. & Brzezinski, M. A. (2000). Silicon metabolism in 

diatoms: implicationsforgrowth.JPhyco/ 36, 821-840. 
Milligan, A. J., Varela, D. E., Brzezinski, M. A. & Morel, F. M. M. (2004). Dynamics 

of silicon metabolism and silicon isotopic discrimination in a marine diatom as a 

function of pC0 2 . Limnol Oceanogr 49, 322-329. 
Mills, M. M., Ridame, C, Davey, M., La Roche, J. & Geider, R. J. (2004). Iron and 

phosphorus co-limit nitrogen fixation in the eastern tropical North Atlantic. Nature 

429, 292-294. 
Morel, F. M. M., Reinfelder, J. R., Roberts, S. B., Chamberlain, C. P., Lee, J. G. & Yee, D. 

(1994). Zinc and carbon co-limitation of marine phytoplankton. Nature 369, 

740-742. 
Morel, F. M. M., Cox, E. H., Kraepiel, A. M. L, Lane, T. W., Milligan, A. J., Schaperdoth, 

I., Reinfelder, J. R. & Tortell, P. D. (2002). Acquisition of inorganic carbon by the 

marine diatom Thalassiosira weissflogii. Funct Plant Biol 29, 301-308. 
Paasche, E. (1964). A tracer study of the inorganic carbon uptake during coccolith 

formation and photosynthesis in the coccolithophorid Coccolithus huxleyi. Physiol 

Plant Suppl\\\, 12-82. 
Paasche, E. (1998). Roles of nitrogen and phosphorus in coccolith formation in Emiliania 

huxleyi (Prymnesiophyceae). EurJPhycol 33, 33-42. 
Paasche, E. (1999). Reduced coccolith calcite production under light-limited growth: a 

comparative study of three clones of Emiliania huxleyi (Prymnesiophyceae). 

Phycologia 38, 508-516. 
Paasche, E. (2001). A review of the coccolithophorid Emilianina huxleyi (Prymnesiophyceae), 

with particular reference to growth, coccolith formation, and calcification- 
photosynthesis interactions. Phycologia 40, 503-529. 



SGM symposium 65 



270 J. A. Raven and others 



Paasche, E. & Bruback, S. (1994). Enhanced calcification in the coccolithophorid Emiliania 

huxleyi (Haptophyceae) under phosphorus limitation. Phycologia 33, 324-330. 
Palenik, B., Brahamsha, B. f Larimer, F. W. & 12 other authors (2003). The genome of a 

motile marine Synechococcus. Nature 424, 1 037-1 042. 
Passow, U. (2004). Switching perspectives: do mineral fluxes determine particulate organic 

carbon fluxes or vice versa? Geochem Geophys Geosyst 5, art. no. Q04002. 
Ptacnik, R., Diehl, S. & Berger, S. (2003). Performance of sinking and nonsinking 

phytoplankton taxa in a gradient of mixing depths. Limnol Oceanogr 48, 1903- 

1912. 
Quigg, A., Finkel, Z. V., Irwin, A. J., Rosenthal, Y, Ho, T.-Y, Reinfelder, J. R., Schofield, 

O., Morel, F. M. M. & Falkowski, P. G. (2003). The evolutionary inheritance of 

elementary stoichiometry in marine phytoplankton. Nature 425, 291-294. 
Raven, J. A. (1987). The role of vacuoles. New Phytol 1 06, 358-422. 
Raven, J. A. (1988). The iron and molybdenum use efficiencies of plant growth with 

different energy, carbon and nitrogen sources. New Phytol 109, 279-287. 
Raven, J. A. (1 990). Predictions of Mn and Fe use efficiencies of phototrophic growth as 

a function of light availability for growth and C assimilation pathway. New Phytol 

116, 1-18. 
Raven, J. A. (1996). The role of autotrophs in global C0 2 cycling. In Microbial Growth on 

C 1 Compounds, pp. 105-144. Edited by M. E. Lidstrom & F. R. Tabita. Dordrecht, 

Netherlands: Kluwer Scientific. 
Raven, J. A. (1997). Phagotrophy in phototrophs. Limnol Oceanogr 42, 198-205. 
Raven, J. A. & Johnston, A. M. (1991). Mechanisms of inorganic-carbon acquisition in 

marine phytoplankton and their implications for the use of other resources. Limnol 

Oceanogr 36, 1701-1714. 
Raven, J. A. & Taylor, R. (2003). Macroalgal growth in nutrient-enriched estuaries: a 

biogeochemical and evolutionary perspective. Water Air Soil Poll ut Focus 3 (1), 7-26. 
Raven, J. A. & Waite, A. M. (2004). The evolution of silicification of diatoms: inescapable 

sinking and sinking as escape? New Phytol 162, 45-61. 
Raven, J. A., Evans, M. C. W. & Korb, R. (1999). The role of trace metals in photosynthetic 

electron transport in 2 -evolving organisms. Photosynth Res 60, 111-1 50. 
Raven, J. A., Johnston, A. M., Kubler, J. E. & 9 other authors (2002). Mechanistic 

interpretation of carbon isotope discrimination by marine macroalgae and sea- 
grasses. Funct Plant Biol 29, 355-378. 
Raven, J. A., Handley, L. L. & Andrews, M. (2004). Global aspects of C/N interactions 

determining plant-environment interactions. J Exp Bot 55, 1 1-25. 
Raven, J. A., Andrews, M. & Quigg, A. S. (2005). The evolution of oligotrophy: 

implications for the breeding of crop plants for low input agricultural systems. Ann 

ApplBiol 146, 261-280. 
Reinfelder, J. R., Kraepiel, A. M. L & Morel, F. M. M. (2000). Unicellular C 4 photo- 
synthesis in a marine diatom. Nature 407, 996-999. 
Reinfelder, J. R., Milligan, A. J. & Morel, F. M. M. (2004). The role of the C 4 pathway 

in carbon accumulation and fixation in a marine diatom. Plant Physiol 135, 2106- 

2111. 
Ridame, C. & Guieu, C. (2002). Saharan input of phosphate to the oligotrophic water 

of the open western Mediterranean Sea. Limnol Oceanogr 47 , 856-869. 
Riebesell, U., Zondervan, I., Rist, B., Tortell, P. D., Zeebe, R. E. & Morel, F. M. M. (2000). 

Reduced calcification of marine phytoplankton in response to increased atmospheric 

C0 2 . Nature 407, 364-367. 



SGM symposium 65 



Fe, N, P and Zn cycling in the ocean 271 



Rocap, G., Larimer, F. W., Lamerdin, J. & 21 other authors (2003). Genome divergence in 

two Prochlorococcus ecotypes reflects oceanic niche differentiation. Nature 424, 

1042-1047. 
Rodriguez, J., Tintore, J., Allen, J. T. & 7 other authors (2001). Mesoscale vertical motion 

and the size of structure of phytoplankton in the ocean. Nature W0, 360-363. 
Sanudo-Wilhelmy, S. A., Kustka, A. B., Gobler, C. J. & 7 other authors (2001). 

Phosphorus limitation of nitrogen fixation by Trichodesmium in the central Atlantic 

Ocean. Nature 41 1, 66-69. 
Sanudo-Wilhelmy, S. A., Tovar-Sanchez, A., Fu, F.-X., Capone, D. G., Carpenter, E. J. & 

Hutchins, D. A. (2004). The impact of surface-absorbed phosphorus on phyto- 
plankton Redfield stoichiometry. Nature 432, 897-901 . 
Schulz, K. G., Zondervan, I., Gerringa, L. J. A., Timmermans, K. R., Veldhuis, M. J. W. & 

Riebesell, U. (2004). Effect of trace metal availability on coccolithophorid 

calcification. Nature 430, 673-676. 
Sekino, K. & Shiraiwa, Y. (1994). Accumulation and utilization of dissolved inorganic 

carbon by a marine unicellular coccolithophorid, Emiliania huxleyi. Plant Cell Physiol 

35,353-361. 
Shiraiwa, Y. (2003). Physiological regulation of carbon fixation in the photosynthesis 

and calcification of coccolithophorids. Comp Biochem Physiol B Biochem Mol Biol 

136,775-783. 
Sterner, R. W. & Elser, J. J. (2002). Ecological Stoichiometry. Princeton, NJ: Princeton 

University Press. 
Sterner, R. W., Smutka, T. M., McKay, R. M. L, Qin, X., Brown, E. T. & Sherrell, R. M. 

(2004). Phosphorus and trace metal limitation of algae and bacteria in Lake Superior. 

Limnol Oceanogr 49, 495-507. 
Strzepek, R. F. & Harrison, P. J. (2004). Photosynthetic architecture differs in coastal and 

oceanic diatoms. Nature 431, 689-692. 
Sunda, W. G., Swift, D. G. & Huntsman, S. A. (1991). Low iron requirement for growth 

in oceanic phytoplankton. Nature 351, 55-57. 
Takeda, S. (1998). Influence of iron availability on nutrient consumption ratio of diatoms in 

oceanic waters. Nature 393, 774-777. 
Thornton, D. C. O. (2002). Individuals, clones or groups? Phytoplankton behaviour and 

units of selection. Ethol Ecol Evol 14, 165-173. 
Ting, C. S., Rocap, G., King, J. & Chisholm, S. W. (2002). Cyanobacterial photosynthesis in 

the oceans: the origin and significance of divergent light-harvesting strategies. 

Trends Microbiol 1 0, 1 34-1 42 . 
Tyrrell, T. (1999). The relative influences of nitrogen and phosphorus on oceanic primary 

production. Nature 400, 525-531 . 
van Oijen, T., van Leeuwe, M. A., Granum, E., Weissing, F. J. f Bellerby, R. G. J., Gieskes, 

W. W. C. & de Baar, H. J. W. (2004). Light rather than iron controls photosynthate 

production and allocation in Southern Ocean phytoplankton populations during 

the austral autumn. J Plankton Res 26, 885-900. 
Vitousek, R M., Cassman, K., Cleveland, C. & 8 other authors (2001). Towards an eco- 
logical understanding of biological nitrogen fixation. Biogeochemistry 57/58, 1-45. 
Volker, C. & Wolf-Glad row, D. A. (1999). Physical limits on iron uptake mediated by 

siderophores or surface reductases. Mar Chem 65, 227-244. 
Walker, D. I., Campey, M. L. & Kendrick, G. A. (2004). Nutrient dynamics in two seagrass 

species, Posidonia coriacea and Zostera tasmanica on Success Bank, Western 

Australia. Estuar Coastal Shelf Sci 60, 251-260. 



SGM symposium 65 



272 J. A. Raven and others 



Watson, A. J., Bakker, D. C. E. ; Ridgwell, A. J., Boyd, P. W. & Law, C. S. (2000). Effect of 

iron supply on Southern Ocean C0 2 uptake and implications for glacial atmospheric 

C0 2 . Nature 407, 730-733. 
Wetz, M. S. & Wheeler P. A. (2003). Production and partitioning of organic matter during 

simulated phytoplankton blooms. Limnol Oceanogr 48, 1808-1817. 
Wolfe, G. V., Steinke, M. & Kirst, G. O. (1997). Grazing-activated chemical defence in a 

unicellular marine alga. Nature 387, 894-897. 
Wu, J., Sunda, W., Boyle, E. A. & Karl, D. M. (2000). Phosphate depletion in the western 

North Atlantic Ocean. Science 289, 759-762. 
Young, E. B. & Beardall, J. (2005). Modulation of photosynthesis and inorganic carbon 

acquisition in a marine microalga by nitrogen, iron and light availability. Can J Sot (in 

press). 



SGM symposium 65 



Mechanisms and environmental 
impact of microbial metal 
reduction 

Jonathan R. Lloyd 

The Williamson Research Centre for Molecular Environmental Studies and the School of Earth, 
Atmospheric and Environmental Sciences, University of Manchester, Manchester M1 3 9PL, UK 



INTRODUCTION 

Although it has been known for over a century that micro-organisms have the potential 
to reduce metals, more recent observations showing that a diversity of specialist 
bacteria and archaea can use such activities to conserve energy for growth under 
anaerobic conditions have opened up new and fascinating areas of research with 
potentially exciting practical applications (Lloyd, 2003). Micro-organisms have 
also evolved metal-resistance processes that often incorporate changes in the oxida- 
tion state of toxic metals. Several such resistance mechanisms, which do not support 
anaerobic growth, have been studied in detail by using the tools of molecular biology. 
Three obvious examples include resistance to Hg(II), As(V) and Ag(II) (Bruins et aL, 
2000). The molecular bases of respiratory metal-reduction processes have not, however, 
been studied in such fine detail, although rapid advances are expected in this area 
with the imminent availability of complete genome sequences for key metal-reducing 
bacteria, in combination with genomic, proteomic and metabolomic tools. This 
research is being driven forward both by the need to understand the fundamental 
basis of a range of biogeochemical cycles, and also by the possibility of harnessing 
such activities for a range of biotechnological applications. These include the bio- 
remediation of metal-contaminated land and water (Lloyd 6c Lovley, 2001), the 
oxidation of xenobiotics under anaerobic conditions (Lovley 6c~ Anderson, 2000), 
metal recovery in combination with the formation of novel biocatalysts (Yong et aL, 
2002a) and even the generation of electricity from sediments (Bond et aL, 2002). The 
aim of this review is to give an overview of the range of metals (and metalloids) reduced 
by micro-organisms, the mechanisms involved and the environmental impact of 

SGM symposium 65: Micro-organisms and Earth systems - advances in geomicrobiology. 
Editors G. M. Gadd, K. T. Semple & H. M. Lappin-Scott. Cambridge University Press. ISBN 521 86222 1 ©SGM 2005 



274 J. R. Lloyd 



such transformations. Where appropriate, possible applications for these processes will 
also be discussed. 

REDUCTION OF Fe(lll) AND Mn(IV) 

A wide range of archaea and bacteria are able to conserve energy though the reduction 
of Fe(III) (ferric iron) to Fe(II) (ferrous iron). Many of these organisms are also able to 
grow through the reduction of Mn(IV) to Mn(II). The environmental relevance of 
Fe(III) and Mn(IV) reduction has been well-documented (Thamdrup, 2000), whilst 
geochemical and microbiological evidence suggests that the reduction of Fe(III) may 
have been an early form of respiration on Earth (Vargas et al., 1998). Some workers 
have even proposed Fe(III) reduction as a candidate process for the basis of life on 
other planets (Nealson 6c Cox, 2002). On modern Earth, Fe(III) can be the dominant 
electron acceptor for microbial respiration in many subsurface environments (Lovley 
& Chapelle, 1995). As such, Fe(III)-reducing communities can be responsible for 
the majority of organic matter oxidized in such environments. Recent studies have 
shown that a range of important xenobiotics that contaminate aquifers can also be 
degraded under anaerobic conditions by Fe(III)-reducing micro-organisms (Lovley, 
1997; Anderson et ai, 1998; Lovley &; Anderson, 2000). Transformations of inorganic 
contaminants are also possible, and Fe(III)-reducing micro-organisms can also have an 
impact on the fate of other high-valency contaminant metals through direct enzymic 
reduction or via indirect reduction catalysed by biogenic Fe(II). 

Focusing on the enzymic transformations of Fe(III) and Mn(IV), these organisms 
can also influence the mineralogy of sediments through the reductive dissolution of 
insoluble Fe(III) and Mn(IV) oxides (Fig. la). These processes can result in the release 
of potentially toxic levels of reduced Fe(II) and Mn(II), and also trace metals that were 
bound by the host Fe(III) or Mn(IV) minerals. Depending on the chemistry of the water, 
a range of reduced minerals can also be formed, including, for Fe(III) reduction, 
magnetite (Fe 3 4 ), siderite (FeC0 3 ) and vivianite [Fe 3 (P0 4 ) 2 . 8H 2 0] (see Fig. lb-d), 
resulting in a change in structure of the sediments. There is a considerable level of 
interest in these end products of Fe(III) reduction, as they are nanoscale and have 
properties that may make them useful for a range of biotechnological processes. For 
example, large quantities of regular-shaped, nanosized (5-10 nm) crystals of the 
ferromagnetic mineral magnetite (Fe 3 4 ) are formed when Fe(II)-reducing bacteria 
respire by using Fe(III) oxides in laboratory culture (Fig. 2). By controlling and 
manipulating this process of biomineral production, a low-cost, low-energy, environ- 
mentally friendly method of manufacture of nanoparticles could be developed to 
replace existing practices. Of particular appeal is the potential to convert natural and 
waste Fe(III) oxides (e.g. from mining/water industries), which are bulky and difficult 
to handle, to a high-value product that is easy to process. Indeed, ferrite spinels such as 

SGM symposium 65 



Microbial metal reduction 275 



Acetate 




Fig. 1. Mechanism of Fe(lll) reduction (a) and the biogenic Fe(ll)-bearing minerals magnetite (b), 
vivianite (c) and siderite (d). Bars, 50 nm (b); 10 pm (c); 20 pm (d). Figures were generously provided by 
M. Wilkins, V. Coker, F. Islam and L. Adams (University of Manchester, UK). 



magnetite have low coercivity (i.e. low power needed to magnetize or demagnetize 
them), high permeability, high magnetic saturation and low conductivity. These 
properties make them ideal for use in ultrahigh-density data-storage media, frequency- 
selective circuits, radio-receiver antennae, microwave waveguides and other high- 
frequency devices. Indeed, recent work from our laboratory has also demonstrated that 
'designer' magnets can be made by Fe(III) -reducing bacteria by incorporating other 
transition metals into the spinel structure in place of Fe, in some cases enhancing 
the magnetic properties of the biomineral (Coker et al., 2004). 

Diversity of Fe(lll)-reducing organisms 

The first organisms were shown to conserve energy for growth through the reduction of 
Fe(III) [and Mn(IV)] in the 1980s. These model organisms were Shewanella oneidensis 
(formerly Alteromonas putrefaciens and then Shewanella putrefaciens) and Geobacter 
metallireducens (formerly strain GS-15) (Lovley et al., 1987, 1989a; Myers & Nealson, 
1988). Earlier studies had focused on organisms that grow predominantly via ferment- 
ation of sugars such as glucose, with metals utilized as minor electron acceptors 
(Roberts, 1947); typically, <5 % of the reducing equivalents is used for metal reduction 
(Lovley, 1991). 

Over the last 20 years, numerous organisms have been isolated that can grow by using 
Fe(III) as an electron acceptor; more than 90 are listed in a recent review (Lovley et al., 
2004). In many freshwater subsurface environments, the most abundant seem to be 



SGM symposium 65 



276 J. R. Lloyd 




Oxygen 




Octahedral sites 



Tetrahedral sites 




Fig. 2. Production (a) and structure (b) of magnetite produced by Geobacter sulfurreducens. 
Magnetite exhibits a classical spinel crystalline structure, with the oxygen ions forming a compact, 
face-centred cubic assembly and the iron cations occupying octahedral and tetrahedral sites. 
Figures were generously provided by V. Coker and R. Pattrick (University of Manchester, UK). 

relatives of Geobacter metallireducens and fall within the family Geobacteraceae, in 
the delta subdivision of the Proteobacteria (e.g. Rooney-Varga etaL, 1999; Snoeyenbos- 
West et ai, 2000; Roling et al., 2001; Stein et al., 2001; Holmes et al., 2002; Islam et al., 
2004). This group comprises the genera Geobacter, Desulfuromonas, Desulfuromusa 
and Pelobacter (Lovley et al., 2004). These organisms, with the exception of Pelobacter 
species, are able to oxidize a wide range of organic compounds completely, including 
acetate, when respiring by using Fe(III); Pelobacter species are more restricted in the 
range of electron donors utilized, although they can couple hydrogen oxidation to 
metal reduction. Some members of the family Geobacteraceae are also able to use 



SGM symposium 65 



Microbial metal reduction 277 



aromatic compounds, including toluene, phenol and benzoate, as electron donors 
for metal reduction. This is in contrast to Shewanella oneidensis and close relatives 
in the gamma subdivision of the Proteobacteria (a range of Shewanella, Ferrimonas 
and Aeromonas species) that are generally able to use only a restricted range of small 
organic acids and hydrogen as electron donors for Fe(III) and Mn(IV) reduction. The 
full range of other prokaryotes able to reduce Fe(III) is extensive and increasing 
steadily, and crosses a wide range of environments (including extremes of pH and 
temperature) and, indeed, taxonomic groupings. Although Shewanella and Geobacter 
species have been the most intensively studied model Fe(III)-reducing bacteria, the 
reader is referred to an excellent overview of the wide range of currently identified 
Fe(III) -reducing bacteria (Lovley et al., 2004). Perhaps one of the most interesting new 
isolates that can grow through the reduction of Fe(III) is archaeal 'strain 121', which 
has pushed the upper temperature limit for life to 121 °C (Kashefi & Lovley, 2003). 

Mechanisms of Fe(lll) and Mn(IV) reduction: electron transfer 
to insoluble minerals 

The mechanisms of Fe(III) reduction and, to a lesser degree, Mn(IV) reduction have 
been studied in most detail in Shewanella oneidensis and Geobacter sulfur reducens. 
Indeed, research on these organisms has been given added impetus through the 
availability of their genome sequences (available at http://www.tigr.org) and suitable 
genetic systems for the generation of deletion mutants for both of these organisms 
(Myers & Myers, 2000; Coppi et al., 2001). Although the terminal reductase has yet to 
be identified unequivocally in either organism, the involvement of c-type cytochromes 
has been implicated in electron transport to Fe(III) and Mn(IV) by several studies 
(Myers & Myers, 1993, 1997; Gaspard et al., 1998; Magnuson et al., 2000; Beliaev et 
al., 2001; Lloyd et al., 2003). In some examples, activities have also been localized to 
the outer membrane or surface of the cell, consistent with a role in direct transfer of 
electrons to Fe(III) and Mn(IV) oxides that are highly insoluble at circumneutral pH 
(Myers & Myers, 1992, 2001; Gaspard et al., 1998; DiChnstina et al., 2002; Lloyd et al., 
2002). In addition to the proposed direct transfer of electrons to Fe(III) and Mn(IV) 
minerals, soluble 'electron shuttles' are also able to transfer electrons between metal- 
reducing prokaryotes and the mineral surface. This mechanism alleviates the 
requirement for direct contact between the micro-organism and mineral. For example, 
humics and other extracellular quinones are utilized as electron acceptors by Fe(III)- 
reducing bacteria (Lovley et al., 1996) and the reduced hydroquinone moieties are able 
to transfer electrons abiotically to Fe(III) minerals. The oxidized humic is then avail- 
able for reduction by the micro-organism, leading to further rounds of electron 
shuttling to the insoluble mineral (Nevin 6c Lovley, 2002). Very low concentrations of 
an electron shuttle, e.g. 100 nM of the humic analogue anthraquinone-2,6-disulfonate 
(AQDS), can rapidly accelerate the reduction of Fe(III) oxides (Lloyd et al., 1999a) and 

SGM symposium 65 



278 J. R. Lloyd 



possibly other insoluble metal oxides, such as Mn(IV). The environmental significance 
of such processes, however, remains to be confirmed. The secretion of soluble electron 
shuttles by actively respiring Fe(III) and Mn(IV) reducers has also been proposed for 
both Shewanella oneidensis and Geobacter sulfurreducens, and remains hotly debated 
in Geobacter species. Early studies suggested the release of a small, soluble, otype 
cytochrome by Geobacter sulfurreducens (Seeliger et al., 1998), but more recent studies 
have suggested that this protein is not an effective electron shuttle (Lloyd et al., 1999a). 
Studies have also suggested that a small, quinone-containing, extracellular electron 
shuttle is released by Shewanella oneidensis and may also promote electron transfer to 
Fe(III) and Mn(IV) minerals (Newman 6c Kolter, 2000). Finally, an important new 
discovery was made recently when it was shown that Geobacter metallireducens 
synthesized pili and flagella when grown on insoluble Fe(III) or Mn(IV) minerals, but 
not soluble forms of the metals (Childers et al., 2002). These results suggest that 
Geobacter species sense when soluble electron acceptors are depleted and synthesize the 
appropriate appendages that allow movement to Fe(III) and Mn(IV) minerals and 
subsequent attachment. Pili may also play a direct role in electron transfer to the 
extracellular electron acceptor (Reguera et al., 2005). 

REDUCTION OF OTHER TRANSITION METALS 

In addition to Fe(III) and Mn(IV), dissimilatory metal-reducing prokaryotes are able 
to respire by using a wide range of transition metals, including high-valency ions 
of vanadium, chromium, molybdenum, cobalt, palladium, silver, gold and mercury. In 
many cases, reduction leads to a dramatic change in solubility, can potentially lead 
to the precipitation of metal-containing ores and may also offer routes to the bio- 
remediation of metal-contaminated water. 

Vanadium reduction 

Early studies showed V(V) reduction by 'Micrococcus lactilyticus', Desulfovibrio 
desulfuricans and Clostridium pasteurianum (Woolfolk & Whiteley, 1962), followed by 
the observation that the ability to reduce V(V) was widespread amongst soil bacteria 
and fungi (Bautista & Alexander, 1972). More recent work has focused on two 
pseudomonads: 'Pseudomonas vanadiumreductans' and 'Pseudomonas isachenkovii\ 
isolated from a waste stream from a ferrovanadium factory and sea water, respectively 
(Yurkova 6c Lyalikova, 1991). Anaerobic cells were able to utilize a wide range of 
electron donors, including hydrogen, sugars and amino acids. V(V) was reduced to 
blue-coloured V(IV) and possibly further to V(III), the latter indicated by the formation 
of a black precipitate and by its reaction with the reagent Tairon (Yurkova 6c Lyalikova, 
1991). Geobacter metallireducens also reduces V(V) and this form of metabolism has 
been suggested as a mechanism for remediating vanadium-contaminated water (Ortiz- 
Bernad <?£#/., 2004a). 

SGM symposium 65 



Microbial metal reduction 279 

Chromium reduction 

The widespread use of chromium in the metal industries and subsequent contamin- 
ation problems have led to a lot of interest in this metal. Although trace quantities are 
required for some metabolic activities, e.g. glucose and lipid metabolism, chromium 
is considered toxic and is designated a priority pollutant in many countries. Two 
oxidation states dominate: Cr(VI) is the most toxic and mobile form encountered 
commonly, with Cr(III) being less soluble and less toxic. Indeed, Cr(III) is considered 
1000 times less mutagenic than Cr(VI) (Wang, 2000). Current treatment involves 
reduction of Cr(VI) to Cr(III) by using chemical reductants at low pH, followed by 
adjustment to near-neutral pH and subsequent precipitation of Cr(III). Recent studies, 
however, have shown that micro-organisms can also reduce Cr(VI) efficiently at circum- 
neutral pH, and could be used to treat Cr(VI)-contaminated water. In most cases, 
Cr(VI) reduction does not support anaerobic growth, although a few publications have 
suggested that conservation of energy is possible through this form of anaerobic 
metabolism (e.g. Tebo 6c Obraztsova, 1998). 

A wide range of facultative anaerobes are able to reduce Cr(VI) to Cr(III), including 
Escherichia coli, pseudomonads, Shewanella oneidensis and Aeromonas species 
[see Wang (2000) for a more exhaustive list]. Anaerobic conditions are generally 
required to induce maximum activity against Cr(VI), but some enzyme systems operate 
under aerobic conditions, e.g. the soluble NAD(P)H-dependent reductases of l Pseudo- 
monas ambigucf G-l (Suzuki et al., 1992) and Pseudomonas putida (Park et al., 2000). 
Obligate anaerobes are also able to reduce Cr(VI) enzymically and anaerobic growth 
coupled to Cr(VI) reduction has been reported for a sulfate-reducing bacterium (Tebo 
& Obraztsova, 1998). The reduction of Cr(VI) by sulfate-reducing bacteria is parti- 
cularly well-studied (e.g. Lloyd et al., 2001) and has been shown to be catalysed by 
cytochrome c 3 (Lovley & Phillips, 1994). Other studies have also implicated the involve- 
ment of cytochromes in Cr(VI) reduction by bacteria: cytochrome c in Enterobacter 
cloacae (Wang et al., 1989) and cytochromes b and d in Escherichia coli (Shen & Wang, 
1993). Environmental factors that affect Cr(VI) reduction include competing electron 
acceptors, pH, temperature, redox potential and the presence of other metals (Wang, 
2000). A recent study has also demonstrated that the presence of complexing agents 
can promote Cr(VI) reduction, possibly through protection of the metal reductase by 
chelation of Cr(III) or intermediates formed (Mabbett et al., 2002). The type of electron 
donor supplied can also have an effect on the rate and extent of Cr(VI) reduction. 
Optimal electron donors, in keeping with other dissimilatory metal-reduction pro- 
cesses described in this review, are low-molecular-mass carbohydrates, amino acids and 
fatty acids. Finally, indirect mechanisms that also promote Cr(VI) reduction in con- 
taminated sediments are catalysed by biogenic sulfide (Smillie et al., 1981; Fude et al., 
1994) and Fe(II) (Fendorf & Li, 1996). Experiments using contaminated sediments from 

SGM symposium 65 



280 J. R. Lloyd 



Norman, OK, USA, have, however, confirmed that indirect mechanisms may not always 
be the critical control on chromium solubility, with direct enzymic Cr(VI) reduction by 
a consortium of methanogens being implicated (Marsh et al., 2000). 



Molybdenum reduction 

Although microbial Mo (VI) reduction could play a role in the molybdenum cycle, for 
example leading to the concentration of insoluble molybdenum in anaerobic marine 
sediments and reduction spots in rocks (Lovley, 1993), comparatively few studies have 
addressed this process. Early work suggested that 'Pseudomonas guillermondi? and a 
Micrococcus species could reduce Mo (VI) to molybdenum blue (Bautista 6c Alexander, 
1972). More recently, similar activities have been identified in cultures of a molyb- 
denum-resistant Enterobacter species (Ghani et al., 1993). The organism was grown 
under anaerobic conditions in glucose-containing medium supplemented with 200 mM 
Mo (VI). Reduction of Mo (VI) was accompanied by a change in colour, as Mo(V) 
formed and complexed with phosphate in the medium to form methylene blue (Ghani 
et al., 1993). The use of metabolic inhibitors suggested that the electrons for Mo (VI) 
reduction were derived from the glycolytic pathway, whilst NADH functioned as an 
electron donor in broken cells in vitro (Ghani etal., 1993). The ability to reduce Mo (VI) 
has also been identified in pre-grown cells of the sulfate-reducing bacterium Desulfo- 
vibrio desulfuricans, both through direct enzymic mechanisms and indirectly via sulfide 
under sulfate-reducing conditions (Tucker et al., 1997). Cells of Desulfovibrio desulfur- 
icans immobilized in a bioreactor have also been used to remove Mo (VI) from solution 
at high efficiency (Tucker et al., 1998). The organism was unable to grow using Mo (VI) 
as an electron acceptor (Tucker et al., 1997) and it is unlikely that actively growing 
cultures of sulfate-reducing bacteria would play a direct role in reducing high con- 
centrations of Mo (VI) in the environment, given the toxicity of molybdate to these 
organisms (Oremland & Capone, 1988). 



Cobalt reduction 

The reduction of Co (III) has received recent attention because radioactive 60 Co can be a 
problematic contaminant at sites where radioactive waste has been stored. Co (III) is 
especially mobile when complexed with EDTA and several studies have focused on the 
ability of Fe (III) -reducing bacteria to retard the mobility of the metal through reduction 
to Co(II) (Caccavo et al., 1994; Gorby et al., 1998). The Co(II) formed does not asso- 
ciate strongly with EDTA [it is over 25 orders of magnitude less thermodynamically 
stable than Co (III) EDTA] and absorbs to soils, offering the potential for in situ immo- 
bilization of the metal in contaminated soils. The precise mechanisms of dissimilatory 
Co (III) reduction remain to be investigated. 

SGM symposium 65 



Microbial metal reduction 281 

Palladium reduction 

The reduction of soluble Pd(II) to insoluble Pd(0) has also attracted interest, as this 
enzymic process may be used to recover palladium from industrial catalysts (Lloyd et 
al., 1998a) and may also be used to synthesize nanoscale bioinorganic catalysts of 
considerable commercial potential (Yong et al., 2002a). Interest in this area is driven 
by the widespread use of platinum-group metals (PGMs), including palladium, in 
automotive catalytic converters required to reduce gaseous emissions, and problems 
associated with their recycling. With approximately 5 g PGM per catalytic converter, 
the consumption of PGMs was altogether 70312-5 kg in 1994, with only 11 250 kg 
recovered (Lloyd et al., 1998a). The lifetime of a catalytic converter is only approx- 
imately 80000 km (50000 miles), although many fail sooner, and future shortages and 
higher prices may be predicted. Chemical and electrochemical treatments are made 
difficult by complex solution chemistry. 

Initial experiments aimed at reducing and recovering palladium were based on the use 
of Desulfovibrio desulfuricans because it is active against a wide range of metals, 
including Fe (III), Mn(IV), U(VI), Cr(VI) and Tc (VII), via hydrogenase or cytochrome c 3 
(Lloyd et al., 1998a). Cells were able to reduce 0-5 mM Pd(II) [as Pd(NH 3 ) 4 Cl] with a 
range of electron donors, including pyruvate, formate and H 9 . Although the enzyme 
responsible for Pd(II) reduction has not been identified, the involvement of a peri- 
plasmic hydrogenase is implicated by the use of hydrogen as electron donor and 
inhibition by treatment with 0-5 mM Cu 2+ (Lloyd et ai, 1998a). Transmission electron 
microscopic studies, in combination with energy-dispersive X-ray microanalysis, 
confirmed precipitation in the periplasm, with X-ray diffraction studies confirming 
reduction to Pd(0). More recent studies have focused on the recovery of Pd(II) at a range 
of pH values and also from acid (aqua regia) leachates from spent automotive catalysts 
(Yong et al., 2002b). Inhibition by chloride ions was reported and this may therefore 
necessitate leaching from catalysts to minimize the formation of PdCl^ - prior to 
bioreduction and recovery. Delivery of reducing power to an immobilized biocatalyst 
has also been studied in a novel electrobioreactor (Yong et ai, 2002b) containing a 
biofilm of Desulfovibrio desulfuricans immobilized on a Pd-Ag membrane that trans- 
ported atomic hydrogen to the cells, minimizing loss of gaseous hydrogen. Pd(0) 
recovered in the electrobioreactor proved a better catalyst that its chemical counterpart, 
as determined by hydrogen liberation from hypophosphite (Yong et al., 2002a) and 
reduction of several target organics (L. E. Macaskie, personal communication). 

Gold and silver reduction 

It has been argued that Fe(III)-reducing bacteria may play a role in the deposition 
of gold ores, as these organisms are present in high- and moderate-temperature 
sedimentary environments where gold deposits have been recovered (Kashefi et al., 

SGM symposium 65 



282 J. R. Lloyd 



2001). Indeed, several dissimilatory Fe(III)-reducing bacteria and archaea, including 
the hyper thermophilic archaea Pyrobaculum islandicum and Pyro coccus furiosus, the 
hyperthermophilic bacterium Thermotoga maritima and the mesophilic bacteria 
Shewanella algae and Geovibrio ferrireducens, were shown to reduce Au(III) (as gold 
chloride) to insoluble Au(0) (Kashefi et aL, 2001). The ability to reduce Au(III) seems 
to be species-specific, and closely related organisms with similar activities against a 
range of other metals have differing activities against Au(III) (Kashefi et aL, 2001). For 
example, unlike Pyrobaculum islandicum, a close relative, Pyrobaculum aeropbilum, is 
unable to reduce Au(III). Also, there is an obligate requirement for hydrogen as an 
electron donor in organisms that can reduce Au(III), suggesting the involvement of a 
hydrogenase. Given the direct reduction of other metals by hydrogenase (Lloyd et aL, 
1997), it is tempting to hypothesize that hydrogenases may play a direct role in Au(III) 
reduction. 



Microbial reduction of Ag(I) has also been studied, but in little detail. Early reports 
noted that the reduction of Ag(I) may account for resistance to silver in some micro- 
organisms (Belly 6c Kydd, 1982), but more recent studies have uncovered alternative 
strategies for resistance to Ag(I) in organisms isolated from hospital burns wards, where 
silver may be used as a biocide (Gupta et aL, 1999), with a recent review of this new area 
presented by Silver (2003). Several studies, including that by Fu et al. (2000), have shown 
biosorption of Ag(I) to the surface of cells (in this case, a Lactobacillus species), 
followed by reduction to Ag(0). Here, the mechanism for Ag(I) reduction remains 
unknown. 



Mercury reduction 

A well-studied metal-resistance system is encoded by genes of the mer or mercury- 
resistance operon, which relies upon the reduction of Hg(II) (mercuric ions). Here, 
Hg(II) is transported into the cell via the MerT transporter protein and detoxified by 
reduction to relatively non-toxic, volatile elemental mercury by an intracellular 
mercuric reductase (MerA) (Hobman & Brown, 1997). The biotechnological potential 
of this process has been described recently, focusing on the use of mercury-resistant 
bacteria and the proteins that they encode (Lloyd et aL, 2004). Applications include the 
bioremediation of mercury-contaminated water and the development of Hg(II)- 
detecting biosensors. Finally, in addition to the MerA-mediated mechanism of mercury 
reduction, other enzymes are also able to reduce Hg(II). A novel, Fe 2+ -dependent 
mechanism for mercury reduction has been characterized in the membrane fraction of 
Tbiobacillus ferrooxidans, which may involve cytochrome c oxidase (Iwahori et aL, 
2000), and c-type cytochromes of Geobacter metallireducens also reduce Hg(II) (Lovley 

SGM symposium 65 



Microbial metal reduction 283 



et al., 1993). Results from our laboratory have also shown that whole cells of Geobacter 
sulfur re due ens are able to reduce Hg(II) without the involvement of mer mercury- 
resistance genes (N. Law & J. R. Lloyd, unpublished data). 



REDUCTION OF METALLOIDS 

Arsenic reduction: a role in mass poisoning worldwide? 

Contamination of groundwaters, abstracted for drinking and irrigation, by sediment- 
derived arsenic threatens the health of tens of millions of people worldwide, most 
notably in Bangladesh and West Bengal (Chakraborti et al., 2002; Smedley &C Kinni- 
burgh, 2002). Despite the calamitous human-health impacts arising from the extensive 
use of arsenic-enriched groundwaters in these regions, the mechanisms of arsenic 
release from sediments remain poorly characterized and have been a topic of intense 
debate (Nickson et al., 1998; Chowdhury et al., 1999; Oremland & Stolz, 2003). 
However, recent results have suggested that metal-reducing bacteria may be a major 
cause of this humanitarian catastrophe. 

The immediate source of arsenic in these groundwaters in Bengal is widely considered 
to be the host sediments, which are transported by the rivers Ganges, Brahmaputra and 
Meghna and derived from the weathering of the Himalayas within the l-5x 10 6 km 2 
catchment area of these three great river systems. Depending upon the sediment type, 
arsenic typically occurs at concentrations of 2-100 parts per million and is found in, 
and adsorbed onto, a variety of mineralogical hosts, including hydrated ferric oxides, 
phyllosilicates and sulfide minerals (Smedley & Kinniburgh, 2002). The mechanism of 
arsenic release from contaminated sediments remains controversial, but microbially 
mediated release of arsenic from hydrated ferric oxides is gaining the consensus as the 
dominant mechanism for the mobilization of arsenic into groundwater systems of the 
Ganges delta. For example, we recently provided direct evidence for the role of 
indigenous metal-reducing bacteria in the formation of toxic, mobile As(III) in sedi- 
ments from the Ganges delta (Islam et al., 2004). In this study, sediment samples were 
collected at a depth of 13 m from a site in West Bengal known to have relatively high 
concentrations of arsenic in the groundwater. Arsenic mobilization was optimal under 
anaerobic conditions, and addition of acetate to anaerobic sediments, as a proxy for 
organic matter and a potential electron donor for metal reduction, resulted in stimu- 
lation of microbial reduction of Fe(III), followed by As(V) reduction and release 
of mobile As (III). Microbial communities responsible for metal reduction and arsenic 
mobilization in the stimulated anaerobic sediments were analysed by using molecular 
(PCR-based) and cultivation-dependent techniques. Both approaches confirmed an 
increase in numbers of Fe (III) -reducing bacteria and suggested a vital role for metal- 

SGM symposium 65 



284 J. R. Lloyd 



reducing bacteria in mediating arsenic release. Indeed, the PCR studies showed that 
the microbial communities in these sediments were dominated by Geobacter species. 
We have also recently obtained similar results using Cambodian sediments liberating 
arsenic into groundwater (H. A. L. Rowland, R. L. Pederick, D. A. Polya, R. D. Pancost, 
B. E. van Dongen, A. G. Gault, J. M. Charnock, D. J. Vaughan & J. R. Lloyd, unpubl- 
ished data). However, we have also noted recently that Geobacter species (e.g. 
Geobacter sulfurreducens) cannot reduce As(V) enzymically and may actually promote 
the retention of arsenic in sediments, through the formation of Fe (II) minerals that also 
sorb the metalloid (F. S. Islam, R. L. Pederick, A. G. Gault, L. K. Adams, D. A. Polya, 
J. M. Charnock & J. R. Lloyd, unpublished data). These results also show that the 
reduction of Fe(III) alone is not sufficient to mobilize sorbed arsenic, suggesting that 
specialized As(V)-respiring bacteria play a very important role in the reductive 
mobilization of arsenic as mobile, toxic As (III). 

Several organisms capable of growing through dissimilatory reduction of As(V) have 
now been isolated, although none from south-east Asian aquifers. Chrysiogenes arsena- 
tes is a strict anaerobe that was isolated from wastewater from a gold mine (Macy et aL, 
1996). This organism contains a periplasmic arsenate reductase consisting of two sub- 
units (masses 87 and 29 kDa) that contain molybdenum, iron, sulfur and zinc co-factors 
(Krafft & Macy, 1998). Sulfuro spirillum arsenophilum (previously strain MIT 13 T ) is 
a Gram-negative, vibrioid, microaerobic, sulfur-reducing bacterium isolated from 
arsenic-contaminated watershed sediments in eastern Massachusetts, USA. It is a very 
close relative of Sulfuro spirillum barnesii, which can also reduce As(V), and has a 
broad activity against metals, including Fe(III) and Se(VI) (see sections on these metals/ 
metalloids). A recent review has mentioned the preliminary purification and characteri- 
zation of an arsenate reductase from this organism, which constituted a trimeric 
complex of 120 kDa, consisting of subunits of 65, 31 and 22 kDa (Oremland & Stolz, 
2000). A Gram-positive, sulfate-reducing bacterium, Desulfotomaculum auripig- 
mentum, has also been described, which reduces As(V) followed by sulfate, resulting in 
the formation of orpiment (As 2 S 3 ) (Newman et aL, 1997). 

Reduction of As(V) to As(III) by the ArsC reductase also forms the basis of a well- 
studied microbial arsenic-resistance mechanism, preceding efflux of As (III) from the 
cell (Mukhopadhyay et aL, 2002). It should be noted, however, that ArsC-mediated 
As(V) reduction does not support microbial growth and there is currently little evidence 
linking this mechanism of As(V) reduction to the biogeochemical cycling of arsenic. 
Finally, on this note, it is worth mentioning that a novel organism has recently been 
isolated that is able to use As (III) as an electron donor for aerobic growth (Santini et aL, 
2000), thus closing the biological arsenic cycle. 

SGM symposium 65 



Microbial metal reduction 285 



Finally, As(V) is actually a minor (but important) electron acceptor for anaerobic 
respiration in aquifers. However, in some environments, such as Mono Lake, CA, USA, 
As(V) can be the dominant electron acceptor for carbon oxidation (Oremland et aL, 
2000) and here, the role of microbial metabolism in controlling arsenic speciation is far 
clearer. This fascinating environment and the organisms that it supports have been 
reviewed recently (Oremland et al., 2004). 

Reduction of Se(VI) and Se(IV) and other group VIB elements 

In contrast to As(V) reduction, the biotransformation of Se(VI) and Se(IV) to relatively 
unreactive Se(0) results in its removal from water. Several studies have demonstrated 
that these transformations can be catalysed by microbes (Oremland & Stolz, 2000). For 
example, the ability to reduce Se(VI) is widespread in sediments, with biological 
reduction demonstrated unequivocally in 10 out of 11 sediment types (Steinberg & 
Oremland, 1990) . Also, Se(VI) is not reduced chemically under physiological conditions 
of pH and temperature, and Se(VI) reduction is inhibited by autoclaving of sediments. 

Organisms that are known to reduce Se(VI) enzymically include Wolinella succinogenes 
(Tomei et aL, 1992), Desulfovibrio desulfuricans (Tomei et aL, 1995), Pseudomonas 
stutzeri (Lortie et aL, 1992), Enterobacter cloacae (Losi & Frankenberger, 1997) and 
Escherichia coli (Avazeri et aL, 1997). In these examples, Se(VI) reduction does not 
support growth and seems to be incidental to the physiology of the organism. In at least 
one organism (Escherichia coli), the involvement of broad-specificity nitrate reductases 
is implicated by biochemical studies (Avazeri et aL, 1997). In addition to these rather 
non-specific reactions, specialist organisms are known to conserve energy through 
Se(VI) reduction, including Thauera selenatis (Macy & Lawson, 1993), Sulfuro- 
spirillum barnesii [originally 'Geo spirillum barnesif strain SES-3 (Stolz et aL, 1997)] 
and two bacilli (Bacillus arseniciselenatis and Bacillus selenitireducens) , both isolated 
from Mono Lake, CA, USA (Switzer Blum et aL, 1998). Of these four model organisms, 
the mechanism of Se(VI) reduction is best understood in Thauera selenatis (Schroder 
et aL, 1997). A periplasmic complex of approximately 180 kDa (with subunits of 
masses 96, 40 and 23 kDa) has been characterized and shown to contain molybdenum, 
iron and acid-labile sulfur. Specificity for Se(VI) is high, with a K m of 16 jiM. The 
enzyme was unable to reduce nitrate, nitrite, chlorate, chlorite or sulfate. Biochemical 
studies are not so advanced in Sulfuro spirillum barnesii, although the enzyme activity 
contrasts with that characterized in Thauera selenatis, as it is localized in the membrane 
fraction and may have a wider substrate specificity (Stolz et aL, 1997). 

Tellurite (TeOj~) reduction has also been studied in several organisms, mainly in the 
context of resistance to this toxic oxyanion. Indeed, the antibacterial properties of 
Te(IV) have been known for more than 70 years: in the pre-antibiotic era, Te(IV) was 

SGM symposium 65 



286 J. R. Lloyd 



used to treat a range of bacterial infections and Te(IV) remains an ingredient of several 
selective media (e.g. for verocytotoxigenic Escherichia coli 0157). Basal levels of 
resistance to toxic Te(IV) have been attributed to the activity of a membrane-bound 
nitrate reductase in Escherichia coli (Avazeri et aL, 1997). An additional Te(IV) 
reductase was detected in the soluble fraction of anaerobically grown cells. Growth by 
using Te(IV) as an electron acceptor was also reported in an engineered strain over- 
expressing nitrate reductase, but was not thought to be physiologically relevant in wild- 
type cells of Escherichia coli (Avazeri et aL, 1997). Rhodobacter sphaeroides has also 
been reported to reduce Te(IV) (as well as oxyanions of selenium), with an absolute 
requirement for a functional photosynthetic electron-transfer chain under photo- 
synthetic (anaerobic) growth conditions, or functional cytochromes hc x and c 2 under 
aerobic growth conditions (Moore 6c Kaplan, 1992). Again, metal reduction was 
discussed in the context of resistance to the metalloids. Finally, plasmids are known to 
encode several distinct resistance determinants for Te(IV) and, again, Te(IV) reduction 
is implicated as the resistance mechanism, as elemental tellurium is deposited within 
tellurite-resistant bacteria (Taylor, 1999). However, other mechanisms of resistance, 
involving cysteine-metabolizing enzymes and methyl transferases, may be important 
(Taylor, 1999). Finally, Te(IV) [and Se(VI)/Se(IV)] reduction and precipitation by 
sulfate-reducing bacteria has also been reported, in the order Te(IV) >Se(VI) >Se(IV), 
which is in contrast to that predicted by the redox potentials alone (Lloyd et aL, 2001). 
To date, there have been no reports of microbial growth coupled to the reduction 
of Te(IV) by non-genetically engineered micro-organisms. 

REDUCTION OF ACTINIDES AND FISSION PRODUCTS AND THE 
BIOREMEDIATION OF RADIOACTIVE WASTE 

The release of radionuclides from nuclear sites and their subsequent mobility in the 
environment is a subject of intense public concern and has prompted much recent 
research on the environmental fate of key radionuclides (Lloyd & Renshaw, 2005a). The 
major burden of anthropogenic environmental radioactivity is from the controlled 
discharge of process effluents produced by industrial activities allied to the generation 
of nuclear power, although significant quantities of natural and artificial radionuclides 
were also released as a consequence of nuclear weapons testing in the 1950s and 1960s 
via accidental release, e.g. from Chernobyl in 1986, and from the ongoing storage of 
nuclear materials amassed over the last 60 years of nuclear activities. Indeed, the scale 
of our nuclear legacy is enormous, including 120 Department of Energy sites in the 
USA alone, and other facilities in Europe and the former USSR (Lloyd & Renshaw, 
2005a). In several cases, storage has been compromised, leading to contamination of 
trillions of litres of groundwater and millions of cubic metres of contaminated soil and 
debris. The costs of cleaning up these sites are estimated to be in excess of a trillion US 
dollars in the USA alone, and 50 billion pounds sterling in the UK. Given these high 

SGM symposium 65 



Microbial metal reduction 287 



costs and the technical limitations of current chemical-based approaches, there has 
been an unprecedented interest in the interactions of micro-organisms with key 
radionuclides, in the hope of developing cost-effective bioremediation approaches 
for decontamination of sediments and waters affected by nuclear waste (Lloyd et aL, 
2004). 

Because many radionuclides of concern are both redox-active and less soluble when 
reduced, bioreduction offers much promise for controlling the solubility and mobility 
of target radionuclides in contaminated sediments, e.g. the reduction of U(VI) (the 
uranyl ion; UO^ + ) to U(IV) (uraninite; U0 2 ) (Lovley et aL, 1991; Lovley & Phillips, 
1992b) or the reduction of the fission product Tc(VII) (the pertechnetate ion; TCO4) to 
Tc(IV) (Tc0 2 ) (Lloyd et aL, 2000b). Several studies have also addressed the colonization 
of radioactive environments (see Lloyd 6c Renshaw, 2005b) and it would seem that the 
radioactive burden of several nuclear-waste types is not necessarily inhibitory to all 
microbial life. For example, a recent study using pure cultures of bacteria proposed for 
application in bioremediation programmes has addressed the toxicity of actinides, 
metals and chelators. The model organisms tested include Deinococcus radiodurans, 
Pseudomonas putida and Shewanella putrefaciens CN32 (Ruggiero et aL, 2005). 
Actinides, including chelated Pu(IV), U(VI) and Np(V), inhibited growth at millimolar 
concentrations, suggesting that actinide toxicity is primarily chemical (not radio- 
logical) and that radiation resistance (e.g. in Deinococcus species) does not necessarily 
ensure radionuclide tolerance. The author proposes that actinide toxicity will not 
impede bioremediation using naturally occurring bacteria, although the toxicity of 
these key radionuclides remains to be determined in sedimentary environments under 
field conditions. 

Uranium reduction 

The first demonstration of dissimilatory U(VI) reduction was by Lovley et aL (1991), 
who reported that the Fe (III) -reducing bacteria Geobacter metallireducens (previously 
designated strain GS-15 T ) and Shewanella oneidensis (formerly Alteromonas putre- 
faciens and then Shewanella putrefaciens) can conserve energy for anaerobic growth via 
the reduction of LI (VI). It should be noted, however, that the ability to reduce U(VI) 
enzymically is not restricted to Fe (III) -reducing bacteria. Other organisms, including a 
Clostridium species (Francis, 1994) and the sulfate-reducing bacteria Desulfovibrio 
desulfuricans (Lovley & Phillips, 1992a) and Desulfovibrio vulgaris (Lovley & Phillips, 
1994), also reduce uranium, but are unable to conserve energy for growth via this 
transformation. To date, the enzyme system responsible for U(VI) reduction has been 
best studied in Desulfovibrio vulgaris. Purified tetrahaem cytochrome c 3 was shown to 
function as a U(VI) reductase in vitro, in combination with hydrogenase, its physio- 
logical electron donor (Lovley & Phillips, 1994). In vivo studies using a cytochrome c 3 

SGM symposium 65 



288 J. R. Lloyd 




r« 










ft? 







'^Extracellular 
'/' -U(IV) 



*0i# 




/ 

Periplasmic U(IV) 



Acetate 



CO- 




U(VI) 

soluble 



(b) 




U(V) "=<^ Disproportionation 

unstable 



iO.» 



OMfi u(iv) 

p _ ■ * * insoluble 



Fig. 3. Reduction of U(VI) to insoluble U(IV) (electron-dense deposits) in thin sections of Geobacter 
sulfurreducens, viewed by using transmission electron microscopy (a). U(VI) reduction by this organism 
is via unstable U(V), which disproportionates to give insoluble U(IV) [and U(VI) for further reduction] 
(b). Transmission electron microscopy by S. Glasauer (University of Guelph, Canada). Bar, 0-5 \xm. 



mutant of the close relative Desulfovibrio desulfuricans strain G20 confirmed a role for 
cytochrome c 3 in hydrogen-dependent U(VI) reduction, but suggested additional 
pathways from organic electron donors to U(VI) that bypassed the cytochrome (Payne 
et al., 2002). Similar cytochrome-mediated mechanisms have been proposed in 
Geobacter species, whilst U(VI) reduction in a Shewanella putrefaciens strain shares 
components of the nitrite-reducing pathway in this organism (Wade ck: DiChristina, 
2000). The mechanism of U(VI) reduction by Geobacter sulfurreducens has also been 
studied recently in detail by using X-ray absorbance spectroscopy, which showed the 
formation of an unstable U(V) intermediate (Renshaw et al., 2005). This organism was 
unable to reduce the stable analogue Np(V), suggesting that the further reduction of 
U(V) is by disproportionation to U(IV) [with U(VI) generated available for further 
reduction] (Fig. 3). This surprising level of specificity of hexavalent actinides illustrates 



SGM symposium 65 



Microbial metal reduction 289 



a need for detailed investigations on the impact of micro-organisms on complex, 
actinide-containing wastes. 

Field studies on uranium bioreduction in situ 

There has been a considerable level of interest in harnessing the metabolism of U(VI)- 
reducing bacteria for the bioremediation of uranium-contaminated aquifers. For 
example, recent studies focused on biostimulation of U(VI) -reducing bacteria at a 
'Uranium Mill Tailings Remedial Action' (UMTRA) site in Rifle, CO, USA, through 
the injection of an electron donor (acetate) into the subsurface (Anderson et al., 2003). 
The decrease in soluble U(VI) was coincident with an increase in Fe(II) in the 
groundwater and a significant enrichment of Geobacter species. However, after 39 days, 
the composition of the microbial community began to change as sulfate-reducing 
organisms dominated, and soluble U(VI) increased with a decrease in Fe(II), acetate and 
sulfate and an accumulation of sulfite. Thus, the precise constituents of the microbial 
communities present in the sediments clearly exert control on U(VI) speciation and 
require careful optimization. The geochemistry of the groundwaters should also not 
be overlooked. For example, Ca~ + cations at millimolar concentrations cause a signifi- 
cant decrease in the rate and extent of bacterial U(VI) reduction by a range of metal- 
reducing bacteria, suggesting that U(VI) is a less effective electron acceptor when 
present as the Ca 2 U02(C0 3 )3 complex (Brooks et al., 2003). High nitrate concen- 
trations can also inhibit U(VI) reduction by acting as a competing electron acceptor, 
and may even promote reoxidation of reduced U(IV) (Istok et al., 2004). Mineralogical 
constraints can also be important in controlling the end points for uranium bio- 
remediation. For example, although much work has focused on the reduction of soluble 
U(VI), the fate of sorbed U(VI), which can be appreciable in sediments, is also 
potentially important. For example, although soluble U(VI) was reduced in a slurry 
prepared from sediments from Rifle, CO, USA, sorbed U(VI) was not reduced and was 
deemed 'not bioavailable' for microbial reduction (Ortiz-Bernad et al., 2004b). 

Reduction of other actinides (plutonium and neptunium) 

Although 2o8 U remains the priority pollutant in most medium- and low-level radio- 
active wastes, other actinides, including 230 Th, 237 Np, 241 Pu and 241 Am, can also be 
present (Macaskie, 1991; Lloyd & Macaskie, 2000). Th(IV) and Am (III) are stable 
across most E h values encountered in radionuclide-contaminated waters, but the 
potentials for Pu(V)/Pu(IV) and Np(V)/Np(IV), in common with that of U(VI)/U(IV), 
are more electropositive than the standard redox potential of ferrihydrite/Fe~ + (approx. 
V; Thamdrup, 2000). Thus, Fe(III)-reducing bacteria have the metabolic potential 
to reduce these radionuclides enzymically or via Fe(II) produced from the reduction of 
Fe(III) oxides. This is significant because the tetravalent actinides are amenable to 
bioremediation, due to their high ligand-complexing abilities (Lloyd & Macaskie, 

SGM symposium 65 



290 J. R. Lloyd 



2000), and are also immobilized in sediments containing active biomass (Peretrukhin 
et al, 1996). Thus, although it is possible for Fe(III)-reducing bacteria to reduce and 
precipitate actinides directly, e.g. the reduction of soluble U(VI) to insoluble U(IV) (see 
above), some transformations do not result in formation of an insoluble mineral phase, 
but in the formation of a cation more amenable to bioprecipitation. This is illustrated 
when considering highly soluble Np(V) (NpO?), which was reduced to soluble Np(IV) 
by Shewanella putrefaciens, with the Np(IV) removed as an insoluble phosphate 
biomineral by a phosphate-liberating Citrobacter species (Lloyd et al, 2000a). This is in 
sharp contrast to the case in Geobacter sulfur re due ens, which is unable to reduce 
Np(V) (see above). Also, some studies have suggested that the reduction of Pu(IV) to 
Pu(III) can be achieved by Fe (III) -reducing bacteria, although the Pu(III) was reported 
to reoxidize spontaneously (Rusin et al, 1994). Although this may lead to solublization 
of sediment-bound Pu(IV), it will yield a trivalent actinide that is also amenable to 
bioremediation by using a range of microbially produced ligands (Lloyd & Macaskie, 
2000). The biochemical basis of these transformations remains uncharacterized. 

Technetium reduction 

The fission product technetium is another long-lived radionuclide that is present in 
nuclear waste and has attracted considerable recent interest. This is due to a combin- 
ation of its mobility as the soluble pertechnetate ion [Tc(VII); TCO4], bioavailability as 
an analogue of sulfate and long half-life (2-13 x 10 5 years) (Wildung et al, 1979). Like 
Np(V), Tc(VII) has weak ligand-complexing capabilities and is difficult to remove from 
solution by using conventional 'chemical' approaches. Several reduced forms of the 
radionuclide are insoluble, however, and metal-reducing micro-organisms can reduce 
Tc(VII) and precipitate the radionuclide as a low-valency oxide [Tc(IV); Tc0 2 ]. 

In an early study on Tc(VII) bioreduction, a novel phosphorimaging technique was used 
to show reduction of the radionuclide by Shewanella putrefaciens and Geobacter 
metallireducens, with similar activities subsequently detected in laboratory cultures of 
Rbodobacter sphaeroides, Paracoccus denitrificans, some pseudomonads (Lloyd et al, 
2002), Escherichia coli (Lloyd et al., 1997) and a range of sulfate-reducing bacteria 
(Lloyd et al., 1998b, 1999b, 2001). Other workers have used this technique to show that 
Thiobacillus ferrooxidans and Thiobacillus thiooxidans (Lyalikova & Khizhnyak, 
1996) and the hyperthermophile Pyrobaculum islandicum (Kashefi & Lovley, 2000) are 
also able to reduce Tc(VII). It should be stressed that Tc(VII) reduction has not been 
shown to support growth in any of these studies, and seems to be a fortuitous 
biochemical side reaction in the organisms studied to date. Recent work has also shown 
that Tc(VII) can be reduced through indirect microbial processes via, for example, 
biogenic sulfide (Lloyd et al, 1998b), Fe(II) (Lloyd et al, 2000b) or U(IV) (Lloyd et al, 
2002). Tc(VII) reduction and precipitation by biogenic Fe(II) are particularly efficient 

SGM symposium 65 



Microbial metal reduction 291 



and may offer a potentially useful mechanism for the remediation of technetium- 
contaminated sediments containing active Fe (III) -reducing bacteria (Lloyd et al., 
2000b). 

This latter point has been confirmed by several recent studies using a range of sediment 
materials. In one study, sediments from the Humber Estuary, UK, were left to age and 
exhibited a clear progression of terminal electron-accepting processes (Burke et al., 
2005). The reduction and precipitation of Tc(VII) were associated with the formation 
of biogenic Fe(II) and were catalysed by pure cultures of Fe(III)-reducing prokaryotes 
inoculated in sterilized microcosms. Technetium solubility has also been studied by 
using core samples from a shallow, sandy aquifer located on the US Atlantic Coastal 
Plain (Wildung et al., 2004). The dominant electron donor in the sediments was Fe(II) 
(0-5 M HCl-extractable), with Tc(IV) hydrous oxide being the major solid-phase reduc- 
tion product. The authors noted presumptive evidence for direct enzymic reduction 
in only a few key sand samples. The potential for biogenic Fe (II) -mediated reduction of 
Tc(VII) has also been assessed in detail in other studies, e.g. by using sediments from the 
US Department of Energy's Hanford and Oak Ridge sites (Fredrickson et al., 2004). 

The biochemical basis of Tc(VII) reduction has been best studied in Escherichia coli. 
Initial studies demonstrated that anaerobic, but not aerobic, cultures of Escherichia 
coli reduced Tc(VII), with the reduced radionuclide precipitated within the cell (Lloyd 
et al., 1997). Results obtained from studies conducted with wild-type cells and 34 
defined mutants defective in the synthesis of regulatory or electron-transfer proteins 
were used to construct a model for Tc(VII) reduction by Escherichia coli. The central 
tenet of this model is that the hydrogenase 3 component of formate hydrogenlyase 
catalyses the transfer of electrons from dihydrogen to Tc(VII) (Fig. 4). According to this 
model, the formate dehydrogenase component (FdhH) is required only if formate, or a 
precursor, is supplied as an electron donor for Tc(VII) reduction in place of hydrogen. 
This model has been validated by the observations that a mutant unable to synthesize 
hydrogenase 3 was unable to reduce Tc(VII) when either hydrogen or formate was 
supplied as an electron donor (Lloyd et al., 1997). Hydrogenase-mediated Tc(VII) 
reduction has also been noted in sulfate-reducing bacteria (Lloyd et al., 1999b; De Luca 
et al., 2001) and Geobacter sulfur re ducens (J. C. Renshaw 6c J. R. Lloyd, unpublished 
observations). 



DEGRADATION OFXENOBIOTICS BY METAL-REDUCING 
BACTERIA 

Some metal-reducing bacteria, most notably Geobacter species, have the ability to 
couple Fe(III) reduction to the complete oxidation of aromatic contaminants (Lovley et 
al., 1989b; Lovley 6c Lonergan, 1990; Coates et al., 2001). It is possible to stimulate 

SGM symposium 65 



292 J. R. Lloyd 



(a) 



Formate 



FdhH 




FNR 




2H 



+ 



Hydrogenase 3 




H. 



Mo 



Tc(VII) 



Tc(IV) 




25- 














(c) 






Cu 




"D 20 








c 
o 


Tc cl 






u 

"> 15- 









QJ 








Q. 










£ 10 










D 










O 


Cu P 








5- 
0.. 


i — <— 


u I 

— i — i — i — i — « 


i — . — , — . — , — , — .- 


Tc 

JU 







10 



15 



20 



Energy (keV) 



Fig. 4. Mechanism of Tc(VII) reduction by E. coli. Hydrogenase 3 of the formate hydrogenlyase 
complex is able to catalyse the reduction of soluble Tc(VII) to insoluble Tc(IV) (a). Hydrogen or formate 
are suitable electron donors; with the latter a formate dehydrogenase (FdhH) is also required for 
reduction of Tc(VII). The insoluble Tc(IV) is precipitated within the cell, visible as an electron dense 
deposit in TEM images of thin sections of the cells (b) and confirmed by EDS analysis (c). Bar, 1 jim. 

these activities by increasing the bioavailability of Fe(III) oxides in subsurface sediments 
by, for example, using Fe(III) chelators that solublize Fe(III) (Lovley et al., 1994) or 
humic acids that can act as electron shuttles between the Fe(III) -reducing species and 
Fe(III) oxides (Lovley et al., 1996). Both approaches eliminate the need for the organism 
to contact the insoluble Fe(III) oxide directly to reduce it. 



SGM symposium 65 



Microbial metal reduction 293 



(a) 



Xenobiotic hydrazone and azo bonds are part of the chromophore 




H0 3 SOCH 2 CH 2 -S— t ^-N— N y J" 2 ^=N— I Vs-CH.CH.OSO^ 



H0 3 S 




« 



SO3H 



Xenobiotic aromatic sulfonic acid groups make the dye highly soluble 



(b) 




Fig. 5. Reduction and decoloration of the azo dye remazol black B (a) by anaerobically grown cells of 
Shewanella sp. J 1 81 43 (b). Figures were generously provided by C. Pearce (University of Manchester, 
UK). 



In addition to coupling the oxidation of aromatics to metal reduction, Fe (III) -reducing 
bacteria can also reduce redox-active organic xenobiotics in lieu of Fe(III). A good 
example is the reduction of recalcitrant, coloured azo dyes, the disposal of which poses 
a considerable problem to textile-dyeing industries worldwide. Due to the relatively low 
levels of dye-fibre fixation in current reactive dyeing processes, as much as 50 % of the 
dye that is present in the original dyebath is lost to the wastewater. Physical and/or 
chemical processes are available, but are costly and can generate problematic 
concentrated sludges for disposal. An alternative procedure utilizes anaerobically 
grown cultures of a Shewanella species (designated strain J18143), isolated from soil 
contaminated with textile dyes, to reduce and decolorize textile wastewaters (Nelson et 
al., 2000). This highly efficient biocatalyst has been incorporated into the recently 
developed BIOCOL commercial process for the treatment of azo dyes (Conlon & 
Khraisheh, 2002). In this process, the bacterial cells are immobilized on an activated 



SGM symposium 65 



294 J. R. Lloyd 



carbon support that adsorbs the target dye molecules and the potentially toxic amine 
breakdown products for further biodegradation. Recent studies from our laboratory 
have studied the underlying physiology of this organism, which is able to reduce a 
wide range of azo compounds (e.g. remazol black B; see Fig. 5) and is compatible with 
realistic process conditions, including moderate temperatures and alkaline pH 
(C. Pearce, J. Guthrie & J. R. Lloyd, unpublished data). 

CONCLUSIONS 

Although the environmental relevance of microbial metal-reduction processes has only 
recently become apparent, rapid advances in the understanding of these important 
biotransformations have been made. However, we still have much to learn about the 
precise mechanisms involved and the full impact of such reactions on a range of 
biogeochemical cycles. Given the availability of genomic sequences for key metal- 
reducing micro-organisms, new post-genomic approaches and the possibility of 
combining these tools with advanced techniques from other branches of science and 
technology (e.g. isotopic, spectroscopic and computational tools), rapid advances in 
these areas are predicted. 

ACKNOWLEDGEMENTS 

The author thanks the UK Natural Environment Research Council, Biotechnology and 
Biological Sciences Research Council and Engineering and Physical Sciences Research Council 
and the Natural and Accelerated Bioremediation Research (NABIR) programme of the US 
Department of Energy for financial support. 

REFERENCES 

Anderson, R. T., Rooney-Varga, J. N., Gaw, C. V. & Lovley, D. R. (1998). Anaerobic 

benzene oxidation in the Fe(lll) reduction zone of petroleum-contaminated aquifers. 

Environ Sci Technol 32, 1 222-1 229. 
Anderson, R. T, Vrionis, H. A., Ortiz-Bernad, I. & 10 other authors (2003). Stimulating 

the in situ activity of Geobacter species to remove uranium from the groundwater of 

a uranium-contaminated aquifer. Appl Environ Microbiol 69, 5884-5891 . 
Avazeri, C, Turner, R. J., Pommier, J., Weiner, J. H., Giordano, G. & Vermeglio, A. 

(1997). Tellurite reductase activity of nitrate reductase is responsible for the basal 

resistance of Escherichia coli to tellurite. Microbiology 143, 1181-11 89. 
Bautista, E. M. & Alexander, M. (1972). Reduction of inorganic compounds by soil 

microorganisms. Soil Sci Soc Am Proc 36, 918-920. 
Beliaev, A. S., Saffarini, D. A., McLaughlin, J. L & Hunnicutt, D. (2001). MtrC, an outer 

membrane decahaem c cytochrome required for metal reduction in Shewanella 

putrefaciens MR-1 . Mol Microbiol 39, 722-730. 
Belly, R. T. & Kydd, G. C. (1982). Silver resistance in microorganisms. Dev I nd Microbiol 23, 

567-577. 
Bond, D. R., Holmes, D. E., Tender, L. M. & Lovley, D. R. (2002). Electrode-reducing 

SGM symposium 65 



Microbial metal reduction 295 



microorganisms that harvest energy from marine sediments. Science 295, 483- 

485. 
Brooks, S. C, Fredrickson, J. K., Carroll, S. L, Kennedy, D. W., Zachara, J. M., Plymale, 

A. E., Kelly, S. D., Kemner, K. M. & Fendorf, S. (2003). Inhibition of bacterial U(VI) 

reduction by calcium. Environ Sci Technol 37, 1 850-1 858. 
Bruins, M. R., Kapil, S. & Oehme, F. W. (2000). Microbial resistance to metals in the 

environment. Ecotoxicol Environ Saf 45, 1 98-207. 
Burke, I. T., Boothman, C, Lloyd, J. R., Mortimer, R. J. G., Livens, F. R. & Morris, K. 

(2005). Effects of progressive anoxia on the solubility of technetium in sediments. 

Environ Sci Technol 39, 41 09-^1 1 6. 
Caccavo, F., Jr, Lonergan, D. J., Lovley, D. R., Davis, M., Stolz, J. F. & Mclnerney, M. J. 

(1994). Geobacter sulfurreducens sp. nov., a hydrogen- and acetate-oxidizing 

dissimilatory metal-reducing microorganism. Appl Environ Microbiol 60, 3752-3759. 
Chakraborti, D., Rahman, M. M., Paul, K., Chowdhury, U. K., Sengupta, M. K., 

Lodh, D., Chanda, C. R., Saha, K. C. & Mukherjee, S. C. (2002). Arsenic calamity 

in the Indian subcontinent: what lessons have been learned? Talanta 58, 3-22. 
Childers, S. E., Ciufo, S. & Lovley, D. R. (2002). Geobacter metallireducens accesses 

insoluble Fe(lll) oxide by chemotaxis. Nature 416, 767-769. 
Chowdhury, T. R., Basu, G. K., Mandal, B. K. & 11 other authors (1999). Arsenic 

poisoning in the Ganges delta. Nature 401, 545-546. 
Coates, J. D., Bhupathiraju, V. K., Achenbach, L. A., Mclnerney, M. J. & Lovley, D. R. 

(2001). Geobacter hydrogenophilus, Geobacter chapel lei and Geobacter grbiciae, 

three new, strictly anaerobic, dissimilatory Fe(lll)-reducers. Int J Syst Evol Microbiol '51 , 

581-588. 
Coker, V., Pattrick, R. A. D., van der laan, G. & Lloyd, J. R. (2004). Use of Bacteria to 

Produce Spinel Nanopartides. UK Patent 0424636.9. 
Conlon, M. & Khraisheh, M. (2002). Bioadsorption Process for the Removal of Colour from 

Textile Effluent. Patent WO0242228. 
Coppi, M. V., Leang, C, Sandler, S. J. & Lovley, D. R. (2001). Development of a genetic 

system for Geobacter sulfurreducens. Appl Environ Microbiol 67, 3 1 80-3 1 87. 
De Luca, G., de Philip, P., Dermoun, Z., Rousset, M. & Vermeglio, A. (2001). Reduction 

of technetium(VII) by Desulfovibrio fructosovorans is mediated by the nickel-iron 

hydrogenase. Appl Environ Microbiol 67, 4583-4587. 
DiChristina, T. J., Moore, C. M. & Haller, C. A. (2002). Dissimilatory Fe(lll) and Mn(IV) 

reduction by Shewanella putrefaciens requires ferE, a homolog of the pulE (gspE) 

type II protein secretion gene.JBacteriol 184, 142-1 51 . 
Fendorf, S. E. & Li, G. (1996). Kinetics of chromate reduction by ferrous iron. Environ Sci 

Technol 30, 1614-1617. 
Francis, A. J. (1 994). Microbial transformations of radioactive wastes and environmental 

restoration through bioremediation. J Alloys Comp 213/214, 226-231 . 
Fredrickson, J. K., Zachara, J. M., Kennedy, D. W., Kukkadapu, R. K., McKinley, J. P., 

Heald, S. M., Liu, C. & Plymale, A. E. (2004). Reduction of TCO4 by sediment- 
associated biogenic Fe(ll). Geochim Cosmochim Acta 68, 3171-3187. 
Fu, J. K., Liu, Y. Y, Gu, P. Y, Tang, D. L, Lin, Z. Y, Yao, B. X. & Weng, S. Z. (2000). 

Spectroscopic characterization on the biosorption and bioreduction of Ag + by 

Lactobacillus A09. Acta Phys-Chim Sin 1 6, 779-782. 
Fude, L, Harris, B., Urrutia, M. M. & Beveridge, T. J. (1994). Reduction of Cr(VI) by a 

consortium of sulfate-reducing bacteria (SRB III). Appl Environ Microbiol 60, 

1525-1531. 



SGM symposium 65 



296 J. R. Lloyd 



Gaspard, S., Vazquez, F. & Holliger, C. (1998). Localization and solubilization of the iron 

reductase of Geobacter sulfurreducens. Appl Environ Microbiol 64, 3 1 88-3 1 94. 
Ghani, B., Takai, M., Hisham, N. Z., Kishimoto, N., Ismail, A. K. M., Tano, T. & Sugio, T. 

(1993). Isolation and characterization of a Mo 6+ -reducing bacterium. Appl Environ 

Microbiol 59, 1176-1180. 
Gorby, Y. A., Caccavo, R, Jr & Bolton, H., Jr (1998). Microbial reduction of cobalt IH EDTA" 

in the presence and absence of manganese(IV) oxide. Environ Sci Techno! 32, 

244-250. 
Gupta, A., Matsul, K., Lo, J.-F. & Silver, S. (1999). Molecular basis for resistance to silver 

cations in Salmonella. Nat Med 5, 1 83-1 88. 
Hobman, J. L. & Brown, N. L. (1997). Bacterial mercury- resistance genes. Met Ions Biol Syst 

34, 527-568. 
Holmes, D. E., Finneran, K. T., O'Neil, R. A. & Lovley, D. R. (2002). Enrichment of 

members of the family Geobacteraceae associated with stimulation of dissimilatory 

metal reduction in uranium-contaminated aquifer sediments. Appl Environ Microbiol 

68, 2300-2306. 
Islam, F. S., Gault, A. G., Boothman, C, Polya, D. A., Charnock, J. M., Chatterjee, D. & 

Lloyd, J. R. (2004). Role of metal-reducing bacteria in arsenic release from Bengal 

delta sediments. Nature 430, 68-71 . 
Istok, J. D., Senko, J. M., Krumholz, L. R., Watson, D., Bogle, M. A., Peacock, A., 

Chang, Y.-J. & White, D. C. (2004). In situ bioreduction of technetium and uranium 

in a nitrate-contaminated aquifer. Environ Sci Technol 38, 468-475. 
Iwahori, K., Takeuchi, F., Kamimura, K. & Sugio, T. (2000). Ferrous iron-dependent 

volatilization of mercury by the plasma membrane of Thiobacillus ferrooxidans. Appl 

Environ Microbiol 66, 3823-3827. 
Kashefi, K. & Lovley, D. R. (2000). Reduction of Fe(lll), Mn(IV), and toxic metals at 100 °C 

by Pyrobaculum islandicum. Appl Environ Microbiol 66, 1 050-1 056. 
Kashefi, K. & Lovley, D. R. (2003). Extending the upper temperature limit for life. Science 

301,934. 
Kashefi, K., Tor, J. M., Nevin, K. P. & Lovley, D. R. (2001). Reductive precipitation of gold 

by dissimilatory Fe(lll)-reducing Bacteria and Archaea. Appl Environ Microbiol 67, 

3275-3279. 
Krafft, T. & Macy, J. M. (1998). Purification and characterization of the respiratory arsenate 

reductase of Chrysiogenes arsenatis. EurJBiochem 255, 647-653. 
Lloyd, J. R. (2003). Microbial reduction of metals and radionuclides. FEMS Microbiol Rev 27, 

411-425. 
Lloyd, J. R. & Lovley, D. R. (2001). Microbial detoxification of metals and radionuclides. 

CurrOpin Biotechnol 12, 248-253. 
Lloyd, J. R. & Macaskie, L. E. (2000). Bioremediation of radionuclide-containing 

wastewaters. In Environmental Microbe-Metal Interactions, pp. 27/ '-327 '. Edited by 

D. R. Lovley. Washington, DC: American Society for Microbiology. 
Lloyd, J. R. & Renshaw, J. C. (2005a). Microbial transformations of radionuclides: 

fundamental mechanisms and biogeochemical implications. Met Ions Biol Syst 

44, 205-240. 
Lloyd, J. R. & Renshaw, J. C. (2005b). Bioremediation of radioactive waste: radionuclide- 

microbe interactions in laboratory and field-scale studies. Curr Opin Biotechnol 

16,254-260. 
Lloyd, J. R., Cole, J. A. & Macaskie, L. E. (1997). Reduction and removal of heptavalent 

technetium from solution by Escherichia coli.J Bacteriol 179, 2014-2021 . 



SGM symposium 65 



Microbial metal reduction 297 



Lloyd, J. R., Yong, P. & Macaskie, L. E. (1998a). Enzymatic recovery of elemental palladium 

by using sulfate-reducing bacteria. Appl Environ Microbiol 64, 4607^4609. 
Lloyd, J. R., Nolting, H.-R, Sole, V. A., Bosecker, K. & Macaskie, L E. (1998b). 

Technetium reduction and precipitation by sulfate-reducing bacteria. Geomicrobiol J 

15,45-58. 
Lloyd, J. R., Blunt-Harris, E. L & Lovley, D. R. (1999a). The periplasmic 9-6-kilodalton 

c-type cytochrome of Geobacter sulfurreducens is not an electron shuttle to Fe(lll). 

JBacteriol 181, 7647-7649. 
Lloyd, J. R., Ridley, J., Khizniak, T., Lyalikova, N. N. & Macaskie, L E. (1999b). Reduction 

of technetium by Desulfovibrio desulfuricans: biocatalyst characterization and use in 

a flowthrough bioreactor. Appl Environ Microbiol 65, 2691-2696. 
Lloyd, J. R., Yong, P. & Macaskie, L. E. (2000a). Biological reduction and removal of Np(V) 

by two microorganisms. Environ Sci Technol 34, 1 297-1 301 . 
Lloyd, J. R., Sole, V. A., Van Praagh, C. V. G. & Lovley, D. R. (2000b). Direct and Fe(ll)- 

mediated reduction of technetium by Fe(lll)-reducing bacteria. Appl Environ 

Microbiol 66, 3743-3749. 
Lloyd, J. R., Mabbett, A. N., Williams, D. R. & Macaskie, L E. (2001). Metal reduction 

by sulphate-reducing bacteria: physiological diversity and metal specificity. 

Hydrometallurgy 59, 327-337. 
Lloyd, J. R., Chesnes, J., Glasauer, S., Bunker, D. J., Livens, F. R. & Lovley, D. R. (2002). 

Reduction of actinidesand fission products by Fe(lll)-reducing bacteria. Geomicrobiol 

719, 103-120. 
Lloyd, J. R., Leang, C, Hodges Myerson, A. L., Coppi, M. V., Ciufo, S., Methe, B., 

Sandler, S. J. & Lovley, D. R. (2003). Biochemical and genetic characterization 

of PpcA, a periplasmic c-type cytochrome in Geobacter sulfurreducens. Biochem J 

369, 153-161. 
Lloyd, J. R., Lovley, D. R. & Macaskie, L. E. (2004). Biotechnological application of metal- 
reducing microorganisms. Adv Appl Microbiol 53, 85-128. 
Lortie, L, Gould, W. D., Rajan, S., McCready, R. G. L & Cheng, K.-J. (1992). Reduction 

of selenate and selenite to elemental selenium by a Pseudomonas stutzeri isolate. 

Appl Environ Microbiol 58, 4042-4044. 
Losi, M. E. & Frankenberger, W. T., Jr (1997). Reduction of selenium oxyanions 

by Enterobacter cloacae SLD1a-1 : isolation and growth of the bacterium and its 

expulsion of selenium particles. Appl Environ Microbiol 63, 3079-3084. 
Lovley, D. R. (1991). Dissimilatory Fe(lll) and Mn(IV) reduction. Microbiol Rev 55, 259-287. 
Lovley, D. R. (1993). Dissimilatory metal reduction. Annu Rev Microbiol 47, 263-290. 
Lovley, D. R. (1997). Potential for anaerobic bioremediation of BTEX in petroleum- 
contaminated aquifers. J Ind Microbiol Biotechnol 18, 75-81 . 
Lovley, D. R. & Anderson, R. T. (2000). Influence of dissimilatory metal reduction on fate of 

organic and metal contaminants in the subsurface. Hydrogeol J 8, 77-88. 
Lovley, D. R. & Chapel le, F. H. (1995). Deep subsurface microbial processes. Rev Geophys 

33, 365-382. 
Lovley, D. R. & Lonergan, D. J. (1990). Anaerobic oxidation of toluene, phenol, and 

p-cresol by the dissimilatory iron-reducing organism, GS-1 5. Appl Environ Microbiol 

56, 1858-1864. 
Lovley, D. R. & Phillips, E. J. P. (1992a). Reduction of uranium by Desulfovibrio 

desulfuricans. Appl Environ Microbiol 58, 850-856. 
Lovley, D. R. & Phillips, E. J. P. (1992b). Bioremediation of uranium contamination with 

enzymatic uranium reduction. Environ Sci Techno! 26, 2228-2234. 



SGM symposium 65 



298 J. R. Lloyd 



Lovley, D. R. & Phillips, E. J. P. (1994). Reduction of chromate by Desulfovibrio vulgaris and 
its c 3 cytochrome. Appl Environ Microbiol 60, 726-728. 

Lovley, D. R., Stolz, J. R, Nord, G. L & Phillips, E. J. P. (1987). Anaerobic production of 
magnetite by a dissimilatory iron-reducing microorganism. Nature 330, 252-254. 

Lovley, D. R., Phillips, E. J. P. & Lonergan, D. J. (1989a). Hydrogen and formate oxidation 
coupled to dissimilatory reduction of iron or manganese by Alteromonas putre- 
faciens. Appl Environ Microbiol 55, 700-706. 

Lovley, D. R., Baedecker, M. J., Lonergan, D. J., Cozzarelli, I. M., Phillips, E. J. P. & 
Siegel, D. I. (1989b). Oxidation of aromatic contaminants coupled to microbial iron 
reduction. Nature 339, 297-300. 

Lovley, D. R., Phillips, E. J. P., Gorby, Y. A. & Landa, E. R. (1991). Microbial reduction of 
uranium. Nature 350, 41 3-41 6. 

Lovley, D. R., Giovannoni, S. J., White, D. C, Champine, J. E., Phillips, E. J. P., Gorby, 
Y. A. & Goodwin, S. (1993). Geobacter metallireducens gen. nov. sp. nov., a micro- 
organism capable of coupling the complete oxidation of organic compounds to the 
reduction of iron and other metals. Arch Microbiol 1 59, 336-344. 

Lovley, D. R., Woodward, J. C. & Chapelle, F. H. (1994). Stimulated anoxic biodegradation 
of aromatic hydrocarbons using Fe(lll) ligands. Nature 370, 128-131 . 

Lovley, D. R., Coates, J. D., Blunt-Harris, E. L, Phillips, E. J. P. & Woodward, J. C. (1996). 
Humic substances as electron acceptors for microbial respiration. Nature 382, 
445-448. 

Lovley, D. R., Holmes, D. E. & Nevin, K. P. (2004). Dissimilatory Fe(lll) and Mn(IV) 
Reduction . Adv Microb Physiol 49, 2 1 9-286. 

Lyalikova, N. N. & Khizhnyak, T. V. (1996). Reduction of heptavalent technetium by acido- 
philic bacteria of the genus Thiobacillus. Mikrobiologiia 65, 468-473 (in Russian). 

Mabbett, A. N., Lloyd, J. R. & Macaskie, L. E. (2002). Effect of complexing agents 
on reduction of Cr(VI) by Desulfovibrio vulgaris ATCC 29579. Biotech nol Bioeng 
79, 389-397. 

Macaskie, L. E. (1991). The application of biotechnology to the treatment of wastes 
produced from the nuclear fuel cycle: biodegradation and bioaccumulation as 
a means of treating radionuclide-containing streams. Crit Rev Biotech nol 11, 41- 
112. 

Macy, J. M. & Lawson, S. (1993). Cell yield (Y M ) of Thauera selenatis grown anaerobically 
with acetate plus selenate or nitrate. Arch Microbiol 160, 295-298. 

Macy, J. M., Nunan, K., Hagen, K. D., Dixon, D. R., Harbour, P. J., Cahill, M. & Sly, 
L. I. (1996). Chrysiogenes arsenatis gen. nov., sp. nov., a new arsenate-respiring 
bacterium isolated from gold mine wastewater. Int J Syst Bacteriol 46, 1 1 53-1 1 57. 

Magnuson, T. S., Hodges-Myerson, A. L. & Lovley, D. R. (2000). Characterization of 
a membrane-bound NADH-dependent Fe 3+ reductase from the dissimilatory 
Fe 3+ -reducing bacterium Geobacter sulfurreducens. FEMS Microbiol Lett 185, 205- 
211. 

Marsh, T. L., Leon, N. M. & Mclnerney, M. J. (2000). Physiochemical factors affecting 
chromate reduction by aquifer materials. Geomicrobiol J 17, 291-303. 

Moore, M. D. & Kaplan, S. (1992). Identification of intrinsic high-level resistance to rare- 
earth oxides and oxyanions in members of the class Proteobacteria: characterization 
of tellurite, selenite, and rhodium sesquioxide reduction in Rhodobacter sphaeroides. 
J Bacteriol 17 4, 1505-1514. 

Mukhopadhyay, R., Rosen, B. P., Phung, L. T. & Silver, S. (2002). Microbial arsenic: from 
geocycles to genes and enzymes. FEMS Microbiol Rev 26, 31 1-325. 



SGM symposium 65 



Microbial metal reduction 299 



Myers, C. R. & Myers, J. M. (1992). Localization of cytochromes to the outer membrane of 

anaerobically grown Shewanella putrefaciens MR-1 J Bacterid 174, 3429-3438. 
Myers, C. R. & Myers, J. M. (1993). Ferric reductase is associated with the membranes of 

anaerobically grown Shewanella putrefaciens MR-1. FEMS Microbiol Lett 108, 

15-22. 
Myers, C. R. & Myers, J. M. (1997). Cloning and sequence of cymA, a gene encoding a 

tetraheme cytochrome c required for reduction of iron(lll), fumarate, and nitrate by 

Shewanella putrefaciens MR-1 . J Bacteriol 1 79, 1 1 43-1 1 52. 
Myers, C. R. & Nealson, K. H. (1988). Bacterial manganese reduction and growth with 

manganese oxide as the sole electron acceptor. Science 240, 1319-1 321 . 
Myers, J. M. & Myers, C. R. (2000). Role of the tetraheme cytochrome CymA in anaerobic 

electron transport in cells of Shewanella putrefaciens MR-1 with normal levels of 

menaq ui none. J Bacteriol 182, 67-75. 
Myers, J. M. & Myers, C. R. (2001). Role of outer membrane cytochromes OmcA and OmcB 

of Shewanella putrefaciens MR-1 in reduction of manganese dioxide. Appl Environ 

Microdot 67, 260-269. 
Nealson, K. H. & Cox, B. L. (2002). Microbial metal-ion reduction and Mars: extraterrestrial 

expectations? CurrOpin Microdot 5, 296-300. 
Nelson, G., Wilmott, N. & Guthrie, J. (2000). Degradative Bacteria. UK patent 

GB2316684B. 
Nevin, K. P. & Lovley, D. R. (2002). Mechanisms for Fe(lll) oxide reduction in sedimentary 

environments. Geomicrodol J 19, 141-1 59. 
Newman, D. K. & Kolter, R. (2000). A role for excreted quinones in extracellular electron 

transfer. Nature 405, 94-97. 
Newman, D. K., Kennedy, E. K., Coates, J. D., Ahmann, D., Ellis, D. J., Lovley, D. R. & 

Morel, F. M. M. (1997). Dissimilatory arsenate and sulfate reduction in Desulfo- 

tomaculum auripigmentum sp. nov. Arch Microdot 168, 380-388. 
Nickson, R., McArthur, J., Burgess, W., Ahmed, K. M. f Ravenscroft, P. & Rahman, M. 

(1998). Arsenic poisoning of Bangladesh groundwater. Nature 395, 338. 
Oremland, R. S. & Capone, D. G. (1988). Use of "specific" inhibitors in biogeochemistry 

and microbial ecology. Adv Microb Ecol 10, 285-383. 
Oremland, R. S. & Stolz, J. F. (2000). Dissimilatory reduction of selenate and arsenate 

in nature. In Environmental Microbe-Metal Interactions, pp. 199-224. Edited by 

D. R. Lovley. Washington, DC: American Society for Microbiology. 
Oremland, R. S. & Stolz, J. F. (2003). The ecology of arsenic. Science 300, 939-944. 
Oremland, R. S., Dowdle, P. R., Hoeft, S. & 7 other authors (2000). Bacterial dissimilatory 

reduction of arsenate and sulfate in meromictic Mono Lake, California. Geochim 

Cosmochim Acta 64, 3073-3084. 
Oremland, R. S., Stolz, J. F. & Hollibaugh, J. T. (2004). The microbial arsenic cycle in Mono 

Lake, California. FEMS Microbiol Ecol 48, 1 5-27. 
Ortiz-Bernad, I., Anderson, R. T., Vrionis, H. A. & Lovley, D. R. (2004a). Vanadium 

respiration by Geobacter metallireducens: novel strategy for in situ removal of 

vanadium from groundwater. Appl Environ Microbiol 70, 3091-3095. 
Ortiz-Bernad, I., Anderson, R. T., Vrionis, H. A. & Lovley, D. R. (2004b). Resistance of 

solid-phase U(VI) to microbial reduction during in situ bioremediation of uranium- 
contaminated groundwater. Appl Environ Microbiol 70, 7558-7560. 
Park, C. H., Keyhan, M., Wielinga, B., Fendorf, S. & Matin, A. (2000). Purification to 

homogeneity and characterization of a novel Pseudomonas putida chromate 

reductase. Appl Environ Microbiol 66, 1 788-1 795. 



SGM symposium 65 



300 J. R. Lloyd 



Payne, R. B., Gentry, D. M., Rapp-Giles, B. J., Casalot, L. & Wall, J. D. (2002). Uranium 
reduction by Desulfovibrio desulfuricans strain G20 and a cytochrome c 3 mutant. 
Appl Environ Microbiol 68, 3129-3132. 

Peretrukhln, V. R, Khizhnyak, N. N., Lyalikova, N. N. & German, K. E. (1996). 
Biosorption of technetium-99 and some actinides by bottom sediments of Lake 
Belsso Kosino of the Moscow region. Radiochemistry 38, 440-443. 

Reguera, G., McCarthy, K. D., Mehta, T., Nlcoll, J. S., Tuominen, M. T. & Lovley, D. R. 
(2005). Extracellular electron transfer via microbial nanowires. Nature 435, 
1098-1101. 

Renshaw, J. C, Butchins, L. J. C, Livens, F. R., May, I., Charnock, J. M. & Lloyd, J. R. 
(2005). Bioreduction of uranium: environmental implications of a pentavalent 
intermediate. Environ Sci Technol (in press). 

Roberts, J. L. (1947). Reduction of ferric hydroxide by strains of Bacillus polymyxa. Soil Sci 
63, 135-140. 

Roling, W. F. M., van Breukelen, B. M., Braster, M. f Lin, B. & van Verseveld, H. W. 
(2001). Relationships between microbial community structure and hydrochemistry 
in a landfill leachate-polluted aquifer. Appl Environ Microbiol 67, 4619-4629. 

Rooney-Varga, J. N., Anderson, R. T., Fraga, J. L, Ringelberg, D. & Lovley, D. R. (1999). 
Microbial communities associated with anaerobic benzene degradation in a petro- 
leum-contaminated aquifer. Appl Environ Microbiol 65, 3056-3063. 

Ruggiero, C. E., Boukhalfa, H., Forsythe, J. H., Lack, J. G., Hersman, L. E. & Neu, M. P. 
(2005). Actinide and metal toxicity to prospective bioremediation bacteria. Environ 
Microbiol 7 ,88-97. 

Rusin, P. A., Quintana, L., Brainard, J. R., Strietelmeier, B. A., Tait, C. D., Ekberg, S. A., 
Palmer, P. D., Newton, T. W. & Clark, D. L. (1994). Solubilization of plutonium 
hydrous oxide by iron-reducing bacteria. Environ Sci Technol 28, 1 686-1 690. 

Santini, J. M., Sly, L. I., Schnagl, R. D. & Macy, J. M. (2000). A new chemolithoautotrophic 
arsenite-oxidizing bacterium isolated from a gold mine: phylogenetic, physiological, 
and preliminary biochemical studies. Appl Environ Microbiol 66, 92-97 '. 

Schroder, I., Rech, S., Krafft, T. & Macy, J. M. (1997). Purification and characteri- 
zation of the selenate reductase from Thauera selenatis. J Biol Chem 272, 23765- 
23768. 

Seeliger, S., Cord-Ruwisch, R. & Schink, B. (1998). A periplasmic and extracellular 
c-type cytochrome of Geobacter sulfurreducens acts as a ferric iron reductase and 
as an electron carrier to other acceptors or to partner bacteria. J Bacteriol 180, 3686- 
3691. 

Shen, H. & Wang, Y.-T. (1993). Characterization of enzymatic reduction of hexavalent 
chromium by Escherichia coli ATCC 33456. Appl Environ Microbiol 59, 3771-3777. 

Silver, S. (2003). Silver resistance. FE MS Microbiol Rev 27, 341-353. 

Smedley, P. L. & Kinniburgh, D. G. (2002). A review of the source, behaviour and 
distribution of arsenic in natural waters. Appl Geochem 17, 517-568. 

Smillie, R. H., Hunter, K. & Loutit, M. (1981). Reduction of chromium(VI) by bacterially 
produced hydrogen sulphide in a marine environment. Water Res 15, 1351-1354. 

Snoeyen bos- West, O. L, Nevin, K. P., Anderson, R. T. & Lovley, D. R. (2000). Enrichment 
of Geobacter species in response to stimulation of Fe(lll) reduction in sandy aquifer 
sediments. Microb Ecol 39, 1 53-1 67. 

Stein, L. Y, La Due, M. T., Grundl, T. J. & Nealson, K. H. (2001). Bacterial and archaeal 
populations associated with freshwater ferromanganous micronodules and sedi- 
ments. Environ Microbiol 3, 10-18. 



SGM symposium 65 



Microbial metal reduction 301 



Steinberg, N. A. & Oremland, R. S. (1990). Dissimilatory selenate reduction potentials in a 

diversity of sediment types. Appl Environ Microbiol 56, 3550-3557. 
Stolz, J. R, Gugliuzza, T., Switzer Blum, J., Oremland, R. S. & Murillo, F. M. (1997). 

Differential cytochrome content and reductase activity in Geospirillum barnesii strain 

SeS3. Arch Microbiol 1 67, 1 -5. 
Suzuki, T., Miyata, N., Horitsu, H., Kawai, K., Takamizawa, K., Tai, Y. & Okazaki, M. 

(1992). NAD(P)H-dependent chromium (VI) reductase of Pseudomonas ambigua 

G-1 : a Cr(V) intermediate is formed during the reduction of Cr(VI) to Cr(lll). J Bacteriol 

174,5340-5345. 
Switzer Blum, J., Burns Bindi, A., Buzzelli, J., Stolz, J. F. & Oremland, R. S. (1998). 

Bacillus arsenicoselenatis, sp. nov., and Bacillus selenitireducens sp. nov.: two halo- 

alkaliphiles from Mono Lake, California, that respire oxyanions of selenium and 

arsenic. Arch Microbiol 171,1 9-30. 
Taylor, D. E. (1999). Bacterial tellurite resistance. Trends Microbiol 7, 111-11 5. 
Tebo, B. M. & Obraztsova, A. Ya. (1998). Sulfate-reducing bacterium grows with 

Cr(VI), U(VI), Mn(IV), and Fe(lll) as electron acceptors. FE MS Microbiol Lett 162, 193- 

198. 
Thamdrup, B. (2000). Bacterial manganese and iron reduction in aquatic sediments. Adv 

MicrobEcol 16, 41-84. 
Tomei, F. A., Barton, L. L, Lemanski, C. L. & Zocco, T. G. (1992). Reduction of selenate 

and selenite to elemental selenium by Wolinella succinogenes. Can J Microbiol 38, 

1328-1333. 
Tomei, F. A., Barton, L. L, Lemanski, C. L, Zocco, T. G., Fink, N. H. & Sillerud, L. O. 

(1995). Transformation of selenate and selenite to elemental selenium by Desulfo- 

vibrio desulfuricans. J Ind Microbiol 1 4, 329-336. 
Tucker, M. D., Barton, L. L. & Thomson, B. M. (1997). Reduction and immobilization of 

molybdenum by Desulfovibrio desulfuricans. J Environ Qual 26, 1 146-1 1 52. 
Tucker, M. D., Barton, L. L. & Thomson, B. M. (1 998). Reduction of Cr, Mo, Se and U 

by Desulfovibrio desulfuricans immobilized in polyacrylamide gels. J Ind Microbiol 

Biotechnol 20, 13-19. 
Vargas, M., Kashefi, K., Blunt-Harris, E. L. & Lovley, D. R. (1998). Microbiological 

evidence for Fe(lll) reduction on early Earth. Nature 395, 65-67. 
Wade, R., Jr & DiChristina, T. J. (2000). Isolation of U(VI) reduction-deficient mutants of 

Shewanella putrefaciens. FEM5 Microbiol Lett 1 84, 1 43-1 48. 
Wang, Y.-T. (2000). Microbial reduction of chromate. In Environmental Microbe-Metal 

Interactions, pp. 225-235. Edited by D. R. Lovley. Washington, DC: American Society 

for Microbiology. 
Wang, P.-C, Mori, T, Komori, K., Sasatsu, M., Toda, K. & Ohtake, H. (1989). Isolation 

and characterization of an Enterobacter cloacae strain that reduces hexavalent 

chromium under anaerobic conditions. Appl Environ Microbiol 55, 1 665-1 669. 
Wildung, R. E., McFadden, K. M. & Garland, T. R. (1979). Technetium sources and 

behaviour in the environment. J Environ Qual 8, 1 56-1 61 . 
Wildung, R. E., Li, S. W., Murray, C. J., Krupka, K. M., Xie, Y, Hess, N. J. & Roden, E. E. 

(2004). Technetium reduction in sediments of a shallow aquifer exhibiting 

dissimilatory iron reduction potential. FEMS Microbiol Ecol 49, 1 51-162. 
Woolfolk, C. A. & Whiteley, H. R. (1962). Reduction of inorganic compounds with 

molecular hydrogen by Micrococcus lactilyticus. I. Stoichiometry with compounds 

of arsenic, selenium, tellurium, transition and other elements. J Bacteriol 84, 647- 

658. 



SGM symposium 65 



302 J. R. Lloyd 



Yong, P., Farr, J. P. G., Harris, I. R. & Macaskie, L E. (2002a). Palladium recovery by 
immobilized cells of Desulfovibrio desulfuricans using hydrogen as the electron 
donor in a novel electrobioreactor. Biotechnol Lett 24, 205-212. 

Yong, P., Rowson, N. A., Farr, J. P. G., Harris, I. R. & Macaskie, L. E. (2002b). 

Bioaccumulation of palladium by Desulfovibrio desulfuricans. J Chem Techno! 
Biotechnol 77, 593-601. 
Yurkova, N. A. & Lyalikova, N. N. (1991). New vanadate-reducing facultative 
chemolithotrophic bacteria. Mikrobiologiia 59, 672-677 (in Russian). 



SGM symposium 65 



New insights into the physiology 
and regulation of the anaerobic 
oxidation of methane 

Martin Kruger 1 - 2 and Tina Treude 2 

1 Federal Institute for Geosciences and Resources (BGR), Stilleweg 2, D-30655 Hannover, Germany 
2 Max-Planck-lnstitute for Marine Microbiology, Celsiusstrasse 1, D-28359 Bremen, Germany 



INTRODUCTION 

Methane is an important link within the global carbon cycle and has become a major 
focus for scientific investigations over the last decades, especially since the discovery of 
large deposits of methane hydrates in continental margins. The majority of recent 
methane production is biogenic, i.e. produced either by thermogenic transformation of 
organic material or by methanogenesis as the final step in fermentation of organic 
matter carried out by methanogenic archaea in anoxic habitats (Reeburgh, 1996). There 
are also abiotic sources of methane, e.g. at mid-oceanic ridges, where serpentinization 
takes place. In marine environments, the bulk of the methane is produced in shelf and 
upper continental-margin sediments, which receive large amounts of organic matter 
from deposition (Reeburgh, 1996). As methane builds up, it migrates upwards and may 
reach the sediment surface. Here, its ebullition and oxidation can lead to the formation 
of complex geostructures, such as pockmarks or carbonate chimneys and platforms, as 
well as large-scale topographies, such as mud volcanoes and carbonate mounds (Ivanov 
et al., 1991; Milkov, 2000). In most of the deeper continental margin and the abyssal 
plain sediments, methane production is low, as only 1-5 % of the surface primary 
production reaches the bathyal and abyssal seabed, due to degradation processes in the 
water column (Gage & Tyler, 1996). 

Despite the high rates of methane production in shallow marine regions, the contri- 
bution of the oceans, with around 3-5 % to the global methane emission into the 
atmosphere, is extremely low compared with major methane sources, such as wetlands, 
rice fields or ruminants (IPCC, 1994; Reeburgh, 1996). The reason for this is the 

SGM symposium 65: Micro-organisms and Earth systems - advances in geomicrobiology. 
Editors G. M. Gadd, K. T. Semple & H. M. Lappin-Scott. Cambridge University Press. ISBN 521 86222 1 ©SGM 2005 



304 M. Kruger and T. Treude 



.2 



SOJ (tnmol m 3 ) 

o 10 20 ao 



Depth 
(cm) 






i 


I 










50 






^sol 




i 
100 






\CH 4 




150 














I 


i i 


11 



Sulfate-dependent 
AOM 







2 4 

CH 4 (mmoidm 3 ) 



Fig. 1. Scheme depicting typical profiles of sulfate and methane concentrations in porewater of an 
anoxic marine sediment, indicating a distinct zone of AOM activity. 



presence of sulfate in marine systems, which is an electron acceptor used for microbial 
sulfate reduction (Jorgensen & Fenchel, 1974; Jorgensen, 1982). As long as sulfate is 
present in the sediment, methanogenesis is shifted to deeper sediment layers due to 
substrate competition between sulfate-reducing bacteria and methanogens (Zehnder, 
1988). When migrating towards the sediment surface, methane is consumed by two 
microbial pathways: anaerobic oxidation of methane (AOM) and aerobic oxidation 
of methane. Reeburgh (1996) proposed that, by these processes, up to 80 % of the 
methane produced in the marine sediments is oxidized prior to its release into 
the hydrosphere. 

Unlike freshwater and terrestrial habitats, aerobic oxidation of methane is less import- 
ant in marine ecosystems, as oxygen availability in the sediments is low compared to 
sulfate, the electron acceptor for AOM (D'Hondt et al., 2002). However, information 
about rates and micro-organisms involved in aerobic oxidation of methane in the ocean 
is scarce. So far, only a few studies have been published (e.g. Lidstrom, 1988; Holmes et 
al., 1996; Valentine et al., 2001; Kruger et aL, 2005), of which the majority deal mainly 
with pelagic processes. 

In marine sediments, the bulk of upward-migrating methane is consumed during AOM 
using sulfate instead of oxygen as electron acceptor (equation 1; Zehnder & Brock, 
1980; Hoehler et al., 1994; Reeburgh, 1996; Valentine & Reeburgh, 2000; Hinrichs & 



SGM symposium 65 



Physiology and regulation of AOM 305 



Table 1. Characteristics of habitats investigated for AOM (see also Fig. 1). 



Characteristic 


Hydrate Ridge 


Eckernforde 


Bay 


Chilean 
continental margin 


Type of seep 


Gas seep/gas 
hydrates 


Gassy coastal 

sediment 




Diffusive system 


Methane transport 


Advective 


Diffusive/advective 


Diffusive 


Sulfate penetration depth (cm) 


10 


30 




150-350 


Sediment depth at which 
methane is completely 
consumed (cm) 


Methane reaches 
hydrosphere 


0-5 




110-360 


Thickness of AOM zone (cm) 


0-10 


20-25 




4-40 


Measured areal AOM rate 
(mmol m -2 day -1 ) 


56-100 


0-4-1-5* 




0-2-5-7 


Calculated methane flux 
(mmol rrr 2 day -1 ) 


56-200t 


0-6-1-3* 




0-07-0-13 


Retention of methane in 
sediment (%) 


50-100 


1 00 (except gas) 
bubbles 


100 



* Rates determined by Treude etal. (2005b). 

tCalculated from rate measurements (Treude etal., 2003) and methane effluxes (Torres etal., 2002). 

t- Estimated from the methane profiles of Abegg & Anderson (1 997). 

Boetius, 2002). The high availability of sulfate enables methane consumption far below 
the sediment-water interface (Fig. 1; Table 1). 



CH 4 +S0 4 



2- 



HCOI+H.O + HS" 



(equation 1) 



It has been proposed that the only major escape route for methane into the hydrosphere 
or atmosphere is ebullition of free gas or gas hydrates floating from the seabed. 
Wherever sulfate is available, AOM communities control the release of dissolved 
methane from the sediment (Treude, 2003). Without this mechanism, the contribution 
of the world's oceans to global methane emission would approximately equal the 
amount of methane emanating from ruminants, one of the biggest sources of today's 
methane emission into the atmosphere (Reeburgh, 1996). 

After a historical introduction, this review focuses on recent findings concerning 
the micro-organisms involved in AOM and its environmental regulation and new 
information concerning the mechanism of this still enigmatic process. 

HISTORY OF AOM 

The first published investigation of AOM was by Martens & Berner (1974), who 
described conspicuous methane and sulfate profiles in organic-rich sediments (Fig. 1), 



SGM symposium 65 



306 M. Kruger and T. Treude 

in which methane did not accumulate before sulfate depletion from porewater. From 
the decrease of methane concentrations in the sulfate-reducing zone, they concluded 
that methane must be consumed with sulfate instead of oxygen as terminal electron 
acceptor. 

Since then, more biogeochemical evidence has been published confirming that the 
process of methane consumption in marine sediments takes place at the base of 
the sulfate zone, linked directly or indirectly to the activity of sulfate-reducing bacteria 
(Reeburgh, 1980; Iversen & Jorgensen, 1985; Hoehler et al., 1994; Hansen et al., 1998; 
Niewohner et al., 1998; Borowski et al., 2000). These findings were based on depth 
profiles of methane and sulfate/sulfide (Fig. 1), 13 C : 12 C ratios in carbon dioxide and 
methane in sediment profiles, and labelling studies with sediment samples (e. g. Iversen 
& Blackburn, 1981; Alperin & Reeburgh, 1985; Iversen &C Jorgensen, 1985; Hoehler 
et al., 1994). However, Zehnder &C Brock (1979, 1980) were the first to demonstrate 
methane oxidation under anoxic conditions by methanogenic archaea and hypo- 
thesized a coupled two-step mechanism of AOM. They postulated that methane is first 
activated by methanogenic archaea working in reverse, leading to the formation of 
intermediates, e.g. acetate or methanol. In a second step, the intermediates are oxidized 
to C0 2 under concurrent sulfate reduction by other non-methanogenic members of the 
microbial community 

Since these pioneering studies, knowledge of AOM has increased substantially, 
involving biogeochemical, microbiological and molecular methods. Radiotracer 
measurements enabled the first direct quantification of AOM and concurrent sulfate 
reduction in anoxic marine sediments (Fig. 2; Reeburgh, 1976; Iversen & Blackburn, 
1981; Devol, 1983). Iversen & Blackburn (1981) measured a 1 : 1 ratio of AOM and 
sulfate reduction in the sulfate-methane transition zone of Danish sediments, 
demonstrating the close coupling between these processes. 

Hoehler et al. (1994) confirmed by thermodynamic modelling that a consortium of 
methanogenic archaea and sulfate-reducing bacteria could gain energy from AOM 
(Fig. 3). In situ and inhibitor studies (Hoehler et al., 1994), as well as laboratory 
experiments with growing methanogens converting [ 14 C]methane to 14 C0 9 during 
methanogenesis (Harder, 1997; Zehnder & Brock, 1979, 1980), further stimulated 
the discussion about AOM being a reverse process of methanogenesis. 

MICRO-ORGANISMS INVOLVED IN AOM 

It has only been during the last 5-10 years that the identification of the micro- 
organisms responsible for AOM has been possible. It was advanced by investigations of 
lipid biomarkers in sediments from methane seeps. In the search for these organisms, 

SGM symposium 65 



Physiology and regulation of AOM 307 



(TJ 



E 
u 

O 

E 



8 



B 

03 



10 000 



1 000 



100 



10 



Methane seeps 



Gassy 
sediment 



Diffusive system 





Hydrate Ridge 



Black Sea 
reef 



Eckernf6rde 

Bay 



Chile 800 m Chile 1160 m 



Fig. 2. Rates of AOM at methane hot spots in selected habitats differing in methane availability and 
flux rates (mean ±sem, n = 3-1 8). Data are from Treude eta/. (2003) (Hydrate Ridge), Michaelisefa/. 
(2002) (Black Sea), Treude era/. (2005b) (Eckernforde Bay) andTreudeef al. (2005a) (Chile). 




CH 4 



o » 



t 



o n - methanotroph' 

o 



Sulfate reducer 



Gas hydrates/seeps 



Fig. 3. Interactions between methanotrophic archaea and sulfate-reducing bacteria according to 
the hypothesis of Hoehleref al. (1994). Scheme by K. Nauhaus. 

scientists found methanogen-associated lipids, named crocetane, archaeol and hydroxy- 
archaeol, in active methane seeps, revealing extremely light d u C values down to 
-110 %o, giving evidence for an involvement of archaea in methane consumption (Elvert 
et d. y 1999; Hinrichs et al, 1999; Pancost et al., 2000; Thiel et al., 2001; Schouten et al., 
2003). These archaeal lipids were also found in association with isotopically light 
bacterial lipids, commonly found in sulfate-reducing bacteria (Hinrichs et al., 2000; 
Hinrichs & Boetius, 2002; Elvert et al., 2003). Similar to archaeal lipids, this relative 
enrichment in 12 C indicated the incorporation of methane-derived carbon into 
bacterial cells. 



SGM symposium 65 



308 M. Kruger and T. Treude 

Boetius et al. (2000) presented the first microscopic pictures of an AOM consortium 
visualized by fluorescence in situ hybridization (FISH), showing aggregates of archaeal 
cells surrounded by a shell of sulfate-reducing bacteria. The aggregates grow to a size 
of about 6-10 ^m before they break apart into subaggregates, implicating the need to 
keep short distances between cells during substrate exchange. These consortia were 
discovered in surface sediment overlying methane hydrates at Hydrate Ridge, where 
they represented >90 % of the microbial biomass. 

After revealing the conspicuous morphology of the AOM consortium, further methods 
were used to obtain direct evidence for methanotrophy of the AOM consortium. A 
combination of FISH and secondary-ion mass spectrometry allowed the measurement 
of (5 13 C profiles of the biomass of single aggregates (Orphan et al., 2001a). A high 
depletion in 13 C, with values down to -96 and -62 %o, was detected in archaeal and 
bacterial cells, respectively. These results confirmed the assimilation of isotopically 
light methane by the consortium. 

Molecular studies showed that the anaerobic methanotrophs (ANME) were affiliated 
most closely with methanogenic archaea of the order Methanosarcinales (Hinrichs 
et al., 1999; Orphan et al., 2001b) and have frequently been associated with sulfate- 
reducing bacteria of the genera Desulfosarcina and Desulfococcus (Boetius et al., 2000; 
Michaelis et al., 2002; Knittel et al., 2003). Today, three major groups of methane- 
oxidizing archaea have been identified: ANME-1, ANME-2 and ANME-3 (Hinrichs 
et al, 1999; Boetius et al., 2000; Orphan et al., 2001b; Knittel et al., 2005) . ANME-2 and 
ANME-3 belong to the order Methanosarcinales. The ANME-2 group has recently 
been divided into three subgroups, ANME-2a to -2c (Knittel et al., 2005), which seem to 
exhibit differences in environmental preferences and the structure of their aggregation 
with sulfate reducers. ANME-1 is distinct from, but related to, methanogenic archaea 
of the orders Metbanomicrobiales and Methanosarcinales. 

The bacterial diversity at methane seeps is high, especially within the d-proteo bacteria 
(Knittel et al., 2003), including members of the genera Desulfosarcina, Desulfo- 
rhopalus, Desulfocapsa and Desulfobulbus. Comprehensive overviews on the diversity 
and phylogeny of archaea and sulfate reducers involved in AOM and associated with 
methane seeps have been published recently by Knittel et al. (2003, 2005), respectively 

HOT SPOTS FOR THE STUDY OF AOM 

In general, AOM can be expected wherever methane and sulfate coexist in anoxic 
environments. One main factor determining the magnitude of AOM is the methane 
supply, because methane-turnover rates were found to increase with methane concen- 
tration and methane flux (Nauhaus et al., 2002; Treude, 2003). Hot spots for AOM have 

SGM symposium 65 



Physiology and regulation of AOM 309 

been found in diverse habitats, characterized by a wide range of environmental 
characteristics. Hinrichs & Boetius (2002), Treude (2003) and Kriiger et al. (2005) have 
reviewed AOM rates in marine sediments of different water depths, as well as methane 
seeps. These first compilations of AOM field measurements and modelling have 
suggested a direct coupling between methane supply and methane consumption in the 
habitat. At methane seeps of ancient reservoirs or gas-hydrate locations, AOM rates 
were found to be 10-100 times higher than those in non-seep regions. However, the 
data of environmental AOM rates are still fragmentary. Below, two methane-rich 
marine environments, which have so far played a major role during the investigation 
of AOM, are briefly introduced: they are located at Hydrate Ridge, off the coast of 
Oregon, USA (Boetius & Suess, 2004 and references therein) and on the north-western 
shelf of the Black Sea (Michaelis et al., 2002). 

Hydrate Ridge 

At Hydrate Ridge, gas-hydrate deposits are located a few centimetres below the sedi- 
ment surface in a water depth of 600-800 m, corresponding to the hydrate-stability 
zone (Suess et al., 1999). These layers lead to very high methane fluxes (up to 200 mmol 
m~ 2 day _1 ; Table 1), which fuel a diverse, seep-associated community (Sahling et al., 
2002; Treude et al., 2003), including zones covered by thick mats of sulfide-oxidizing 
members of the genus Beggiatoa or inhabited by different dwelling clams, such as Cal- 
yptogena and Acharax species. In these surface sediments, aggregates of methano- 
trophic ANME-2 and sulfate-reducing DesulfococcuslDesulfosarcina cells dominate 
the microbial biomass (Boetius et al., 2000). Their maximum abundance is located 
within the upper 10 cm below the sea floor, where methane and sulfate meet (Treude 
et al., 2003). In this zone, some of the highest densities of ANME cells and methane- 
turnover rates known from marine environments have been found (Boetius et al., 2000; 
Boetius & Suess, 2004). Elevated HCOT concentrations caused by this high AOM 
activity result in an increase in alkalinity and support carbonate precipitation, forming 
large carbonate landscapes at Hydrate Ridge. 

Black Sea 

In the north-western Black Sea, hundreds of active gas seeps occur along the shelf edge 
west of the Crimea peninsula, at water depths between 35 and 800 m (Ivanov et al., 
1991). Within the anoxic zone, massive carbonate accumulations up to 4 m high and 
1 m in diameter have been found associated with these seeps (Pimenov et al., 1997; Thiel 
et al., 2001; Lein et al., 2002; Michaelis et al., 2002; Blumenberg et al., 2004). These 
build-ups are covered by up to 10 cm thick microbial methanotrophic mats. From holes 
in these structures, streams of gas bubbles emanate into the water column. Strong 13 C 
depletions indicate an incorporation of methane carbon into carbonates, bulk micro- 
bial biomass and specific lipids. The main matrix of the microbial mats consists of 

SGM symposium 65 



310 M. Kruger and T. Treude 

densely aggregated cells of methanotrophic ANME-1 and sulfate-reducing Desulfo- 
coccuslDesulfosarcina, but many other bacteria of unknown diversity and function co- 
occur in the mats (Michaelis et aL, 2002; Blumenberg et aL, 2004; Knittel et aL, 2005). 
The physiology of the methanotrophic mats has been studied in greater detail (Pimenov 
et aL, 1997; Michaelis et aL, 2002; Treude, 2003; Nauhaus et aL, 2005), as described 
below. 

ENVIRONMENTAL REGULATION OF AOM 

The most important questions regarding the functioning of AOM in the ocean concern 
its regulation and the growth and environmental adaptation of the communities 
mediating AOM. According to thermodynamic calculations, whether free energy is 
available from AOM depends on the environmental settings (Zehnder 6c Brock, 1980; 
Iversen & Blackburn, 1981). So far, only limited experimental data are available to 
investigate the effect of variable environmental factors on the efficiency of AOM 
(Valentine & Reeburgh, 2000; Nauhaus et aL, 2002, 2005). For example, the balance 
between the concentrations of sulfate and sulfide might be an important factor 
regulating microbial methane consumption (Treude, 2003; Treude et aL, 2003). 

So far, no methanotrophic archaea are available for cultivation, probably because of 
their extremely slow growth rate (Girguis et aL, 2003). Hence, the investigation of their 
physiological capabilities and adaptations is only possible by in vitro studies with 
naturally enriched samples from the environment. At many sites studied extensively for 
AOM, including Hydrate Ridge, the Gulf of Mexico and the Black Sea, ANME-1 and 
ANME-2 have been found to co-occur. However, the dominance of either group varies; 
for example, in Hydrate Ridge samples, ANME-2 populations dominated the com- 
munity (Boetius et aL, 2000) whereas, in the Black Sea mats, ANME-1 far outnumbered 
ANME-2 (Knittel et aL, 2005). These differences in the community composition might 
be due to differences in environmental parameters, such as temperature or the avail- 
ability of methane and sulfate. 

The stoichiometry of AOM, which has been estimated from porewater studies and 
rate measurements in the field (Iversen & Blackburn, 1981), has been confirmed by in 
vitro studies with environmental samples. Simultaneous measurements of methane 
oxidation and sulfide production in samples from Hydrate Ridge and the Black Sea 
have revealed a molar ratio of 1 : 1 between the two processes (Nauhaus et aL, 2002, 
2005). Interestingly, the comparison of methane-driven sulfate reduction per cell 
revealed that ANME-2 communities (Hydrate Ridge) were up to 20 times more active 
than the ANME-1 communities in the microbial mats from the Black Sea. However, it is 
not known whether all methanotrophic cells within a cell aggregate or within a mat are 
equally active, as cells with no contact to the bacterial partner might be inactive if not 

SGM symposium 65 



Physiology and regulation of AOM 31 1 

situated within an optimal substrate-concentration range required for sufficient energy 
conservation (Sorensen et al., 2001). Therefore, it remains an interesting question for 
future research, ideally with pure cultures, whether this difference in cell-specific 
activity between ANME-1 and -2 can be attributed to specific substrate kinetics or 
enzymic mechanisms of the ANME groups. 

It can be assumed that the efficiency of AOM in mitigating methane emissions is 
influenced by environmental parameters, such as pH, temperature and methane and 
sulfate fluxes (Joye et al., 2004). Methane availability in situ depends on the methane 
flux from subsurface reservoirs, as well as methane solubility, which is in turn influenced 
by hydrostatic pressure and temperature (Yamamoto et al., 1976). Consequently, it 
is essential to investigate the response of the AOM organisms to changes in these 
parameters, to be able to estimate the effects of environmental or climatic changes 
on AOM efficiency. An increase of methane partial pressure from 0-1 to 14 MPa 
resulted in a fivefold increase of AOM rates in ANME-2-dominated samples (Nauhaus 
et al., 2002) and a twofold increase in ANME-1-dominated samples (Nauhaus et al., 
2005). A similar stimulation was also detected in samples from shallow water depths 
(Kriiger et al., 2005), which generally do not encounter such high concentrations of 
methane. It is remarkable that it was also possible to induce AOM in formerly inactive 
sediments by increasing the methane availability (Girguis et al., 2003; Kriiger et al., 
2005). 

The free gas ebullition observed at seeps such as the microbial reefs in the Black Sea and 
at Hydrate Ridge indicates methane saturation, with theoretical values of 2-3 MPa 
(40 mM) and 8 MPa (140 mM) for the Black Sea and Hydrate Ridge, respectively. 
Consequently, the rates observed in vitro at only 14 MPa must still represent sub- 
stantial underestimations of rates occurring under in situ conditions. This inability in 
reaching environmental methane concentrations might also inhibit attempts to culture 
these organisms in the laboratory (see below). 

The pH of the environment might also influence activities, as well as the composition 
of microbial communities. For example, Nauhaus et al. (2005) showed that ANME-2 
from Hydrate Ridge sediments had a distinct maximum of AOM rates at pH 7-4 
(pH 7-7-5), whilst the pH optimum was broader for the ANME-1 community from the 
Black Sea, ranging from pH 6-8 to 84. These small differences in environmental 
preferences might contribute to the development of either ANME-1- or -2-dominated 
communities. 

Besides the pH, temperature is another important environmental factor influencing 
micro-organisms. Despite a difference of only 4°C between in situ temperatures of 

SGM symposium 65 



312 M. Kruger and T. Treude 

habitats investigated in the Black Sea and at Hydrate Ridge, the ANME-2 community 
of Hydrate Ridge was more active at low temperatures (8-12 °C), whereas the ANME-1 
community in the Black Sea was mesophilic, with highest AOM activities between 16 
and 24 °C (Nauhaus et al., 2005). These distinct temperature optima for AOM of 
environmental samples dominated by ANME-1 or ANME-2 may indicate a selective 
advantage for either population. Further temperature optima for AOM ranged from 
4°C for sediment from the Haakon Mosby Mud Volcano (-1-5 °C in situ) to 25 °C 
in the Baltic Sea (between 4 and 16 °C in situ) (Kruger et al., 2005). So far, the different 
temperature optima of AOM reflected the different in situ temperatures of the habitat. 
Especially at shallow sites, seasonal temperature changes might also cause changes in 
AOM activity. Indeed, such seasonality of AOM activity has been reported from studies 
in Eckernforde Bay, Baltic Sea (Treude et al., 2005b) and Cape Lookout Bight, USA 
(Hoehler *?**/., 1994). 

So far, reports on the discovery of AOM have been restricted to habitats with sufficient 
concentrations of sulfate available for the microbial partners of the methanotrophic 
archaea (Hinrichs & Boetius, 2002 and references therein), predominantly in marine 
habitats. However, the question is still pending whether, under specific environmental 
conditions, AOM might also proceed with other alternative electron acceptors. In a 
recent study with samples from Hydrate Ridge and the Black Sea, Nauhaus et al. (2005) 
observed no AOM activity without sulfate. Instead, both ANME-1 and -2 communities 
oxidized methane with similar rates at sulfate concentrations ranging from 10 to 
100 mM. Other electron acceptors for AOM, such as nitrate, sulfur, ferric iron and 
manganese oxide, were also reduced in Hydrate Ridge sediments by the indigenous 
microbial population. However, this reduction was not coupled to AOM. 

In summary, the ecological niches occupied more frequently by ANME-1 or ANME-2 
seem to be defined mainly by temperature and the simultaneous availability of methane 
and sulfate. Nevertheless, the factor(s) leading to the dominance of either group remain 
to be identified. 

MECHANISM OF AOM 

The invesigation of mechanistic details of AOM is still hindered by the lack of pure 
cultures of anaerobic methanothophs. Only a few studies are available, which have been 
conducted with environmental samples naturally enriched in methanotrophic biomass 
(Hoehler et al., 1994; Nauhaus et al., 2002, 2005). An important question regarding the 
mechanism of AOM is whether the two reactions involved, i.e. methane oxidation and 
sulfate reduction, are indeed carried out by a consortium of methanotrophic archaea 
and associated bacteria (Fig. 3), as indicated by the striking structural features revealed 
by microscopic analysis (Boetius et al., 2000; Orphan et al., 2001b; Michaelis et al., 

SGM symposium 65 



Physiology and regulation of AOM 313 

2002), or whether the entire process is mediated by a single organism. The latter is 
indicated by repeated findings of ANME (mainly ANME-1) without contact to sulfate- 
reducing bacteria (Orphan et al., 2001a; Michaelis et al., 2002; Joye et al., 2004; Treude 
etai, 2005b). 

Besides the microscopic evidence for a consortium of methanotrophic archaea and 
sulfate-reducing bacteria mediating AOM, further evidence was gained by inhibition 
experiments. In these experiments on environmental regulation of AOM, molybdate 
and bromoethanesulfonate (BES) as inhibitors for sulfate reduction and methano- 
genesis, respectively, have been used (Alperin 6c Reeburgh, 1985; Hoehler et al., 1994; 
Hansen et al., 1998; Nauhaus et al., 2005). In both ANME-1- and -2-dominated sam- 
ples, AOM was inhibited completely by BES. This inhibition was reversible and AOM 
activity was resumed after removal of BES. Molybdate inhibited AOM completely in 
ANME-2-dominated samples, but only partially in ANME-1-dominated samples 
(Nauhaus et al., 2005), which was explained by strong adsorption of molybdate to 
extracellular polysaccharide. These results are in good agreement with previous studies, 
in which a partial to complete inhibition of AOM by these compounds was observed 
(Alperin & Reeburgh, 1985; Hoehler et al., 1994). In conclusion, the application of a 
specific inhibitor, i.e. BES for methanogens and molybdate for sulfate-reducing bacteria, 
inhibited AOM. However, this inhibition is not absolute proof for syntrophy or a 
symbiotic association between ANME and sulfate-reducing bacteria. A complete inhi- 
bition would also be observed if methane oxidation and sulfate reduction were carried 
out by a single organism. 

Nevertheless, the inhibition of AOM by a methanogen-specific inhibitor provides 
important evidence for the hypothesis that AOM is taking place via a reversal of 
methanogenesis using similar enzymes (Zehnder 6c Brock, 1980; Hoehler et al., 1994). 
Recently, an abundant protein closely resembling methyl coenzyme M reductase, the 
terminal enzyme in methanogenesis (Thauer, 1998), was detected in ANME-1 cells 
from Black Sea samples (Kriiger et al., 2003). This protein represents a likely candidate 
for the initial step in AOM (Fig. 4). Furthermore, Hallam et al. (2003) found gene 
sequences of the same enzyme in sediments with AOM activity, which they assigned to 
ANME-2. In a subsequent paper, even the almost-complete methanogenic enzymic 
system was attributed to ANME, based on the analysis of a metagenomic library from 
an AOM site (Fig. 4) (Hallam et al., 2004). 

If, indeed, a syntrophic mechanism is necessary, the question arises as to what the exact 
mechanism of this process might be and which intermediates are exchanged within the 
consortium (Fig. 3). The nature of the cooperation between the archaeal and bacterial 
partners in AOM has not yet been elucidated (Sorensen et al., 2001; Nauhaus et al., 
2002). 

SGM symposium 65 



314 M. Kruger and T. Treude 



CO* 



formyl MF 

If 

formyl H 4 HPMT 

If 

methenyl-H 4 HMPT 

26 ">2e- 

methylene H 4 MPT 

methyl H 4 MPT 

It 

methyl Co M 



Formylmethanofuran dehydrogenase* 
(fwd/fmd) 

Formylmethanofuran H 4 MPT 
W-formy transferase* (ftr) 

Methenyl H 4 MPTcyclohydrolase* (mch) 

Methylene H 4 MPT dehydrogenase* 
(mtd) 

Methylene H 4 MPT reductase (mer) 
H 4 MPT-S-methy (transferase (mtr) 



I coenzyme M reductase* (mcr) 



CH ( 



Fig. 4. Activities (marked with an asterisk) and genes for methanogenic enzymes found in 
methanotrophic archaea (AN ME). Activities are after Kruger eta/. (2003) and genes are after Hallam 
eta/. (2004). 



The addition of exogenous electron donors in the form of methanogenic substrates or 
other C 1 -C 3 compounds, including acetate, formate, hydrogen and methanol, did not 
stimulate sulfate reduction in the absence of methane (Nauhaus et al., 2002, 2005). In 
theory, if the sulfate-reducing bacteria are adapted to the substrate supplied by the 
methane-oxidizing partner, they should respond immediately to the addition of 
potential AOM intermediates, with higher activity 

As an alternative to hydrogen or carbon compounds, the possibility of electron transfer 
between the archaeal and sulfate-reducing partners has been discussed for AOM 
consortia (Sorensen et al., 2001; Nauhaus et al., 2002). Such an electron transfer to an 
external acceptor has been reported previously for different anaerobic reactions 
(Seeliger et al., 1998; Schink 6c Stams, 2001). However, the addition of several com- 
pounds able to capture electrons - phenazines, AQDS (anthroquinone disulfonate) and 
humic acids - did not replace the function of the sulfate-reducing bacterium. Neither of 
the added compounds induced AOM (Nauhaus et al., 2005). One explanation for this 
might be that, besides their important role in membrane-bound electron transport, 
phenazines might also have toxic effects on micro-organisms (Ingram & Blackwood, 
1970; Abken et al., 1998). Even though Straub & Schink (2003) showed that AQDS 
may serve as an electron shuttle for iron-reducing bacteria, the redox conditions in 
the incubations might change upon the addition of AQDS or humic acids, perhaps 
suppressing AOM (Hernandez &C Newman, 2001). In summary, these experiments did 



SGM symposium 65 



Physiology and regulation of AOM 315 

not provide direct evidence for a methanogenic or sulfidogenic substrate or electrons 
as an intermediate in AOM. Consequently, despite the conspicuous aggregation of 
archaea and sulfate-reducing bacteria in the environment, it might be possible that 
AOM is carried out by one organism alone (see below). 

GROWTH OF AOM MICRO-ORGANISMS IN THE LABORATORY 

Pure cultures of anaerobic methanotrophic archaea have not yet been isolated. 
Consequently, physiological studies have only been possible on environmental samples 
naturally enriched in methanotrophic archaea and their sulfate-reducing partners 
(Hoehler et al., 1994; Blumenberg et ai, 2004, 2005; Nauhaus et al., 2002, 2005). 
However, sample availability and quality have limited the scope of these experiments. 
For major questions regarding the mechanism and regulation of AOM, it seems - 
despite substantial progress - inevitable to work with pure cultures. Nevertheless, 
recent studies on microbial communities with a limited diversity have shown that there 
is the possibility to work on these aspects of AOM by using metagenomic (Hallam et 
al., 2004) or biochemical (Kriiger et ai, 2003) approaches. 

Recently, Girguis et al. (2003) developed a novel continuous-flow system to study AOM, 
which simulates the in situ conditions and could support the growth of anaerobic, 
methanotrophic archaea. The major limitation of this system was that it only operated 
under 0-1 MPa methane pressure and did not resemble in situ conditions on the sea 
floor. Consequently, thermodynamic calculations showed that the Gibbs free-energy 
yield for AOM at 1 atm methane pressure was low. In contrast, high in situ methane 
concentrations and a close physical association between AOM partners can produce 
energy yields sufficient to support biosynthesis (Sorensen et al., 2001; Nauhaus et ai, 
2002). Therefore, it would be advantageous to use high-pressure systems, as described 
by Nauhaus et al. (2002), for future research on the growth of AOM micro-organisms, 
thus allowing the application of elevated methane partial pressures. 

The means by which methanotrophic archaea are capable of growing at atmospheric 
methane pressures are difficult to understand. High rates of AOM observed in the deep 
oceans require high dissolved-methane concentrations (Nauhaus et al., 2002), which 
can only be sustained at high pressures. Nevertheless, there is evidence for AOM at 
atmospheric methane pressure in different shallow-water ecosystems (Iversen & 
Jorgensen, 1985; Martens et al., 1999; Kriiger et al., 2005). 

CONCLUSIONS 

Microbially mediated AOM significantly influences biological and biogeochemical 
processes on local to global scales. The process reduces methane flux into the water 
column, stimulates subsurface microbial metabolism and also supports rich deep-sea 

SGM symposium 65 



316 M. Kruger and T. Treude 

chemosynthetic communities that derive energy from one of its by-products, hydrogen 
sulfide. However, advective systems such as methane seeps might still represent a 
significant source of methane emission from the ocean to the hydrosphere, as recent 
data have shown that AOM is only able to inhibit methane emission into the water 
column completely in diffusive systems (Table 1; Treude, 2003). 

The range of datasets available today (Orphan et al., 2001b; Hinrichs & Boetius, 2002; 
Teske et al., 2002; Knittel et al., 2003; Treude et al., 2003; Boetius 6c Suess, 2004; 
Kallmeyer & Boetius, 2004) suggests that the co-occurrence of methane and sulfate is 
the major environmental factor defining the ecological niche occupied by AOM 
communities. So far, no environment is known in which only one of the ANME groups 
occurs. The significant dominance of either group points to the presence of defined 
environmental niches within AOM zones, which have not been distinguished so far. 
Further experimental studies of ANME-enriched samples from different environments 
with a range of habitat characteristics are needed to answer the question of niche 
selection. 

Finally, it seems that, although a lot is known about AOM and the respective microbial 
consortia, there are still many questions to be answered. For example, the intermediate 
that is exchanged between archaea and sulfate-reducing bacteria is still unknown. 
Thermodynamic considerations show that the archaea alone might not gain free energy 
by this process, thus depending on cooperation with the sulfate-reducing bacteria. 
Despite recent progress in the fields of biochemistry and metagenomics, this and other 
questions will probably not be answered until pure cultures become available. 

ACKNOWLEDGEMENTS 

We especially thank Dr Gundula Eller for critical reading of the manuscript. This is publication 
no. GEOTECH-129 of the programme GEOTECHNOLOGIEN of the Bundesministerium fiir 
Bildung und Forschung and Deutsche Forschungsgemeinsch aft . Within this programme, the 
study is part of projects MUMM-1, -2 and GHOSTDABS. Further support came from the Max 
Blanch. Society. 

REFERENCES 

Abegg, F. & Anderson, A. L. (1997). The acoustic turbid layer in muddy sediments of 
Eckernfoerde Bay, Western Baltic: methane concentration, saturation and bubble 
characteristics. MarGeol 137, 137-147. 

Abken, H.-J., Tietze, M., Brodersen, J., Baumer, S. ; Beifuss, U. & Deppenmeier, U. 
(1998). Isolation and characterization of methanophenazine and function of 
phenazines in membrane-bound electron transport of Methanosarcina mazei G61 . 
JBacteriol 180, 2027-2032. 

Alperin, M. J. & Reeburgh, W. S. (1985). Inhibition experiments on anaerobic methane 
oxidation. Appl Environ Microbiol 50, 940-945. 

SGM symposium 65 



Physiology and regulation of AOM 317 



Blumenberg, M., Seifert, R., Reitner, J., Pape, T. & Michaelis, W. (2004). Membrane lipid 
patterns typify distinct anaerobic methanotrophic consortia. Proc Natl Acad Sci U 5 A 
101, 11111-11116. 

Blumenberg, M., Seifert, R., Nauhaus, K., Widdel, R, Pape, T. & Michaelis, W. (2005). 
In vitro study of lipid biosynthesis in an anaerobically methane oxidizing microbial 
mat. Appl Environ Microbiol (in press). 

Boetius, A. & Suess, E. (2004). Hydrate Ridge: a natural laboratory for the study of micro- 
bial life fueled by methane from near-surface gas hydrates. Chem Geo/ 205, 291-31 0. 

Boetius, A., Ravenschlag, K., Schubert, C. J. & 7 other authors (2000). A marine micro- 
bial consortium apparently mediating anaerobic oxidation of methane. Nature 407, 
623-626. 

Borowski, W. S., Hoehler, T. M., Alperin, M. J., Rodriguez, N. M. & Paull, C. K. (2000). 
Significance of anaerobic methane oxidation in methane-rich sediments overlying 
the Blake Ridge gas hydrates. In Proceedings of the Ocean Drilling Program, Scientific 
Results, vol. 164, pp. 87-99. Edited by C. K. Paull, R. Matsumoto, P. J. Wallace & 
W. P. Dillon. College Station, TX: Texas A&M University. 

Devol, A. H. (1983). Methane oxidation rates in the anaerobic sediments of Saanich Inlet. 
Limnol Oceanogr 28, 738-742. 

D'Hondt, S., Rutherford, S. & Spivack, A. J. (2002). Metabolic activity of subsurface life in 
deep-sea sediments. Science 295, 2067-2070. 

Elvert, M., Suess, E. & Whiticar, M. J. (1999). Anaerobic methane oxidation associated 
with marine gas hydrates: superlight C-isotopes from saturated and unsaturated C 20 
and C 25 irregular isoprenoids. Naturwissenschaften 86, 295-300. 

Elvert, M., Boetius, A., Knittel, K. & Jorgensen, B. B. (2003). Characterization of specific 
membrane fatty acids as chemotaxonomic markers for sulfate-reducing bacteria 
involved in anaerobic oxidation of methane. Geomicrobiol J 20, 403^-19. 

Gage, J. D. & Tyler, P. A. (1 996). Deep-Sea Biology: a Natural History of Organisms at the 
Deep-Sea Floor. Cambridge: Cambridge University Press. 

Girguis, P. R., Orphan, V. J., Hallam, S. J. & DeLong, E. F. (2003). Growth and methane 
oxidation rates of anaerobic methanotrophic archaea in a continuous-flow bio- 
reactor. Appl Environ Microbiol 69, 5472-5482. 

Hallam, S. J., Girguis, P. R., Preston, C. M., Richardson, P. M. & DeLong, E. F. (2003). 
Identification of methyl coenzyme M reductase A (mcrA) genes associated with 
methane-oxidizing archaea. Appl Environ Microbiol 69, 5483-5491 . 

Hallam, S. J., Putnam, N., Preston, C. M., Detter, J. C, Rokhsar, D., Richardson, P. M. & 
DeLong, E. F. (2004). Reverse methanogenesis: testing the hypothesis with environ- 
mental genomics. Science 305, 1457-1462. 

Hansen, L. B., Finster, K., Fossing, H. & Iversen, N. (1998). Anaerobic methane oxidation 
in sulfate depleted sediments: effects of sulfate and molybdate additions. Aquat 
Microb Ecol 14, 195-204. 

Harder, J. (1997). Anaerobic methane oxidation by bacteria employing 14 C-methane 
uncontaminated with 14 C -carbon monoxide. MarGeol 137, 13-23. 

Hernandez, M. E. & Newman, D. K. (2001). Extracellular electron transfer. Cell Mol Life 
Sci 58, 1562-1571. 

Hinrichs, K.-U. & Boetius, A. (2002). The anaerobic oxidation of methane: new insights in 
microbial ecology and biogeochemistry. In Ocean Margin Systems, pp. 457-477. 
Edited by G. Wefer, D. Billett, D. Hebbeln, B.B. J0rgensen, M. Schluter & T van 
Weering. Berlin: Springer. 

Hinrichs, K.-U., Hayes, J. M., Sylva, S. P., Brewer, P. G. & DeLong, E. F. (1999). Methane- 



SGM symposium 65 



318 M. Kruger and T. Treude 



consuming archaebacteria in marine sediments. Nature 398, 802-805. 
Hinrichs, K.-U., Summons, R. E., Orphan, V., Sylva, S. P. & Hayes, J. M. (2000). Molecular 

and isotopic analysis of anaerobic methane-oxidizing communities in marine sedi- 
ments. Organic Geochem 31,1 685-1 701 . 
Hoehler, T. M., Alperin, M. J., Albert, D. B. & Martens, C S. (1994). Field and laboratory 

studies of methane oxidation in an anoxic marine sediment: evidence for a methano- 

gen-sulfate reducer consortium. Global Biogeochem Cycles 8, 451^-63. 
Holmes, A. J., Owens, N. J. P. & Murrell, J. C. (1996). Molecular analysis of enrichment 

cultures of marine methane oxidising bacteria. J Exp Mar Biol Ecol 203, 27-28. 
Ingram, J. M. & Blackwood, A. C. (1970). Microbial production of phenazines. /\oV Appl 

Microbiol 13, 267-282. 
IPCC (1994). Climate Change 1994: Radiative Forcing of Climate Change and an Evaluation 

of the IPCC 1592 Emission Scenarios. Cambridge: Cambridge University Press. 
Ivanov, M. V., Polikarpov, G. G., Lein, A. Y. & 7 other authors (1991). Biogeochemistry 

of the carbon cycle in the zone of Black Sea methane seeps. Dokl Akad Nauk 555R 

320, 1235-1240. 
Iversen, N. & Blackburn, T. H. (1981). Seasonal rates of methane oxidation in anoxic 

marine sediments. Appl Environ Microbiol 41 , 1 295-1 300. 
Iversen, N. & Jorgensen, B. B. (1985). Anaerobic methane oxidation rates at the sulfate- 

methane transition in marine sediments from Kattegat and Skagerrak (Denmark). 

Limnol Oceanogr 30, 944-955. 
Jorgensen, B. B. (1 982). Mineralization of organic matter in the sea bed - the role of sulfate 

reduction. Nature 296, 643-645. 
Jorgensen, B. B. & Fenchel, T. (1974). The sulfur cycle of a marine sediment model system. 

Mar Biol 24, 189-210. 
Joye, S. B., Boetius, A., Orcutt, B. N., Montoya, J. P., Schulz, H. N., Erickson, M. J. & 

Lugo, S. K. (2004). The anaerobic oxidation of methane and sulfate reduction 

in sediments from Gulf of Mexico cold seeps. Chem Geo/ 205, 219-238. 
Kallmeyer, J. & Boetius, A. (2004). Effects of temperature and pressure on sulfate 

reduction and anaerobic oxidation of methane in hydrothermal sediments of 

Guaymas Basin . Appl Environ Microbiol 70, 1 23 1 -1 233 . 
Knittel, K., Boetius, A., Lemke, A., Eilers, H., Lochte, K., Pfannkuche, O., Linke, P. & 

Amann, R. (2003). Activity, distribution, and diversity of sulfate reducers and other 

bacteria in sediments above gas hydrates (Cascadia Margin, Oregon). Geomicrobiol J 

20, 269-294. 
Knittel, K.„ Losekann, T., Boetius, A., Kort, R. & Amann, R. (2005). Diversity and 

distribution of methanotrophic archaea at cold seeps. Appl Environ Microbiol 71 , 

467^79. 
Kruger, M., Meyerdierks, A., Glockner, F. O. & 8 other authors (2003). A conspicuous 

nickel protein in microbial mats that oxidize methane anaerobically. Nature 426, 

878-881. 
Kruger, M., Treude, T., Wolters, H., Nauhaus, K. & Boetius, A. (2005). Microbial 

methane turnover in different marine habitats. Palaeogeogr Palaeoclimatol Palaeo- 

ecol (in press). 
Lein, A. Y, Ivanov, M. V., Pimenov, N. V. & Gulin, M. B. (2002). Geochemical 

characteristics of the carbonate constructions formed during microbial oxidation of 

methane under anaerobic conditions. Mikrobiologiia 71, 89-102 (in Russian). 
Lidstrom, M. E. (1988). Isolation and characterization of marine methanotrophs. Antonie 

van Leeuwenhoek 54, 1 89-1 99. 



SGM symposium 65 



Physiology and regulation of AOM 319 



Martens, C. S. & Berner, R. A. (1974). Methane production in the interstitial waters of 

sulfate-depleted marine sediments. Science 185, 1 167-1 169. 
Martens, C. S., Albert, D. B. & Alperin, M. J. (1999). Stable isotope tracing of anaerobic 

methane oxidation in the gassy sediments of Eckernforde Bay, German Baltic Sea. 

Am J Sci 299, 589-610. 
Michaelis, W., Seifert, R., Nauhaus, K. & 14 other authors (2002). Microbial reefs in the 

Black Sea fueled by anaerobic oxidation of methane. Science 297, 1013-101 5. 
Milkov, A. V. (2000). Worldwide distribution of submarine mud volcanoes and associated 

gas hydrates. Mar Geo/ 167, 29-42. 
Nauhaus, K., Boetius, A., Kruger, M. & Widdel, F. (2002). In vitro demonstration of 

anaerobic oxidation of methane coupled to sulfate reduction in sediment from a 

marine gas hydrate area. Environ Microbiol 4, 296-305. 
Nauhaus, K., Treude, T., Boetius, A. & Kruger, M. (2005). Environmental regulation of the 

anaerobic oxidation of methane: a comparison of ANME-I and ANME-II com- 
munities. Environ Microbiol 7, 98-1 06. 
Niewohner, C, Hensen, C, Kasten, S., Zabel, M. & Schulz, H. D. (1998). Deep sulfate 

reduction completely mediated by anaerobic methane oxidation in sediments of the 

upwelling area off Namibia. Geochim Cosmochim Acta 62, 455^-64. 
Orphan, V. J., House, C. H. # Hinrichs, K.-U., McKeegan, K. D. & DeLong, E. F. (2001a). 

Methane-consuming archaea revealed by directly coupled isotopic and phylogenetic 

analysis. Science 293, 484-487. 
Orphan, V. J., Hinrichs, K.-U., Ussier, W., Ill, Paull, C. K., Taylor, L T, Sylva, S. P., Hayes, 

J. M. & DeLong, E. F. (2001b). Comparative analysis of methane-oxidizing archaea 

and sulfate-reducing bacteria in anoxic marine sediments. Appl Environ Microbiol 67, 

1922-1934. 
Pancost, R. D., Slnninghe Damste, J. S., de Lint, S., van der Maarel, M. J. E. C. & 

Gottschal, J. C. (2000). Biomarker evidence for widespread anaerobic methane 

oxidation in Mediterranean sediments by a consortium of methanogenic archaea 

and bacteria. Appl Environ Microbiol 66, 1 1 26-1 1 32. 
Pimenov, N. V., Rusanov, I. I., Poglazova, M. N., Mityushina, L. L, Sorokln, D. Y, 

Khmelenina, V. N. & Trotsenko, Yu. A. (1997). Bacterial mats on coral-like struc- 
tures at methane seeps in the Black Sea. Mikrobiologiia 66, 354-360 (in Russian). 
Reeburgh, W. S. (1976). Methane consumption in Cariaco Trench waters and sediments. 

Earth Planet Sci Lett 28, 337-344. 
Reeburgh, W. S. (1980). Anaerobic methane oxidation: rate depth distributions in Skan Bay 

sediments. Earth Planet Sci Lett 47 , 345-352. 
Reeburgh, W. S. (1996). "Soft spots" in the global methane budget. In Eighth International 

Symposium on Microbial Growth on C. Compounds, pp. 334-342. Edited by 

M. E. Lidstrom & F. R. Tabita. Dordrecht: Kluwer. 
Sahling, H., Rickert, D., Lee, R. W., Linke, P. & Suess, E. (2002). Macrofaunal community 

structure and sulfide flux at gas hydrate deposits from the Cascadia convergent 

margin, NE Pacific. Mar Ecol Prog Ser 231, 121-138. 
Schink, B. & Stams, A. J. M. (2001). Syntrophism among prokaryotes. In The Prokaryotes: 

an Evolving Electronic Resource for the Microbiological Community. Edited by 

M. Dworkin, S. Falkow, E. Rosenberg, K.-H. Schleifer & E. Stackebrandt.Heidelberg: 

Springer (http://www.prokaryotes.com). 
Schouten, S., Wakeham, S. G., Hopmans, E. C. & Sinninghe Damste, J. S. (2003). 

Biogeochemical evidence that thermophilic archaea mediate the anaerobic oxidation 

of methane. Appl Environ Microbiol 69, 1 680-1 686. 



SGM symposium 65 



320 M. Kruger and T. Treude 



Seeliger, S. f Cord-Ruwisch, R. & Schink, B. (1998). A periplasmic and extracellular c-type 

cytochrome of Geobacter sulfurreducens acts as a ferric iron reductase and as an 

electron carrier to other acceptors or to partner bacteria. J Bacteriol 180, 3686-3691 . 
Sorensen, K. B., Finster, K. & Ramsing, N. B. (2001). Thermodynamic and kinetic 

requirements in anaerobic methane oxidizing consortia exclude hydrogen, acetate 

and methanol as possible electron shuttles. Microb Ecol 42, 1-10. 
Straub, K. L. & Schink, B. (2003). Evaluation of electron-shuttling compounds in microbial 

ferric iron reduction. FEM 5 Microbiol Lett 220, 229-233. 
Suess, E., Torres, M. E., Bohrmann, G. & 8 other authors (1999). Gas hydrate 

destabilization: enhanced dewatering, benthic material turnover and large methane 

plumes at the Cascadia convergent margin. Earth Planet Sci Lett 170, 1-15. 
Teske, A., Hinrichs, K.-U., Edgcomb, V., Gomez, A. de V., Kysela, D., Sylva, S. P., Sogin, 

M. L. & Jannasch, H. W. (2002). Microbial diversity of hydrothermal sediments in 

the Guaymas Basin: evidence for anaerobic methanotrophic communities. Appl 

Environ Microbiol 68, 1 994-2007. 
Thauer, R. K. (1998). Biochemistry of methanogenesis: a tribute to Marjory Stephenson. 

Microbiology 144, 2377-2406. 
Thiel, V., Peckmann, J., Richnow, H. H., Luth, U., Reitner, J. & Michaelis, W. (2001). 

Molecular signals for anaerobic methane oxidation in Black Sea seep carbonates and 

a microbial mat. Mar Chem 73, 97-1 12. 
Torres, M. E., McManus, J., Hammond, D. E., de Angelis, M. A., Heeschen, K. U., 

Colbert, S. L, Tryon, M. D., Brown, K. M. & Suess, E. (2002). Fluid and chemical 

fluxes in and out of sediments hosting methane hydrate deposits on Hydrate Ridge, 

OR, I: Hydrological provinces. Earth Planet Sci Lett 201, 525-540. 
Treude, T. (2003). Anaerobic oxidation of methane in marine sediments. PhD thesis, 

Universitat Bremen, Bremen, Germany (http://elib.suub.uni-bremen.de/publications/ 

dissertations/E-Diss84 5_treude.pdf). 
Treude, T., Boetius, A., Knittel, K., Wallmann, K. & Jorgensen, B. B. (2003). Anaerobic 

oxidation of methane above gas hydrates at Hydrate Ridge, NE Pacific Ocean. Mar 

Ecol Prog Ser 264, 1-14. 
Treude, T., Niggemann, J., Kallmeyer, J., Wintersteller, P., Schubert, C. J., Boetius, A. & 

Jorgensen, B. B. (2005a). Anaerobic oxidation of methane and sulfate reduction 

along the Chilean continental margin. Geochim Cosmochim Acta 69, 2767-2779. 
Treude, T., Kruger, M., Boetius, A. & Jorgensen, B. B. (2005b). Environmental control on 

anaerobic oxidation of methane in the gassy sediments of Eckernforde Bay (German 

Baltic). Limnol Oceanogr(\n press). 
Valentine, D. L. & Reeburgh, W. S. (2000). New perspectives on anaerobic methane 

oxidation. Environ Microbiol 2, A11-A&A. 
Valentine, D. L, Blanton, D. C, Reeburgh, W. S. & Kastner, M. (2001). Water column 

methane oxidation adjacent to an area of active hydrate dissociation, Eel River Basin. 

Geochim Cosmochim Acta 65, 2633-2640. 
Yamamoto, S., Alcauskas, J. B. & Crozier, T. E. (1976). Solubility of methane in distilled 

water and seawater. J Chem Eng Data 21 , 78-80. 
Zehnder, A. J. B. (1988). Biology of Anaerobic Microorganisms. New York: Wiley. 
Zehnder, A. J. B. & Brock, T. D. (1979). Methane formation and methane oxidation by 

methanogenic bacteria. J Bacteriol 137, 420-432. 
Zehnder, A. J. B. & Brock, T. D. (1980). Anaerobic methane oxidation: occurrence and 

ecology. Appl Environ Microbiol 39, 1 94-204. 



SGM symposium 65 



Biogeochemical roles of fungi in 
marine and estuarine habitats 

Nicholas Clipson, 1 Eleanor Landy 2 and Marinus Otte 3 

1 ' 3 Department of Industrial Microbiology 1 and Department of Botany 3 , University College Dublin, 
Belfield, Dublin 4, Ireland 

2 School of Biomedical and Molecular Sciences, University of Surrey, Guildford GU2 7XH, UK 



INTRODUCTION 

A fungal component of the marine biota was only recognized as recently as 1944 
(Barghoorn & Linder, 1944), and it was not until the 1960s that studies commenced to 
assess the extent and diversity of fungi in marine systems. Since this time, considerable 
effort has been exerted to uncover marine fungal diversity, with high decadal discovery 
indices in the 1970s and 80s (Hawksworth, 1991), resulting in around 1000 fungal 
species known today from marine environments. Nevertheless, it is hardly surprising 
that, with the extent of marine environments globally, we probably have a very in- 
complete view of fungal diversity, together with their frequency and function in these 
ecosystems. The objective of this review is to assess the extent of our present knowledge 
and to highlight future directions to further elucidate their biology and ecology 

THE NATURE OF MARINE ENVIRONMENTS 

Marine ecosystems are globally extensive, and account for around 70 % of global 
surface area. They can be defined generally as aquatic systems influenced by substantial 
concentrations of salts, particularly sodium chloride, from existing oceanic systems. 
Seas and oceans divide between regions bordering and influenced by terrestrial regions 
and the open ocean, which is strongly zoned through the water column. These broad 
boundaries are illustrated in Fig. 1, which also details linkages between marine com- 
partments. At the sea surface, light is an important factor driving primary productivity 
in the photic zone, to a depth of around 200 m (Dring, 1992), with carbon and other 
nutrients moving downwards through the water column to stimulate food webs within 
benthic zones, finally adding to sediments. 

SGM symposium 65: Micro-organisms and Earth systems - advances in geomicrobiology. 
Editors G. M. Gadd, K. T. Semple & H. M. Lappin-Scott. Cambridge University Press. ISBN 521 86222 1 ©SGM 2005 



322 N. Clipson, E. Landy and M. Otte 



r* 



A 



V-i 



.<*> 



CD 

> 

be 



Terrestrial 



> 







\3 

Estuary <^ 

~^r> 



v 

Sediments 



Atmosphere 
/\ 



1 




Coastal waters 

=> < 



Open ocean, pelagic 



c> 




c> 




Deep ocean layer, -_ 

sub-pelagic * J 



Near ocean floor, 
benthic 



a 



Hydrothermal 
vents 



it 



J 



Sediments 



> 



Fig. 1. Interactions between and within marine ecosystems. While most interactions are bi-directional, 
interactions between rivers and estuaries and between hydrothermal vents and their surroundings are 
unidirectional. Circular arrows within oceanic compartments indicate cycles that are largely confined 
within those compartments (after Wangersky, 1 980; Salomons & Forstner, 1 984). 



Salinity concentrations within open ocean systems are characteristically very stable, at 
around 500 mM NaCl, although these are lower in enclosed seas such as the Baltic and 
Black Seas, where NaCl concentrations are diluted by the effects of freshwater inputs 
from riverine discharge and ice melt and may be in the region of 50-200 mM. Typical 
open ocean sea water ion concentrations are given in Table 1. The open ocean water 
column changes with depth, with productivity declining dramatically as light and 
temperature decrease. The ocean floor is an area of low primary productivity and low 
biomass density (Nybakken, 1997) except for deep-sea hydrothermal vent and cold seep 
communities, which are characterized by temperatures in the range 8-16 °C (as 
opposed to the normal 2 °C of most of the deep sea) and high pressures exerted by the 
overlying water column (Ballard, 1977). These environments are generally found at 
the edges of tectonic plates and are typically very rich in reduced sulfur compounds. 
Primary productivity is driven by prokaryal chemolithoautotrophs, which drive 



SGM symposium 65 



Roles of fungi in marine and estuarine habitats 323 
Table 1. Concentrations of major and selected minor constituents of sea water 

Data were compiled from Martin (1 970) and Dring (1 992). 



Constituent 


Concentration 


Constituent 


Concentration 




(g kg 1 ) 


(mM) 


(Mg I" 1 ) 


Chloride 


19-35 


548 


Silicon 


0-4900 


Sodium 


10-76 


470 


Nitrogen (combined) 


0-560 


Sulfate 


2-71 


28 


Phosphorus 


0-90 


Magnesium 


1-29 


54 


Aluminium 


0-10 


Calcium 


0-413 


10 


Iron 


0-1-62 


Potassium 


0-387 


10 


Zinc 


1-48 


Bicarbonate 


0-142 


2 


Iodine 


48-80 


Bromide 


0-067 


0-8 


Copper 


0-5-27 


Strontium 


0-008 


1-5 


Manganese 


0-2-8-6 


Boron 


00045 


0-4 


Cobalt 


0-005-4-1 


Fluoride 


0-001 


0-07 







relatively rich food webs with high productivity (Nybakken, 1997). Cold seeps are often 
associated with hypersaline brines or hydrocarbon seeps. 

Near-shore marine ecosystems are much more variable and dynamic, particularly being 
affected by terrestrial and meteorological influences. The most extensive and productive 
of these environments are saline wetlands, which characteristically form as salt marshes 
in polar and temperate zones dominated by low-growing herbs and mangrove 
swamps in tropical regions dominated by salt-tolerant trees (Packham 6c Willis, 1997). 
Both systems are affected by diurnal tidal inundation (much less so in enclosed seas, 
where diurnal variation in tidal height is very small), which makes soil and sediment 
environmental parameters highly variable, particularly for soil salinity, oxygenation 
and drainage (Jefferies et aL, 1979; Armstrong et al., 1985). Terrestrial influences can 
also be important, resulting in inputs of nutrients and freshwater as a result of riverine 
deposition and run-off. Other salt-affected coastal environments might include dune 
and shingle systems and rocky foreshores, which are influenced as spray communities 
from sea water (Packham &C Willis, 1997). Adjacent to coastal regions, and where 
continental shelves are shallow, coastal sea communities form, including coral reefs, 
which are found in both tropical and cold seas. Marine communities can also be asso- 
ciated with former oceanic activity, generally where seas have become separated from 
present oceanic areas. These can either form highly saline inland seas such as the Dead 
Sea, where saline concentrations can reach 4-5 M (Buchalo et al., 1998), or evaporate 
completely to form salt pans and salt deserts. Many of these highly saline environments 
are dominated by ions other than sodium and chloride, including magnesium, calcium 



SGM symposium 65 



324 N. Clipson, E. Landy and M. Otte 

and potassium, with the counter ions carbonate, bicarbonate and sulfate (Flowers 
et ai, 1986) . Perhaps 10 % of global land surface area is dominated by these hypersaline 
environments. 

For a detailed view of the elemental contents of the major environments associated 
with marine systems, see Hutzinger (1980) and Salomons & Forstner (1984). In general 
terms, open oceans are more dilute than areas above continental shelves, adjacent to 
land masses or estuarine. Terrestrial and near-shore systems are usually well connected 
via rivers, leading to rapid elemental transfer from terrestrial to marine ecosystems. 
Transport to the open ocean and its underlying layers is much slower. In both cases, 
many elements are immobilized in sediments, which act as elemental sinks after 
deposition. These can be remobilized back slowly to the water column, particularly for 
phosphorus and metals, where transfer to the atmosphere is negligible (Elmsley, 1980; 
Salomons & Forstner, 1984). Some elements have a large atmospheric component, such 
as carbon, nitrogen and sulfur, which are much more rapidly cycled in these systems. 
This may also include some metals and metalloids such as mercury, arsenic and 
selenium, which can be methylated and lost to the atmosphere. 

Although open ocean systems are more stable than marine systems at terrestrial 
boundaries, important dynamics do occur through the water column, largely reflecting 
oceanic mixing by circulatory systems. Through the oceans there are local and depth- 
delineated changes in factors such as temperature, salinity, oxygen and nutrient levels. 
Examples are water-column transect studies across the Atlantic Ocean (Lavin et al., 
2003) and near-shore studies across the Strait of Georgia (British Columbia) (Masson, 
2002). In the open ocean North Atlantic, Lavin et al. (2003) found a mean salinity of 
around 3-5 %, varying between 3-45 and 3-7 % between 6000 m depth and the surface. 
Mean surface oxygen was 200 ^mol kg -1 , declining to around 150 |imol kg -1 at 1000 m 
but increasing to around 250 ^imol kg -1 at 6000 m. Silicates, nitrates and phosphates 
were all depleted in the surface oceanic layers, increasing to average levels at around 
1000 m and below. Oxygen and nutrient levels are also heavily influenced by the 
circulation of currents through these systems, leading to localized variations. 



MARINE FUNGAL DIVERSITY IN MARINE ECOSYSTEMS 

The presence of fungi in marine ecosystems was first recognized by Barghoorn & 
Linder (1944), who proposed the term 'marine fungi' for fungal species living in these 
systems. Although there has been considerable effort since the early 1960s to elucidate 
marine fungal biodiversity, present estimates probably grossly underrepresent its true 
extent. This is for a number of reasons. Firstly, the oceans are exceptionally extensive, 
both on the basis of their global surface area and as they are compartmentalized 

SGM symposium 65 



Roles of fungi in marine and estuarine habitats 325 

through the water column. Many of these environments are inaccessible and a 
challenge to meaningful sampling. Secondly, total fungal diversity is itself grossly 
underestimated, with many geographical areas undersampled, many species lying 
hidden in interactions with other organisms such as insects or plants and the problems 
of isolating and culturing most species. Overall, Hawksworth (2001) estimates that 
around 80000 species have been identified and classified in a background of possibly in 
excess of 1-5 million species. It is within this background that marine fungal diversity is 
probably heavily underestimated. Generally, two approaches are used to assess marine 
fungal diversity. Firstly, direct observation of fungal structures on natural substrates or 
baited substrates such as wood blocks or litter bags can be made. Secondly, isolations 
can be made onto traditional agar-based cultural media. Certainly, in the latter case, 
culturability of environmental isolates of fungi probably falls into the region of that for 
bacteria, in the region of 1-5 %. There have been very few studies (e.g. Buchan et aL, 
2003) to test whether conventionally made diversity estimates in a given location corres- 
pond to molecular estimates. In the case of fungal diversity during the degradation of 
the salt-marsh halophyte Spartina alterniflora, Buchan et al. (2003) found Phaeo- 
sphaeria spartinicola, Phaeosphaeria halima and a Mycosphaerella sp. to be dominant 
using terminal restriction fragment length polymorphism. 

There are also problems in defining what a marine fungus is. Kohlmeyer 6c Kohlmeyer 
(1979) advanced three ecological groupings of fungi occurring in marine environments, 
including obligate marine fungi, facultative marine fungi and terrestrial fungi (see also 
Kohlmeyer 6c Volkmann-Kohlmeyer, 2003). Obligate marine fungi were defined by 
their inability to complete their life cycle outside the marine environment, whereas 
facultative species could complete life cycles in either marine or terrestrial systems but 
were of terrestrial origin. Terrestrial fungi do not have the ability to complete life cycles 
in marine environments and, although commonly isolated, might just exist in a dor- 
mant state until more favourable conditions prevail for spore germination (Kohlmeyer 
6c Kohlmeyer, 1979). On a physiological basis, it is unclear what confers the marine 
lifestyle of fungal organisms. The principal adaptation required by marine fungi is 
probably salt-tolerance. Many terrestrial species, for example most aspergilli or 
penicillia (e.g. Beever 6c Laracy, 1986), are exceptionally salt-tolerant both vegetatively 
and for sporulation, but are not generally found to be active in marine ecosystems. 
Some terrestrial Aspergillus species have been found in marine ecosystems, such as 
Aspergillus sydowii and Aspergillus fumigatus, which cause disease in the Caribbean 
sea fan Gorgonia ventalina. Disease-causing populations are probably maintained by 
continual influx of vegetative material from neighbouring terrestrial environments 
(Smith et al., 1996; Geiser et al., 1998). The question concerning the nature of the 
marine fungal lifestyle remains. It might relate to the substrate conditions existing 
in marine ecosystems, sensitivity at a certain lifestyle stage (for example sporulation or 

SGM symposium 65 



326 N. Clipson, E. Landy and M. Otte 

spore germination), the oligotrophic nature of the marine environment, the selection of 
appropriate media for physiological studies or the ability to attach to substrate. 

Biogeographically, and in terms of marine ecosystem-type, fungi appear to be ubiqui- 
tous. Nevertheless, some caution must be expressed with this view, as there are sub- 
stantial areas of the globe which have not been surveyed for this group. Marine fungi 
appear to be found worldwide, including polar regions (Pugh & Jones, 1986; Grasso 
et al.j 1997), although probably the highest levels of diversity so far have been found in 
the tropics associated with coastal ecosystems, especially mangroves. This may reflect 
that most effort has been exerted within these regions. Additionally, coastal and near- 
shore habitats are much better studied than oceanic habitats because of the expense of 
boat time. Benthic and deep-sea fungal communities have only been sampled on a very 
few occasions (e.g. Alongi, 1987; Nagahama et al., 2001). Another important issue 
leading to diversity underestimates is the issue of culturability. Traditionally, marine 
fungal diversity has been assessed almost exclusively by culture-based isolation 
approaches or direct observation of natural substrates. 

The most comprehensive assessment of known marine fungal diversity to date was 
made by Jones &C Mitchell (1996), taking into account those that had been described 
and estimating species waiting for full description and publication, including thrausto- 
chytrids and lower fungi. They advanced a total of around 1400 species. Another 
approach to assessing marine fungal diversity has been to compile checklists of species 
occurring either on particular substrates or in known areas. Schmit & Shearer (2003) 
compiled a checklist from available literature of mangrove-associated fungi, recording a 
total of 625 species, with both marine and terrestrial species recorded. Advances in 
producing checklists of fungal species illustrate how knowledge of their diversity has 
been extended. Kohlmeyer (1969) could list 76 mangrove-fungi species, Hyde 6c Jones 
(1988) listed 90 species, with Jones & Alias (1997) extending this to 268 species. These 
latter two studies speculated that mangrove-fungal diversity is most extensive in South- 
East Asia, reflecting greater substrate diversity, although this probably also reflects 
greater numbers of isolation studies. 

Two studies have aimed to produce checklists on an area basis. Gonzalez et al. (2001) 
have produced a checklist of higher marine fungi of Mexico, listing 62 species. They 
believed that this number significantly underestimates the true diversity of Mexican 
marine fungi. In a literature search of marine fungi identified within the seas and 
coastal regions of the European Union (EU), Clipson et al. (2001) found 318 species 
described. Interestingly, all three checklists, together with that of Jones & Mitchell 
(1996), found that marine fungi were predominantly either ascomycetes (particularly 
the Halosphaeriales) or mitosporic fungi, with details given in Table 2. Many of the 

SGM symposium 65 



Roles of fungi in marine and estuarine habitats 327 



Table 2. Marine fungal diversity, based upon a global estimate and literature 
searches for the European Union area, Mexico and on mangrove substrates 

Data were taken from Jones & Mitchell (1 996) (global), Clipson etal. (2001 ) (EU marine), Gonzalez ef 
a/. (2001 ) (Mexico marine) and Schmit & Shearer (2003) (mangrove). Species awaiting description and 
publication are indicated in parentheses. 



Fungal group 


Global 


EU marine 


Mexico marine 


Mangrove 


Lower fungi 


100 (+32) 


12 




14 


Thraustochytrids 


40 








Ascomycetes 




214 


47 


278 


Hemiascomycetes 


50 








Euascomycetes 


305 (+143) 








Deuteromycotina 


79 (+200) 


84 


14 


277 


Basidiomycotina 


7 (+7) 


10 


1 


30 


Trichomycetes 


23 








Lichens 


18 (+410) 









genera found have common characteristics (reproduction in aquatic habitats; thin- 
walled, unitunicate, deliquescing asci; central pseudoparenchyma in immature asci; 
hyaline, bicelled ascospores with polar or equatorial appendages) which Jones (1995) 
considers as adaptations to aquatic habitats. The large number of mitosporic fungi 
found is problematic because little is known about their relationship with their sexual 
stages. Currently, the EU list is being updated, with 338 marine species being recognized 
by 2005 (E. Landy, unpublished). 

Schmit &C Shearer (2003) also explored substrate preferences of mangrove fungi, with 
marine ascomycetes being intertidal or submerged species and mitosporic species being 
associated with sediments and possibly of terrestrial origin. Most marine ascomycetes 
possess appendages or gelatinous sheaths to aid attachment, making them particularly 
adapted to the forces operating in intertidal zones. Also, there is a further problem in 
that marine fungi are defined only imprecisely, with single mangrove trees possessing 
populations of both marine and terrestrial mangrove fungi. At what position on a 
single tree, and exposed to what outside environmental conditions, is a fungal species 
marine? At best, this must represent a rather arbitrary division. As Schmit & Shearer 
(2003) point out, we need to know much more about the evolution and ecology (related 
to their physiology and genetics) to understand the determinants of distribution of 
mangrove fungi. The same is true for marine fungal distribution in all other marine 
habitats. To date, only coastal habitats have been explored in any detail, leaving fungal 
diversity in the vast bulk of the open oceans and their depths largely undetermined. It is 
known that yeasts are frequent in open seas (e.g. Van Uden & Fell, 1968). In a study 
based in the Atlantic south of Portugal, Gadanho et al. (2003) identified 31 yeast taxa 



SGM symposium 65 



328 N. Clipson, E. Landy and M. Otte 

using microsatellite-primed PCR. Yeast cell densities typically decline with increased 
depth and distance from land (Gadanho et al., 2003), presumably reflecting substrate 
availability and favourable environmental conditions. Yeasts are also known in the deep 
sea, with Nagahama et al. (2002) reporting the presence of red yeasts from abyssal 
zones. There are also some reports of fungi from hypersaline habitats. From the Dead 
Sea, Kritzman (1973) reported an osmophilic yeast, and Buchalo et al. (1998) made the 
first reports of filamentous fungi, including species of Gytnnascella, Ulocladium and 
Penicillium, for this hypersaline habitat. A number of species of melanized yeast-like 
fungi have been reported from hypersaline salterns (Gunde-Cimermann et al., 2000). 

FUNGAL ADAPTATION TO MARINE ENVIRONMENTS 

That a specialized mycoflora has become representative of marine ecosystems would 
suggest that specialized adaptations are necessary for the marine fungal habit. Clearly, 
adaptation to a combination of factors such as salinity, redox potential, pressure, 
substrate availability and quality and temperature, throughout the fungal life cycle, 
determines fungal diversity and activity in individual marine ecosystems. 

The best-studied fungal response to such factors is that to salinity, which is ubiquitous 
in marine ecosystems. Much of our understanding of the mechanisms by which marine 
fungi adapt to salt comes from the physiological studies on the marine hyphomycete 
Dendryphiella salina by Jennings and co-workers in the 1980s and 1990s and from 
subsequent genetic studies on this and other non-marine fungi (e.g. Clement et al., 
1999; and see Bohnert et al., 2001; Hooley et al., 2003). Integrating information 
between these marine species and other osmotolerant but non-marine species has led to 
a model of fungal cellular adaptation to salinity Central to osmotic adaptation by fungi 
is maintenance of osmotic gradients across the hyphal membrane to maintain inwardly 
directed water fluxes. Water (=osmotic) potential of sea water is around -24 MPa, 
meaning that cellular water potentials have to be maintained more negative. Fungal 
cells have rigid cell walls and possess turgor pressure generated as the difference 
between cellular water potential and cellular osmotic potential (generated by osmo- 
tically active solutes), turgor pressure being an important component of apical growth 
and expansion processes in fungi (Money, 1997). To generate hyphal water potentials, 
fungal cells have to maintain substantial concentrations of intracellular solutes. 
The fungal cell has two potential sources of osmotically active solutes, ions taken up 
from the external environment and organic solutes either synthesized from metabolism 
or taken up externally if available. Both these strategies are potentially problematic 
to fungal growth; many ions, particularly sodium and chloride, are toxic to cellular 
metabolic processes, and organic solutes for osmotic purposes can divert carbon from 
growth. In vitro studies have shown that many enzymes involved in fungal primary 
metabolism are inhibited substantially by sodium chloride levels well below (approx. 

SGM symposium 65 



Roles of fungi in marine and estuarine habitats 329 

100 mM) those of sea water (Pa ton & Jennings, 1988), although this may be altered in 
vivo by either upregulation of protein systems or by effects of the counter-anion 
environment (Nilsson & Adler, 1990; Hooley et al., 2003). 

In Dendryphiella salina, a balance of these two approaches has been found to generate 
hyphal osmotic potential (Clipson &C Jennings, 1992). X-ray microanalysis studies 
of hyphal cells growing at sea-water concentrations (500 mM NaCl) indicated 
cytoplasmic concentrations of around 170 mM sodium (Clipson et al., 1990). This is 
probably still compatible with cytosolic enzyme activity, albeit with some inhibition of 
more sensitive enzymes. There was also no evidence of preferential accumulation 
of ions in vacuoles, which are much less involved in metabolism than is the cytoplasm 
(Clipson et al., 1990). Many algal and plant halophiles have large vacuoles which are 
used as ion stores for osmotic adjustment, in conjunction with low cytoplasmic ion 
levels (Flowers et al., 1986). Although Dendryphiella salina does not have large vacu- 
oles, vacuolar localization of ions may be important in those fungal species that do. 

In Dendryphiella salina, the balance of osmotic potential is generated through 
compatible solutes, either accumulated or synthesized. These solutes are 'compatible' 
in that they do not interfere with cellular metabolism. In fungi, the principal compatible 
solutes are polyols, including glycerol, mannitol, arabinitol and erythritol (Blomberg & 
Adler, 1993). In most fungi, it is mannitol which is the primary osmoresponsive polyol, 
with others generally increasing as cultures age. For species such as Dendryphiella 
salina and the osmophilic yeasts Debaryomyces hansenii and Zygosaccharomyces 
rouxii, where polyol concentrations have been measured under saline conditions 
in vitro, polyols generated between 36 and 75 % of cellular osmotic potential (Hooley 
et ai, 2003) . Other compatible solutes may also be important in marine fungi, including 
proline and trehalose (Jennings & Burke, 1990; Davis et ai, 2000). 

Clearly, these osmotic mechanisms are under genetic control, although precise details 
are not understood at present due to the difficulty of genetic manipulation in marine 
fungal species. The genetics of osmotolerance in yeast is quite well understood and has 
been reviewed recently by Yale & Bohnert (2001) and Hooley et al. (2003). Saccharo- 
myces cerevisiae is not a good model for halophilic fungi as it is relatively salt-sensitive, 
with the halophilic yeast Debaryomyces hansenii being more representative. A number 
of genes involved in stress responses by the marine yeast Debaryomyces hansenii have 
been characterized, including the cloning of the genes for superoxide dismutase 
(Hernandez-Saavedro & Romero-Geraldo, 2001), a homologue of the high-osmolarity 
glycerol response gene HOG1 encoding a MAP kinase (Bansal & Mondal, 2000) 
and the identification of several ion-transport-related genes including DhENAl and 
DhENA2 for sodium ATPases (Almagro et al., 2001). Nevertheless, there is still a very 

SGM symposium 65 



330 N. Clipson, E. Landy and M. Otte 

incomplete view of the genetics of salt tolerance in halophilic fungi in general and 
filamentous marine fungi in particular. 

So far, an overview of fungal salt tolerance has been presented. In most marine environ- 
ments, particularly the open ocean, salt concentrations do not fluctuate appreciably or 
rapidly, and adaptation is probably a continuous and gradual process presenting little 
problem to fungal organisms. In coastal regions, the influence of tidal effects, evapo- 
transpiration and freshwater inputs makes salt concentrations much more unstable 
(Armstrong et al., 1985). Although the mechanisms are not known, presumably fungal 
species living in these environments are adapted to adjust osmotically rather quickly in 
response to such fluctuations. Such an ability might differentiate these species from 
terrestrial fungi. Another important issue is tolerance to marine environmental factors 
through the full fungal life cycle, particularly spore germination and early hyphal 
development, and sporulation processes. These have not been examined to any extent, 
although it is well known that many non-marine osmophilic fungi, particularly 
aspergilli and penicillia, are tolerant through these stages and certainly in the labora- 
tory can complete life cycles on highly saline media (Hooley et al., 2003). 

Marine fungi, dependent upon the habitat they occupy, may have to be adapted to other 
adverse environmental factors. On salt marshes, sediments and soils tend to be anoxic, 
with highly negative redox potentials. Most fungal species are aerobic (Carlile et al., 
2001) and salt-marsh soils probably do not form an important habitat for fungi. 
Although fungal species can be cultured from marine soils, these are probably transient 
species only present as spores and not vegetationally active. 

Open oceans tend to be relatively oligotrophic except where ocean circulatory systems 
cause upwelling of nutrients, for example the Benguela System off south-west Africa. 
This may lead to planktonic blooms in the photic zone, giving localized regions with 
higher potential substrate concentrations for degrading organisms such as fungi and 
bacteria. Within the water column, it might have been thought that low oxygen contents 
might be inhibitory to fungal activity, although recent oceanological studies would 
indicate that deep-sea water oxygen concentrations are probably close to saturated 
(Lavin et al., 2003). Lorenz & Molitoris (1997) tested the effects of simulated deep-sea 
conditions on a number of marine yeasts, finding that many grew under pressures 
equivalent to a depth of 4000 m. 

MARINE BIOGEOCHEMICAL PROCESSES 

Our view of the role of fungi in biogeochemical processes remains rather limited, 
perhaps with the exception of some coastal ecosystems. This mirrors diversity esti- 
mates, particularly with regard to fungal distribution in the open ocean away from 

SGM symposium 65 



Roles of fungi in marine and estuarine habitats 331 

coastal regions. In terms of substrate availability, there are radical differences between 
the marine-terrestrial interface, perhaps one of the most productive ecosystems on 
Earth (Newell, 1996), and the open ocean, which is relatively oligotrophic. Fungi are 
heterotrophs and their abundance is likely to reflect availability of carbon-based 
substrates. Knowledge concerning the fungal contribution to marine cycling processes 
also mirrors research effort and ease of study Coastal environments are generally close 
to research centres and require little specialist equipment to sample. In contrast, studies 
performed at sea require expensive boat time, detailed planning and, particularly for 
studies in the benthic and abyssal zones, sophisticated sampling equipment. 
Nevertheless, this has been carried out widely and successfully for marine bacteria (e.g. 
Eilersrt*/.,2000). 

Biogeochemical cycling 

Biogeochemical cycling refers to all processes and pathways involved in the cycling 
and turnover of elements. This includes not only chemical transformations, but 
also physical processes such as adsorption/desorption and biological transfer within 
(compartmentation) and between organisms. The best studied and most relevant 
elements important in marine biocycling are carbon, nitrogen, phosphorus and sulfur, 
for which overviews are given in Fig. 2. For all four elements, the fluxes between and 
within the compartments are very small compared with the sizes of the terrestrial 
and oceanic pools. The fluxes of carbon and nitrogen between the oceans and the 
atmosphere are bi-directional and of equal size, while the fluxes of phosphorus and 
sulfur show a net deposition. The oceanic pools of nitrogen and phosphorus are of the 
same order of magnitude as the terrestrial pools. In contrast, the oceanic pools for 
carbon and sulfur are several orders of magnitude larger than the terrestrial pools. 
While data available for carbon indicated that there is a distinct difference in the sizes of 
the carbon pools between the surface layer and deep ocean, no such data are available 
for the other elements. From a global perspective, all compartments are connected, 
but this does not apply at the organismal (fungal) level (Hutzinger, 1980), with little 
connectedness between open ocean near-surface ecosystems and the ocean floor. 
Within the marine environment, a number of more or less distinct systems can be 
identified that each have their own biogeochemical cycles (see Fig. 1). The focus of the 
ensuing sections will be on biogeochemical cycling of individual elements, particularly 
carbon, nitrogen, phosphorus and sulfur. Although these will be considered separately 
for clarity, it must be noted that they are intrinsically linked within biogeochemical 
cycling processes, all being basic constituents of organic matter. 

Carbon, nitrogen and phosphorus. The quality and availability of organic 
compounds based upon carbon skeletons is central to fungal distribution and function 
within marine ecosystems. Fungi develop a range of responses to organic matter within 

SGM symposium 65 



332 N. Clipson, E. Landy and M. Otte 



Terrestrial 
2500 




face 725 
40 



o 



Open ocean 
Deep 38 000 




0-47 x10 6 




N 



0-16x10 6 



o 



1000 



0-13 x10 6 



0-27 x10 6 



210 




95 



o 



1-3x10 9 



140 



Fig. 2. Fluxes and cycling of carbon, nitrogen, phosphorus and sulfur through the oceans. Pools 
(boxes) are expressed in 10 15 g for carbon and 10 12 g for nitrogen, phosphorus and sulfur; fluxes 
(arrows) are in 10 15 gyear 1 for carbon and 10 12 g year 1 for nitrogen, phosphorus and sulfur. Circular 
arrows within oceanic compartments indicate cycling through biota, but no such data were available 
for sulfur (after Pierrou, 1976; Richey, 1983; Rosswall, 1983; Freneyefa/., 1983). 



ecosystems, dependent upon the condition of organic matter, and are capable of 
deriving nutrition saprotrophically from non-living organic matter or either biotrophi- 
cally (symbiosis) or pathogenically from other living organisms. Certainly, we know 
most about how fungi derive nutrition saprotrophically; nevertheless, there are many 
examples of fish, invertebrate and plant pathogens in both open ocean and coastal 
systems (see Porter, 1986, for overview) and examples of symbionts, particularly as 
mycorrhizas and endophytes of marine plants (Cooke et al., 1993; Cornick et aL, 2005). 
Availability of hosts will be important in determining the distribution of such fungal 
interactions, although this has not been studied to any great extent. 



SGM symposium 65 



Roles of fungi in marine and estuarine habitats 333 



Table 3. Production of exoenzymes for the breakdown of substrate polymers 
demonstrated in isolates of filamentous marine fungi 

Data were taken from Torzilli (1982), Schaumann ef a/. (1986), Rohrmann ef a/. (1992) and Bucheref 
a/. (2004). 



Marine fungal species 



Exoenzyme produced 



Lulworthia spp. 

Halocyphina villosa, Lulworthia spp. 

Halocyphina villosa 

Halocyphina villosa, Lulworthia spp., Corollospora maritima, 
Digitatispora marina, Cirrenalia pygmea, Varicosporina ramulosa 

Lulworthia spp. 

Pleospora pelagica, Pleospora vagans, Phaeosphaeria typharum 

Halocyphina villosa, Pleospora pelagica, Pleospora vagans, 
Phaeosphaeria typharum 

Lulworthia spp. 

Lulworthia spp. 

Lulworthia spp. 

Lulworthia spp. 

Corollospora maritima, Cirrenalia pygmea, Varicosporina ramulosa, 
Pleospora pelagica, Pleospora vagans, Phaeosphaeria typharum 

Lulworthia spp. 

Lulworthia spp. 

Ascocratera manglicola, Astrosphaeriella striatispora, 
Cryptovalsa halosarceicola, Linocarpon bipolaris, Rhizophila marina 



Carrageenase 
Laminarinase 
Gelatinase 
Caseinase 

Alginase 

Pectinase 

Xylanase 

Agarase 
Lipase 
Amylase 
Chitinase 
Cellulase complex 

Laccase 

Tyrosinase 

Ligninase 



Central to fungal nutrition is the production and extracellular secretion of degradatory 
enzymes, making fungi particularly effective in the breakdown of complex polymeric 
compounds. A number of studies has examined the range of exoenzymes produced 
by marine fungal saprotrophs in vitro (Torzilli, 1982; Torzilli & Andrykovitch, 1986; 
Molitoris & Schaumann, 1986; Schaumann et al., 1986; Lorenz & Molitoris, 1992; 
Rohrmann et al., 1992), which are summarized in Table 3. 

Although there is little quantitative information except for salt-marsh systems, fungal 
biomass will reflect substrate availability This differs markedly between substrate-rich 
coastal ecosystems and the generally oligotrophic open ocean. In the open ocean, 
fungal communities are largely yeast-dominated, as opposed to comprising filamentous 
species (Sieburth, 1979). In a study in the Pacific, Van Uden & Fell (1968) found viable 
yeast cell numbers varying between 13 and 274 cells 1 _1 . In general, yeast cell densities 
decline with increased depth and distance from land (Gadanho et al., 2003), pre- 
sumably reflecting substrate availability and favourable environmental conditions. 



SGM symposium 65 



334 N. Clipson, E. Landy and M. Otte 

Substrate differences do occur in the open ocean, particularly where the upwelling of 
colder waters creates regions of high productivity Presumably, the activity of recycling 
organisms such as fungi will be higher in these regions. Nevertheless, we know very 
little about the role of yeasts in biogeochemical cycling in such systems. 

Coastal zones are highly productive and rich in vegetation, which provides complex 
polymeric material for recycling by fungi and other degraders. Fungi are highly effective 
degraders of lignocellulosic material and may be central in commencing mixed degra- 
dative processes for this material (see Newell, 1996). Many fungal isolates from marine 
systems have been classified as white rot (lignin-degrading) organisms (Molitoris & 
Schaumann, 1986; Bucher et al., 2004), particularly where isolates have been made from 
woody material. Overall, it is clear that fungi of coastal zones are able to produce a very 
extensive battery of exoenzymes, making them important components of recycling 
activity 

Nitrogen is essential to fungal growth and its availability is central to fungal activity in 
all environments, including the marine environment (Sguros 6c Simms, 1963; Jennings, 
1989; Feeney et ai, 1992; Edwards et ai, 1998; Ahammed & Prema, 2002). In general, 
fungi can utilize both inorganic and organic forms of nitrogen, depending upon species 
and environment (Carlile et ai, 2001). Little is known about how marine fungi derive 
nitrogen from their environments or contribute to nitrogen cycling, although both 
generalist and specialist utilizers will be present. In marine systems, total inorganic 
nitrogen (as combined nitrogen) ranges between and 560 jig 1 _1 (Dring, 1992). This is 
likely to limit fungal growth very much in the same way that marine primary produc- 
tivity is limited by oceanic nitrogen and phosphorus concentrations (Dring, 1992). 
Fungi probably derive the bulk of their nitrogen requirements from organic substrate 
resulting from primary production, probably as reduced nitrogen, with close linkage 
between factors affecting primary productivity and fungal biomass. 

There are some examples of fungal species in marine habitats deriving nitrogen from 
quite specific organic sources. For example, the trichomycete fungus Entromyces 
callianassae occurs exclusively in the foregut lining of the callianassid shrimp 
Nihonotrypaea harmandi and appears to be involved in the hydrolysis of certain 
nitrogen-containing compounds (Kimura et al., 2002). Like their terrestrial counter- 
parts, some marine fungi have been found to degrade nitrogen-rich materials such as 
chitin which are not easily utilized by other organisms. Kirchner (1995) found that the 
yeast-like fungus Aureobasidium pullulans was involved in the degradation of chitin in 
the moults and carcasses of the marine copepod Tisbe holothuriae. In general, marine 
fungi may play a role in the degradation of more recalcitrant nitrogen-containing 
compounds, which may be only slowly degraded by other recyclers. Although quanti- 

SGM symposium 65 



Roles of fungi in marine and estuarine habitats 335 

fying the contribution of marine fungi to nitrogen cycling is extremely difficult, Newell 
(1996) considered that, on salt marshes, fungi could account for nearly all the nitrogen 
present in decaying standing biomass of salt-marsh plants. Certainly, the role of fungi 
in the marine nitrogen cycle may have been grossly underestimated, particularly at the 
marine-terrestrial interface; our knowledge of the fungal contribution in the open 
ocean is almost entirely absent. 

Open ocean phosphorus concentrations are very low, ranging from to 90 ^ig l -1 (Dring, 
1992). Similarly to nitrogen, these are likely to be too low to support appreciable fungal 
growth, with most phosphorus derived from organic substrates related to primary 
productivity However, very few reports exist on marine fungi in relation to phosphorus. 
Bongiorni & Dini (2002) described the habitat, seasonality and species distribution of 
thraustochytrids, marine fungoid protists, and found that, among the nutrients assessed 
in the water column, only total phosphorus was related to variations in thraustochytrid 
densities. 

On salt marshes, fungi may be involved in phosphorus cycling as mycorrhizas, with 
such associations reported in mangroves (Sengupta & Chaudhuri, 2002) and some 
salt-marsh plant species (Carvalho et ai, 2001). Mycorrhizal colonization of salt-marsh 
halophytes tends to be quite poorly distributed, with some of the most widespread salt- 
marsh plant species such as the Chenopodiaceae and many of the Graminae not being 
strongly mycorrhizal. In contrast, halophytic members of the Asteraceae do tend to be 
functionally mycorrhizal (Rozema et ai, 1986; Carvalho et ai, 2001) . In the open ocean, 
a much more oligotrophic environment than coastal waters, phosphorus availability 
is known to limit bacterial productivity (Van Wambeke et ai, 2002), although effects 
on yeast populations are not known. 

Sulfur. While sulfur plays an important role in global biogeochemical cycling in 
general, and in marine environments in particular, little is known about fungal 
involvement in sulfur cycling. Sulfate is a major constituent of sea water, which typically 
contains around 2-6 gl _1 (Dring, 1992). Also, sediments and hydro thermal vent regions 
tend to be rich in sulfides (Dring, 1992). Fungi are known to be able to oxidize inorganic 
sulfur (Wainwright, 1989) and to produce compounds such as dimethyl sulfide 
(Slaughter, 1989), which is an important compound in the cycling of sulfur in marine 
environments. Fungi are therefore likely to be involved in biogeochemical cycling of 
sulfur, particularly in marine sediments. 

A few studies report on fungal sulfur metabolism of inorganic and organic sources. 
Phae 6c Shoda (1991) reported a fungal species able to degrade hydrogen sulfide, 
methanethiol, dimethyl sulfide and dimethyl disulfide, while Faison et al. (1991) 

SGM symposium 65 



336 N. Clipson, E. Landy and M. Otte 

reported on a coal-solubilizing Paecilomyces sp. able to degrade organic sulfur com- 
pounds such as ethyl phenyl sulfide, diphenyl sulfide and dibenzyl sulfide. Reports 
on the involvement of fungi associated with marine organisms in relation to sulfur are 
scant. Zande (1999) described an ascomycete on the gills of the gastropod Bathynerita 
naticoidea, which the author suggested could be involved in detoxification of the 
abundant sulfide compounds in its habitat. Also associated with sulfide-rich marine 
environments is the fungus Fusarium lateritium, which is able to degrade dimethyl- 
sulfoniopropionate derived from algae and the salt-marsh grass Spartina alterniflora 
(Bacic&Yoch,1998). 

FUNGAL CONTRIBUTION TO CYCLING IN MARINE 
ECOSYSTEMS 

For most marine ecosystems, the contribution by fungi to biogeochemical cycling can at 
best only be speculated upon. As considered above in the discussion of the effect of 
carbon availability on marine fungal processes, fungal biomass appears to be maximal 
(and presumably most active) where organic substrates are at high levels. This gives a 
fundamental difference between the generally oligotrophic open ocean systems and the 
land-ocean interface, where some of the most globally productive ecosystems exist. For 
the latter, salt-marsh ecosystems are well studied and much is known about the role 
of fungal activity in these systems (see Newell, 1993a, 1996). 

Salt-marsh ecosystems have high rates of primary productivity resulting from nutrient 
recycling, riverine external loading, tidally mediated net export of organic matter, 
balancing of nutrient inputs and outputs as well as aerobic and anaerobic microbial 
activity (Adam, 1990). The most intensively studied of these systems are those of the 
eastern seaboard of the United States, where a sophisticated overview of the role of 
fungal activity is now available (Gessner & Kohlmeyer, 1976; Gessner, 1977; Newell 
et al., 2000), particularly for the salt-marsh cordgrass Spartina alterniflora Loisel. 
Spartina alterniflora is an abundant and highly productive plant species in this system 
and has frequently been used as the model for understanding fungal-plant interactions 
in this unusual environment (Newell, 1993a, b). 

Spartina alterniflora is highly lignocellulosic (>70 %) and is extensively colonized by 
fungi during decomposition (Newell et al., 1996). Newell (1996) considered that there 
are four main strategies by which marine microbes capture metabolizable organic 
carbon: (i) maximized use of surface area coupled with high substrate affinity and 
suitable enzymes that are capable of diffusing into solid particles; (ii) penetration by 
tunnelling or surface erosion; (iii) penetration by absorptive ectoplasmic nets or 
rhizoids; and (iv) pervasion via networks of self-extending tubular flowing cytoplasm 
within rigid microfibrillar tubes of chitin laminaran or cellulose laminaran. Marine 

SGM symposium 65 



Roles of fungi in marine and estuarine habitats 337 

fungi contribute to the degradation of Spartina alterniflora through the production of 
lignocellulosic enzymes, which commence the breakdown of this material as plants 
senesce leading to a subsequent mixed degradation. The life cycle of Spartina alterni- 
flora involves shoot growth from perennial rhizomes from April to July, maturation and 
seed production from July to October, frost-mediated death in October and an ensuing 
decomposition period aided by weathering, shearing and tidal action. Fungi appear 
to be central in commencing the decomposition process (Newell, 1993a), efficiently 
breaking down lignin and cellulose into smaller, low-molecular-mass fragments 
available either for conversion to fungal biomass or for further decomposition by 
bacteria over the ensuing months. A number of distinct laccase sequence types have 
been identified from a group of fungi isolated from this salt-marsh decay system (Lyons 
etaL, 2003). 

Some of the fungi reported to be involved in this process include Dreschslera halodes, 
Halospbaeria hamata, Pbaeospbaeria typbarum, Stagonospora sp., Buergenerula 
spartinae, Alternaria alternata, Epicoccum nigrum, Claviceps purpurea, Leptospbaeria 
obiones, Leptospbaeria pelagica, Pleospora pelagica, Pleospora vegans and Lulwortbia 
sp. (Gessner, 1977). More recently, ascomycetous fungi have been reported as the major 
secondary producers within standing decaying leaves of S. alterniflora L. throughout its 
geographical range (Newell et al., 2000; Newell, 2001). This has been further supported 
by work on the fungal internal transcribed spacer sequences of these species by Buchan 
et al. (2003). The four ascomycetes that appear to be most involved in this system are 
Pbaeospbaeria spartinicola Leuchtmann, Mycospbaerella sp. 2 (of Kohlmeyer & 
Kohlmeyer, 1979), Pbaeospbaeria halima Johnson and Buergenerula spartinae Kohlm. 
et Gessner (Newell, 2001). 

The fungal signal molecule ergosterol has also been used extensively to detail fungal 
biomass in Spartina alterniflora-dom'mated salt-marsh systems. Significant differences 
in ergosterol content of decaying leaves occurred, which were interpreted to indicate 
differences in nitrogen (but not phosphorus) availability (Newell, 2001). Nitrogen is 
thought to be a limiting nutrient for both Spartina alterniflora (Mendelssohn 6c 
Morris, 2000) and its associated fungal populations (Newell et al., 1996). Looking for 
seasonal differences, Newell's group also found that the ratio between fungal produc- 
tivity and mean ergosterol contents was significantly higher in winter/spring than in 
summer/autumn. 

The situation in open oceans differs greatly from that in coastal regions. Organic 
matter is rarely abundant, except where planktonic blooms occur, often resulting from 
nutrient upwelling. Low substrate availability generally leads to very low fungal bio- 
mass, of yeasts rather than the filamentous fungi characteristic of coastal ecosystems. 

SGM symposium 65 



338 N. Clipson, E. Landy and M. Otte 

There are at present no quantitative data linking primary production to fungal numbers 
in these systems. Beneath the photic zones in the open ocean, degraders presumably 
become more important as organic material sediments through the water column. It is 
known that bacterial numbers reduce with depth, with numbers at 5000 m being 
around 10 % of those at 75 m (Dring, 1992), together with much lower bacterial growth 
rates. It might be expected that, at some depth, particularly below the photic zone, 
micro-organisms and fungi are the most important organisms (relative to total bio- 
mass), growing on the detritus precipitating down from the ocean surface layers. At 
great depths, a surprising abundance of life can be found near 'smokers' (deep-ocean 
hydrothermal vents) and it is very likely that fungi form a significant component of 
these systems (Duhig et al., 1992; Burgath & Von Stackelberg, 1995; Nagahama et aL, 
2001). 



CONCLUSIONS 

In comparison with terrestrial environments, we have a rather restricted view of the role 
of fungi in marine ecosystems. This partially reflects the extent of marine environments 
on a global basis, together with their remoteness and difficulty for sampling. There is a 
better understanding of land-sea interfaces, particularly from the studies of Newell 
and co-workers in the eastern USA, although these studies need to be taken up for other 
salt marsh/mangrove systems in other parts of the world. Although there has been 
considerable effort to uncover fungal diversity, marine mycologists have been slow to 
apply molecular methods to explore diversity. Whereas these would be routine for 
marine bacteriologists, fungal molecular studies are few (e.g. Buchan et al., 2003). Our 
knowledge of fungal diversity in open ocean systems is even more restricted, but could 
more frequently be incorporated into programmes examining marine bacterial 
populations, particularly as the predominant fungi in these systems are yeasts. We 
urgently need to know more about fungi existing in benthic and deep-sea (hydro- 
thermal) zones. Although we know something about marine fungal diversity in general, 
little is known about their function except in salt marshes. We need to know to what 
extent fungi contribute to food webs and fluxes of organic matter in marine systems 
other than salt marshes, in order to give fuller views of these systems in general. 
Exploration of new marine environments and determining the role of fungi in these 
environments is the next challenge for marine mycology. 



REFERENCES 



Adam, P. (1990). Saltmarsh Ecology. Cambridge: Cambridge University Press. 
Ahammed, S. & Prema, P. (2002). Influence of media nutrients on synthesis of lignin 
peroxidase from Aspergillus sp. Appl Biochem Biotechnol 1 02/1 03, 327-336. 



SGM symposium 65 



Roles of fungi in marine and estuarine habitats 339 



Almagro, A., Prista, C, Benito, B. ; Loureiro-Dias, M. C. & Ramos, J. (2001). Cloning 

and expression of two genes coding for sodium pumps in the salt-tolerant yeast 

Debaryomyces hansenii. J Bacteriol 183,3251-3255. 
Alongi, D. M. (1987). The distribution and composition of deep-sea microbenthos in a 

bathyal region of the western Coral Sea. Deep Sea Res 34, 1 245-1 254. 
Armstrong, W., Wright, E. J., Lythe, S. & Gaynard, T. J. (1985). Plant zonation and 

the effects of the spring-neap tide cycle on soil aeration in a Humber salt marsh. 

J Ecol 73, 323-339. 
Bacic, M. K. & Yoch, D. C. (1998). In vivo characterization of dimethylsulfoniopropionate 

lyase in the fungus Fusarium lateritium. Appl Environ Microbiol 64, 1 06-1 1 1 . 
Ballard, R. D. (1977). Notes on a major oceanographic find (marine animals near hot-water 

vents at ocean bottom). Oceanus 20, 35—44. 
Bansal, P. K. & Mondal, A. K. (2000). Isolation and sequence of the HOG1 homologue 

from Debaryomyces hansenii by complementation of the hog 1 Delta strain of 

Saccharomyces cerevisiae. Yeast 16, 81-88. 
Barghoorn, E. S. & Under, D. H. (1944). Marine fungi: their taxonomy and biology. 

Farlowia I, 395-467. 
Beever, R. E. & Laracy, E. P. (1986). Osmotic adjustment in the filamentous fungus 

Aspergillus nidulans. J Bacteriol 1 68, 1 3 58-1 365 . 
Blomberg, A. & Adler, L. (1993). Tolerance of fungi to NaCI. In Stress Tolerance of Fungi, 

pp. 233-256. Edited by D. H. Jennings. New York: Marcel Dekker. 
Bohnert, H. J., Ayoubi, P., Borchert, C. & 23 other authors (2001). A genomics approach 

towards salt stress tolerance. Plant Physiol Biochem 39, 295-31 1 . 
Bongiorni, L. & Dini, F. (2002). Distribution and abundance of thraustochytrids in different 

Mediterranean coastal habitats. Aquat Microb Ecol 30, 49-56. 
Buchalo, A. S., Nevo, E., Wasser, S. P., Oren, A. & Molitoris, H. P. (1998). Fungal life in 

the extremely hypersaline water of the Dead Sea: first records. Proc Biol Sci 265, 

1461-1465. 
Buchan, A., Newell, S. Y, Butler, M., Biers, E. J., Hollibaugh, J. T. & Moran, M. A. 

(2003). Dynamics of bacterial and fungal communities on decaying salt marsh grass. 

Appl Environ Microbiol 69, 6676-6687. 
Bucher, V. V. C, Hyde, K. D., Pointing, S. B. & Reddy, C. A. (2004). Production of wood 

decay enzymes, loss of mass and lignin solubilization in wood by diverse marine 

fungi. Fungal Divers 15, 1-14. 
Burgath, K. P. & Von Stackelberg, U. (1995). Sulfide-impregnated volcanics and ferro- 

manganese incrustations from the southern Lau basin (Southwest Pacific). Mar 

Georesour Geotechnol 1 3, 263-308. 
Carlile, M. J., Watkinson, S. C. & Gooday, G. W. (2001). The Fungi, 2nd edn. San Diego: 

Academic Press. 
Carvalho, L. M., Cacador, I. & Martins-Loucao, M. (2001). Temporal and spatial variation 

of arbuscular mycorrhizas in salt marsh plants of the Tagus estuary (Portugal). 

Mycorrhiza 1 1 , 303-309. 
Clement, D. J., Stanley, M. S., O'Neil, J., Woodcock, N. A., Fincham, D. A., Clipson, 

N. J. W. & Hooley, P. (1999). Complementation cloning of salt tolerance determ- 
inants from the marine hyphomycete Dendryphiella salina in Aspergillus nidulans. 

Mycol Res 103, 1252-1258. 
Clipson, N. J. W. & Jennings, D. H. (1992). Dendryphiella salina and Debaryomyces 

hansenii: models for ecophysiological adaptation to salinity by fungi that grow in the 

sea. Can J Bot 70, 2097-2105. 



SGM symposium 65 



340 N. Clipson, E. Landy and M. Otte 



Clipson, N. J. W., Hajibagheri, H. A. & Jennings, D. H. (1990). X-ray microanalysis of the 

marine fungus Dendryphiella salina at different salinities. J Exp Bot 41 , 1 99-202. 
Clipson, N. J. W., Landy, E. T. & Otte, M. L. (2001). Fungi. In European Register of 

Marine Species: a Checklist of the Marine Species in Europe and a Bibliography 

of Identification Guides, Collection Patrimoines Naturels, vol. 50, pp. 1 5-19. Edited 

by M. J. Costello, C. S. Emblow & R. White. Paris: Publications Scientifiques 

duMNHN. 
Cooke, J. C, Butler, R. H. & Madole, G. (1993). Some observations on the vertical 

distribution of vesicular arbuscular mycorrhizae in roots of salt marsh grasses 

growing in saturated soils. Mycologia 85, 547-550. 
Cornick, J., Standwerth, A. & Fisher, P. J. (2005). A preliminary study of fungal endophyte 

diversity in a stable and declining bed of Spartina anglica Hubbard. Mycologist 19, 

24-29. 
Davis, D. J., Burlak, C. & Money, N. P. (2000). Osmotic pressure of fungal compatible 

osmolytes. Mycol Res 104, 800-804. 
Dring, M. J. (1992). The Biology of Marine Plants, 2nd edn. Cambridge: Cambridge 

University Press. 
Duhig, N. C, Davidson, G. J. & Stolz, J. (1992). Microbial involvement in the formation of 

Cambrian sea-floor silica-iron oxide deposits, Australia. Geology 20, 511-514. 
Edwards, J., Chamberlain, D., Brosnan, G., West, D., Stanley, M. S., Clipson, N. J. W. & 

Hooley, P. (1998). A comparative physiological and morphological study of 

Dendryphiella salina and D. arenaria in relation to adaptation to life in the sea. Mycol 

Res 102, 1198-1202. 
Eilers, H., Pernthaler, J., Glockner, F. O. & Amann, R. (2000). Culturability and in situ 

abundance of pelagic bacteria from the North Sea. Appl Environ Microbiol 66, 

3044-3051. 
Elmsley, J. (1 980). The phosphorus cycle. In The Handbook of Environmental Chemistry, vol. 

1 , part A, The Natural Environment and the Biogeochemical Cycles, pp. 1 47-1 67. 

Edited by O. Hutzinger. Berlin: Springer. 
Faison, B. D., Clark, T. M., Lewis, S. N., Ma, C. Y, Sharkey, D. M. & Woodward, C. A. 

(1991). Degradation of organic sulfur compounds by a coal-solubilizing fungus. 

Appl Biochem Biotechnol 28/29, 237-251 . 
Feeney, N., Curran, P. M. T. & O'Muircheartaigh, I. G. (1992). Biodeterioration of woods 

by marine fungi and Chaetomium globosum in response to an external nitrogen 

source. Int Biodeterior Biodegrad 29, 1 23-1 33. 
Flowers, T. J., Hajibagheri, M. A. & Clipson, N. J. W. (1986). Halophytes. Q Rev Biol 6, 

313-337. 
Freney, J. R., Ivanov, M. V. & Rodie, H. (1983). The sulphur cycle. In The Major Biogeo- 
chemical Cycles and their Interactions, pp. 46-50. Edited by B. Bolin & R. B. Cook. 

Chichester: Wiley. 
Gadanho, M., Almeida, J. M. & Sampaio, J. P. (2003). Assessment of yeast diversity in a 

marine environment in the south of Portugal by microsatellite-primed PCR. Antonie 

van Leeuwenhoek 84, 2 1 1-221 . 
Geiser, D. M., Taylor, J. W., Ritchie, K. B. & Smith, G. W. (1998). Cause of sea fan death 

in the West Indies. Nature 394, 1 37-1 38. 
Gessner, R. V. (1977). Seasonal occurrence and distribution of fungi associated with 

Spartina alterniflora from a Rhode Island estuary. Mycologia 69, 477-491 . 
Gessner, R. V. & Kohlmeyer, J. (1976). Geographical distribution and taxonomy of fungi 

from salt marsh Spartina. Can J Bot 54, 2023-2037. 



SGM symposium 65 



Roles of fungi in marine and estuarine habitats 341 



Gonzalez, M. C, Hanlin, R. T. & Ulloa, M. (2001). A checklist of higher marine fungi of 

Mexico. Mycotaxon 80, 241-253. 
Grasso, S., Bruni, V. & Maio, G. (1997). Marine fungi in Terra Nova Bay (Ross Sea, 

Antarctica). New Microbiol 20, 371-376. 
Gunde-Cimermann, N. f Zalarb, P., de Hoogc, S. & Plemenitasd, A. (2000). Hypersaline 

waters in salterns - natural ecological niches for halophilic black yeasts. FEMS 

Microbiol Ecol 32, 235-240. 
Hawksworth, D. L. (1991). The fungal dimension of biodiversity: magnitude, significance, 

and conservation. Mycol Res 95, 641-655. 
Hawksworth, D. L. (2001). The magnitude of fungal diversity: the 1-5 million species 

estimate revisited. Mycol Res 105, 1422-1432. 
Hernandez-Saavedra, N. Y. & Romero-Geraldo, R. (2001). Cloning and sequencing 

the genomic encoding region of copper-zinc superoxide dismutase enzyme from 

several marine strains of the genus Debaryomyces (Lodder & Kreger-van Rij). Yeast 

18, 1227-1238. 
Hooley, P., Fincham, D. A., Whitehead, M. P. & Clipson, N. J. W. (2003). Fungal 

osmotolerance. AdvAppI Microbiol 53, 1 77-2 1 1 . 
Hutzinger, O. (editor) (1980). The Handbook of Environmental Chemistry, vol. 1, part A, 

The Natural Environment and the Biogeochemical Cycles. Berlin: Springer. 
Hyde, K. D. & Jones, E. B. G. (1988). Marine mangrove fungi. Mar Ecol 9, 15-33. 
Jefferies, R. L, Davy, A. J. & Rudmik, T. (1979). The growth strategies of coastal 

halophytes. In Ecological Processes in Coastal Environments, pp. 243-268. Edited 

by R. L. Jefferies & A. J. Davy. Oxford: Blackwell. 
Jennings, D. H. (1989). Some perspectives on nitrogen and phosphorus metabolism in 

fungi. In Nitrogen, Phosphorus and Sulphur Utilization by Fungi, pp. 1-32. Edited 

by L. Boddy, R. Marchant & D. J. Read. Cambridge: Cambridge University Press. 
Jennings, D. H. & Burke, R. M. (1990). Compatible solutes -the mycological dimension 

and their role as physiological buffering agents. New Phytol 116, 277-283. 
Jones, E. B. G. (1995). Ultrastructure and taxonomy of the aquatic ascomycetous order 

Halosphaeriales. Can J Bot 73, S790-S801. 
Jones, E. B. G. & Alias, S. A. (1997). Biodiversity of mangrove fungi. In Diversity of 

Tropical Fungi, pp. 177-186. Edited by K. D. Hyde. Hong Kong: Hong Kong 

University Press. 
Jones, E. B. G. & Mitchell, J. I. (1996). Biodiversity of marine fungi. In Biodiversity 

International Biodiversity Seminar, pp. 31-42. Edited by A. Cimermann & N. Gunde- 
Cimermann. Ljublijana: National Institute of Chemistry and Slovenia National 

Commission for UNESCO. 
Kimura, H., Harada, K., Hara, K. & Tamaki, A. (2002). Enzymatic approach to fungal 

association with arthropod guts: a case study for the crustacean host, Nihonotrypaea 

harmandi, and its foregut fungus, Enteromyces callianassae. Mar Ecol 23, 1 57- 

183. 
Kirchner, M. (1995). Microbial colonization of copepod body surfaces and chitin 

degradation in the sea. Helgol Meeresunters 49, 2001-2012. 
Kohlmeyer, J. (1969). Ecological notes on fungi in mangrove forests. Trans Br Mycol Soc 53, 

237-250. 
Kohlmeyer, J. & Kohlmeyer, E. (1979). Marine Mycology: the Higher Fungi. New York: 

Academic Press. 
Kohlmeyer, J. & Volkmann-Kohlmeyer, B. (2003). Fungi from coral reefs: a commentary. 

Mycol Res 107, 386-387. 



SGM symposium 65 



342 N. Clipson, E. Landy and M. Otte 



Kritzman, G. (1973). Observations on the microorganisms in the Dead Sea. MSc thesis, 

Hebrew University of Jerusalem (in Hebrew). 
Lavin, A. M., Bryden, H. L. & Parilla, G. (2003). Mechanisms of heat, freshwater, oxygen 

and nutrient transports and budgets at 24-5° N in the subtropical North Atlantic. 

Deep Sea Res Part I Oceanogr Res Pap 50, 1 099-1 1 28. 
Lorenz, R. & Molitoris, H.-P. (1992). Combined influence of salinity and temperature 

(Phoma pattern) on growth of marine fungi. Can J Bot 70, 2111-2115. 
Lorenz, R. & Molitoris, H.-P. (1997). Cultivation of fungi under simulated deep sea 

conditions. Mycol Res 101, 1355-1365. 
Lyons, J. I., Newell, S. Y, Buchan, A. & Moran, M. A. (2003). Diversity of ascomycete 

laccase gene sequences in a southeastern US salt marsh. Microb Ecol 45, 270-281 . 
Martin, D. F. (1970). Marine Chemistry, vol. 2. New York: Marcel Dekker. 
Masson, D. (2002). Deep water renewal in the Strait of Georgia. Estuar Coast Shelf Sci 54, 

115-126. 
Mendelssohn, I. A. & Morris, J. T. (2000). Eco-physiological controls on the primary 

productivity of Spartina alterni flora. In Concepts and Controversies in Tidal 

Marsh Ecology, pp. 59-80. Edited by M. P. Weinstein & D. A. Kreeger. Dordrecht: 

Kluwer. 
Molitoris, H.-P. & Schaumann, K. (1986). Physiology of marine fungi: a screening 

programme for growth and enzyme production. In The Biology of Marine Fungi, 

pp. 35-48. Edited by S. T. Moss. Cambridge: Cambridge University Press. 
Money, N. P. (1997). Wishful thinking of turgor revisited: the mechanics of fungal growth. 

Fungal Genet Biol 21,173-187. 
Nagahama, T., Hamamoto, M., Nakase, T., Takami, H. & Horikoshi, K. (2001). 

Distribution and identification of red yeasts in deep-sea environments around the 

northwest Pacific Ocean. Antonie van Leeuwenhoek 80, 101-110. 
Newell, S. Y. (1993a). Decomposition of shoots of a saltmarsh grass methods and dynamics 

of microbial assemblages. Adv Microb Ecol 13, 301-326. 
Newell, S. Y (1993b). Membrane containing fungal mass and fungal specific growth rate 

in natural samples. In Handbook of Methods in Aquatic Microbial Ecology, 

pp. 579-586. Edited by P. F. Kemp, B. F Sherr, E. B. Sherr & J. J. Cole. Boca Raton, FL: 

Lewis Publishers. 
Newell, S. Y (1996). Established and potential impacts of eukaryotic mycelial decomposers 

in marine/terrestrial ecotones. J Exp Mar Biol Ecol 200, 1 87-206. 
Newell, S. Y (2001). Multiyear patterns of fungal biomass dynamics and productivity within 

naturally decaying smooth cordgrass shoots. Limnol Oceanogr 46, 573-583. 
Newell, S. Y, Porter, D. & Lingle, W. L. (1996). Lignocellulosis by ascomycetes (fungi) on a 

saltmarsh grass (smooth cordgrass). Microsc Res Tech 33, 32-46. 
Newell, S. Y, Blum, L. K. # Crawford, R. E., Dai, T. & Dionne, M. (2000). Autumnal 

biomass and potential productivity of salt marsh fungi from 29° to 43° north latitude 

along the United States Atlantic coast. Appl Environ Microbiol 66, 1 80-1 85. 
Nilsson, A. & Adler, L. (1990). Purification and characterisation of glycerol-3-phosphate 

dehydrogenase (NAD + ) in the salt-tolerant yeast Debaryomyces hansenii. Biochim 

BiophysActa 1034, 180-185. 
Nybakken, J. W. (1997). Marine Biology - an Ecological Approach. Boston: Addison- 

Wesley. 
Packham, J. R. & Willis, A. J. (1997). Ecology of Dunes, Salt Marsh and Shingle. London: 

Chapman & Hall. 
Paton, F. M. & Jennings, D. H. (1988). Effect of sodium and potassium chloride and polyols 



SGM symposium 65 



Roles of fungi in marine and estuarine habitats 343 



on malate and glucose 6-phosphate dehydrogenase from the marine fungus 

Dendryphiella salina. Trans Br Mycol Soc 91 , 205-2 1 5. 
Phae, C. G. & Shoda, M. (1991). A new fungus which degrades hydrogen sulfide, 

methanethiol, dimethyl sulfide and dimethyl disulfide. Biotechnol /.eft 13, 375-380. 
Pierrou, U. (1976). The global phosphorus cycle. In The Major Biogeochemical Cycles and 

their Interactions, pp. 46-50. Edited by B. Bolin & R. B. Cook. Chichester: Wiley. 
Porter, D. (1986). Mycoses of marine organisms: an overview of pathogenic fungi. In The 

Biology of Marine Fungi, pp. 141-1 54. Edited by S. T. Moss. Cambridge: Cambridge 

University Press. 
Pugh, G. J. F. & Jones, E. B. G. (1 986). Antarctic marine fungi: a preliminary account. In The 

Biology of Marine Fungi, pp. 323-330. Edited by S. T. Moss. Cambridge: Cambridge 

University Press. 
Richey, J. E. (1983). The phosphorus cycle. In The Major Biogeochemical Cycles and Their 

Interactions, pp. 51-56. Edited by B. Bolin & R. B. Cook. Chichester: Wiley. 
Rohrmann, S., Lorenz, R. & Molltoris, H. P. (1992). Use of natural and artificial seawater 

for the investigation of growth, fruit body production, and enzyme activities in 

marine fungi. Can J Bot 70, 2 1 06-2 1 1 0. 
Rosswall, T. (1983). The nitrogen cycle. In The Major Biogeochemical Cycles and Their 

Interactions, pp. 46-50. Edited by B. Bolin & R. B. Cook. Chichester: Wiley. 
Rozema, J., Arp, W., Van Diggelen, J., Van Esbroek, M., Broekman, R. & Punte, H. 

(1986). Occurrence and ecological significance of vesicular-arbuscular mycorrhiza in 

the salt marsh environment. Acta Bot Neerl 35, 457-467. 
Salomons, W. & Forstner, U. (1984). Metals in the Hydrocyde. Berlin: Springer. 
Schaumann, K., Mulach, W. & Molitorls, H.-P. (1986). Comparative studies on growth 

and exoenzyme production of different Lulworthia isolates. In The Biology of Marine 

Fungi, pp. 49-60. Edited by S. T Moss. Cambridge: Cambridge University Press. 
Schmlt, J. P. & Shearer, C. A. (2003). A checklist of mangrove-associated fungi, their 

geographical distribution and known host plants. Mycotaxon 85, 423-477. 
Sengupta, A. & Chaudhurl, S. (2002). Arbuscular mycorrhizal relations of mangrove plant 

community at the Ganges river estuary in India. Mycorrhiza 12, 169-174. 
Sguros, P. L. & Simms, J. (1963). Role of marine fungi in biochemistry of oceans. 2. Effect of 

glucose, inorganic nitrogen, and tris (hydroxy methyl) amino methane on growth and 

pH changes in synthetic media. Mycologia 55, 728-741 . 
Sieburth, J. M. (1979). Sea Microbes. Oxford: Oxford University Press. 
Slaughter, J. C. (1989). Sulphur compounds in fungi. In Nitrogen, Phosphorus and Sulphur 

Utilization by Fungi, pp. 91-105. Edited by L. Boddy, R. Marchant & D. J. Read. 

Cambridge: Cambridge University Press. 
Smith, G. W., Ives, L D., Nagelkerken, I. A. & Ritchie, K. B. (1996). Caribbean sea-fan 

mortalities. Nature 383, 487. 
Torzilli, A. P. (1982). Polysaccharidase production and cell wall degradation by several salt 

marsh fungi. Mycologia 74, 297-302. 
Torzilli, A. P. & Andrykovitch, G. (1986). Degradation of Spartina lignocellulose by 

individual and mixed cultures of salt-marsh fungi. Can J Bot 64, 221 1-2215. 
Van Uden, N. & Fell, J. W. (1968). Marine yeasts. In Advances in the Microbiology of the 

Sea, vol. 1 , pp. 1 67-202. Edited by M. R. Droop & E. J. F. Wood. New York: Academic 

Press. 
Van Wambeke, F., Christaki, U., Giannakourou, A., Moutin, T. & Souvemerzoglou, K. 

(2002). Longitudinal and vertical trends of bacterial limitation by phosphorus and 

carbon in the Mediterranean Sea. Microb Ecol 43, 1 19-133. 



SGM symposium 65 



344 N. Clipson, E. Landy and M. Otte 



Wainwright, M. (1989). Inorganic sulphur oxidation by fungi. In Nitrogen, Phosphorus and 

Sulphur Utilization by Fungi, pp. 73-90. Edited by L. Boddy, R. Marchant & D. J. Read. 

Cambridge: Cambridge University Press. 
Wangersky, P. J. (1980). Chemical oceanography. In The Handbook of Environment 

Chemistry, vol. 1 , part A, The Natural Environment and the Biogeochemical Cycles, 

pp. 51-68. Edited by O. Hutzinger. Berlin: Springer. 
Yale, J. & Bohnert, H. J. (2001). Transcript expression in Saccharomyces cerevisiae at high 

salinity. J Biol Chem 276, 1 5996-1 6007. 
Zande, J. M. (1999). An ascomycete commensal on the gills of Bathynerita naticoidea, the 

dominant gastropod at Gulf of Mexico hydrocarbon seeps. Invertebr Biol 118, 

57-62. 



SGM symposium 65 



Role of micro-organisms in 
karstification 

Philip C. Bennett 1 and Annette Summers Engel 2 

department of Geological Sciences, The University of Texas at Austin, Austin, TX78712, USA 

department of Geology and Geophysics, Louisiana State University, Baton Rouge, LA 70803, 
USA 



INTRODUCTION 

Whilst chemolithoautotrophic micro-organisms are found in nearly every environment 
on Earth, they are more abundant in dark habitats where competition by photo- 
synthetic organisms is eliminated. Caves, particularly, represent dark but accessible 
subsurface habitats where the importance of microbial chemolithoautotrophy to 
biogeochemical and geological processes can be examined directly At Lower Kane 
Cave, WY, USA, hydrogen sulfide-rich springs provide a rich energy source for chemo- 
lithoautotrophic micro-organisms, supporting a surprisingly complex consortium of 
micro-organisms, dominated by sulfur-oxidizing bacteria. Several evolutionary lineages 
within the class 'Epsilonproteobacteria' dominate the biovolume of subaqueous 
microbial mats, and these microbes support the cave ecosystem through chemo- 
lithoautotrophic carbon fixation. The anaerobic interior of the cave microbial mats is a 
habitat for anaerobic metabolic guilds, dominated by sulfate-reducing and -fermenting 
bacteria. Biological controls of speleogenesis had not been considered previously and it 
was found that cycling of carbon and sulfur through the different microbial groups 
directly affects sulfuric acid speleogenesis and accelerates limestone dissolution. This 
new recognition of the contribution of microbial processes to geological processes 
provides a better understanding of the causal factors for porosity development in 
sulfidic groundwater systems. 

Karst landscapes form where soluble carbonate rocks dissolve by chemical solution 
(karstification), resulting in numerous geomorphic features, including caves and 
subterranean-conduit drainage systems (e.g. White, 1988; Ford & Williams, 1989). This 

SGM symposium 65: Micro-organisms and Earth systems - advances in geomicrobiology. 
Editors G. M. Gadd, K. T. Semple & H. M. Lappin-Scott. Cambridge University Press. ISBN 521 86222 1 ©SGM 2005 



346 P. C. Bennett and A. S. Engel 

has traditionally been viewed as an abiotic, chemical process that occurs near the water 
table, with biologically produced C0 2 as the principal reactive component. 



This model fits well with the classic view of terrestrial subsurface microbial environ- 
ments as a 'top-downward' model, where photosynthetically fixed, reduced carbon (C) 
supplies the subsurface community with substrates for cell mass and energy (Kinkle & 
Kane, 2000), because microbial processes occurring in the absence of light energy have 
generally been considered insufficient to support ecosystem-level processes. In the past, 
the subsurface community has been portrayed as a mixed community of heterotrophs, 
with only hydrogen-oxidizing methanogens typically recognized as a significant auto- 
trophic population (Stevens & McKinely, 1995; Chapelle et aL, 2002). However, the 
absence of light energy does not preclude life, as subsurface environments couple 
relatively constant temperatures and protection from surface conditions with an 
abundance of dissolved inorganic solutes and exposed mineral surfaces that can serve 
as sources of energy and nutrients (Maher & Stevenson, 1988; Stevens 6c McKinley, 
1995; Ghiorse, 1997; Bachofen et aL, 1998; Rogers et aL, 1999; Bennett et aL, 2000; 
Ben-Ari, 2002). Consequently, in subsurface environments, chemosynthetic micro- 
organisms are vitally important to global chemical and ecosystem processes because 
they gain cellular energy from the chemical oxidation of inorganic compounds and 
fixing inorganic C, and serve as catalysts for reactions that would otherwise not occur 
or would proceed slowly over geological time. 



Sulfur-based microbial ecosystems include communities that oxidize energy-rich, non- 
surface-derived, reduced S substrates [S(-II)], coupled to molecular-oxygen reduction, 
to fix C0 2 . In near-surface environments, such as caves and mines in shallow bedrock 
(<200 m depth), these communities can be examined directly and chemolithoauto- 
trophic S oxidizers form isolated but diverse and highly productive microbial-mat 
communities that support higher trophic levels within the subsurface ecosystems (e.g. 
Sarbu et aL, 1996; Angert et aL, 1998; Hose et aL, 2000; Sarbu et aL, 2000; Engel et aL, 
2001). The ubiquity of reduced S compounds in anaerobic groundwater suggests 
that microbial communities that utilize reduced S are dispersed more widely in the 
terrestrial subsurface than is currently recognized. Equally important, S(-II) oxidation 
to sulfate generates excess protons that accelerate rock weathering, porosity develop- 
ment and mobilization of toxic metals, demonstrating a geological significance of 
terrestrial S-based microbial ecosystems beyond the descriptive characterization 
of novel populations. Understanding the dynamics of S-based subsurface communities 
and the coupling of the S and C cycles is a step toward linking microbial ecology to 
geological phenomena. 

SGM symposium 65 



Role of micro-organisms in karstification 347 

Karst aquifers 

Caves act as portals to subsurface karst environments, with the typical habitat charac- 
teristics being complete darkness, nearly constant air and water temperatures and 
usually oligotrophic nutrient conditions, due to a limited supply of allochthonous 
organic material (Kinkle & Kane, 2000; Poulson &c Lavoie, 2000). The classic model for 
karst development (speleogenesis) involves carbonic acid dissolution, usually at and 
rarely below the water table. More recently, sulfuric acid speleogenesis was proposed 
from work in Lower Kane Cave, WY, USA (Egemeier, 1981). Based on observations of 
H 2 S-bearing thermal springs, extensive gypsum mineral deposits and gypsum-replaced 
carbonate rock walls, Egemeier (1981) put forth the sulfuric acid speleogenesis model, 
which included volatilization of H 2 S from the sulfidic groundwater to the cave atmos- 
phere and H 2 S autoxidation to sulfuric acid on the moist cave walls: 

H 2 S+20 2 ^H 2 S0 4 (equation 1) 

where the acid then reacted with and replaced the carbonate with gypsum: 

H 2 S0 4 + CaC0 3 + H 2 -* CaS0 4 . 2H 2 + C0 2 (equation 2) 

Gypsum is dissolved easily into groundwater and the net result is mass removal and an 
increase in void volume. Cave formation due to sulfuric acid has now been recognized 
in several active sulfidic cave systems (e.g. Hose et al., 2000; Sarbu et al., 2000) and has 
been a process linked to the development of at least 10 % of carbonate caves worldwide 
(Palmer, 1991). 

Sulfur-based cave ecosystems 

Microbial communities colonizing sulfidic cave habitats have received recent attention 
due to their chemolithoautotrophic metabolism (e.g. Sarbu et al., 1996) and their 
geological impact due to acid production (Vlasceanu et al., 2000; Engel et al., 2001; 
Northup & Lavoie, 2001). Early studies of sulfuric acid speleogenesis (Egemeier, 1981) 
assumed that abiotic H 2 S autoxidation was the important process for cave formation, 
and did not consider microbial S(-II) oxidation. In some of the active systems, e.g. in 
the Movile Cave (Romania), the Frasassi Caves (Italy) and Parker Cave (USA), 
filamentous aerobic to microaerophilic S-oxidizing bacteria (SOB) dominate sub- 
aqueous microbial mats (Angert et al., 1998; Hose et al., 2000) and were found to fix 
inorganic C, providing energy to sustain complex cave ecosystems (Sarbu et al., 1996, 
2000). Movile Cave provides us with an extreme example of a highly evolved, terrestrial, 
chemolithoautotrophically based ecosystem (Sarbu et al., 1996), where 33 endemic, 
cave-adapted, invertebrate taxa have been identified (from 48 total taxa), including 
24 terrestrial and nine aquatic animal species. 

SGM symposium 65 



348 P. C. Bennett and A. S. Engel 

Culture-independent and -dependent approaches have been used to characterize 
the microbial communities, and 16S rRNA gene-based phylogenetic analyses of the 
microbial mats from Movile Cave, Parker Cave, Cueva de Villa Luz (Mexico), the 
Frasassi Caves and Cesspool Cave (USA) show diverse SOB, belonging to distinct 
lineages within different classes of the proteobacteria and including the genera 
Thiothrix, Thiobacillus, Thiomonas, Tbiomicrospira, Thiovulum and Achromatium 
(Sarbu et al., 1996; Angert et al., 1998; Hose et al., 2000; Vlasceanu et al., 2000; Engel 
et al., 2001). The predominant microbial groups in many of the investigated sulfidic 
caves were found to belong to novel lineages within the class 'Epsilonproteobacteria' 
(Angert et al., 1998; Engel et al., 2001, 2003) and, in addition to caves, closely related 
organisms have been described from many S-rich, oligotrophic natural habitats, 
including sulfidic springs, groundwater associated with oilfields, marine waters and 
sediments, deep-sea hydrothermal-vent sites and vent-associated metazoans (Engel 
et al., 2004b). Despite the unique S-based microbial diversity of these habitats, there 
are few detailed studies describing the occurrence of epsilonproteobacteria from 
terrestrial environments. Moreover, little is known about either the ecophysiology of 
most epsilonproteobacteria, as many assemblages have not been cultured; from the 
work that has been done, most of these bacteria cycle S as SOB (e.g. Takai et al., 2003). 
The SOB communities, including epsilonproteobacteria, occupy the redox boundary 
where reduced S mixes with aerobic water, and these microbes can take advantage of a 
potential energy gradient to produce sulfate and protons as the ultimate end products 
of their metabolism. 

Colonization of the sulfidic caves by non-SOB microbial groups has rarely been 
addressed. Heterotrophic and anaerobic micro-organisms related to the phylum 
i Bacteroidetes i have been reported from some of the caves from 16S rRNA gene-clone 
libraries (Angert et al., 1998; Engel et al., 2001), and dissimilatory sulfate-reducing 
bacteria (SRB) and fermenting bacteria have been characterized from molecular 
investigations of few sulfidic aquifers, including those associated with oilfields (Voor- 
douw et al., 1996; Ulrich et al., 1998). Based on geochemical data, SRB and SOB have 
been characterized from Yucatan cenotes open to phototrophic activity (Stoessell et al., 
1993). However, the importance of SRB for recycling S compounds by generating 
supplemental H ? S that SOB can use (e.g. Widdel & Bak, 1992) has only recently been 
addressed (Engel et al., 2004b). 

Coupled C and S metabolism 

Whilst the microbiology and biochemistry of an S-based system are complex, only a 
few aspects relevant to S(-II) oxidation in a non-photosynthesizing environment are 
reviewed here. SOB use reduced S compounds as electron donors (Madigan et al., 
1997): 

SGM symposium 65 



Role of micro-organisms in karstification 349 

H 2 S+20 2 -*S05" +2H + (-798 kj mol" 1 ) (equation 3) 

HS-+ V20 2 +H + - S°+H 2 (-209 kj mol" 1 ) (equation 4) 

S°+H 2 + 1Vz0 2 -> SOj-+2H + {-5S7 kj mol" 1 ) (equation 5) 

S 2 0|-+H 2 0+20 2 ^2SO^ +2H + (-823 kj mol- 1 ) (equation 6) 

These general reactions proceed both biotically and abiotically, with abiotic autoxi- 
dation of H 2 S proceeding spontaneously in aerobic aqueous systems (Millero et al., 
1987). Almost all non-photosynthesizing, chemolithoautotrophic SOB are obligate 
aerobes, requiring 7 as the electron acceptor for S(-II) oxidation (Ehrlich, 1996). The 
mechanisms of biological S oxidation are still being elucidated. For example, whilst 
thiobacilli oxidize sulfide to sulfate directly (i.e. via equation 3), many SOB initially 
form S° as an intermediate species (equation 4) that is then stored intracellularly and 
further oxidized (equation 5) during periods of limiting sulfide (Ehrlich, 1996). Others 
oxidize reduced-sulfur intermediates, such as thiosulfate (equation 6). Thiobacillus 
denitrificans (Justin & Kelly, 1978), Thermotbrix thiopara (Caldwell et al., 1976) and 
some epsilonproteobacteria (Moyer et al., 1995) couple S oxidation to nitrate reduction 
in dysoxic environments, whereas Thiobacillus ferrooxidans couples the anaerobic 
oxidation of S° to iron reduction (Corbett & Ingledew, 1987). Chemo-organotrophs in 
dysoxic marine environments conserve energy from S° conversion to sulfuric acid and 
H 2 S by electron transfer in the presence of a sulfide scavenger (e.g. Fe^ + ) without 1 
(sulfur disproportionation) (Thamdrup et al., 1993): 

4S°+4H 2 O^SOf+3H 2 S+2H + (equation 7) 

Anaerobic microbes are common in marine and freshwater habitats, from anoxic sedi- 
ments and bottom waters. Of the anaerobic microbial guilds, specifically SRB, many 
have not been studied in detail from groundwater and karst springs. SRB, obligate 
anaerobes that use molecular hydrogen to reduce sulfate to H ? S, are divided into two 
broad physiological subgroups: those that oxidize acetate and those that do not. 
Certain species of each subgroup are capable of chemolithoautotrophic growth with 
CO? as the sole C source. If sulfate concentrations are high, SRB oxidize fermentation 
by-products completely to C0 2 . In low-sulfate environments, however, SRB compete 
with methanogens for hydrogen and organic compounds (Widdel Sc Bak, 1992; 
Scholten^^/,,2002). 

MICROBIAL GEOCHEMISTRY METHODS 

An interdisciplinary approach was used to examine the diversity of cave microbial 
communities and their roles in rock modification and karstification. By combining 
detailed characterization of both the geochemical characteristics of the habitat (Engel 
et al., 2004a) and the composition and structure of the microbial community (Engel et 

SGM symposium 65 



350 P. C. Bennett and A. S. Engel 

al., 2003, 2004b), the link between microbial metabolism and geochemical interactions 
can be examined. Recent research on the S-based ecosystem in Lower Kane Cave reveals 
complex microbial communities that tightly cycle S and C compounds across redox 
boundaries, whilst generating acidity and accelerating rock weathering (Engel et al., 
2004a). 

Geochemical characterization 

The objective of the geochemical characterization was to define the microbial habitat 
and to quantify the changes in solute concentration or speciation that are indicators 
of biogeochemical interactions. Water samples were collected for complete analyses of 
dissolved constituents, isotopic ratios and geochemical parameters by using standard 
methods and methods developed specifically for sulfidic cave systems (Engel et al., 
2004a). Field parameters included pH, temperature, specific conductance and field 
alkalinity on a filtered sample. Dissolved oxygen (DO) is a critical habitat constraint 
and was surveyed by using multiple methods, including electrochemical, microelec- 
trode fluorescence quenching, colorimetric analyses and gas chromatography (GC). 
Dissolved Fe 2+ is a sensitive indicator of the presence of dissimilatory iron-reducing 
bacteria (Lovley & Phillips, 1988) and this was measured immediately in the field by the 
ferrozine method to prevent sample degradation. Total and dissolved metal concen- 
trations are used to characterize the geochemical consequences of the microbial activity 
and were determined from a filtered acid-preserved sample by inductively coupled 
plasma-mass spectrometry (ICP-MS). Dissolved anions, nutrients and organic acids 
were measured by ion chromatography (IC) and total inorganic and dissolved organic 
C (DOC) were measured by a carbon analyser. Stream water-dissolved solute speciation 
and activity, equilibrium gas partial pressure and saturation state with respect to 
mineral phases were calculated by using the geochemical speciation model phreeqc 
(Parkhurst & Appelo, 1999). Sediments and core were characterized by X-ray diffrac- 
tion (XRD) analysis of dried samples and total elemental composition by dissolution 
and ICP-MS analysis. Sediment and core samples were also examined by conventional 
and environmental scanning electron microscopy (CSEM and ESEM, respectively) for 
the presence of micro-organisms on mineral surfaces, mineral composition by energy- 
dispersive X-ray spectrometry (EDX) analysis and for evidence of mineral weathering. 

Microbial-community analysis 

Both culture-dependent and -independent methods were used to characterize the 
metabolically active microbes in Lower Kane Cave. Typically, only 1 % of environ- 
mental bacteria can be cultured (Leff et al., 1995) and standard culturing methods often 
introduce a selective bias toward micro-organisms that are able to grow quickly and to 
utilize the substrates provided in the medium (McDougald et al., 1998). However, 
culturing allows for quantification of metabolically active organisms in the samples 

SGM symposium 65 



Role of micro-organisms in karstification 351 

and can lead to species identification through phylogenetic analyses (Palleroni, 1997). 
Laboratory strain isolation is essential to quantify the potential geological significance 
of microbes from the different habitats and can be used to help target unknown 
environmental samples. In contrast, molecular methods allow for characterization of a 
microbial community that may have populations that are difficult, if not impossible, to 
cultivate (Head etal., 1998). 

Biomass determination. Biomass, both viable and non-viable, is a basic attribute for 
characterizing mineral-microbe interactions and for understanding the implications of 
the more detailed metabolic and phylogenetic analyses. Biomass of filtered samples was 
determined by 4,6-diamidino-2-phenylindole (DAPI) cell counts (Gough & Stahl, 
2003). This method is excellent for individual cells, but it is less reliable for filamentous 
organisms where the definition of a single cell is problematic. For filamentous mat 
samples, total organic C content of an acidified sample (to remove carbonate-mineral 
material) was analysed by using a carbon analyser and the total cell-count estimate was 
done by using the conversion factor of 350 fg C per cell (Bratbak & Dundas, 1984). This 
method offers a maximum biomass, as it assumes that all of the organic C is associated 
with biomass. Parallel determination of bacterial biomass can also be done by using 
phosphatidyl ester-linked fatty acid (PLFA) analysis (Vestal & White, 1989). Phospho- 
lipids are membrane lipids that are turned over rapidly during metabolism (White et al., 
1979). Consequently, phospholipids indicate viable microbial biomass at the moment 
that the microbial mats were extracted and are not subject to the positive artefact of the 
DAPI cell-count or C biomass-estimate methods. 

Molecular methods. Microbial-community composition and structure were assessed 
by using established methods and a full-cycle rRNA approach (Engel et al., 2003, 
2004b). Total micro bial-mat environmental DNA was extracted and near-full-length 
16S rRNA gene sequences (rDNA) were obtained by PCR amplification using either 
archaeal and bacterial universal primers or lineage-specific primers (e.g. Ausubel et al., 
1990; Lane, 1991; Engel et al., 2003). A TA cloning kit (Invitrogen) was used to facilitate 
PCR-product transformation and cloning procedures. Sequence inserts from clone 
plasmids were PCR-amplified with plasmid-specific primers, purified by using Sepha- 
dex columns and sequenced by using an automated ABI sequencer (e.g. ABI Prism 
377X Perkin Elmer sequencer at Brigham Young University, Provo, UT, USA). 16S 
rDNA clone libraries were constructed from samples throughout Lower Kane 
Cave and, to facilitate community-composition analyses, clone sequence inserts were 
screened by using restriction-fragment length-polymorphism (RFLP) analysis, where 
selected clones from each representative RFLP pattern were sequenced (Engel et al., 
2004b). Results were analysed by using several phylogenetic-analysis methods after 
closely related rRNA gene sequences obtained from the Ribosomal Database Project 

SGM symposium 65 



352 P. C. Bennett and A. S. Engel 

were aligned by using clustal_x (Thompson et al., 1997) and then adjusted manually 
based on conserved primary and secondary gene structure. Phylogenetic methods have 
included neighbour joining by phylip (Felsenstein, 1993), maximum likelihood, 
minimum evolution and maximum parsimony by paup* (Swofford, 2002) and Bayesian 
inference in MrBayes (Ronquist & Huelsenbeck, 2003). 

To identify and reliably quantify micro-organisms directly from the microbial mats, 16S 
rRNA-targeted oligonucleotide probes were used for fluorescence in situ hybridization 
(FISH) (Aim et al., 1996). In addition to general archaeal and bacterial probes, gene 
probes were designed to target two novel groups of epsilonproteobacteria found in 
Lower Kane Cave (Engel et al., 2003). 

Metabolic survey. Microbial groups were enriched in pre-reduced anaerobic media 
specific for different metabolic groups, including chemolithoautotrophic, lactate/ 
formate- or acetate-utilizing SRB and fermenting bacteria. Enumeration of each meta- 
bolic group was done by using the most probable number (MPN) estimates of tenfold 
serial dilutions of cultivable micro-organisms from each enrichment medium. All 
enrichments were incubated in the dark at room temperature in a Coy anaerobic 
chamber with N ? : H 2 mixed gases to maintain anaerobic conditions. Following enrich- 
ment, growth was monitored by measuring OD 590 , evolved gases produced by the 
mixed cultures (e.g. CH 4 for methanogens) were measured by GC and SRB were 
screened for by measuring H 9 S production by GC. 

In situ microcosms 

The microbial influence on speleogenesis was evaluated by using the in situ microcosm 
approach that has been used extensively to examine silicate colonization and weather- 
ing (Hiebert & Bennett, 1992; Roberts et al., 2004). Sterile and non-sterile field 
chambers (in situ microcosms) containing 0-5-1-0 cm 3 chips of Iceland Spar calcite 
(Wards Scientific) and native limestone were deployed in the cave to test whether micro- 
organisms or the bulk cave stream-water chemistry controlled carbonate dissolution. 
Microcosms were constructed from 2-5 x 5 cm PVC pipes with screw caps on both ends. 
Sterile microcosms had 0-1 |im PVDF hydrophilic filters on the end, whilst non-sterile 
microcosms had 0-5 mm polyethylene mesh on either end to allow for fluid flow and 
also for microbial colonization of the chips. Paired sterile and non-sterile microcosms 
were placed throughout the cave stream and within the microbial mats, and remained in 
the cave for 2 weeks to 9 months. A technique similar to the buried-slide technique was 
also used, such that thin, polished wafers (surface areas of 1-2 cmj of limestone and 
pieces of Iceland Spar were attached to glass slides and then surrounded by a 0-5 mm 
mesh sack. At each microcosm site, a mesh sack was also deployed. Chips and wafers 
were examined by ESEM, confocal laser-scanning microscopy (CLSM) for FISH and 

SGM symposium 65 



Role of micro-organisms in karstification 353 

laser ablation (LA)-ICP-MS. A parallel interaction between the habitat and the micro- 
bial communities is the control exerted by the geological matrix, and colonization and 
weathering patterns from the carbonate surfaces were compared to other mineral 
substrates, such as chert or insoluble residues specific to the habitat being studied. 
Results from other environments (Hiebert & Bennett, 1992; Rogers et aL, 1998; Bennett 
et aL, 2000; Roberts et aL, 2004) suggest that colonization of a particular mineral 
substratum is influenced directly by the chemical composition of that mineral. 

Unpreserved chips were examined by using a Philips XL30 ESEM; chamber conditions 
varied from 10 to 95 % relative humidity, using a Peltier cooling stage, and variable water 
pressure [0-9-64 torr (120-853 Pa), with corresponding accelerating voltages from 4 to 
20 kV]. For FISH, chips and wafers were fixed in two ways within 12 h of collection: (i) 
with 4 % paraformaldehyde for 3 h; or (ii) with 50 % ice-cold ethanol (Engel et aL, 2003). 
Fixed samples were air-dried and dehydrated by sequential washes with 50, 80 and 
100 % ethanol for 3 min prior to hybridization. Gene probes specific for Lower Kane 
Cave epsilonproteobacterial lineages, as well as probes targeting other microbial 
groups, were applied to the rock surfaces and surfaces were examined by using CLSM. 

The example of Lower Kane Cave 

S-oxidizing bacteria are an underappreciated geological weathering force, particularly 
in subsurface karst systems. In simple microbial systems, such as planktonic organisms 
in groundwater, the production of acidity can accelerate rock weathering, but in 
the complex, stratified microbial-mat communities that can attach to rock surfaces, the 
microbial influences on weathering can be more difficult to elucidate. Influences may 
involve local production of H 2 S or other S gases and transfer to the vapour phase, with 
the net effect of transferring protons (or dissolution capacity) from aqueous to sub- 
aerial environments. 

Lower Kane Cave is located in the Bighorn Basin near Lovell, WY, USA, adjacent to 
oilfields and localized sulfidic thermal and non-thermal springs (Egemeier, 1981). The 
cave is formed in the Little Sheep Mountain Anticline along the Bighorn River in the 
Madison Limestone and is ~350 m long. Thick, white, filamentous microbial mats, 
3-10 cm thick and interconnected with white, web-like films (Fig. 1), are associated 
with four sulfidic springs that discharge into the cave, and mats stretch for up to 20 m in 
the cave stream below the springs (Engel et aL, 2003, 2004a, b). The biomass of the 
filamentous mats was 10 12 cells ml -1 , based on the total organic C-estimate method. 

The cave's springs are all of the calcium/bicarbonate/sulfate water type. Although the 
cave is forming from sulfuric acid speleogenesis, the spring and stream pH is buffered to 
circumneutral by ongoing carbonate dissolution. The dissolved sulfide concentration of 

SGM symposium 65 



354 P. C. Bennett and A. S. Engel 




Fig. 1. Photograph of subaqueous microbial mats in Lower Kane Cave. Image width is 50 cm, water 
depth is 10 cm. 

incoming spring water was ~38 ^mol 1 (speciated as ~60 % HS~:40 % H 2 S, based on 
the pH of the springs, pK 7-04), with non-detectable DO. The concentration of DO and 
sulfide changed downstream, such that, at the end of the microbial mats, DO exceeded 
40 ^moll -1 and sulfide was non-detectable. DOC (including methane) in all of the 
waters was extremely low, at < 80 jimol l -1 . 

Microbial-mat diversity, characterized by 16S rDNA clone library construction and 
sequence analyses, revealed the presence of several different bacterial phyla. The 
surface of the mats, predominantly of white filament bundles, was oxygenated based on 
microelectrode DO profiles, whereas the grey mat interior was anoxic ~3 mm beneath 
the surface of the mat-water interface. The majority of the sequences retrieved from 
the white filaments belonged to the phylum Proteobacteria, specifically the classes 
'Epsilonproteobacteria' (68%), 'Gammaproteobacteria' (12-2%), 'Betaproteo- 
bacteria 9 (11-7 %) and 'Deltaproteobacteria' (0-8 %), as well as other bacterial phyla, 
including the class Acidobacteria (5-6%) and the BacteroideteslChlorobi groups 
(1-7 %). In comparison, ~50 % of the retrieved 16S rDNA clone sequences from the 
interior mats were related to groups of SRB affiliated to the class 'Deltaproteobacteria' 
and to unculturable members of the phylum Chloroflexi. Rarer clones belonged to the 
phyla 'Gammaproteobacteria\ Planctomycetes, BacteroideteslChlorobi, 'Beta- 
proteobacteria\ 'Epsilonproteobacteria', Verrucomicrobia, ' Alphaproteobacteria\ 
Spirocbaetes, Actinobacteria, Acidobacteria and different OP candidate groups. FISH 
of the microbial mats revealed that ~70 % of the total bacterial biovolume was 
dominated by filamentous epsilonproteobacteria and specifically dominated by one 
novel lineage, referred to as LKC group II (Engel et ai, 2003) (Fig. 2). The microbial 

SGM symposium 65 



Role of micro-organisms in karstification 355 






Fig. 2. FISH image of microbial-mat sample, (a) Overlapping probes; green, EUB338l-lllmix probes for 
all eubacteria; light blue, GAM42a for gammaproteobacteria (e.g. Thiothrix species). Bar, 20 urn. (b) 
EUB338l-lllmix-targeted filaments, being mostly epsilonproteobacteria belonging to the Lower Kane 
Cave group II evolutionary lineage (Engel etal., 2003). (c) GAM42a probe only. All images were taken 
by using CLSM. Photograph scale is the same in all images. 



SGM symposium 65 



356 P. C. Bennett and A. S. Engel 

mats from Lower Kane Cave represent the first non-marine natural system demon- 
strably driven by the activity of filamentous epsilonproteobacteria. 

The high concentrations of S(-II) in the cave springs provide a rich energy source for 
SOB, and the concentration of dissolved sulfide decreased rapidly through the micro- 
bial mats. Abiotic sulfide autoxidation is extremely slow in disaerobic water at pH ~7-4 
[autoxidation half-life was calculated to be > 800 h (Engel et al., 2004a)] and sulfide 
volatilization from the water to the cave atmosphere accounts for <8 % of the sulfide 
loss in the stream, based on gas-flux experiments (Engel et al., 2004a). With no other 
mechanism for S(-II) loss, there would be, for example, much higher sulfide concen- 
trations at the end of the microbial mats. The only other loss mechanism, microbial 
consumption under microaerophilic conditions, could cause the observed rapid loss 
in S(-II). Because epsilonproteobacterial filaments were found to dominate the micro- 
bial mats, these organisms are presumed to be SOB. Future culturing of these microbial 
groups will verify these observations. 

The epsilonproteobacteria influence mat redox chemistry and ecosystem function 
directly by colonizing the initially nutrient-poor habitat, providing an energy source 
through chemolithoautotrophic C fixation, forming a dense mat and consuming 2 . 
Chemolithoautotrophy in the cave serves as the base for the cave food web and bulk 
white mats had a mean <5 13 C value of -36 %o, demonstrating chemolithoautotrophic 
fractionation against 13 C from an inorganic C source in the cave-stream water of 
-8-9 %o (Engel et al., 2004b). Most SOB and the epsilonproteobacterial filaments form 
complex microbial mats and consumption of DO by the SOB creates an anaerobic 
habitat within the mat interior (Engel et al., 2004b). The MPN method was used to 
estimate the biomass of anaerobic metabolic guilds; up to 10 6 cells ml -1 were found, 
with SRB and fermenters being the dominant culturable groups; iron reducers, S° 
reducers and methanogens were in low abundance overall (<10 3 cells ml -1 ). Redox 
stratification of the mats spatially separates metabolic guilds, such as SOB from SRB, 
and nutrients are consequently cycled between the multiple ecosystem components, 
whilst also being advected downstream. Based on the full-cycle rRNA approach, 
microbial diversity is greater within the mat interior and downstream portions of the 
mat, as mat density and organic C availability increase due to chemolithoautotrophic 
input and heterotrophic C degradation. The mat possibly terminates when H 2 S is 
consumed and 2 is too high for SOB metabolic function. 

The dominant mechanism for S(-II) loss is subaqueous microbial oxidation and most 
SOB oxidize reduced S compounds completely to sulfate, with a substantial energy 
yield (e.g. equation 3). As a result of the energetic oxidation reactions, acidity is gener- 
ated in the form of sulfuric acid, which can attack the geological matrix supporting the 

SGM symposium 65 



Role of micro-organisms in karstification 357 




Fig. 3. ESEM image of a limestone fragment collected from the thick microbial-mat region. The 
surface is deeply corroded, consistent with chemical weathering. Smooth crystals are authigenic 
gypsum crystals. Filamentous microbial biofilm is barely visible covering the specimen under the 

imaging conditions used. Bar, 5 urn. 

microbial community. Although some SOB can be acidophiles, most are neutrophilic 
and colonization of carbonate surfaces or habitats may buffer excess acidity and 
maintain pH homeostasis. Consequently, observed deeply corroded native carbonate 
rocks in the cave stream (Fig. 3), with dissolution effects only on surfaces exposed to the 
stream water and filamentous microbial mats, can be attributed to the activity of SOB. 
Examination of the surfaces by CSEM and ESEM revealed a complex reacting environ- 
ment, with dissolving carbonate, secondary gypsum mineral deposits and a thick cover 
of predominantly filamentous microbes in an exopolysaccharide matrix. 

The results from the experimental in situ microcosm chips show that the filamentous 
sulfide-oxidizing bacteria colonize the calcite surface (Fig. 4a). Over time, the sulfide 
oxidation results in chemical corrosion of the surface, first producing shallow, etched 
trenches along the filament (Fig. 4b) and culminating in a broadly etched surface 
resembling the native limestone fragments (Fig. 5). 

FISH probes were applied to the experimental microcosm carbonate surfaces to 
determine the identities of the filaments colonizing the surfaces (Engel et aL, 2004a). 
Gene probes targeting the LKC group II (' Epsilonproteobacteria') , 'Gammaproteo- 
bacteria' (e.g. Tbiotbrix species) and all eubacteria were used. Positive hybridization 
signals were observed for filamentous organisms and most of the filaments on the 
carbonate surfaces hybridized simultaneously with the general eubacterial probe and 

SGM symposium 65 



358 P. C. Bennett and A. S. Engel 




Fig. 4. ESEM image of chips of calcite collected from in situ microcosms exposed to microbial 
colonization for 3 months, (a) Image of S-oxidizing Thiothrix sp. filament with abundant stored 
elemental sulfur, (b) Thiothrix sp. filament on calcite surface, with visible etching of the mineral 
along a filament trench forming in the mineral. Bars, 5 urn. 



the LKC group II probe. Exceptionally bright hybridization signals for each of the 
probes indicated high rRNA content and suggested that the microbes were active when 
the microcosms were retrieved. 



SGM symposium 65 



Role of micro-organisms in karstification 359 




Fig. 5. CSEM photomicrograph of calcite chip from in situ microcosm, showing dissolution of the 
mineral surface (lower left) in association with filamentous micro-organisms that formed a thick 
biofilm on the mineral surface. Bar, 20 urn. 

The rapid loss of sulfide from the cave stream, carbonate dissolution associated with 
microbial filaments and the dominance of epsilonproteobacteria on the experimental 
carbonate surfaces support the hypothesis that these organisms, as SOB, have a direct 
role in sulfuric acid speleogenesis in Lower Kane Cave. The epsilonproteobacteria 
generate sulfuric acid as a by-product of their metabolism and locally depress the pH at 
limestone surfaces, which subsequently focuses carbonate dissolution. The previous 
cave-formation model was based on H 7 S volatilization from the cave stream to the 
atmosphere, but negligible volatilization and abiotic autoxidation of S(-II) were found 
in the cave. Instead, S(-II) was consumed by subaqueous SOB and cave enlargement 
occurs via microbially enhanced dissolution of the cave floor. 

CONCLUSIONS 

The subsurface is a varied habitat for diverse microbial communities, as reactive rock 
surfaces and mineral-rich groundwater provide a variety of energy sources for micro- 
organisms, especially chemolithoautotrophs. Stream geochemistry and the spatial 
relationships of aerobic and anaerobic guilds allow for tight cycling of C and S, 
constrained by redox boundaries within the mat environment. In Lower Kane Cave, the 
bulk of the subaqueous microbial mats were dominated by novel evolutionary lineages 
of the class 'Epsilonproteobacteria', representing the first non-marine system driven by 
the activity of this group. Ecologically, the epsilonproteobacteria are chemolitho- 
autotrophic SOB and their genetic and metabolic diversity creates habitats within the 
microbial-mat interior for other microbial groups, such as SRB and other anaerobic 
metabolic guilds. 



SGM symposium 65 



360 P. C. Bennett and A. S. Engel 

Nearly all of the S(-II) coming into the Lower Kane Cave is consumed by SOB within 
the subaqueous microbial mats, and these bacteria drive subaqueous sulfuric acid 
speleogenesis by attachment to carbonate surfaces and generation of sulfuric 
acid, which focuses local carbonate undersaturation and dissolution. Prior to this work, 
sulfuric acid speleogenesis was considered to be an abiotic, subaerial process or to be 
limited to shallow groundwater depths because of oxygen requirements for abiotic 
autoxidation of S(-II). However, chemolithoautotrophy linked to S(-II) oxidation 
under microaerophilic conditions extends the phreatic depths to which porosity and 
conduit enlargement can occur. The recognition of the geomicrobiological contri- 
butions to subaqueous carbonate dissolution fundamentally changes the model for 
sulfuric acid speleogenesis and supplements how subsurface porosity may develop in 
karst aquifers. 

ACKNOWLEDGEMENTS 

We thank the US Bureau of Land Management for permission to access this field site. We are 
grateful to M. L. Porter, S. A. Engel, T. J. Dogwiler, K. Mabin, M. Edwards, R. Payn, J. Deans and 
H. H. Hobbs, III, for field assistance. Funding for this work was provided by the Life in Extreme 
Environments (LExEn) program of the US National Science Foundation (EAR-0085576), the 
National Speleological Society, the Geological Society of America and the Geology Foundation 
of the University of Texas. 

REFERENCES 

Aim, E. W., Oerther, D. B., Larsen, N., Stahl, D. A. & Raskin, L. (1996). The oligo- 
nucleotide probe database. Appl Environ Microbiol 62, 3557-3559. 
Angert, E. R., Northup, D. E., Reysenbach, A.-L, Peek, A. S., Goebel, B. M. & Pace, 

N. R. (1998). Molecular phylogenetic analysis of a bacterial community in Sulphur 

River, Parker Cave, Kentucky. Am Miner 83, 1 583-1 592. 
Ausubel, R, Brent, R., Kingston, R., Moore, D., Seidman, J., Smith, H. & Strujil, K. 

(1990). Current Protocols in Molecular Biology, pp. 1 .6.1-1 .6.2. New York: Greene 

Publishing Associates and Wiley-lnterscience. 
Bachofen, R., Ferloni, P. & Flynn, I. (1998). Microorganisms in the subsurface. Microbiol 

Res 153, 1-22. 
Ben-Ari, E. T. (2002). Microbiology and geology: solid marriage made on Earth. ASM News 

68, 13-18. 
Bennett, P. C, Hiebert, F. K. & Rogers, J. R. (2000). Microbial control of mineral- 

groundwater equilibria: macroscale to microscale. Hydrogeol J 8, 47-62. 
Bratbak, G. & Dundas, I. (1984). Bacterial dry matter content and biomass estimations. 

Appl Environ Microbiol 48, 755-757. 
Caldwell, D. E., Caldwell, S. J. & Laycock, J. P. (1976). Thermothrix thiopara gen. et sp. 

nov., a facultatively anaerobic facultative chemolithotroph living at neutral pH and 

high temperature. Can J Microbiol 22, 1 509-1 517. 
Chapelle, F. H., O'Neil, K., Bradley, P. M., Methe, B. A., Ciufo, S. A., Knobel, L L & 

SGM symposium 65 



Role of micro-organisms in karstification 361 



Lovley, D. R. (2002). A hydrogen-based subsurface microbial community dominated 

by methanogens. Nature 415, 312-315. 
Corbett, C. M. & Ingledew, W. J. (1987). Is Fe 3+/2+ cycling an intermediate in sulphur 

oxidation by Fe 2+ -grown Thiobacillus ferrooxidans? FEMS Microbiol Lett 41 , 1-6. 
Egemeier, S. J. (1981). Cavern development by thermal waters. Nat Speleol Soc Bull 43, 

31-51. 
Ehrlich, H. L. (1996). Geomicrobiology, 3rd edn. New York: Marcel Dekker. 
Engel, A. S., Porter, M. L, Kinkle, B. K. & Kane, T. C. (2001). Ecological assessment and 

geological significance of microbial communities from Cesspool Cave, Virginia. 

Geomicrobiol J 18, 259-274. 
Engel, A. S., Lee, N., Porter, M. L, Stern, L A., Bennett, P. C. & Wagner, M. (2003). 

Filamentous " Epsilonproteobacteria" dominate microbial mats from sulfidic cave 

springs. Appl Environ Microbiol 69, 5503-551 1 . 
Engel, A. S., Stern, L. A. & Bennett, P. C. (2004a). Microbial contributions to cave 

formation: new insights into sulfuric acid speleogenesis. Geology 32, 369-372. 
Engel, A. S., Porter, M. L, Stern, L. A., Quinlan, S. & Bennett, P. C. (2004b). Bacterial 

diversity and ecosystem function of filamentous microbial mats from aphotic (cave) 

sulfidic springs dominated by chemolithoautotrophic "Epsilonproteobacteria" . FEMS 

Microbiol Ecol 51, 31 -53. 
Felsenstein, J. (1993). phylip (phylogeny inference package), version 3.5c. Department of 

Genome Sciences, University of Washington, Seattle, USA. 
Ford, D. C. & Williams, P. W. (1989). Karst Geomorphology and Hydrology. London: Unwin 

Hyman. 
Ghiorse, W. C. (1997). Subterranean life. Science 275, 789-790. 
Gough, H. L. & Stahl, D. A. (2003). Optimization of direct cell counting in sediments. 

J Microbiol Methods 52, 39-56. 
Head, I. M., Saunders, J. R. & Pickup, R. W. (1998). Microbial evolution, diversity, and 

ecology: a decade of ribosomal RNA analysis of uncultivated microorganisms. Microb 

Ecol 35, 1-21. 
Hiebert, F. K. & Bennett, P. C. (1992). Microbial control of silicate weathering in organic- 
rich ground water. Science 258, 278-281 . 
Hose, L D., Palmer, A. N., Palmer, M. V., Northup, D. E., Boston, P. J. & DuChene, H. R. 

(2000). Microbiology and geochemistry in a hydrogen-sulphide-rich karst 

environment. Chem Geol 169, 399^-23. 
Justin, P. & Kelly, D. P. (1978). Growth kinetics of Thiobacillus denitrificans'm anaerobic and 

aerobic chemostat culture. J Gen Microbiol 1 07, 1 23-1 30. 
Kinkle, B. & Kane, T. C. (2000). Chemolithoautotrophic microorganisms and their potential 

role in subsurface environments. In Ecosystems of the World, vol. 30, pp. 309-318. 

Edited by H. Wilkens, D. C. Culver & W. F. Humphreys. Amsterdam: Elsevier. 
Lane, D. J. (1991). 16S/23S rRNA sequencing. In Nucleic Acid Technigues in Bacterial 

Systematics, pp. 1 1 5-175. Edited by E. Stackebrandt & M. Goodfellow. Chichester: 

Wiley. 
Leff, L G., Dana, J. R., McArthur, J. V. & Shimkets, L J. (1995). Comparison of methods 

of DNA extraction from stream sediments. Appl Environ Microbiol 61, 1 141-1 143. 
Lovley, D. R. & Phillips, E.J. P. (1988). Novel mode of microbial energy metabolism: organic 

carbon oxidation coupled to dissimilatory reduction of iron or manganese. Appl 

Environ Microbiol 54, 1 472-1 480. 
Madigan, M. T., Martinko, J. M. & Parker, J. (1997). Brock's Biology of Microorganisms, 

8th edn. Upper Saddle River, NJ: Prentice Hall. 



SGM symposium 65 



362 P. C. Bennett and A. S. Engel 



Maher, K. A. & Stevenson, D.J. (1988). Impact frustration of the origin of life. Nature 331, 

612-614. 
McDougald, D., Rice, S. A., Weichart, D. & Kjelleberg, S. (1998). Nonculturability: 

adaptation or debilitation? FEM5 Microbiol Ecol 25, 1-9. 
Millero, F. J., Hubinger, S., Fernandez, M. & Garnett, S. (1987). Oxidation of H 2 S in 

seawater as a function of temperature, pH, and ionic strength. Environ Sci Technol 

21,439-443. 
Moyer, C. L, Dobbs, F. C. & Karl, D. M. (1995). Phylogenetic diversity of the bacterial 

community from a microbial mat at an active, hydrothermal vent system, Loihi 

Seamount, Hawaii. Appl Environ Microbiol 61 , 1 555-1 562. 
Northup, D. E. & Lavoie, K. H. (2001). Geomicrobiology of caves: a review. Geomicrobiol J 

18, 199-222. 
Palleroni, N. J. (1997). Prokaryotic diversity and the importance of culturing. Antonie van 

Leeuwenhoek 72, 3-1 9. 
Palmer, A. N. (1991). Origin and morphology of limestone caves. Geol Soc Am Bull 103, 

1-21. 
Parkhurst, D. L & Appelo, C. A. J. (1999). Users guide to phreeqc (version 2) -a computer 

program for speciation, batch-reaction, one-dimensional transport, and inverse 

geochemical calculations. In Water- Resources Investigations Report 99-4259. 

Washington, DC: US Geological Survey. 
Poulson, T. L. & Lavoie, K. H. (2000). The trophic basis of subsurface ecosystems. In 

Ecosystems of the World, vol. 30, pp. 231-249. Edited by H. Wilkens, D. C. Culver & 

W. F. Humphreys. Amsterdam: Elsevier. 
Roberts, J. A., Bennett, P. C, Gonzalez, L. A., Macpherson, G. L. & Milliken, K. L. 

(2004). Microbial precipitation of dolomite in methanogenic groundwater. Geology 

32, 277-280. 
Rogers, J. R., Bennett, P. C. & Choi, W. J. (1998). Feldspars as a source of nutrients for 

microorganisms. Am Miner 83, 1 532-1 540. 
Rogers, J. R., Bennett, P. C. & Hiebert, F. K. (1999). Patterns of microbial colonization on 

silicates. In US Geological Survey Toxic Substances Hydrology Program: Proceedings 

of the Technical Meeting, pp. 237-242. Edited by D. W. Morganwalp & H. T. Buxton. 

Charleston, SC, USA, 8-12 March 1999. 
Ronquist, F. & Huelsenbeck, J. P. (2003). MrBayes 3: Bayesian phylogenetic inference 

under mixed models. Bioinformatics 19, 1 572-1 574. 
Sarbu, S. M., Kane, T. C. & Kinkle, B. K. (1996). A chemoautotrophically based cave 

ecosystem. Science 272, 1 953-1 955. 
Sarbu, S. M., Galdenzi, S., Menichetti, M. & Gentile, G. (2000). Geology and biology 

of Grotte di Frasassi (Frasassi Caves) in central Italy, an ecological multi-disciplinary 

study of a hypogenic underground karst system. In Ecosystems of the World, vol. 30, 

pp. 361-381 . Edited by H. Wilkens, D. C. Culver & W. S. Humphreys. Amsterdam: 

Elsevier. 
Scholten, J. C. M., van Bodegom, P. M., Vogelaar, J., van Ittersum, A., Hordijk, K., 

Roelofsen, W. & Stams, A. J. M. (2002). Effect of sulfate and nitrate on acetate 

conversion by anaerobic microorganisms in a freshwater sediment. FEMS Microbiol 

Ecol 42, 375-385. 
Stevens, T. O. & McKinley, J. P. (1995). Lithoautotrophic microbial ecosystems in deep 

basalt aquifers. Science 270, 450^55. 
Stoessell, R. K., Moore, Y. H. & Coke, J. G. (1993). The occurrence and effect of sulfate 



SGM symposium 65 



Role of micro-organisms in karstification 363 



reduction and sulfide oxidation on coastal limestone dissolution in Yucatan cenotes. 

Ground Water 31, 566-575. 
Swofford, D. L. (2002). paup*: Phylogenetic analysis using parsimony (*and other methods) 

(version 4). Sunderland, MA: Sinauer Associates. 
Takai, K., Inagaki, R, Nakagawa, S., Hirayama, H., Nunoura, T., Sako, Y, Nealson, K. H. 

& Horikoshi, K. (2003). Isolation and phylogenetic diversity of members of 

previously uncultivated £-Proteobacteria in deep-sea hydrothermal fields. FEM5 

Microbiol Lett 218, 167-174. 
Thamdrup, B., Finster, K., Hansen, J. W. & Bak, F. (1993). Bacterial disproportionation of 

elemental sulfur coupled to chemical reduction of iron or manganese. Appl Environ 

Microbiol 59, 101-108. 
Thompson, J. D., Gibson, T. J., Plewniak, F., Jeanmougin, F. & Higgins, D. G. (1997). The 

clustal_x windows interface: flexible strategies for multiple sequence alignment 

aided by quality analysis tools. Nucleic Acids Res 25, 4876-4882. 
Ulrich, G., Martino, D., Burger, K., Routh, J., Grossman, E. L, Ammerman, J. W. & 

Suflita, J. M. (1998). Sulfur cycling in the terrestrial subsurface: commensal 

interactions, spatial scales, and microbial heterogeneity. Microb Ecol 36, 141-1 51 . 
Vestal, J. R. & White, D. C. (1989). Lipid analysis in microbial ecology: quantitative 

approaches to the study of microbial communities. Bioscience 39, 535-541 . 
Vlasceanu, L, Sarbu, S. M., Engel, A. S. & Kinkle, B. K. (2000). Acidic cave-wall biofilms 

located in the Frasassi Gorge, Italy. Geomicrobiol J 17, 125-139. 
Voordouw, G., Armstrong, S. M., Reimer, M. F., Fouts, B., Telang, A. J., Shen, Y. & 

Gevertz, D. (1996). Characterization of 16S rRNA genes from oil field microbial 

communities indicates the presence of a variety of sulfate-reducing, fermentative, 

and sulfide-oxidizing bacteria. Appl Environ Microbiol 62, 1 623-1 629. 
White, W. B. (1988). Geomorphology and Hydrology of Karst Terrains. New York: Oxford 

University Press. 
White, D. C, Davis, W. M., Nickels, J. S., King, J. D. & Bobbie, R. J. (1979). Determination 

of the sedimentary microbial biomass by extractable lipid phosphate. Oecologia 40, 

51-62. 
Widdel, F. & Bak, F. (1992). Gram-negative mesophilic sulfate-reducing bacteria. In The 

Prokaryotes, pp. 3352-3378. Edited by A. Balows, H. Truper, M. Dworkin, W. Harder 

& K.-H. Schleifer. New York: Springer. 



SGM symposium 65 



INDEX 



References to tables/figures are shown in italics 



Acetate oxidation 191 
acetogenesis 111,189,241 
Acbarax 309 
Acbromatium 35—37 

A. oxaliferum 36 

A. uolutans 53 

calcite deposition 51,52,53 

genetic and ecological diversity 60—61, 62, 63 

RY5, RYKS, RY8 spp. 61 

sulfur cycle 46,50,51,348 
acid mine drainage (AMD) 12, 14, 23, 25, 26 
Acidobacteria 354 

acidolysis, by fungi 207, 208-211, 214, 215, 218 
Acineto b acte r sp. B0064 117,118 
Acremomium -like hyphomycete 215—216 
actinides, microbial reduction 286—291 
Actinobacteria 354 
Aeromonas 277 ', 279 

Ag 

fungal interactions 211, 215 

microbial reduction 282 

microbial resistance 115,273 
aggregation 

diffusion-limited (DLA) 142-143 

reaction-limited (RLA) 142-143 
Al, fungal interactions 212, 216 
' Alpbaproteobacteria' 354 
Alternaria alternata 337 
Alteromonas putrefaciens see Sbewanella 

one id en sis 
americium (Am), microbial reduction 289 
ammonia 

anaerobic oxidation 152,164 

marine cycling 250—251 
ammonium, in primary production 181 
anaerobic methanotrophs (ANMEs) 308, 310, 311, 

312,313,314,316 
Antarctic dry valleys 71—73 

location 72 

microbial activity in 73, 75—76 

organisms in 73, 74 

sources of resources 76—81 
AOM see methane (anaerobic oxidation) 
apatite 211 
archaeol 307 
archaeal 'strain 121' 277 
Artbrobacter 113,115,117,119 
As 

fungal interactions 215 



groundwater contamination 283—284 

microbial influences 22 

microbial reduction 164,283—286 

microbial resistance 273, 284 
Ascocratera manglicola 333 
ascomycetes 206, 326, 327, 337 
Aspergillus 218,220,330 

A. fumigatus 325 

A. nidulans 215 

A.niger 208,218 

A. sydowii 325 
P37 sp. 215 

Astrospbaeriella striatispora 333 
ATPases, P IB -type 115-122 
Aureobasidium pullulans 334 
autunite 212 
azo dyes 293-294 
Azotobacter vinelandii 120 

Bacillus 

B. arseniciselenatis 285 
B. selenitireducens 285 
B. sp. U26 123 
B.sp.V6 123-124 

B. subtilis 87-88, 89-90 
bacterial cell wall 85-104 

cell surface complexity 94—99 

cryo-transmission electron microscopy 86—93 

hydrophobicity studies 99—104 

metal— ion interaction and mineralization 
91-93 

potentiometric properties 93—99 
'Bacteroidetes' 348, 354 
basalt 205, 208, 234 
Basidiomycotina 327 
Beauveria caledonica 209, 210, 212-213 
Beggiatoa 37-38 

B.alba 45 

carbon cycle 55, 56 

nitrogen cycle 53, 54 

sulfur cycle 44,45,46,47,309 

within-genus diversity 57—59 
benthic communities 

organic matter content 1 78 

organic matter degradation 1 78, 179, 180, 
182-183, 188-191 

oxygen uptake 182 

respiration 186 
bentonite 141,217,215,219 



SGM symposium 65 



366 Index 



beta-imaging 5 
'Betaproteobacteria* 354 
biocatalysts 273,281,283-294 
biofilms 

existence, structure and function 22—23, 24 
formation 24 

in mycotransformation of minerals 205, 206 
intraterrestrial environment 239 
metabolic links to metals 23 
metal geochemistry interactions 11—28 
aqueous, fundamentals 12—18 
emerging technologies 19,27—28 
environmental aspects 23, 24, 25, 26 
important solid phases 13—15 
microbial functional metabolism 19—22 
role of redox status 18 
sorption reactions 15, 1 6, 17— 18 
biogeochemical cycling, in coastal zones 173—193 
functional groups 188—191 
organic matter degradation and recycling 

182-183 
physico-chemical differences in latitudinal 

regions 173—175 
primary production 174, 175—182, 183 
role of coastal wetlands 191—193 
temperature— substrate concentration 
interaction 183-188 
biogeochemical cycling, marine ecosystems 
fungal contribution 336—338 
role of fungi 330—336 
C,NandP 331-335 
sulfur 331,332,335-336 
biomagnetite 160—161 
biomarkers, lipid 133,306—308 
biomass, microbial 

in intraterrestrial environment 239 
marine, resource availability 261—265 
subsurface microbial mats 110, 351, 353, 356 
biomineralization 

accidental biominerals 156, 157—158 
bacterial cell wall interactions see bacterial 

cell wall 
biofilms 23—26 
biomagnetite 160—161 
eukaryotic 159-160 
prokaryotic 157-159, 160 
purposeful biominerals 158, 159—160 
see also biosilification, mycotransformation 
bio remediation 
fungal role 220 
metal reduction 273, 278, 279, 282, 287, 289, 

291 
metal resistance 112—114 
radioactive waste 280, 286-291 



biosensors, Hg 282 
biosignatures 165—166, 216 
biosilification 131—145 

chemistry of silica 133—138 
opal-A 137,138 

precipitation rate, calculation 136 
saturation state 135—136, 137 
cyanobacterial biomineralization pathways 

and colloid aggregation 141—144 
cyanobacterial surface properties and 
function 138-141 
biotite 208-209,216 
bioweathering, defined 203 

see also weathering 
Black Sea 

AOM studies 307,309-310,311,312 
redox zones 165,166 
borehole techniques 236—238 
Botrytis 218 

bromoethanesulfonate (BES) 313 
Buergenerula spartinae 337 
building stone, deterioration 208—211 
butyrate-oxidizing bacteria, identification 4 

C cycle 

biomineralization and 159 

fungal interactions 210 

giant sulfur bacteria 55—56 

in Antarctic dry valleys 76—81 

in coastal zones 187, 189, 192-193 

intraterrestrial environment 241—243 

life detection and definition 162 

marine, fungal role 331—335 

subsurface ecosystems 345, 346, 353, 356 
coupled C and S metabolism 348—349 

see also primary production 
Ca 

fungal interactions 209, 210-211, 212, 
213-214, 214-215, 216, 218, 219 
caesium trifluoroacetate density-gradient 

centrifugation 4 
calcite 

coccolithophores 158, 262—265 

fungi 214-215 

sulfur-oxidizing bacteria 35, 51, 52, 53, 
214-215, 357-358, 359 
calcium carbonate 210—211 
calcium oxalate 209, 212, 213-214, 215 
calcretes 210, 214 
Calotlmx 95-96 
Calyptogena 309 
carbonates 

metal geochemistry 13, 14 

precipitation 212,214—215 



SGM symposium 65 



Index 367 



carbon-concentrating mechanisms (CCMs) 

257-258, 259-261 
carbonic acid/carbonate dissolution 347, 352, 353, 

357, 359, 360 
carbonic anhydrases 257—258, 259, 261 
cave ecosytem see subsurface, cave ecosystem 
Cd 

fungal interactions 211, 214, 220 

in marine primary production 257, 258 

resistance of subsurface bacteria 115 
cell wall see bacterial cell wall 
chasmolithic organisms 203, 204, 210 
chemical gradients 165—166 
chemo-autotrophy, hydrogen-driven 111, 112 
chemolithoautotrophic metabolism 111—112, 345, 

346,347,349,352,356 
chemotropism 207 

Chilean coastal sediments 50, 54—55, 58 
Chilean continental margin 305, 307 
chlorite 208,216 
Cblorobi 354 
Chlorofiexi 354 
Cbrysiogenes arsenates 284 
Cirrenalia pygme 333 
Citrobacter 290 

Cladosporium cladosporioides 218 
Claviceps purpurea 337 
clay— fungal interactions 216—218,220 
Clostridium 287 

C. pasteurianum 278 
Co 

fungal interactions 214, 220 

in marine primary production 257, 259 

microbial reduction 280 
C0 2 

biogeochemical cycling in coastal zones 173, 
189, 192-193 

in marine primary production 247, 259, 263 

intra terrestrial environment 241 
coal-solubilizing fungus 336 

coastal wetlands, biogeochemical cycling 191—193 
coastal zones, defined 173 

see also biogeochemical cycling in coastal 
zones 
coccolithophores 52, 158, 159,262-265 
cold- seep environments 55—56 
cold water paradigm 183 
Comamonas 116,117,118 

sp. B0173 123 
community gene arrays (CGA) 122 
community structure, isotope labelling 1—8 
complexolysis, by fungi 207, 209, 219 
concrete biodeterioration 219 
contact angle measurement (CAM) 101-102, 103 



contact guidance 207 

contamination see bioremediation, horizontal gene 

transfer, radioactive waste 
core-drilling 236-238 
Corollospora maritima 333 
Cortinarius glaucopus 208 
Cr 

fungal interactions 220 
metal geochemistry 18,22 
microbial reduction 279—280 
resistance of subsurface bacteria 113, 114 
crassulacean acid metabolism (CAM) 257 
crocetane 307 

cryo- transmission electron microscopy (cryoTEM) 
bacterial cell wall 86-93 

freeze-substitution 86-88, 90-92 
frozen hydrated thin sections 88—91 
Gram-positive cell walls 87-88, 89-90 
metal— ion interaction and mineralization 
91-93 
cryptoendolithic organisms 203, 204, 206, 210 
Cryptoualsa balosarceicola 333 
Cu 

fungal interactions 211-212, 214, 216, 220 
resistance of subsurface bacteria 114, 115 
cyanobacteria 

gene transfer 121 
in biosilification 133 

biomineralization pathways and colloid 

aggregation 141—144 
sheath 139-140, 141-143 
surface properties and function 138-141 
marine primary productivity 249, 251, 
252-253 
cytochromes 55—56 
b 279,286 

c 279,282-283,286,287-288 
c-type 277,278,282-283 
d 279 



Dating methods 

groundwater 235 

metabolic abilities 161—166 
day length, coastal zones 174, 175, 176 
Dead Sea, marine fungi 328 
Debaryomyces bansenii 329 
defence compounds 262 
Deinococcus radiodurans 119, 120, 121, 287 
'Deltaproteobacteria' 354 
Dendrypbiella salina 328, 329 
denitrification 

energy yield 47-49 

giant sulfur bacteria 47-49, 53—55 



SGM symposium 65 



368 Index 



in coastal zones 189—190 
desert varnish 215—216 
Desulfobacter 191 
Desulfobulbus 308 
Desulfocapsa 308 
Desulfococcus 308,309,310 
Desulforbopalus 308 
Desulfuromonas 191 
Desulfosarana 308,309,310 
Desulfotalea/Desulforhopalus group 190 
Desulfotomaculum 191 

D. auripigmentum 284 
Desulfovibrio 190 

D. desulfuricans 114, 278, 280, 285, 287, 288 

D. vulgaris 287 
deuteromycetes (Fungi imperfecti) 206 
Deuteromycotina 327 

diatoms, silicification by 262—265 

diazotrophy 250,253,255 

Digitatispora marina 333 

DNA-SIP 3 

DNRA, giant sulfur bacteria 47-49, 54-55 

dolocretes 210 

dolomite 210 

Drechslera balodes 337 

dry valleys see Antarctic dry valleys 

dunite 208 

EckernfordeBay 305,312 

edge effects 16, 17 

electricity generation 273 

electrobioreactor 281 

electron shuttles 277-278, 292, 314 

electrophoresis, denaturing-gradient gel (DGGE) 

4 
electrostatic interaction chromatography (ESIC) 

102 
endolithic organisms 203, 204, 206, 209, 210, 219 
energy sources 

intraterrestrial environment 241—242 

microbial 153-154, 155, 256, 157 

solar, in coastal zones 173—175 
energy yield 

dentrification 47^9 

sulfur oxidation 47—49 
Enterobacter 280 

E. cloacae 279,285 
Entromyces callianassae 334 

environmental aspects, biofilm metal geochemistry 

23,24,25,26 
environmental genomics 8 
environmental proteomics 8 
Epicoccum nigrum 337 
epilithic organisms 203, 204, 206, 209 



'Epsilonproteobacteria' 37, 43, 56, 345, 348, 349, 

352, 353, 354, 355, 356, 357-358, 359 
ergosterol 337 

Escherichia coli 215, 279, 285, 286, 290, 291, 292 
estuarine flushing 174—175 
Etbmodiscus 264 
eukaryotes 

anaerobic 154—155 

extremophily 166—167 

genetic diversity 152, 1 53 

metabolic diversity 152, 154-155, 159-160 

mineral formation 159—160 
euoendolithic organisms 203, 204 
exchangeable fraction 17—18 
exobiology 168 
exoenzymes, fungal 333,334 
extremophiles 166—167, 206 



anaerobic oxidation 1 64 
biofilms 20,21,22,24,25,26 
fungal interactions 212, 213, 215, 216, 218 
in marine primary productivity 252—255 
assimilation of inorganic C 259—261 
fate of biomass 261—265 
interactions with N, P, Zn and other resources 

258-265 
limitation, biogeochemical cycling 180-181, 

182 
microbial reduction 152, 164, 274-278 
degradation of xenobiotics 274 
diversity of organisms 275—278 
in coastal zones 189, 191 
mechanisms 277—278 
subsurface bacteria 111 
redox status 18 
solid phase reactions 13, 14 
sorption reactions 15, 16, 17 
feedback, geochemical 21,22,27 
feldspar 208,209,218 
Ferrimonas 277 

FISH (fluorescence in situ hybridization) 308, 
354-356 
FISH-MAR 5,6,7 
fission products, microbial reduction 286—291 
flavodoxin 253 
food reserve concept 187 
formate hydrogenlyase 291,292 
forsterite 216 

freeze-substitution 86-88, 90-93 
freshwater sediments, giant sulfur bacteria 50, 59 
frozen hydrated thin sections 88—91 
functional gene array (FGA) 122 



SGM symposium 65 



Index 369 



fungi, in marine ecosystems 

contribution to cycling 336—338 
exoenzymes 333—334 
fungal adaptation 328—330 
fungal diversity 324—325 
in biogeochemical processes 330—336 
marine fungus, definition 325 
osmotic adaptation 328—330 
salt tolerance 328-330 
fungi in rock, mineral and soil transformations 
201-221 
as biosorbents 211—212,220 
environmental biotechnology 219—220 

bioremediation 220 
fungal— clay interactions 216—218, 220 
in terrestrial environment 205—206 
metal binding and accumulation 211—212 
mycogenic mineral formation 212—216 
carbonate precipitation 214—215 
oxalate precipitation 212, 213—214 
reduction or oxidation of metals/ 
metalloids 215—216 
processes influenced by minerals 205 
rock and mineral habitats 203—205 
tropic responses 207 
weathering processes 202—203, 206—208 
biochemical 207-208 
biomechanical 206—207 
clay and silicate 217—218 
concrete biodeterioration 219 
rock and building stone 208—211 
Fungi imperfecti 206 
Fusarium lateritium 336 

' 'Gammaproteobacteria' 6, 37, 38, 42, 43, 56, 60, 

354, 355, 357 
genetic diversity 152, 153, 240 

see also horizontal gene transfer 
genome, P content 256 
Geobacter 191,276, 277-278, 284, 288, 290, 291 

G. metallireducens 115, 276, 278, 282-283, 
287, 290 

G. sulfurreducens 291 

strain GS-15 see G. metallireducens 
Geobacteraceae 111 
Geo spirillum barnesii strain SES-3 see 

Sulfuro spirillum barnesii 
Geovibrio ferrireducens 282 
giant sulfur bacteria 35—64 

descriptions of genera 35—43 

energetic considerations of sulfur oxidation 
44-49 

energy and geochemical significance 43—56 

evolutionary and ecological diversity 56—61 



in Acbromatium 60-61, 62, 63 
metagenomics 61-64 
within-genus diversity 57—60 
geochemical significance 49—56 
carbon cycle 55—56 
nitrogen cycle 53—55 
sulfur cycle 49—53 
phylogenetic tree 3 7 
glass, deterioration 208 
gold (Au) reduction 281-282 
Gram-negative bacteria 
cell surface 101,102 
cell wall 94,95,99 

metal homeostasis genes 115, 118, 124 
see also cyanobacteria 
Gram-positive bacteria 

cell wall 87-88,90,91-93,99 
metal homeostasis genes 115, 119,120, 123, 124 
granite 209,218 
grazers, marine 261—262 
groundwater 235—236 
contamination 112 
As 283-284 
radioactive 286—287 
dating methods 235 
mixing processes 235—236 
organisms 239—241 
sampling techniques 236—238 
Gymnascella 328 
gypsum 216, 347, 357 

Halocypbina uillosa 333 

Halospbaeria bamata 337 

Halosphaeriales 326 

heavy metals 

interactions of subsurface bacteria 112—114 
subsurface contamination 109, 112—114 

bioremediation 112—114 
see also specific metals 

Heleboma 217 

HetC0 2 -MAR 7 

heterotrophic leaching 207 

heterotrophic metabolism, subsurface 110—112 

Hg 

biosensors 282 
fungal interactions 215 
microbial reduction 282—283 
microbial resistance 113, 273, 282—283 

homeoviscous model 184 

horizontal gene transfer (HGT) 

in life detection and definition 163—164 
metal homeostasis genes 114—124 
evolution 114-115 
mechanisms 114—115 



SGM symposium 65 



370 Index 



microarray 122-124 

P IB -typeATPases 115-122 
subsurface environment 114—124 
hornblendes 209 
hot-spring environment 

biosilification 131-133, 144-145 

substrate sequestration 184 
HRABT (high- resolution acid— base titrations) 

93-94, 95-96, 97-98 
Hyalodendron 217 

Hydrate Ridge 305, 307, 309, 310, 311, 312 
hydrogen 

intraterrestrial environment 241 
hydrogen-driven biosphere hypothesis 241 
hydrogenase 3 291, 292 
hydrophilicity, microbial surface 100, 103 
hydrophobic interaction chromatography (HIC) 

101, 102-103 
hydrophobicity, microbial surface 99—104 
hydroxy-archaeol 307 
Hymenoscypbus ericae 210, 217 
hyperthermophiles 238,290 
hypolithic organisms 203,204 
Hysterangium crassum 218 



Insolation, coastal zones 173-174, 176, 179-180 
interfacial reactions see sorption reactions 
intraterrestrial environment 233—243 

definition 234 

organisms 238—239 
activity 242-243 
energy sources 241—242 
species diversity 240—241 

range of biomass 239 

strategies for exploration 236—238 

variability 234-236 
invertebrates, in Antarctic dry valleys 73, 74 
ion concentrations, marine 323, 324 
ionic strength 15, 16, 17—18 
Iron Mountain ecosystem 62-63 
isotope-array approach 5 
isotopic labelling methods 1—8, 27—28 

DNA-SIP 3 

FISH-MAR 5,6,7 

HetC0 2 -MAR 7 

intraterrestrial environment 242 

life detection and definition 162 

RNA-SIP 3-5 
isotopic signals for soils 77, 78 



K 



fungal interactions 209 
kaolin 208 
kaolinite 217, 218 



karst landscape formation 345—346 
kinetic biosignatures 165—166 

Lac car ia laccata 210 

laccase-like multicopper oxidase 216 

Lactococcus lactis 115 

layered communities 165—166 

legacy model 76, 79 

Legionella pneumopbila 119, 120 

Leptospbaeria 

L. obiones 337 

L. pelagica 337 
Leucotbrix 41-42 
Licbenotbelia 215 
lichens 73,201,207,208,214,216 

marine 327 
life detection and definition 161—166 
lignin-degrading organisms 334 
lignocellulosic enzymes 337 
limestone 209, 210, 211, 214-215, 352-353, 

357-358 
Linocarpon bipolaris 333 
Lower Kane Cave 350,352,353-359 
Lulwortbia 333,337 

Magnetite 160-161,274,275,276 

magnetosomes 1 60, 161 

Magnetospirillitm magnetotacticum 160 

magnets, 'designer' 275 

mangroves 191-192, 323, 326, 327, 335 

marble 209 

marine ecosystems 

characteristics 321—324 

ion concentrations 323,324 

fungal adaptation 328—330 

fungal contribution to cycling 336—338 

fungal diversity 324—328 

fungi in biogeochemical processes 330—336 

interactions 321,322 

primary productivity 322—323, 336—338 
marine snow 248, 264 
mass spectrometry, secondary-ion 308 
mats, microbial 

methanotrophic 309—310 

subsurface ecosystems 345, 346, 347—348, 352, 
353-359 
biomass determination 351,353 

see also Beggiatoa, Tbioploca 
melanin 207,211,212 
membrane fluidity 184 
mercury-resistance operon 282 
mesophiles 184,186 
metabolic diversity 151—168 

extremophily 166—167 



SGM symposium 65 



Index 371 



kinetic biosignatures and layered communities 

165-166 
microbial diversity 152—157 
mineral formation 157—161 
relevance to exobiology 152, 168 
timing emergence of metabolic abilities 
161-166 
metabolism, microbial 
functional 19—22 
subsurface 110—112 
metagenomics, giant sulfur bacteria 61-64 
metal accumulation, by fungi 207, 209, 211—212, 

217, 220 
metal geochemistry 

aqueous, fundamentals 12—18 
important solid phases 13—15 
role of redox status 18 
sorption reactions 15, 16, 17—18 
bacterial cell wall interactions see bacterial 

cell wall 
biofilms see biofilms — metal geochemistry 

interactions 
overlap with microbiology 19—27 
biofilm metabolic links 23 
environmental aspects 23, 24, 25, 26 
microbial functional metabolism 19—22 
metal homeostasis genes, in subsurface microbial 

communities 114—124 
metalloids, reduction 283—286 
metal reduction see specific metals 
metal resistance, subsurface bacteria 112—114 
metal sequestration, fundamentals 12, 13—18, 25 
metatorbenite 211—212 
metazeunerite 211 
methane, aerobic oxidation 304 
methane, anaerobic oxidation (AOM) 152, 159, 
164,303-316 
consortium 306,307,308,312,313 
environmental regulation 310—312 
history 305-306 
laboratory growth 315 
mechanism 312—315 
micro-organisms 306—308 

ANME groups 308,309,310,311,312, 
313,314,316 
study environments 308—310 

Black Sea 307,309-310,311,312 
Hydrate Ridge 305,307,309,310,311,312 
methane metabolism, giant sulfur bacteria 56 
methane production 303—304 
Metbanococcoides 190 
methanogenesis 
AOM and 306 
in Antarctic dry valley soils 80—81 



in coastal zones 189, 191 

intraterrestrial 241 

subsurface bacteria 111—112 
methanogenic enzymic system 313, 314 
Metbanomicrobiales 190,308 
Metbanosarcinales 190,308 
Mg 

fungal interactions 208-209, 216, 218 
mica 216,218 
microarray 

community gene arrays (CGA) 122 

functional gene array (FGA) 122 

horizontal gene transfer (HGT) 122—124 

phylogenetic oligonucleotide array (POA) 122 
microautoradiography see FISH-MAR, HeC0 2 - 

MAR 
microbial communities, isotopic labelling 1—8 
'Microbulbiferflagellatus'' 120 
microcline 208-209 
Micrococcus 280 

'M. lactilyticus* 278 
microcolonial fungi 206, 209 
microelectrophoresis see zeta-potential analysis 
microfossils 85—86 

formation 145 

in life detection and definition 162—163 
mineralized component 14 
minerals 

influence on microbial processes 205 

mycotransformation 201—221 

biochemical deterioration 202, 207—211 
environmental biotechnology 219—220 
fungal— clay interactions 216—218 
mycogenic formation 212—216 

secondary 202—203 

see also biomineralization 
mitosporic fungi 326, 327 
Mn 

biofilms 24 

fungal interactions 214, 215—216 

in marine primary production 258 

microbial reduction 152, 164, 189, 191, 
274-275, 277-278 

redox status 18 

solid phase reactions 13, 14 

sorption reactions 16,17 
Mo, microbial reduction 280 
molybdate 313 
montmorillonite 217 
Mucor 218 
muscovite 208 
My cena galop vis 208 
Mycospbaerella 325 

sp.2 337 



SGM symposium 65 



372 Index 



N cycle 

giant sulfur bacteria 47-49, 53-55 
in Antarctic dry valleys 76, 77, 78-81 
in coastal zones 187, 188, 189-190, 191-192 
in marine primary productivity 249—252 
assimilation of inorganic C 259—261 
fate of biomass 261—265 
fungal role 331—335 
interactions with Fe, P, Zn and other 

resources 258—265 
recycled production 250—251 
Namibian coastal sediments 49—50 
natural organic matter (NOM) 

biogeochemical cycling in coastal zones 176, 

175,179,182-183,188 
in Antarctic dry valleys 73, 74, 75-76, 78-81 
in metal geochemistry 13, 15, 18, 25 
nematodes, in Antarctic dry valleys 73, 74 
nepheline 208 

neptunium (Np) 287,289-290 
Ni, fungal interactions 211,220 
nitrates 

ammonification see DNRA 
biogeochemical cycling in coastal zones 

181-182, 189-190, 192 
reduction see denitrification 
storage by giant sulfur bacteria 38, 39, 53—55, 
57-58 
nitrification, Antarctic dry valleys 78, 79 
nuclear magnetic resonance (NMR) 96, 97, 99 
nutrient fluxes, coastal zones 174—175, 176 
nutrient sources, coastal wetlands 191—192 

Oidiodendron mains 210 

olivine 208,216 

organic acids, fungal production 207—211, 214, 

215,218 
organic matter see natural organic matter 
osmotic adaptation by fungi 328—330 
outwelling, coastal wetlands 191 
oxalate precipitation 222,213—214 
oxalic acid 209,210,214,215,218 
oxidants, use by life forms 154 
oxygen concentration/flux, biofilms 20-21, 24 
oxyhydroxide minerals 13, 14, 16, 17, 18, 20, 22 

P cycle 

in coastal zones 187, 188, 192 
in marine primary productivity 255—257 
assimilation of inorganic C 259—261 
fate of biomass 261—265 
interactions with Fe, N, Zn and other 
resources 258—265 
marine, fungal role 331—335 



Pacific coastal sediments 49 

Paecilomyces 218,336 

palladium (Pd) 281 

palygorskite 217,218 

Paracoccus denitrificans 290 

parasitism, marine 248, 262 

PAR (photosynthetically active radiation) 247, 260 

Paxillus involutus 210,211 

Pb 

fungal interactions 209, 210, 211, 214, 220 

resistance of subsurface bacteria 113, 114, 
115,116 
pelagic primary production 176—179, 179—180, 181 
Pelobacter 191,276 
Penicillium 218,220,325,328,330 

P. expansum 208 

P. frequent ans 218 

P. simplicissimum 208 
perchlorate reduction 1 64 
periplasmic space 90,91,93 
pH 

effects on AOM 311 

in trace-metal sorption 17 

metal binding by fungi 211 
Pbaeospbaeria 

P. balima 325, 337 

P. spartinicola 325, 337 

Ptypbarum 333,337 
phosphite oxidation 164 
phosphorimaging technique 290 
photosynthesis 155, 156, 157, 176, 243 

marine primary production 247, 248, 251, 
259, 260 
phototrophic component 24, 25 
phylogenetic analysis 

microbial mats 348, 352 
phylogenetic oligonucleotide array (POA) 122 
phylogenetic tree 

life detection and definition 164—165 

microbial diversity 152, 253, 164—165 
phytoextraction 220 
phytoremediation 220 
pili, in electron transfer 278 
Piloderma 208,209 
Planctomycetes 354 
platinum-group metals (PDMs) 281 
Pleospora 

P. pelagica 333, 337 

P. vegans 333,337 
plutonium (Pu) 287,289-290 
poikilophilic organisms 216 
poikilotrophic organisms 205 
polar coastal zones 

benthic oxygen uptake 2 82 



SGM symposium 65 



Index 373 



biogeochemical cycling 173—193 

primary production rates 174, 175—182 

seasonal variation in characteristics 173—175, 
176-180 
polyols 329 
potentiometric properties of cell surfaces 93—99, 

99 
primary production 

in coastal zones 173, 174, 175-182, 183 

marine 

assimilation of inorganic C 259—261 

constraints 249-258 

fate of biomass 261—265 

Fe, N, P and Zn cycling 247-265 

interactions among resources 258—265 

marine ecosystems 322—323 

fungal contribution 336—338 

seasonal variation 176—180 
Procbloro coccus 251, 252, 253, 256 
prokaryotes 

extremophily 166—167 

genetic diversity 152, 1 53 

metabolic abilities 164 

metabolic diversity 152-154, 157-159, 160 

mineral formation 157—159, 160 
Proteobacteria 354 

metal homeostasis genes 113, 116, 117,118, 
119,120 
protonation see acidolysis 
Pseudomonas 118,279,290 

P aeruginosa 94,95,96-97,120 

P ambigua 279 

P. guillermondii 280 

P. isacbenkouii 278 

Pputida 279,287 

Pstutzeri 285 

P. vanadium reductans 278 
psychrophiles 184-185, 186 
psychro tolerance 184 
Pyrobaculum 

Paeropbilum 120,121,282 

Pislandicum 282,290 
Pyro coccus 

'P. abyss? 120,121 

Pfunosus 282 
pyromorphite 209,210 

Quartz 208 

Radioactive waste 219 

microbial degradation 280, 286—291 
radiotropism 219 
Kalstonia 

R.B0665 116 



R. metallidurans 115 
redox chemistry 157 

metabolic diversity 151,153—154 
redox ladder 242,243 
redoxolysis, by fungi 207, 211 
redox status 18 

Acbromatium communities 61 
redox zones, Black Sea 165,166 
refractory component 14 
remazol black B 293 
Rbizoctonia solani 217 
rhizoferrin 208 
Rbizopbila marina 333 
Rbizosolenia 264 
Rbodobacter spbaeroides 286, 290 
rhodochrosite 215 
Rbodoferax ferrireducens 111 
rhodopsin 155 
RNA, as biomarker 3—5 
RNA-SIP 3-5 
rock 

hard 234,239,242 

intraterrestrial environment 234—243 

microbial habitats 203—206 

mycotransformation of minerals 201—221 
weathering processes 202—203, 206—211 

sedimentary 234,239,242 
rosette formation 41—42, 57 
Rubisco 258,261 

Saccbaromyces cerevisiae 215, 329 
salinity 322, 323-324 

fungal adaptation 328—330 
Salmonella typbimurium 120 
salt marshes 191-192 

marine fungi 323, 326, 335, 336, 337 
salt-tolerant fungi 325, 328-330 
sandstone 209,211,218 
Sargasso Sea, metagenomic analysis 62, 63 
Scopulariopsis 218 

S. brevicaulis 208 
Scottnema lindsayae 73, 74 
scytonemin 139 
Se 

fungal interactions 215 

microbial reduction 285, 286 
sedimentary pools 12, 13—15 
seep sites, giant sulfur bacteria 56 
selenate reduction 1 64 
serpentine 208 
Sbewanella 

cell surface properties 96, 97, 98-99, 101-102, 
103 

metal homeostasis genes 114 



SGM symposium 65 



374 Index 



metal reduction 277 

S. alga BrY 97, 98-99, 102 

S. algae 282 

S. oneidensis 98-99, 101, 102, 275, 277, 278, 
279,287 

S. putrefaciens 99, 103, 157-158, 287, 288, 290 
see also S. oneidensis 

strain J18143 293 
Si cycle 

biomineralization 159 

fungal interactions 213 
siderite 215,274,275 
siderophores 207,208,252-253 
silica 

amorphous (defined) 133 

chemistry 133—138 
silicates 

clay— fungal interactions 216—218 

fungal deterioration 208 
silicification by diatoms 262—265 
silification see biosilification 
sinter formation see biosilification 
soil 

amelioration, fungi 220 

Antarctic dry valleys 73, 75—81 
microbial activity 73,75—76 
organisms in 73, 74, 206 
properties 73,74,75—76 
resources 76—81 
respiration 75—76 

clay— fungal interactions 216—217 

radioactive contamination 286—287 
sorption reactions in biofilms 12—13, 15, 16, 17—18 
Spartina alterniflora 325, 336-337 
spatial-subsidy model 79, 80, 81 
speleogenesis see subsurface, cave ecosystems 
Spbingomonas 118 

S.sp.F199 118 
Spirocbaetes 354 
spodumene 208 
Sr, fungal interactions 214 
Stagonospora 37 
Staphylococcus aureus 115 
Stenotropbomonas maltopbilia 115, 119, 120 
stress avoidance, fungal 207 
string-of-pearls morphology 42 
substrate utilization rate 186—187 
subsurface, cave ecosystem 345—360 

coupled C and S metabolism 348—349 

karst aquifers 347 

microbial-community analysis 350—352 
microbial mat diversity 354—356 

microbial geochemistry methods 349—359 
geochemical characterization 350 



in situ microcosms 352—353, 357—358, 

359 
Lower Kane Cave 353—359 

sulfur-based 347-348 
subsurface, deep terrestrial 109—110 

interactions of bacteria with heavy metals 
112-114 

metal homeostasis genes in microbial 
communities 114—124 

microbial biomass and diversity 110 

microbial metabolism in 110—112 
Suillus 

S. bovinus 210 

S. granulatus 211 

S. luteus 210 
sulfate 

in AOM 304-306, 307, 309, 310, 312-315, 316 

reduction 

coastal zones 178 
in coastal zones 189, 190 
subsurface bacteria 111 
sulfides, crystalline metal 161 
sulfidic minerals 

fundamental reactions 13, 14 

in biofilm formation 24 

weathering and oxidation 25, 26 
sulfites 

reduction, in coastal zones 191 
sulfur cycle 

in life detection and definition 162, 163 

marine, fungal role 331, 332, 335—336 

subsurface ecosystems 345—360 

coupled C and S metabolism 348—349 

see also giant sulfur bacteria 
Sulfuribydrogenibium subterraneum 112 
Sulfuro spirillum 

S. arsenopbilum 284 

S. barnesii 284, 285 

strain MIT-13 see S. arsenopbilum 
sulfur-oxidizing bacteria 

energy considerations 44-49 

giant sulfur bacteria 44—49 

subsurface ecosystems 347—349, 356—357, 358, 
359 
sulfur-reducing bacteria 

subsurface ecosystems 348, 349, 352, 354, 359 
sunlight, as energy source 155, 156, 157 
superoxide dismutases 257, 329 
Synecbococcus 251,252,253 
syntrophism 153 

Tailings 25,26 

Tc, microbial reduction 290—291 

Te, microbial reduction 215, 285—286 



SGM symposium 65 



Index 375 



Telepbora terrestris 210 
temperate coastal zones 

benthic oxygen uptake 1 82 

biogeochemical cycling 173—193 

primary production rates 174, 175—182 

seasonal variation in characteristics 173—175, 
176-180 
temperature 

effects on AOM 311-312 

intra terrestrial environment 238 
temperature— substrate concentration interaction 

183-188 
TEM (transmission electron microscopy) see 

cryo-transmission electron microscopy 
tepius 216 
textile dyes 293-294 
Tbalassiosira 

T. pseudonana 253 

T. weissfiogii 257—258 
Tbauera selenatis 285 
Tbermob acillus ferrooxidans 349 
thermophiles 184,186,238 
Tbermotoga maritima 282 
Tbermotbrix tbiopara 349 
thigmotropism 207 
Tbiobacillus 348,349 

T. denitrificans 45, 349 

T. ferrooxidans 26,282,290,349 

T. tbiooxidans 290 
Tbiomargarita 42^13, 45, 50, 57, 58, 59-60 

T. namibiensis 57, 59 
Tbiomicrospora 348 
Tbiomonas 348 
Tbioploca 38-39 

carbon cycle 55, 56 

nitrogen cycle 54 

short-cell morphotype (SCM) 58 

sulfur cycle 45,47,49,51 

T. araucae 58 

T. cbilieae 58 

within-genus diversity 57, 58 
Tbwtbrix 40-42, 47, 60, 348, 357-358 

T. defluuii 41 

T. eikelboomii 41 

T. unzii 60 
Tbiovulum 43,56,348 

T. ma jus 43 

T. minus 43 
thorium (Th), microbial reduction 289 
thraustochytrids 326, 327, 335 
Tokyo Bay sediments 55, 58 
toxic metals see specific metals 
Trapelia involuta 211-212 
Tricboderma 218 



Tricbodesmium 249,255 

Trichomycetes 327 

tropical coastal zones 

benthic oxygen uptake 1 82 
biogeochemical cycling 173—193 
primary production rates 1 74, 175—182 
seasonal variation in characteristics 173—175, 
176-180 

tropic responses, of fungi 219 

tunnelling, feldspar 209 

U 

binding by fungi 212 

fungal interactions 211—212 

microbial influences 22 

microbial reduction 287—289 

resistance of subsurface bacteria 114 
Ulocladium 328 
uncultured micro-organisms, isotopic labelling 

1-8 
underground ecosystems see intraterrestrial 

environment, subsurface 
underground laboratory 236, 238 
Ureaplasma parvum 119,120 

Vanadium (V), reduction 278 
Varicosporina ramulosa 333 
veil formation 43 
vermiculite 216 
Verrucomicrobia 354 
vivianite 274, 275 

Waste rock material, weathering 25, 26 
water temperature 

coastal zones 174, 175, 176, 179-180, 183-188 
temperature— substrate concentration 
interaction 183-188 
weathering processes 
carbonates 210—211 
role of fungi 

building stone 208-211 
clay and silicate 217—218 
concrete biodeterioration 219 
rock 202-203,206-211 
waste rock material 25, 26 
weddelite 212,213-214 
whewelite 212,213-214,215 
Wolinella succinogenes 285 
wood ash 211 

Xenobiotics, microbial degradation 273, 274, 
291-294 

Yeasts, marine 327-328, 329, 330, 333-334 



SGM symposium 65 



376 Index 



Zeta-potential analysis 97, 102 fate of biomass 261-265 

Zn interactions with N, P, Fe and other 

fungal interactions 211, 212-213, 214,218,220 resources 258-265 

in marine primary productivity 257-258 resistance of subsurface bacteria 114, 115 

assimilation of inorganic C 259-261 Zygosaccbaromyces rouxii 329 



SGM symposium 65