Skip to main content

Full text of "كتب في علم الميكروبيولوجي"

See other formats


-4wcmV 



i 

3 



Bacterial Invasion 
of Host Cells 



Edited by 
Richard |. Lament 




CI.A' * *•• 



CAMBRIDG 



www.cambr idgeorgft /ao-j 21 009542 



Bacterial Invasion of Host Cells 

This book concerns the intimate association between bacteria and host cells. 
Many bacterial pathogens are able to invade and survive within cells at mu- 
cosal membranes. Remarkably, the bacteria themselves orchestrate this pro- 
cess through the exploitation of host cellular signal transduction pathways. 
Intracellular invasion can lead to disruption of host tissue integrity and per- 
turbation of the immune system. An understanding of the molecular basis 
of bacterial invasion and of host cell adaptation to intracellular bacteria will 
provide fundamental insights into the pathophysiology of bacteria and the 
cell biology of the host. The book details specific examples of bacteria that are 
masters of manipulation of eukaryotic cell signaling and relates these events 
to the broader context of host-pathogen interaction. Written by experts in 
the field, this book will be of interest to researchers and graduate students in 
microbiology, immunology, and biochemistry, as well as molecular medicine 
and dentistry. 

richard j. lam o nt is Professor of Oral Biology in the College of Dentistry, 
University of Florida, and he works on the adhesion and invasion of oral 
pathogens, biofilm formation, and intercellular communication. 



Published titles 

1. Bacterial Adhesion to Host Tissues. Edited by Michael Wilson 0521801079 

2. Bacterial Evasion of Host Immune Responses. Edited by Brian Henderson and 
Petra Oyston 0521801737 

3. Dormancy and Low-Growth States in Microbial Disease. Edited by Anthony 
R.M. Coates 0521809401 

4. Susceptibility to Infectious Diseases. Edited by Richard Bellamy 0521815258 

Forthcoming titles in the series 

Mammalian Host Defence Peptides. Edited by Deirdre Devine and Robert 

Hancock 0521822203 
The Dynamic Bacterial Genome. Edited by Peter Mullany 0521821576 
Bacterial Protein Toxins. Edited by Alistair Lax 052 18209 IX 
The Influence of Bacterial Communities on Host Biology. Edited by Margaret 

McFall Ngai, Brian Henderson, and Edward Ruby 0521834651 
The Yeast Cell Cycle. Edited by Jeremy Hyams 0521835569 
Salmonella Infections. Edited by Pietro Mastroeni and Duncan Maskell 

0521835046 



Over the past decade, the rapid development of an array of techniques in 
the fields of cellular and molecular biology has transformed whole areas of 
research across the biological sciences. Microbiology has perhaps been influ- 
enced most of all. Our understanding of microbial diversity and evolutionary 
biology and of how pathogenic bacteria and viruses interact with their animal 
and plant hosts at the molecular level, for example, have been revolutionized. 
Perhaps the most exciting recent advance in microbiology has been the de- 
velopment of the interface discipline of Cellular Microbiology, a fusion of 
classic microbiology, microbial molecular biology, and eukaryotic cellular 
and molecular biology. Cellular Microbiology is revealing how pathogenic 
bacteria interact with host cells in what is turning out to be a complex evo- 
lutionary battle of competing gene products. Molecular and cellular biology 
>- are no longer discrete subject areas but vital tools and an integrated part 
O of current microbiological research. As part of this revolution in molecular 
O biology, the genomes of a growing number of pathogenic and model bac- 
CQ teria have been fully sequenced, with immense implications for our future 
rv understanding of microorganisms at the molecular level. 
— Advances in Molecular and Cellular Microbiology is a series edited by re- 

^~ searchers active in these exciting and rapidly expanding fields. Each volume 
-< will focus on a particular aspect of cellular or molecular microbiology and 
13 will provide an overview of the area; it will also examine current research. 
— i This series will enable graduate students and researchers to keep up with the 
LJ rapidly diversifying literature in current microbiological research. 



AM CM 



Q Series Editors 

"; Professor Brian Henderson 

•^ University College London 

<^j Professor Michael Wilson 

— i University College London 
O 

^ Professor Sir Anthony Coates 

= St. George's Hospital Medical School, London 

{j Professor Michael Curtis 

Z St. Bartholomew's and Royal London Hospital, London 



Advances in Molecular and Cellular Microbiology 5 



Bacterial Invasion 
of Host Cells 



EDITED BY 

Richard J. Lamont 

University of Florida 



"BTff 



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.cambridge.org 

Information on this title: www.cambridge.org/9780521809542 

© Cambridge University Press 2004 

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 

iSBN-13 978-0-511-18514-4 eBook (NetLibrary) 
isbn-io 0-511-18514-6 eBook (NetLibrary) 

ISBN-13 978-0-521-80954-2 hardback 
isbn-io 0-521-80954-1 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 ix 

Preface xiii 

1 Invasion mechanisms of Salmonella 1 

Beth A. McCormick 

2 Shigella invasion 25 

Chihiro Sasakawa 

3 How Yersinia escapes the host: To Yop or not to Yop 59 

Geertrui Denecker and Guy R. Cornells 

4 Stealth warfare: The interactions of EPEC and EH EC with 

host cells 87 

Emma Allen-Vercoe and Rebekah DeVinney 

5 Molecular ecology and cell biology of Legionella 

pneum op hi la 1 2 3 

Maelle Molmeret, Dina M. Bitar, and YousefAhu Kwaik 

6 Listeria monocytogenes invasion and intracellular growth 161 

Kendy K.Y. Wong and Nancy E. Freitag 

7 N. gonorrhoeae: The varying mechanism of pathogenesis in 

males and females 203 

Jennifer L. Edwards, Hillery A. Harvey, and Michael A. Apicella 

8 Group A streptococcal invasion of host cells 239 

Harry S. Courtney and Andreas Podhielski 





H 

W 
H 

O 
U 



9 Invasion of oral epithelial cells by Actinobacillus 
actinomycetemcomitans 275 

Diane Hutchins Meyer, Joan E. Lippmann, and Paula Fives-Taylor 

10 I n vasi o n by Porphyron) onas gingiva I is 29 5 

Ozlem Yilmaz and Richard J. Lamont 

Index 3 1 5 



Contributors 




Emma Allen- Vercoe 

Department of Microbiology and Infectious Diseases 

University of Calgary 

Health Sciences Centre 

Calgary, Alberta 

Canada 

Michael A. Apicella 

Department of Microbiology 
The University of Iowa 
Iowa City, Iowa 52242 
USA 

Dina M. Bitar 

Department of Microbiology and Department of Medical Microbiology 

and Immunology 
Faculty of Medicine 
Al-Quds University 
Jerusalem, 19356 
Israel 

Guy R. Cornells 

Biozentrum 

70 Klingelbergstrasse 

CH 4056 Basel 

Switzerland 




CO 



Harry S. Courtney 

Research Service (151) 
Veterans Affairs Medical Center 
Memphis, Tennessee 38104 
USA 

Geertrui Denecker 

Biozentrum 

70 Klingelbergstrasse 

CH 4056 Basel 

Switzerland 

Rebekah DeVinney 

Department of Microbiology and Infectious Diseases 



g University of Calgary 

5 Health Sciences Centre 

g Calgary, Alberta 

o Canada 

u 

Jennifer L. Edwards 

Department of Microbiology 
The University of Iowa 
Iowa City, Iowa 52242 
USA 

Paula Fives-Taylor 

Department of Microbiology and Molecular Genetics 
University of Vermont 
Burlington, Vermont 05405 
USA 

Nancy E. Freitag 

Seattle Biomedical Research Institute and the Department of Pathobiology 
University of Washington 
Seattle, Washington 98109 
USA 

Hillery A. Harvey 

Department of Microbiology 
The University of Iowa 
Iowa City, Iowa 52242 
USA 



Yousef Abu Kwaik 

Department of Microbiology and Immunology 
University of Kentucky Chandler Medical Center 
Lexington, Kentucky 40536-0084 
USA 

Richard J. Lamont 

Department of Oral Biology 
University of Florida 
Gainesville, Florida 32610 
USA 

Joan E. Lippmann 

Department of Microbiology and Molecular Genetics 
University of Vermont 




n 

o 
Z 

H 

Burlington, Vermont 05405 2 

td 

C 
H 
O 

on 



USA 

Beth McCormick 

Department of Pediatric Gastroenterology and Nutrition 

Mucosal Immunology Laboratory 

Massachusetts General Hospital 

Charles town, Massachusetts 02129-4404 

USA 

Diane Hutchins Meyer 

Department of Microbiology and Molecular Genetics 
University of Vermont 
Burlington, Vermont 05405 
USA 

Maelle Molmeret 

Department of Microbiology and Immunology 
University of Kentucky College of Medicine 
Lexington, Kentucky 40536 
USA 

Andreas Podbielski 

Department of Medical Microbiology & Hospital Hygiene 
University Hospital Rostock 
D-18057 Rostock 
Germany 



Chihiro Sasakawa 

Institute of Medical Science, University of Tokyo, 4-6-1, Shirokanedai 

Minato-ku 

Tokyo 108-8639 

Japan 

Kendy K.Y. Wong 

Seattle Biomedical Research Institute and the Department of Pathobiology 
University of Washington 
Seattle, Washington 98109 
USA 

Ozlem Yilmaz 

Department of Pathobiology 
3 University of Washington 

£ Seattle, Washington 98195 

3 USA 

H 

O 

u 




Preface 



Few microbiologists are likely to forget the moment when the sheer scale 
and diversity of the microbial world became apparent to them. Similarly, it 
is a sobering thought that, in (or more accurately on) the human body, bac- 
teria outnumber human cells by at least 10 to 1. Fortunately, most of these 
bacteria behave as good guests should, and they are content to remain on the 
other side of the physical barriers that separate us from the outside world. 
It is inevitable that at such a large gathering, some guests, the pathogens, 
will misbehave, and worse, start a fight. For many years it was thought that 
the battle between host and pathogen at the mucosal membranes was fought 
at arm's length. The bacteria lobbed toxins and other noxious agents at the 
host and the host returned the favor with antibodies. Bacteria that ventured 
too close were rapidly dispatched by the professional killing machines, the 
phagocytic cells. Some bacteria, such as Mycobacterium tuberculosis, however, 
proved inconveniently recalcitrant to intracellular killing and could become 
permanent guests within macrophages. Nonetheless, despite the apprecia- 
tion that mitochondria originate from intracellular bacteria, the notion that 
mucosal pathogens could be intimately involved with nonprofessional phago- 
cytes is relatively new. 

We now know that a wide variety of organisms are capable of direct- 
ing their own entry into epithelial cells and other host nonphagocytic cells. 
These bacteria engage in a remarkably sophisticated molecular dialogue with 
host cells in order to manipulate signal transduction pathways and effectuate 
bacterial entry. In such an immunologically protected, nutrient-rich envi- 
ronment, bacteria can thrive. At first glance the burden of an intracellular 
bacterial load would appear to bode ill for the host cell. However, a long 
evolutionary relationship has resulted, in many cases, in a more balanced 
encounter between invasive organism and host, whereby the bacteria have 




adapted to minimize the degree of damage to the host. In other words, both 
bacteria and host try to make the best of their enforced cohabitation. 

The long-term consequences of such an arrangement can only be specu- 
lated on. Does the presence of intracellular bacteria help the host by providing 
a continuous low level stimulation of the immune system? Could intracellular 
bacteria aid in normal development of the epithelium as has been suggested 
for Bacteroides species in the intestine? Conversely, is it possible that sup- 
pression of apoptotic cell death by some organisms could be a cofactor in 
cancer development? While indulging speculation, it is interesting that, in 
what is called the Penrose- Hameroff orchestrated objective reduction model, 
microtubule stability can perform a cognitive role through quantum compu- 
tations within the neurons of the brain. If bacteria that can impinge upon 
microtubule stability were to gain access to the brain, the results could be 
g startling! 

In the following chapters the molecular bases of invasion, and the out- 
comes for host and bacterium, are discussed for variety of Gram-negative 
and Gram-positive species. Also included are organisms that subvert host 
cell signaling but do not appear to locate intracellularly in large numbers, 
or indeed use these pathways to prevent uptake into phagocytic cells. In a 
volume of this size, obviously we cannot be comprehensive; nonetheless, 
although some specific invasive species are not included, the themes that 
are developed are broadly applicable. In that vein, each of the authors has 
addressed the topic according to his or her own perspective, so individual 
chapters are uniquely informative. 




< 

w 
ft 



Bacterial Invasion of Host Cells 



CHAPTER 1 

Invasion mechanisms of Salmonella 

Beth A. McCormick 



Salmonella enterica serovar Typhimurium is a facultative intracellular pa- 
thogen that causes gastroenteritis in humans and a systemic disease similar 
to typhoid fever in mice. Following oral ingestion, bacteria colonize the in- 
testinal tract and then penetrate the lymphatic and blood circulation systems. 
Passage of eukaryotic organisms through the intestinal epithelium is thought 
to be initiated by bacterial invasion into M cells and enterocytes. The process 
of epithelial cell invasion can be studied experimentally because S. enterica 
serovar Typhimurium invades cultured epithelial cells in vitro. Many of the 
genes required for epithelial invasion have been found within eukaryotic 
pathogenicity island 1 (SPI-1), which is a contiguous 40-kb region at centro- 
some 63 of the chromosome. SPI-1 genes encode a bacterial type III secretion 
apparatus and several effectors, which contribute to pathogenesis through an 
interaction with eukaryotic proteins. The type III secretion apparatus is a mul- 
tiprotein complex that is thought to build a contiguous channel across both 
the bacterial and epithelial cell membranes, resulting in efficient translo- 
cation of bacterial effectors directly into the cytosol of epithelial cells. The 
secreted effectors are thought to interact with eukaryotic proteins to activate 
signal transduction pathways and rearrange the actin cytoskeleton, leading to 
membrane ruffling and engulfment of the bacterium. This chapter discusses 
the mechanism by which S. typhimurium enter into host cells. 

CLINICAL DESCRIPTION 

S. enterica, gram-negative bacteria of the family Enterobacteriaceae, cause 
a variety of diseases in humans and other animal hosts. Salmonella serovars 
fall into two general categories: those that cause enteric (typhoid) fever in 
humans (S. typhi and S. paratyphi), and those that do not (S. enteriditis and 



S. typhimurium) . Enteric fever is a systemic illness characterized by a high, 
sustained fever, abdominal pain, and weakness. Millions of cases of enteric 
fever are reported annually throughout the world, and, without antimicro- 
bial therapy, the mortality rate is 15% (Hook, 1990). Nontyphoidal serovars 
of S. typhimurium produce an acute gastroenteritis, characterized by intesti- 
nal pain and usually nonbloody diarrhea, which is a serious public health 
problem in developing countries (Hook, 1990). Approximately 40,000 cases 
of nontyphoidal Salmonella infection are reported annually in the United 
States, and the nontyphoidal Salmonella are responsible for more deaths in 
the United States than any other foodborne pathogen, with a mortality rate 
of approximately 1% (Hohmann, 2001). Because S. typhimurium is usually 
self-limiting in healthy adults, it is often not reported to public health offi- 
cials, and the total number of annual cases in the United States has been 
g estimated to be 1-3 million. However, in infants, in the elderly, in immuno- 

s compromised individuals, or in response to certain serovars, nontyphoidal 

8 Salmonella, such as S. choleraesuis, infection can also result in bacteremia and 

^ establishment of secondary focal infections (i.e., meningitis, septic arthritis, 




< 



x or pneumonia). Antibiotic treatment of nontyphoidal salmonellosis is not 



H 



£ recommended, because it may prolong the illness; this is most likely due to 

the killing of the normal gut microbiota, which exert a protective effect on the 
intestine (Hohmann, 2001). Bacteremia and its associated secondary infec- 
tions can be treated with antibiotics, such as ampicillin or cephalosporins, 
but this treatment has recently become a serious problem as a result of the 
rapidly growing number of Salmonella isolates that are multidrug resistant. 
Most infections with Salmonella (typhoidal and nontyphoidal) are con- 
tracted through contaminated food or water, and, although it is very rare, 
direct person-to-person transmission can occur. Studies with healthy vol- 
unteers have demonstrated that 10 6 -10 9 organisms are necessary to cause 
symptomatic illness (Hook, 1990). Most organisms are killed by the low pH of 
the stomach, but those that persist target the colon and small intestine (distal 
ileum) as their portal of entry into the host. The surviving bacteria then di- 
rect their internalization by the epithelial cells (both enterocytes and M cells) 
lining the intestine. S. typhimurium pass through the intestinal epithelial bar- 
rier to gain access to the lymphoid follicles, from which point the bacteria 
are eventually transported to the bloodstream and more distant sites of in- 
fection, such as the liver and spleen (Hook, 1990). Nontyphoidal Salmonella 
remain primarily at the level of the intestinal epithelium and submucosa (ex- 
cept during bacteremia) , where they elicit an acute inflammatory response 
that manifests itself as fever, diarrhea, and abdominal cramping (Hohmann, 
2001). Although in healthy adults the symptoms usually abate within a few 



days, it can take at least 1 month to fully clear all Salmonella from the gastroin- 
testinal tract, reflecting the fitness of these bacteria for their gastrointestinal 
niche (Hohmann, 2001). 

SALMONELLA ENTRY: SALMONELLA PATHOGENICITY ISLAND-1 

One of the first noteworthy in vitro activities observed regarding 
Salmonella-host cell interactions was the ability of this organism to induce 
its own uptake into epithelial cells, which are not normally phagocytic. This 
unusual phenotype, termed invasion, allowed for the identification and char- 
acterization of invasion genes associated with SPI-1. Galan and Curtiss (1989) 
were the first to characterize the S. typhimurium invasion locus, inv, which was 
identified by complementation of a noninvasive mutant of S. typhimurium. 
In particular, using an in vitro system of cultured epithelial cells, these in- 5 

vestigators discovered that highly virulent S. typhimurium strains carrying & 

inv mutations were defective for entry into but not attachment to Henle-407 * 

cells. Moreover, when administered perorally to Balb/c mice, the inv mutants w 

of S. typhimurium had higher 50% lethal doses than their wild-type parents > 

(Galan and Curtiss, 1989). £ 

Since this original observation, significant progress has been made to- £ 

ward understanding the molecular mechanisms that lead to Salmonella entry c* 

into cells. As subsequently discussed in detail, contributions by a variety of Eg 

o 
laboratories have established that the key ingredient of the machinery used g 

by Salmonella to gain access into nonphagocytic cells is the type III secretion £ 

system encoded at centrosome 63 of its chromosome. Type III secretion sys- 
tems are widely distributed among plant and animal pathogenic bacteria that 
share the property of engaging host cells in an intimate manner. Composed 
of more than 20 proteins, these systems are regarded as one of the most com- 
plex protein secretion systems discovered. Such complexity is largely due to 
their specialized function, which is not only to secrete proteins from the bac- 
terial cytoplasm but also to deliver them to the inside of the eukaryotic host 
cell. Adding to this level of complexity is the temporal and spatial restrictions 
that govern their activity. 



The regulation of Salmonella pathogenicity islands 

Salmonella have evolved spatially and temporally regulated systems, 
which secrete proteins that allow for the microorganism to invade the intesti- 
nal epithelium. In Salmonella, these delivery systems are encoded on regions 
of the bacterial chromosome termed pathogenicity islands (Darwin and Miller, 



< 



a 




o 



w 



CQ 



u 



o* 



c* 



cc 







03 

8- 



.O 
0) 



•9- 



* i 8- 



CQ 



C3) 

.03 




a 



* 



□ 









03 



to 

0) 

c 
o 

1_ 

Q. 
CO 

-C 
^O 

~co 


CO 

03 

o 
_o 

CO 

c 

03 



CO 



o 





■*— • •*—• 

CD 

i_ i— 

o o 



CO CO 



□ 



CO 

13 
■*-• 

03 

i_ 

03 
Q. 
Q. 
03 

C 

o 

"■4-" 



i— 

O 

CO 




Q. 

>^ 

•*— • 





CO 



o 


i_ 




( ) 


(/) 


-i—' 


*— • 


03 


{ 







_3 


r 


U) 


o 




1_ 


Li. 




h 


03 


n 


C 


o 


u 




-1— > 


03 


Q. 



o 

13 
i_ 
■*—• 

CO 

□ 



o 

CO 

c 

03 



PQ 




1999a; Hansen-Wester and Hensel, 2001). Many of the genes required for 
epithelial invasion have been found within SPI-1, which is a contiguous 40-kb 
region at centrosome 63 of the chromosome (Mills et al., 1995). SPI-1 genes 
encode a bacterial type III secretion apparatus and several effectors, which 
contribute to pathogenesis through an interaction with eukaryotic proteins 
(Fig. 1.1 A) . Salmonella invasion gene expression is known to be modulated by 
multiple environmental signals, including osmolarity, oxygen tension, pH, 
and stage of growth (Lee and Falkow, 1990). 

Specifically, the expression of SPI-1 genes appears to be regulated at 
several stages in a complex manner by regulators within SPI-1, including 
HilA and InvF, and those outside SPI-1, such as the two-component reg- 
ulators, the flagella associated genes, and the small DNA binding proteins 
(Fig. 1.1B). Salmonella does not constitutively express the virulence pheno- 
types associated with the SPI-1 type III secretion system (TTSS-1). In vitro 5 

inducing conditions that result in optimal expression of TTSS-1 include high & 

osmolarity, low oxygen tension, slightly basic pH, and the growth rate of the z 

bacteria (Lee and Falkow, 1990). The primary mechanism for controlling the w 

production of TTSS-1 factors in response to environmental and physiological > 

cues is by transcriptional regulation. The expression of genes encoding the K 

TTSS-1 apparatus and most of the effectors requires HilA, a transcription £ 

factor encoded on SPI-1. HilA, a member of the OmpR/ToxR family, directly c* 

binds to and activates the promoter of SPI-1 operons and functions as a cen- Eg 

o 
tral regulator of invasion gene expression (Lostroh et al., 2000). Osmolarity, g 



H 



oxygen, and pH coordinately affect the transcription of hilA, and changes in £ 

the level of HilA mediate the regulation of the SPI-1 TTSS-1 by the same 

Figure 1.1. (facing page). Type III secretion genes of SPI-1 and their regulation. (A) An 
overview of the type III secretion system encoded on SPI-1 includes subunits of a type III 
secretion apparatus, effectors secreted by the apparatus, factors required for their efficient 
translocation, and transcriptional regulators. The part of the island encoding a 
high-affinity iron transporter (sitBCDA) is not depicted. (B) Sequential upregulation by 
factors encoded on SPI-1 leads to expression of the type III secretion system. When in 
vitro environmental conditions are favorable for invasion gene expression (low pH, low 
oxygen tension, and high osmolarity), HilD derepresses the hilA promoter. The straight, 
solid arrows show that HilA protein directly activates the expression of structural genes 
such as the prgs and another regulatory gene, invF. HilA transcription initiated at Pi nv p 
results in a long mRNA that continues through sic A. Although not illustrated, InvF, as a 
complex with SicA, directly activates the expression of effectors such as SipB, SopE, and 
SopB/SigD. HilD also makes a small direct contribution to invF expression by slightly 
upregulating the activity of a promoter far upstream of the start of invF translation. 
(Adapted from P.C. Lostroh and C.A. Lee, Microh. Infect. 3: 1281-1291, 2001.) 



environmental conditions. Notably, no environmental condition has ever 
been documented that affects HilA-dependent TTSS-1 genes without also 
affecting hilA expression. 

SPI-1 encodes four additional transcriptional regulators besides HilA; 
these include InvF, HilD, SprB, and SprA/SirC. To date, genetic evidence 
implicates a cascade of transcriptional activation in which HilD, HilA, and 
InvF act in sequence to stimulate TTSS-1 gene expression in vitro (Fig. 1.1B). 
Initially, HilD, an AraC/XylS family member, binds directly to several sites 
within PhM and derepresses hilA expression. HilA then binds to conserved 
sequences located between — 54 and — 37 relative to both invF and prgH start 
sites. Activation of P prg H a ^d Pi nv p results in the transcription of prgHIJK- 
orgAB and InvFGEABCspaMNOPQRS. Therefore, HilA directly activates the 
expression of the structural type III secretion genes as well as the transcription 
g factor InvF. InvF, an AraC-like transcriptional regulator, promotes expres- 

s sion of HilA-activated effector genes by inducing their transcription from a 

8 second HilA-independent promoter (Darwin and Miller, 1999b). That is, InvF 

^ activates a promoter upstream of sicA, causing additional expression of sicA- 




< 



X sip BCD A. Furthermore, it has been demonstrated that two transcriptional 



H 



£ regulators of SPI-1, HilC and HilD, allow the expression of hilA by coun- 

teracting the action of an unknown repressor (Lucas and Lee, 2001). These 
complex regulations appear to ensure that invasion genes are appropriately 
expressed when Salmonella infects the host. 

SPI-2 also encodes for a protein secretion system similar to that encoded 
by SPI-1. Unlike SPI-1, which is found in all Salmonella lineages, SPI-2 is 
found only in S. enterica serovars and its acquisition most likely led to the 
divergence of S. enterica and S. bongori. SPI-2 is located at minute 30 of the 
S. typhimurium chromosome and has been implicated in the systemic phase 
of infection (Hensel et al., 1998; Shea et al., 1999). Expression of SPI-2 is 
induced within macrophages (Cirillo et al., 1998), and it appears to mediate 
bacterial survival within macrophages through evasion of NADPH oxidase- 
dependent killing (Vasquez-Torres et al., 2000) , and interference with cellular 
trafficking of Salmonella containing vacuoles (Uchiya et al., 1999). 

The type III secretion complex 

A TTSS is a complex organelle composed of more than 20 proteins (Fig. 
1.2). Subsets of these proteins are organized into a supramolecular structure, 
termed the needle complex, which spans the bacterial envelope (Kubori et al., 
1998). This structure resembles the basal body of flagella, suggesting an 
evolutionary relationship between these two organelles. Indeed, components 



InvE 

SpaM/InvI 

SpaO 

OrgO 

OrgB 



Outer membrane 



Periplasm 
Inner membrane 




PrgI 

PrgJ 

(needle) 



InvG InvH 
'(secretin pore) 



OP 



InvA, SpaPQRS 
(export apparatus) 



o 



V 




PrgK (top inner membrane ring) 
PrgH (bottom inner membrane ring) 



InvC/SpaL 
(ATPase) 



SpaN/InvJ 
(specificity regulator) 



Figure 1.2. Schematic representation of the SPI-1 type III secretion system. The secretion 
machinery is made up of approximately 20 proteins that span the inner and outer 
membranes and direct the secretion of proteins without the classical sec-dependent signal 
sequence. 

of the TTSS share amino acid similarity to flagellar proteins. The needle 
complex of the SPI-1 encoded TTSS has been characterized in some detail 
(Kubori et al., 2000; Zhou and Galan, 2001). It consists of a mul tiring base 
composed of the SPI-1 encoded proteins PrgHIJK. PrgH alone multimerizes 
into a tetrameric structure, but when complexed with PrgK it oligomerizes 
into ring-shaped structures that resemble the base of the needle complex 
of flagella. InvG has also been reported to form part of the base, and this is 
consistent with the observation that InvG forms a ring in the outer membrane 
in the presence of a helper lipoprotein, InvH. PrgI and PrgJ appear to form 
part of the bore of the needle. The core components, which comprise the 
needle-like complex, are highly conserved among different gram-negative 
pathogens, suggesting a common mode of operation. However, regulation 
of the secretion event is not well understood except that it requires energy in 
the form of ATP (Eichelberg et al., 1994). Further, unlike the sec-dependent 
pathway of bacterial protein secretion, type III secretion does not require a 
signal sequence on the protein to be exported, and in this respect it shares 
similarities with the bacterial flagellar export system. 



< 
> 

on 
O 

z 
g 

w 
n 

X 
> 

X 

in 

O 
►fl 

> 
H 

o 

H 

> 



The needle complex is constructed in an orderly manner (Kubori et al., 
1998, 2000; Zhou and Galan, 2001) . The proteins that make up the base of the 
complex are secreted through the inner membrane by the sec-mediated path- 
way. Once in the periplasm, the proteins form a complex that associates with 
a set of inner membrane proteins that share extensive sequence similarity 
to the components of the flagellar export apparatus. The resulting complex 
is restricted to the export of only the proteins that are necessary to make the 
needle structure. Once this foundation is made, the type III secretion appa- 
ratus becomes competent for the export of other type III secreted proteins, 
including those that are targeted to the inside of host cells. 

Although much progress has been made in characterizing the secretion 
apparatus itself, little is known about how the effector proteins are subse- 
quently translocated across the eukaryotic cell membrane. To date, three 
g proteins, SipB, SipC, and SipD, are required for the translocation of effector 

s proteins into the host cell, although the mechanisms by which SipB, SipC, 

8 and SipD exert their functions is not understood. Although these three pro- 

^ teins have been shown to be required for translocating effector proteins into 




< 



x the cytoplasm of the host cells, SipB, SipC, and SipD are not essential for 



H 



£ the secretion process (Collazo and Galan, 1997). At least 13 proteins that 

are delivered by the SPI-1 TTSS have been identified: AvrA, SipA, SipB, 
SipC, SipD, SlrP, SopA, SopB, SopD, SopE/E2, SptP, and SspHl (see ref- 
erences in Zhou and Galan, 2001). During the infection process, these pro- 
teins are presumably translocated into the cytosol of the host cell, where 
they engage host cell components to induce host cellular responses and pro- 
mote bacterial uptake. Although some of these effector proteins are encoded 
within SPI-1, several effector proteins are encoded outside this pathogenicity 
island. 



INTERNALIZATION OF SALMONELLA BY THE HOST EPITHELIUM 

Animal models of Salmonella infection 

A key feature of Salmonella pathogenesis is the ability of these bacteria 
to induce their own internalization by the normally nonphagocytic epithelial 
cells that line the intestine. Interactions between Salmonella and the intesti- 
nal epithelium were first described by Takeuchi, who orally infected guinea 
pigs with S. typhimurium (Takeuchi, 1967). From this early work it was de- 
termined that bacteria that closely contact the epithelial cells lining the intes- 
tine, primarily the ileum, elicit the local degeneration of filamentous actin in 
apical microvilli and the underlying terminal web. The morphology of other 



areas of the apical surface, either on the same cell or adjacent enterocytes, 
remains unaffeced. Subsequently, extruded membrane (described as mem- 
brane ruffles) surrounds the bacteria, resulting in their internalization into 
membrane-bound vacuoles. Once the bacteria are internalized, the overlying 
apical membrane regains its microvillar morphology, and despite these dras- 
tic changes to the apical cytoarchitecture and the presence of intracellular 
S. typhimurium, infected enterocytes remain healthy. Interestingly, although 
some bacteria become internalized by enterocytes, the majority remain in 
the intestinal lumen (Watson et al., 1995). Similar observations have been 
reported in other animals, including calves, pigs, and primates, all of which 
present with a diarrheal gastroenteritis in response to S. typhimurium and 
other related Salmonella strains (Bolton et al., 1999; Rout et al., 1974; Wallis 
and Galyov, 2000). 

< 
> 

Cell culture models of Salmonella infection S 

As a way to investigate in more detail the changes to the host intesti- w 

nal epithelium during early Salmonella-host cell interactions, a number of > 

cell culture models have been developed. Initial studies were performed with K 

epithelial cell lines that, when cultured on porous filter supports, establish £ 

electrically resistant epithelial monolayers with full apical-basolateral po- c* 

larity. Two polarized cell lines used in these early studies were the Madin § 

o 
Darby canine kidney cell line (M DCK) , derived from dog kidney distal tubule g 

cells, and the Caco-2 cell line, derived from human colonic epithelia. Polar- £ 

ized cells presented with nontyphoidal Salmonella on the apical cell surface 
exhibit similar features to enterocytes in the guinea pig model: microvilli be- 
come disassembled, and the resulting membrane extrusions internalize the 
bacteria (Finlay and Falkow, 1990; Finlay et al., 1988). Thus, S. typhimurium 
contacting the apical plasma membrane were observed to induce ruffling 
of the membrane at sites of bacterial-epithelial cell contact, providing the 
driving force for bacterial internalization. The ability of S. typhimurium to 
induce contact-dependent membrane ruffling as a means of gaining entry 
into the host cell suggests that the bacteria recapitulate a process resembling 
phagocytosis in these normally nonphagocytic cells. This process has been 
termed macropinocytosis to reflect its resemblance to pinocytosis, or fluid up- 
take into cells, but with the engulfment of much larger particles (Francis 
et al., 1993). Interestingly, although Salmonella initially interact with their 
animal hosts at the apical surface of the intestine, studies with the T84 cell 
polarizing cell line have revealed that they can invade from the basolateral cell 
surface at the same frequency as from the apical surface (Criss et al., 2003). 




u 



o 
u 
u 



< 



H 



However, the significance of this observation in the in vivo setting remains 
to be determined. 

S. typhimurium invasion is not restricted to epithelial cells. In vivo the bac- 
teria also invade macrophages, and in vitro they infect a variety of eukaryotic 
cells, except yeast and erythrocytes (Finlay et al., 1991). Subsequent invasion 
assays with HeLa cells and the Chinese hamster ovary (CHO) fibroblast cell 
line demonstrated that less polarized cells are more effectively infected by S. 
typhimurium, suggesting that the rigid cytoarchitecture of polarized epithelial 
cells is a hindrance to bacterial internalization in vivo. These studies implied 
that S. typhimurium utilize the same strategy to enter both polarized and 
nonpolarized cells, and the latter has gained widespread use for studies of 
the molecular regulation of S. typhimurium invasion. 



INVOLVEMENT OF THE HOST CELL CYTOSKELETON IN 
S BACTERIAL INTERNALIZATION 



^ The distinct morphologic changes occurring to the apical enterocyte 



x membrane upon binding of S. typhimurium suggested that host cell microfil- 



£ aments (composed of actin) or microtubules might be involved in the forma- 

tion of membrane ruffles. Finlay and Falkow were the first to report that treat- 
ment with cytochalasins, drugs that prevent F -actin polymerization, inhibits 
Salmonella invasion of multiple cultured cell lines. In contrast, microtubule- 
depolymerizing agents do not block bacterial internalization, suggesting that 
the actin cytoskeleton, but not the microtubual network, plays an active role 
in bacterial entry into host cells (Finlay and Falkow, 1988). Moreover, pre- 
treatment with cytochalasin D does not prevent bacterial attachment to the 
host cell surface, indicating that actin-dependent cytoskeletal rearrangements 
and membrane ruffling follow initial bacterial binding (Francis et al, 1993). 
Immunofluorescence microscopy later demonstrated that bacteria recruit fil- 
amentous actin to sites of active bacterial invasion. Confocal laser scanning 
microscopy revealed that several actin-binding proteins, including a-actinin, 
tropomysin, and talin, are recruited to the S. typhimurium-induced ruffles in 
cultured cells (Finlay et al., 1988). Remarkably, Salmonella do not disrupt the 
actin cytoarchitecture in other regions of the cell, including cortical actin bun- 
dles or stress fibers (Finlay et al., 1991). S. typhi also induce actin-dependent 
ruffling during invasion, suggesting that this aspect of bacterial invasion is 
conserved regardless of eventual disease outcome (Mills and Finlay, 1994). 
Because ruffle formation is essential to the invasion process, understanding 
the development of these structures is critical to understanding Salmonella 
pathogenesis as a whole. 



GTP 




GDP 




RadlGD^ 





Figure 1.3. Nucleotide cycling of monomelic GTPases: In the resting state, the 
monomeric GTPase (shown here as Racl) is in the GDP-bound, inactive conformation. 
Upon stimulation, a GEF catalyzes the release of GDP from the GTPase, followed by 
binding of GTP. This places the GTPase in the active conformation, where it can interact 
with effector proteins. To turn off the signal, a GAP enhances the GTPase's intrinsic 
hydrolysis rate, leading to GTPase inaction. S. typhimurium encodes two related GEFs for 
Rho GTPases, that is, SopE and SopE2, as well as one GAP for these GTPases, that is, 
SptP. 

INVOLVEMENT OF RHO GTPase IN S. TYPHIMURIUM INVASION 
OF NONPHAGOCYTIC CELLS 

Over the past 10 years, it has been demonstrated that the formation 
of actin-based cytoskeletal structures, which occurs in response to growth 
factors and other extracellular stimuli, is regulated by monomeric guano- 
sine triphosphatases (GTPases) of the Rho family (Hall, 1998; van Aelst and 
D'Souza-Schorey, 1997). Rho proteins are members of the Ras superfamily 
of monomeric GTPases, and, like all Ras superfamily members, they cycle 
between active (GTP-bound) and inactive (GDP-bound) conformations (Fig. 
1.3). Members of this family include RhoA-B-C-D-E-G, Racl-2, Cdc42, and 
TC10; however, RhoA, Racl, and Cdc42 have been the most extensively stud- 
ied. In vitro, both GTP binding and hydrolysis activities of the GTPases are 
extremely low; therefore, accessory factors are required to facilitate these pro- 
cesses. Guanine nucleotide exchange factors (GEFs) catalyze the release of 
GDP and binding of GTP, which activates the GTPase, while GTPase activat- 
ing proteins (GAPs) stimulate the GTP hydrolysis rate, thereby promoting 
their inactivation (Fig. 1.3). In fibroblasts, activation of RhoA promotes for- 
mation of stress fibers and focal contacts; Racl activation yields lamellipodia 



< 
> 

on 
O 

z 
g 

w 
n 

X 
> 

X 

in 

O 
►fl 

> 
H 

o 

H 

> 



and dorsal ruffles; and Cdc42 activation leads to the extension of filopodia 
(KozmaetaL, 1995; Nobes and Hall, 1995; Ridley and Hall, 1992). During cell 
spreading, Rho family members function sequentially, with initial activation 
of Cdc42 followed by Racl and RhoA (Nobes and Hall, 1995; Ridley and Hall, 
1992). In other actin-dependent processes, distinct subsets of Rho GTPases 
become activated, often in a cell-type specific manner. 

The involvement of Rho GTPases in S. typhimurium invasion was initially 
examined in nonpolarized cell lines of both epithelioid (HeLa and COS-1) and 
fibroblastic lineages. In these cells, Chen et al. (1996) demonstrated that inva- 
sion of Salmonella was primarily dependent on Cdc42. In this model, expres- 
sion of a point mutant of Cdc42 unable to bind GTP (which acts in a dominant 
inhibitory manner) prevented bacterial entry, whereas expression of domi- 
nant negative Racl partially inhibited internalization but not as effectively as 
g the Cdc42 mutant. The result of this study correlated with previous analysis 

s of Rho GTPases during Fc receptor-mediated phagocytosis in macrophages. 

8 Particularly, expression of dominant negative mutants of either Rac or Cdc42 , 

^ but not Rho, blocks phagocytosis of IgG-opsonized particles by having unique 




< 



x but complementary effects on localized actin polymerization at the plasma 



H 



£ membrane (Caron and Hall, 1998; Cox et al., 1997; Massol et al., 1998). 

Moreover, in their activated form, Cdc42 and Rac have been shown to in- 
duce actin polymerization through the activation of N-WASP and the Arp2/3 
complex. At present it is not known whether S. typhimurium direct their mor- 
phological changes in the actin cytoskeleton by using a similar activation 
strategy. 

Nonetheless, as a result of the unique structure of the enterocyte brush 
border, the cytoskeletal regulatory factors co-opted by Salmonella during in- 
vasion in polarized epithelia are different from those identified in studies 
with nonpolarized cells (Criss et al., 2001). Dominant negative Racl, but not 
Cdc42, significantly inhibited bacterial entry at the apical aspect of polarized 
cells. In this in vitro model of Salmonella - enterocyte interaction, the bacteria 
elicit actin reorganization and membrane ruffling at the apical surface in a 
manner that is morphologically indistinguishable from ruffling in nonpo- 
larized cell lines. However, during entry at the apical pole of epithelial cells, 
Salmonella encounter a complex, highly organized actin cytoskeleton unlike 
any other cell surface they invade. At the apical domain, polymerized actin 
is organized into rigid microvilli and the underlying terminal web, a cross- 
linked meshwork of actin filaments that attaches to intercellular junctional 
complexes (Fath et al., 1993). Accordingly, the ability of Salmonella to reorga- 
nize the apical plasma membrane and its underlying actin architecture may 
require the mobilization of a unique set of cellular regulatory factors. 




S. TYPHIMURIUM GENES THAT REGULATE EPITHELIAL 
CELL INVASION 

Several S. typhimurium gene products secreted via the SPI-1 encoded 
TTSS have been found to participate in the process of bacterial uptake by 
epithelial cells. These gene products fall into two categories: those that affect 
Rho GTPase activity, and those that directly affect host actin dynamics. 

SopE/SopE2 

SopE was first identified as a protein secreted by the SPI-1 TTSS of 
S. dublin, and it was subsequently found in S. typhimurium. Initial studies 
determined that deletion of sop E reduces invasiveness to 40%-60% of wild- 
type levels, presumably as a result of a reduced capacity of the pathogen to 

elicit plasma membrane ruffling, which can be rescued by complementation 3 

of the sopE locus (Wood et al., 1996; Hardt et al., 1998). Subsequently, S. s 

o 

typhimurium SopE was found to have GDP-GTP nucleotide exchange activity * 

on Rho family GTPases in vitro (i.e., it acts like a GEF; see Hardt et al., 1998; » 

Rudolph et al., 1999). Ectopic expression of SopE protein in mammalian cells > 

elicits membrane ruffling over the surface of the cell in a Racl- and Cdc42- £ 

dependent manner (Hardt et al., 1998). SopE is not encoded within SPI-1 £ 

but is instead found on a lysogenic bacteriophage, which is only possessed g 

by a subset of Salmonella spp. However, possession of the SopE phage does § 

not correlate with invasiveness or pathogenicity (Mirold et al., 1999). Since § 

this initial report, it was found that S. typhimurium possesses a homolog of £ 

SopE called SopE2, which has approximately 69% identity to SopE and is also 
secreted by the SPI-1 TTSS. A mutant strain deleted in SopE2 has reduced 
invasiveness relative to wild-type bacteria, but, unlike SopE, SopE2 is found 
in all pathogenic strains of Salmonella examined (Bakshi et al., 2001; Stender 
et al., 2000). These findings implicate SopE/E2 in the formation of the actin 
rearrangements necessary for membrane ruffling on the host cell surface 
and subsequent bacterial internalization. 

It is interesting to note that SopE can activate Cdc42 despite its lack of 
sequence similarity to Dbl-like proteins, the Rho-specific eukaryotic GEFs. 
Recent investigations focusing on the mechanism by which SopE mediates 
guanine nucleotide exchange have determined that SopE binds to and locks 
the switch I and switch II regions of Cdc42 in a conformation that promotes 
guanine nucleotide exchange (Buchwald et al., 2002). Although this confor- 
mation resembles that of Racl in a complex with the eukaryotic Dbl-like 
exchange factor Tiam 1, the catalytic domain of SopE has an entirely differ- 
ent architecture from that of Tiam 1; furthermore, it interacts with the switch 




regions by means of different amino acids. In this regard, SopE is the first 
example of a non-Dbl-like protein capable of inducing guanine nucleotide 
exchange in Rho family proteins. 



SopB 

SopB exhibits potent phosphoinositide phosphatase activity in vitro and 
is capable of mediating pronounced inositol phosphate fluxes in vivo (Galyov 
et al., 1997). In addition, SopB has been found to stimulate Cdc42-dependent 
rearrangements of the actin cytoskeleton that are a prerequisite for cellular 
invasion. The ability of SopB to activate Cdc42 is dependent on its phos- 
phatase activity, because a phosphatase-defective SopB in which a critical 
active-site cysteine residue was changed to serine lost its ability to activate 
g Cdc42. Because inositol -based molecules can directly affect Cdc42 activity, it 

s is thought that SopB activates Cdc42 and Racl indirectly by fluxing cellular 

8 phosphoinositides (Zhou et al., 2001). 

^ The activation of Cdc42 and Racl triggers a series of signal transduction 

X events that lead to actin cytoskeleton rearrangements. Despite their different 

£ biochemical activities, SopE/E2 and SopB exert at least partially redundant 

functions during Salmonella invasion. Thus, introduction of a loss-of- function 
mutation in the genes that encode either one of these proteins results in a 
minor defect in Salmonella entry. However, the simultaneous inactivation of 
SopE/E2 and SopB results in a very severe entry defect. 



SptP 

Cells infected with S. typhimurium quickly recover from the dramatic 
actin cytoskeletal rearrangements and regain their normal cellular architec- 
ture. SptP was identified as a S. typhimurium protein with homology in its 
carboxy-terminal to both prokaryotic and eukaryotic phosphatases, and it was 
demonstrated to possess tyrosine phosphatase activity (Kaniga et al., 1996). 
Although SptP mutants do not have an invasion deficiency, cells infected with 
sptP-deficient S. typhimurium do not exhibit normal recovery of their actin 
cytoskeleton following bacterial entry. Sequence scanning of SptP revealed 
a region in its amino terminus with homology to GAPs for Rho proteins, 
which is also possessed by other bacterial pathogens (ExoS of Pseudomonas 
spp. and YopE of Yersinia spp.), as well as by eukaryotes. SptP behaves as a 
GAP for Cdc42 and Racl, but not Rho A or Ras. A mutation of arginine to ala- 
nine within the proposed catalytic arginine finger abrogated GAP activity (Fu 
and Galan, 1999). These results suggest that SopE/E2 and SptP coordinately 



control the GDP-GTP cycle of Rac and Cdc42 in host cells, thereby modu- 
lating the actin cytoskeleton. Thus, SptP's GAP activity opposes the Cdc42 
and Racl activating function of SopE, SopE2, and SopB to help the host cell 
rebuild its actin cytoskeletal network. How these proteins are regulated in 
vivo so that their activities do not nullify each other is not yet clear, but it may 
be due to differential secretion or activation of SptP by its chaperone, SicP 
(Fu and Galan, 1998). In addition to its GAP activity located within the amino 
terminus, the carboxy-terminal domain of SptP possesses potent tyrosine 
phosphatase activity. Such tyrosine phosphatase activity of SptP is not only 
involved in reversing the MAP kinase activation that results from Salmonella 
invasion but also targets the intermediate filament vimentin, which is re- 
cruited to the membrane ruffles stimulated by Salmonella (Murli et al., 2001) . 



SipC 




< 
> 

on 

SipC has been reported to nucleate and bundle actin in vitro. The z 

bundling and nucleation activities are located at different domains of SipC. w 

The precise role of these activities in vivo is unknown because the necessary > 

experiments to address this important issue are hampered by the fact that £ 

SipC is required for the translocation of effector proteins into host cells. SipC £ 

has been identified along with SipB as a general chaperone for the transloca- c* 

tion of other SPI-1 type III secreted effector proteins into the host cell (Carlson Eg 

o 
and Jones, 1998). In addition, SipC becomes translocated into the host cell, g 

where it has a bipartite ability to modulate actin polymerization directly. In £ 

vitro, the C -terminus of SipC aids in the nucleation of new actin filaments 
(the rate-limiting step in actin polymerization), whereas the N-terminal half 
facilitates filament bundling. Accordingly, microinjection of purified SipC 
protein into HeLa cells induces actin polymerization, but rather than induc- 
ing ruffles like SopE, it promotes the condensation of filamentous actin into 
large aggregates (Hayward and Koronakis, 1999). The physical function of 
these aggregates is unclear. 

SipA 

SipA is also encoded within and secreted by the SPI-1 TTSS. It is thought 
that SipA affects actin dynamics in cells by initiating actin polymerization at 
the site of Salmonella entry by lowering the critical concentration of actin 
required for polymerization (Zhou et al., 1999a). A sip A mutant strain of S. 
typhimurium has a minor invasion deficiency that is only detectable at very 
early time points (up to 20 min) of bacterial entry. Furthermore, although the 



sip A mutant elicits actin-dependent membrane ruffling, these ruffles are less 
localized to sites of internalization than those induced by wild-type bacteria 
(Zhou et al., 1999a). In additon, SipA was found to bind filamentous actin 
in an in vitro binding assay and induce formation of actin bundles at sites of 
bacterial internalization (Hayward and Koronakiz, 1999). In vivo, SipA may 
additionally affect actin dynamics by binding to, and enhancing the activity 
of, the bundling protein T-plastin (Zhou et al., 1999a, 1999b). Furthermore, 
McGhie et al. (2001) recently determined that SipA potentiates the effects of 
SipC on filamentous actin nucleation and bundling in vitro. Thus, it appears 
that SipA modulates the internalization process by decreasing the critical 
concentration for actin polymerization, inhibiting depolymerization of actin 
filaments, and increasing the bundling activity of T-plastin. 

SipA is also unique in that interaction of this effector protein at the apical 
g surface of intestinal epithelial cells is sufficient to initiate the proinflamma- 

s tory signal transduction pathway that leads to polymorphonuclear leukocyte 

8 (PMN) transepithelial migration. The recruitment of PMN to the intestinal 

^ epithelium is a key virulence determinant underlying the development of 




< 



x Salmonella-elicited enteritis. Purified SipA applied to the apical surface of 



H 



£ intestinal epithelial cells initiates an ADP ribosylating factor 6 (ARF6) lipid- 

signaling cascade, which, in turn directs the activation of protein kinase C 
(PKC) and subsequent PMN transepithelial migration (Lee et al., 2000; Criss 
et al., 2001). This demonstrates that some SPI-1 effector proteins involved in 
the invasion of epithelial cells may have additional roles that do not require 
their introduction directly into the cytosol of the host. Moreover, the signifi- 
cance of these results has been confirmed by the finding that SipA plays an 
important role in eliciting proinflammatory responses, such as PMN influx, 
during Salmonella infection of calves - a relevant in vivo model system used 
to study human enterocolitis (Zhang et al., 2002). 

ROLE OF SPI-1 IN PATHOGENESIS 

To understand the role of invasion in Salmonella pathogenesis, re- 
searchers have investigated the in vivo phenotypes of invasion gene mutants. 
Most in vivo studies have used the murine model of typhoid fever, in which 
orally introduced S. typhimurium causes a systemic illness in Balb/c mice. 
To induce systemic illness in these animals, S. typhimurium first colonize 
the distal ileum, and, after successful colonization, a subpopulation of S. ty- 
phimurium can be found in the gut-associated lymphatic tissues. Still later, 
host death can occur in response to high numbers of bacteria found within 
deep lymphoid-rich organs such as the spleen and liver (Carter and Collins, 



1974). Invasion, per se, has long been thought to be important for this pro- 
cess because mutants that are noninvasive in vitro are less able to reach the 
spleen and liver and so are attenuated in orally infected mice. However, if 
introduced systemically, noninvasive mutants are as virulent as the wild-type 
(Galan and Curtiss, 1989; Ahmer et al., 1999). Many groups have interpreted 
such data to mean that invasion of nonphagocytic cells allows S. typhimurium 
to access the lymphatics, especially through the Peyer's patches that underlie 
M cells in the distal ileum (Penheiter et al., 1997). 

In spite of this, several recent observations suggest that invasion, per se, 
is not always required for S. typhimurium to access privileged sites within 
the host. For example, S. typhimurium can reach the spleen in an invasion- 
independent manner by residing inside CD 18+ macrophages (Vazquez- 
Torres et al., 1999). In yet another example, it has been postulated that CD18- 
exp res sing phagocytes are involved in an alternate route for bacterial invasion 5 

(Rescigno et al., 2001). Among CD18+ cells, dendritic cells are migratory and & 

phagocytic cells that are ideally located for antigen sampling in tissues that in- z 

terface with the external environment, where they perform a sentinel function w 

for incoming pathogens. With the use of polarized monolayers of the intesti- > 

nal epithelial cell Caco-2, it has been shown that dendritic cells are able to K 

open up the tight junctions between epithelial cells, send dendrites outside £ 

the epithelium, and directly sample bacteria. Because dendritic cells express c* 

tight junction proteins (i.e., claudin-1, occludin, and ZO-1) the integrity of § 

o 
the mucosa is preserved. This unique cell-cell interaction allows dendritic g 




cells to sample the environmental microorganisms without compromising £ 

the barrier function and to deliver the organisms to the lymphoid tissues 
where an efficient immune response can be mounted. Thus, this identifies 
a new mechanism for bacterial uptake in the mucosa tissue. 

In a different study it was demonstrated that S. typhimurium lacking 
the entirety of SPI-1 cannot invade tissue cells but nevertheless still dis- 
seminates to systemic organs in Balb/c mice following intragastric infection 
(Murray and Lee, 2001) . This observation highlights a new important concept 
in Salmonella pathogenesis establishing that TTSS-1 may also have activities 
aside from inducing invasion-associated rearrangements inside nonphago- 
cytic epithelial cells. It is now apparent that S. typhimurium attracts, kills, 
and parasitizes different immune cells, and some of these activities require 
TTSS-1. 

Therefore, the in vivo significance of invasion is likely to vary depending 
on the particular host-bacterial interaction. For instance, invasion may be re- 
quired for some aspects of virulence but not for access to deeper tissues. It is 
also likely that there are situations in which invasion-independent TTS S S P I - 1 




w 






phenotypes contribute significantly to salmonellosis and others in which 
invasion-dependent systemic dissemination is more critical. Although inva- 
sion can be uncoupled from some pathogenesis-associated phenotypes in 
vitro, it is not currently feasible to uncouple invasion from other TTSS SPI-1 
phenotypes in vivo. Thus, the relative contribution of different SPI-1 pheno- 
types remains to be elucidated. 



HISTORICAL PERSPECTIVE OF SPI-1 

The genes that comprise the SPI-1 are not present in the genome of 
Escherichia coli K-12, but groups of similarly organized genes with related 
sequences occur on the virulence plasmids of the invasive enteric pathogens 
Shigella and Yersinia and in the genome of certain plant and animal pathogens 
2 of the genera Erwinia, Pseudomonas, and Xanthomonas (Li et al., 1995). And, 

rt as already mentioned, there are similarities between certain SPI-1 genes and 

u loci involved in biogenesis of flagella in a variety of bacteria. 

^ Perhaps the best characterized example of the functional and structural 

h conservation in TTSS-1 is between Salmonella and Shigella. The inv/spa genes 

of SPI-1 are homologous to the Shigella mxi/spa genes; as a consequence, it 
is not surprising that the Salmonella and Shigella TTSS not only exhibit signi- 
ficant similarities in the primary sequence of their determinants (Hermant 
et al., 1995) but also complement each other functionally for secretion in 
vitro (Rosqvist et al., 1995). Perhaps they even share essentially the same 
macro molecular structure. Moreover, there is also significant structural and 
functional conservation between the Sip proteins encoded on the SPI-1 of 
S. typhimurium and the Ipa proteins encoded on the Shigella mxi/spa locus, 
suggesting that the entry processes engaged by these two enteric pathogens 
are promoted by similar effectors (Hermant et al., 1995). 

In consideration of the base compositions, genomic locations (i.e., chro- 
mosome vs. plasmid), and phylogenic distribution of these genes, it is un- 
likely that the SPI-1 TTSS complex was ancestral in the Enterobacteriaceae. 
In addition, given their relatively low G + C content in S. enterica (46%), 
Li et al. (1995) proposed that the SPI-1 genes were horizontally transferred 
from Yersinia. However, because of their occurrence in all the subspecies of 
S. enterica and the overall similarity of their evolutionary diversification to 
that of housekeeping genes (Li et al., 1995), it is more likely that they were 
already present in the last common ancestor of the contemporary lineages 
of the salmonellae. Thus, it is generally agreed that Salmonella, Yersinia, and 
Shigella independently acquired these genes from another source. 




REFERENCES 

Ahmer, B.M., van Reeuwijk, J., Watsom, P.R., Wallis, T.S., and Heffron, F. (1999). 
Salmonella SirA is a global regulator of genes mediating enteropathogenesis. 
Mol. Microbiol 31, 971-982. 

Bakshi, C.S., Singh, V.P., Wood, M.W., Jones, P.W., Wallis, T.S., and Galyov, E.E. 
(2000). Identification of SopE2, a Salmonella secreted protein which is highly 
homologous to SopE and involved in bacterial invasion of epithelial cells. J. 
Bacteriol. 182, 2341-2344. 

Bolton, A.J., Osborne, M.P., Wallis, T.S., and Stephen, J. (1999). Interaction of 
Salmonella choleraesuis, Salmonella duhlin, and Salmonella typhimurium with 
porcine and bovine terminal ileum in vivo. Microbiology 145, 2431-2441. 

Buchwald, G., Friebel, A., Galan, J.E., Hardt, W.D., Wittinghofer, A., and Schef- 
fzek, K. (2002). Structural basis for the reversible activation of a Rho protein 
by the bacterial toxin SopE. EMBO J. 21, 3286-3295. < 

Carlson, S.A. and Jones, B.D. (1998). Inhibition of Salmonella typhimurium in- 3 

vasion by host cell expression of secreted bacterial invasion proteins. Infect. g 

Immun. 66, 5295-5300. £ 

Caron, E. and Hall, A. (1998). Identification of two distinct mechanisms of phago- g 

cytosis controlled by different GTPases. Science 282, 1717-1721. S 

Carter, P.B. and Collins, F.M. (1974). The route of enteric infection in normal 2 

mice. J. Exp. Med. 139, 1189-1203. > 

Chen, L.-M., Hobbie, S., and Galan, J.E. (1996). Requirement of CDC42 for o 

Salmonella-induced cytoskeletal and nuclear responses. Science 274, 2115- h 

2118. * 

Cirillo, D.M., Valdivia, R.H., Monack, D.M., and Falkow, S. (1998). Macrophage- 
dependent induction of the Salmonella pathogenicity island 2 type III secre- 
tion system and its role in intracellular survival. Mol. Microbiol. 30, 175-188. 

Collazo, CM. and Galan, J.E. (1997). The invasion-associated type III system of 
Salmonella typhimurium directs the translocation of Sip proteins into the host 
cell. Mol. Microbiol. 24, 747-756. 

Cox, D., Chang, P., Zhang, Q, Reddy, P.G., Bokoch, G.M., and Greenberg, S. 
(1997). Requirement for both Racl and Cdc42 in membrane ruffling and 
phagocytosis in leukocytes. J. Exp. Med. 186, 1487-1494. 

Criss, A.K., Ahlgren, D.M., Jou, T-S., McCormick, B.A., and Casanova, J.E. (2001). 
The GTPase Racl selectively regulates Salmonella invasion at the apical 
plasma membrane of polarized epithelial cells. J. Cell Sci. 114, 1331-1341. 

Criss, A.K. and Casanova, J.E. (2003). Coordinate regulation of Salmonella enterica 
serovar Typhimurium invasion of epithelial cells by the Arp 2/3 complex and 
Rho GTPases. Infect. Immun. 71, 2885-2891. 



Criss, A.K., Silva, M., Casanova, J.E., and McCormick, B.A. (2001). Regula- 
tion of Salmonella-induced neutrophil transmigration by epithelial ADP- 
ribosylation factor 6. J. Biol. Chem. 276, 48,431-48,439. 
Darwin, K.H. and Miller, V.L. (1999a). Molecular basis of the interaction of 

Salmonella with the intestinal mucosa. Clin. Microbiol. Rev. 12, 405-428. 
Darwin, K.H. and. Miller, V.L. (1999b). InvF is required for expression of genes 
encoding proteins secreted by the SPI1 type III secretion apparatus in 
Salmonella typhimurium. J. Bacteriol. 181, 4949-4954. 
Eichelberg, K., Ginocchio, C.C., and Galan, J.E. (1994). Molecular and functional 
characterization of the Salmonella typhimurium invasion genes invB and invC: 
homology of InvC to the FOF1 ATPase family of proteins. J. Bacteriol. 176, 
4501-4510. 
Fath, K.R., Mamajiwalla, S.N., and Burgess, D.R. (1993). The cytoskeleton in 
g development of epithelial cell polarity. J. Cell Sci. Suppl. 17, 65-73. 

| Finlay, B.B. and Falkow, S. (1988). Comparison of the invasion strategies used 

u by Salmonella cholerae-suis, Shigella jlexneri and Yersinia enterocolitica to enter 

^ cultured animal cells: endosome acidification is not required for bacterial 




< 



E invasion or intracellular replication. Biochimie. 70, 1089-1099. 



H 



£ Finlay, B.B. and Falkow, S. (1990). Salmonella interactions with polarized human 

intestinal Caco-2 epithelial cells. J. Infect. Dis. 162, 1096-1106. 

Finlay, B.B., Gumbiner, B., and Falkow, S. (1988). Penetration of Salmonella 
through a polarized Madin-Darby canine kidney epithelial cell monolayer. J. 
Cell Biol. 107, 221-230. 

Finlay, B.B., Ruschkowski, S., and Dedhar, S. (1991). Cytoskeletal rearrangements 
accompanying Salmonella entry into epithelial cells. J. Cell Sci. 99, 283-296. 

Francis, C.L., Ryan, T.A., Jones, B.D., Smith, S.J., and Falkow, S. (1993). Ruffles 
induced by Salmonella and other stimuli direct macropinocytosis of bacteria. 
Nature 364, 639-642. 

Fu, Y. and Galan, J.E. (1998). Identification of a specific chaperone for SptP, 
a substrate of the centrisome 63 type III secretion system of Salmonella 
typhimurium. J . Bacteriol. 180, 3393-3399. 

Fu, Y. and Galan, J.E. (1999). A Salmonella protein antagonizes Rac-1 and cdc42 
to mediate host recovery after bacterial invasion. Nature 401, 293-297. 

Galan, J.E. and Curtiss, R. (1989). Cloning and molecular characterization of 
genes whose products allow Salmonella typhimurium to penetrate tissue cul- 
ture cells. Proc. Natl. Acad. Sci. USA 86, 6383-6387. 

Galyov, E.G., Wood, M.W., Rosqvist, R., Mullan, P.B., Watson, P.R., Hedges, 
S., and Wallis, T.S. (1997). A secreted effector protein of Salmonella dublin 
is translocated into eukaryotic cells and mediates inflammation and fluid 
secretion in infected ileal mucosa. Mol. Microbiol. 25, 903-912. 



Hall, A. (1998). Rho GTPases and the actin cytoskeleton. Science 270, 509-514. 

Hansen-Wester, I. and Hensel, M. (2001). Salmonella pathogenicity island encod- 
ing type III effector systems. Microbes Infect. 3, 549-559. 

Hardt, W.D., Chen, L.M., Schuebel, K.E., Bustelo, X.R., and Galan, J.E. (1998). S. 
typhimurium encodes an activator of Rho GTPases that induces membrane 
ruffling and nuclear responses in host cells. Cell 93, 815-826. 

Hayward, R.D. and Koronakis, V. (1999). Direct nucleation and bundling of actin 
by the SipC protein of invasive Salmonella. EMBO J. 18, 4926-4934. 

Hensel M., Shea, J.E., Waterman, S.R., Mundy, R., Nikolaus, T., Banks, G., 
Vazquez-Torres, A., Gleeson, C, Fang, F.C., and Holden, D.W. (1998). 
Genes encoding putative effector proteins of the type III secretion system 
of Salmonella pathogenicity island 2 are required for bacterial virulence and 
proliferation in macrophages. Mol. Microbiol. 30, 163-174. 

Hermant, D., Menard, R., Arricau, N., Parsot, C, and Popoff, M.Y. (1995). Func- 5 

tional conservation of the Salmonella and Shigella effectors of entry into ep- % 

ithelial cells. Mol. Microbiol. 17, 781-789. § 

Hohmann, E.L. (2001). Nontyphoidal salmonellosis. Clin. Infect. Dis. 32, 263- w 

269. SS 

Hook, E.W. (1990). Salmonella species (including typhoid fever). In Principles and K 

Practice of Infectious Diseases, ed. G.L. Mandell, R.G. Douglas, and J.E. Bennet, £ 

pp. 1700-1716. New York: Churchill Livingstone. ^ 

Kaniga, K., Uralil, J., Bliska, J.B., and Galan, J.E. (1996). A secreted tyrosine phos- g 

o 
phate with modular effector domains in the bacterial pathogen Salmonella g 




H 



typhimurium. Mol. Microbiol. 21, 633-641. £ 

Kozma, R., Ahmed, S., Best, A., and Lim, V. (1995). The Ras-related protein 
Cdc42Hs and bradykinin promote formation of peripheral actin microspikes 
and filopodia in Swiss 3T3 fibroblasts. Mol. Cell. Biol. 15, 1942-1952. 

Kubori, T., Matsushima, Y., Nakamura, D., Uralil, J., Lara-Tejero, M., Sukhan, 
A., Galan, J.E., and Aizawa, S.I. (1998). Supramolecular structure of the 
Salmonella typhimurium type III protein secretion system. Science 280, 602- 
605. 

Kubori, T., Sukhan, A., Aizawa, S.I., and Galan, J.E. (2000). Molecular character- 
ization and assembly of the needle complex of the Salmonella typhimurium 
type III protein secretion system. Proc. Natl. Acad. Sci. USA 97, 10,225- 
10,230. 

Lee, C.A. and Falkow, S. (1990). The ability of Salmonella to enter mammalian 
cells is affected by bacteria growth state. Proc. Natl. Acad. Sci. USA 87, 4304- 
4308. 

Lee, C.A., Silva, M., Siber, A.M., Kelly, A.J., Galyov, E., and McCormick, B.A. 
(2000). A secreted Salmonella protein induces a proinflammatory response 




in epithelial cells, which promotes neutrophil migration. Proc. Natl. Acad. 
Sci. USA97, 12,283-12,288. 

Li, J., Ochman, H., Groisman, E.A., Boyd, E.F., Solomon, F., Nelson, K., and 
Selander, R.K. (1995). Relationship between evolutionary rate and cellular 
location among the Inv/Spa invasion proteins of Salmonella enterica. Proc. 
Natl. Acad. Sci. USA 92, 7252-7256. 

Lostroh, C.P., Bajaj, V., and Lee, C.A. (2000). The cis requirement for transcrip- 
tional activation by HilA, a virulence determinant encoded on SPI1. Mol. 
Microbiol. 37, 300-315. 

Lucas, R.L. and Lee, C.A. (2001). Roles ofhilC and hilD in regulation of hilA ex- 
pression in Salmonella enterica serovar typhimurium. J. Bacteriol. 183, 2733- 
2745. 

Massol, P., Montcourrier, P., Guilemot, J.C., and Chavier, P. (1998). Fc receptor- 
g mediated phagocytosis requires CDC42 and Racl. EMBO J. 17, 6219-6229. 

| McGhie, E.J., Hayward, R.D., and Koronakis, V. (2001). Cooperation between 

u actin-binding proteins of invasive Salmonella: SipA potentiates SipC nucle- 

s ation and bundling of actin. EMBO J. 20, 2131-2139. 

X Mills, D.M., Bajaj, V., and Lee, C.A. (1995). A 40 kb chromosomal fragment encod- 

£ ing Salmonella typhimurium invasion genes is absent from the corresponding 

region of the Escherichia K-12 chromosome. Mol. Microbiol. 15, 749-759. 

Mills, S.D. and Finlay, B.B. (1994). Comparison of Sal monella typhi and Salmonella 
typhimurium invasion, intracellular growth and localization in cultured hu- 
man epithelial cells. Microb. Pathog. 17, 409-423. 

Mirold, S., Rabsch, W., Rohde, M., Stender, S., Tschape, H., Russmann, H., Igwe, 
E., and Hardt, W.D. (1999). Isolation of a temperate bacteriophage encoding 
the type III effector protein SopE from an epidemic Salmonella typhimurium 
strain. Proc. Natl. Acad. Sci. USA 96, 9845-9850. 

Murli, S., Watson, R.O., and Galan, J.E. (2001). Role of tyrosine kinases and the 
tyrosine phosphatase SptP in the interaction of Salmonella with host cells. 
Cell. Microbiol. 3, 795-810. 

Murray, R.A. and Lee, C.A. (2000). Invasion genes are not required for Salmonella 
enterica serovar typhimurium to breach the intestinal epithelium: evidence 
that Salmonella pathogenicity island 1 has alternative functions during infec- 
tion. Infect. Immun. 68, 5050-5055. 

Nobes, CD. and Hall, A. (1995). Rho, Rac, and Cdc42 GTPases regulate the 
assembly of multimolecular focal adhesion complexes associated with actin 
stress fibers, lamellipodia and filopodia. Cell 81, 53-62. 

Penheiter, K.L., Mathur, N., Giles, D., Fahlen, T., and Jones, B.D. (1997). Non- 
invasive Salmonella typhimurium mutants are avirulent because of an inability 




to enter and destroy M cells of ileal Peyer's patches. Mol. Microbiol. 24, 697- 
709. 

Rescigno, M., Urbano, M., Valzasina, B., Francolini, M., Rotta, G., Bonasio, 
R., Granucci, F., Kraehenbuhl, J-P., and Ricciardi-Castagnoli, P. (2001). 
Dendritic cells express tight junction proteins and penetrate gut epithelial 
monolayers to sample bacteria. Nature Immunol. 2, 361-367. 

Ridley, A.J. and Hall, A. (1992). The small GTPase binding protein rho regulates 
the assembly of focal adhesions and actin stress fibers in response to growth 
factors. Cell 70, 389-399. 

Rosqvist, R., Hakansson, S., Forsberg, A., and Wolf-Watz, H. (1995). Functional 
conservation of the secretion and translocation machinery for virulence pro- 
teins of Yersiniae, Salmonellae, and Shigellae. EMBO J. 14, 4187-4195. 

Rout, W.R., Formal, S.B., Dammin, G.J., and Giannella, R.A. (1974). Pathophys- 
iology of Salmonella diarrhea in the Rhesus monkey: intestinal transport, 5 
morphological and bacteriological studies. Gastroenterology 67, 59-70. ^ 

Rudolph, M.G., Weise, C., Mirold, S., Hillenbrand, B., Bader, B., Wittinghofer, s 

A., and Hardt, W.D. (1999). Biochemical analysis of SopE from Salmonella w 

typhimurium, a highly efficient guanosine nucleotide exchange factor for Rho > 

GTPases. J. Biol. Chem. 274, 30,501-30,509. £ 

Shea, J.E., Beuzon, C.R., Gleeson, C., Mundy, R., and Holden, D.W. (1999). Influ- £ 

ence of the Salmonella typhimurium pathogenicity island 2 type III secretion ^ 

system on bacterial growth in the mouse. Infect. Immun. 67, 213-219. g 

Stender, S., Friebel, A., Linder, S., Rohde, M., Mirold, S., and Hardt, W.D. (2000). g 

Identification of SopE2 from Salmonella typhimurium, a conserved guanine £ 

nucleotide exchange factor for Cdc42 of the host cell. Mol. Microbiol. 36, 
1206-1221. 

Takeuchi, A. (1967). Electron microscope studies of experimental Salmonella 
infection. I. Penetration into the intestinal epithelium by Salmonella ty- 
phimurium. Am. J. Pathol. 50, 109-136. 

Uchiya, K., Barbieri, M.A., Funato, K., Shah, A.H., Stahl, P.D., and Groisman, 
E.A. (1999). A Salmonella virulence protein that inhibits cellular trafficking 
EMBO J. 18, 3926-3933. 

Van Aelst, L. and D'Souza-Schorey, C. (1997). Rho GTPases and signaling net- 
works. Genes Dev. 11, 2295-2322. 

Vazquez -Torres A., Jones-Carson, J., Mastroeni, P., Ischiropoulos, H., and Fang, 
F.C. (2000). Antimicrobial actions of the NADPH phagocyte oxidase and 
inducible nitric oxide synthase in experimental salmonellosis. I. Effects on 
microbial killing by activated peritoneal macrophages in vitro. J. Exp. Med. 
192, 227-236. 




Vazquez-Torres, A., Jones-Carson, J., Baumler, A.J., Falkow, S., Valdivia, R., 
Brown, W., Le, M., Berggren, R., Parkos, W.T., and Fang, F.C. (1999). Ex- 
traintestinal dissemination of Salmonella by CD-18-expressing phagocytes. 
Nature 401, 804-808. 
Wallis, T.S. and Galyov, E.E. (2000). Molecular basis of Sa/mone//a-induced en- 
teritis. Mol. Microbiol. 36, 997-1005. 
Watson, P.R., Paulin, S.M., Bland, A.P., Jones, P.W., and Wallis, T.S. (1995). Char- 
acterization of intestinal invasion by Salmonella typhimurium and Salmonella 
dublin and effect of a mutation in the invH gene. Infect. Immun. 63, 2743- 
2754. 
Wood, M., Rosqvist, W.R., Mullan, P.B., Edwards, M.H., and Galyov, E.E. (1996). 
SopE, a secreted protein of Salmonella dublin, is translocated into the target 
eukaryotic cell via a sip-dependent mechanism and promotes bacterial entry. 
g Mol. Microbiol. 22, 327-338. 

| Zhang, S., Santos, R.L., Tsolis, R.M., Stender, S., Hardt, W.D., Baumler, A.J., and 

u Adams, L.G. (2002). The Salmonella enterica Serotype Typhimurium effec- 

^ tor proteins SiPA, SoPA, SopB, SopD, and SopE2 act in concert to induce 

E diarrhea in calves. Infect. Immun. 70, 3843-3855. 

£ Zhou, D. and Galan, J.E. (2001). Salmonella entry into host cells: the work in 

concert of type III secreted effector proteins. Microb. Infect. 3, 1293-1298. 
Zhou, D., Chen, L.L.H., Shears, B.S., and Galan, J.E. (2001). A Salmonella in- 
ositol polyphosphatase acts in conjunction with other bacterial effectors to 
promote host-cell actin cytoskeleton rearrangements and bacterial internal- 
ization. Mol. Microbiol. 39, 248-259. 
Zhou, D., Mooseker, M.S., and Galan, J.E. (1999a). An invasion-associated 
Salmonella protein modulates the actin-bundling activity of plastin. Proc. Natl. 
Acad. Sci. USA 96, 10,176-10,181. 
Zhou, D., Mooseker, M.S., and Galan, J.E. (1999b). Role of S. typhimurium actin- 
binding protein SipA in bacterial internalization. Science 283, 2092-2095. 



CHAPTER 2 

Shigella invasion 

Chihiro Sasakawa 



INVASION 

Shigella invasion and the host inflammatory responses 

Shigella cause bacillary dysentery (shigellosis), a disease provoking a se- 
vere inflammatory diarrhea in humans and primates. In tropical areas of 
developing countries, shigellosis is endemic and a major killer of children 
under 5 years of age. Shigellosis occurs following ingestion of a very small 
number (100-1000) of bacteria, thus permitting easy spread of the disease by 
person-to-person contact as well as by the drinking of contaminated water. 

Shigella, a Gram-negative bacillus, comprises four species - S. dysenteriae, 
S.flexneri, S. boydii, and S. sonnei (Pupo et al., 2000; Lan and Reeves, 2002). 
Shigella is now recognized as a member of Escherichia coli; however, the group 
of bacteria causing shigellosis is idiomatically called Shigella in this chapter. 
Shigellosis is also caused by enteroinvasive E. coli (EIEC), a pathogenic E. coli. 
Shigella and EIEC possess a large 210- to 230-kb plasmid on which the major 
virulence functions are encoded. Because Shigella has neither adhesins for 
upper GI tract cells nor flagella, after infection by means of the fecal-oral 
route the bacteria reach the colon and rectum directly, where they translocate 
through the epithelial barrier by means of the M cells overlaying the solitary 
lymphoid nodules (Fig. 2.1; also see Wassef et al., 1989; Sansonetti et al., 

1991, 1996). Once they have reached the underlying M cells, Shigella infect 
the resident macrophages and multiply. Within the macrophages Shigella 
secrete IpaB, which specifically binds to, cleaves, and activates caspase-1, 
thus leading to macrophage cell death through apoptosis (Zychlinsky et al., 

1992, 1994, 1996). 

The stimulation of caspase-1 in infected macrophages causes the pro- 
duction of large amounts of lL-l/3 and IL-18, thus eventually leading to an 




actin tail 



bacteria 



Mcell 



enterocyte 




< 

GO 

< 

GO 

O 

i— i 

X 
i— i 

X 
u 



apical side 



AIWWMWWVW1 




nucleus 



basolateral side 



macrophage 



Figure 2.1. A simplified model for the infection of colonic epithelial cells by Shigella (refer 
to the text for details). 



increase in the permeability of the epithelial barrier to Shigella and migra- 
tion of polymorphonuclear leukocytes (PMNs; see Zychlinsky et al., 1992; 
Zychlinsky and Sansonetti 1997). Meanwhile, the bacteria released from the 
dead macrophages immediately enter surrounding enterocytes from the ba- 
solateral surface by directing large-scale membrane ruffling, which finally 
leads to phagocytic events. Though the invading bacterium is entrapped by 
a phagocytic membrane, Shigella immediately disrupts the membrane and 
escapes into the cytoplasm (Fig. 2.1; also see Sansonetti et al., 1986). Within 
the cytoplasm, Shigella multiply and induce actin polymerization at one pole 
of the bacterium, by which the intracellular bacterium can gain a propulsive 
force to move intracellularly and intercellularly (Fig. 2.1; also see Bernardini 
etal., 1989). 

Internalized Shigella release a large amount of lipopolysaccharide (LPS) 
into the host cytoplasm, where the LPS binds Nodi, a member of the CED4/ 
Apaf-1 superfamily, eventually leading to activation of NF-/cB by means of 
the stimulation of the bipartite CARD-kinase protein, RICK (Inohara et al., 
2000; Girardin et al., 2001). In response to the activation of NF-/cB, colonic 
epithelial cells express a large array of proinflammatory cytokines, especially 
IL-8, thus further promoting local inflammation and attracting more PMNs 
(Zychlinsky and Sansonetti, 1997). Therefore, the predominant pathogenic 
feature of Shigella is the ability to invade macrophages as well as epithelial 
cells, including subsequent dissemination into adjacent epithelial cells. In 




H 



< 
> 

CO 

o 



this chapter, I mostly focus on the bacterial system involved in the invasion 
of epithelial cells and subsequent intracellular and intercellular spreading 
processes. 

Basolateral entry into polarized cells 

Shigella infection of polarized epithelial cells such as Caco-2 or Madin 
Darby Canine Kidney (MDCK) cells reveals that the bacteria invade from the 
basolateral surface into the cells (Mounier et al., 1992). This characteristic 
entry can be seen when Shigella infect rabbit ligated ileal loops, where bac- 
teria move to the basolateral surface by means of the M cells (Wassef et al., 
1989; Sansonetti et al., 1991, 1996), indicating that Shigella have an affin- 
ity to the basolateral surface of the polarized enterocytes to effectuate entry. 
Shigella are capable of inducing a highly dynamic rearrangement of actin 8 

and tubulin cytoskeletons during entry, and this leads to a large-scale phago- £ 

cytic event. Shortly after coming into contact with epithelial cells, Shigella 
induce the formation of focal adhesion-like actin-dense patches beneath the 
bacterial contact point (Tran Van Nhieu and Sansonetti, 1999) and trigger 
local destruction of the microtubule structure (Yoshida et al., 2002), which 
is followed by the protrusion of large-scale membrane ruffles. The invading 
Shigella are finally enclosed by a large membrane vacuole, but the pathogens 
immediately escape into the host cell cytoplasm, where they elicit intracellular 
and intercellular movement (Fig. 2.1). 

Invasion-associated genes on the large plasmid 

The invasion of epithelial cells by Shigella requires many genes mostly 
confined to the 31-kb pathogenicity island (PAI) on the large virulence plas- 
mid, which is highly conserved among Shigella spp. (Sasakawa et al., 1988; 
also see Fig. 2.2). The PAI of S.Jlexneri contains 28 genes bracketed by several 
IS elements and vestigial DNA sequences, where the 28 genes are arranged in 
several transcribed regions, encoding the components of the type III secretion 
system (TTSS), secreted effector proteins, chaperone proteins, and regulatory 
proteins (Fig. 2.2; also see Buchrieser et al., 2000; Venkatesan et al., 2001). 
ipaBCDA encode secreted effectors such as IpaA, IpaB and IpaC required for 
invasion of epithelial cells, whereas mxi and spa mostly encode components 
of the TTSS including the TTSS-associated secreted proteins. ipgA, ipgC, and 
spa 15 encode IpgA, with IpgC and Spa 15 that act as chaperones for IcsB, 
IpaB/IpaC, and IpgBl/IpaA, respectively (Menard et al., 1994a, 1994b; Page 
et al., 2002; Ogawa, unpublished results). The PAI possesses two regulatory 




< 




<N 




CO 



3 

^ 

'h^ 1 



CO 





On 






V> 






1 



9 



.8 



1 



•9 



.& 



.1 






Q 



.& 



i 



pa 










0\ 



I 



I 







o 



a 




2 s - 



S, 









3 



Q 



0) 

c 

o 

E 

i= 

.2 
+■» 

cd 

CD 

2 

CD 
Q. 

ns 
o 

1 

it 

CD 

a 

i 

=3 

u> 

CD 

a 



Figure 2.2. {facing page). The genetic structure of the ipa/mxi/spa PAI encoding the 
Shigella TTSS. (A) A simplified structure of the ipa/mxi/spa PAI. Arrows represent IS 
elements and the vestigial DNA sequences. (B) The genetic constitution of the ipa/mxi/spa 
PAI. (For details refer to the text and to Buchrieser et al., 2000 and Venkatesan et al., 2001.) 




genes, virB and mxiE, required for the expression of the virulence-associated 
genes in the 31-kb PAI (Adler et al., 1989; Dorman and Porter, 1998; Mavris 
etal.,2002). 

In addition, another plasmid-encoded gene, virF, codes for the essential 
regulatory protein VirF, an AraC-like transcriptional regulator, which directly 
binds the virB promoter to activate the virB gene. The VirB protein in turn 
activates the promoters for several transcribed regions in the 31-kb PAI. On 
the large plasmid, there is another effector gene called virA (Uchiya et al., 
1995), located near the virG (icsA) gene; together these form a PAI (Venkate- 
san et al., 2001). VirA has recently been shown to be essential for evoking 
membrane ruffling in epithelial cells and promoting Shigella entry into host 
cells (Yoshida et al., 2002). Recently the whole genomic sequence of the large 
plasmid (pWRlOO) ofS.flexneri serotype 5 was determined (Buchrieser et al., 
2000; Venkatesan et al, 2001). g 

Examination of the repertoire of proteins secreted by means of the £ 

TTSS under conditions that activate the TTSS revealed that 15 proteins 
(IcsB, IpaH9.8, IpaH7.8, IpaH4.5, MxiC, MxiL, Spa32, OspCl, OspB, IpgBl, 
OspDl, OspEl, OspF, OspG, and VirA) plus IpaA, IpaB, IpaC, IpaD, and 
IpgD can be secreted from Shigella by means of the TTSS into the medium 
(Buchrieser et al., 2000; Ogawa, unpublished results). Although the proteins 
potentially secreted by means of the TTSS such as Spa32, MxiC, or MxiL 
(Tamano et al., 2002; Tamano unpublished results) do not necessarily serve 
as effectors, studies have clearly indicated that Shigella secrete a diverse array 
of effectors into the external medium and target host cells. 

Type III secretion system 

The TTSS is a highly sophisticated bacterial effector protein delivery 
system. Upon contact with the host cells, a set of effector proteins is delivered 
from the infecting bacteria to the host cells by means of the TTSS. These 
translocated proteins have a variety of effects on host cells and are necessary 
for bacterial attachment, invasion, trafficking, and avoidance from the host 
defense systems. Although the mechanisms of protein export by means of 
the TTSS, including the biosynthesis of the secretion machinery, are still to 



H 



< 
> 

CO 

o 




< 

GO 

< 

O 



U 




2 A Scale bar, 100 nm 



Figure 2.3. The supramolecular structure of the Shigella type III secretion machinery. 
(A) An electron micrograph of the purified type III secretion machinery from S.jiexneri. 

be elucidated, some common characteristics of this secretion system have 
emerged (Hueck, 1998). 

Genetic and functional studies have indicated that the TTSS is encoded 
by more than 20 genes. The subset of genes together with other genes coding 
for the secreted effectors, chaperones, and regulators exist as a PAI, which 
potentially transposes horizontally in different species of bacteria, thus dis- 
tributing among many Gram-negative pathogenic bacteria, where some of 
the PAI are located on the chromosome and some on a plasmid. In fact, 
there is considerable homology between the proteins of the TTSS in differ- 
ent pathogens (Hueck, 1998). Of note, some of the proteins of TTSSs also 
share significant similarity to the components of the bacterial flagella ex- 
port machinery. For example, some of the putative components of the type 
III secretion complexes such as S.jiexneri MxiJ, Spa47, Spa33, Spa24, Spa9, 
Spa29, Spa40, and MxiJ share significant similarity to FlhA, Flil, FliN, FliP, 
FliQ, FliR, FlhB, and FliF of the Salmonella flagellar export system, respec- 
tively (Hueck, 1998; Cornells and Van Gijsegem, 2000; Piano et al., 2001). 

Furthermore, the TTSS is functionally and structurally similar to the 
flagellar export system. Secretion of a set of proteins by means of the TTSS 



effector 



!h 



cytoplasmic membrane 




host cell cytoplasm 

iff! 



IpaC 



needle 



IpaB 



upper rings 



outer membrane 



periplasm 



inner membrane 



B 




bacterial cytoplasm 



Spa47 ATPase 



Figure 2.3. (cont.) (B) Model of the structure of the type III secretion machinery, including 
the translocation of effectors from bacterial cytoplasm into the host cell cytoplasm. 




ir, 
H 



< 
> 

o 



is dependent on the energy supply mediated by the Fl-type ATPase asso- 
ciated with the secretion apparatus. The same is true for secretion of the 
extracellular flagellar components, which form the hook, cap, and flagella 
filament by means of the flagella export system. The supramolecular struc- 
tures of the TTSS of Salmonella typhimurium and S.flexneri have recently 
been elucidated, and they share similarity with that of the flagellar basal body 
(Fig. 2.3B; also see Kubori et al., 1998; Blocker et al., 1999; Tamano et al., 
2000). As already mentioned for the Shigella TTSS, the expression of genes 
encoding the TTSS as well as the flagella export system is under stringent 
control that is mediated by complicated regulatory networks in each of the 
bacteria. 

Despite the structural and functional similarities of the TTSS in each 
pathogen, the proteins delivered by means of the TTSS are quite diverse. 
For example, Shigella potentially deliver ~20 proteins by means of the TTSS; 
some share similarity to secreted proteins from different pathogens, whereas 
others are unique to Shigella. Studies of the proteins secreted from Yersinia, 



Salmonella, and Shigella have indicated that some have a role in linking the 
secretion complex to the target host plasma membrane, whereas others serve 
as effectors to modulate host cell functions. 

The purified type III secretion complexes from S. typhimurium and S. 
flexneri as reported by Kubori et al. (1998) and Tamano et al. (2000), respec- 
tively, contained four major components. The Salmonella type III secretion 
complex contained InvG, PrgH, PrgI, and PrgK proteins, whereas the Shigella 
type III secretion complex contained MxiD, MxiG, MxiH, and MxiJ. Recently, 
these type III secretion complexes have been shown to contain an additional 
component, which is PrgJ in Salmonella and Mxil in Shigella. The supramolec- 
ular structures of the type III secretion complexes of each bacterium observed 
by electron microscopy are similar, being composed of two distinctive parts, 
the needle and basal parts (Fig. 2.3B). The needle of the S. typhimurium and 
g S. flexneri type III secretion complexes consists mainly of PrgI and MxiH, 




< 



% respectively (Kubori et al., 2000; Tamano et al., 2000). The basal part of the S. 



< 

GO 



X 

U 



< typhimurium type III complex is composed of InvG, PrgH, and PrgK, whereas 

§ that of the S. flexneri complex is composed of MxiD, MxiG, and MxiJ. Further- 

x more, the supramolecular structures of the basal portion of both type III com- 

plexes share significant similarity to that of the Salmonella flagella basal body. 
Indeed, the basal part of the type III secretion complex possesses two 
pairs of rings, referred to as the upper and lower rings (Fig. 2.3B). Because 
the basal portion was observed by electron microscopy to be embedded in the 
osmotic-shocked bacterial envelope, similar to the flagellar basal complex, 
the two pairs of rings are thus assumed to be anchored to the inner and 
outer membranes of bacteria. The flagellar hook forms a curved protruding 
structure from which a long flagella filament is extended; the type 1 1 1 secretion 
complex forms a straight needle protruding from the basal part. The length 
of the type III needle of wild type S. flexneri is estimated to be 45 nm and 
distributed in a narrow range with a standard deviation of 3.3 nm (Tamano 
et al., 2000). The length of the basal body of the Shigella type III secretion 
complex is estimated to be approximately 31 nm, which is consistent with 
the thickness of the Gram-negative bacterial envelope (approximately 25 nm) 
(Tamano et al., 2000). This suggests that the type III secretion complex spans 
both the outer and inner membranes. Although the number of needles per 
bacterium has not been accurately determined, on the basis of the distribution 
of the type III secretion structures in a field of osmotically shocked bacterial 
envelope as observed by electron microscopy, it is estimated to be around 
50-60 per bacterium. 

Genetic and functional studies of TTSSs have strongly suggested that the 
basic morphological features displayed by Salmonella and Shigella would be 



conserved among other pathogens. Indeed, recently studies have shown that, 
in enteropathogenic E. coli (EPEC), a long (50-700 nM) filamentous structure 
protrudes from the tip of the TTSS needles; it is composed of EspA, which is 
encoded by the espA gene located on the locus of enterocyte effacement (LEE) 
PAI, downstream of the region encoding genes of the TTSS (Sekiya et al., 
2001; Daniell et al., 2001). Previous studies indicated that EspA forms a fila- 
mentous structure that assembles as a physical bridge between the bacteria 
and host cell surface, which then functions as a conduit for the transloca- 
tion of bacterial effectors into host cells (Knutton et al., 1998). In fact, the 
espA mutant of EPEC has been shown to be deficient in forming a long fila- 
mentous structure and delivering effectors such as Tir, EspB, and EspD into 
the host cells, thus becoming a nonadherent mutant. Similarly, some plant 
pathogens such as Pseudomonas syringae and Ralstonia solanacearum form 
a filamentous appendage called the Hrp-pilus, which consists of HrpA in % 

P. syringae or HrpY in R. solanacearum (Van Gijsegem et al., 2000). S 



The effectors delivered from Shigella trigger host cellular signal pathways 
to direct its own internalization event (Fig. 2.4). One signal transduction 
pathway is linked to the interaction of secreted IpaB and IpaC with putative 
host surface receptors, and the others are evoked by intracellular effectors 
such as IpaA, IpaB, IpaC, IpgD, and VirA. Although the roles of IpaB and 
IpaC during invasion are important, their functions are complicated. For 
example, IpaB and IpaC act as secreted effector proteins in the target host 
cells to stimulate caspase-1 and Rho GTPases, respectively, whereas IpaB and 
IpaD act as a molecular plug for the TTSS at the tip of the TTSS needle (Figs. 
2.3B and 2.4; also see Menard et al., 1996; Tran Van Nhieu and Sansonetti, 
1999). Like Yersinia YopB and YopD (Hakansson et al., 1996) or Salmonella 
SipB and SipC (Collazo and Galan, 1996), upon contact between Shigella and 
the host cell, IpaB and IpaC serve as a membrane pore located at the tip 
of the TTSS needle in the host cell plasma membrane, thus allowing the 
translocation of secreted effector proteins into the host cells (Fig. 2.3B; also 
see Blocker etal., 1999). 

The IpaB and IpaC proteins secreted into the culture supernatant 
form high molecular matrix-like structures, which promote Shigella invasion 
through interaction with CD44 (the hyaluronan receptor belonging to the 
immunoglobulin superfamily) and a 5 /3i integrin (Fig. 2.4; also see Skoudy 
et al., 2000; Watarai et al., 1996). Because both molecules are distributed on 
the basolateral surface of the polarized epithelial cells, these interactions are 




H 



z 



Secreted effector proteins induce host cellular signaling & 



o 












/ 


CO 


\ 




CO 


r< 




CvJ 




'co 


o 




Q. 




o 

-1— ' 


.c 
DC 




\ 


</ 


o 
o 










D) 








CO 






CM / \ 


.c 






LU/ c\ 


0. 






> = 




"D 

C 






§ Q- 


CO 




<>•■ 


S\ / - 


c 


^~ 




LO } 







o 


^^a 


Q. 


^■"^ ^^^ 


H— 


CO 


^^ 


0) 


^^Y 


=3 


DC 




DC 




CL 






o 
CO 
DC 




CD 

c 

CO 
-Q 

E 

CD 










^ 


CM 






^1- 






o 










O 





H 



< 
> 

CO 

o 



thus assumed to contribute to the basolateral entry of Shigella into the epithe- 
lial cells by mediating outside-in signaling to induce local rearrangement of 
the actin cytoskeleton (Skoudy et al., 2000). as Pi integrin localizes at focal ad- 
hesions, and the cytoplasmic domain of the fi\ integrin acts as a cytoskeleton 
linker by means of focal adhesion proteins; the cytoplasmic moiety of CD44 
also acts as a cytoskeleton linker by means of association with ERM (ezrin- 
radixin-moesin) proteins. Ezrin is shown to be recruited at the periphery of 
the extended membrane ruffles at the point of Shigella entry (Skoudy et al., 
1999). A recent study has shown that IpaB can bind CD44, where IpaB parti- 
tions during Shigella invasion within specialized membrane microdomains 
enriched in cholesterol and sphingolipids, called rafts (Lafont et al., 2002). 
CD44 is known to participate in signaling responses regulating the reorgani- 
zation of the cytoskeleton (Hirao et al., 1996). Early in the invasion (~15 min 
after contact), an accumulation of cholesterol and raft-associated proteins % 

such as GPI-anchored proteins can be observed at Shigella entry foci. £ 

In agreement with this, bacterial entry is impaired upon cholesterol de- 
pletion with methyl-/? -cyclodextrin at the site of bacterial entry (Lafont et al., 
2002). Therefore, the binding of IpaB at the tip of the TTSS needle to clus- 
tered CD44 is thought to increase binding affinity; thus rafts and clustered 
CD44 seem to be crucial for efficient Shigella entry. Meanwhile, a 5 /?i integrin 
accumulates together with F-actin, vinculin, talin, a-actinin, and tyrosine- 
phosphorylated FAK, which are major scaffolding components of focal ad- 
hesions (Watarai et al., 1996, 1997). However, these interactions alone are 
insufficient to elicit the phagocytic events required for bacterial uptake by 
epithelial cells (Menard et al., 1996). For induction of large-scale actin re- 
arrangements and large membrane protrusions sufficient to engulf several 
bacterial particles simultaneously, cellular signals evoked by Shigella effectors 
such as IpaA, IpaC, IpgD, and VirA appear to be necessary (Fig. 2.4). 

In addition, Shigella require the activation of FAK and Src tyrosine ki- 
nase to induce cytoskeletal rearrangements during entry (Watarai et al., 1997; 
Dumenil et al., 1998). Overexpression of a dominant interfering form of 
pp60 c " Src leads to inhibition of Shigella-induced cytoskeletal rearrangements 
and decreases the phosphorylation of cortactin (Dumenil et al., 1998), a 
Src substrate recruited at the site of Shigella entry (Dehio et al., 1995). Because 
focal adhesion formation is dependent on the activation of Rho and Src, an 
early cellular event triggered by Shigella contact may be associated with such a 
cellular function (Adam et al., 1996; Menard et al., 1996; Watarai et al., 1997). 



Figure 2.4. (facing page). A model for cytoskeletal rearrangements induced during Shigella 
invasion of epithelial cells, and the roles of effector proteins (refer to the text for details). 




u 



I pa A 

IpaA secreted via the TTSS upon cell contact has been indicated to mod- 
ify the Shigella-induced entry foci, and an ipaA mutant induced disorganized 
filopodial protrusions (Fig. 2.4; also see Bourdet-Sicard et al., 1999). IpaA has 
a high affinity for N-terminal residues 1-265 of vinculin. In cosedimentation 
and solid-phase assays, IpaA binding to vinculin increases the association of 
vinculin with F-actin, which in turn promotes depolymerization of F-actin 
associated with the IpaA-vinculin complex. Vinculin is a cytoskeletal protein 
present at focal adhesion and cell-cell adhesion structures, involved in cell ad- 
hesion to the extracellular matrix, cell motility, and tumorigenesis (Jockusch 
and Rudiger, 1996). Vinculin is composed of N-terminal head and C-terminal 
tail domains. It has multiple functional domains; talin and a-actinin inter- 
act with the N-terminal domain, whereas F-actin binds to the C-terminal 
| tail domain. Importantly, the interactions with proteins are determined by 

gj the conformation of vinculin. The unfolded form is the activated form and 

in 

w interacts with talin, a-actinin, and F-actin through the exposed binding do- 

g mains, whereas, in the folded conformation, the tail domain interacts with 

£ the head domain, resulting in the masking of the binding domains (Jockusch 

and Rudiger, 1996). Importantly, the 1-258 domain of vinculin has been in- 
dicated to be involved in the intramolecular association with the C-terminal 
tail domain. 

The binding of IpaA to vinculin was found to be strong, with a Kd of 
5 nM; therefore, it is likely that the binding of IpaA to the head domain dis- 
rupts the head-tail interaction, thus allowing vinculin to open up and link 
to F-actin. The vinculin-IpaA complex promotes F-actin depolymerization 
in vitro and in vivo. Indeed, microinjection of IpaA into HeLa cells induces 
a rapid (within 40 s) cell retraction with the disappearance of actin stress 
fibers. When IpaA is comicroinjected with the MBP-fused vinculin 1-265 
moiety, cell retraction induced by IpaA can be blocked and actin stress fibers 
along with vinculin-containing focal complexes are still intact, indicating 
that the vinculin-IpaA binding domain has a functional role in inducing 
the IpaA-induced cytoskeletal rearrangements. Although the mechanism for 
stimulating F-actin depolymerization is speculative, the activity of IpaA for 
induction of actin depolymerization by means of the vinculin-IpaA com- 
plex seems to be important in regulating the formation of protrusions that 
might promote detachment of Shigella. Alternatively, the IpaA-induced acti- 
vation of vinculin and recruitment of focal adhesion components may lead 
to the formation of a specific focal adhesion-like structure that might help to 
maintain bacterial contact with the epithelial surface (Bourdet-Sicard et al., 
1999). 




m 



< 
> 

CO 



IpaC 

IpaC has an activity to modulate actin dynamism, because the forma- 
tion of filopodia and lamellipodia is induced when purified IpaC protein is 
added to semipermeabilized Swiss 3T3 cells or a ipaC clone is transfected into 
HeLa cells (Tran van Nhieuet al., 1999). The IpaC-induced membrane protru- 
sions have been implicated in the activation of Cdc42, which in turn activates 
Racl, suggesting that IpaC can somehow act as the effector for promoting 
Shigella invasion of epithelial cells (Fig. 2.4). In addition, a recent study has 
indicated that Shigella direct their own entry into macrophages by exploit- 
ing IpaC to stimulate macrophage phagocytic activity (Kuwae et al., 2001). 
Indeed, Shigella invade murine macrophages such as J 774 more efficiently 
than the noninvasive ipaC mutants. Wild type Shigella can induce large-scale 
lamellipodial extensions including ruffle formation around the bacteria. In 
contrast, when macrophages are infected with the noninvasive ipaC mutant, S 

the invasiveness and induction of membrane extension are dramatically re- g 

duced. Shigella infection of J 774 cells causes tyrosine phosphorylation of 
several proteins, including paxillin and c-Cbl, and the profile of phosphory- 
lated protein is distinctive from that stimulated by S. typhimurium or phorbol 5 

ester. Upon addition of a recombinant IpaC into the external medium of 
J 774, membrane extensions were rapidly induced, and this also promoted 
uptake of E. coli. Importantly, the exogenously added IpaC was shown to be 
integrated into the macrophage plasma membrane. An analysis of the IpaC 
sequence (382 amino acids) with TMpred, a program for the prediction of pu- 
tative membrane-spanning regions, predicts that the residues encompassing 
121-139 and 169-191 are the putative transmembrane domains, whereas the 
remaining N-terminal and C-terminal regions are predicted to be presented 
on the external side of the plasma membrane. 

With the use of three IpaC antibodies that recognize three distinctive 
regions in IpaC, it has been suggested that the residues 140-168 of IpaC 
exist as the cytoplasmic loop, whereas the preceding N-terminal portion 
of TM1 (transmembrane 1) and the following C-terminal portion of TM2 
(transmembrane 2) exist as the external membrane domains (Kuwae et al., 
2001) . Although the mechanisms underlying the IpaC-mediated macrophage 
spreading event are still to be elucidated, some surface receptors such as 
Mac-1 (oiuPi integrin) and Fey R on macrophages seem to be involved in the 
phagocytic event. 

Incubation of J774 cells with wild type Shigella, but not with the ipaC 
deletion mutant, generates Mac-1 and FcyR foci at the site of bacterial 
contact. Importantly, the Mac-1 foci were observed along the periphery of 
the extended cell membrane, strongly indicating that an integrin-dependent 



adhesion event, which is a prominent feature of the Mac- 1 -mediated 
macrophage adherence, had occurred. The observed macrophage response is 
consistent with the study by Renesto et al. (1996), in which PMNs suspended 
in medium became adherent onto serum-coated wells when wild type Shigella 
but not the ipaC mutant was added to the external medium. 

Clustering of Mac-1 and FcyR has been indicated to stimulate protein 
tyrosine phosphorylation and local rearrangement of the actin cytoskeleton. 
Because some truncated versions of IpaC capable of associating with the host 
plasma membrane are able to stimulate macrophage cell spreading (Kuwae 
et al., 2001), IpaC might have an association with or effect on some pu- 
tative host receptor(s), such as otuPi integrin, mediating the induction of 
membrane extension by means of activation of a cellular signal transduction 
pathway such as the activation of Cdc42. The IpaC-induced membrane pro- 
g trusions from HeLa cells can be inhibited by a dominant interfering form of 




< 



i— i 

X 
u 



m Cdc42 , whereas a dominant interfering form of Rac results in inhibition of the 

< lamellipodium formation, further supporting the notion that IpaC has activ- 

§ ity to stimulate Cdc42 activity. This in turn causes Racl activation (Mounier 

et al., 1999; Tran van Nhieu et al., 1999). Importantly, recent studies have 
suggested that activated Racl stimulates WAVE2, a WASP (Wiskott-Aldrich 
syndrome protein) family protein, via IRSp53, a substrate for the insulin re- 
ceptor, by forming a Racl-IRSp53-WAVE2 complex, which recruits Arp2/3 
complex including profilin, thus evoking membrane ruffling (Mild et al., 
2000; Takenawa and Miki, 2000; Krugmann et al, 2001). Because WAVE2 
can be detected around the area of protruded membrane ruffles induced 
by Shigella (Suzuki, unpublished data), the Racl-IRSp53-WAVE2 complex 
might take part in the formation of membrane protrusion induced by Shigella 
invasion (Fig. 2.4). 



IpgD 

IpgD, a 69-kDa protein encoded by ipgD and located upstream of 
ipaBCDA, is secreted by the TTSS in amounts similar to the Ipa proteins. Like 
the Ipa proteins, IpgD is stored in the Shigella cytoplasm unless the TTSS 
is stimulated such as by incubation in conditioned medium (Niebuhr et al., 
2000). The storage of IpgD in the cytoplasm requires its association with 
a cytoplasmic chaperone, IpgE, encoded by the gene located immediately 
downstream of ipgD. Interestingly, after secretion, IpgD forms a complex 
with IpaA in the extracellular medium, although the biological significance 
is unclear. An ipgD mutant still enters host cells; however, in comparison 
with that directed by the wild type, the morphology of the membrane ruffle is 
altered. For example, scanning electron microscopic analysis reveals that the 




H 



< 
> 

CO 

o 



ipgD mutant provokes fewer actin rearrangements and less membrane ruf- 
fling on the target cell surface, suggesting that IpgD is involved in the process 
of invasion of epithelial cells and that the protein serves as a translocated ef- 
fector (Niebuhr et al., 2000). A protein homologous to IpgD, SopB (also called 
SigD), has been identified in S. duhlin, and it is involved in invasion, because 
a sigD mutant affected Salmonella invasion of CHO and HEp-2 cells (Galyov et 
al., 1997; Hong and Miller 1998). SopB has sequence homology with mam- 
malian inositol polyphosphate 4-phosphatase, and recombinant SopB pro- 
tein prepared from S. dublin shows inositol phosphate phosphatase activity 
required for promoting membrane fission during invasion (Norris et al., 
1998; Terebiznik et al., 2002). 

Similarly, the sequence of IpgD has been suggested to have the active 
site of mammalian inositol polyphosphate 4-phospatase, which specifically 
dephosphorylates PtdIns(4,5)P 2 into PtdIns(5)P (Niebuhr et al, 2000, 2002). g 

In mammalian cells this enzyme plays key roles in many processes, includ- £ 

ing reorganization of the actin cytoskeleton, and cytoskeleton-plasma mem- 
brane linkage. The importance of the enzymatic activity encoded by IpgD has 
been shown in promoting detachment of the plasma membrane from the cy- 
toskeleton to facilitate extension of membrane filopodia and ruffles evoked 
by Shigella invasion of epithelial cells. Indeed, continuous ectopic expression 
of IpgD in HeLa cells increases membrane detachment and causes formation 
of membrane blebs (Niebuhr et al., 2002). 

VirA 

An examination of the cytoskeletal architecture around invading Shigella 
by confocal microscopy indicated that the local microtubule network beneath 
the protruding ruffles undergoes remarkable destruction (Yoshida et al., 
2002). This finding, together with the increase in Shigella invasiveness in 
host cells treated with microtubule -de stabilizing agents such as nocodazole, 
suggests that the bacteria have the ability to modulate tubulin dynamics. 
VirA has activity to trigger microtubule dynamic instability in vitro and in 
vivo, which can stimulate Racl activity, thus leading to membrane ruffling 
(Fig. 2.4). 

VirA is a 45-kDa protein composed of 401 amino acids; it is able to 
bind a/3-tubulin dimers but not microtubules. In an in vitro tubulin poly- 
merization assay system, purified VirA showed activity to inhibit the poly- 
merization of tubulin and stimulate microtubule de stabilization. Interest- 
ingly, a portion of VirA, encompassing residues 224 to 315, involved in 
the interaction with tubulin heterodimers shares significant (>40%) amino 
acid homology with a portion of EspG encoded by the espG gene in the 




LEE of EPEC or enterohemorrhagic E. coli, as well with as some other un- 
characterized bacterial proteins such as NMB0928 (Neisseria menigitidis) or 
Cjl457c (Campylobacter jejuni) . Indeed, the expression of EspG in a Shigella 
virA mutant can rescue invasiveness, suggesting that EspG and VirA share 
an essential function (Elliott et al., 2001). The expression of VirA in mam- 
malian cells such as HeLa, COS-7, or Swiss3T3 cells allows for the forma- 
tion of membrane ruffling, though the scale of ruffles is smaller than that 
evoked by Shigella. Microinjection of VirA into HeLa cells also induces a 
localized membrane ruffling in a few minutes, whereas overexpression of 
VirA in host cells causes the destruction of microtubules and protruding 
membrane ruffles. Importantly, the VirA-induced membrane ruffling is de- 
pendent on the host Racl activity, because when VirA is coexpressed with a 
dominant negative Racl mutant in the cells, the appearance of ruffles can be 
g shut off. 

m In agreement with this, wild type S. jlexneri, but not the virA mutant, 

< stimulates Racl and induces the formation of membrane ruffles in infected 

§ HeLa cells (Yoshida et al., 2002). These observations suggest that the desta- 

i bilization of microtubules induced by VirA secreted from Shigella into host 

u cells can provoke the formation of membrane ruffles, thus stimulating bacte- 

rial entry (Fig. 2.4). Although it is still unclear whether or not other invasive 
bacteria are able to stimulate host microtubule dynamic instability, a simi- 
lar activity to Shigella VirA may also be found to be involved in some other 
bacterial infections of host cells. 

The microtubule network is dynamic in migrating or growing cells in 
which the microtubules undergo growth and shortening, called microtubule 
dynamic instability, which is mediated by various factors including micro- 
tubule stabilizing and destabilizing factors. For migrating cells, the interplay 
between the microtubule and actin cytoskeletal systems is though to be cru- 
cial. Indeed, recent studies have strongly indicated that microtubule growth 
and shortening participate in the activation of Racl and RhoA signaling, 
respectively, to control actin dynamics. 

Waterman- Storer et al. (1999) revealed that when microtubule growth 
in host cells is stimulated by pretreatmenting with nocodazole followed by 
washing out the drug, Racl is activitated, thus leading to the formation of 
lamellipodial protrusions in fibroblasts. Enomoto (1996) originally showed 
that microtubule disruption by colcemid or vinblastine, but not taxol, rapidly 
and reversibly induced the formation of actin stress fibers and focal adhe- 
sions, which was accompanied by activated cell motility. Consistent with 
that study, Krendel et al. (2002) have observed that RhoA activity in fibro- 
blasts can be stimulated by nocodazole. These studies have suggested that 
the depolymerization or shortening of microtubules can somehow trigger the 



stimulation of Rho activity, such as by releasing factors bound to the micro- 
tubules into the cytosol, and these factors are then required for activating Rho 
GTPases beneath the plasma membrane (Fig. 2.4). 

Of note, this notion has recently been supported by the finding of func- 
tional involvement of microtubules in regulating the Rho guanine nucleotide 
exchange factor GEF-H1 and Rho activity itself (Krendel et al., 2002). Al- 
though the association of microtubules and Racl or GEF-H1 has been in- 
dicated, the function of this interaction is only recently becoming clear. In- 
terestingly, recent studies have strongly indicated that Racl and RhoA have 
some functional linkage to each other, where the enhancement of one ac- 
tivity downregulates activity of the other. Therefore, the cross-talk between 
Racl and RhoA activities may account for the microtubule instability-induced 
membrane ruffling in mammalian cells as well as for the ruffling induced 
by Shigella VirA (Fig. 2.4). £ 



CELL-CELL SPREADING 

After escaping from phagocytic vacuoles, Shigella multiply and move 
within the cytoplasm. The ability of Shigella to move within the host cyto- 
plasm, and the subsequent cell-cell spreading, is a prerequisite for shigel- 
losis. Intracellular Shigella exploit actin polymerization at one pole of the 
bacterial surface, through which the bacterium gains a propulsive force to 
spread within the cytoplasm and into adjacent epithelial cells (Fig. 2.5). Under 
optimum conditions, intracellular bacterial motility is around 1 5-20 /xm/min 
(Mimuro et al., 2000). The actin-based motility of Shigella is dependent on 
VirG (IcsA) encoded by the virG gene on the large plasmid (Makino et al., 
1986; Bernardini et al, 1989; Lett et al, 1989). 

VirG 

VirG (IcsA) is a surface-exposed outer membrane protein, which accu- 
mulates at one pole of the bacterium (Fig. 2.6; also see Goldberg et al., 1993; 
Goldberg, 2001). VirG is composed of 1102 amino acids and contains three 
distinctive domains: the N-terminal signal sequence (residues 1-52), the 706- 
amino-acid a -domain (residues 53-758), and the 344-amino-acid C-terminal 
/3-core (residues 759-1102; see Goldberg et al., 1993; Suzuki et al., 1995). 
The a-domain is exposed on the surface of bacteria, whereas the /?-core is 
embedded in the outer membrane to form a membrane pore. The a -domain 
is translocated through the membrane pore onto the bacterial surface by a 
typical autotransporter mechanism as represented by the IgA protease of 
N. gonorrhoeae (Pohlner et al., 1987). 



H 



< 
> 

o 



'■•' ' *■ 






>' 



^-it^i 



— 







f" 



>>'> 









- 








.. 




B 




1*1 "• 




< 
< 

O 

i— i 

X 
i— i 

X 
u 






7 * : • ' ■ ' -V-^^l&ftV 1 ' 




. 




1 


^^^ 1 i M-V^C-' -V...., 




V ■■'/;■'* ■ - 




^w 




^k 


' 


XA 


'.*;•■- ■'•: 






V#- '■■ 


1 ■ -' -/". 




1 


..***?^-^ ...■■' 



Figure 2.5. Electron micrograph of motile Shigella in epithelial cells. (A) Bacteria escaping 
from the phagocytic vacuoles. (B) Multiplied bacteria within the host cell cytoplasm. (C) A 
motile bacterium forming a long actin tail. (D) A motile bacterium entering the neighbor 
host cell. 



The asymmetric distribution of VirG along the bacterial body is a pre- 
requisite for the polar movement of Shigella in mammalian cells, including 
bacterial spreading between epithelial cells (Goldberg et al., 1993; Suzuki 
et al., 1995; Goldberg, 2001). Although the mechanisms are still specula- 
tive, recently studies have suggested that the unipolar localization of VirG 
results from its direct targeting of the pole following diffusion laterally in 
the outer membrane (Goldberg, 2001; Charles et al., 2001; Robbins et al., 
2001). Interestingly, when a VirG-GFP fusion protein is expressed in E. coli, 
S. typhimurium, Yersinia pseudotuberculosis, or Vibrio cholerae, the protein al- 
ways localizes at one pole, suggesting that the mechanism of polar targeting 
for VirG is not unique to Shigella. Several factors including its own VirG a 
portion have been implicated in the establishment or maintenance of the 
asymmetric distribution (Suzuki et al., 1995; Steinhauer et al., 1999). The 
N -terminal two thirds of the VirG a -domain, which contains six glycine-rich 
repeats, is essential for mediating actin assembly, because the domain serves 



Shigella 



profilin 




▲ 



profilin 



O 



o 



G-actin 



direction of 
bacterial movement 



Arp2/3 



Figure 2.6. Current model for VirG-induced actin polymerization on Shigella in infected 
epithelial cells. VirG accumulated at one pole of bacterium recruits vinculin and N-WASP. 
The vinculin recruits profilin by means of binding to VASP; the N-WASP activated upon 
binding by Cdc42 allows recruitment of the Arp2/3 complex, with profilin. The activated 
Arp2/3 complex, with the aid of profilin, can catalyze rapid actin nucleation and 
elongation. 




ir, 
H 



< 
> 

o 



to interact with host proteins such as vinculin and neural WASP (N-WASP; 
see Suzuki et al., 1996, 1998; Suzuki and Sasakawa, 2001). The C-terminal 
one third of the a -domain is required for VirG to distribute asymmetrically, 
because S.flexneri expressing a VirG mutant with a deletion in this region no 
longer displays polar movement; instead, it is surrounded by an actin cloud 
(Suzuki etal, 1996). 

Interestingly, LPS plays a role in either the establishment or maintenance 
of VirG at one pole of Shigella (Okada et al., 1991). A number of genes 



involved in the biosynthesis of LPS have been shown to affect the localization 
of VirG (Rajakumar et al., 1994). Indeed, removal of the O-side chain from 
the LPS of S.flexneri results in an aberrant localization of VirG, causing a 
circumferential distribution over the whole bacterial body (Goldberg, 2001). 
SopA (also called IcsP), an outer membrane protease, has also been indicated 
to be involved in the asymmetric distribution of VirG by cleaving laterally 
diffused VirG protein along the bacterial body (Egile et al., 1997; Shere et al., 
1997). Finally, the absence of OmpT, an outer membrane protease encoded 
by the ompT gene in E. coli, is crucial for VirG to be maintained on the cell 
surface, because OmpT specifically cleaves at Arg758-Arg 7 59 of VirG, causing 
degradation of the a -domain of VirG on bacteria (Nakata et al., 1993). In fact, 
none of the Shigella and EIEC strains examined has the ompT region, thus 
ensuring the VirG a-domain is expressed and maintained on the bacterial 
g surface (Nakata et al., 1993). 




< 

< 

00 

< 

00 

O 

i— i 

X 
i— i 

X 
u 



VirG ligands 

VirG can interact with at least two host proteins, vinculin and N-WASP 
(Fig. 2.6). Vinculin, a protein linking focal adhesions and actin filaments, 
interacts directly with a portion of the VirG a -domain that spans residues 
103-508. As already mentioned (see the subsection on IpaA), the function 
of vinculin in mammalian cells is regulated by PtdIns(4,5)P2. In the inac- 
tive state, the N -terminal globular head domain interacts with the C -terminal 
elongated tail domain, and this interaction is disrupted by the binding of 
PtdIns(4,5)P2- The exposed head and tail domains become activated to inter- 
act with other molecules. In epithelial cells infected with Shigella, vinculin 
is recruited to the bacterial surface as well as to the actin comet tail elon- 
gated from motile bacteria in infected cells (Suzuki et al., 1996). Laine et al. 
(1997) revealed that the recruited vinculin is cleaved, leaving the head portion, 
which interacts with VirG along with vasodilator stimulating phosphoprotein 
(VASP) and pro film. Thus, the complex formed in the vicinity of the bac- 
terium seems to contribute to enhancing the growth of barbed ends of actin 
filaments. Microinjection of the vinculin head portion into Shigella-infected 
cells stimulates the bacterial motility (Laine et al., 1997). In fact, the speed at 
which E. coli expressing VirG induce formation of the actin tail in vinculin- 
depleted Xenopus egg extracts is shown to be significantly decreased to less 
than 30% of the original level (Suzuki, unpublished data). Thus, vinculin 
appears to contribute to Shigella-mducing actin assembly such as through 
interaction with VASP, because VASP recruits profilin (Fig. 2.6). Alterna- 
tively, existing actin filaments bound by vinculin at the bacterial surface may 



N-WASP 



WASP 



WAVE2 




Basic 



VCA 



Figure 2.7. Functional domains of N-WASP, WASP, and WAVE2 and molecules that 
interact with these proteins (Takenawa and Miki, 2000). 

facilitate actin nucleation mediated by the Arp2/3 complex interacting with 
the VirG-N-WASP complex. 




ir, 
H 



< 
> 

o 



N-WASP 

N-WASP is a member of the WASP and WAVE family, which includes 
human WASP, Saccharomyces cerevisiae WASP-like protein Lasl7p/Beelp, 
and WAVE1, WAVE2, and WAVE3 proteins (Takenawa and Miki, 2000). 
WASP and WAVE family proteins integrate upstream signaling events with 
changes in the actin cytoskeleton by means of the Arp2/3 complex. Fig- 
ure 2.7 shows the functional structures of N-WASP, WASP, and WAVE2, 
and molecules to which they bind (Takenawa and Miki, 2000). The expres- 
sion of WASP is limited to hematopoietic cells, whereas N-WASP and WAVEs 
are ubiquitously expressed in host cells including epithelial cells. N-WASP 
has been implicated both in the formation of filopodia and in the actin-based 
motility of intracellular Shigella, whereas WAVEs are involved in formation 
of lamellipodia and membrane ruffling. N-WASP and WASP possess several 
distinctive domains: a homology (WH1) domain that binds PtdIns(4,5)P 2 , a 
domain composed of basic amino acids, a GTPase binding domain (GBD) 



that binds Cdc42 , a proline-rich region ( P RR) , a G-actin-binding verprolin ho- 
mology (V) domain, a domain (C) with homology to the actin-depolymerizing 
protein cofilin, and finally a C-terminal acidic (A) segment (Fig. 2.7). The 
C-terminal VCA domain mediates the interaction with the Arp2/3 com- 
plex, by which the Arp2/3 complex is activated, thus mediating actin 
polymerization. 

In Shigella-infected cells, N-WASP, but not the other members of the 
WASP family, accumulates at the pole of the intracellular bacterium assem- 
bling an actin comet tail (Suzuki et al., 2002). Functional assays using the 
ectopic expression of dominant negative N-WASP in mammalian cells or 
immunodepletion in Xenopus egg extracts revealed that N-WASP is an essen- 
tial host component for mediating the actin-based motility of intracellular 
Shigella (Suzuki et al., 1998). Of note, none of the WASP family proteins as- 
g sociate with the surface of intracellular I. monocytogenes including the actin 




< 



i— i 

X 
u 



m tails (Suzuki et al., 2002). The binding of Shigella VirG to WASP family pro- 

< teins is limited to only N-WASP. With the use of a series of chimeras obtained 

§ by swapping the N-WASP and WASP domains, the specificity of VirG to in- 

teract with the N-terminal WH1 region of N-WASP was found to serve as 
the critical ligand (Suzuki et al., 2002). Consistent with this, hematopoietic 
cells such as J 774 cells (mouse macrophages), human monocytes, PMNs, or 
platelets express WASP predominantly but not N-WASP, which cannot sup- 
port the actin-based movement of intracellular S.flexneri (Egile et al., 1999; 
Suzuki et al., 2002). This was also confirmed by use of N-WASP-deficient 
embryos (Snapper et al., 2001). 



Cdc42 

In vitro studies have indicated that activation of N-WASP in cells re- 
quires the binding of Cdc42 to the GBD motif of N-WASP (Mild et al., 1996; 
Rohatgi et al., 1999, 2000). This binding inhibits the intramolecular interac- 
tion between the C-terminal acidic amino acids and the basic amino acids 
near the GBD, thus causing the unfolding of N-WASP into the activated form. 
Furthermore, when N-WASP interacts with a fragment of VirG encompass- 
ing residues 53-503 of VirG, the N-WASP-Arp2/3 complex-mediated actin 
nucleation can also be stimulated without Cdc42 in vitro (Egile et al., 1999). 
The ability of VirG to activate N-WASP without Cdc42 was also observed by 
using Clostridium difficile Tcd-10463, which inhibits Rho GTPases; however, 
the actin tails are significantly shorter in the presence of the exotoxin than 
in infected cells without the toxin, implying that actin assembly by Shigella 
is partly affected by toxin (Mounier et al., 1999). A later study, however, 
strongly indicated that cellular Cdc42 is required for the actin-based motility 
of Shigella in infected cells (Suzuki et al., 2000). Microinjection of activated 



Cdc42 accelerates Shigella motility, whereas inhibiting Cdc42 activity, for 
example by adding Rho GDI, a guanine nucleotide dissociation inhibitor, 
into cell extracts greatly reduces bacterial motility. In pyrene actin polymer- 
ization assays, VirG-N-WASP-Arp2/3 complex is insufficient to express the 
full activity for polymerizing actin; rather, in the presence of activated Cdc42, 
the actin nucleation activity is remarkably stimulated. In fact, Cdc42 can be 
accumulated at one pole of Shigella in the process of initiating movement 
in infected cells. Importantly, Cdc42 is not accumulated on motile Shigella 
possessing an actin tail in the infected cells, implying that Cdc42 seems to be 
no longer necessary after a steady speed has been reached, at which stage the 
VirG-N-WASP-Arp2/3 complexes would be constitutively activated. These 
studies have thus indicated that Cdc42 takes part in initiating the actin-based 
motility of intracellular Shigella in epithelial cells. 



Co 
H 



> 

CO 

o 



Arp2/3 

VirG expressed on the bacterial surface in host cells can directly recruit < 

and activate N-WASP, which in turn recruits and activates the Arp2/3 com- 
plex. Consequently, the VirG-N-WASP-Arp2/3 complex formed at one pole 
of the bacterium can mediate actin nucleation and elongation (Fig. 2.6). To 
initiate actin nucleation, the Arp2/3 complex is activated upon physical inter- 
action with the VCA region of N-WASP. With the aid of other host factors, the 
VirG-N-WASP-Arp2/3 complex mediates rapid actin filament growth at the 
barbed end, including cross-linking between the elongated actin filaments. 
In this way Shigella can gain a propulsive force in the host cytoplasm, where 
some motile bacteria impinge on the host plasma membrane, leading to the 
extension of membranous protrusions. Through these protrusions that pen- 
etrate neighboring cells, Shigella further move into adjacent epithelial cells by 
disrupting the double membranes with the secreted proteins (Suzuki et al., 
1994; Schuch et al., 1999). In a reconstitution experiment supporting the 
actin-based motility of Shigella with pure proteins, host factors required 
for Shigella movement were confirmed to include actin, Arp2/3 complex, 
and N-WASP (Loisel et al., 1999). In addition, actin depolymerization factor 
(ADF)/cofilin, capping protein, and profilin are also indicated to be involved 
in the regulation of actin turnover and stabilization of the actin tail (Loisel 
et al., 1999; Mimuro et al., 2000). Several other actin-associated proteins, such 
as plastin (fimbrin), filamin, VASP, zyxin, ezrin, CapZ, Nek, and WASP in- 
teracting protein (WIP), have also been identified as being localized to the 
actin tail or at the posterior end of intracellular bacteria. However, whether 
or not these host factors are functionally required for Shigella movement in 
infected cells awaits further study. 



Profilin 

Profilin binding to actin facilitates the formation of ATP-actin 
monomers, the form of actin to be assembled into the free barbed ends of 
actin filaments. Profilin can interact with various proteins, notably proteins 
with a proline-rich sequence, such as N-WASP, VASP, MENA, pl40mDia, 
WAVE/Scar, and Arp2/3 complex (Suzuki and Sasakawa, 2001). Inareconsti- 
tution assay in vitro, profilin and VASP (for Listeria) were shown to enhance 
bacterial motility but were not essential, suggesting that recruited profilin 
helps to increase the local concentration of ATP-actin (Loisel et al., 1999) . Pro- 
filin exists in two isoforms in mammalian cells, profilin I and II, with profilin 
I having a greater affinity for N-WASP (Kd = 60 nM) than profilin II (Kd = 400 
nM). Hence, the role of profilin I in the actin-based motility of intracellular 
Shigella has recently been investigated (Mimuro et al., 2000). Upon overex- 
| pression of a profilin H133S mutant defective in interaction with the PRR of 

gj N-WASP including poly-L-proline, Shigella motility is significantly decreased. 

w Similarly, the depletion of profilin from Xenopus egg extracts results in a de- 




o 



g crease in bacterial motility that is rescued by adding back profilin I but not 

£ H133S mutant. Consistent with this, on overexpression of an N-WASP mu- 

tant lacking the PRR unable to interact with profilin, the actin tail formation of 
intracellular Shigella is abolished. In N-WASP-depleted extracts, the addition 
of wild type N-WASP but not the N-WASP mutant restores bacterial motility, 
indicating that profilin associated with N-WASP is an essential host factor for 
supporting rapid movement of intracellular Shigella (Mimuro et al., 2000). 



CONCLUSION 

Invasion of epithelial cells by Shigella is a highly dynamic cellular event 
that occurs through the complicated interaction between bacterial effectors 
and target host factors. Shigella-directed internalization requires large-scale 
membrane ruffling, which eventually leads to a phagocytic event. In pro- 
moting the host cellular events including the subsequent infectious steps, 
the roles of effectors such as IpaA, IpaB, IpaC, IpgD, and VirA delivered 
by means of the TTSS are crucial. However, many other putative effectors 
secreted by the TTSS (Buchrieser et al., 2000) must also be involved in al- 
most the entire stage of infection, including modulation of host immune 
responses. Furthermore, it is thought that targeting of the host factors by 
bacterial effectors during infection would appropriately be operated by the 
pathogen, for which the timing and amounts of effectors to be secreted by 
means of the TTSS may be stringently controlled at a posttranslational or 



cotranslational level, depending on the stage of Shigella infection (Blocker et 
al., 2003) . Clearly, we must await further study to elucidate the role of all of the 
effectors in each stage of Shigella infection, including the secretion control 
system. Such information is needed for better understanding of the sophis- 
ticated bacterial infectious strategy and host inflammatory responses, which 
are prerequisite for the development of both a novel safer Shigella vaccine 
and a suitable animal model to study the disease. 



REFERENCES 

Adam, T., Giry, M., Boquet, P., and Sansonetti, P.J. (1996). Rho-dependent mem- 
brane folding causes Shigella entry into epithelial cells. EMBO J. 15, 3315— 
3321. 

Adler, B., Sasakawa, C, Okada, N., Makino, S., and Yoshikawa, M. (1989). A 
dual transcriptional activation system for the 230 kb plasmid genes coding 
for virulence-associated antigens of Shigella flexneri. Mol. Microbiol. 3, 627- 
635. 

Bernardini, M.L., Mounier, J., d'Hauteville, H., Coquis-Randon, M., and San- 
sonetti, P.J. (1989). Identification of icsA, a plasmid locus of Shigella flexneri 
that governs bacterial intra- and intercellular spreading through interaction 
with F-actin. Proc. Natl. Acad. Sci. USA 86, 3867-3871. 

Blocker, A., Gounon, P., Larquest, K., Neibuhr, V., Cabiaux, C, Parsot, C, and 
Sansonetti, P.J. (1999). The tripartite type III secretion of Shigella flexneri 
inserts IpaB and IpaC into host membranes. J. Cell Biol. 147, 683-693. 

Blocker, A., Komoriyama, K., and Aizawa, S. (2003). Type III secretion systems 
and bacterial flagella: insights into their function from structural similarities. 
Proc. Natl. Acad. Sci. USA 100, 3027-3030. 

Bourdet-Sicard, R., Rudiger, M., Jockusch, B.M., Gounon, P., Sansonetti, P. J., and 
Tran van Nhieu, G. (1999). Binding of the Shigella protein IpaA to vinculin 
induces F-actin depolymerization. EMBO J. 18, 5853-5862. 




ACKNOWLEDGMENTS 

I am grateful to Drs. Toshihiko Suzuki and Reiko Akakura for their 
critical review of the manuscript and to all of the laboratory members who 
contributed to the project. Original research in the author's laboratory is 

supported by a Grant-in Aid for Scientific Research on Priority Areas entitled 5 

o 

on "Infection and Host Responses" from the Japanese Ministry of Education, 5 

It* 

Science, Technology, Sport and Culture. 



m 



< 
> 

o 



Buchrieser, C, Glaser, P.P., Rusniok, C, Nedjari, H., d'Hauteville, H., Kunst, 
F., Sansonetti, P.J., and Parsot, P.J. (2000). The virulence plasmid pRWIOO 
and the repertoire of proteins secreted by the type III secretion apparatus of 
Shigella flexneri. Mol. Microbiol. 38, 760-771. 

Charles, M., Perez, M., Kobil, J.H., and Goldberg, M.B. (2001). Polar targeting 
of Shigella virulence factor IcsA in Enterobacteriacae and Vibrio. Proc. Natl. 
Acad. Sci. USA 98, 9871-9876. 

Collazo, CM. and Galan, J.E. (1996). Requirement of exported proteins for se- 
cretion through the invasion-associated Type III system in Salmonella ty- 
phimurium. Infect. Immun. 64, 3524-3531. 

Cornelis, G.R. and Van Gijsegem, F. (2000). Assembly and function of type III 
secretory systems. Anna. Rev. Microbiol. 30, 47-56. 

Daniell, S.H., Takahashi, N., Wilson, R., Friedberg, D., Rosenshine, I., Boody, 
| F.P., Shaw, R.K., Knutton, S., Frankel, G., and Aizawa, S. (2001). The fila- 




< 



m mentous type III secretion translocon of enteropathogenic Escherichia coli. 

< Cell. Microbiol. 3, 865-871. 



< 

GO 

O 



« Dehio, C., Prevost, M.C., and Sansonetti, P.J. (1995). Invasion of epithelial cells by 



i— i 

X 
u 



:— src 



Shigella flexneri induces tyrosine phosphorylation of cortactin by a pp60 c ~ 
mediated signalling pathway. EMBOJ. 14, 2471-2482. 

Dorman, C.J. and Porter, M.E. (1998). The Shigella virulence gene regulatory 
cascade: a paradigm of bacterial gene control mechanisms. Mol. Microbiol. 
29, 677-684. 

Dumenil G., Olivo, J.C., Pellegrini, S., Fellous, M., Sansonetti, P. J., and Tran van 
Nhieu, G. (1998). Interferon a; inhibits a Src-mediated pathway necessary for 
Shigella-induced cytoskeletal rearrangements in epithelial cells. J. Cell Biol. 
143, 1003-1012. 

Egile, C, d'Hauteville, H., Parsot, C, and Sansonetti, P.J. (1997). SopA, the outer 
membrane protease responsible for polar localization of IcsA in Shigella 
flexneri. Mol. Microbiol. 23, 1063-1073. 

Egile, C, Loisel, T.P., Laurent, V., Li, R., Pantaloni, D., Sansonetti, P. J., and ear- 
lier, M-F. (1999). Activation of the CDC42 effector N-WASP by the Shigella 
flexneri IcsA protein promotes actin nucleation by Arp2/3 complex and bac- 
terial actin-based motility. J. Cell Biol. 146, 1319-1332. 

Elliott, S.J., Krejany, E.O., Mellies, J.L., Robins-Browne, R.M., Sasakawa, C, and 
Kaper, J.B. (2001). EspG a novel type III system-secreted protein from en- 
teropathogenic Escherichia coli with similarities to VirA of Shigella flexneri. 
Infect. Immun. 69, 4027-4033. 

Enomoto, T. (1996). Microtuble disruption induces the formation of actin stress 
fibers and focal adhesions in cultured cells: possible involvement of Rho 
signal cascade. Cell Struc. Func. 5, 317-326. 




H 



< 
> 

CO 

o 



Galyov, E.E., Wood, M.W., Rosquist, R., Mullan, P.B., Watson, P.R., Hedges, 
S., and Wallis, T.S. (1997). A secreted effector protein of Salmonella duhlin 
is translocated into eukaryotic cells and mediates inflammation and fluid 
secretion in infected ileal mucosa. Mol. Microbiol. 25, 903-912. 

Girardin, S.E., Tournebize, R., Mavris, M., Page, A-L., Li, X., Stark, G.R., Bertin, 
J., DiStefano, P.S., Yaniv, M., Sansonetti, P.J., and Philpott, D.J. (2001). 
CARD/Nodl mediates NF-kB and JNK activation by invasive Shigella jlexneri. 
EM BO Reports 21, 736-742. 

Goldberg, M.B. (2001). Actin-based motility of intracellular microbial pathogens. 
Microbiol. Mol. Biol. Rev. 65, 595-626. 

Goldberg, M.B., Barzu, O., Parsot, C, and Sansonetti, P.J. (1993). Unipolar lo- 
calization and ATPase activity of IcsA, a Shigella jlexneri protein involved in 
intracellular movement. J. Bacteriol. 175, 2189-2196. 

Hakansson, S., Schesser, K., Person, C, Galyov, E.E., Rosqvist, R., Homble, F., £ 

and Walf-Watz, H. (1996). The YopB protein of Yersinia pseudotuberculosis is £ 

essential for the translocation of Yop effector proteins across the target cell 
plasma membrane and displays a contact-dependent membrane disrupting 
activity. EMBOJ. 15, 5812-5823. 

Hirao, M., Sato, N., Kondo, T., Yonemura, S., Monden, M., Sasaki, T., 
Takai, Y., and Tsukita, S. (1996). Regulation mechanisms of ERM 
(ezrin/radixin/moesin) protein/plasma membrane association: possible in- 
volvement of phosphatidylinositol turnover and Rho -dependent signaling 
pathway. J. Cell Biol. 135, 37-51. 

Hong, K.H. and Miller, V.L. (1998). Identification of a novel Salmonella invasion 
locus homologous to Shigella ipgDE. J. Bacteriol. 180, 1793-1802. 

Hueck, C.J. (1998). Type III protein secretion systems bacterial pathogens of 
animal and plants. Microbiol. Mol. Biol. Rev. 62, 379-433. 

Inohara, N., Ogura, Y., Chen, F.F., Muto, A., and Nunez, G. (2001). Human Nodi 
confers responsiveness to bacterial lipopolysaccharides. J. Biol. Chem. 276, 
2551-2554. 

Jockusch, B.M. and Rudiger, M. (1996). Crosstalk between cell adhesion 
molecules: vinculin as a paradigm for regulation by conformation. Trends 
Cell Biol. 6,311-315. 

Knutton, S., Rosenshine, I., Pallen, M.J., Nisan, I., Neves, B.C., Bain, C, Wolf, 
C, Dougan, G., and Frankel, G. (1998). A novel EspA-associated surface or- 
ganelle of enteropathogenic Escherichia coli involved in protein translocation 
into epithelial cells. EMBOJ. 17, 2166-2176. 

Krendel, M., Zenke, F.T., and Bokoch, G.M. (2002). Nucleotide exchange factor 
GEF-H1 mediates cross-talk between microtubles and the actin cytoskeleton. 
Nat. Cell Biol. 4, 294-301. 



Kubori, T., Matsushima, Y., Nakamura, D., Uralil, J., Lara-Tejero, M., Sukhan, A., 
Galan, J.E., and Aizawa, S. (1998). Supramolecular structure of the Salmonella 
typhimurium type III protein secretion system. Science 280, 602-605. 

Kubori, T., Shkan, A., Aizawa, S., and Galan, J.E. (2000). Molecular characteri- 
zation and assembly of the needle complex of the Salmonella typhimurium 
type III protein secretion system. Proc. Natl. Acad. Sci. USA 97, 10,225- 
10,230. 

Krugmann, S.K., Jordens, I., Gevaert, K., Driessens, M., Vandekerckhove, J., and 
Hall, A. (2001). Cdc42 induces filopodia by promoting the formation of an 
IRSp53:Mena complex. Curr. Biol. 11, 1645-1655. 

Kuwae, A., Yoshida, S., Tamano, K., Mimuro, H., Suzuki, T., and Sasakawa, C. 
(2001). Shigella invasion of macrophage requires the insertion of IpaC into 
the host plasma membrane. J. Biol. Chem. 276, 32,230-32,239. 
g Lafont, F., Tran van Nhieu, G., Hanada, K., Sansonetti, P.J., and Gisou van der 




< 



m Goot, F. (2002). Initial steps of Shigella infection depend on the cholesterol/ 



< 

GO 



< spingolipid raft-mediated CD44-IpaB interaction. EM BO J. 21, 4449- 



8 4457. 

i— i 

x 
i— i 

X 
u 



Laine, R.O., Zeile, W., Kang, F., Purich, D.L., and Southwick, F.S. (1997). Vinculin 

proteolysis unmasks an ActA homolog for actin-based Shigella motility. J. Cell 

Biol. 138, 1255-1264. 
Lan, R. and Reeves, P.R. (2002). Escherichia coli in disguise: molecular origins of 

Shigella. Microb. Infect. 4, 1125-1132. 
Lett, M-C., Sasakawa, C., Okada, N., Sakai, T., Makino, S., Yamada, M., Komatsu, 

K., and Yoshikawa, M. (1989). virG, a plasmid-coded virulence gene of Shigella 

flexneri: identification of the virG protein and determination of the complete 

coding sequence. J. Bacteriol. 171, 353-359. 
Loisel, T.P., Boujemaa, R., Pantaloni, D., and Carlier, M-F. (1999). Reconstitution 

of actin-based motility of Listeria and Shigella using pure proteins. Nature 401 , 

613-616. 
Makino S., Sasakawa, C., Kamata, T., Kurata, T., and Yoshikawa, M. (1986). A 

genetic determinant required for continuous reinfection of adjacent cells on 

a large plasmid in Shigella flexneri 2a. Cell 46, 551-555. 
Mavris, M., Page, A.L., Tournebize, R., Demers, B., Sansonetti, P. J., and Parsot, 

C. (2002). Regulation of transcription by the activity of the Shigella flexneri 

type III secretion apparatus. Mol. Microbiol. 43, 1543-1553. 
Menard, R., Prevost, M.C., Gounon, P., Sansonetti, P. J., and Dehio, C. (1996). 

The secreted Ipa complex of Shigella flexneri promotes entry into mammalian 

cells. Proc. Natl. Acad. Sci. USA 93, 1254-1258. 
Menard, R., Sansonetti, P.J., and Parsot, C. (1994a). The secretion of the Shigella 

flexneri Ipa invasins is induced by the epithelial cells and controlled by IpaB 

and IpaD. EM BO J. 13, 5293-5302. 



> 

CO 

O 



Menard, R., Sansonetti, P. J., Parsot, C, and Vasselon, T. (1994b). Extracellular 
association and cytoplasmic partioning of the IpaB and IpaC invasins of 
Shigella jlexneri. Cell 79, 515-525. 

Mild, H., Miura, K., and Takenawa, T. (1996). N-WASP, a novel actin- 
depolymerizing protein, regulate the cortical cytoskeletal rearrangement in a 
PIP2-dependent manner downstream of tyrosine kinase. EM BO J. 15, 5326- 
5335. 

Mild, H., Yamaguchi, H., Suetsugu, S., and Takenawa, T. (2000). IRSp53 is an es- 
sential intermediate between Rac and WAVE in the regulation of membrane 
ruffling. Nature 408, 732-735. 

Mimuro, H., Susuki, T., Suetsugu, S., Mild, H., Takenawa, T., and Sasakawa, 
C. (2000). Profilin is required for sustaining efficient intra- and intercellular 
spreading of Shigella jlexneri. J. Biol. Chem. 275, 28,893-28,901. 

Mounier, J., Laurent, V., Hall, A., Fort, P., Calier, M-F., Sansonetti, P. J., and Egile, £ 

C. (1999). Rho family GTPases control entry of Shigella Jlexneri into epithelial S 

cells but not intracellular motility. J. Cell Sci. 112, 2069-2080. > 

Mounier, J., Vasselon, T., Hellio, R., Lesourd, M., and Sansonetti, P.J. (1992). <j 

Shigella jlexneri enters human colonic Caco-2 epithelial cells through the 
basolateral pole. Infect. Immun. 60, 237-248. 

Nakata, N., Tobe, T., Fukuda, I., Suzuki, T., Komatsu, K., Yoshikawa, M., and 
Sasakawa, C. (1993). The absence of surface protease, OmpT, determines 
the intercellular spreading ability of Shigella: the relationship between the 
ompT and kcpA loci. Mol. Microbiol. 9, 459-468. 

Niebuhr, K., Jouihri, N., Allaoui, A., Gounon, P., Sansonetti, P.J., and Parsot, C. 
(2000). IpgD, a protein secreted by the type III secretion machinery of Shigella 
jlexneri, is chaperoned by IpgE and implicated in entry focus formation. Mol. 
Microbiol. 38, 8-19. 

Niebuhr, K., Giuriato, S., Pedron, T., Philpott, D.J., Gaits, F., Sable, J., Sheetz, 
M.P., Parsot, C., Sansonetti, P. J., and Payrastre, B. (2002). Conversion of 
PtdIns(4,5)P2 into PtdIns(5)P by the S. jlexneri effector IpgD reorganizes 
host cell morphology. EMBOJ. 21, 5069-5078. 

Norris, F.A., Wilson, M.P., Wallis, T.S., Galyov, E.E., and Majerus, P.W. (1998). 
SopB, a protein required for virulence of Salmonella dublin, is an inositol 
phosphate phosphatase. Proc. Natl. Acad. Sci. USA 95, 14,057-14,059. 

Okada, N., Sasakawa C, Tobe, T., Yamada, M., Nagai, S., Talkder, K., Komatsu, K., 
Kanegasaki, S., and Yoshikawa, M. (1991). Virulence-associated chromoso- 
mal loci of Shigella jlexneri identified by random Tn5 insertion mutagenesis. 
Mol. Microbiol. 5, 887-893. 

Page, A-L., Sansonetti, P. J., and Parsot, C. (2002). Spal5 of Shigella jlexneri, a 
third type of chaperone in the type III secretion pathway. Mol. Microbiol. 43, 
1533-1542. 




Piano G.V., Day, J.B., and Ferracci, F. (2001). Type III export: new uses for an old 
pathway. Mol. Microbiol. 40, 284-293. 

Pohlner, J., Halter, K., Beyreuther, K., and Meyer, T.F. (1987). Gene structure 
and extracellular secretion of Neisseria gonorrhoeae IgA protease. Nature 325, 
458-462. 

Pupo, G.M., Lan, R., and Reeves, P.R. (2000). Multiple independent origins of 
Shigella clones of Escherichia coli and convergent evolution of many of their 
characteristics. Proc. Natl Acad. Sci. USA 97, 10,567-10,572. 

Rajakumar, R., Jost, B.H., Sasakawa, C., Okada, N., Yoshikawa, M., and Adler, B. 
(1994). Nucleotide sequence of the rhamnose biosynthetic operon of Shigella 
flexneri 2a and role of lipopolysacchride in virulence. J. Bacteriol. 176, 2364- 
2373. 

Renesto, P., Mounier, J., and Sansonetti, P.J. (1996). Induction of adherence and 
g degranulation of polymorphonuclear leukocytes: A new expression of the 

m invasive phenotype of Shigella flexneri. Infect. Immun. 64, 719-723. 

< Robbins, J.R., Monack, D., McCallum, S.J., Vegas, A., Pham, E., Goldberg, M.B., 

« and Theriot, J. A. (2001). The making of a gradient: IcsA (VirG) polarity in 

Shigella flexneri. Mol. Microbiol. 41, 861-872. 
u Rohatgi, R., Ho, H.Y., and Kirschner. M.W. (2000). Mechanisms of N-WASP 

activation by CDC42 and phosphatidylinositol 4,5-bisphosphate. J. Cell Biol. 
150, 1299-1310. 

Rohatgi, R., Ma, H., Mild, H., Lopez, M., Kirchhausen, T., Takenawa, T., and 
Kirschner, M.W. (1999). The interaction between N-WASP and the Arp2/3 
complex links Cdc42-dependent signals to actin assembly. Cell 97, 221- 
231. 

Sansonetti, P.J., Arondel, J., Fountaine, A., D'Hauteville, H., and Bernardini, L. 
(1991). ompB (osmo-regulation) and icsA (cell to cell spreading) mutants of 
Shigella flexneri: vaccine candidates and probes to study the pathogenesis of 
shigellosis. Vaccine 9, 416-422. 

Sansonetti, P. J., Arondel, J.R., Prevost, M.C., and Huerre, M. (1996). Infection 
of rabbit Peyer's patches by Shigella flexneri: effect of adhesive or invasive 
bacterial phenotypes on follicle-associated epithelium. Infect. Immun. 64, 
2752-2764. 

Sansonetti, P. J., Ryter, A., Clerc, P., Maurelli, A.T., and Mounier, J. (1986). Multi- 
plication of Shigella flexneri within HeLa cells: lysis of the phagocytic vacuole 
and plasmid-mediated contact hemolysis. Infect. Immun. 51, 461-469. 

Sasakawa, C., Kamata, K., Sakai, T., Makino, S., Yamada, H., Okada, N., and 
Yoshikawa, M. (1988). Virulence-associated genetic regions comprising 31 
kilobases of the 230-kilobase plasmid in Shigella flexneri 2a. J. Bacteriol. 170, 
2480-2484. 




> 

CO 

O 



Schuch, R., Sandlin, R.C., and Maurelli, A.T. (1999). A system for identifying 
post-invasion functions of invasion genes: requirements for the Mxi-Spa 
type III secretion pathway of Shigella flexneri in intercellular dissemination. 
Mol. Microbiol. 34, 675-689. 

Sekiya, K., Ohishi, M., Ogino, T., Tamano, K., Sasakawa, C, and Abe, A. (2001). 
Supermolecular structure of the enteropathogenic Escherichia coli type III se- 
cretion system and its direct interaction with the EspA-sheath-like structure. 
Proc. Natl. Acad. Sci. USA 98, 11,638-11,643. 

Shere, K.D., Sallustion, S., Manessis, A., D'Aversa, T.G., and Goldberg, M.B. 
(1997). Distribution of IcsP, the major Shigella protease that cleaves IcsA, 
accelerates actin-based motility. Mol. Microbiol. 25, 451-462. 

Skoudy, A., Mounier, J., Aruffo, A., Ohayon, H., Gounon, P., Sansonetti, P. J., 
and Tran van Nhieu, G. (2000). CD44 binds to the Shigella IpaB protein and 
participates in bacterial invasion of epithelial cells. Cell. Microbiol. 2, 19-33. £ 

Skoudy, A., Tran van Nhieu, G., Mantis, N., Aprin, M., Mounier, }., Gounon, P., S 

and Sansonetti, P.J. (1999). A functional role for ezrin during Shigella flexneri > 

entry into epithelial cells. J. Cell Sci. Ill, 2059-2068. < 

Snapper, S.B., Takeshima, F., Anton, I., Liu, C.H., Thomas, S.M., Nguyen, D., 
Dudley D., Fraser, H., Purich, D., Lopez-Llasaca, M., Klein, C., Davidson, L., 
Bronson, R., Mulligan, R., Southwick F., Geha, R., Goldberg, M.B., Rosen, 
F.S., Hartwig, J.H., and Alt, F.W. (2001). N-WASP deficiency reveals distinct 
pathways for cell surface projections and microbial actin-based motility. Nat. 
Cell Biol. 3, 897-904. 

Steinhauer, J., Agha, R., Andrew, T.P., Varga, W., and Goldberg, B. (1999). The 
nuipolar Shigella surface protein IcsA is targeted directly to the bacterial 
old pole: IcsP cleavage of IcsA occurs over the entire bacterial surface. Mol. 
Microbiol. 32, 367-377. 

Suzuki, T., Lett, M-C., and Sasakawa, C. (1995). Extracellular transport of VirG 
protein in Shigella. J. Biol. Chem. 270, 30,874-30,880. 

Suzuki, T., Mild, T., Takenawa, T., and Sasakawa, C. (1998). Neural Wiskott- 
Aldrich syndrome protein is implicated in actin-based motility of Shigella 
flexneri. EMBOf. 17, 2767-2776. 

Suzuki, T., Mimuro, H., Suetsugu, S., Miki, H., Takenawa, T., and Sasakawa, C. 
(2002). Neural Wiskott-Aldrich syndrome protein (N-WASP) is the specific 
ligand for Shigella VirG among the WASP family and determines the host 
cell type allowing actin-based spreading. Cell. Microbiol. 4, 223-233. 

Suzuki, T., Mimuro, H., Miki, H., Takenawa, T., Sasaki., Nakanishi, H., Takai, 
Y., and Sasakawa, C. (2000). Rho family GTPase Cdc42 is essential for the 
actin-based motility of Shigella in mammalian cells. ]. Exp. Med. 191, 1905- 
1920. 



Suzuki, T., Murai, T., Fukuda, I., Tobe, T., Yoshikawa, M., and Sasakawa, C. 
(1994). Identification and characterization of a chromosomal virulence gene, 
vac], required for intercellular spreading of Shigella flexneri. Mol. Microbiol. 
11, 31-41. 

Suzuki, T., Saga, S., and Sasakawa, C. (1996). Functional analysis of Shigella VirG 
domains essential for interaction with vinculin and actin-based motility. J. 
Biol. Chem. 271, 21,878-21,885. 

Suzuki, T. and Sasakawa, C. (2001). Molecular basis of the intracellular spreading 
of Shigella. Infect. Immun. 69, 5959-5966. 

Tamano, K., Aizawa, S., Katayama, E., Nonaka, T., Imajo-Ohmi, S., Kuwae, A., 
Nagai, S., and Sasakawa, C. (2000). Supramolecular structure of the Shigella 
type III secretion machinery: the needle part is changeable in length and 
essential for delivery of effectors. EMBOJ. 19, 3876-3887. 
g Tamano, K., Eisaku, K., Toyotome, T., and Sasakawa, C. (2002). Shigella Spa32 




< 



x 

u 



m is an essential secretory protein for functional type III secretion machinery 

< and uniformity of its needle length. J. Bacteriol. 184, 1244-1252. 

S Takenawa, T. and Mild, H. (2000). WASP and WAVE family proteins: key 

^ molecules for rapid rearrangment of cortical actin filaments and cell move- 

ment./. Cell Sci. 114, 1801-1809. 

Terebiznik, M.R., Vieira, O.V., Marcus, S.L., Slade, A., Yip, CM., Trimble, W.S., 
Meyer, T., Finlay, B.B., and Grinstein, S. (2002). Elimination of host cell 
Ptdlns(4, 5)P 2 by bacterial SigD promotes membrane fission during invasion 
by Salmonella. Nature Cell Biol. 4, 766-773. 

Tran van Nhieu, G., Caron, E., Hall, A., and Sansonetti, P.J. (1999). IpaC induces 
actin polymerization and filopodia formation during Shigella entry into ep- 
ithelial cells. EMBOJ. 18, 3249-3262. 

Tran van Nhieu, G. and Sansonetti, P.J. (1999). Mechanism of Shigella entry into 
epithelial cells. Curr. Opin. Microbiol. 2, 51-55. 

Van Gijsegem, F., Vasse, J., Camus, J-C., Marenda, M., and Boucher, C. (2000). 
Ralstonia solanacearum produces Hrp-dependent pili that are required for 
PopA secretion but not for attachment of bacteria to plant cells. Mol. Microbiol. 
36, 249-260. 

Venkatesan, M.M., Goldberg, M.B., Rose, D.J., Grotbeck, E.J., Burland, V., and 
Blattner, F.R. (2001). Complete DNA sequence and analysis of the large 
virulence plasmid of Shigella flexneri. Infect. Immun. 69, 3271-3285. 

Wassef, J.S., Keren, D.F., and Mailloux, J.L. (1989). Role of M cells in initial 
antigen uptake and in ulcer formation in the rabbit intestinal loop model of 
shigellosis. Infect. Immun. 57, 858-863. 

Watarai, M., Funato, S., and Sasakawa, C. (1996). Interaction of Ipa proteins 
of Shigella flexneri with aspi integrin promotes entry of the bacteria into 
mammalian cells. J. Exp. Med. 183, 991-999. 




Watarai, M., Kamata, Y., Kozaki, S., and Sasakawa, C. (1997). Rho, a small GTP- 
binding protein, is essential for Shigella invasion of epithelial cells. J. Exp. 
Med. 185, 281-292. 

Waterman- Storer, CM., Wothylake, R.A., Liu, B.P., Burridge, K., and Salmon, 
E.D. (1999). Microtubule growth activates Racl to promote lamellipodial pro- 
trusion in fibroblasts. Nat. Cell Biol. 1, 45-50. 

Uchiya, K., Tobe, T., Komatsu, K., Suzuki, T., Watarai, M., Fukuda, I., Yoshikawa, 
M., and Sasakawa, C. (1995). Identification of a novel virulence gene, virA, on 
the large plasmid of Shigella, involved in invasion and intercellular spreading. 
Mol. Microbiol. 17, 241-250. 

Yoshida, S., Katayama, E., Kuwae, A., Mimuro, H., Suzuki, T., and Sasakawa, C. 
(2002) . Shigella deliver an effector protein to trigger host microtubule destabi- 
lization, which promotes Racl activity and efficient bacterial internalization. 
EMBOJ. 21,2923-2935. £ 

Zychlinsky, A., Fitting, C, Cavaillon, J.M., and Sansonetti, P.J. (1994). Intereukin S 

1 is released by murine macrophages during apoptosis induced by Shigella > 

flexneri.]. Clin. Invest. 94, 1328-1332. < 

Zychlinsky, A., Prevost, M.C., and Sansonetti, P.J. (1992). Shigella flexneri induces 
apoptosis in infected macrophages. Nature 358, 167-169. 

Zychlinsky, A. and Sansonetti, P.J. (1997). Apoptosis as a proinflammatory event: 
what can we learn from bacteria-induced cell death? Trends Microbiol. 5, 201- 
204. 

Zychlinsky, A., Thirumalai, K., Arondel, J., Cantey, J.R., Aliprantis, A.O., and 
Saonsonetti, P.J. (1996). In vivo apoptosis in Shigella flexneri infection. Infect. 
Immun. 64, 5357-5365. 



m 



> 

CO 

O 



CHAPTER 3 

How Yersinia escapes the host: To Yop 
or not to Yop 

Geertrui Denecker and Guy R. Cornells 



The genus Yersinia contains three species of Gram-negative bacteria that are 
pathogenic for humans: Y. pestis, the agent of bubonic plague; Y. pseudotu- 
berculosis, causing mesenteric adenitis and septicemia; and Y. enter ocolitica, 
causing gastrointestinal syndromes (enteritis and mesenteric lymphadeni- 
tis) . Bacteria from these three species have a tropism for lymphoid tissues and 
share the common capacity to resist the innate immune response. Whereas 
Y. pestis is generally inoculated by a fleabite or aerosol, Y. enterocolitica and 
Y. pseudotuberculosis are foodborne pathogens, which gain access to the under- 
lying lymphoid tissue (e.g., Peyer's patches) of the intestinal mucosa through 
M cells (Fig. 3.1; see Autenrieth and Firsching, 1996; Perry and Fetherston, 
1997). Once Yersinia has entered the lymphoid system, it overcomes the pri- 
mary immune response of the host by using the type III secretion system 
(TTSS) (Cornells et al., 1998; Cornells, 2002). TTSS is a sophisticated viru- 
lence mechanism by which Gram-negative pathogens inject effector proteins 
directly into host cells. Currently, more than 20 different TTSSs have been 
described in animal, plant, and even insect pathogens (Hueck, 1998; Galan 
and Collmer, 1999; Cornells, 2000; Buttner and Bonas, 2002). 

Depending on the effectors injected, the employment of the TTSS will 
have a different outcome. Some, like the Mxi-Spa system of Shigella Jlexneri 
or the Salmonella pathogenicity island 1 (SPI-1) system of Salmonella enterica, 
make use of the innate immune system of the host to enhance the proinflam- 
matory response and to trigger phagocytosis by normally nonphagocytic cells, 
whereas others, such as the pathogenic Yersinia Ysc-Yop system, essentially 
paralyze the innate immune response of the host (Galan, 2001; Sansonetti, 
2001; Cornells, 2002; Juris et al., 2002). The Yersinia TTSS becomes acti- 
vated upon contact with eukaryotic cells and directs effector proteins - called 
Yops - over the bacterial membranes. Some of the Yops form a kind of 




Y. pestis 




CO 

I— I 

1-1 
w 

z 

O 
U 



Q 

< 

w 
W 
U 

w 

w 
Q 
i— i 

Pi 

w 
w 
O 



V. enterocolitica 

Y. pseudotuberculosis 



Contaminated food 



Lymph\ 
node( ^) 




Epithelial cells M-cel 

\W\A/\W\AA Q. 



O 



O 



CD 



(WWTWWV 



o 



^ 



> . 



o 



o 



Antiphagocytosis: 
YopE, -T, -0, -H 




Macrophages 



Local and 

systemic dissemination 




Macrophage apoptosis, 
Inhibition of proinflammatory 
response 
3 YopP, -H 



Figure 3.1. Model showing the interaction of Yersinia with an eukaryotic cell. 
At 37° C and upon contact of Yersinia with its eukaryotic target cell, the adhesins YadA or 
Inv interact with the ^i-integrins and other extracellular matrix proteins at the cell surface, 
and Yersinia attaches tightly to the cell membrane. The Ysc injectisome is installed and 
the Yop translocators and effectors, some of which are intrabacterially capped with a 
chaperone, are transported through the bacterial inner and outer membranes by the Ysc 
injectisome. The translocators Yops, YopB, YopD, and LcrV form a pore in the target cell 
membrane, through which the effector Yops are translocated into the cell cytosol. Four 
effector Yops with different enzymatic functions (YopE, a Rho GTPase-activating protein, 
YopT, a cysteine protease, YopO/YpkA, a serine/threonine kinase, and YopH, a pro- 
tein tyrosine phosphatase) will cooperatively lead to the destruction of the actin cytoskeleton, 

(cont.) 



translocation pore in the eukaryotic target cell membrane, whereas the other 
Yops are effector proteins that are delivered through this pore into the cy- 
tosol of the target cell. At least six different Yop effectors are injected by 
the secretion translocation apparatus, five of which have been shown to play 
an important role in the defense against the innate immune response (Fig. 
3.2). In this chapter we discuss how Yersinia escapes the host immune re- 
sponse after initial invasion of the intestinal mucosa by (i) its strong resis- 
tance to phagocytosis, which is caused by the concerted action of Yop E, YopH, 
YopO (called YpkA in Y. pseudotuberculosis and Y. pestis), and YopT, (ii) its ca- 
pacity to block the proinflammatory response induced by different immune 
cells, caused by the action of Yop P (called Yop J in Y. pseudotuberculosis and 
Y. pestis) and YopH, and (iii) its ability to induce cell death of the macrophage, 
promoted by YopP/J (DeVinney et al., 2000; Aepfelbacher and Heesemann, 
2001; Cornells, 2002; Juris et al, 2002; Orth, 2002). g 



Figure 3.1. (cont. ) and by doing so contribute to the antiphagocytic action of Yersinia. Two 
Yops (YopH and YopP/J) are involved in the downregulation of the proinflammatory 
response of the immune cells, and YopP/J will also lead to the induction of apoptosis in 
macrophages. YopM, a protein containing several leucine-rich repeats, is translocated to 
the nucleus; however, its function remains unclear. PP = Peyer's patches. 




a 

ir, 



> 

W 

on 

H 



THE YERSINIA VIRULENCE SYSTEM 

Ysc-Yop III secretion system 5j 

The Ysc-Yop type III secretion machinery present in all pathogenic 

Yersinia is encoded by a 70-kDa virulence plasmid, which harbors the genes jjj 

for the Ysc (for Yop secretion) secretion apparatus or Ysc injectisome, for g 

an array of proteins secreted by this apparatus - called Yops (for Yersinia ri 

outer proteins) - and for a set of proteins controlling the system (Cornells ^ 

et al., 1998). The Ysc injectisome is composed of a large dual-ring structure * 

o 
spanning the bacterial inner and outer membrane, which resembles the flag- * 

ellum basal body. This structure is associated with a needle-like complex that § 

extends outside the bacterium. It mediates the secretion of the Yop effector o 

proteins (YopE, YopH, YopO/ YpkA, YopT, YopP/J, and YopM), a structural 
component of the needle (YscF), and the components of a translocation appa- 
ratus, which are YopB, YopD, and LcrV (Cornells et al., 1998; Cornells, 2002; 
also see Fig. 3.2). The latter proteins are inserted into the host membranes 
and form a kind of pore, which allows the delivery of the Yop effector proteins 
into the cytosol of the cell. Whether the Ysc injectisome and the translocation 



o 




CO 

I— I 

1-1 
w 

z 

O 
U 

& 



Q 
% 

< 

w 
W 
U 

w 

w 
Q 
i— i 

& 
Pi 

w 
w 
O 




YopTg/^^/W YopB, YopD 
YopHdV} Y °P E CZ^YopM 

YopP YopO 



Yersinia 



Figure 3.2. Schematic representation of entry routes during Yersinia infection in humans. 
Y. pestis is generally inoculated by fleabites or aerosol and enters the bloodstream directly. 
Y. enterocolitica and Y. pseudotuberculosis are both foodborne pathogens and enter the 
underlying lymphoid tissue (e.g., Peyer's patches) of the intestinal mucosa through M 
cells. Once Yersinia has reached the lymphoid system, the plasmid-encoded effector Yops 
allow the bacteria to avoid phagocytosis and actively downregulate the proinflammatory 
response in order to promote their extracellular survival. OM = outer membrane; 
P = periplasm; IM = inner membrane; LRR = leucine-rich repeat. 

pore form a continuous channel connected by the needle is currently being 
investigated. 

Type Ill-dependent protein secretion in Yersinia is a tightly regulated 
process, and several regulatory circuits control both the expression of the 
injection system and the injection of the Yop effector proteins itself. The first 
level of regulation involves temperature. Although growth of Yersinia is un- 
affected by low temperatures, such as those found in contaminated food 
(Y enterocolitica and Y. pseudotuberculosis) or the stomach of the fleas 
(Y pestis), the expression of Yop effector proteins is repressed at these low 
temperatures. It is only at 37° C that a stock of intracellular Yops is synthe- 
sized and the Ysc injectisome is installed. However, the injectisome remains 
closed and a mechanism of feedback inhibition prevents a deleterious accu- 
mulation of Yops (Cornells et al., 1987). This first level of regulation involves 
at least two proteins: a plasmid-encoded transcriptional activator, VirF, and 
a chromosome-encoded histone-like protein, YmoA (Cornells et al., 1991; 



Lambert de Rouvroit et al., 1992; Rohde et al., 1999). A second level of regu- 
lation is close contact with the host cell membrane, which is established by 
the bacterial adhesins Inv, YadA, and Ail (Pettersson et al., 1996). It is only at 
37° C and upon contact with the eukaryotic cell that the injectisome is opened, 
the negative feedback regulation is relieved, and Yersinia starts to inject its 
effectors into the cytosol (Francis et al., 2002; Miller, 2002). Yop proteins 
destined to be secreted have no classical signal sequence that is cleaved off 
during secretion, but nevertheless their N-terminal part (~15 amino acids 
or codons) contains the information that is necessary for secretion (Michiels 
et al., 1990; Sory et al., 1995; Anderson and Schneewind, 1997). 

Furthermore, some secreted proteins require the binding of a special- 
ized cytosolic chaperone, called Syc (specific Yersinia chaperone), to be se- 
creted (Wattiau et al., 1996). The loss of a chaperone results in the inefficient 
secretion of its cognate partner, while the secretion of other proteins re- § 

mains unaffected. Chaperones do not share a general homology, but they ^ 

do share common properties of being small (less than 20 kDa), being acidic 8 

(pi ~4-5), and possessing a C-terminal amphipatic helix. The Syc chaper- 2 

ones have been proposed to act as (i) bodyguards, preventing degradation or " 

premature association of their target; (ii) secretion pilots, being part of the % 

signal for recognition of their substrates by the export machinery; (iii) hier- h 

x 

archy factors, establishing a hierarchy for Yop delivery into its host cell; or 
(iv) antifolding factors, maintaining their substrate in a secretion-competent 
state (Frithz-Lindsten et al., 1995; Cheng and Schneewind, 1999; Boyd et al., h 



Other virulence mechanisms and TTSS present 
in Yersinia species 

In addition to the 70-kb virulence plasmid encoding the Ysc-Yop TTSS, 
Y. pestis harbors two unique plasmids encoding essential virulence determi- 
nants. The 9.5-kb plasmid (pPst/pPCPl) contains the Pla protease, which 
enables the spread of Y. pestis from subcutaneous infection sites into the 
circulation (Perry and Fetherston, 1997). Pla has been shown to exhibit 
coagulase activity and can also activate plasminogen into plasmin. It has 
also been reported that Pla can serve as an adhesion-promoting factor for 
Y. pestis (Cowan et al., 2000). The 100- to 110-kb plasmid (pFra/pMTl) en- 
codes the murine toxin Ymt, a phospholipase D family member, and the 
fraction 1 (Fl) capsule-like protein (Perry and Fetherston, 1997). Ymt has 




O 

on 
H 



O 



2000; Stebbins and Galan, 2001; Lee and Schneewind, 2002; Wulff-Strobel g 



o 



et al., 2002; Feldman et al., 2002). However, at this stage it is premature to o 



w 



decide which of these functions accounts for the need of chaperones. 2 



H 

H 
O 

*< 

O 



recently been shown to be an essential factor for the survival of Y. pestis 
in the midgut of the flea (Hinnebusch et al., 2002), and Fl has been sug- 
gested to be involved in the antiphagocytic activity of Y. pestis and to re- 
duce the number of bacteria that interact with the macrophages (Du et al., 
2002). 

In mammals, the level of free iron is too low to sustain bacterial growth; 
therefore, pathogens possess siderophores that can solubilize the iron bound 
to host proteins and transport it to the bacteria. Y. pestis, Y. pseudotuberculosis, 
and high-virulence 7. enterocolitica strains carry a chromosomally encoded 
high pathogenicity island (H PI), which comprises genes involved in the syn- 
thesis of a siderophore called yersiniabactin (Heesemann et al., 1993; Carniel, 
2001). This capacity to acquire iron is an essential virulence determinant for 
the invading Yersinia bacteria, and it endows them with the ability to multiply 
in the host and cause systemic infections. 
S Recently a second TTSS of Y. enterocolitica, called Ysa (for Yersinia 

secretion apparatus) and its substrates for secretion - Ysp proteins - has 




CO 

I— I 

h-l 



Pi 

O 
U 

& been described (Haller et al., 2000; Foultier et al., 2002). Interestingly, the 



o 

g chromosome-encoded Ysa-Ysp TTS S of Y. enterocolitica is similar to the Mxi- 

^ Spa TTSS of Shigella and to the SPI-1 encoded TTSS of S. enterica, but it 

g is different from another chromosome-encoded TTSS of Y. pestis (Parkhill 

g et al., 2001). In addition, the ysa locus is only present in the high -virulence 

biotype IB strains of Y. enterocolitica and, at least in laboratory conditions, 

h is only operational at low temperature (Haller et al., 2000; Foultier et al., 

pi 

S 2002) . Whether this Ysa TTSS plays a role in the high -virulence phenotype of 

Y. enterocolitica or in a yet to be identified cold-blooded host is unclear at the 
moment and awaits further in vivo experiments. 



FIRST CONTACT 

Interaction of the enteropathogenic Y. enterocolitica and 
Y. pseudotuberculosis with M cells 

Y. enterocolitica and Y. pseudotuberculosis possess two different adhesins: 
the chromosomally encoded Inv (Invasin) and the pYV plasmid-encoded 
YadA (Yersinia adherence protein A; see Boland and Cornells, 2000) . They me- 
diate initial adhesion, uptake, and translocation of the bacteria through the 
M cells, covering the Peyer's patches, to the underlying lymphoid tissues, 
where the bacteria remain extracellular, multiply, and eventually migrate to 
deeper tissues such as liver and spleen (Fig. 3.1; also see Sansonetti, 2002). 
The Inv protein has been shown to be important for the initial step of inva- 
sion by its strong interaction with host /^-integrin expressed on the apical 



membranes of the M cells (Pepe and Miller, 1993; Berton and Lowell, 1999; 
Schulte et al. , 2000) . The cytoplasmic domain of integrins will transmit signals 
to the cell cytoskeleton that mediate internalization of Yersinia by a "zipper- 
ing" process (Isberg et al., 2000). As epithelial cells only express integrins at 
their basal membrane, the enterocytes are not expected to be heavily invaded 
during oral infection. Indeed, an analysis of intestines of infected mice shows 
that Y enterocolitica is only found in sections that contain Peyer's patches. 
This indicates that M cells, rather than enterocytes, form the major port of 
entry for Yersinia. 

After this initial step of invasion, the YadA protein seems to be the pre- 
dominant adhesin, mediating adherence through interaction with extracel- 
lular matrix proteins such as fibronectin and collagen (El Tahir and Skurnik, 
2001). YadA also protects Y. enterocolitica against the bactericidal and opso- 
nizing action of complement by binding complement factor H (China et al., § 
1993). Once the dome is reached, Yersiniae survive attack by professional ^ 
macrophages by injecting antiphagocytic Yops (see what follows) that disrupt £ 
the cytoskeleton. Yersinia will thus essentially remain extracellular, which al- 2 
lows its survival and possible Inv-mediated entry into nonphagocytic cells, " 
but this is not well documented. % 

w 

on 
H 

X 

M 

Y. pestis enters the bloodstream immediately g 

GO 

H 

Y. pestis is a pathogen primarily affecting rodents, which is usually trans- h 

mitted to humans by a fleabite (Fig. 3.1). When a flea ingests a blood meal har- g 

boring Y. pestis, the ingested Yersinia secretes a coagulase that clots the blood o 



w 



and thus prevents the flea from swallowing the bacteria. Ymt, a plasmid- 2 



o 

encoded and intrabacterially expressed phospholipase D, protects the bac- h 

o 

terium from a cytotoxic digestion product of blood plasma in the flea gut g 

(Hinnebusch et al., 1998; Hinnebusch et al., 2002). After multiplying in the 
clotted blood, Y. pestis is transmitted efficiently into a human host when 
the hungry flea repeatedly attempts to feed and the blood clot is regurgi- 
tated into the host (Perry and Fetherston, 1997; Cole and Buchrieser, 2001). 
The bacterium then spreads from the site of infection to the regional lymph 
nodes, where it grows to high numbers and causes swelling of the lymph 
node (bubo), resulting in bubonic plague. If the lymphatic system becomes 
overwhelmed, the infection rapidly spreads into the lymphstream and blood- 
stream, causing fatal blood poisoning, followed by colonization of all the main 
organs (including the lungs). It is notable that Y pestis lacks functional YadA 
and Inv, which are present in its enteropathogenic relatives. However, some 
studies indicate that Y. pestis may invade and cause systemic infection from 
digestive and aerogenic routes of infection. 



Q. 

O 



o 
'cd 

CO 





o 



CO 

Q_ 

o 

>- 

E 
co 

"c 

CO 

.c 
o 

CD 

E 

=3 
O 
CD 



CO 
O 

o 

Q_ 
O 

>- 



_c 




_c 








_C 




V4- © 






*— 










■4— 1 

o 




"■4— ' 

o 








"■4— ' 
O 




o g 






o 










CO 




CO 








CO 




« R g 






c 















CD 

















o 










.c 




-C 








SZ 




O iE <n 






o 

13 






c 
o 




-4— 1 




-4-^ 








■!—• 




^^2 












o 


c 


o 


c 






o 


c 


°1> 




i*," 
C" c 


V4_ 


CO 




-4— • 

o 




c 


o 


c 


o 






c 


o 


c ■- o 


O 

c 


o 


•^ 


o 


CD 


CO 




'■4—' 


-4— > 




o 

"■4— » 


-f— • 







o 

"-4— • 


-4— • 




O - - -4= 

5 C (0 


-4-J 

CO 





.co 


CO 

sz 


c 




o 





o 









o 





E co c 

+-• (D E 


g 


E 
E 


CO 


CO 


Q_ 


5 




Z3 

-f— • 


CO 


2 

-4— ' 


CO 






13 

-4— • 


CO 


'■4— ' 

'n 


c 
o 


o 

■4— 1 

Q_ 


o 


o 

c 




CO 


o 


CO 


o 






CO 


O 


CO r CO 




_C0 


S- ^ 


5 


.*: 






Q 


>> 

o 


CD 

Q 


o 








Q 


-4— ' 


^ "O ^ 

Q co .£ 




^c 


CO 



CO 


CO 

E 


c 
D 




o 


o 

■4— ' 




6 










c 
o 














o 


D_ 




1_ 


-4—" 

o 












co 






^ 










"■4— 1 

CO 

> 

'■4— 1 

o 

CO 


h- 
o 

o 


o 

c 
o 



O 










°--r 

1- 

-D CD 
CO >% 




"D 


c 
o 


T3 

- C 








■4— < 

c 

CD 
CO 

c 

CO 


CO 

co 
>^ 

o 

SZ 


'-4— ' 

CO 

> 

-1— • 

o 

CO 

c 

CD 


8 i 

2 

o H_ 

9. E 




-4-< 



CD 

i_ 

CO 

■*-■ 

o 





> 

-4— • 

CO 

-4—' 

13 

Q. 

CO 

-4— • 


■4^ 

o 

c 

13 

H — 

c 
o 

'-4—' 


'E 


CO 

c 

"co 
o 

■4— • 

o 




CO 

13 

O 

13 
C 




-1— 1 


CD 

c 
'o 

13 


.Q 








-o 




CO LL 

-4-J 


sz 


CO 


o 


-4— i 

Q. 






.C 




c 

'CD 

-4— ' 


'co 

CD 


o 



-i— i 


CO 





c 

o 

c 

v 




5.E 


zz 
o 

sz 

-4— > 


'-4—' 

'l3 


c 
o 


o 

CL 
CO 




-4— > 

o 

-4— ■ 




o 


T3 


> 
CD 


o 

Q. 


CO 

CO 







CO X 

o CO 


f .1 


O 
Q. 




CO 



-4— > 




CD 


-Q 


CD 


n 







C 
13 







CO 



"D 


o 

CO 




SZ 




CO 




"■4— < 


C\] 


CO 


C\J 


sz 




. . 




c * 


CO 




-4— • 




E 
c 



-4— > 

o 

Q. 

CE 
CE 




CO 
> 

■4— 1 

o 

CO 
1 

CD 

CO 

CO 

Q_ 

H 


O 
T3 

O 

CO 

cr 

o 

sz 


CO ■*■ 

CD O 

2 o 

C CO 


o 

sz 

co" 

13 
C 

E 




C 
CO 

-Q 

E 





CO 
CO 

c 
i_ 






CO < 

O |^ 



O 
Q. 



C 



CO 


° £ 

o o 

11 


o 

c 
o 

"■4— • 

CO 
> 

o 

CO 


CD 
■D 

CO 

§ 

CO 

o 




OCCO O 


cc 




-4— 1 


E 


CO 




0- O 


O 


CO 


_c 


;~ 


_l 


















3i 

O n 
















CC 
















s « 
















CL 
















o 
















_l 
















CO 
















CD 
















cr 
































< 
















3 
















o 
































cc 
















"co 
















o 
















U) 




































c 




































T3 
















T3 


eo^a - 


















C 
















in 






CD 

CO 
CO 
0- 

h- 














CD 

c 
2 






w.sa ■ 

99SH ■ 




01 

co 

CO 


OLZQ - 




ID 

CO 

CO 

c 






0- 








0) 
CO 
CO 






CD 

Q. 

CO 

x: 
O 


frt-iy - 




< 6SL0 - 
CD 




o 

Q. 
(0 












0) 


szi-o H 

60LH - 




0) 

o 

Q. 






□ 








O 






IB 
CO 






rfl — - 


9c3Lz 


- 




CO 
>> 






CO 






















2 O) 








O 






c 






















to ^ 














D) 






















JD T3 




































3 c 














"CO 




>\ 




k \ 






A 






A 


CO 1q 






A 






A 


c 
o 
























~D 










LU 




h- 






o 


< 


X 






a: 






^ 


2 




Q. 




Q. 






Q. 


.*: 


Q. 






Q. 






Q. 


CD 




O 




O 






O 


Q. 


O 






o 






O 


cy) 




>- 




>- 






>- 


>- 


>- 






>- 






>- 


CX] 



YopE 



Figure 3.3. (facing page). Structural diagrams of the Yop effectors and their enzymatic 
function. Except for YopM, which is an LRR protein of unknown function, the enzymatic 
function of all the effector Yops has been identified, as indicated. The amino acids 
important for their catalytic function are depicted. All six Yop effectors have a short 
N-terminal secretion signal (~15 amino acids or codons), which is necessary for their 
secretion. Three of them (YopE, YopT, and YopH) also contain a chaperone-binding site. 




HOW YERSINIA ESCAPES THE HOST: TO YOP OR NOT TO YOP 
Inhibition of phagocytosis 

When a nonpathogenic bacterium enters the host organism, it is usually 
engulfed by professional phagocytes, such as macrophages, neutrophils, or 
dendritic cells (May and Machesky, 2001; Underhill and Ozinsky, 2002a). 
Phagocytosis of a bacterium is usually preceded by the activation of many 
signaling pathways, causing rearrangement of the actin cyto skeleton, exten- 
sion of the plasma membrane, and finally engulfment. Members of the Rho 
family GTPases (Cdc42, Rac, and Rho) play a central role in this process, as 
they are key regulators of the actin cytoskeleton dynamics associated with ad- 
hesion, membrane ruffling, and stress fiber formation (Hall, 1998; Bar-Sagi 
and Hall, 2000; Chimini and Chavrier, 2000). Furthermore, the formation of 
focal adhesion complexes, mediated by the action of paxillin, pl30Cas, and g 

focal adhesion kinase (FAK) , at points of contact with bacteria may also play a ^ 

role in phagocytosis (Greenberg et al., 1993; Allen and Aderem, 1996; Berton g 

and Lowell, 1999). Finally, the phosphoinositide 3-kinase (PI3K), phospholi- 2 

pase C (PLC), and protein kinase C (PKC) signaling pathways are integration S 

points for regulating phagocytosis. Pathogenic Yersiniae subvert several of £ 

these pathways by injecting YopE, YopT, YopO/YpkA, and YopH. This en- h 

sures an extracellular lifestyle and propagation in the host. It was recently M 

x 
demonstrated that deletion of either YopE, YopT, YopO/YpkA, or YopH ren- g 

H 

ders Yersinia more susceptible to phagocytosis by macrophages and PMNs, h 

and thus that the concerted action of all four antiphagocytic Yops is necessary g 

for full protection against phagocytosis (Grosdent et al., 2002). o 

z 
o 

H 
H 

Translocation of YopE into mammalian cells leads to a cytotoxic re- * 

sponse, characterized by rounding up of the cells and detachment from the 
extracellular matrix (Rosqvist et al., 1990). YopE is one of the earliest iden- 
tified Yop effectors that has an inhibitory effect on the actin cytoskeleton by 
inactivation of the Rho family GTPases (Figs. 3.3 and 3.4A; also see Black and 



o 




CO 

I— I 

1-1 
w 

z 

O 
U 

& 



Q 

w 
W 

u 
w 

w 
Q 

i— i 

& 
Pi 

w 
w 





FR 



u 



GTP 

Rho YopT 
Cdc42 
Rac 




Actin 



YopE 
YopT 
YpkA 



Inhibition of 
phagocytosis 



Rho 



Filopodia 

and 

microspikes 



w 



GTP . GTP 

Rac Cdc42 



r* 



Lamellipodia 

and 

membrane ruffles 



Stress fiber 
formation 



Actin-based 
Cytoskeletal rearrangements 




Figure 3.4. Molecular mechanism of the Yop effectors in the host cell. (A) Antiphagocytic 
action of YopE, YopT, and YopO/YpkA. Upon contact of Yersinia with a phagocytic 
receptor, the Rho family members (Rho, Rac, and Cdc42) are targeted to the cell 
membrane and converted to their GTP-activated state, which promotes actin 
polymerization and facilitates phagocytosis. YopE, acting as a GAP, will transiently 

(cont.) 



Bliska, 2000; Von Pawel-Rammingen et al., 2000). Rho GTPases are reg- 
ulated at different levels (Hall, 1998; Ridley, 2001). Cytosolic GDP-bound 
Rho proteins are normally posttranslationally modified by prenylation at the 
C-terminus, which is important for their translocation to the cell membrane. 
The guanine nucleotide exchange factors (GEFs) induce the release of bound 




Figure 3.4. (cont. ) downregulate Rho, Rac, and Cdc42 by converting them into the 

inactive GDP-bound state. The YopT cysteine protease cleaves the C-terminus of Rho, 

Rac, and Cdc42, removing the prenyl group of the Rho GTPases and liberating them from 

the plasma membrane. The YopO/YpkA serine/threonine kinase becomes 

autophosphorylated upon contact with actin and interacts with Rho and Rac; however, its 

real cellular target is unknown. The concerted action of these three Yops will lead to a 

destruction of the actin cytoskeleton network and in this way inhibit phagocytosis. FR, ^ 

phagocytic receptor; ECM, extracellular matrix. (B) Antiphagocytic and anti-inflammatory ^ 

action of YopH. The interaction between Yersinia and the eukaryotic cell surface causes a 3 

rapid tyrosine phosphorylation of adhesion complexes to mediate uptake of Yersinia. To j3 

exert its antiphagocytic role, the phosphotyrosine phosphatase YopH is targeted to the £ 

focal adhesion complexes, where it dephosphorylates proteins such as the FAK, pl30Cas, 



M 

on 

n 
> 
and paxillin, and to other not yet well-characterized adhesion-regulated complexes in g 



on 



macrophages to dephosphorylate Fyb and SKAP-HOM. By inhibition of the PI3K/Akt h 

signaling pathway, YopH also contributes to the downregulation of the inflammatory M 

response. Upon infection, Akt is activated and will phosphorylate several proteins that are o 

H 

involved in apoptosis, cellular proliferation, and cytokine/chemokine production, such as ^ 

glycogen synthase kinase 3 (GSK3) or the transcription factors of the forkhead family (e.g., ^ 

o 
FKHR), and locks them in their phosphorylated inactive state. Inhibition of this pathway ^ 

by YopH is presumably responsible for the inhibition of T-cell proliferation, IL-2 w 

production, and MCP-1 production. PTEN, phosphatase and tensin homologue deleted on § 

chromosome 10; PDK1, phosphoinositide-dependent kinase-1; RTK, receptor tyrosine o 

kinase; TCR, T-cell receptor. (C) Model showing the anti- inflammatory and proapoptotic o 

role of YopP/J. YopP/J downregulates of the inflammatory response by binding to and 

preventing the activation of members of the MAP kinase kinase (MKK) family and of 

IKK/3. By blocking both these pathways, YopP/J efficiently shuts down multiple kinase 

cascades and the cytokine induction required by the host cell to respond to a bacterial 

infection. YopP/J is also responsible for the induction of apoptosis of macrophages, which 

probably involves both the downregulation of survival genes and the activation of the 

apoptotic cascade upstream of Bid, presumably by interfering with a signaling pathway 

triggered from the TLRs. The cysteine protease activity of YopP/J is necessary for both the 

downregulation of the inflammatory response and the induction of apoptosis of 

macrophages. However, exactly how the YopP-de-sumoylating (de-ubiquitinylating?) 

activity is interrelated with the inhibition of the MKKs and IKK/) and the induction of 

apoptotic pathways awaits further research. See color section. 



GDP and thereby allow binding of GTP. This results in the activation of the 
Rho proteins at the cell membrane and binding to their downstream target. 
Inactivation of Rho GTPases is regulated by guanine nucleotide dissociation 
inhibitors (GDIs) and GTPase-activating proteins (GAPs). 

The former produce an inactive complex with the Rho proteins in the 
cytosol by masking the prenyl group, and the latter induce the hydrol- 
ysis of the bound GTP to GDP, thereby returning the Rho proteins to 
their inactive form. During phagocytosis the reorganization of the actin cy- 
toskeleton is orchestrated by these Rho family GTPases: Rho controls stress 
fiber formation and actin-myosin-based contractility; Cdc42 drives the for- 
mation of actin-rich filopodia; and Rac promotes the formation of lamel- 
lipodia and membrane ruffles (Hall, 1998). Therefore, they represent ideal 
targets for bacterial virulence factors, as their inactivation would block phago- 
cytosis and allow the extracellular survival of bacterial. The C-terminal effec- 
S tor domain of YopE mimics the activity of eukaryotic GAP, which results 

in a transient downregulation of the Rho GTPases, and in this way leads to 




CO 

I— I 

h-l 



Pi 

O 
U 

P the disruption of the actin cytoskeleton and consequently the inhibition of 



o 



o 

g phagocytosis (Black and Bliska, 2000; Von Pawel-Rammingen et al., 2000). 

^ However, although it was shown in vitro that YopE has GAP activity 

g toward Rho, Rac, and Cdc42, whether these three Rho proteins are all inacti- 

g vated in every cell type, or whether there might be other Rho family members 

that could be in vivo substrates, awaits further analysis. One clue for the speci- 

h ficity of YopE came from a study on human umbilical vein endothelial cells, 

pi 

S where it was shown that Yop E acted selectively on the Rac-mediated pathways 

but had no effect on Cdc42- or Rho-dependent signaling (Andor et al., 2001). 
It should be noted that YopE shares a high degree of structural similarity 
with the GAP domains of Exoenzyme S (ExoS) of Pseudomonas aeruginosa 
and SptP from S. typhimurium, but it has no obvious structural similarity 
with known mammalian functional GAP homologs (Evdokimov et al., 2002), 
suggesting that they could have evolved separately. 

As mentioned before, the Yersinia type III weapon includes a pore, nec- 
essary to translocate the effectors into the host (Cornells et al., 1998; Cornells, 
2002). In a current model the translocation pore is filled by the Yop effectors 
themselves (Hakansson et al., 1996b). However, recently it has been pro- 
posed that apart from its antiphagocytic role, injected YopE would also play 
a role in minimizing plasma membrane damage caused by pore formation 
(Viboud and Bliska, 2001). The GAP function of YopE was demonstrated to 
be necessary in preventing pore formation, suggesting that pore formation 
itself needs the activation of Rho GTPases. 




YopT 

The most recently identified effector modulating the Rho family of 
GTPases is YopT (Figs. 3.3 and 3.4A). Infection of mammalian cells with 
a Y. enterocolitica strain only expressing the YopT effector leads to rounding 
up of the cell and disruption of the cytoskeleton, which contributes to the 
antiphagocytic activity of YopT (Iriarte and Cornells, 1998; Grosdent et al., 
2002). Translocation of YopT into host cells leads to a modification of RhoA, 
resulting in an acidic shift in its pi and redistribution of membrane-bound 
RhoA toward the cytosol (Zumbihl et al., 1999). In addition, incubation of 
purified cell membranes or artificial lipid vesicles containing RhoA with 
purified YopT leads to the release of RhoA to the supernatant (Sorg et al., 
2001). The mechanism of action of YopT was recently unraveled by Shao and 
coworkers, who demonstrated that Yersinia YopT, as well as its homologue 
AvrPphB from P. aeruginosa, belong to a family of cysteine proteases (Shao o 

et al., 2002). YopT recognizes the posttranslational modified Rho GTPases -< 

(Rho, Rac, and Cdc42) and proteolytically cleaves them near the C-terminus, <* 

which leads to their release from the cell membrane. This cleavage removes £ 

the prenyl group of the Rho GTPases and results in an irreversible inactiva- g 

tion of the targeted Rho GTPases, whereas the action of YopE (GAP) can be £ 

reverted by the GEFs within the cell. g 

O 

on 

YopO/YpkA h 

The YopO/YpkA effector is an autophosphorylating serine/threonine o 

protein kinase that modulates the cytoskeleton dynamics (Figs. 3.3 and 3.4A) o 

and also contributes to resistance to phagocytosis (Galyovet al., 1993; Hakans- g 

son et al., 1996a; Grosdent et al., 2002). YopO/YpkA is produced as an in- h 

active kinase, which becomes activated after translocation into the host cell g 

upon binding to actin (Dukuzumuremyi et al., 2000; Juris et al., 2000). The 
N-terminal part of YopO/YpkA contains the kinase domain, whereas the C- 
terminal part of the kinase binds to actin. The C-terminal part also contains 
sequences that bear similarity to several eukaryotic RhoA-binding kinases, 
and it binds to RhoA and Rac but not to Cdc42 (Barz et al., 2000; Dukuzu- 
muremyi et al., 2000). The kinase domain is required to localize YopO/YpkA 
to the plasma membrane, whereas the C-terminal part is responsible for its 
effect on the actin cytoskeleton in Yersmia-infected cells (Dukuzumuremyi 
et al., 2000; Juris et al., 2000). However, although YopO/YpkA has a clear 
effect on the actin cytoskeleton, its real cellular target is unknown and awaits 
further investigation. 



YopH 

The fourth antiphagocytic Yop is the multifunctional YopH. The 51-kDa 
YopH effector protein is composed of two functional domains: the C-terminal 
part (residues 206-468) has a structure similar to the one of mammalian 
phosphotyrosine phosphatases (PTPases; see Guan and Dixon, 1990), and the 
N-terminal part (residues 1-130) is the binding site of YopH to its substrate 
(Fig. 3.3; also see Black etal., 1998). Part of this latter domain (residues 20-69) 
is also the binding domain for its chaperone SycH, necessary for translocation 
into the eukaryotic cell (Wattiau et al., 1996). Upon interaction of the Yersinia 
surface protein Inv with fi\ -integrin on the cell surface of epithelial cells, there 
is a rapid tyrosine phosphorylation of proteins of focal adhesion complexes 
(Perssonetal, 1997). 

Translocation of YopH into epithelial cells leads to the dephosphory- 
lation of proteins from these focal adhesion complexes, such as the dock- 
| ing proteins pl30Cas and paxillin and the FAK (Black and Bliska, 1997; 

u Persson et al., 1997). Similarly, in macrophages, contact also induces tyrosine 

g phosphorylation of proteins of adhesion complexes (Andersson et al., 1996). 

§ Targeting of YopH into macrophages also leads to a rapid dephosphorylation 

rt of pl30Cas and paxillin (Hamid et al., 1999). In addition, in macrophages 

u YopH will also lead to dephosphorylation of the Fyn -binding protein (FBP) 

Q 




CO 

I— I 

h-l 
w 



p 



and the scaffolding protein SKAP-HOM, which have been shown to be part of 
a novel adhesion-regulated signaling complex (Hamid et al., 1999; Black et al., 

h 2000) . Dephosphorylation of these proteins contributes to the antiphagocytic 

g activity of YopH (Fig. 3. 4B). 

Besides its role as an antiphagocytic factor, YopH has also been shown 
to interfere in other signaling pathways of the immune defense system, such 
as downregulating the Fc-mediated oxidative burst in macrophages and neu- 
trophils (Bliska and Black, 1995; Ruckdeschel et al., 1996) and blocking cal- 
cium signaling in neutrophils (Andersson et al., 1999). Finally, during infec- 
tion of macrophages with Y. enter ocolitica, the PI3K/Akt pathway is rapidly 
activated and then inactivated in a YopH -dependent way, which possibly con- 
tributes to an anti-inflammatory role of YopH (Fig. 3.4B; also see following 
subsection; also Sauvonnet et al., 2002a). 

Inhibition of the inflammatory response and 
induction of apoptosis 

When a pathogen interacts with a mammalian cell, multiple receptors 
will simultaneously recognize these pathogens both through direct binding 
and by binding to opsonins on the microbe surface (Underhill and Ozinsky, 



2002a) . One of the key mediators of microbe detection is the Toll-like re- 
ceptor (TLR) family, which plays an important role in signaling toward in- 
flammation and apoptosis (Akira et al., 2001; Imler and Hoffmann, 2001). 
Ten mammalian TLRs now have been identified. Five of them (TLR2, 
TLR4, TLR5, TLR6, and TLR9) have been shown to respond to an array of 
different microbial components, such as lipopolysaccharide (LPS), lipopep- 
tides, peptidoglycans (PGNs), lipoteichoic acid (LTA), flagellin, and CpG 
motifs in DNA. By using a combination of different invariant TLRs, the im- 
mune system can recognize a broad spectrum of pathogens. Stimulation of 
TLR2 and TLR4 leads to the recruitment of the adaptor molecule MyD88 and 
the serine kinase I L-l -receptor-associated kinase (IRAK; see Underhill and 
Ozinsky, 2002b). Together with TRAF-6, this multiprotein assembly me- 
diates the activation of (i) the I/cB kinase (IKK) complex, which leads to 
activation of the nuclear factor /cB (NF-/cB), and (ii) the mitogen-activated § 

protein kinase (MAPK) kinase family (MKKs), which also leads to the ^ 

activation of different transcription factors, such as activator protein- 1 £ 

(AP-1). | 



YopP/J 




w 

oo 

n 

> 

M 
oo 



W 
O 

00 

H 



As a Gram-negative bacterium, Yersinia is endowed with several compo- h 
nents capable of activating the TLR system and stimulating a proinflam- 
matory response. Indeed, during Yersinia infection of macrophages, the 

proinflammatory response is initially upregulated; however this is quickly h 

counteracted by the YopP/J effector protein (Ruckdeschel et al., 1997, 1998). g 

YopP/J has been shown to cause a variety of anti-inflammatory effects in vitro, o 

such as suppression of tumor necrosis factor-a (TNF-a?) and interleukin-6 (IL- 3 

6) and IL-8 production; downregulation of intercellular adhesion molecule- 1 h 

o 

(ICAM-1) expression; and blocking of the activation of MAPK pathways, in- g 

eluding extracellular signal-regulated kinases (ERK) , p38, jun amino-terminal 
kinase (JNK), and NF-/cB (Boland and Cornells, 1998; Palmer et al., 1998; 
Schesser et al., 1998; Palmer et al., 1999; Denecker et al., 2002). The first 
clue as to how YopP/J could simultaneously block these multiple signaling 
pathways was elucidated by Orth et al. (1999; also see Fig. 3.4C). 

In a two-hybrid system, YopJ from Y. pseudotuberculosis bound multi- 
ple members of the MAPK kinase superfamily, including MKKs and IKK/3, 
thereby preventing their phosphorylation and subsequent activation. The in- 
teraction of its counterpart YopP from Y. enterocolitica was confirmed by 
coimmunoprecipitation experiments in HEK293T cells and macrophages 
(Denecker et al., 2001; Ruckdeschel et al., 2001a). By blocking both the con- 
served family of MKKs and IKK/?, YopP/ J efficiently shuts down multiple 



kinase cascades, resulting in a downregulation of the inflammatory response 
of its host cells. 

The second clue about the mechanism of action of YopP/J was based on 
structural similarities. It has been suggested that YopP/J belongs to a family 
of cysteine proteases related to the ubiquitin-like protein proteases (Figs. 3.3 
and 3.4B; also see Orth et al., 2000; Orth, 2002). Amino acid alignment of the 
adenoviral protease AVP, YopP/J, its effector homologues, and Ulpl, a yeast 
ubiquitin-like protease, revealed the catalytic triad necessary for the cysteine 
protease activity, which is conserved between all these proteins. Mutation 
of the YopP/J hypothetical catalytic cysteine- 172, which presumably results 
in the loss of its cysteine protease activity, hampers its capacity to inhibit 
the NF-/cB and MAPK signaling cascades (Orth et al., 2000; Denecker et al., 
2001). Ubiquitin-like protein proteases cleave the C-terminus of an 11-kDa 
small ubiquitin-related modifier, SUMO-1 (Yeh et al., 2000). YopJ has been 
S shown to reduce the cellular concentration of SUMO-1 -conjugated proteins 

in an overexpression experiment; however, no direct substrate of YopJ has 




CO 

I— I 

h-l 



Pi 

O 
U 

& been identified (Orth et al., 2000). Furthermore, it was recently suggested by 

g Orth (2002) that overexpression of YopJ also leads to a decrease in ubiquiti- 

^ nated proteins. Thus the exact mechanism of blockage of phosphorylation of 

g MKKs and IKK/? through the cysteine protease function of YopP/J remains 

fc unresolved. 

In addition, it is possible that an additional cytosolic factor is needed for 

h functional activity: (i) with the use of in vitro kinase reaction experiments, 

pi 

S it was not possible to demonstrate that YopJ could prevent phosphorylation 

of MKK1 (Orth et al., 1999); (ii) recombinant protein preparations of YopJ 
were catalytically inactive when assayed with a variety of radiolabeled or flu- 
orometric peptides (Orth et al., 2000); and (iii) the viral AVP protease also 
requires an additional cofactor (Mangel et al., 1993). Besides cysteine-172, 
it was recently demonstrated that arginine-143, present in all high -virulence 
Y. enterocolitica strains and in YopJ ( Y. pestis and Y. pseudotuberculosis) , plays a 
major role in determining the inhibitory impact of YopP on the suppression 
of NF-/cB activation and survival of macrophages (Ruckdeschel et al., 2001b; 
Denecker et al., 2002). 

YopP/J is not only responsible for the inhibition of the inflammatory 
response of the host but also induces apoptosis in macrophages, although 
not in other cell types (Mills et al., 1997; Monack et al., 1997). Two different 
mechanisms of YopP/J-dependent apoptosis have been proposed. In the first 
hypothesis, YopP/ J would act as a direct activator of the cell death machinery, 
which involves an early - presumably caspase-8-dependent - cleavage of Bid 




to its proapoptotic truncated form, followed by the release of cytochrome c 
from the mitochondria, leading to the activation of procaspase-9, -3, and -7 
(Denecker et al., 2001). The point at which YopP/J interferes with the apop- 
totic signaling cascade is a subject for future studies. In the second hypothesis, 
YopP/ J -induced apoptosis of macrophages would merely result from its in- 
hibition of NF-/cB activation, thus blocking the host cell survival pathways, in 
combination with TLR stimulation (Ruckdeschel et al., 1998, 2001a, 2002). 
The two hypotheses could be combined in a model in which the YopP/ J 
targets both antiapoptotic and proapoptotic pathways in macrophages (Fig. 
3.4C). In this model, YopP/J might downregulate the expression of some 
antiapoptotic genes and at the same time alter a TLR-dependent signaling 
cascade in a manner allowing procaspase-8 activation and subsequent Bid 
processing. 

o 

YopH | 

YopH could also contribute to the downregulation of the inflammatory | 

response (Fig. 3.4B). Indeed, YopH has recently been shown to suppress 5 

the Yersinia-induced activation of the PI3K/Akt signaling pathway, which *t 

could be correlated with the downregulation of monocyte chemoattractant h 
protein-1 (MCP-1) mRNA levels (upregulation of MCP-1 is dependent on the 
PI3K/Akt signaling cascade; see Alberta et al., 1999; Scheid and Woodgett, 

2001; Sauvonnet et al., 2002a). By inhibition of MCP-1 production, YopH h 

would inhibit the recruitment of other macrophages to the site of infection, g 

which would allow Yersinia to colonize the lymphoid system. Besides its role o 

in the innate immune system, YopH also contributes to the downregulation 3 

of the adaptive immune response. It was demonstrated that T-cell cytokine h 

o 

production and proliferation, and expression of the B-cell costimulatory re- g 

ceptor B7.2, in response to antigen stimulation were inhibited after transient 
exposure to Yersinia (Yao et al., 1999; Sauvonnet et al., 2002a). This inhibition 
of antigen-specific T- and B-cell activation occurred in a YopH -dependent way 
by interfering with the phosphorylation of tyrosine-phosphorylated compo- 
nents associated with the T- and B-cell antigen receptor signaling complex 
(e.g., Fyb and SKAP-HOM), and most probably also by interfering with the 
PI3K/Akt pathway (Fig. 3.4B; also see Hamid et al., 1999; Yao et al, 1999; 
Black et al., 2000; Sauvonnet et al., 2002a). Thus, Yersinia possesses different 
elements that have the capacity to downregulate the inflammatory response. 
The relevance of the anti -inflammatory role played by these different ele- 
ments during infection is a matter for future in vivo studies. 



O 

on 
H 




Effectors other than Yops that help to defeat the 
immune response 

Analysis of the transcriptome alterations in infected mouse macrophages 
revealed that several genes involved in the inflammatory response of a 
macrophage to a bacterial infection are downregulated by the action of pYV- 
encoded factors other than YopP (Sauvonnet et al., 2002b). As already men- 
tioned, YopH is a good candidate (Yao et al., 1999; Sauvonnet et al., 2002a). In 
addition, LcrV may represent another factor, as it was recently demonstrated 
that LcrV-induced IL-10 release could inhibit TNF-a? production in zymosan 
A-stimulated macrophages (Sing et al., 2002). According to the currently ac- 
cepted type III-secretion-translocation model (Cornelis et al., 1998), LcrV is 
part of the translocation machinery that delivers Yop effectors into the eu- 
karyotic cell, but this does not exclude a possible role of its own, independent 
£ of the rest of the injectisome. 

g In addition, for the induction of cell death, YopP might not be the sole 

* factor. For Y. pseudotuberculosis it was recently demonstrated that the Inv 

^ protein could cause a rapid apoptotic-necrotic caspase-independent cell death 

^ in T lymphocytes (Arencibia et al. , 2002) . This process was mediated by means 

w of ^1-integrins and was independent of the Yop-Ysc TTSS of Yersinia. 

u 

w 

2 YopM 

h YopM belongs to a growing family of type III effectors that has several 

pi 

£ representatives in Shigella (ipaH multigene family) and Salmonella (SspH; see 

o 

Kobe and Kajava, 2001). It is a strongly acidic protein composed almost en- 
tirely of 20/22 residue leucine-rich repeats (LRR; see Fig. 3.3). The repeating 
LRR unit of YopM is the shortest among all LRRs known to date, and, depend- 
ing on the Yersinia species, the amount of copies can vary between 13 and 20 
repeats. The crystal structure has revealed that the LRRs, consisting of paral- 
lel p -sheets, form a crescent shape, which is flanked by an a -helical hairpin at 
the N-terminus (Evdokimov et al., 2001). The latter domain has been shown 
to be part of the signal necessary to target YopM for translocation into eukary- 
otic cells (Boland et al., 1996). Intriguingly, individual YopM molecules form 
a tetramer in the crystal, creating a hollow cylinder with an inner diameter 



of 35 A (Evdokimov et al., 2001). YopM has been shown to traffic to the nu- 
cleus via a vesicle-associated pathway, but its action in the nucleus remains 
unknown (Skrzypek et al., 1998). New insights about the role of YopM came 
from a recent study in which a transcriptome analysis of Yersima-infected 
macrophages revealed that YopM may control the expression of genes in- 
volved in the cell cycle and in cell growth (Sauvonnet et al., 2002b). 



REFERENCES 

Aepfelbacher, M. and Heesemann, J. (2001). Modulation of Rho GTPases and the 

actincyto skeleton by Yersinia outer proteins (Yops). Int. J. Med. Microbiol. 291, 

269-276. 
Akira, S., Takeda, K., and Kaisho, T. (2001). Toll-like receptors: critical proteins 

linking innate and acquired immunity. Nat. Immunol. 2, 675-680. 
Alberta, J.A., Auger, K.R., Batt, D., Iannarelli, P., Hwang, G., Elliott, H.L., Duke, 

R., Roberts, T.M., and Stiles, CD. (1999). Platelet-derived growth factor 




CONCLUSION 

After initial invasion of its host cells, pathogenic Yersinia remain extra- 
cellular because they have the capacity to resist phagocytosis. This resistance 
depends on the type III Ysc-Yop system, which upon close contact with a 
target cell injects six different effector Yops into the cytosol of the cell. As a 
result of this "tranquilizing" injection, phagocytosis is inhibited, the onset 
of the proinflammatory response is slowed down, and most probably lym- 
phocyte proliferation is prevented. Four Yop effectors (YopE, YopT, YopO, 
and YopH) contribute to the antiphagocytic action of Yersinia, as their con- 
certed action leads to the complete destruction of the actin cytoskeleton. 
YopE (GAP) and YopT (cysteine protease) target the Rho family of GTPases 
directly and will inhibit their activation. YopO/YpkA also interacts with the 
Rho family of GTPases. However, although this leads to a partial destruc- 
tion of the cytoskeleton, currently no direct target could be identified. Lastly, 
the antiphagocytic function of the tyrosine phosphatase YopH is to disas- 3 

semble adhesion complexes at the cell membrane. Although YopH plays a 
crucial role in the antiphagocytic protection, it has recently become clear that 
YopH may have other important roles during infection, such as inhibiting jj 

the proliferation of lymphocytes. YopP/J (cysteine protease) efficiently shuts g 

i_i 

down multiple kinase cascades, and in this way it may be responsible for the ffi 

downregulation of the inflammatory response, required by the host cell to K 

respond to a bacterial infection. YopP/J is also responsible for the induction ri 

of apoptosis of macrophages. ° 

Understanding the role of the YopM effector is still a challenge and is a § 

topic for further research. Thus, the action of different Yops may converge w 

into single key issues, but one Yop may have different effects. Surprisingly, § 

LcrV, one of the proteins that are involved in translocation of the effectors o 

across the host cell membrane, has an anti-inflammatory role on its own, % 
without being injected into the cell cytosol. 



x 

o 
3 



ir, 



Hi 

on 




CO 

I— I 

h-l 
w 



stimulation of monocyte chemoattractant protein-1 gene expression is me- 
diated by transient activation of the phosphoinositide 3-kinase signal trans- 
duction pathway. J. Biol. Chem. 274, 31,062-31,067. 

Allen, L.A. and Aderem, A. (1996). Mechanisms of phagocytosis. Curr. Opin. 
Immunol. 8, 36-40. 

Anderson, D.M. and Schneewind, O. (1997). A mRNA signal for the type III 
secretion of Yop proteins by Yersinia enterocolitica. Science 278, 1140-1143. 

Andersson, K., Carballeira, N., Magnusson, K.E., Persson, C, Stendahl, O., Wolf- 
Watz, H., and Fallman, M. (1996). YopH of Yersinia pseudotuberculosis in- 
terrupts early phosphotyrosine signalling associated with phagocytosis. Mol. 
Microbiol. 20, 1057-1069. 

Andersson, K., Magnusson, K.E., Majeed, M., Stendahl, O., and Fallman, M. 
(1999). Yersinia pseudotuberculosis-induced calcium signaling in neutrophils 
is blocked by the virulence effector YopH. Infect. Immun. 67, 2567-2574. 

5 Andor, A., Trulzsch, K., Essler, M., Roggenkamp, A., Wiedemann, A., Heese- 
8 mann, J., and Aepfelbacher, M. (2001). YopE of Yersinia, a GAP for Rho 

6 GTPases, selectively modulates Rac-dependent actin structures in endothe- 
g Hal cells. Cell. Microbiol. 3, 301-310. 

rt Arencibia, I., Frankel, G., and Sundqvist, K.G. (2002). Induction of cell death 

g in T lymphocytes by invasin via beta 1-integrin. Eur. J. Immunol. 32, 1129- 

| 1138. 

Autenrieth, LB. and Firsching, R. (1996). Penetration of M cells and destruction 
h of Peyer's patches by Yersinia enterocolitica: an ultrastructural and histological 

g study. J. Med. Microbiol. 44, 285-294. 

Bar-Sagi, D. and Hall, A. (2000). Ras and Rho GTPases: a family reunion. Cell 

103, 227-238. 
Barz, C., Abahji, T.N., Trulzsch, K., and Heesemann, J. (2000). The Yersinia 
Ser/Thr protein kinase YpkA/YopO directly interacts with the small GTPases 
RhoA and Rac-1. FEBS Lett. 482, 139-143. 
Berton, G. and Lowell, C.A. (1999). Integrin signalling in neutrophils and 

macrophages. Cell. Signal. 11, 621-635. 
Black, D.S. and Bliska, J.B. (1997). Identification of pl30Cas as a substrate of 
Yersinia YopH (Yop51), a bacterial protein tyrosine phosphatase that translo- 
cates into mammalian cells and targets focal adhesions. EM BO J. 16, 2730- 
2744. 
Black, D.S. and Bliska, J.B. (2000). The RhoGAP activity of the Yersinia pseudo- 
tuberculosis cytotoxin YopE is required for antiphagocytic function and viru- 
lence. Mol. Microbiol. 37, 515-527. 
Black, D.S., Marie-Cardine, A., Schraven, B., and Bliska, J.B. (2000). The Yersinia 
tyrosine phosphatase YopH targets a novel adhesion-regulated signalling 
complex in macrophages. Cell. Microbiol. 2, 401-414. 



Black, D.S., Montagna, L.G., Zitsmann, S.,andBliska, J.B. (1998). Identification of 
an amino -terminal substrate-binding domain in the Yersinia tyrosine phos- 
phatase that is required for efficient recognition of focal adhesion targets. 
Mol. Microbiol. 29, 1263-1274. 

Bliska, J.B. and Black, D.S. (1995). Inhibition of the Fc receptor-mediated oxidative 
burst in macrophages by the Yersinia pseudotuberculosis tyrosine phosphatase. 
Infect. Immun. 63, 681-685. 

Boland, A. and Cornells, G.R. (1998). Role of YopP in suppression of tumor 
necrosis factor alpha release by macrophages during Yersinia infection. Infect. 
Immun. 66, 1878-1884. 

Boland, A. and Cornells, G.R. (2000). Interaction of Yersinia with host cells. Sub- 
cell. Biochem. 33, 343-382. 

Boland, A., Sory, M.P., Iriarte, M., Kerbourch, C, Wattiau, P., and Cornells, G.R. 

(1996). Status of YopM and YopN in the Yersinia Yop virulon: YopM of Y. § 

enterocolitica is internalized inside the cytosol of PU5-1.8 macrophages by ^ 

the YopB, D, N delivery apparatus. EMBOJ. 15, 5191-5201. g 

Boyd, A.P., Lambermont, I., and Cornells, G.R. (2000). Competition between the h 

Yops of Yersinia enterocolitica for delivery into eukaryotic cells: role of the " 

SycE chaperone binding domain of YopE. J. Bacteriol. 182, 4811-4821. *t 



w 



Buttner, D. and Bonas, U. (2002). Port of entry- the type III secretion translocon. h 

Trends Microbiol. 10, 186-192. 



O 

on 
H 



Carniel, E. (2001). The Yersinia high-pathogenicity island: an iron-uptake island. 

Microbes Infect. 3, 561-569. h 

Cheng, L.W. and Schneewind, O. (1999). Yersinia enterocolitica type III secretion. g 



^ 



On the role of SycE in targeting YopE into HeLa cells. J. Biol. Chem. 274, o 

22,102-22,108. z 

o 

Chimini, G. and Chavrier, P. (2000). Function of Rho family proteins in actin H 

_ o 

dynamics during phagocytosis and engulfment. Nat. Cell. Biol. 2, E191-E196. g 

China, B., Sory, M.P., N'Guyen, B.T., De Bruyere, M., and Cornells, G.R. (1993). * 

Role of the YadA protein in prevention of opsonization of Yersinia enterocol- 
itica by C3b molecules. Infect. Immun. 61, 3129-3136. 
Cole, S.T. and Buchrieser, C. (2001). Bacterial genomics. A plague o' both your 

hosts. Nature 413, 467, 469-470. 
Cornells, G. (2002). Yersinia type III secretion: send the effectors. J. Cell Biol. 158, 

401-408. 
Cornells, G., Vanootegem, J.C., and Sluiters, C. (1987). Transcription of the yop 

regulon from Y. enterocolitica requires trans acting pYV and chromosomal 

genes. Microb. Pathog. 2, 367-379. 
Cornells, G.R. (2000). Type III secretion: a bacterial device for close combat with 

cells of their eukaryotic host. Philos. Trans. R. Soc. Lond. B. Biol. Sci. 355, 

681-693. 




CO 

I— I 

h-l 
w 



Cornells, G.R., Boland, A., Boyd, A.P., Geuijen, C, Iriarte, M., Neyt, C, Sory, 
M.P., and Stainier, I. (1998). The virulence plasmid of Yersinia, an antihost 
genome. Microbiol. Mol. Biol. Rev. 62, 1315-1352. 
Cornells, G.R., Sluiters, C, Delor, I., Geib, D., Kaniga, K., Lambert de Rouvroit, C, 
Sory, M.P., Vanooteghem, J.C., and Michiels, T. (1991). ymoA, a Yersinia en- 
terocolitica chromosomal gene modulating the expression of virulence func- 
tions. Mol. Microbiol. 5, 1023-1034. 
Cowan, C, Jones, H.A., Kaya, Y.H., Perry, R.D., and Straley, S.C. (2000). Invasion 
of epithelial cells by Yersinia pestis: evidence for a Y. pestis-specific invasin. 
Infect. Immun. 68, 4523-4530. 
Denecker, G., Declercq, W., Geuijen, C.A., Boland, A., Benabdillah, R., van Gurp, 
M., Sory, M.P., Vandenabeele, P., and Cornells, G.R. (2001). Yersinia entero- 
colitica YopP-induced apoptosis of macrophages involves the apoptotic sig- 
naling cascade upstream of Bid. J. Biol. Chem. 276, 19,706-19,714. 
S Denecker, G., Totemeyer, S., Mota, L.J., Troisfontaines, P., Lambermont, I., 

8 Youta, C, Stainier, I., Ackermann, ML, and Cornells, G.R. (2002). Effect 

P of low- and high-virulence Yersinia enterocolitica strains on the inflamma- 

g tory response of human umbilical vein endothelial cells. Infect. Immun. 70, 

rt 3510-3520. 

g DeVinney, I., Steele-Mortimer, I., and Finlay, B.B. (2000). Phosphatases and ki- 

rn 

g nases delivered to the host cell by bacterial pathogens. Trends Microbiol. 8, 

2 29-33. 

h Du, Y., Rosqvist, R., and Forsberg, A. (2002). Role of fraction 1 antigen of Yersinia 

S pestis in inhibition of phagocytosis. Infect. Immun. 70, 1453-1460. 

Dukuzumuremyi, J.M., Rosqvist, R., Hallberg, B., Akerstrom, B., Wolf-Watz, 
H., and Schesser, K. (2000). The Yersinia protein kinase A is a host factor 
inducible Rho A/ Rac -binding virulence factor. J. Biol. Chem. 275, 35,281- 
35,290. 

El Tahir, Y. and Skurnik, M. (2001). YadA, the multifaceted Yersinia adhesin. Int. 
J. Med. Microbiol. 291, 209-218. 

Evdokimov, A.G., Anderson, D.E., Routzahn, K.M., and Waugh, D.S. (2001). Un- 
usual molecular architecture of the Yersinia pestis cytotoxin YopM: a leucine- 
rich repeat protein with the shortest repeating unit. J. Mol. Biol. 312, 807-821. 

Evdokimov, A.G., Tropea, J.E., Routzahn, K.M., and Waugh, D.S. (2002). Crystal 
structure of the Yersinia pestis GTPase activator YopE. Protein Sci. 11, 401- 
408. 

Feldman, M.F., Muller, S., Wriest, E., and Cornells, G.R. (2002). SycE allows 
secretion of YopE-DHFR hybrids by the Yersinia enterocolitica type III Ysc 
system. Mol. Microbiol. 46, 1183-1197. 

Foultier, B., Troisfontaines, P., Muller, S., Opperdoes, F., and Cornells, G.R. 
(2002). Characterization of the ysa pathogenicity locus in the chromosome of 



Yersinia enterocolitica and phylogenic analysis of type III secretion systems. 

J. Mol. Evol. 55, 37-51. 
Francis, M.S., Wolf-Watz, H., and Forsberg, A. (2002). Regulation of type III 

secretion systems. Curr. Opin. Microbiol. 5, 166-172. 
Frithz-Lindsten, E., Rosqvist, R., Johansson, L., and Forsberg, A. (1995). The 

chaperone-like protein YerA of Yersinia pseudotuberculosis stabilizes YopE 

in the cytoplasm but is dispensible for targeting to the secretion loci. Mol. 

Microbiol. 16, 635-647. 
Galan, J.E. (2001). Salmonella interactions with host cells: type III secretion at 

work. Annu. Rev. Cell. Dev. Biol. 17, 55-68. 
Galan, J.E. and Collmer, A. (1999). Type III secretion machines: bacterial devices 

for protein delivery into host cells. Science 284, 1322-1328. 
Galyov, E.E., Hakansson, S., Forsberg, A., and Wolf-Watz, H. (1993). A secreted 

protein kinase of Yersinia pseudotuberculosis is an indispensable virulence § 

determinant. Nature 361, 730-732. ^ 

Greenberg, S., Chang, P., and Silverstein, S.C. (1993). Tyrosine phosphorylation S 

is required for Fc receptor-mediated phagocytosis in mouse macrophages. J. S 

Exp. Med. 177, 529-534. 5 

Grosdent, N., Maridonneau-Parini, I., Sory, M.P., and Cornells, G.R. (2002). Role *t 

of Yops and adhesins in resistance of Yersinia enterocolitica to phagocytosis. h 

Infect. Immun. 70, 4165-4176. 
Guan, K.L. and Dixon, J.E. (1990). Protein tyrosine phosphatase activity of an 

essential virulence determinant in Yersinia. Science 249, 553-556. h 



O 

on 
H 



O 



Hakansson, S., Galyov, E.E., Rosqvist, R., and Wolf-Watz, H. (1996a). The Yersinia g 

YpkA Ser/Thr kinase is translocated and subsequently targeted to the inner o 

surface of the HeLa cell plasma membrane. Mol. Microbiol. 20, 593-603. 3 

Hakansson, S., Schesser, K., Persson, C, Galyov, E.E., Rosqvist, R., Homble, F., H 

o 

and Wolf-Watz, H. (1996b). The YopB protein of Yersinia pseudotuberculosis g 

is essential for the translocation of Yop effector proteins across the target cell 
plasma membrane and displays a contact-dependent membrane disrupting 
activity. EMBOJ. 15, 5812-5823. 

Hall, A. (1998). Rho GTPases and the actin cytoskeleton. Science 279, 509-514. 

Haller, J.C., Carlson, S., Pederson, K.J., and Pierson, D.E. (2000). A chromosoma- 
lly encoded type III secretion pathway in Yersinia enterocolitica is important 
in virulence. Mol. Microbiol. 36, 1436-1446. 

Hamid, N., Gustavsson, A., Andersson, K., McGee, K., Persson, C, Rudd, C.E., 
and Fallman, M. (1999). YopH dephosphorylates Cas and Fyn-binding pro- 
tein in macrophages. Microb. Pathog. 27, 231-242. 

Heesemann, J., Hantke, K., Vocke, T., Saken, E., Rakin, A., Stojiljkovic, I., 
and Berner, R. (1993). Virulence of Yersinia enterocolitica is closely associ- 
ated with siderophore production, expression of an iron-repressible outer 




CO 

I— I 

h-l 
w 



membrane polypeptide of 65,000 Da and pesticin sensitivity. Mol. Microbiol. 8, 

397-408. 

Hinnebusch, B.J., Fischer, E.R., and Schwan, T.G. (1998). Evaluation of the role of 

the Yersinia pestis plasminogen activator and other plasmid-encoded factors 

in temperature -dependent blockage of the flea. J. Infect. Dis. 178, 1406-1415. 

Hinnebusch, B.J., Rudolph, A.E., Cherepanov, P., Dixon, J.E., Schwan, T.G., and 

Forsberg, A. (2002). Role of Yersinia murine toxin in survival of Yersinia pestis 

in the midgut of the flea vector. Science 296, 733-735. 

Hueck, C.J. (1998). Type III protein secretion systems in bacterial pathogens of 

animals and plants. Microbiol. Mol. Biol. Rev. 62, 379-433. 
Imler, J.L. and Hoffmann, J.A. (2001). Toll receptors in innate immunity. Trends 

Cell Biol. 11, 304-311. 
Iriarte, M. and Cornells, G.R. (1998). YopT, a new Yersinia Yop effector protein, 
affects the cytoskeleton of host cells. Mol. Microbiol. 29, 915-929. 
S Isberg, R.R., Hamburger, Z., and Dersch, P. (2000). Signaling and invasin- 

8 promoted uptake via integrin receptors. Microbes Infect 2, 793-801. 

B Juris, S.J., Rudolph, A.E., Huddler, D., Orth, K., and Dixon, J.E. (2000). A distinc- 

g tive role for the Yersinia protein kinase: actin binding, kinase activation, and 

J5 cytoskeleton disruption. Proc. Natl. Acad. Sci. USA 97, 9431-9436. 

g Juris, S.J., Shao, F., and Dixon, J.E. (2002). Yersinia effectors target mammalian 

g signalling pathways. Cell. Microbiol. 4, 201-211. 

Kobe, B. and Kajava, A.V. (2001). The leucine-rich repeat as a protein recognition 

h motif Curr. Opin. Struct. Biol. 11, 725-732. 

ph 

S Lambert de Rouvroit, C, Sluiters, C, and Cornells, G.R. (1992). Role of the tran- 

o 

scriptional activator, VirF, and temperature in the expression of the pYV 

plasmid genes of Yersinia enterocolitica. Mol. Microbiol. 6, 395-409. 
Lee, V.T. and Schneewind, O. (2002). Yop fusions to tightly folded protein do- 
mains and their effects on Yersinia enterocolitica type III secretion./. Bacteriol. 

184, 3740-3745. 
Mangel, W.F., McGrath, W.J., Toledo, D.L., and Anderson, C.W. (1993). Viral 

DNA and a viral peptide can act as cofactors of adenovirus virion proteinase 

activity. Nature 361, 274-275. 
May, R.C. and Machesky, L.M. (2001). Phagocytosis and the actin cytoskeleton. J. 

Cell Sci. 114, 1061-1077. 
Michiels, T., Wattiau, P., Brasseur, R., Ruysschaert, J.M., and Cornells, G. (1990). 

Secretion of Yop proteins by Yersiniae. Infect. Immun. 58, 2840-2849. 
Miller, V.L. (2002). Connections between transcriptional regulation and type III 

secretion? Curr. Opin. Microbiol. 5, 211-215. 
Mills, S.D., Boland, A., Sory, M.P., van der Smissen, P., Kerbourch, C, Finlay, 

B.B., and Cornells, G.R. (1997). Yersinia enterocolitica induces apoptosis in 



macrophages by a process requiring functional type III secretion and translo- 
cation mechanisms and involving YopP, presumably acting as an effector 
protein. Proc. Natl. Acad. Sci. USA94, 12,638-12,643. 

Monack, D.M., Mecsas, J., Ghori, N., and Falkow, S. (1997). Yersinia signals 
macrophages to undergo apoptosis and YopJ is necessary for this cell death. 
Proc. Natl. Acad. Sci. USA 94, 10,385-10,390. 

Orth, K. (2002). Function of the Yersinia effector YopJ. Curr. Opin. Microbiol. 5, 
38-43. 

Orth, K., Palmer, L.E., Bao, Z.Q., Stewart, S., Rudolph, A.E., Bliska, J.B., and 
Dixon, J.E. (1999). Inhibition of the mitogen-activated protein kinase kinase 
superfamily by a Yersinia effector. Science 285, 1920-1923. 

Orth, K., Xu, Z., Mudgett, M.B., Bao, Z.Q., Palmer, L.E., Bliska, J.B., Mangel, W.F., 
Staskawicz, B., and Dixon, J.E. (2000). Disruption of signaling by Yersinia 
effector YopJ, a ubiquitin-like protein protease. Science 290, 1594-1597. § 

Palmer, L.E., Hobbie, S., Galan, J.E., and Bliska, J.B. (1998). YopJ of Yersinia 
pseudotuberculosis is required for the inhibition of macrophage TNF-alpha 
production and downregulation of the MAP kinases p38 and JNK. Mol. Mi- S 

crobiol. 27, 953-965. £ 

n 

Palmer, L.E., Pancetti, A.R., Greenberg, S., and Bliska, J.B. (1999). YopJ of Yersinia *t 

spp. is sufficient to cause downregulation of multiple mitogen-activated pro- h 

tein kinases in eukaryotic cells. Infect. Immun. 67, 708-716. M 

Parkhill, J., Wren, B.W., Thomson, N.R., Titball, R.W., Holden, M.T., Prentice, g 

M.B., Sebaihia, M., James, K.D., Churcher, C, Mungall, K.L., Baker, S., h 




a 

ir, 



O 



Basnam, D., Bentley, S.D., Brooks, K., Cerendo-Tarraga, A.M., Chilling- g 

to 

worth, T., Cronin, A., Davies, R.M., Davis, P., Dougan, G., Feltwell, T., o 

to 

Hemlin, N., Holroyd, S., Jagels, K., Karlyshev, A.V., Leather, S., Moule, S., 3 

Oyston, P.C., Quail, M., Rutherford, K., Simmons, M., Skelton, J., Stevens, H 

o 

K., Whitehead, S., Barrell, B.G. (2001). Genome sequence of Yersinia pestis, g 

the causative agent of plague. Nature 413, 523-527. 

Pepe, J.C. and Miller, V.L. (1993). Yersinia enterolitica invasin: a primary role in 
the initiation of infection. Proc. Natl. Acad. Sci. USA 90, 6473-6477. 

Perry, R.D. and Fetherston, J.D. (1997). Yersinia pestis - etiologic agent of plague. 
Clin. Microbiol. Rev. 10, 35-66. 

Persson, C., Carballeira, N., Wolf-Watz, H., and Fallman, M. (1997). The PTPase 
YopH inhibits uptake of Yersinia, tyrosine phosphorylation of pl30Cas and 
FAK, and the associated accumulation of these proteins in peripheral focal 
adhesions. EMBOJ. 16, 2307-2318. 

Pettersson, J., Nordfelth, R., Dubinina, E., Bergman, T., Gustafsson, M., Mag- 
nusson, K.E., and Wolf-Watz, H. (1996). Modulation of virulence factor ex- 
pression by pathogen target cell contact. Science 273, 1231-1233. 




CO 

I— I 

h-l 
w 



Ridley, A.J. (2001). Rho proteins, PI 3-kinases, and monocyte/macrophage motil- 
ity. FEBS Lett. 498, 168-171. 

Rohde, J.R., Luan, X.S., Rohde, H., Fox, J.M., and Minnich, S.A. (1999). The 
Yersinia enterocolitica pYV virulence plasmid contains multiple intrinsic DNA 
bends which melt at 37 degrees C. J. Bacteriol. 181, 4198-4204. 

Rosqvist, R., Forsberg, A., Rimpilainen, M., Bergman, T., and Wolf-Watz, H. 
(1990). The cytotoxic protein YopE of Yersinia obstructs the primary host 
defence. Mol. Microbiol. 4, 657-667. 

Ruckdeschel, K., Harb, S., Roggenkamp, A., Hornef, M., Zumbihl, R., Kohler, 
S., Heesemann, J., and Rouot, B. (1998). Yersinia enterocolitica impairs acti- 
vation of transcription factor NF-kappa B: involvement in the induction of 
programmed cell death and in the suppression of the macrophage tumor 
necrosis factor alpha production. J. Exp. Med. 187, 1069-1079. 

Ruckdeschel, K., Machold, J., Roggenkamp, A., Schubert, S., Pierre, J., Zumbihl, 

5 R., Liautard, J. P., Heesemann, J., and Rouot, B. (1997). Yersinia enterocolit- 
8 ica promotes deactivation of macrophage mitogen-activated protein kinases 

6 extracellular signal-regulated kinase-1/2, p38, and c-Jun NH2-terminal ki- 
g nase. Correlation with its inhibitory effect on tumor necrosis factor-alpha 
J5 production. J. Biol. Chem. 272, 15,920-15,927. 

w 

g Ruckdeschel, K., Mannel, O., Richter, K., Jacobi, C, Triilzsch, K., Rouot, B., and 

g Heesemann, J. (2001a). Yersinia outer protein P of Yersinia enterocolitica si- 

multaneously blocks the nuclear factor-/c B pathhway and exploits lipopolysac- 
h charide signaling to trigger apoptosis in macrophages. J. Immunol. 166, 1823- 

g 1831. 

o 

Ruckdeschel, K., Mannel, O., and Schrottner, P. (2002). Divergence of apoptosis- 
inducing and preventing signals in bacteria-faced macrophages through 
myeloid differentiation factor 88 and IL-1 receptor-associated kinase mem- 
bers. J. Immunol. 168, 4601-4611. 

Ruckdeschel, K., Richter, K., Mannel, O., and Heesemann, J. (2001b). Arginine- 
143 of Yersinia enterocolitica YopP crucially determines isotype -related NF- 
kappa B suppression and apoptosis induction in macrophages. Infect. Immun. 
69, 7652-7662. 

Ruckdeschel, K., Roggenkamp, A., Schubert, S., and Heesemann, J. (1996). Dif- 
ferential contribution of Yersinia enterocolitica virulence factors to evasion of 
microbicidal action of neutrophils. Infect. Immun. 64, 724-733. 

Sansonetti, P. (2002). Host-pathogen interactions: the seduction of molecular 
cross talk. Gut 50 (Suppl 3), III2-III8. 

Sansonetti, P.J. (2001). Microbes and microbial toxins: paradigms for microbial- 
mucosal interactions III. Shigellosis: from symptoms to molecular patho- 
genesis. Am. J. Physiol. Gastrointest. Liver Physiol. 280, G319-G323. 




3 

a 



Sauvonnet, N., Lambermont, I., van der Bruggen, P., and Cornells, G. (2002a). 
YopH prevents monocyte chemoattractant protein 1 expression in macro- 
phages and T-cell proliferation through inactivation of the phosphatidylinos- 
itol 3-kinase pathway. Mol. Microbiol. 

Sauvonnet, N., Pradet-Balade, B., Garcia-Sanz, J.A., and Cornells, G.R. (2002b). 
Regulation of mRNA expression in macrophages following Yersinia entero- 
colitica infection: role of different Yop effectors. J. Biol. Chem. 2, 2. 

Scheid, M.P. and Woodgett, J.R. (2001). PKB/AKT: functional insights from ge- 
netic models. Nat. Rev. Mol. Cell. Biol. 2, 760-768. 

Schesser, K., Spiik, A.K., Dukuzumuremyi, J.M., Neurath, M.F., Pettersson, S., 
and Wolf-Watz, H. (1998). The yop J locus is required for Yers wia-mediated 
inhibition of NF-kappa B activation and cytokine expression: YopJ contains 
a eukaryotic SH2-like domain that is essential for its repressive activity. Mol. 
Microbiol. 28, 1067-1079. g 

Schulte, R., Kerneis, S., Klinke, S., Bartels, H., Preger, S., Kraehenbuhl, J. P., 
Pringault, E., and Autenrieth, LB. (2000). Translocation of Yersinia entro- 
colitica across reconstituted intestinal epithelial monolayers is triggered by S 

Yersinia invasin binding to beta 1 integrins apically expressed on M-like cells. " 

Cell. Microbiol. 2, 173-185. % 

Shao, F., Merritt, P.M., Bao, Z., Innes, R.W., and Dixon, I.E. (2002). A Yersinia h 

effector and a Pseudomonas avirulence protein define a family of cysteine ^ 

proteases functioning in bacterial pathogenesis. Cell 109, 575-588. 8 

Sing, A., Roggenkamp, A., Geiger, A.M., and Heesemann, J. (2002). Yersinia h 

enterocolitica evasion of the host innate immune response by V antigen- g 

to 

induced IL-10 production of macrophages is abrogated in IL-10-deficient o 

to 

mice. J. Immunol. 168, 1315-1321. 2 

J o 

Skrzypek, E., Cowan, C, and Straley, S.C. (1998). Targeting of the Yersinia pestis H 

o 

YopM protein into HeLa cells and intracellular trafficking to the nucleus. g 

Mol. Microbiol. 30, 1051-1065. * 

Sorg, I., Goehring, U.M., Aktories, K., and Schmidt, G. (2001). Recombinant 

Yersinia YopT leads to uncoupling of RhoA-effector interaction. Infect. Im- 

mun. 69, 7535-7543. 
Sory, M.P., Boland, A., Lambermont, I., and Cornells, G.R. (1995). Identification 

of the YopE and YopH domains required for secretion and internalization 

into the cytosol of macrophages, using the cyaA gene fusion approach. Proc. 

Natl. Acad. Sci. USA 92, 11,998-12,002. 
Stebbins, C.E. and Galan, J.E. (2001). Maintenance of an unfolded polypeptide by 

a cognate chaperone in bacterial type III secretion. Nature 414, 77-81. 
Underhill, D.M. and Ozinsky, A. (2002a). Phagocytosis of microbes: complexity 

in action. Annu. Rev. Immunol. 20, 825-852. 



Underbill, D.M. and Ozinsky, A. (2002b). Toll -like receptors: key mediators of 

microbe detection. Curr. Opin. Immunol. 14, 103-110. 
Viboud, G.I. and Bliska, J.B. (2001). A bacterial type III secretion system in- 
hibits actin polymerization to prevent pore formation in host cell membranes. 
EMBOJ. 20, 5373-5382. 
Von Pawel-Rammingen, U., Telepnev, M.V., Schmidt, G., Aktories, K., Wolf- 
Watz, H., and Rosqvist, R. (2000). GAP activity of the Yersinia YopE cytotoxin 
specifically targets the Rho pathway: a mechanism for disruption of actin 
microfilament structure. Mol. Microbiol. 36, 737-748. 
Wattiau, P., Woestyn, S., and Cornells, G.R. (1996). Customized secretion chap- 

erones in pathogenic bacteria. Mol. Microbiol. 20, 255-262. 
Wulff-Strobel, C.R., Williams, A.W., and Straley, S.C. (2002). LcrQ and SycH 
function together at the Ysc type III secretion system in Yersinia pestis to 
impose a hierarchy of secretion. Mol. Microbiol. 43, 411-423. 
S Yao, T., Mecsas, J., Healy, J. I., Falkow, S., and Chien, Y. (1999). Suppression of T 

8 and B lymphocyte activation by a Yersinia pseudotuberculosis virulence factor, 

S yopH. J. Exp. Med. 190, 1343-1350. 

g Yeh, E.T., Gong, L., and Kamitani, T. (2000). Ubiquitin-like proteins: new wines 

rt in new bottles. Gene 248, 1-14. 

g Zumbihl, R., Aepfelbacher, M., Andor, A., Jacobi, C.A., Ruckdeschel, K., Rouot, 

g B., and Heesemann, J. (1999). The cytotoxin YopT of Yersinia enterocolitica 

induces modification and cellular redistribution of the small GTP -binding 
£ protein RhoA. J. Biol. Chem. 274, 29,289-29,293. 

Oh 
W 

w 

O 




CO 

I— I 

h-l 
w 



CHAPTER 4 

Stealth warfare: The interactions of EPEC 
and EH EC with host cells 

Emma Allen- Vercoe and Rebekah DeVinney 



Although the Gram-negative bacterium Escherichia coli is normally consid- 
ered to be a harmless commensal of the gastrointestinal flora, there are some 
exceptions to this rule. In the past few decades it has become increasingly 
evident that there are serotypes off. coli that may cause disease in susceptible 
hosts. Disease states range from the invasive infections of the urinary tract 
caused by uropathogenic E. coli (UPEC) to the more typical diarrheal disease 
caused by several groups of E. coli serotypes, including enteropathogenic 
E. coli (EPEC) and enterohemorrhagic E. coli (EHEC). There is an increas- 
ing realization that bacteria-host interactions with pathogenic E. coli are far 
more complex and intricate than originally imagined. Studying the ways that 
bacteria such as EHEC and EPEC are able to subvert host cell functions to 
their own ends can be thought of as a "window" through which we are able 
to view the inner workings of the eukaryotic cell. In this chapter we examine 
some of the mechanisms by which EHEC and EPEC are able to coerce their 
host; we also examine the sequelae of these interactions. 

EPEC usually refers to E. coli serotypes 055:[H6], 086:H34, 0111:[H2], 
OH4:H2, OH9:[H6], 0127:H6, OU2:H6, and 0142:H34, where square 
brackets indicate the occurrence of nonmotile strains (Nataro and Kaper, 
1998). Infection with EPEC typically causes a chronic watery diarrhea, ac- 
companied by a low-grade fever and nausea. These symptoms can lead to 
rapid dehydration of the patient, and, because of this, EPEC is a leading 
cause of infant mortality, particularly in developing countries where rehydra- 
tion therapy may be difficult. In the Western world, isolated incidences of 
EPEC infection are often associated with daycares and nurseries (Nataro and 
Kaper, 1998), where young children closely associate with one another and 
facilitate the spread of infection. 





EHEC usually refers to serotype 0157:H7 and less commonly to serotype 
OllliH - . An EHEC infection is often heralded by the onset of watery diar- 
rhea, which progresses rapidly to severe bloody diarrhea (hemorrhagic coli- 
tis) in many patients, regardless of age (Nataro and Kaper, 1998). In the very 
young and very old, as well as in immunocompromised patients, the dis- 
ease can be complicated by the onset of hemolytic-uremic syndrome (HUS), 
which is characterized by hemolytic anemia, thrombocytopenia, and renal 
failure. HUS is caused by the secretion by EHEC of shiga-like toxin (SLT), 
a potent cytotoxin with a predilection for human kidney cells. A description 
of the mechanisms of action of SLT and the many effects on the host cell is 
beyond the scope of this chapter but is reviewed by O'Loughlin and Robins- 
Browne, 2001. The onset of HUS, even with rapid treatment, can prove fatal 
to a patient. Outbreaks of EHEC infection are becoming increasingly high 
g profile in North America and Europe. The low infectious dose required to 

g cause infection may allow the organism to spread rapidly. A recent outbreak 

g of EHEC 0157:H7 in Walkerton, Ontario infected close to 2000 people and 

3 led to seven deaths (Kondro, 2000). 

g Neither EPEC nor EHEC are generally regarded as invasive pathogens; 

rt however, some limited evidence has suggested that EPEC in particular may, 

5j in fact, be able to invade host cells under the right conditions (Czerucka et al., 

o 2000). Additionally, both EPEC and EHEC use a highly specialized mecha- 

n nism that allows the bacteria to adhere tightly to the epithelial cell surface and 

i subvert normal host cell functions by direct interaction. In this chapter, the 

jj mechanisms and consequences of both EPEC and EHEC infections of the 

2 host are discussed, and the subtle differences between EPEC and EHEC inter- 

w action with host cells, which are beginning to emerge as research progresses, 

are highlighted. 

ADHERENCE MECHANISMS REQUIRED FOR INITIAL 
ATTACHMENT OF EPEC AND EHEC TO HOST CELLS 

EPEC appear to have a predilection for the human ileum (Nataro and 
Kaper, 1998), and the first stage of adherence of EPEC to host cells is thought 
to involve bundle -forming pili (BFP; see Giron et al., 1991). BFP are type-4 
pili encoded by a cluster of 14 genes found on a 92 -kb plasmid known as the 
EPEC adherence factor plasmid (EAF; see Tobe et al., 1999) . EPEC-expressing 
BFP form dense microcolonies on the surface of tissue culture monolayers 
in a pattern known as localized adherence (LA). Whether BFP are involved 
more in interbacterial adherence or adherence to the host cell is currently 
under debate. However, there is no doubt that BFP expression represents a 



significant virulence factor for EPEC. Mutant EPEC unable to produce BFP 
are significantly reduced in their ability to adhere to tissue culture monolayers 
(Frankel et al., 1998). Furthermore, human volunteers given an inoculum of 
mutant EPEC defective for the expression of BFP were far less likely to develop 
diarrhea than were volunteers given the wild-type strain (Bieber et al., 1998). 
Together, these factors indicate that the initial adherence of EPEC to host 
cells is a critical step in the subsequent development of disease. 

The flagella of Gram-negative bacteria are being increasingly recognized 
as effectors of bacterial cell adhesion to host cells, such as in the binding 
of Salmonella to chick gut explants (Allen-Vercoe and Woodward, 1999). Re- 
cently, the flagella of EPEC have also been shown to be important in the 
adherence of the bacteria to epithelial cells in vitro, and experimental evi- 
dence suggested that a eukaryotic signal can induce flagellar expression and 
thus enhance adhesion (Giron et al., 2002). Additionally, it was found that h 

EPEC serotypes previously classified as nonmotile could, in fact, elaborate g 

flagella under the correct conditions (Giron et al., 2002). ^ 

In addition to its role in pedestal formation (discussed in what follows), g 

> 

the EPEC and EH EC outer membrane protein intimin may play a role in g 

adherence host cells. The family of intimins from attaching/effacing (A/E) g 




2 



lesion-forming bacteria has been divided into at least five distinct types that 
differ from each other at the amino acid level, particularly across the putative 8 

binding domain located at the C-terminus (Frankel et al., 1995). EPEC and n 

EH EC encode different intimin types, depending on their serotype. Intimin- ° 

a and intimin-/? are usually associated with EPEC strains, whereas EHEC 

0157:H7 specifically expresses intimin-y . Interestingly, intimin may be re- « 

sponsible for the tissue tropism exhibited by different pathogenic E. coli. n 

Infection with an EHEC strain expressing EPEC intimin-a shows a pattern § 

of adherence to human gut explants that resembles that of EPEC (Fitzhenry £ 

etal.,2002). <j 

The eukaryotic receptor(s) for intimin has yet to be identified. However, h 

two targets have been suggested: /3i-integrins (Frankel et al., 1996) and nu- ffi 

cleolin (Sinclair and O'Brien, 2002). The expression of intimin by EPEC and h 

EHEC is an important virulence determinant. Inhuman volunteers, an EPEC ? 

intimin - mutant was shown to be significantly attenuated compared with its 
wild-type parent strain (Donnenberg et al., 1993). In addition, antibodies to 
both EPEC and EHEC intimins are often found at a high titer in individuals 
who have been infected with these pathogens (Loureiro et al., 1998; Jenkins 
etal.,2000). 

The preferred site for colonization of EHEC is the colon, and the initial 
adherence of EHEC to host cells is thought to be quite different from that 



in 



of EPEC. EHEC does not possess the EAF plasmid and thus does not elabo- 
rate BFP. Hence, the pattern of EHEC adherence to epithelial cells in tissue 
culture is more diffuse than that of EPEC, and no microcolonies are formed. 
In the absence of BFP, the Tir (translocated intimin receptor) -independent 
adhesion mediated by intimin and described herein is likely to become more 
important to EHEC for colonization of host cells than it is for EPEC. Indeed, 
an intimin - mutant of EHEC 0157:H7 failed to colonize human intestinal 
explants (Fitzhenry et al., 2002). 

More recently, further EHEC adhesins have been proposed. Iha (IrgA 
homologue adhesin) is an outer membrane protein (OMP) with externally 
directed domains. Its corresponding gene, iha, confers an adhesive phenotype 
when transferred to a normally nonadherent E. coll K12 laboratory strain, 
although its role in EHEC adhesion remains to be elucidated (Tarr et al., 
£ 2000). The chromosomal gene efal (EHEC factor for adherence) has also been 

g recently implicated in the initial adherence of non-0157 EHEC to host cells, 

g and by virtue of its low G + C content may represent part of a pathogenicity 

3 islet acquired by horizontal transmission. The product of the efal gene, Efal, 

g has been shown to play a role in influencing the colonization of the bovine 

rt gut in vivo by non-0157 EHEC strains (Stevens et al., 2002). Although first 

5j characterized in EHEC, an Efa 1 homologue, LifA, has also been found in 

o EPEC (Nicholls et al., 2000), where it confers upon the bacteria the ability to 

n modulate host mucosal immunity in the gut (Klapproth et al., 1996; Malstrom 

i and James, 1998; also discussed later in this chapter). 




i-i 

< 



< 

s INTIMATE ATTACHMENT OF EHEC AND EPEC TO HOST CELLS, 



w 



AND THE TYPE III SECRETORY APPARATUS 

A hallmark of disease caused by EHEC and EPEC is the formation of 
A/E lesions (Fig. 4.1). These lesions are defined by the intimate attachment 
of the bacteria to the host cell surface and the localized destruction (efface- 
ment) of the brush border microvilli (Moon et al., 1983). The bacteria induce 
an accumulation of actin beneath their site of attachment, and this results 
in the formation of a pseudopod, or pedestal, upon which the bacteria are 
located (Fig. 4.2). A/E lesion-forming bacteria such as EHEC and EPEC uti- 
lize a Gram-negative specific mechanism, called a type III secretion system 
(TTSS) , in order to inject effector molecules directly into the host cell (Hueck, 
1998). 

For EPEC, a region of the chromosome known as the locus for entero- 
cyte effacement (LEE) is sufficient for A/E lesion formation. The LEE is a 
36-kb region containing 41 open reading frames (ORFs) organized into five 



EPEC 






J V 



12 3 4 

Figure 4.1. Simplified schematic of pedestal formation by EPEC. (1) EPEC uses BFP to 
adhere to the host cell microvilli, leading to effacement of the brush border (2), 
whereupon the bacterium uses its TTSS to inject Tir (black bulb shape), into the host cell, 
where it is phosphorylated (asterik) (3) Tir presents its intimin-binding domain to intimin 
expressed on the bacterial surface. The binding of Tir to intimin leads to a cascade of 
signaling events (4), which act to stimulate actin polymerization. A pedestal is formed on 
the host cell membrane, on which the bacterium resides. 



polycistronic operons, termed LEE1-LEE5 (Elliott et al., 1998; Mellies et al., 
1999). LEE1-LEE3 encode genes encoding the TTSS apparatus. LEE4 con- 
tains the genes required for pedestal formation. These include tat, which 
codes for intimin; tir, which codes for Tir; and a gene that codes for the 
molecular chaperone for Tir, cesT (Abe et al., 1999). LEE 5 contains genes for 
the type Ill-secreted proteins EspA, EspB, and EspD, which are necessary for 
pedestal formation. EspB and EspD are thought to form pore-like structures 
in the host cell membrane (Warawa et al., 1999; Kresse et al., 1999; Wachter 
et al., 1999) that interact with the filamentous EspA structures on the bac- 
terial cell to allow the close association of the type III secretory apparatus 
with the host cell membrane, and the subsequent insertion of Tir (Knutton 
et al., 1998). Both EHEC and EPEC EspB have recently been found to interact 
directly with the host protein a-catenin, a cytoskeleton-associated molecule, 
in a Tir-independent way, and the formation of A/E lesions has been found 
to be dependent on this interaction (Kodama et al., 2002). 

Recently, the supermolecular structure of the assembled EPEC TTSS 
apparatus has been elucidated by electron microscopy (Sekiya et al., 2001). It 




on 
H 
w 

> 
M 
H 

X 

i 

> 
w 

H 

X 
w 
i— i 

Z 
H 
M 

n 

H 

p— i 

O 

Z 

m 
O 

Tl 
W 

n 

> 

o 

M 

X 
w 
n 



H 

X 

X 
o 

m 
H 

n 

M 
t~< 

IS) 




w 

i— i 
> 
m 

Q 
X 

$ 
w 
M 
w 

Q 

< 
w 
O 

u 

> 

I 

w 

i-l 

w 





Figure 4.2. A cluster of pedestal structures induced by EPEC on the surface of a HeLa cell. 
Indirect immunofluorescence microscopy was used to demonstrate the formation of 
pedestal structures by EPEC. Long stalks of actin (green) are capped at the tip by Tir (red), 
upon which a single bacterium (blue) sits. The boxed region has been enlarged to clearly 
show a single pedestal. See color section. 



was reported that the EspA protein, which has no homology to any known 
protein in other bacterial TTSSs other than that for EH EC, may be a major 
component of a sheath -like structure that is assembled by the bacteria to act 
as a physical bridge between bacteria and host cells. This, in turn, indicates 
that the assembly of EHEC and EPEC type III secretons may be carried out 
somewhat differently compared with other pathogens. 



The Tir protein, once introduced into the host cell membrane by means 
of the TTSS, serves as the receptor for intimin and allows the intimate as- 
sociation of bacteria with host. Whereas EPEC adherence by means of BFP 
permits the bacteria to exist within 100-300 nm of the host cell surface, the 
distance between bacterial cell membrane to host cell membrane during Tir- 
intimin attachment can be as little as 10nm(NisanetaL, 1998). EPEC Tir was 
originally thought to be a host cell protein, and its identification as a bacterial 
protein, injected by the bacteria in order to effect intimate adhesion by in- 
timin, represents the first discovered example of host cell subversion of this 
type (Kenny et al., 1997). Work done in animal infection studies with the nat- 
ural rabbit pathogen, REPEC O103:H2, has demonstrated that intimin, Tir, 
EspA, and EspB are essential for diarrheal disease (Abe et al., 1998; Marches 
etal.,2000). 

Whereas some of the components of the TTSS are highly conserved be- h 

tween EPEC and EHEC, eae, espA, espB, espD, and tir are much less highly g 

conserved, although the mechanism of Tir translocation is thought to be ^ 

identical for both pathogens (Frankel et al., 1998). However, although it is g 

> 

possible to induce nonpathogenic laboratory strains of E. coli to form A/E le- g 

sions by introducing into them plasmids carrying the LEE region from EPEC g 




2 



(McDaniel and Kaper, 1997), the reciprocal experiment with the EHEC LEE 
region does not support either A/E lesion formation or Esp protein secretion, 8 

suggesting that, for EHEC, further factors are involved that are outside of the n 

LEE (Elliott et al., 1999; DeVinney et al., 2001). These unidentified factors ° 

required by EHEC are a current focus for research into EHEC pathogenicity. 

The first gene of LEE1 encodes the Ler regulator, an H-NS-like protein « 

that activates all other genes in the LEE for both EPEC and EHEC, and is n 

required for pedestal formation (Friedberg et al., 1999; Elliott et al., 2000). § 

The PerABC proteins encoded by the pEAF plasmid of EPEC positively reg- £ 

ulate BFP expression (Donnenberg et al., 1992) and may also regulate the 2 

transcription of LEE genes through ler (Mellies et al., 1999). h 

The regulation of the LEE appears to be quite different for EPEC and ffi 

EHEC. Both EHEC and EPEC utilize two quorum-sensing systems, termed h 

1 and 2, central to which are proteins termed autoinducer (AI)-l and AI-2, ^ 

respectively. During transition from the late exponential to the stationary 
growth phase, both EPEC and EHEC LEEs are activated by AI-2 (Sperandio 
et al., 1999). However, during the stationary phase, EHEC AI-1 appears to be 
downregulated by the expression of LEE-encoded genes through the action 
of a quorum-sensing regulator called SdiA (Kanamaru et al., 2000). In con- 
trast, the AI-1 of EPEC does not appear to be affected in this manner (Abe 
et al., 2002). The differences in EPEC and EHEC LEE gene expression can 



in 




< 



be illustrated by the finding that, in vitro, EH EC Tir synthesis and secretion 
requires specific growth conditions that do not apply to EPEC Tir expression 
(DeVinneyetaL, 1999). 



HOST CELL RESPONSES TO EHEC AND EPEC INFECTION 
Pedestal formation 

The formation of actin pedestals beneath adherent A/E bacteria is a host 
response so unusual that the presence of actin pedestals in infected tissue cul- 
ture monolayers can be used as a diagnostic test for EPEC and EHEC (Knutton 
et al., 1989). Pedestals may grow to 10 /im above the host cell surface, at a 
rate of around 1 /xm/min, and appear to be able to shorten and lengthen 
£ while remaining attached to the host cell (Rosenshine et al., 1996; Sanger 

g et al., 1996). It has been found that pedestals themselves are able to pro- 

g pel their attached bacteria along the host cell surface at an estimated rate 

3 of 70 nm/s (Sanger et al., 1996). The exact purpose of an actin pedestal is 

g puzzling. It may serve as a mechanism for the bacteria to avoid the host di- 

rt arrheal response, or as an avoidance tactic for internalization into the host 

§ cell. However, in tissue culture, cells infected with A/E bacteria eventually 

o round up and detach, suggesting that host cells may not be able to tolerate 

n prolonged contact with these pathogens (Baldwin et al., 1993). 

i A chief effector of pedestal formation, Tir, is a 72- to 78-kDA protein that 

jj is inserted into the host cell membrane by means of the TTSS apparatus. 



2 EPEC and EHEC strains that are unable to secrete Tir do not form pedestals 

w (DeVinney et al., 1999). Tir contains two predicted transmembrane domains 

that are required for the stable insertion of the protein into the host cell 
plasma membrane (Gauthier et al., 2000). The amino- and carboxy-termini 
of Tir are intracellular to the host cell, and intimin binds to a hydrophobic 
extracellular domain called the intimin-binding domain (IBD; see DeVinney 
et al., 1999; Kenny, 1999; Luo et al, 2000). 

The process of pedestal formation is best understood for EPEC (Fig. 4.3). 
EPEC Tir is tyrosine phosphorylated at position 474 (Y474) after insertion 
into the host cell membrane. Whereas Tir phosphorylation is required for 
pedestal formation (Kenny, 1999; Goosney et al., 2000), Tir translocation is 
thought to be completely independent of host cell modifications. In studies of 
EPEC attachment to red blood cells (RBCs), tyrosine phosphorylation of Tir 
and pedestal formation was not seen, although Tir was inserted correctly into 
the RBC membrane and was able to bind intimin (Shaw et al., 2002). The Tir 
chaperone, CesT, functions to direct Tir to the translocation apparatus and 
may help to maintain Tir in a secretion-competent state (Abe et al., 1999). 




Pedestal 
elongation 



Figure 4.3. Simplified schematic of EP EC-induced signaling events leading to pedestal 
formation. The binding of intimin to translocated Tir leads to tyrosine phosphorylation of 
the protein at position 474. This phosphorylation is required for the direct binding of the 
cellular adaptor Nek, which signals N-WASP recruitment to Tir, and the subsequent 
recruitment of the Arp2/3 complex to allow pedestal formation. The cytoskeletal proteins 
a-actinin, villin, cortactin, and vinculin all bind directly to Tir and are thought to play a 
role in pedestal formation through their F-actin binding sites. 

Several host cell structural proteins have been implicated in pedestal 
formation on the basis of observations made by use of immunofluorescence 
microscopy. Pedestals are predominantly formed from filamentous actin, 
with micro filament-associated proteins such as a-actinin, talin, ezrin, and 
villin also being found along the length of the pedestal (Goosney et al., 2001). 
The actin-stabilizing proteins tropomyosin and nonmuscle myosin II can be 
seen at the base of a pedestal (Sanger et al., 1996). The Tir protein serves 
directly as a conduit between host cell and bacteria, and much research has 
been focused on the mechanisms that enable this interaction, because this 
enigmatic protein offers us a unique window into the dynamics of host cell 
cytoskeletal rearrangement. 



on 
H 
w 

> 
M 
H 

X 

i 

> 

M 
H 

w 

I— I 

Z 
H 

M 

n 

H 

p— i 

O 

Z 

m 

O 
►n 

w 
w 

n 

> 

o 

M 

X 
w 
n 



H 

K 

o 

m 
H 

n 

M 








► 


Acquired 

immunity 




^ 



Tyrosine phosphorylation of Tir plays an essential role in pedestal for- 
mation by facilitating the direct binding of the cellular adaptor protein Nek 
(Gruenheid et al., 2001; Campellone et al., 2002). Nek binding leads to the re- 
cruitment of N-WASP, and the Arp2/3 complex, which in turn allows pedestal 
elongation (Kalman et al., 1999) . Cultured cells defective for the expression of 
N-WASP do not support the formation of E PEC actin pedestals, highlighting 
the central role played by this protein (Lommel et al., 2001). The critical role 
played by N-WASP suggests an involvement of Rho family GTPases. Sur- 
prisingly, these do not appear to play a role in pedestal formation (Ben-Ami 
etal, 1998; Ebel et al., 1998). 

Tir has also been demonstrated to bind directly to the focal adhesion (FA) 
proteins a-actinin, talin, and vinculin. FAs are anchoring structures that al- 
low the close association of epithelial cells and fibroblasts to the extracellular 
matrix (ECM) . Tir binds directly to a-actinin by means of its amino terminus h 

(Goosney et al., 2000; Huang et al., 2002), in a manner that is tyrosine phos- ^ 

phorylation independent (Goosney et al., 2000). Tir can also bind to talin, ^ 

and this association has been shown to be essential for correct F-actin focus- g 

> 

ing beneath adherent cells, possibly through the nucleation of nascent actin g 

filaments (Cantarelli et al., 2001). Vinculin has also been found to bind to g 



Figure 4.4. (facing page). EPEC-induced signaling events in the host cell. The interaction 
of EPEC effector molecules (gray boxes) with host proteins (white boxes) leads to the 
exploitation of various signaling pathways, and an array of outcomes for the host (rounded 
boxes). For details of individual pathways, see text. (For clarity, some less understood 
pathways have been omitted.) 



2 



Tir directly (Freeman et al., 2000). The association of FA proteins with the 

pedestal structure supports the notion that pedestals are composites of FAs 8 

and microvilli (Freeman et al., 2000; Goosney et al., 2001). Recent research n 

has shown that EPEC can disrupt FAs, by specifically dephosphorylating the ° 

FA kinase, FAK. FAK dephosphorylation requires a functional TTSS and is 

enhanced, but not absolutely dependent on, pedestal formation (Shifrin et al., « 

2002). Cortactin, an F-actin-binding protein, has been shown to accumulate n 

underneath adherent EPEC in its tyrosine-dephosphorylated form (Cantarelli § 

et al., 2002) . In this form, cortactin is able to efficiently bind other cytoskeletal £ 

proteins and to cross-link F-actin (Huang et al., 1997), and it may thus play 2 

an important role in pedestal formation. h 

Several signaling molecules are thought to be involved in pedestal forma- ffi 

tion (Fig. 4.4). Phospholipase C (PLC), phosphtidyinositol 3-kinase (PI3K), h 

and 5 -lipoxygenase have all been shown to be important for the accumulation g 
of a-actinin beneath adherent EPEC, because inhibitors of these molecules 



in 



also inhibited a-actinin accumulation in a concentration-dependent man- 
ner (Johnson-Henry et al., 2001). Recent evidence also suggests that the F- 
actin-binding protein, annexin 2, is recruited underneath the sites of EPEC 
attachment in epithelial monolayers (Zobiack et al., 2002) and may act as a 
bridge between actin and the host cell membrane (Gerke and Moss, 1997). 
Recruitment of annexin 2 by EPEC was found to be Tir independent, rais- 
ing the possibility that cytoskeletal rearrangements by EPEC may occur by at 
least two distinct pathways - one that is Tir mediated and another that is not 
(Zobiack etal, 2002). 

Although EH EC and EPEC both induce pedestal formation, EH EC uses 
methods distinct from EPEC to recruit the actin-nucleating machinery. EHEC 
Tir is less than 60% identical to EPEC Tir, with the C-terminal domain 
showing the least homology (Paton et al., 1998). Indeed, the absence of 
£ the tyrosine residue that is targeted for phosphorylation in EPEC Tir cor- 

g relates with the finding that EHEC is able to induce pedestal formation in 

g the absence of Tir tyrosine phosphorylation (Ismaili et al., 1995; DeVinney 

3 et al., 2001). Although EHEC Tir is not tyrosine phosphorylated and does not 

g bind Nek, both N-WASP and the Arp2/3 complex are still recruited to the 

rt EHEC pedestal (Goosney et al., 2001), which suggests that EHEC and EPEC 

5j modulate different pathways in order to form pedestals (DeVinney et al., 

o 2001). 

n A further major difference is that EHEC, but not EPEC, can form 

i pedestals with mutant EPEC Tir in which the Y474 has been replaced by 

jj phenylalanine (DeVinney et al., 2001; Kenny, 2001), suggesting that addi- 




< 



2 tional EH EC-specific bacterial factor(s) may be involved in pedestal forma- 

w tion by EHEC that is independent of Tir tyrosine phosphorylation (DeVinney 

et al., 2001). Although only a small proportion of EHEC strains, including 
EHEC 0157, form pedestals that do not contain tyrosine-phosphorylated Tir, 
several non-0157 serotypes that are designated as EHEC have Tir sequences 
similar to that of EPEC and are tyrosine phosphorylated in a similar way to 
EPEC Tir (DeVinney et al., 2001). Because the 0157:H7 serotype correlates 
closely with human EHEC infections , it is tempting to speculate that serotypes 
that form pedestals in a tyrosine-independent manner represent the more ef- 
ficient pathogens, as they are less reliant on the signaling capacities of the 
host cell in order to subvert host cell processes to their own ends. 



Inhibition of phagocytosis 

Several A/E pathogens, including EPEC, EHEC, and RE PEC, are able to 
inhibit uptake by specialist phagocytic cells such as M cells and macrophages, 



and they can even colonize the niches rich in these cells (Inman and Cantey, 
1983; Kresse et al., 2001). An ability to inhibit phagocytosis would clearly be 
advantageous to the bacteria by delaying the host immune response. 

The EPEC antiphagocytic phenotype has been demonstrated to be reliant 
on a functional TTSS and is independent of pedestal formation (Goosney 
et al., 1999). This study demonstrated that host protein tyrosine dephospho- 
rylation could be observed in type III secretion-competent EPEC, and that 
mutants that had nonfunctional type III secretion systems were unable to 
elicit host protein modifications, suggesting that antiphagocytosis is reliant 
on a bacterial protein effector that is secreted through the type III mecha- 
nism. More recent research was able to show that EPEC blocks its uptake by 
inhibition of a PI3K-mediated pathway (Celli et al., 2001). Whether EHEC 
and other A/E pathogens utilize the same mechanism for antiphagocytosis 
is unknown (although considered likely). EPEC, however, has the distinction h 

of being the first pathogen for which an antiphagocytic pathway of this type 
has been described. 



Inhibition of mitosis 

Certain strains of RE PEC and EPEC (although not the reference strain 




> 

M 

H 

X 

i 

> 

M 
H 

w 
I— I 



E2 348/69) are able to induce in HeLa cells an irreversible cytopathic effect, 3 

characterized by the progressive recruitment of stress fibers and FAs and n 

associated with the prevention of cell proliferation (De Rycke et al., 1997; 2 

Nougayrede et al., 1999). Such effects are not related to any toxin secretion, 

are dependent on a functional TTSS, and require the secreted proteins EspA, « 

EspB, and EspD, but not Tir or intimin (De Rycke et al., 1997; Nougayrede n 

et al., 1999; Marches et al., 2000). A closer examination of the observed cyto- § 

static effect revealed that the cell cycle is arrested in the G2 /M phase, and that £ 

this mitotic inhibition is not likely to be a consequence of the cytoskeletal 2 

arrangements themselves (Nougayrede et al., 2001). h 

The host cell protein, Cdkl, together with the protein cyclin Bl, forms a ffi 

complex that governs the transition from G2 to mitosis. In normal cells, Cdkl h 

is present at a constant level throughout the cell cycle, and its dephosphory- ? 

lation is required for entry into mitosis (Norbury and Nurse, 1992). Strains 
of RE PEC and EPEC that exert a cytostatic effect appear to affect Cdkl de- 
phosphorylation by an as yet unknown mechanism, arresting the cell cycle 
in the G2 phase (Nougayrede et al., 2001). Whether mitotic inhibition plays a 
role in EPEC pathogenesis, and by which exact method, can only be guessed 
at pending further research into this effect. However, it has been suggested 
that a cell-cycle arrest of the stem cells that supply cells to the intestinal villi 



in 



may delay the shedding of the epithelia, thus allowing the bacteria to remain 
attached to its host for a prolonged period (Nougayrede et al., 2001). 

Disruption of tight junctions 

Although as yet incompletely understood, the pathogenesis of EPEC- 
induced diarrhea may involve not only the formation of A/E lesions but also 
disruption of epithelial barrier function. Tight junctions are specialized inter- 
cellular structures that form "gaskets" around epithelial cells that prevent the 
leakage of fluid through the intercellular gaps, and the resultant diffusion of 
membrane proteins. Several studies have demonstrated that EPEC infection 
of host cells in tissue culture consistently leads to a significant decrease in 
transepithelial resistance (TER), a measure of tight junction integrity, in a 
£ time-dependent manner (Spitz et al., 1995). This host response has also been 

jz; 

g shown to be dependent on a functional TTSS (Canil et al., 1993; Philpott 
g et al. , 1 996) , although attachment of bacteria mediated by Tir-intimin interac- 
ts tions was not in itself sufficient to trigger host cell membrane depolarization 
g (Stein etal, 1996). 

rt EPEC-induced phosphorylation of myosin light chain (MLC) seems to 

§ play a central role in the pathway that leads to eventual loss of cell membrane 

o integrity (Manjarrez-Hernandez et al., 1996). Both EPEC and EHEC infec- 

n tions in vitro have been shown to activate protein kinase C (PKC), which is 

i likely to lead to MLC phosphorylation (Crane and Oh, 1997; Philpott et al., 

jj 1998). Activation of MLC-kinase (MLCK) leads to further phosphorylation of 




< 



2 MLC (Manjarrez-Hernandez et al., 1996) and contraction of the cytoskeletal 

w rings underlying tight junctions, and thus it leads to a loss of tight junction 

integrity. 

The host cell protein occludin is thought to play an important role in 
maintaining the integrity of tight junctions (McCarthy et al., 1996). EPEC 
infection has been shown to result in dephosphorylation of occludin and its 
resulting dissociation from tight junctions into the intracellular space, which 
would very likely contribute to tight junction permeability (Simonovic et al., 
2000). Whether this occludin dephosphorylation is linked to the pathway in- 
volving the phosphorylation of MLC, or whether the two pathways are sepa- 
rate, is unknown. Occludin is one of many proteins identified in the makeup 
of tight junctions, and whether EPEC targets any other tight junction pro- 
teins has not yet been determined. Interestingly, EHEC has been shown to 
alter the distribution of another important tight junction-associated protein, 
ZO-1, again through the action of PKC (Philpott et al., 1998), which directly 
targets ZO-1 for phosphorylation (Stuart and Nigam, 1995). 



Recently it was found that EPEC mutants deficient in their ability to ex- 
press EspF, a protein that is translocated into the host cell cytoplasm by the 
TTSS, were deficient in their abilities to disrupt electrical resistance across 
polarized epithelial cells (McNamara et al., 2001). EspF appears to rely on a 
chaperone protein, CesF, for full translocation efficiency (Elliott et al., 2002). 
Once translocated into the host cell, EspF is not found distributed homoge- 
nously; instead it appears to be sequestered, perhaps by the host cell protein(s) 
with which it interacts (McNamara et al., 2001). Whether EspF does indeed 
modulate host functions through interaction with a host cell protein remains 
to be elucidated. 



Upregulation of chloride secretion and adenosine 
triphosphate (ATP) release 




on 
H 
w 

> 



One of the hallmark symptoms of EPEC infection is the onset of severe, g 

watery diarrhea. PKC activation in the intestine triggers a rapid secretion ^ 

of ions (particularly chloride ions, Cl~) and fluid into the lumen (Beubler g 

> 

and Schirgi-Degen, 1993). Thus it was not surprising to find that EPEC is g 

able to activate PKC (Crane and Oh, 1997), which is likely to result in the g 

upregulation of Cl~ secretion and the subsequent development of watery ~ 

diarrhea. 3 

A further mechanism whereby watery diarrhea might be induced in- n 

volves the galanin-1 receptors found on the surface of epithelial cells lining g 

the human GI tract. These receptors, on activation, cause Cl~ secretion. Ex- 

pression of the galanin-1 receptor is transcriptionally regulated by nuclear « 

factor (NF)-/cB, which is in turn activated during infection with both EH EC n 

and EPEC (Savkovic et al., 1997; Dahan et al., 2002). Indeed, it was demon- § 

strated that infection with both EH EC and EPEC increased the number of £ 

galanin-1 receptors by means of the activation of NF-/cB, and thus elevated 2 

Cl~ secretion (Hecht et al., 1999), a factor that contributes to the etiology of h 

infectious watery diarrhea. ffi 

In addition to NF-kB-PKC activation, a further hypothesis for the patho- h 

genesis of diarrhea promoted by EPEC infection has emerged. It was recently ? 
demonstrated that polarized cells could be induced to release a large concen- 
tration of ATP into the apical culture medium upon infection with EPEC 
(Crane et al., 2002). This release of ATP is TTSS dependent, and, in partic- 
ular, the EspF protein appears to play an important role (Crane et al., 2002). 
ATP release in response to EPEC infection may be a direct result of leakage 
from the ATP-rich cytosol through pores formed by the bacterial TTSS. The 

o 

pore created by EPEC in the host cell membrane is between 30 and 50 A (Ide 



in 




et al., 2001), which would allow for easy passage of the ATP molecule. ATP 
release may also be effected by the modulation by EPEC of the cystic fibrosis 
transmembrane regulator (CFTR; see Crane et al., 2002), which acts as a cel- 
lular "valve" to allow for the normal release of ATP from the host cell cytosol 
during cell stretching or swelling or as a response to cAMP stimulation (Jiang 
etal., 1998). 

Once extracellular, ATP is rapidly broken down into less phosphory- 
lated nucleotides and adenosine. Adenine nucleotides released from EPEC- 
infected cells could spread to neighboring cells and may trigger a fluid se- 
cretory response. This could explain the apparent paradox that EPEC prefer- 
entially adheres to the villi of the gut epithelia, yet it is the crypt cells of the 
gut that have the capacity to generate watery diarrhea through ion secretion 
(Crane et al., 2002). The exact mechanism of the stimulation of ATP release 
g by EPEC, and whether the same mechanism is utilized by EH EC, remains to 

fc be elucidated. 

g The secretion of Cl~ and the subsequent development of watery diarrhea 

3 may be considered an innate host defense mechanism that has evolved to 

g rid the host of enterobacterial pathogens such as EHEC and EPEC. For the 

rt pathogen itself, this defense mechanism has the advantage that it may allow 

5j for increased spread of the bacteria from host to host. 

w 

O 
U 

* The Ca 2 + controversy 

w 

jj Early research into the effects on host cells of A/E lesion-forming EPEC 

2 demonstrated that near to the sites of bacterial attachment there seemed 

w to be a localized elevation of Ca 2+ within the host cell (Baldwin et al., 1991), 

coupled with the phosphorylation of (among other proteins) MLC (Manjarrez- 
Hernandez et al., 1991). Subsequent to this, increased levels of inositol 
triphosphate (IP) were detected in EPEC-infected cells (Dytoc et al., 1994; 
Foubister et al., 1994), and thus it seemed plausible that the PLC pathway, 
which generates IP and in turn a release of Ca 2+ from intracellular stores, 
was being stimulated by EPEC infection. Additionally, it was found that the 
buffering of intracellular free Ca 2+ by use of the chelating agent BAPTA 
could prevent A/E lesion formation by EPEC (Baldwin et al., 1991; Dytoc 
etal., 1994). 

Recent work using more sensitive methods has been able to measure 
both temporal and spatial measurements of Ca 2+ in live cells infected with 
both EPEC and EHEC (Bain et al., 1998). In contrast to the earlier studies 
just described, no increase in Ca 2+ levels could be measured, and the addi- 
tion of BAPTA to host cells was not seen to prevent the formation of A/E 
lesions (Bain et al., 1998). It was suggested that the alterations in cell Ca 2+ 



levels seen in previous studies was related to the cytotoxic effects of E PEC 
(Bain et al., 1998). Clearly, more research has to be done in this area to deter- 
mine whether or not EPEC and EH EC can modulate host intracellular Ca 2+ 
levels. 

Activation of mitogen-activated protein (MAP) kinase cascades 

The MAP kinases group includes extracellular signal-related protein ki- 
nases 1 and 2 (ERK1 and ERK2), p38 MAP kinase, and c-Jun N-terminal 
kinase (JNK). As well as the classical mitogenic response, MAP kinases 
have also been implicated in the regulation of stress responses, cytokine 
and chemokine responses, and cytoskeletal reorganization (Garrington and 
Johnson, 1999). Indeed, EH EC is able to induce the expression of the 
chemokine interleukin-8 ( I L-8) in host cells via MAP kinase pathways ( Dahan h 

et al., 2002; discussed later). g 

MAP kinases play a role in the host cell response to infection with inva- ^ 

sive bacteria such as Salmonella (Hobbie et al., 1997). Although invasion is g 

> 
not thought to be a central process required for EPEC pathogenicity, EPEC g 

was found to be able to invade T84 epithelial cells in culture, and this inva- g 



w 



2 



sion relied on the activation of ERK1/2 MAP kinase (Czerucka et al., 2000). 

Additionally, it was found that the intimate attachment of EPEC to host cells 8 

was able to activate ERK1/2, p38, and JNK signaling pathways in a manner n 

that was dependent on both a functional TTSS and the expression of intimin ° 

(Czerucka et al., 2001) . The MAP kinase-activated pathways proceed through 

a cascade of phosphorylation events leading to the transcriptional activation « 

of certain genes. Recently EPEC was found to upregulate the transcription n 

of the gene coding for the early growth response factor-1 (Egr-1) through § 

the ERK1/2 pathway, the p38 pathway, or both (de Grado et al., 2001). Egr-1 £ 

is a protein that is activated in most cell types in response to stress, and it 2 

was found that the mouse egr-1 gene was upregulated during in vivo infection h 

with the EPEC-related pathogen, Citrobacter rodentium (de Grado et al., 2001), ffi 

demonstrating that this effect is not isolated to cultured cells. The functional h 

consequences of upregulation of Egr-1 production by infected cells remain ? 
to be determined and may prove to be a key part of the pathogenic strategy 
of these bacteria. 



Modulation of apoptosis 

The induction of apoptosis plays an important role in the regulation of the 
immune response to bacterial infections, although the process of apoptosis 
may be hijacked by the pathogen to its own advantage (Muller and Rudel, 



in 




2001). Several enteric pathogens are known to trigger apoptosis in host cells 
both in vitro and in vivo (Muller and Rudel, 2001). 

For EPEC, the induction of apoptosis in host cells would seem to be 
counterintuitive because the bacteria rely on their adhesion to the host ep- 
ithelia in order to effect A/E lesion formation. Several studies have found 
that EPEC may actually act to slow down the apoptotic process during in- 
fection. For example, although EPEC-infected cells eventually succumbed to 
cell death that had hallmark features of apoptosis, such as early expression of 
phosphatidylserine (PS) on host cell surfaces and internucleosomal cleavage 
of DNA, the time taken for this effect to become apparent was much greater 
than for apoptosis induced by invasive bacteria (Crane et al., 1999). More 
recently, a comprehensive in vivo study in rabbits infected with RE P EC 103 
demonstrated that although apoptosis physiologically occurs in the rabbit 
g ileum, particularly at the tips of the absorptive villi, infection with REPEC 

g did not promote the rate of apoptosis but may actually have diminished it 

g (Heczko et al., 2001). Thus the modulation of apoptosis during infection by 

3 EPEC and EH EC could be an important bacterial strategy used both to in- 

g crease the likelihood of bacterial attachment and to control the rate of (and 

* downstream effects of) host cell death. 

5j The stimulation of tyrosine kinases, PKC, and the transcription factor NF- 

o k B all suppress cell death, and E PEC is able to effectively activate these factors 

n (Crane and Oh, 1997; Savkovic et al., 1997). Interestingly, an EPEC secreted 

i protein, EspF, has been shown to effect cell death in a manner that has fea- 

jj tures compatible with pure apoptosis (Crane et al., 2001). Conversely, EPEC 

2 BFP may play a role in the induction of apoptosis, because nonpathogenic 

w E. coli strains expressing BFP genes induced significant levels of apoptosis 

in host epithelial cells (Abul-Milh et al, 2001). 

The role that EHEC plays in the modulation of apoptosis has not been 
so extensively studied, although from the limited data available it appears 
that EHEC is also able to effect apoptosis in a manner that is dependent on 
bacterial attachment and occurs in a similar timeframe to apoptotic cell death 
induced by EPEC (Barnett Foster et al., 2000). The best-characterized toxin of 
EHEC, SLT, is also known to induce apoptosis (Jones et al., 2000), although 
the rate of this toxin-mediated apoptosis is much slower than that triggered 
by EPEC and EHEC attachment (Barnett Foster et al., 2000). Interestingly, 
this study also pointed to a rationale for EPEC and EHEC adhesion, because 
one of the physiological effects of apoptotic cell death is the upregulation of 
levels of phosphatidylethanolamine (PE) in the cell membrane, a molecule 
used by both EPEC and EHEC for bacterial attachment (Barnett Foster et al., 
1999). 




Disruption of mitochondrial membrane potentia 

In agreement with the observation of host cell apoptotis in response to 
EPEC infection, it has recently emerged that the gene orfl9, upstream of tir, 
codes for a protein that seems to interfere with the mitochondrial ability to 
maintain its membrane potential, a proapoptotic trigger (Kenny and Jepson, 
2000). The precise mechanism of action of the Orfl9 protein (since renamed 
Map, for mitochondrial-associated protein) and the consequences of its in- 
teractions with mitochondria have yet to be elucidated, although the delivery 
of Map into the host cell cytoplasm by means of the TTSS and its subsequent 
targeting directly to the mitochondria represents the first discovered exam- 
ple of bacteria-host cell interactions of this type (Kenny and Jepson, 2000). 
Whether EH EC is able to target mitochondria in a similar manner remains to 
be elucidated, but this seems likely if one considers the possession by EH EC 
of a Map homologue (Perna et al., 2001). > 

Although it may seem counterintuitive for EPEC to possess mechanisms x 

that act both to prevent and induce apoptosis, in fact it demonstrates the > 

extraordinary ability of this pathogen to control a central host cell function. > 

The downstream effects of modulation of apoptosis are diverse and may be ^ 

advantageous to the bacterium at different stages of infection. It is possible M 

that bacterial genes that affect apoptosis are regulated in response to diverse g 

environmental stimuli encountered during the disease process. g 

n 

H 
i— i 

O 

Z 

m 

Modulation of Cdc42 activity ° 



on 
H 






Early during infection, EPEC is able to trigger transient filopodia-like n 



> 



cytoskeletal rearrangements distinct from those stimulated by Tir-intimin § 

pedestal formation (Kenny et al., 2002). Interestingly, it was found that if £ 

Map expression was increased by expression of orfl 9 from a plasmid, then 2 

the formation of filopodia in infected cells was stimulated, although this stim- h 

ulation was independent of any observable effect on the mitochondria ( Kenny x 

et al., 2002). This second role for Map was demonstrated to be dependent on h 

the activation of the host GTPase, Cdc42 (Kenny et al., 2002), a molecule that ? 

is not required for pedestal formation triggered by Tir and intimin (Kalman 
et al., 1999), although it plays a role in cytoskeletal modification in uninfected 
cells. 

Surprisingly, the transient nature of filopodia formation was found to be 
a result of further interactions of Cdc42 with Tir, which serve to deactivate 
the host GTPase. In this respect, Tir may also have a dual function, behaving 
as a GTPase-activating protein (GAP) and a signal-transducing molecule, 



in 




w 

i— i 
> 
m 

Q 



as well as a receptor for intimin. To modulate Cdc42 activity, Tir seems to 
require its interaction with intimin, although this function of the Tir-intimin 
complex is separate from its ability to induce pedestal formation (Kenny et al., 
2002). 

This fascinating interplay between two bacterial proteins that seem to act 
on host cell functions in an opposing manner appears to be counterintuitive 
at first. It is thought that the Cdc42 -dependent signaling mediated by Map 
is inhibitory to pedestal formation (Kenny et al., 2002), and because pedestal 
formation is important to virulence, a bacterial mechanism to downregulate 
the Cdc42 -mediated pathway is necessary. Why Map induces Cdc42 in the 
first place is a mystery that remains to be solved, although Map is undoubtedly 
important to the infection process (because, if it were not, its function would 
likely have been abolished by natural selection) . 

Modification of host immune responses 



x Innate immunity 

g Infection of host cells with microbial pathogens, including EPEC and 

rt EHEC, leads to a proinflammatory response by means of the triggering of 

§ various cellular signals that activate the NF-/cB pathway and stimulate ex- 

o pression of IL-8 (Kresse et al., 2001; Savkovic et al., 2001). EPEC has been 

n found to exploit the ERK1/2 pathway and to induce inflammation via these 

g routes (Czerucka et al., 2000, 2001; Savkovic et al., 2001). EHEC infection in 

jj vitro results in phosphorylation of ERK1/2 as well as two further groups of the 

2 MAPK family, p38 and JNK, which activate NF-/cB as well as the transcription 

w factor, AP-1 (Dahan et al., 2002). These transcription factors in turn regulate 

the expression of IL-8. It is thought that the stimulation of these pathways is 

likely to be effected by the secretion of as yet unknown bacterial factors by 

means of the EPEC-EHEC TTSS (Dahan et al, 2002). 

When histological specimens from the intestines of animals infected 
with EPEC are examined, there is evidence for a dramatic infiltration by 
inflammatory cells, in particular, polymorphonuclear leukocytes (PMNs) 
(Moon et al., 1983). The exact advantage to the pathogen of PMN attraction 
to the site of infection remains to be elucidated, although it is known that 
recruited PMNs release prostaglandins, which in turn increase the activity of 
adenylate cyclase in intestinal cells. As a result, cAMP levels increase, leading 
to a release of Cl~ by the cells. As already discussed, release of Cl~ contributes 
to the development of watery diarrhea, a hallmark of both EPEC and EHEC 
infection and a likely mechanism by which the bacteria facilitate their spread 
from host to host. 




One further advantage of PMN infiltration, to SLT-producing EHEC in 
particular, has recently been described. The genes encoding SLT in EHEC are 
carried as part of lysogenic bacteriophages, and it is thought that the induc- 
tion of these phages during infection contributes to EHEC pathogenesis by 
increasing the transcription and copy number of the phage genes (Plunkett 
et al., 1999). It was recently found that phage induction during EHEC infec- 
tion may be due, at least in part, to the hydrogen peroxide released by neu- 
trophils as a component of their antibacterial arsenal (Wagner et al., 2001). 
Indeed, SLT has been shown to inhibit neutrophil apoptosis (Liu et al., 1999), 
which may prolong H2O2 release and promote production of SLT. 

Adaptive immunity 

The severity of E PEC and EHEC infections tends to correlate with the age 
of the patient, with infants and young children being most at risk for the de- h 

velopment of disease . It has been found that infants are more likely to develop g 

X 

diarrhea during their first colonization with EPEC than they are during sub- ^ 

sequent exposures to the pathogen (Cravioto et al., 1990), although whether g 

> 

this is also true of EHEC infections of infants is unknown. In adult volun- g 

teers infected with EPEC, there was no specific effect of prior EPEC infection g 

on the incidence of diarrhea, although disease severity was reduced in indi- ~ 

viduals who were reinfected with a homologous EPEC strain (Donnenberg 3 

et al., 1998). Effects such as these may be the result of acquired (adaptive) n 

immunity in the host. ° 

Several studies have been undertaken in an effort to identify the antibody 

responses generated to specific EPEC and EHEC antigens during the acute « 

and later stages of disease. Given the key roles that EHEC and EPEC factors, n 

such as Esps, Tir, intimin, and (for EPEC) BFP, play in the virulence of § 

these pathogens, studies have usually focused on the immune responses £ 

generated to these bacterial factors in both natural and volunteer infection 
studies. 

Human volunteers experimentally infected with EPEC developed a x 

strong IgG response against intimin (Nataro and Kaper, 1998), although in h 

subsequent studies of natural infections of both EHEC and EPEC, an anti- ? 

body response to intimin has varied in strength from strong to undetectable 
(Li et al., 2000). Immune responses to EspA and EspB were similarly variable 
in patients naturally infected with EPEC (Martinez et al., 1999). In sera from 
patients infected with EHEC0157:H7, there was little reactivity to EspA and 
EspB during the acute phase of illness, although this titer was increased in 
the later stages of disease. Tir was shown to induce a significant antibody 
response in EHEC infections (Li et al., 2000), and this response was found 



n 



H 

X 



in 



to persist after infection (Li et al., 2000). BFP was found to elicit an IgG res- 
ponse, but not an IgA response, in children infected with BFP + EPEC strains 
(Martinez et al, 1999). 

Interestingly, a study of secretory IgA purified from breast milk from 
mothers who were living in areas where EPEC was endemic showed that 
these antibodies prevented the adherence of EPEC to cultured epithelial cells 
(Cravioto et al., 1991). The antibodies reacted to many EPEC proteins, several 
of them unknown (Manjarrez-Hernandez et al., 2000) . This illustrates, in part, 
the complexity of the human immune response to EPEC (which is likely to be 
reflected by that to EH EC) and indicates that there is still much to be learned 
before a suitable molecular target is identified for vaccine production. 

Lysates of both EPEC and EHEC are known to inhibit lymphokine pro- 
duction by lymphoid cells from multiple sites (Malstrom and James, 1998). 
£ The nature of this inhibition was recently partially characterized when it 

g was found that a large gene present in EPEC encoded a toxin that specifi- 

g cally acted to inhibit lymphocyte proliferation, and IL-2, IL-4, and gamma 

3 interferon (IFN-y) production (Klapproth et al., 2000). The toxin was named 

g lymphostatin and is one of the largest bacterial toxins known. Homologues of 

rt lifA (encoding lymphostatin) are found in most strains of EPEC, some strains 

5j of EHEC (but not 0157 strains) where it is designated Efal, and C. rodentium 

o (Klapproth et al., 2000), correlating with an A/E ability of the pathogen. 

n The mechanism of action of lymphostatin is yet to be determined because 

i it is difficult to purify (Klapproth et al., 2000). However, its effects are limited 

jj to lymphokine expression and appear to cause no changes in the epithelial cell 




< 



2 cytoskeleton (Klapproth et al., 2000) . A potential consequence of lymphostatin 

w expression may be the suppression of an adaptive immune response to the 

bacteria, thus prolonging the infection and increasing the likelihood that the 

pathogen will be passed on to new hosts. 



CONCLUSIONS 

By now, it should be clear that the pathogenesis of EPEC and EHEC 
infection is not a simple process. EPEC and EHEC have evolved myriad 
strategies to subvert host cell functions that, when combined, give rise to the 
etiology of the diseases that they cause. The host is undoubtedly not passive 
during infection with these bacteria, but the outcome of infection rests on 
the delicate balance of host cell processes, some of which are competing in a 
molecular tug of war controlled by the host on one side and the bacteria on 
the other. With the advent of increasing bacterial resistance to antibiotics, we 
can no longer rely on antimicrobial compounds for the effective treatment of 




bacterial infections. To design better therapeutics, we will need to target the 
mechanisms the bacteria use to control their hosts. 

Although we have been forced to find alternative methods to treat in- 
fections with bacteria such as EPEC and EH EC, this coercion has not been 
without benefit. The study of bacteria-host cell interactions has given new 
dimension to the field of cell biology, and as we learn more about the ways 
in which bacteria interact with us, we will undoubtedly discover levels of 
complexity within the host cell far beyond those we have studied to date. 
Pathogens themselves have become the cell biologist's working tools and are 
providing fascinating insights into the workings of the eukaryotic cell. 

ACKNOWLEDGMENTS 

As a result of space limitations, we could not reference original sources h 

for all data cited in this review; original sources can be found in the reviews g 

cited. Research in the DeVinney Lab is supported by Canadian Institutes ^ 

for Health Research, and the Alberta Heritage Foundation for Medical Re- g 

> 

search (AHFMR). E. Allen-Vercoe is an AHFMR postdoctoral fellow, and g 

R. DeVinney is an AHFMR scholar. g 

w 

I— I 

Z 
H 

REFERENCES g 

n 

H 
_ i— i 

Abe, A., de Grado, M., Pfuetzner, R.A., Sanchez- Sanmartin, C, Devinney, R., ° 

Puente, J.L., Strynadka, N.C., and Finlay, B.B. (1999). Enteropathogenic Es- o 

cherichia coli translocated intimin receptor, Tir, requires a specific chaperone « 

for stable secretion. Mol. Microbiol. 33, 1162-1175. n 

> 

Abe, A., Heczko, U., Hegele, R.G., and Finlay, B.B. (1998). Two enteropatho- g 

genie Escherichia coli type III secreted proteins, EspA and EspB, are virulence x 

factors. J. Exp. Med. 188, 1907-1916. <j 

Abe, H., Tatsuno, I., Tobe, T., Okutani, A., and Sasakawa, C. (2002). Bicarbon- h 

ate ion stimulates the expression of locus of enterocyte effacement-encoded ffi 

genes in enterohemorrhagic Escherichia coli 0157:H7. Infect. Immun. 70, h 

3500-3509. 

Abul-Milh, M., Wu, Y., Lau, B., Lingwood, C.A., and Foster, D.B. (2001). In- 
duction of epithelial cell death including apoptosis by enteropathogenic 
Escherichia coli expressing bundle -forming pili. Infect. Immun. 69, 7356- 
7364. 

Allen-Vercoe, E. and Woodward, M.J. (1999). The role of flagella, but not fim- 
briae, in the adherence of Salmonella enterica serotype Enteritidis to chick 
gut explant. J. Med. Microbiol. 48, 771-780. 



n 

M 
t- 
M 

in 




Bain, C, Keller, R., Collington, G.K., Trabulsi, L.R., and Knutton, S. (1998). In- 
creased levels of intracellular calcium are not required for the formation of 
attaching and effacing lesions by enteropathogenic and enterohemorrhagic 
Escherichia coli. Infect. Immun. 66, 3900-3908. 
Baldwin, T.J., Lee-Delaunay, M.B., Knutton, S., and Williams, P.H. (1993). 
Calcium-calmodulin dependence of actin accretion and lethality in cultured 
HEp-2 cells infected with enteropathogenic Escherichia coli. Infect. Immun. 
61, 760-763. 
Baldwin, T.J., Ward, W., Aitken, A., Knutton, S., and Williams, P.H. (1991). El- 
evation of intracellular free calcium levels in HEp-2 cells infected with en- 
teropathogenic Escherichia coli. Infect. Immun. 59, 1599-1604. 
Barnett Foster, D., Abul-Milh, M., Huesca, M., and Lingwood, C.A. (2000). En- 
terohemorrhagic Escherichia coli induces apoptosis which augments bacterial 
£ binding and phosphatidylethanolamine exposure on the plasma membrane 

g outer leaflet. Infect. Immun. 68, 3108-3115. 

g Barnett Foster, D., Philpott, D., Abul-Milh, M., Huesca, M., Sherman, P.M., and 

3 Lingwood, C.A. (1999). Phosphatidylethanolamine recognition promotes en- 

g teropathogenic E. coli and enterohemorrhagic E. coli host cell attachment. 

3 Microb. Pathog. 27, 289-301. 

5j Ben-Ami, G., Ozeri, V., Hanski, E., Hofmann, F., Aktories, K., Hahn, K.M., 

o Bokoch, G.M., and Rosenshine, I. (1998). Agents that inhibit Rho, Rac, and 

n Cdc42 do not block formation of actin pedestals in HeLa cells infected with 

fc enteropathogenic Escherichia coli. Infect. Immun. 66, 1755-1758. 

Sj Beubler, E. and Schirgi-Degen, A. (1993). Stimulation of enterocyte protein kinase 

2 C by laxatives in-vitro. J. Pharm. Pharmacol. 45, 59-62. 

w Bieber, D., Ramer, S.W., Wu, C.Y., Murray, W.J., Tobe, T., Fernandez, R., 

and Schoolnik, G.K. (1998). Type IV pili, transient bacterial aggregates, 
and virulence of enteropathogenic Escherichia coli. Science 280, 2114- 
2118. 
Campellone, K.G., Giese, A., Tipper, D.J., and Leong, J.M. (2002). A tyrosine- 
phosphorylated 12-amino-acid sequence of enteropathogenic Escherichia coli 
Tir binds the host adaptor protein Nek and is required for Nek localization 
to actin pedestals. Mol. Microbiol. 43, 1227-1241. 
Canil, C., Rosenshine, I., Ruschkowski, S., Donnenberg, M.S., Kaper, J.B., and 
Finlay, B.B. (1993). Enteropathogenic Escherichia coli decreases the transep- 
ithelial electrical resistance of polarized epithelial monolayers. Infect. Immun. 
61, 2755-2762. 
Cantarelli, V.V., Takahashi, A., Yanagihara, I., Akeda, Y., Imura, K., Kodama, T., 
Kono, G., Sato, Y., and Honda, T. (2001). Talin, a host cell protein, inter- 
acts directly with the translocated intimin receptor, Tir, of enteropathogenic 




00 



> 

H 



Escherichia coli, and is essential for pedestal formation. Cell. Microbiol. 3, 
745-751. 

Cantarelli, V.V., Takahashi, A., Yanagihara, I., Akeda, Y., Imura, K., Kodama, 
T., Kono, G., Sato, Y., Iida, T., and Honda, T. (2002). Cortactin is necessary 
for F-actin accumulation in pedestal structures induced by enteropathogenic 
Escherichia coli infection. Infect. Immun. 70, 2206-2209. 

Celli, J., Olivier, M., and Finlay, B.B. (2001). Enteropathogenic Escherichia coli 
mediates antiphagocytosis through the inhibition of PI 3-kinase-dependent 
pathways. EMBOJ. 20, 1245-1258. 

Crane, J.K., Majumdar, S., and Pickhardt, D.F. Ill (1999). Host cell death due 
to enteropathogenic Escherichia coli has features of apoptosis. Infect. Immun. 
67, 2575-2584. 

Crane, J.K., McNamara, B.P., and Donnenberg, M.S. (2001). Role of EspF in host 

cell death induced by enteropathogenic Escherichia coli. Cell. Microbiol. 3, h 

197-211. 

Crane, J.K. and Oh, J.S. (1997). Activation of host cell protein kinase C by en- ^ 

teropathogenic Escherichia coli. Infect. Immun. 65, 3277-3285. g 

Crane, J.K., Olson, R.A., Jones, H.M., and Duffey, M.E. (2002). Release of ATP g 

during host cell killing by enteropathogenic E. coli and its role as a secretory g 

mediator. Am. J. Physiol. Gastrointest. Liver Physiol. 283, G74— G86. 

Cravioto, A., Reyes, R.E., Trujillo, F., Uribe, F., Navarro, A., De La Roca, J.M., 3 

Hernandez, T.M., Perez, G., and Vazquez, V. (1990). Risk of diarrhea during n 

the first year of life associated with initial and subsequent colonization by § 

specific enteropathogens. Am. J. Epidemiol. 131, 886-904. o 

Cravioto, A., Tello, A., Villafan, H., Ruiz, J., del Vedovo, S., and Neeser, J.R. « 

(1991). Inhibition of localized adhesion of enteropathogenic Escherichia coli n 

to HEp-2 cells by immunoglobulin and oligosaccharide fractions of human § 

colostrum and breast milk. J. Infect. Dis. 163, 1247-1255. x 

Czerucka, D., Dahan, S., Mograbi, B., Rossi, B., and Rampal, P. (2001). Impli- 2 

cation of mitogen-activated protein kinases in T84 cell responses to entero- h 

pathogenic Escherichia coli infection. Infect. Immun. 69, 1298-1305. ffi 

Czerucka, D., Dahan, S., Mograbi, B., Rossi, B., and Rampal, P. (2000). Saccha- h 

romyces boulardii preserves the barrier function and modulates the signal ? 

transduction pathway induced in enteropathogenic Escherichia coii-infected 
T84 cells. Infect. Immun. 68, 5998-6004. 

Dahan, S., Busuttil, V., Imbert, V., Peyron, J.F., Rampal, P., and Czerucka, D. 
(2002). Enterohemorrhagic Escherichia coli infection induces interleukin-8 
production via activation of mitogen-activated protein kinases and the tran- 
scription factors NF-kappa B and AP-1 in T84 cells. Infect. Immun. 70, 2304- 
2310. 



in 



de Grado, M., Rosenberger, CM., Gauthier, A., Vallance, B.A., and Finlay, B.B. 
(2001). Enteropathogenic Escherichia coli infection induces expression of the 
early growth response factor by activating mitogen-activated protein kinase 
cascades in epithelial cells. Infect. Immun. 69, 6217-6224. 
De Rycke, J., Comtet, E., Chalareng, C., Boury, M., Tasca, C., and Milon, A. 
(1997). Enteropathogenic Escherichia coli O103 from rabbit elicits actin stress 
fibers and focal adhesions in HeLa epithelial cells, cytopathic effects that are 
linked to an analog of the locus of enterocyte effacement. Infect. Immun. 65, 
2555-2563. 
DeVinney, R., Puente, J.L., Gauthier, A., Goosney, D., and Finlay, B.B. (2001). 
Enterohaemorrhagic and enteropathogenic Escherichia coli use a different 
Tir-based mechanism for pedestal formation. Mol. Microbiol. 41, 1445- 
1458. 
£ DeVinney, R., Stein, ML, Reinscheid, D., Abe, A., Ruschkowski, S., and Finlay, 

g B.B. (1999). Enterohemorrhagic Escherichia coli 0157:H7 produces Tir, which 

g is translocated to the host cell membrane but is not tyrosine phosphorylated. 

* Infect. Immun. 67, 2389-2398. 

g Donnenberg, M.S., Giron, J.A., Nataro, J. P., and Kaper, J.B. (1992). A plasmid- 

w 

rt encoded type IV fimbrial gene of enteropathogenic Escherichia coli associated 

5j with localized adherence. Mol. Microbiol. 6, 3427-3437. 




o Donnenberg, M.S., Tacket, CO., Losonsky, G., Frankel, G., Nataro, J. P., Dougan, 

n G., and Levine, M.M. (1998). Effect of prior experimental human entero- 

£ pathogenic Escherichia coli infection on illness following homologous and 

Sj heterologous rechallenge. Infect. Immun. 66, 52-58. 

2 Donnenberg, M.S., Tzipori, S., McKee, M.L., O'Brien, A.D., Alroy, J., and Kaper, 

w J.B. (1993). The role of the eae gene of enterohemorrhagic Escherichia coli 

in intimate attachment in vitro and in a porcine model. J. Clin. Invest. 92, 
1418-1424. 
Dytoc, M., Fedorko, L., and Sherman, P.M. (1994). Signal transduction in human 
epithelial cells infected with attaching and effacing Escherichia coli in vitro. 
Gastroenterology 106, 1150-1161. 
Ebel, F., von Eichel-Streiber, C, Rohde, M., and Chakraborty, T. (1998). Small 
GTP-binding proteins of the Rho- and Ras-subfamilies are not involved in 
the actin rearrangements induced by attaching and effacing Escherichia coli. 
FEMS Microbiol. Lett. 163, 107-112. 
Elliott, S.J., O'Connell, C.B., Koutsouris, A., Brinkley, C, Donnenberg, M.S., 
Hecht, G., and Kaper, J.B. (2002). A gene from the locus of enterocyte ef- 
facement that is required for enteropathogenic Escherichia coli to increase 
tight-junction permeability encodes a chaperone for EspF. Infect. Immun. 
70, 2271-2277. 



Elliott, S.J., Sperandio, V., Giron, J.A., Shin, S., Mellies, J.L., Wainwright, L., 
Hutcheson, S.W., McDaniel, T.K., and Kaper, J.B. (2000). The locus of ente- 
rocyte effacement (LEE)-encoded regulator controls expression of both LEE- 
and non-LEE-encoded virulence factors in enteropathogenic and enterohe- 
morrhagic Escherichia coli. Infect. Immun. 68, 6115-6126. 

Elliott, S.J., Wainwright, L.A., McDaniel, T.K., Jarvis, K.G., Deng, Y.K., Lai, L.C., 
McNamara, B.P., Donnenberg, M.S., and Kaper, J.B. (1998). The complete 
sequence of the locus of enterocyte effacement (LEE) from enteropathogenic 
Escherichia coli E2348/69. Mol. Microbiol. 28, 1-4. 

Elliott, S.J., Yu, J., and Kaper, J.B. (1999). The cloned locus of enterocyte efface- 
ment from enterohemorrhagic Escherichia coli 0157:H7 is unable to confer 
the attaching and effacing phenotype upon E. coli K-12. Infect. Immun. 67, 
4260-4263. 

Fitzhenry, R.J., Pickard, D.J., Hartland, E.L., Reece, S., Dougan, G., Phillips, A.D., h 

> 
and Frankel, G. (2002). Intimin type influences the site of human intestinal g 

x 
mucosal colonisation by enterohaemorrhagic Escherichia coli 0157:H7. Gut ^ 

50, 180-185. | 

> 
Foubister, V., Rosenshine, I., and Finlay, B.B. (1994). A diarrheal pathogen, en- g 

teropathogenic Escherichia coli (EPEC), triggers a flux of inositol phosphates g 

in infected epithelial cells. J. Exp. Med. 179, 993-998. " 

Frankel, G., Candy, D.C., Fabiani, E., Adu-Bobie, J., Gil, S., Novakova, M., Phillips, 3 

A.D., and Dougan, G. (1995). Molecular characterization of a carboxy- n 




terminal eukaryotic-cell -binding domain of intimin from enteropathogenic § 

Escherichia coli. Infect. Immun. 63, 4323-4328. o 

Frankel, G., Philips, A.D., Novakova, M., Batchelor, M., Hicks, S., and Dougan, « 

G. (1998). Generation of Escherichia coli intimin derivatives with differing bi- n 

ological activities using site-directed mutagenesis of the intimin C-terminus § 

domain. Mol. Microbiol. 29, 559-570. x 

Frankel, G., Phillips, A.D., Novakova, M., Field, H., Candy, D.C., Schauer, ^ 

D.B., Douce, G., and Dougan, G. (1996). Intimin from enteropathogenic h 

x 

Escherichia coli restores murine virulence to a Citrobacter rodentium eaeA mu- X 

o 

tant: induction of an immunoglobulin A response to intimin and EspB. Infect. h 

Immun. 64, 5315-5325. " 

Frankel, G., Phillips, A.D., Rosenshine, I., Dougan, G., Kaper, J.B., and Knutton, 
S. (1998). Enteropathogenic and enterohaemorrhagic Escherichia coli: more 
subversive elements. Mol. Microbiol. 30, 911-921. 

Freeman, N.L., Zurawski, D.V., Chowrashi, P., Ayoob, J.C., Huang, L., Mittal, B., 
Sanger, J.M., and Sanger, J.W. (2000). Interaction of the enteropathogenic 
Escherichia coli protein, translocated intimin receptor (Tir), with focal adhe- 
sion proteins. Cell Motil. Cytoskeleton 47, 307-318. 



in 




Friedberg, D., Umanski, T., Fang, Y., and Rosenshine, I. (1999). Hierarchy in the 
expression of the locus of enterocyte effacement genes of enteropathogenic 
Escherichia coli. Mol. Microbiol. 34, 941-952. 
Garrington, T.P. and Johnson, G.L. (1999). Organization and regulation of 
mitogen-activated protein kinase signaling pathways. Curr. Opin. Cell. Biol. 
11, 211-218. 
Gauthier, A., de Grado, M., and Finlay, B.B. (2000). Mechanical fractionation 
reveals structural requirements for enteropathogenic Escherichia coli Tir in- 
sertion into host membranes. Infect. Immun. 68, 4344-4348. 
Gerke, V. and Moss, S.E. (1997). Annexins and membrane dynamics. Biochim. 

Biophys. Acta 1357, 129-154. 
Giron, J.A., Ho, A.S., and Schoolnik, G.K. (1991). An inducible bundle -forming 
pilus of enteropathogenic Escherichia coli. Science 254, 710-713. 
£ Giron, J.A., Torres, A.G., Freer, E., and Kaper, J.B. (2002). The flagella of en- 

g teropathogenic Escherichia coli mediate adherence to epithelial cells. Mol. 

g Microbiol. 44, 361-379. 

3 Goosney, D.L., Celli, J., Kenny, B., and Finlay, B.B. (1999). Enteropathogenic 

g Escherichia coli inhibits phagocytosis. Infect. Immun. 67, 490-495. 

rt Goosney, D.L., DeVinney, R., and Finlay, B.B. (2001). Recruitment of cytoskele- 

5j tal and signaling proteins to enteropathogenic and enterohemorrhagic Es- 

o cherichia coli pedestals. Infect. Immun. 69, 3315-3322. 

n Goosney, D.L., DeVinney, R., Pfuetzner, R.A., Frey, E.A., Strynadka, N.C., and 

fc Finlay, B.B. (2000). Enteropathogenic E. coli translocated intimin receptor, 

Sj Tir, interacts directly with alpha-actinin. Curr. Biol. 10, 735-738. 

2 Gruenheid, S., DeVinney, R., Bladt, F., Goosney, D., Gelkop, S., Gish, G.D., 

w Pawson, T., and Finlay, B.B. (2001). Enteropathogenic E. coli Tir binds Nek 

to initiate actin pedestal formation in host cells. Nat. Cell Biol. 3, 856-859. 
Hecht, G., Marrero, J.A., Danilkovich, A., Matkowskyj, K.A., Savkovic, S.D., 
Koutsouris, A., and Benya, R.V. (1999). Pathogenic Escherichia coli increase 
Cl-secretion from intestinal epithelia by upregulating galanin-1 receptor ex- 
pression. J. Clin. Invest. 104, 253-262. 
Heczko, U., Carthy, CM., O'Brien, B.A., and Finlay, B.B. (2001). Decreased apop- 
tosis in the ileum and ileal Peyer's patches: a feature after infection with rabbit 
enteropathogenic Escherichia coli O103. Infect. Immun. 69, 4580-4589. 
Hobbie, S., Chen, L.M., Davis, R.J., and Galan, J.E. (1997). Involvement of 
mitogen-activated protein kinase pathways in the nuclear responses and cy- 
tokine production induced by Salmonella typhimurium in cultured intestinal 
epithelial cells./. Immunol. 159, 5550-5559. 
Huang, C., Ni, Y., Wang, T., Gao, Y., Haudenschild, C.C., and Zhan, X. (1997). 
Down-regulation of the filamentous actin cross-linking activity of cortactin 
by Src-mediated tyrosine phosphorylation. J. Biol. Chem. 272, 13,911-13,915. 



Huang, L., Mittal, B., Sanger, J.W., and Sanger, J.M. (2002). Host focal adhe- 
sion protein domains that bind to the translocated intimin receptor (Tir) of 
enteropathogenic Escherichia coli (EPEC). Cell. Motil. Cytoskeleton 52, 255- 
265. 

Hueck, C.J. (1998). Type III protein secretion systems in bacterial pathogens of 
animals and plants. Microbiol. Mol. Biol. Rev. 62, 379-433. 

Ide, T., Laarmann, S., Greune, L., S chillers, H., Oberleithner, H., and Schmidt, 
M.A. (2001). Characterization of translocation pores inserted into plasma 
membranes by type Ill-secreted Esp proteins of enteropathogenic Escherichia 
coli. Cell. Microbiol. 3, 669-679. 

Inman, L.R. and Cantey, J.R. (1983). Specific adherence of Escherichia coli (strain 
RDEC-1) to membranous (M) cells of the Peyer's patch in Escherichia coli 
diarrhea in the rabbit. J. Clin. Invest. 65, 1-8. 

Ismaili, A., Philpott, D.J., Dytoc, M.T., and Sherman, P.M. (1995). Signal trans- h 

duction responses following adhesion of verocytotoxin-producing Escherichia 

coli. Infect. Immun. 63, 3316-3326. ^ 

Jenkins, C, Chart, H., Smith, H.R., Hartland, E.L., Batchelor, M., Delahay, R.M., g 

> 
Dougan, G., and Frankel, G. (2000). Antibody response of patients infected g 

with verocytotoxin-producing Escherichia coli to protein antigens encoded on g 

the LEE locus. J. Med. Microbiol. 49, 97-101. " 

Jiang, Q., Mak, D., Devidas, S., Schwiebert, E.M., Bragin, A., Zhang, Y., Skach, 3 

W.R., Guggino, W.B., Foskett, J.K., and Engelhardt, J.F. (1998). Cystic fibrosis n 

transmembrane conductance regulator-associated ATP release is controlled § 

in 
O 



1/5 



> 
M 
H 

X 



by a chloride sensor. J. Cell Biol. 143, 645-657. o 

Johnson-Henry, K., Wallace, J.L., Basappa, N.S., Soni, R., Wu, G.K., and Sherman, « 

P.M. (2001). Inhibition of attaching and effacing lesion formation following n 

enteropathogenic Escherichia coli and Shiga toxin-producing E. coli infection. § 

Infect. Immun. 69, 7152-7158. x 

Jones, N.L., Islur, A., Haq, R., Mascarenhas, M., Karmali, M., Purdue, M.H., p 

Z.B.W., and Sherman, P. (2000). Escherichia coli Shiga toxins induce apop- h 

x 

tosis in epithelial cells that is regulated by the Bcl-2 family. Am. J. Physiol. X 

Gastrointest. Liver Physiol. 278, G811-G819. h 

Kalman, D., Weiner, O.D., Goosney, D.L., Sedat, J.W., Finlay, B.B., Abo, A., and g 

Bishop, J.M. (1999). Enteropathogenic E. coli acts through WASP and Arp2/3 
complex to form actin pedestals. Nat. Cell Biol. 1, 389-391. 

Kanamaru, K., Tatsuno, I., Tobe, T., and Sasakawa, C. (2000). Regulation of viru- 
lence factors of enterohemorrhagic Escherichia coli 0157:H7 by self-produced 
extracellular factors. Biosci. Biotechnol. Biochem. 64, 2508-2511. 

Kenny, B. (2001). The enterohaemorrhagic Escherichia coli (serotype 0157:H7) 
Tir molecule is not functionally interchangeable for its enteropathogenic 
E. coli (serotype 0127:H6) homologue. Cell. Microbiol. 3, 499-510. 



in 




Kenny, B. (1999). Phosphorylation of tyrosine 474 of the enteropathogenic Es- 
cherichia coli (EPEC) Tir receptor molecule is essential for actin nucleating 
activity and is preceded by additional host modifications. Mol. Microbiol. 31, 
1229-1241. 

Kenny, B., DeVinney, R., Stein, M., Reinscheid, D.J., Frey, E.A., and Finlay, B.B. 
(1997). Enteropathogenic E. coli (EPEC) transfers its receptor for intimate 
adherence into mammalian cells. Cell 91, 511-520. 

Kenny, B., Ellis, S., Leard, A.D., Warawa, J., Mellor, H., and Jepson, M.A. (2002). 
Co-ordinate regulation of distinct host cell signalling pathways by multifunc- 
tional enteropathogenic Escherichia coli effector molecules. Mol. Microbiol. 44, 
1095-1107. 

Kenny, B. and Jepson, M. (2000). Targeting of an enteropathogenic Escherichia 

coli (EPEC) effector protein to host mitochondria. Cell. Microbiol. 2, 579-590. 

£ Klapproth, J.M., Donnenberg, M.S., Abraham, J.M., and James, S.P. (1996). Prod- 

is 
g ucts of enteropathogenic E. coli inhibit lymphokine production by gastroin- 

g testinal lymphocytes. Am. J. Physiol. 271, G841-G848. 

3 Klapproth, J.M., Scaletsky, I.C., McNamara, B.P., Lai, L.C., Malstrom, C, James, 

g S.P., and Donnenberg, M.S. (2000). A large toxin from pathogenic Escherichia 

rt coli strains that inhibits lymphocyte activation. Infect. Immun. 68, 2148- 

% 2155. 

o Knutton, S., Baldwin, T., Williams, P.H., and McNeish, A.S. (1989). Actin accu- 

n mulation at sites of bacterial adhesion to tissue culture cells: basis of a new 

fc diagnostic test for enteropathogenic and enterohemorrhagic Escherichia coli. 

3 Infect. Immun. 57, 1290-1298. 

2 Knutton, S., Rosenshine, I., Pallen, M.J., Nisan, I., Neves, B.C., Bain, C, Wolff, 

w C, Dougan, G., and Frankel, G. (1998). A novel EspA-associated surface or- 

ganelle of enteropathogenic Escherichia coli involved in protein translocation 
into epithelial cells. EMBOJ. 17, 2166-2176. 
Kodama, T., Akeda, Y., Kono, G., Takahashi, A., Imura, K., Iida, T., and Honda, 
T. (2002). The EspB protein of enterohaemorrhagic Escherichia coli interacts 
directly with alpha-catenin. Cell. Microbiol. 4, 213-222. 
Kondro, W. (2000). E. coli outbreak deaths spark judicial inquiry in Canada. Lancet 

355, 2058. 
Kresse, A.U., Guzman, C.A.,andEbel, F. (2001). Modulation of host cell signalling 
by enteropathogenic and Shiga toxin-producing Escherichia coli. Int. J. Med. 
Microbiol. 291, 277-285. 
Kresse, A.U., Rohde, M., and Guzman, C.A. (1999). The EspD protein of en- 
terohemorrhagic Escherichia coli is required for the formation of bacterial 
surface appendages and is incorporated in the cytoplasmic membranes of 
target cells. Infect. Immun. 67, 4834-4842. 



Li, Y., Frey, E., Mackenzie, A.M., and Finlay, B.B. (2000). Human response to 
Escherichia coli 0157:H7 infection: antibodies to secreted virulence factors. 
Infect. Immun. 68, 5090-5095. 

Liu, J., Akahoshi, T., Sasahana, T., Kitasato, H., Namai, R., Sasaki, T., Inoue, 
M., and Kondo, H. (1999). Inhibition of neutrophil apoptosis by verotoxin 2 
derived from Escherichia coli 0157:H7. Infect. Immun. 67, 6203-6205. 

Lommel, S., Benesch, S., Rottner, K., Franz, T., Wehland, J., and Kuhn, R. (2001). 
Actin pedestal formation by enteropathogenic Escherichia coli and intracel- 
lular motility of Shigella flexneri are abolished in N-WASP -defective cells. 
EMBO Rep. 2, 850-857. 

Loureiro, I., Frankel, G., Adu-Bobie, J., Dougan, G., Trabulsi, L.R., and Carneiro- 

Sampaio, M.M. (1998). Human colostrum contains IgA antibodies reactive 

to enteropathogenic Escherichia coli virulence-associated proteins: intimin, 

BfpA, EspA, and EspB. J. Pediatr. Gastroenterol. Nutr. 27, 166-171. g 

> 
Luo, Y., Frey, E.A., Pfuetzner, R.A., Creagh, A.L., Knoechel, D.G., Haynes, ^ 

x 
C.A., Finlay, B.B., and Strynadka, N.C. (2000). Crystal structure of en- ^ 

teropathogenic Escherichia coli intimin-receptor complex. Nature 405, 1073- g 

1077. g 




M 



Malstrom, C. and James, S. (1998). Inhibition of murine splenic and mucosal g 

lymphocyte function by enteric bacterial products. Infect. Immun. 66, 3120- 
3127. 



z 

H 



Manjarrez-Hernandez, H.A., Amess, B., Sellers, L., Baldwin, T.J., Knutton, S., n 



H 



Williams, P.H., and Aitken, A. (1991). Purification of a 20 kDa phosphopro- § 

tein from epithelial cells and identification as a myosin light chain. Phos- o 

phorylation induced by enteropathogenic Escherichia coli and phorbol ester. « 

FEBS Lett. 292, 121-127. " 

> 

Manjarrez-Hernandez, H.A., Baldwin, T.J., Williams, P.H., Haigh, R., Knutton, § 

S., and Aitken, A. (1996). Phosphorylation of myosin light chain at distinct x 

sites and its association with the cytoskeleton during enteropathogenic Es- 2 

cherichia coli infection. Infect. Immun. 64, 2368-2370. h 

Manjarrez-Hernandez, H.A., Gavilanes-Parra, S., Chavez-Berrocal, E., Navarro- ffi 

Ocana, A., and Cravioto, A. (2000). Antigen detection in enteropathogenic h 

Escherichia coli using secretory immunoglobulin A antibodies isolated from 
human breast milk. Infect. Immun. 68, 5030-5036. 

Marches, O., Nougayrede, J. P., Boullier, S., Mainil, J., Charlier, G., Raymond, I., 
Pohl, P., Boury, M., DeRycke, J., Milon, A., and Oswald, E. (2000). Role of 
tir and intimin in the virulence of rabbit enteropathogenic Escherichia coli 
serotype O103:H2. Infect. Immun. 68, 2171-2182. 

Martinez, M.B., Taddei, C.R., Ruiz-Tagle, A., Trabulsi, L.R., and Giron, J.A. 
(1999). Antibody response of children with enteropathogenic Escherichia coli 



n 

M 
t- 
M 

in 




infection to the bundle -forming pilus and locus of enterocyte effacement- 
encoded virulence determinants. J. Infect. Dis. 179, 269-274. 
McCarthy, K.M., Skare, I.B., Stankewich, M.C., Furuse, M., Tsukita, S., Rogers, 
R.A., Lynch, R.D., and Schneeberger, E.E. (1996). Occludin is a functional 
component of the tight junction. J. Cell Sci. 109, 2287-2298. 
McDaniel, T.K. and Kaper, J.B. (1997). A cloned pathogenicity island from en- 
teropathogenic Escherichia coli confers the attaching and effacing phenotype 
on E. coli K-12. Mol. Microbiol 23, 399-407. 
McNamara, B.P., Koutsouris, A., O'Connell, C.B., Nougayrede, J. P., Donnen- 
berg, M.S., and Hecht, G. (2001). Translocated EspF protein from entero- 
pathogenic Escherichia coli disrupts host intestinal barrier function. J. Clin. 
Invest. 107, 621-629. 
Mellies, J.L., Elliott, S.J., Sperandio, V., Donnenberg, M.S., and Kaper, J.B. (1999). 
£ The Per regulon of enteropathogenic Escherichia coli: identification of a reg- 

g ulatory cascade and a novel transcriptional activator, the locus of enterocyte 

g effacement (LEE)-encoded regulator (Ler). Mol. Microbiol. 33, 296-306. 

3 Moon, H.W., Whipp, S.C., Argenzio, R.A., Levine, M.M., and Giannella, R.A. 

g (1983). Attaching and effacing activities of rabbit and human enteropatho- 

rt genie Escherichia coli in pig and rabbit intestines. Infect. Immun. 41, 1340- 

% 1351. 

o Muller, A. and Rudel, T. (2001). Modification of host cell apoptosis by viral and 

w bacterial pathogens. Int.]. Med. Microbiol. 291, 197-207. 

fc Nataro, J. P. and Kaper, J.B. (1998). Diarrheagenic Escherichia coli. Clin. Microbiol. 

3 Rev. 11, 142-201. 

2 Nicholls, L., Grant, T.H., and Robins-Browne, R.M. (2000). Identification of a 

w novel genetic locus that is required for in vitro adhesion of a clinical isolate 

of enterohaemorrhagic Escherichia coli to epithelial cells. Mol. Microbiol. 35, 
275-288. 
Nisan, I., Wolff, C., Hanski, E., and Rosenshine, I. (1998). Interaction of en- 
teropathogenic Escherichia coli with host epithelial cells. Folia Microbiol. 43, 
247-252. 
Norbury, C. and Nurse, P. (1992). Animal cell cycles and their control. Annu. Rev. 

Biochem. 61, 441-470. 
Nougayrede, J. P., Boury, M., Tasca, C., Marches, O., Milon, A., Oswald, E., and 
De Rycke, J. (2001). Type III secretion-dependent cell cycle block caused 
in HeLa cells by enteropathogenic Escherichia coli O103. Infect. Immun. 69, 
6785-6795. 
Nougayrede, J. P., Marches, O., Boury, M., Mainil, J., Charlier, G., Pohl, P., De 
Rycke, J., Milon, A., and Oswald, E. (1999). The long-term cytoskeletal rear- 
rangement induced by rabbit enteropathogenic Escherichia coli is Esp depen- 
dent but intimin independent. Mol. Microbiol. 31, 19-30. 



O'Longhlin, E.V. and Robins-Browne, R.M. (2001). Effect of Shiga toxin and 
Shiga-like toxins on eucaryotic cells. Microbes Infect. 3, 493-507- 

Paton, A.W., Manning, P.A., Woodrow, M.C., and Paton, J.C. (1998). Translo- 
cated intimin receptors (Tir) of Shiga-toxigenic Escherichia coli isolates be- 
longing to serogroups 026, Olll, and 0157 react with sera from patients 
with hemolytic-uremic syndrome and exhibit marked sequence heterogene- 
ity. Infect. Immun. 66, 5580-5586. 

Perna, N.T., Plunkett, G. Ill, Burland, V., Mau, B., Glasner, J.D., Rose, D.J., 
Mayhew, G.F., Evans, P.S., Gregor, J., Kirkpatrick, H.A., Posfai, G., Hackett, 
J., Klink, S., Boutin, A., Shao, Y., Miller, L., Grotbeck, E.J., Davis, N.W., Lim, 
A., Dimalanta, E.T., Potamousis, K.D., Apodaca, J., Anantharaman, T.S., Lin, 
J., Yen, G., Schwartz, D.C., Welch, R.A., and Blattner, F.R. (2001). Genome 
sequence of enterohaemorrhagic Escherichia coli 0157:H7. Nature 409, 529- 

Philpott, D.J., McKay, D.M., Mak, W., Perdue, M.H., and Sherman, P.M. (1998). q 

x 
Signal transduction pathways involved in enterohemorrhagic Escherichia coli- ^ 

induced alterations in T84 epithelial permeability. Infect. Immun. 66, 1680- g 

1687. g 

Philpott, D.J., McKay, D.M., Sherman, P.M., and Perdue, M.H. (1996). Infection g 

of T84 cells with enteropathogenic Escherichia coli alters barrier and transport 
functions. Am. J. Physiol. 270, G634-G645. 8 

Plunkett, G. Ill, Rose, D.T., Durfee, T.J., and Blattner, F.R. (1999). Sequence of n 

Shiga toxin 2 phage 933W from Escherichia coli 0157:H7: Shiga toxin as a § 

phage late-gene product. J. Bacteriol. 181, 1767-1778. o 

Rosenshine, I., Donnenberg, M.S., Kaper, J.B., and Finlay, B.B. (1992). Signal « 

transduction between enteropathogenic Escherichia coli (EPEC) and epithelial n 

cells: EPEC induces tyrosine phosphorylation of host cell proteins to initiate § 

cytoskeletal rearrangement and bacterial uptake. EMBO J. 11, 3551-3560. x 

Rosenshine, I., Ruschkowski, S., Stein, M., Reinscheid, D.J., Mills, S.D., and 2 

Finlay, B.B. (1996). A pathogenic bacterium triggers epithelial signals to form h 

a functional bacterial receptor that mediates actin pseudopod formation. ffi 

EMBO J. 15, 2613-2624. h 

Sanger, J.M., Chang, R., Ashton, F., Kaper, J.B., and Sanger, J.W. (1996). Novel g 

form of actin-based motility transports bacteria on the surfaces of infected 
cells. Cell. Motil. Cytoskeleton 34, 279-287. 

Savkovic, S.D., Koutsouris, A., and Hecht, G. (1997). Activation of NF-kappa B in 
intestinal epithelial cells by enteropathogenic Escherichia coli. Am. J. Physiol. 
273, C1160-C1167. 

Savkovic, S.D., Ramaswamy, A., Koutsouris, A., and Hecht, G. (2001). EPEC- 
activated ERK1/2 participate in inflammatory response but not tight junction 
barrier disruption. Am. J. Physiol. Gastrointest. Liver Physiol. 281, G890-G898. 



in 



Sekiya, K., Ohishi, M., Ogino, T., Tamano, K., Sasakawa, C, and Abe, A. (2001). 
Supermolecular structure of the enteropathogenic Escherichia coli type III se- 
cretion system and its direct interaction with the EspA-sheath-like structure. 
Proc. Natl. Acad. Sci. USA 98, 11,638-11,643. 
Shaw, R.K., Daniell, S., Frankel, G., and Knutton, S. (2002). Enteropathogenic 
Escherichia coli translocate Tir and form an intimin-Tir intimate attachment 
to red blood cell membranes. Microbiology 148, 1355-1365. 
Shifrin, Y., Kirschner, J., Geiger, B., and Rosenshine, I. (2002). Enteropathogenic 
Escherichia coli induces modification of the focal adhesions of infected host 
cells. Cell. Microbiol. 4, 235-243. 
Simonovic, I., Rosenberg, J., Koutsouris, A., and Hecht, G. (2000). Entero- 
pathogenic Escherichia coli dephosphorylates and dissociates occludin from 
intestinal epithelial tight junctions. Cell. Microbiol. 2, 305-315. 
£ Sinclair, J.F. and O'Brien, A.D. (2002). Cell surface -localized nucleolin is a eu- 

g karyotic receptor for the adhesin intimin-gamma of enterohemorrhagic Es- 

g cherichia coli 0157:H7. J. Biol. Chem. 277, 2876-2885. 

3 Sperandio, V., Mellies, J.L., Nguyen, W., Shin, S., and Kaper, J.B. (1999). Quorum 

g sensing controls expression of the type III secretion gene transcription and 

w 

rt protein secretion in enterohemorrhagic and enteropathogenic Escherichia 

* coli. Proc. Natl. Acad. Sci. USA 96, 15,196-15,201. 




o Spitz, J., Yuhan, R., Koutsouris, A., Blatt, C, Alverdy, J., and Hecht, G. (1995). En- 

n teropathogenic Escherichia coli adherence to intestinal epithelial monolayers 

fc diminishes barrier function. Am. J. Physiol. 268, G374-G379. 

3 Stein, M.A., Mathers, D.A., Yan, H., Baimbridge, K.G., and Finlay, B.B. (1996). 

2 Enteropathogenic Escherichia coli markedly decreases the resting membrane 

w potential of Caco-2 and HeLa human epithelial cells. Infect. Immun. 64, 4820- 

4825. 
Stevens, M.P., van Diemen, P.M., Frankel, G., Phillips, A.D., and Wallis, T.S. 
(2002). Efal influences colonization of the bovine intestine by Shiga toxin- 
producing Escherichia coli serotypes 05 and Olll. Infect. Immun. 70, 5158- 
5166. 
Stuart, R.O. and Nigam, S.K. (1995). Regulated assembly of tight junctions by 

protein kinase C. Proc. Natl. Acad. Sci. USA 92, 6072-6076. 
Tarr, P. I., Bilge, S.S., Vary, J.C. Jr., Jelacic, S., Habeeb, R.L., Ward, T.R., 
Baylor, M.R., and Besser, T.E. (2000). Iha: a novel Escherichia coli 0157:H7 
adherence -conferring molecule encoded on a recently acquired chromoso- 
mal island of conserved structure. Infect. Immun. 68, 1400-1407. 
Tobe, T., Hayashi, T., Han, C.G., Schoolnik, G.K., Ohtsubo, E., and Sasakawa, 
C. (1999). Complete DNA sequence and structural analysis of the entero- 
pathogenic Escherichia coli adherence factor plasmid. Infect. Immun. 67, 
5455-5462. 



Wachter, C, Beinke, C, Mattes, M., and Schmidt, M.A. (1999). Insertion of 
EspD into epithelial target cell membranes by infecting enteropathogenic 
Escherichia coli. Mol. Microbiol. 31, 1695-1707. 

Wagner, P.L., Acheson, D.W., and Waldor, M.K. (2001). Human neutrophils 
and their products induce Shiga toxin production by enterohemorrhagic Es- 
cherichia coli. Infect. Immun. 69, 1934-1937. 

Warawa, J., Finlay, B.B., and Kenny, B. (1999). Type III secretion-dependent 
hemolytic activity of enteropathogenic Escherichia coli. Infect. Immun. 67, 
5538-5540. 

Zobiack, N., Rescher, U., Laarmann, S., Michgehl, S., Schmidt, M.A., and Gerke, 
V. (2002). Cell-surface attachment of pedestal-forming enteropathogenic 
E. coli induces a clustering of raft components and a recruitment of annexin 
2. J. Cell Sci. 115, 91-98. 




on 
H 
w 

> 
M 
H 

X 

i 

> 

M 
H 

w 

I— I 

Z 
H 

M 

n 

H 

p— i 

O 

Z 

m 

O 
►n 

w 
w 

n 

> 

O 
M 

W 

n 



H 

K 

o 

en 
H 

n 

M 
M 
M 

CO 



CHAPTER 5 



Molecular ecology and cell biology of 
Legionella pneumophila 

Maelle Molmeret, Dina M. Bitar, and Yousef Abu Kwaik 



Legionella pneumophila, a Gram-negative bacillus that is ubiquitous in aquatic 
environments, is responsible for Legionnaires' disease. It is a facultative in- 
tracellular pathogen that can replicate within eukaryotic host cells such as pro- 
tozoan and macrophages. In water, I. pneumophila grows within protozoan 
hosts. There are at least 13 species of amoebae and 2 species of ciliated pro- 
tozoa that support intracellular replication of I. pneumophila (Fields, 1996). 
Among the most predominant amoebae in water sources are hartmannellae 
and acanthamoebae, which have also been isolated from water sources associ- 
ated with Legionnaires' disease outbreaks (Fields, 1996). Interaction between 
L. pneumophila and protozoa is considered to be central to the pathogenesis 
and ecology of L. pneumophila (Rowbotham, 1986; Harb et al., 2000). In hu- 
mans, L. pneumophila reaches the lungs after inhalation of contaminated 
aerosol droplets (Fields, 1996; Fliermans, 1996; also see Fig. 5.1). The main 
sources of contaminated water droplets are hot water and air-conditioning 
systems, but the bacteria have been isolated from fountains, spas, pools, den- 
tal and hospital units, and other man-made water systems (Fliermans, 1996; 
also see Fig. 5.1). No person-to-person transmission has been described. Once 
in the lungs, L. pneumophila are ingested in alveolar macrophages, the major 
site of bacterial replication. This results in an acute and severe pneumonia. In 
addition to Legionnaires' disease, L. pneumophila also causes Pontiac fever, 
which is a self-limiting flu-like illness that is not well understood but is not 
lethal. Approximately one half of the 48 species of Legionella have been asso- 
ciated with human disease. L. pneumophila is responsible for 90% of cases 
of Legionnaires' disease. However, all the Legionella species under appro- 
priate conditions may be capable of intracellular growth and infliction of 
human disease. Infections that are due to less common species of legionellae 





75 

c 
g 

CD 
<D 



■D 


■= 
♦- Jv CO 

u 5i] 
co J2 <d 

E"cd 

— c 



o 

O CO 

O) 
CD 



T3 
0) 

£ <D © 

JO o 

o © "w 
x c S> 
HI g > 

CD 



W CO O) 

•P > .£ w 
E c = © 

CO CO CO o 

s I El 

»- 3 O CO 
C O Q 




are not frequently diagnosed and reported, and they are less studied than 
L. pneumophila (Fields et al., 2002). 

The unique intracellular fate of L. pneumophila is one of the interesting 
aspects of this organism. Unlike phagosomes containing inert particles or 
avirulent bacteria, the I. pneumophila-containing vacuoles avoid the "normal" 
endocytic pathway, recruiting rough endoplasmic reticulum (RER) and mito- 
chondria, to reside in a specialized vacuole allowing intracellular replication 
(Horwitz and Silverstein, 1980; Horwitz, 1983a, 1983b, 1984; Horwitz and 
Maxfield, 1984; also see Fig. 5.2). The formation of this specialized vacuole is 
directed by the type IV secretion system encoded by the dot/icm genes. The dot 
(defect in organelle trafficking) /icm (intracellular multiplication) loci consist 
of 23 genes located in two chromosomal loci of I. pneumophila. These genes 
have been identified independently by two different laboratories (Segal et al., 
1998; Segal and Shuman, 1998; Vogeletal., 1998). An analysis of the predicted § 

amino acid sequences of the dot/icm genes has revealed several characteris- g 

tics that indicate a role in conjugal transfer of DNA, which has been con- £ 

firmed by conjugation studies (Segal and Shuman, 1998; Vogel et al., 1998). w 

This apparatus has also been shown to be involved in proper maturation ° 

of the I. pneumophila-containing phagosome in mammalian and protozoan 8 

cells, directing the biogenesis of the specialized vacuole in which Legionella * 




replicate (Swanson and Isberg, 1995b; Segal and Shuman, 1999; Molmeret g 

et al., 2002a). The dot/icm genes are also required for macropinocytosis in £ 

A/ J mice macrophages (Watarai et al., 2001b), upregulation of phagocytosis g 



o 



in human-derived macrophages (Hilbi et al., 2001), induction of apoptosis g 



o 



in macrophages (Gao and Abu Kwaik, 1999a, 1999b; Zink et al., 2002), and § 

H 

tn 

O 
H 

o 

; ; * 

Figure 5.1. (facing page). The environmental life of I. pneumophila within protozoa. S 

(1) L. pneumophila from biofilms with other bacteria, or in suspension, infecting protozoa. ^ 

(2) Following entry, L. pneumophila resides in a membrane-bound vacuole that recruits m 

ci 

host cell organelles, such as the mitochondria and the rough endoplasmic reticulum, and ^ 

o or q 

does not fuse with lysosomes. Mutants such as dot/icm mutants fuse to lysosomes (3b); S 

(-I 
L. pneumophila replicates within specialized vacuoles and reaches large numbers (3a). £ 

(4) The infectious particle is not known but may include excreted Legionella-fiWed vesicles, 
intact Legionella-fiWed amoebae, or free legionellae that have lysed their host cell. 

(5) Transmission to humans occurs by mechanical means, such as faucets and 
showerheads. Infection in humans occurs by inhalation of the infectious particle and 
establishment of infection in the lungs. (6) Legionellae that have escaped their host cell 
may survive in suspension for long periods of time, reinfect other protozoa, or recolonize 
biofilms. (This figure was taken from ASM News 66 (10): 609-616, 2000.) 




I— I 

< 

< 

w 
on 

O 

Q 

< 

i— i 
pq 



< 



H 
w 
Pi 
w 

1-1 
o 

w 
1-1 
1-1 

< 







Figure 5.2. Transmission electron micrographs of the infection of Acanthamoeb a 
polyphaga (top panel) and U937 macrophages (bottom panel) by L. pneumophila. Coiling 
phagocytosis (A and B); formation of the RE R- surrounded phagosome (C and D); and late 
stages of the infection (E and F). Note that in E and F there is no intact phagosomal 
membrane. (This figure was adapted from ASM News 66 (10): 609-616, 2000.) 

pore -formation-mediated cytotoxicity in both protozoan and mammalian cells 
(Alii et al, 2000; Gao and Abu Kwaik, 2000; Molmeret et al, 2002a, 2002b). 



ECOLOGY OF LEGIONELLAE 

After isolation of I. pneumophila from the air-conditioning system during 
the first outbreak in Philadelphia, Pennsylvania (1976) , the bacteria have been 
isolated from numerous sources in the environment. Legionellae species have 
been repeatedly shown to be ubiquitous, particularly in aquatic environments 
(Fields, 1996). 



In the environment, legionellae species cannot multiply extracellularly, 
and they have been shown to be parasites of protozoa (Harb et al., 2000). In 
1980, Rowbotham was the first to describe the ability of I. pneumophila to 
multiply intracellularly within protozoa (Rowbotham, 1980). The 13 species 
of amoebae and 2 species of ciliated protozoa that allow intracellular bacterial 
replication have been shown to be potential environmental hosts for legionel- 
lae species (Abu Kwaik et al., 1998). This rather sophisticated host-parasite 
interaction indicates a tremendous adaptation of legionellae to parasitize pro- 
tozoa. This host-parasite interaction is also central to the pathogenesis and 
ecology of these bacteria. 

There are now at least 46 species of legionellae [Benin, et al., 2002]. 
In addition, 12 phylogenetic groups of bacteria belonging to 5 species have 
been designated as Legionella-like amoebal pathogens (LLAP; Adeleke et al., 
1996). The LLAPs are genetically related to legionellae and many of them § 

have been associated with Legionnaires' disease (Birtles et al., 1996; Marrie g 

et al., 2001). In contrast to legionellae species, the LLAPs cannot be cultured £ 

in vitro on artificial media. The LLAPs are isolated by coculture with protozoa w 

(Fields, 1996). LLAPs have been isolated from sputum samples derived from ° 

patients with Legionnaires' disease on the basis of the ability of these bacteria 8 

to multiply in protozoa, because they cannot be grown on artificial media % 




o 
(Adeleke et al. , 1996) . The recent developments in using the polymerase chain g 

reaction for bacterial identification in environmental samples will facilitate £ 

better identification of legionellae species and LLAPs. g 

o 
Many strategies have been used to eradicate legionellae from sources of g 

infection in water and plumbing systems that have been associated with dis- 2 

ease outbreaks. These strategies include chemical biocides such as chlorine, g 

overheating of the water, and UV irradiation (Biurrun et al., 1999; Kool et § 

al., 1999; Muraca et al., 1987). Such interventions have been successful for c 

short periods of time after which the bacteria are again found in these sources g 

(Yamamoto et al., 1991; Biurrun et al., 1999). Thus, eradication of L. pneu- c 

mophila from the environmental sources of infection requires continuous o 

treatment of the water with agents such as monochloramine or copper-silver 2 

ions in addition to maintenance of the water temperature above ^55°C [Kool 
et al., 1999; Kusnetsovet al., 2001]. It is clear that the sophisticated association 
of legionellae with protozoa is a major factor in continuous presence of the 
bacteria in the environment. Compared with L. pneumophila grown in vitro, 
amoebae-grown bacteria have been shown to be highly resistant to chemical 
disinfectants and to treatment with biocides (Barker et al., 1992). Amoebae- 
grown L. pneumophila have been shown to manifest a dramatic increase in 
their resistance to harsh environmental conditions, such as fluctuation in 




temperature, osmolarity, pH, and exposure to oxidizing agents (Abu Kwaik 
et al., 1997). Protozoa have been shown to release vesicles containing L. pneu- 
mophila that are highly resistant to biocides (Berk et al., 1998). 

The ability of I. pneumophila to survive within amoebic cysts further con- 
tributes to the resistance of I. pneumophila to physical and biochemical agents 
used in bacterial eradication (Barker et al., 1992, 1995). It is most likely that 
eradication of the bacteria from the environment should start by preventing 
protozoan infection, an integral part of the infectious cycle of I. pneumophila. 
Further characterization of the mechanisms of bacterial invasion into pro- 
tozoa may allow the design of strategies to block the protozoan receptor 
from attachment to legionellae and thus prevent bacterial entry. Extracellular 
L. pneumophila will be more susceptible to environmental conditions and will 
not be protected from biocides and disinfectants. Furthermore, blockage of 
w bacterial entry into amoebae will render bacteria less infective and virulent 

g to mammalian cells. Alternatively, treatment of water sources contaminated 

§ with I. pneumophila with "safe" agents that block certain essential bacterial 

< 

ph metabolic pathways, such as the peptidoglycan biosynthesis pathway, may 

z prove to be useful (Harb et al., 1998). 

Q It has been proposed that the infectious particle for Legionnaires' disease 

< is amoebae infected with the bacteria (Rowbotham, 1980; also see Fig. 5.1). 

< Although this has not yet been proven, there are many lines of evidence to 
* suggest that protozoa play major roles in the transmission of I. pneumophila. 

< First, many protozoan hosts have been identified that allow intracellular bac- 
S terial replication, which is the only means of bacterial amplification in the 
^ environment (Fields, 1996; Abu Kwaik et al, 1998; Harb et al., 2000). Sec- 
§ ond, in outbreaks of Legionnaires' disease, amoebae and bacteria have been 
| isolated from the same source of infection and the isolated amoebae support 
3 intracellular replication of the bacteria (Fields et al., 1990) . Third, as discussed, 

< following intracellular replication within protozoa, L. pneumophila exhibit a 
dramatic increase in resistance to harsh conditions, such as high temperature, 
acidity, and high osmolarity, which may facilitate bacterial survival in the envi- 
ronment (Abu Kwaik et al., 1997). Fourth, intracellular L. pneumophila within 
protozoa are more resistant to chemical disinfection and biocides compared 
with bacteria grown in vitro bacteria (Barker et al., 1992, 1993, 1995). Fifth, 
protozoa have been shown to release vesicles of respirable size that contain 
numerous L. pneumophila. The vesicles are resistant to freeze-thawing and 
sonication, and the bacteria within the vesicles are highly resistant to biocides 
(Berk et al., 1998). Sixth, following their release from the protozoan host, the 
bacteria exhibit a dramatic increase in their infectivity for mammalian cells 
in vitro (Cirillo et al., 1994). 



In addition, it has been demonstrated that intracellular bacteria within 
Hartmanella vermiformis are dramatically more infectious and are highly 
lethal in mice (Brieland et al., 1997). Seventh, the number of bacteria iso- 
lated from the source of infection of Legionnaires' disease is usually very low 
or undetectable, and thus enhanced infectivity of intracellular bacteria within 
protozoa may compensate for the low infectious dose (O'Brein and Bhopal, 
1993). Eighth, viable but nonculturable L. pneumophila can be resuscitated 
by coculture with protozoa (Steinert et al., 1997). This observation may 
suggest that failure to isolate the bacteria from environmental sources of 
infection may be due to this "dormant" phase of the bacteria that cannot be 
recovered on artificial media. Ninth, there has been no documented case of 
bacterial transmission between individuals. The only source of transmission 
is environmental droplets generated from many human-made devices, 
such as shower heads, water fountains, whirlpools, and cooling towers of § 



ADHERENCE AND ENTRY MECHANISMS 



protozoan hosts 




M 



air-conditioning systems (Fields, 1996). £ 



M 

n 
o 

M 

O 

o 

Initial interactions between L pneumophila and its primitive > 

o 
n 

w 

M 

CO 



Bacterial attachment to H. vermiformis is mediated by adherence to 

a protozoan receptor that has been characterized as a galactose/N -acetyl- 2 

o 
galactosamine (Gal/GalNAc) lectin with similarity to the fii integrin-like g 

Gal/GalNAc lectin of the pathogenic protozoan Entamoebae histolytica (Mann 2 

et al., 1991; Venkataraman et al., 1997; Harb et al., 1998). Integrins are het- g 

erodimeric protein tyrosine kinase receptors that undergo tyrosine phospho- § 

rylation upon engagement to ligands, which subsequently results in recruit- c 

ment and rearrangements of the cytoskeleton. Interestingly, attachment of 5 

L. pneumophila to the Gal/GalNAc of H. vermiformis triggers signal transduc- c 

tion events in H. vermiformis that are manifested in dramatic tyrosine dephos- o 

phorylation of the lectin receptor and other proteins (Venkataraman et al., 2 

1997). Moreover, in addition to these manipulations of the signal transduc- 
tion of H. vermiformis by L. pneumophila, bacterial invasion is also associated 
with specific induction of gene expression in the protozoa, and inhibition 
of this gene expression blocks entry of the bacteria (Abu Kwaik et al., 1994). 
Following this initial host-parasite interaction, uptake of I. pneumophila by 
protozoan cells occurs by both conventional and coiling phagocytosis (in 
which the bacterium is surrounded by a multilayer coil -like structure; Abu 
Kwaik, 1996; Bozue and Johnson, 1996; also see Fig. 5.2). 




Invasion of H. vermiformis by I. pneumophila requires host protein syn- 
thesis, because eukaryotic protein synthesis inhibitors (cycloheximide) block 
the entry process (Abu Kwaik et al., 1994). The uptake of L. pneumophila into 
H. vermiformis mainly occurs through cup-shaped invaginations (or zipper 
phagocytosis) on the surface of the amoeba, although some coiling phagocy- 
tosis also occurs (Venkataraman et al., 1998). Such invaginations are known 
to be microfilament dependent (Venkataraman et al., 1998). However, the 
entry of the bacteria into H. vermiformis is not inhibited by microfilament 
inhibitors such as cytochalasin D (King et al., 1991; Harb et al., 1998). Methy- 
lamine, which is an inhibitor of receptor-mediated endocytosis, inhibits the 
entry of L. pneumophila into H. vermiformis (King et al., 1991). Apparently, 
infection of Acanthamoeha polyphaga by L. pneumophila occurs through a dif- 
ferent mechanism. It is not inhibited by galactose or N-acetylgalactosamine 
« (Harb et al., 1998). The 170-kDa Gal/GalNAc-inhibitable lectin is only mildly 

g dephosphorylated in A. polyphaga upon attachment of I. pneumophila (Harb 

§ et al., 1998). Furthermore, host protein synthesis by A. polyphaga is not re- 

< 

ph quired for invasion by I. pneumophila (Harb et al., 1998). The uptake of the 

s bacteria is not inhibited by cytoskeleton-disrupting agents. The role of this 

Q form of phagocytosis in the intracellular fate of I. pneumophila is not fully 

< understood, because human macrophages are able to phagocytose heat- or 

< formalin-killed I. pneumophila by coiling phagocytosis (Horwitz, 1984). 

i— i 

pq 

i 

2 ATTACHMENT AND ENTRY TO MAMMALIAN CELLS 

S 

w Invasion and intracellular replication of I. pneumophila within pul- 

§ monary cells in the alveoli is the hallmark of Legionnaires' disease (Abu 

° Kwaik, 1998b). These alveolar cells include macrophages, and type I and 

3 II epithelial cells. Attachment of I. pneumophila into macrophages has been 

< shown to be mediated, at least in part, through the attachment of complement- 
is 

coated bacteria to the complement receptor (Payne and Horwitz, 1987), al- 
though non-complement-mediated uptake also occurs (Elliott and Winn, 
1986; Rodgers and Gibson, 1993). The host cell receptor involved in non- 
complement-mediated uptake in macrophages and epithelial cells is not 
known. 

Uptake of I. pneumophila by monocytes and macrophages has been 
shown to occur through conventional and coiling phagocytosis (Horwitz, 
1984; Weinbaum et al., 1984; Elliott and Winn, 1986; Rechnitzer and Blom, 
1989; Dowlingetal., 1992; also see Fig. 5.2). Because heat-killed and formalin- 
killed I. pneumophila are also taken up by coiling phagocytosis (Horwitz, 1984) 
but are targeted to the lysosomes (Horwitz and Maxfield, 1984), this mode of 



uptake may not play a role in subsequent pathogenicity of the bacteria. Many 
clinical isolates of L. pneumophila have been shown to be taken up exclusively 
by conventional phagocytosis (Elliott and Winn, 1986; Rechnitzer and Blom, 
1989). In addition, other species of legionellae, such as L. micdadei, which 
is the second most common species of legionellae that causes Legionnaires' 
disease, is taken up exclusively by conventional phagocytosis (Rechnitzer and 
Blom, 1989). The bacterial ligand that mediates the coiling mode of phagocy- 
tosis is not known. Moreover, the phagocytic receptor that binds the bacteria 
seems to play some role in determining the fate of the intracellular bacteria 
because opsonization with antibodies reduces intracellular growth (Horwitz 
and Silverstein, 1981a; Nash et al., 1984; Payne and Horwitz, 1987). 

Studies have focused on the genetic aspects of the uptake of I. pneu- 
mophila in its host cells. I. pneumophila mutants impaired in different loci, 
such as rtxA and enhC, display significantly reduced entry into host cells, § 

compared with wild-type bacteria (Cirillo et al., 2000). Recently, it has been g 

shown that the enhanced phagocytosis of I. pneumophila by mammalian cells £ 

is dot/icm dependent (Hilbi et al., 2001). Interestingly, the dot/icm genes de- w 

lay uptake and induce macropinocytosis in A/J mice macrophages (Watarai ° 

et al., 2001b). Macropinosomes containing L. pneumophila are induced tran- 8 

siently and shrink rapidly (5-15 min; Watarai et al., 2001b), and this mode of % 




o 
uptake is linked to the Ignl locus on chromosome 13 of mice (Watarai et al., g 

2001b) . With the exception of A/J mice, most of the inbred mouse strains are £ 

not permissive to L. pneumophila infection; neither are macrophages isolated g 

o 
from these mice (Yamamoto et al., 1992; Beckers et al., 1995). The differ- g 

ence between these two mice strains is located on chromosome 13 and is 2 

linked to a single locus, Ignl (Dietrich et al., 1995; Beckers et al., 1997). In g 

macrophages of nonpermissive strains of mice, the macropinocytic uptake § 

of I. pneumophila is reduced (Watarai et al., 2001b). The Ignl allele makes the c 

bacteria behave as if they are lacking the dot/icm system (Watarai et al., 2001b) . g 

Thus, the Ignl allele is required for dot/icm-dependent macropinocytosis and c 

is; 

delayed uptake by mice macrophages (Watarai et al., 2001b). o 

m 



INTRACELLULAR TRAFFICKING 

Intracellular survival and replication within host cells 

During the first few minutes after entry into amoebae, the bacterium is 
enclosed in a phagosome surrounded by mitochondria and host cell vesicles 
(Abu Kwaik, 1996; also see Fig. 5.2). The bacterial phagosome is blocked 
from fusion to the lysosomes (Bozue and Johnson, 1996). In addition, the 




phagosome is surrounded by a multilayer membrane derived from the RER 
of amoebae (Abu Kwaik, 1996; also see Fig. 5.2). Following formation of this 
phagosome within protozoan cells, bacterial replication is initiated. The 4-h 
period prior to initiation of intracellular replication may be the time required 
to recruit these host cell organelles that may be required for replication. Al- 
ternatively, the 4-h period may be a lag phase of metabolic and environmental 
adjustment of the bacteria to a new niche. 

Similar to the protozoan infection, within 5 min following entry of the 
bacteria into macrophages and monocytes, the I. pneumophila phagosome 
is surrounded by host cell organelles such as mitochondria, vesicles, and 
the RER (Horwitz, 1983b; Swanson and Isberg, 1995a; Tilney et al., 2001; 
also see Fig. 5.2). Also similar to the trafficking of I. pneumophila within 
protozoa, the phagosome within mammalian macrophages does not fuse to 

* lysosomes (Horwitz, 1983a, 1984; Horwitz and Maxfield, 1984; also see Fig. 
g 5.2). The role of the RER in the intracellular infection is not known, but the 

§ RE R is not required as a source of protein for the bacteria (Abu Kwaik, 1 998a) . 

< 

ph Interestingly, examination of the intracellular infection of macrophages, alve- 

z olar epithelial cells, and protozoa by another Legionella species, I. micdadei, 

Q showed that, within all of these host cells, the bacteria were localized to RER- 

< free phagosomes (Gao et al., 1999) . Whether other Legionella species replicate 

< within RER-free phagosomes is still to be determined. 

* Macrophages, peripheral blood monocytes, and alveolar epithelial cells 

< support intracellular replication of I. pneumophila (Nash et al., 1984; Gao 
S et al., 1998b). Although alveolar epithelial cells, which constitute more than 
w 95% of the alveolar surface (Gao et al., 1998b), have been shown to allow 
§ intracellular replication of L. pneumophila, their role in the pathogenesis has 
° been largely overlooked. 

w 
i-i 
i-i 

s Role of the dot/icm genes in evasion of the endocytic pathway 

The Dot/icm type IV secretion system is the main virulence system of 
L. pneumophila. Because the Dot/icm secretion system is ancestrally related 
to type IV secretion systems that mediate conjugal DNA transfer between 
bacteria (Christie and Vogel, 2000), I. pneumophila may utilize this trans- 
porter to transfer macromolecules into the host cell to evade endocytic fusion 
(Roy and Tilney, 2002). The dot/icm loci may be involved in the insertion 
of a pore in the host cellular membrane to transfer the effector proteins 
(Kirby and Isberg, 1998; Kirby et al., 1998). The effector molecules involved 
in intracellular trafficking and evasion of the lysosomal fusion within mam- 
malian cells are limited to the phagosome harboring the bacterium, which 



does not alter the biology of endocytic fusion in the rest of the cell (Coers 
et al., 1999). With few exceptions, the function of individual Dot/Icm pro- 
teins is unknown. 

The Dot A protein was the first to be described (Berger et al., 1994). It is a 
polytopic inner membrane protein with eight hydrophobic transmembrane 
domains (Roy and Isberg, 1997). dotA mutants are defective in all virulence 
activities that require the Dot/Icm complex (Berger and Isberg, 1993; Berger 
et al, 1994; Swanson and Isberg, 1995a, 1995b; Kirby et al., 1998; Coers 
et al., 2000). These data are supported by the fact that the DotA sequence 
possesses significant similarities with that of TraY (Segal et al., 1999; 
Komano et al., 2000), a component of the type IV transporter required for 
the conjugal transfer of plasmids ColIbP9 and R64. However, Nagai and 
Roy have shown that the DotA protein is also secreted through the Dot/Icm 
transporter into the culture supernatant during growth of I. pneumophila in § 

liquid broth by means of a functional type IV secretion system (Nagai and g 

Roy, 2001) . Electron micrographs also show that DotA is part of the oligomer £ 

that could be the membrane channel (Roy and Isberg, 1997; Nagai and Roy, w 

2001) . Mutants defective in DotA protein expression are unable to form pores ° 

in host cell membranes (Coers et al., 2000). 8 

As DotH/IcmK and DotO/IcmB in growing I. pneumophila cultures are % 




o 
mainly associated with the membranous fraction, and as dot/icm products g 

may be required during direct contact with host cells, the location on the £ 

surface of L. pneumophila of DotH and DotO proteins has been examined g 

o 
(Watarai et al., 2001a) . These proteins are surface exposed and associated with g 

a fibrous structure on L. pneumophila after exposure to bone-marrow-derived 2 

macrophages. In contrast, during broth culture, this fibrous structure seems g 

to be absent (Watarai et al., 2001a). However, with the use of dotA, dotB, dotH, § 

and dotO mutants, it has been shown that the exposure of DotO/DotH on the c 

bacterial surface is not dependent on other Dot/Icm proteins, including the g 

DotH protein, for the DotO exposure and vice versa (Watarai et al., 2001a). c 

This result shows that the surface exposure of DotH and DotO after contact o 

with macrophages is not dependent on an intact Dot/Icm secretion system 2 

(Watarai et al., 2001a). The surface exposure of these two proteins does not 
involve bacterial contact with the target cell because bacteria incubated in 
medium that have been conditioned by bone-marrow-derived macrophages 
for 24 h yield almost identical results (Watarai et al., 2001a). 

In addition, DotH and DotO surface exposure on L. pneumophila re- 
quires intracellular growth of the bacteria in macrophages and is observed 
late in the infection process, mostly when there are more than 30 bacteria per 
phagosome (Watarai et al., 2001a). In fact, surface-exposed DotH and DotO 




disappear after uptake but reappear following intracellular growth. The expo- 
sure of these two proteins increases L. pneumophila uptake into cells (Watarai 
et al., 2001a) and may be necessary to promote bacterial escape from the 
phagosome and to facilitate the initiation of a new infection in macrophages 
(Watarai et al., 2001a). These data may also suggest that macrophage com- 
ponents are able to induce changes in the L. pneumophila envelope, and 
DotH/DotO export may occur as a response to the target macrophages just 
before uptake to allow efficient initiation of intracellular growth. 

Dot/Icm proteins such as IcmR, IcmQ, IcmX, or IcmW do not present 
sequence homology to other protein components of the type IV secretion sys- 
tem (Segal and Shuman, 1998; Vogel et al., 1998; Christie and Vogel, 2000). 
IcmX, a periplasmic protein, is required for pore-forming activity (Matthews 
and Roy, 2000). The icmXgene is also required for biogenesis of the replica- 

* tive phagosomes containing I. pneumophila (Matthews and Roy, 2000). A 
g truncated IcmX product is secreted into culture supernatants by wild-type 

§ L. pneumophila growing in liquid media, but the transport of the protein into 

< 

ph eukaryotic host cells has not been detected (Matthews and Roy, 2000). 

s icmS and icmW mutants are not defective in pore-forming activities, but 

Q their phagosomes fuse to lysosomes (Coers et al., 2000). IcmS and IcmW 

< are small proteins required for the trafficking of the Legionella-containmg 

< phagosome, and they may function as chaperones to help the secretion of 

* proteins through the type IV secretion system (Coers et al., 2000; Nagai and 

< Roy, 2001). It has also been shown that IcmS and IcmW interact, suggesting 
S that they may be components of a protein complex that is required for modu- 
le lating phagosome biogenesis (Coers et al., 2000). Interestingly, phagosomes 
pi 

g harboring icmS and icmW mutants still recruit host vesicles, including the 

,_i 

2 RER. It is still to be confirmed whether recruitment of host vesicles does not 

3 prevent the phagosome from lysosomal fusion (Roy and Tilney, 2002). 

< The IcmR protein may also be a chaperone (Coers et al., 2000). An icmR 
mutant has undetectable pore-forming activity and can partially evade the 
endocytic pathway, but it eventually fuses with the lysosomes (Coers et al., 
2000). These data suggest that IcmR may be a cofactor for another protein 
effector involved in evasion of lysosomal fusion (Roy and Tilney, 2002). The 
icmR mutant that evades endocytic fusion does not recruit host vesicles, and 
the phagosome is not surrounded by the RER (Roy and Tilney, 2002). Thus, 
effector molecules that recruit host cell vesicles may be different from the 
ones involved in evasion of the endocytic pathway (Roy and Tilney, 2002). 

Similar to dot A and icmX mutants, an icmQ mutant is defective in all 
virulence activities (Coers et al., 2000). Furthermore, the icmQ gene, like 
the icmR and icmS genes, encodes soluble protein. The IcmR and IcmQ 



proteins interact as protein chaperone-substrate. The presence of IcmR, 
which has chaperone characteristics (it is acidic, small, and predicted to have a 
hydrophobic alpha-helix in its C-terminal domain) affects the physical state of 
IcmQ directly (Dumenil and Isberg, 2001). IcmR prevents IcmQ from partic- 
ipating in the formation of high-molecular-weight complexes by dissociating 
IcmQ homopolymers (Dumenil and Isberg, 2001). 

It has been shown that most of the dot/icm genes required for intracel- 
lular growth within human cells are also required for intracellular growth 
in the protozoan host A. castellanii (Segal and Shuman, 1999). In addition, 
enhanced phagocytosis by A. castellanii of wild-type L. pneumophila has also 
been demonstrated to be dependent on dot/icm genes (Hilbi et al., 2001). 
Although some loci have been shown to be essential only for the intracellu- 
lar growth of I. pneumophila in macrophages (Gao et al., 1998a), numerous 
loci have been identified as essential for survival and intracellular replica- § 

tion of I. pneumophila in A. polyphaga or H. vermiformis and in macrophages £ 

(Cianciotto and Fields, 1992; Gao et al, 1997; Stone et al, 1999). £ 

Another host, Dictyostelium discoideum, a unicellular organism that lives w 

in soil, has been shown to support the intracellular multiplication of L. pneu- ° 

mophila (Solomon and Isberg, 2000; Solomon et al., 2000). In the amoebal 8 

form, the cells are highly motile and are very active in phagocytosis. This % 




o 
model is interesting because genetic tools are available for analysis of host- g 

pathogen interactions, such as the existence of plasmids replicating in this £ 

haploid organism as well as extensive sequence DNA information (Solomon g 

o 
and Isberg, 2000; Solomon et al., 2000). The intracellular fate of I. pneu- g 

mophila is very similar in infected D. discoideum to that in macrophages, 2 

including the recruitment of RER and evasion of lysosomal fusion (Solomon g 

et al., 2000) . The growth of I. pneumophila in D. discoideum depends on dot/icm § 

gene functions (Solomon et al. , 2000) . The analysis of growth of wild-type and c 

three isogenic dot/icm mutant strains of I. pneumophila in D. discoideum indi- g 

cates that intracellular growth requires the products of multiple loci, because c 

is; 

dotH, dot I, and dotO mutants all failed to grow and lost viability over the o 

course of 4 days. (Solomon et al., 2000). The similarity between the infection 2 

by L. pneumophila of different protozoa supports the idea that the ability of L. 
pneumophila to parasitize macrophages and hence to cause human disease is 
a consequence of its prior adaptation to intracellular growth within protozoa. 



RECRUITMENT OF RER 

In 1982, Katz and Hashemi showed that the L. pneumophila-containmg 
phagosome resembles the ER membrane. By means of electron microscopy, 




replicating L. pneumophila within macrophages were visualized located 
within organelles morphologically identical with the RER (Katz and Hashemi, 
1982; also see Fig. 5.2). Later, several studies demonstrated that upon 
internalization of L. pneumophila by the host cell, the Legionella-containmg 
vacuole recruits organelles such as vesicles, mitochondria, and ER (Fig. 5.2; 
also see Katz and Hashemi, 1982; Horwitz, 1983b; Bozue and Johnson, 1996). 
With the use of fluorescent markers specific for the ER, it has been 
shown that the L. pneumophila-containing vacuoles may resemble nascent 
autophagosomes (Swanson and Isberg, 1995b). Autophagy is a physiologi- 
cally important cellular process for the degradation of unwanted organelles 
and cellular components, during which the autophagosome is formed from 
RER and fuses to lysosomes (Dorn et al., 2002). The hypothesis that L. pneu- 
mophila exploits the autophagy machinery in host cells and establishes an 

* intracellular niche favorable for replication (Swanson and Isberg, 1995b) has 
g been challenged recently (see the paragraphs that follow; also Tilney et al., 

g 2001; Roy and Tilney, 2002). 

< 

ph Recent studies suggest that fusion (Roy and Tilney, 2002), or exchange 

=; of lipid bilayer with E R vesicles on the L. pneumophila-containing phagosome 

Q (Tilney et al., 2001) occurs, allowing the phagosomal membrane to become 

< as thin as the ER membrane, with similar characteristics (Tilney et al., 2001; 

< Roy and Tilney, 2002). It has been shown that, within 5 min of uptake, host 

* vesicles come into contact with wild-type Legione^a-containing phagosomes 

< and flatten along the surface of the phagosome, and this process is com- 
S pleted within 15 min (Tilney et al., 2001). This does not occur in dot/icm 

w mutant-containing phagosomes (Tilney et al., 2001). This is consistent with 

pi 

g earlier studies that have shown that, after 4 h of infection, there are only 

,_i 

° few vesicles associated with the phagosomal membrane, but there are ribo- 

3 somes studding the phagosomal membrane (Horwitz, 1983b). Interestingly, 

< the thickness of the phagosome membrane becomes similar to the ER mem- 
brane within 15 min (Tilney et al., 2001). Thus, within 15 min of infection, 
the phagosomal membrane resembles that of the ER. The ribosomes at 6 h 
stud the phagosomal membrane, and I. pneumophila is thought to be located 
within the RER (Tilney et al., 2001) . However, these studies rely completely on 
the thickness of the membranes of the ER and the phagosome membranes to 
provide evidence that L. pneumophila is located within the RER (Tilney et al., 
2001). Immunocytochemical studies should be performed to confirm these 
observations. It is likely that the recruitment of the ER may be involved in 
the biogenesis of the phagosome that is dependent on the type IV secretion 
system, because the dot/icm mutants are unable to recruit RER and their 
phagosomes fuse to the lysosomes (Swanson and Isberg, 1995b). 



Interestingly, a recent study showed that recruitment of RER to matur- 
ing phagosomes may be part of regular phagoytosis (Gagnon et al., 2002). 
Electron micrographs of latex beads and pathogens such as Salmonella within 
macrophages have shown that ER membranes fuse with plasmalemma, un- 
derneath the phagocytic cup, and successive waves of ER are recruited to 
the phagosomes of latex beads and bacteria. In neutrophils, the bacteria are 
quickly killed and the ER is not involved in the turnover of membrane for 
phagocytosis (Gagnon et al., 2002). However, because the dot/icm system is 
essential for RER recruitment (Swanson and Isberg, 1995b), it is likely that 
L. pneumophila utilizes a specific pathogen-regulated process to recruit vesi- 
cles and that this process is distinct from regular phagocytosis. 



THE ARF PROTEIN g 

o 

The protein ADP ribosylation factor-1 (ARF1), a highly conserved small £ 

GTP-binding protein that acts as a key regulator of vesicle traffic from ER to £ 

Golgi, is found on phagosomes that contain wild-type I. pneumophila but not w 

dot/icm mutants (Nagai et al., 2002) . These data suggest that a protein injected ° 

through the type IV secretion system may be required for ARF1 recruitment. 8 

Searching the L. pneumophila genome for proteins that have homology to z 




ARF-specific guanine nucleotide exchange factors (GEFs), Nagai et al. have Q 

identified a protein, RalF, that has a sec7-homology domain, known to be £ 

sufficient to stimulate the exchange of GDP for GTP (Nagai et al., 2002). It g 

o 

has been shown that RalF is injected through the phagosomal membrane g 

by a process that requires the dot/icm system (Nagai et al., 2002). However, 2 

phagosomes containing ralF mutants that do not recruit ARF1 evade fusion g 

to lysosomes, and the bacteria replicate intracellularly within macrophages § 

and amoebae (Nagai et al., 2002). Thus, RalF is not essential for transport of c 

L. pneumophila to the ER (Nagai et al., 2002; Roy and Tilney, 2002). g 



CJ 

S 

o 



Pore-forming activity g 

Kirby et al. were the first to describe the pore-forming ability of I. pneu- 
mophila in host cell membranes (Kirby et al., 1998). This ability was doc- 
umented by lysis of macrophages and red blood cells, which is dependent 
on the Dot/icm secretion system (Kirby et al., 1998). Because dot/icm mu- 
tants are defective in evasion of lysosomal fusion, it has been proposed that 
the pore-forming activity is required for export of effector molecules nec- 
essary for evasion of the lysosomal fusion (Kirby et al., 1998). Upon initial 
contact with the host cell, I. pneumophila may insert a pore into the plasma 




membrane to deliver bacterial effector molecules into the host cell (Kirby 
and Isberg, 1998; Kirby et al., 1998). This concept is supported by the fact 
that many dot/icm mutants are defective in both intracellular replication and 
pore -formation-mediated cytotoxicity (Kirby and Isberg, 1998; Kirby et al., 
1998). Moreover, some dot/icm mutants are defective in trafficking and intra- 
cellular replication but not in pore-forming activity (Zuckman et al., 1999). 
Thus, the pore-forming activity is not sufficient for phagosomal trafficking 
(Zuckman etal., 1999). 

We have identified five spontaneous mutants that are unable to egress 
from the macrophages but are able to grow as well as the wild-type strains 
within these cells (Alii et al., 2000). These mutants, designated rib (release 
of intracellular bacteria) , are defective in the pore-forming toxin activity as 
shown by the contact-dependent hemolysis assay and by permeability to 

* propidium idodide upon infection of macrophages and epithelial cells (Alii 
g et al., 2000). The rib mutants are also defective in acute cytotoxic lethality to 

§ A/ J mice and fail to cause alveolar inflammation (Alii et al., 2000; Molmeret 

< 

ph et al., 2002a), thus indicating a key role for the pore-forming toxin in the 

z pulmonary pathology of the bacterium. We have further documented that 

p the Rib toxin is not required for intracellular trafficking and replication (Alii 

< et al., 2000; Gao and Abu Kwaik, 2000; Molmeret et al, 2002b). 

< In addition to defects in evasion of lysosomal fusion, dot/icm mutants are 

* also defective in induction of apoptosis (Zink et al., 2002), enhancement of 

< phagocytosis by human-derived macrophages (Hilbi et al., 2001), and indue - 
S tion of macropinocytic delayed uptake by A/J mice macrophages (Watarai 

m et al., 2001b). In contrast, rib mutants are defective in pore-forming toxin 

pi 

g but are not defective in evasion of lysosomal fusion, intracellular replica- 

,_i 

° tion, or induction of apoptosis (Alii et al., 2000; Molmeret et al., 2002b; Zink 

3 et al., 2002) . These observations may suggest that there are at least two pores 

< inserted into host membranes through the type IV secretion system at differ- 
ent stages of the infection (Molmeret and Abu Kwaik, 2002). The first pore, 
is the "invasion and trafficking pore," which is inserted upon initial contact 
with the host cell to deliver effectors into the host cell cytoplasm and allow 
the establishment of the intracellular infection. The dot/icm genes necessary 
for the assembly of the secretion apparatus are constitutively expressed and 
are required for this step (Hales and Shuman, 1999). The second pore is the 
"egress pore," which is turned off during exponential intracellular replica- 
tion but is triggered upon cessation of replication and is essential for lysis of 
the host cell (Figs. 5.3 and 5.4; also see Alii et al., 2000). This is consistent 
with the fact that, upon transition into the postexponential growth in vitro, 
L. pneumophila becomes cytotoxic (Byrne and Swanson, 1998). Thus, the rib 



24 h 



48 h 




8 '^tfSFfA^' 



5 mB** 







■•■y 












Figure 5.3. The rife mutants' defect in cytolysis of the host cell is due to a defect in 
necrosis-mediated killing. Representative transmission electron micrographs of infected 
U937 macrophages at 24 h and 48 h postinfection by the wild-type strain AA100 and the 
GN229 mutant. The original magnifications were 7,000x and 5,000x for the 24-hp and 
48-h infections, respectively. (This figure was adapted from Alii et al., Infect Immun. 
68 (11): 6431-6441, 2000.) 



mutants retain the "invasion and trafficking pore" but are defective in the 
"cytolysin/egress pore," because they replicate within but fail to egress from 
the host cells (Figs. 5.3 and 5.4). 

The phenotype of the rib mutants is attributable to a point deletion in the 
icmT gene that is predicted to result in a truncated protein of 54 amino acids 
instead of the 86 amino acids of the native protein (Molmeret et al., 2002a, 
2002b). In contrast, an icmT null mutant is defective in both intracellular 
replication and pore formation (Molmeret et al., 2002a, 2002b). It has been 
shown that icmT expression is induced at the stationary phase compared 
with the exponential phase (Gal-Mor et al., 2002). We have shown that IcmT 
is bifunctional and that the carboxy terminus is essential for the pore-forming 
"cytolysin/egress pore" activity and the amino terminus is essential to export 
effectors involved in various pathogenic traits (Molmeret et al., 2002a, 2002b). 

We hypothesized that, upon transition to the postexponential phase of 
growth, the Rib toxin activity is triggered, resulting in insertions of pore first 
in the phagosomal membrane, leading to its disruption and bacterial egress 




o 

M 

W 

o 

a 

M 

n 
o 

M 

O 

o 

> 
z 

o 

n 

w 

M 

CO 
I— I 

o 

M 

O 

o 
o 

H 

tn 

O 
H 

o 

% 

H 



% 

hi 

CJ 

S 

o 
HI 




t 




t 




t 




t 




■■■, % 

< 




<*. 



£ 



so 

c 



£ 'd 

II 

o 



T. 

I 




r-.] 



"d 

a 



O 

■ T— I 
+^ 

rt _ 
u 

• — i (L) 

!-H 



o 

Oh 

D 



■ i— I 
!-H 

CD 

u 

!-H 

rt 

3 

CD 

u 
rt 

!-H 



-a 

CD 
C 

o 

!-H 
!-H 



o 

a 
o 

•i-H 

13 

a 



1 



CD 

!-H 

u 



a a 



CD 
O 

Ph 

'— - 

s 

Oh 

u 

a 

■ i-H 

s 
s 



o 

■ i-H 

w 

J^ 

O 

U 

-M 

CD 

a 

CD 

Oh 

CD 

CD 
w 



o 

!-H 

OX) 

o 

-d 

o 



LO 

(D 

!-H 



CD 

O 


o 

■ T— I 

13 



!-H 



o 

w 

CD 



!-H 



CO 

O 



CD 



OT 

w 

CD 

!-H 



OD 



O 
w 
O 
GL 

Oh m 






C/3 



Cfl 

O 

-M 
Oh 

O 

Oh 
CT3 

u 

•4-H 

n3 



.s s g 



-d 

O 

s 



a ^ 



CD 

M 

CT3 

Oh 

Cfl 

rt 
U 

fctf 

o 

u 
a 

in 

3 



w 

"^H 

o 
-d 

a 

CD 

■a 

o 

'd w 
fl -^ 



c« 

a 

i-H 

o 

Oh 

o 

w 

O 

o 

•i-H 

tn 
CD 
W 

O 



Cfl 

w 

CD 

J-H 

u 

•i-H 

CD 

i-H 

CD 
££ 



U 

CT3 



U 

i-H 
O 
Oh 

O 

H-H 
<H-H 

o 



o 



w 

Th 

rt 
CD 

0D nJ 



El 
o 

o 



a u 



w 



>x 



u 

o 



-a 
a 

Is 

° S fe 



U 

Oh 

nJ 



CD 

Vh 

O 
Oh 

CD 
J3 



O 

d 

o 

M 

CD 

!-H 

Oh 

X 

CD 



I 

u 

o 



CD 
w 
rt 

Oh ^ 
^ Oh 



in 
CT3 



CD 'H 

U * 



u 
u 

n3 

!-H 



c .a 

Oh c3 



03 

•i-H 

o 

CD 

a 

o 

Oh 
X 

a 



w 

o 

_ Oh 

a « 
g | 

o P 

Oh 

X 

o 
&D 

a 



a 
o 



i-H 

... 3 

ti. CD T3 



a 



W 

•--H 

M 

U 



w 

O 



OJD 

a 



tu 



rt 



tn 

03 

"3h 

CD 



w 

CD 

S-H 

O 

Oh 



!-H 

CD 
W 

a 

• i-H 

T3 



w 



w 



5S 



-d 

w 

-M 
•i-H 

o 

a 

•i-H 



u 



tn 

CD 

a 

!-H 



CD 



13 



o 

w 

o 



a 
o 

•i-H 

t) 

CD 
w 

!-H 

u 



CD 
CD 
T3 c/5 

OJ 

CD 



O 
O 
O 



CD 

a 

rt 



O 



CD rn 



o 



00 

si 



a _3 

o 13 



CD 



^ 



CD t -w 
O CD 

13 

rt 



en 

rt 

CD 

i-H 



a « 

o 



<J1 



!-H 

W 



rt 

•i-H 

!-H 

CD 

-t-n 

u 

rt 
^Q 

u 

<?$ 

3 



cd y 

!-H 



CD 

H-> 

o 

CD 
w 
rt 
CD 

Id 

!-H 

-d 

a 

rt 
u 

CD 



5P ^a 

rt T3 

^a uh 

Oh O 



into the cytoplasm ( Figs .5.3 and 5 .4) . To test this hypothesis , we examined late 
stages of the intracellular infection of macrophages and amoebae by electron 
microscopy. Our data showed that disruption of the phagosomal membrane 
was detectable 12 h postinfection in both A. polyphaga and macrophages (un- 
published data). After 12 and 18 h postinfection in both host cells, vesicles 
and organelles from the host cytoplasm entered into the LegioneHa-containing 
phagosome. Between 18 h and 24 h, most of the bacteria present in this dis- 
rupted phagosomal membrane are surrounded by cytoplasmic organelles, 
and not by distinct phagosomal membrane (unpublished data) . These data 
show that the phagosomal membrane is disrupted first; there is no simulta- 
neous lysis of both the phagosomal and the plasma membranes. Whether this 
disruption is the result of a mechanic process or is linked to the pore-forming 
activity of the type IV secretion system (Kirby et al., 1998; Molmeret and Abu 
Kwaik, 2002; Molmeret et al., 2002b) is not known. Our data are consistent ^ 



and outer membrane protein profiles (Barker et al., 1993). After growth in 
host cells, L. pneumophila becomes more resistant to biocides and antibiotics 
(Barker et al., 1992, 1995), more invasive to host cells, and more virulent in 
the guinea pig model (Cirillo et al., 1994, 1999; Brieland et al., 1997). 

In contrast to exponentially growing bacteria, I. pneumophila obtained 
from postexponential cultures expresses traits that are correlated with viru- 
lence, including sodium sensitivity, cytotoxicity, osmotic resistance, motil- 
ity, and the capacity to evade phagosome-lysosome fusion (Rowbotham, 
1980, 1986; Byrne and Swanson, 1998; Hammer and Swanson, 1999). Dur- 
ing the replication phase within host cells, I. pneumophila organisms are 



M 



M 
M 

CO 



with a previous study in which Katz and Hashemi (1982) had observed on £ 

electron micrographs that, when present in macrophages at numbers greater £ 

than 25 per cell, I. pneumophila was usually dispersed within the cytoplasm w 

of the host cell. ° 

o 

o 

> 
GENE REGULATION IN L PNEUMOPHILA £ 

Growth phase and gene regulation 

O 

The phase of growth has a dramatic effect on the phenotype of L. pneu- g 

mophila, whether the bacteria are cultivated in host cells or in microbiological 2 

medium (Rowbotham, 1986; Byrne and Swanson, 1998). Robowtham was the g 

first to observe in amoeba two distinct phenotypic phases of L. pneumophila, § 

the replicative phase and the "active infective phase" (Rowbotham, 1986). £ 

When I. pneumophila is released from eukaryotic cells, it is short, thick, and g 

highly motile, with a smooth and thick cell wall (Rowbotham, 1986) . The cells c 

have a different composition of membrane fatty acids, and they differ in LPS o 

m 




sodium resistant and aflagellated (Rowbotham, 1980, 1986; Byrne and 
Swanson, 1998; Hammer and Swanson, 1999). When I. pneumophila bacteria 
egress from host cells, they are flagellated and sodium sensitive (Rowbotham, 
1980, 1986; Byrne and Swanson, 1998; Hammer and Swanson, 1999). It has 
been hypothesized that amino acid limitation induces the virulent phenotype 
(Rowbotham, 1980, 1986; Byrne and Swanson, 1998; Hammer and Swanson, 
1999). When L. pneumophila enters into the postexponential growth phase 
or is subjected to amino acid limitation, the bacteria accumulate the strin- 
gent response signal, guanosine S'^'-bispyrophosphate (ppGpp), through 
the ppGpp synthetase, RelA (Bachman and Swanson, 2001). When relA from 
Escherichia coli is expressed in L. pneumophila, ppGpp accumulates and the 
bacteria express virulent traits independent of nutrient supply or cell den- 
sity (Bachman and Swanson, 2001). The accumulation of ppGpp increases 

* the amount of the stationary-phase sigma factor RpoS, which triggers the 
g expression of the stationary-phase genes (Bachman and Swanson, 2001). 

§ A rpoS mutant replicates within monocyte HL 60 and THP-1 cells but 

< 

ph is attenuated in A. castellanii (Hales and Shuman, 1999). Therefore, some 

z RpoS -regulated traits could be critical for efficient transmission or infection 

Q in this amoebae model (Hales and Shuman, 1999). Sodium sensitivity and 

< maximal expression of flagellin also requires RpoS (Bachman and Swanson, 

< 2001). I. pneumophila in the postexponenetial phase becomes cytotoxic by 

* an RpoS-independent pathway (Bachman and Swanson, 2001). To replicate 

< efficiently in macrophages, both an RpoS-independent and RpoS-dependent 
S mechanism are utilized by L. pneumophila (Bachman and Swanson, 2001). 
m Thus, when nutrient levels and other conditions are favorable, L. pneumophila 
§ replicates within host cells, and when amino acids become rare, intracellular 
o bacteria express several traits that permit escape from the host cell, survival 

3 in the environment, and transmission to a new host. 

i-i 

:w 

< 

The transmission phenotype 

Because flagellin is expressed at the postexponential phase, Hammer 
et al. (2002) screened for L. pneumophila mutants deficient in flagellin ex- 
pression in order to identify genes involved in the phenotypic transition at 
the postexponential phase. Five activators of virulence have been identified: 
LetA and LetS, a two-component regulator homologous to GacA/S of Pseu- 
domonas, or SirA/BarA of Salmonella, that represses the flagellar regulon; the 
stationary-phase sigma factor RpoS; the flagellar sigma factor FliA, required 
for both motility and contact-dependent cytotoxicity (Heuner et al., 2002) ; and 
a novel locus, letE (Hammer et al., 2002). 




In the postexponential phase, mutants in letA, letS, JliA, and letE are 
nonmotile or show poor motility, are not cytotoxic nor sodium sensitive (ex- 
cept JliA), and are not efficient at infecting macrophages (except letE; Ham- 
mer et al., 2002). In contrast, intracellular growth is independent of these 
genes (Hammer et al., 2002). Amino acid depletion or ppGpp accumulation 
triggers a LetA/ LetS expression and RpoS -dependent cellular differentiation 
(Hammer et al., 2002). The letE locus does not appear to produce any protein 
and may encode for a regulatory RNA that may act as a regulator of letA /let S 
expression (Hammer et al., 2002). 

The relA mutant, producing undetectable levels of ppGpp in the cells 
during the stationary phase, is unable to produce pigment as it becomes 
flagellated. Although a previous study has shown that RpoS, which ac- 
cumulates when RelA is activated, is required for intracellular growth in 
A. castellanii (Hales and Shuman, 1999). Zusman et al. (2002) have shown § 

that relA gene product is dispensable for intracellular growth in H L-60-derived g 

human macrophages and in A. castellanii. Moreover, it has also been shown £ 

that RelA and RpoS have minor effects on expression of some of the dot/icm w 

genes (icmT, icrnR, icmQ icmF, icmM, icm], icmF, icmV, and icmW; see ° 

Zusman et al., 2002). Thus, the role of RpoS in intracellular infection seems 8 

to be specific to the host cell. % 

n 

w 
I- 1 

DNA regulatory elements of the dot/icm genes 2 

O 

Several studies showed that dot/icm expression is regulated during the g 

intracellular infection of L. pneumophila within the host cells. The dot/icm 2 

system upregulates phagocytosis of I. pneumophila (Hilbi et al., 2001), and g 

surface exposure of DotO/DotH on L. pneumophila is induced at the earlier § 

and later stages of the infection (Watarai et al., 2001a). c 

Gal-Mor et al. (2002) examined DNA regulatory elements that may con- g 

trol the expression of the dot/icm genes, using a promoter fusion to lacZ in c 



I. pneumophila. They showed that the expression levels of different dot/icm o 

genes are distinct from one another. The icmR, icmF, icmV, and icm W genes 2 

have high levels of expression in both the exponential and postexponential 
phase of growth, whereas icmR and icmF have higher levels in the stationary 
phase than in the exponential phase. The icmT, icmP, icmQ icmM, and icm] 
genes show low levels of expression in both exponential and postexponential 
phases of growth. However, icmT and icmF have a higher level of expression 
at the stationary phase. 

A total of 12 regulatory elements have been identified (Gal-Mor et al., 
2002); 10 promoter elements of icmT, icmF, icmQ, and icmM genes have low 




expression levels. These promoters contain 6-bp putative consensus sequence 
TATACT, located from 32 to 74 bp from the ATG codons, which is essential 
for their expression. — 10 promoter elements of icm V, icmW, and icmR genes 
that have high expression levels have also been identified. The icmR locus con- 
tains at least three regulatory elements, and regulatory elements have been 
also identified for icmW, icmV, and icmF that have high expression levels. A 
9-bp putative consensus sequence CTATAGTAT has been observed. In addi- 
tion, icm V and icm W have an overlapping regulatory region (Gal-Mor et al., 
2002). 

The reduced effect of an insertion in rpoS or relA on expression of icmP 
only among the 9 icmr.lacZ fusions tested has shown that these two genes are 
not involved in the regulation of the dot/icm genes (Hales and Shuman, 1999; 
Zusman et al., 2002) and that there may be other factors necessary for this 

* regulation. The —10 promoter elements found in some dot/icm genes have 
g extensive homology to one another and are probably recognized by I. pneu- 

§ mophila RpoD (Gal-Mor et al., 2002). An examination of the I. pneumophila 

< 

ph genome sequence has shown that I. pneumophila contains homologs of at 

z least six sigma factors (RpoD, RpoH, RpoF, RpoE, RpoS, and RpoN). The 

Q promoter sequences of RpoH and RpoF are different from the — 10 promoter 

< sequences of the dot/icm genes (Gal-Mor et al., 2002). In addition, the pro- 

< moters recognized by the factors RpoE and RpoN are also different from the 

* promoters of the dot/icm genes (Gal-Mor et al., 2002). As RpoS is not involved 

< in the expression of the dot/icm genes except for the moderate expression of 
S icmP (Zusman et al., 2002), it has been proposed that these —10 regulatory 
w elements of the dot/icm genes are recognized by the vegetative sigma factor 
3 RpoD (Gal-Mor et al, 2002). 

O 

w 
1-1 

: 3 APOPTOTIC OR NECROTIC CELL DEATH 

Induction of apoptosis by L pneumophila in mammalian 
but not protozoan host cells 

Apoptosis requires a cascade of activation of a family of cysteine proteases 
(caspases) that specifically cleave proteins after aspartate residues (Anderson, 
1997) . Among them, caspase-3 plays a central role, allowing caspase-activated 
DNase to enter the nucleus and degrade chromosomal DNA (Enari et al., 
1998). Muller et al. (1996) have shown that L. pneumophila induce apoptosis 
in HL-60 human-derived macrophages after 24-48 h, at a multiplicity of 
infection (MOI) of 10-100. The induction of apoptosis in mammalian cells is 
mediated by activation of caspase-3 that is dose dependent and is maximal at 



3 h postinfection at MOI 50 (Gao and Abu Kwaik, 1999a, 1999b; also see Fig. 
5.4). In alveolar epithelial cells and macrophages, the induction of apoptosis 
is dose dependent but not largely growth phase regulated and can be induced 
by extracellular bacteria (Gao and Abu Kwaik, 1999a, 1999b). The dot/icm 
mutants of L. pneumophila are defective in inducing caspase-3 activation and, 
thus, apoptosis (Gao and Abu Kwaik, 1999a; Zink et al., 2002). Therefore, 
the Dot/icm type IV secretion system of I. pneumophila is essential for the 
induction of apoptosis in mammalian cells (Gao and Abu Kwaik, 1999a; 
Zink et al., 2002). The pore-forming toxin is not required for the induction of 
apoptosis, but upon entry into the postexponential growth phase it enhances 
the ability of the bacteria to induce apoptosis (Zink et al., 2002). 

Interestingly L. pneumophila induces apoptosis in human phagocytes 
but not in protozoan host cells such as A. castellanii (Hagele et al., 1998; 
Gao and Abu Kwaik, 2000). Moreover, although A. polyphaga is capable of § 




undergoing apoptosis, upon stimulation by actinomycin D, I. pneumophila g 

does not induce apoptosis in A. polyphaga (Gao and Abu Kwaik, 1999b, £ 

2000). u 

o 

M 

O 

o 

Induction of necrosis by L pneumophila in mammalian > 

and protozoan cells ° 



w 

M 
M 

CO 



The dot/ icm-regulated Rib pore-forming toxin is essential for L. pneu- 
mophila induction of necrosis, killing, and exiting the host cells (Alii et al., g 
2000; Gao and Abu Kwaik, 2000; Molmeret et al., 2002a). At high MOI, mu- g 
tants defective in the Rib pore-forming toxin replicate like the wild-type strains 2 
within the protozoan and mammalian cells but are trapped within these cells g 
and cannot be released ( Figs .5.3 and 5 .4; also see Alii et al. , 2000; Gao and Abu § 
Kwaik, 2000). The expression of the pore-forming activity by L. pneumophila c 
grown in vitro and within macrophages is completely repressed during expo- 5 

nential growth but is temporally activated upon entry into postexponential c 

is; 

phase (Fig. 5.4; also see Alii et al., 2000). All five rib mutants mentioned o 

earlier that possess identical point deletions following a poly(T) stretch in 2 

the icmT gene are also defective in acute lethality in A/ J mice (Alii et al., 
2000; Molmeret et al., 2002b). Therefore, the Rib pore-forming activity is 
not required for phagosomal trafficking and intracellular multiplication of 
L. pneumophila within the cells, but it is essential for induction of cytolysis of 
the infected macrophages (Alii et al., 2000; Gao and Abu Kwaik, 2000). Thus 
the C-terminus of IcmT is essential for pore-formation-mediated cytolysis. 

L. pneumophila kills A. polyphaga through temporal induction of necro- 
sis, which is mediated by the Rib pore-forming activity of L. pneumophila. 




Wild-type intracellular I. pneumophila causes necrosis-mediated cytolysis of 
A. polyphaga within 48 h after infection (Gao and Abu Kwaik, 2000) . However, 
in A. polyphaga, at low MO I the rib mutants are defective in intracellular mul- 
tiplication (Molmeret et al., 2002b). These results show that the C-terminus 
of IcmT is essential for pore formation and intracellular trafficking and mul- 
tiplication within A. polyphaga, as confirmed by fusion of phagosomes har- 
boring the rib mutants to lysosomes (Molmeret et al., 2002a). Experiments 
performed with an IcmT null mutant that is defective in intracellular traf- 
ficking, intracellular multiplication, and egress from both protozoan and 
mammalian cells suggest that IcmT is bifunctional (Molmeret et al., 2002a, 
2002b). The C-terminus of IcmT is essential for pore-formation-mediated cy- 
tolysis in mammalian cells, and the N-terminus is required for intracellular 
trafficking (Molmeret et al., 2002a). It is unlikely that IcmT is an effector 

* or a common regulator, but it is possible that this protein is a cofactor in- 
g volved in the export of different substrates with roles in pore-forming toxicity 

§ and intracellular trafficking (Molmeret et al., 2002a). Effectors of this type IV 

< 

ph secretion system remain to be identified. 

z The ability to lyse host cells and to egress from them is a fundamental step 

Q in the pathogenic cycle of intracellular bacteria and determines the overall 

< consequences of the infection of an organism. Apoptosis and necrosis are 

< the two commonly observed types of cell death. Necrosis is characterized 

i— i 

* by physical damage that causes cell death. Apoptosis is a regulated suicide 

< program of the cell that manifests morphological and biochemical features 
S distinct from those of necrosis (Cohen, 1993). Killing of mammalian cells 

w by L. pneumophila has been proposed to occur in two phases (Gao and Abu 

pi 

g Kwaik, 1999a, 1999b; also see Fig 5.4). In the first phase, L. pneumophila 

° induces apoptosis in macrophages, monocytes, and alveolar epithelial cells 

3 during the early stages of the infection (Hagele et al., 1998; Gao and Abu 

< Kwaik, 1999b), which is mediated through the activation of caspase-3 (Gao 
and Abu Kwaik, 1999a). Induction of apoptosis is largely independent of the 
bacterial growth phase (Gao and Abu Kwaik, 1999a). The second phase is 
mediated through rapid induction of necrosis by L. pneumophila upon entry 
into the postexponential phase of growth, when the bacteria become cytotoxic 
(Fig. 5.4). Our working model of bacterial egress can be presented in three 
steps (Fig. 5.4). First, upon exiting the exponential phase of intracellular 
growth, an "egress pore" is inserted into the phagosomal membrane, leading 
to its disruption. Second, the bacteria egress into cytoplasm. Third, disruption 
of organelles and the plasma membrane occurs, culminating in lysis of the 
host cells and bacterial egress. 




IMMUNE RESPONSE TO L PNEUMOPHILA INFECTION 

When L. pneumophila is inhaled into the lungs, acute alveolitis and bron- 
chiolitis can be observed (Winn and Myerowitz, 1981). The alveolar exudate 
typically consists of polymorphonuclear cells and macrophages, red blood 
cells, and cellular debris (Glavin et al., 1979). I. pneumophila are mainly in- 
tracellular, and the fate of the bacteria depends on the host cells. Within 
macrophages, most of the bacteria are intact in large vacuoles containing 
a large number of bacilli (Glavin et al., 1979). Within neutrophils, L. pneu- 
mophila most often appears degraded (Katz and Hashemi, 1982). 

The role of alveolar epithelial cells in Legionnaires' disease has not been 
well studied. L. pneumophila has been shown to replicate in a phagosome 
within many epithelial cells in vitro, including type I and II alveolar epithelial 
cells (Gao et al. , 1998b) . There is no reason why L. pneumophila is not expected 
to replicate at the foci of infection within alveolar epithelial cells, particularly 
because these constitute most of the alveolar surface, where the foci of infec- 8 

tion are recognized. In addition, these cells are potential sites of intracellular j* 

replication during activation of macrophages by IFN-y, which inhibits in- g 

tracellular replication of I. pneumophila within monocytes and macrophages o 

(Byrd and Horwitz, 1989) . > 

Although the bacteria bind complement component C3, they are resis- ° 

tant to complement-mediated killing (Horwitz and Silverstein, 1981a, 1981b; m 

Payne and Horwitz, 1987) . I. pneumophila is also resistant when treated with 2 

complement and specific antibodies in presence of polymorphonuclear cells 5 

o 

(Horwitz and Silverstein, 1981a, 1981b). Different results have been obtained ^ 

L_J 

with monocytes in experiments performed at a high multiplicity of infection h 

(Horwitz and Silverstein, 1981a). In the presence of both antibody and com- £ 

plement, phagocytosis of L. pneumophila is more efficient and monocytes 2 
kill approximately half of the inoculum of the bacteria (Horwitz and Silver- 



o 

M 

M 



•fl 



stein, 1981a). In addition, antibodies promote fusion of the infected vacuole § 

to lysosomes as compared with cells treated with complement alone (Horwitz is; 



o 



and Maxfield, 1984). However, adherence studies of I. pneumophila to U937 § 

macrophages, guinea pig alveolar macrophages, and MRC-5 cells in absence 
of serum have shown that neither complement nor antibody is required for 
binding (Gibson et al., 1993). Overall, L. pneumophila is relatively resistant 
to innate and humoral immune responses and the infection is most likely 
controlled by cell-mediated immunity. 

Important roles of Thl-type cytokines, such as tumor necrosis factor 
alpha (TNF-a), IFN-y, and interleukin-12 (IL-12), have been demonstrated 




in the murine A/J model of L. pneumophila infection (Brieland et al., 1994, 
1995; Gebran et al., 1994b). Susa et al. (1998) have examined cytokines and 
the role of CD4 and CD8 cells in Legionnaires' disease. After injection of 
10 6 CFU of I. pneumophila in A/J mice, a challenge that allows the survival 
of the infected animals, inflammatory cells are recruited into the lung on 
the second day, and, by the third day of infection, macrophages, B cells, NK 
cells, and large mononuclear cells are mainly present (Susa et al., 1998). 
T lymphocytes infiltrate subsequently (Susa et al., 1998). The L. pneumophila 
infection results in a rapid upregulation of systemic concentrations of IFN- 
y, TNF-a, IL-1/3, IL-4, and IL-6 (but not IL-2); then these levels decrease on 
the third day (Susa et al., 1998). Recruitment of T cells is necessary for the 
clearance of the bacteria. The depletion of CD4+ and CD8+ T cells leads to 
increased lethality to mice. Moreover, another study has shown that nitric ox- 

* ide (NO) produced by numerous cells, such as macrophages and neutrophils, 
g may regulate IL-6 production in L. pneumophila-infected macrophages 

§ (Yamamoto et al., 1996). Indeed, when macrophages were primed with IFN- 

< 

ph y , bacterial replication was inhibited and NO was produced in large amounts 

=; (Yamamoto et al., 1996) . An examination of cytokine levels in L. pneumophila- 

Q infected macrophages primed with IFN-y revealed a moderate increase of 

< IL-6 production (Yamamoto et al., 1996). 

< After intratracheal injection of I. pneumophila, the increase of bacteria in 

* the lungs by 48 h is accompanied by a massive accumulation of neutrophils 

< (Tateda et al., 2001b). Neutrophils are recruited early during the infection of 
S L. pneumophila in animal models and patients (Brieland et al., 1995). How- 

m ever, Legionella are partially resistant to neutrophil killing, particularly when 

pi 

g the bacteria are opsonized (Horwitz and Silverstein, 1980, 1981a; Katz and 

° Hashemi, 1983). Neutrophils have the ability to produce immunoregulatory 

3 cytokines-chemokines, including IL-12, which may affect Thl/Th2 host re- 

< sponses (Cassatella, 1995). Previous studies have shown a protective role 
for Thl cytokines such as IFN-y and IL-12 during L. pneumophila infection 
(Gebran et al., 1994b; Brieland et al., 1995). In contrast, the Th2 cytokine, 
IL-10, facilitates growth of L. pneumophila in macrophages through the IL- 
10-mediated suppression of Thl cytokines (Tateda et al., 2001a). A recent 
study shows that the CXC chemokine receptor 2, a receptor for chemokine- 
mediated neutrophil accumulation, may play a role in host defense against 
L. pneumophila, because blockade of this receptor enhances mortality in the 
A/J mouse model (Susa et al., 1998). The neutrophils may have a protec- 
tive effect through immunomodulatory actions in L. pneumophila infection 
(Tateda et al., 2001b). Early recruitment of neutrophils may contribute to Tl 
polarization in a murine model (Tateda et al., 2001a). 



IFN-y is recognized as an important cytokine in both innate and cell- 
mediated immune responses. Previous studies have shown that endogenous 
IFN-y is induced in response to Legionella infection (Brieland et al., 1994; 
Susa et al., 1998). Moreover, treatment with IFN-y is able to inhibit the 
intracellular replication of L. pneumophila (Nash et al., 1988). When the cul- 
tures are supplemented with iron-saturated transferrin, the IFN-y effect is 
neutralized (Byrd and Horwitz, 1989). Administration of IFN-y in a murine 
model of L. pneumophila induces the expression of IFN-y and IL-12 from 
natural killer cells (Deng et al., 2001). Moreover, intrapulmonary adenovirus- 
mediated IFN-y gene therapy, in a nonlethal murine model of I. pneumophila 
pneumonia, results in a 10-fold decrease in lung bacterial CFU at 48 h postin- 
fection, compared with controls that did not receive the gene. Thus, even 
in immunocompetent hosts, expression of IFN-y promotes L. pneumophila 
clearance, independent of cell recruitment and proinflammatory cytokine § 

induction. Alveolar macrophages from uninfected animals treated in vivo g 

with the adenovirus -mediated IFN-y inhibit the intracellular growth of £ 

I. pneumophila (Deng et al., 2001). Therefore, IFN-y -secreting cells such w 

as T cells and NK cells may directly contribute to bacteriostasis or killing, ei- ° 

ther by downregulating transferrin receptors (Byrd and Horwitz, 1989, 2000; 8 

Gebran et al., 1994a) or by endogenous NO that may regulate IL-6 production % 



CONCLUSION 

L. pneumophila is an intracellular pathogen that utilizes protozoa in 
aquatic environments to replicate within and to be protected from adverse 
conditions. Legionnaires' disease has become a threat to humans after our 
industrialization and production of man-made devices that generate aerosols, 
which allows transmission of the bacteria to humans. Thus, the normal life 
cycle in protozoa stops when the bacteria encounter human host cells such 
as macrophages. Although I. pneumophila utilizes different ligands to enter 
the host cells and is unable to induce apoptosis in protozoan cells such as 
A. polyphaga or A. castellanii, the remarkable similarities in the intracellular 




o 
(Yamamoto et al., 1996). TNF-a? also helps to control the infection by means g 

of endogenous NO. As a consequence of the I. pneumophila infection, a rapid £ 

nonspecific immune response is followed by a slower-developing specific im- g 

o 
mune response necessary for final eradication of the infection. CD4+ and g 

CD8+ T cells play a role in both phases. During the first phase, T cells might 2 

produce I FN and IL-6, whereas in the second phase they support humoral g 

immunity and specific T-cell-mediated immunity (Kaufmann, 1993; Susa § 

etal, 1998). 






% 

hi 

CJ 

S 

o 
HI 




infection of the two evolutionarily distant host cells (macrophages and proto- 
zoa) suggest that I. pneumophila may utilize similar molecular mechanisms 
to manipulate processes of these host cells (Gao et al., 1997). It has been 
hypothesized that L. pneumophila has evolved as a protozoan parasite in the 
environment and its adaptation to this primitive phagocytic unicellular host 
was sufficient to allow the bacteria to survive and replicate within the biologi- 
cally similar phagocytic cells of the more evolved mammalian host (Cianciotto 
and Fields, 1992; Abu Kwaik, 1996). 

The type IV secretion system is the key virulence factor of this organism, 
allowing it to invade the host cells, replicate, evade the endocytic pathway, 
induce apoptosis, and egress from the host cells (Gao and Abu Kwaik, 1999a, 
1999b, 2000; Alii et al, 2000; Hilbi et al, 2001; Watarai et al., 2001b; Molmeret 
et al., 2002a, 2002b; Zink et al, 2002). Not all the functions of the Dot/Icm 
* proteins have been understood, and no effectors other than RalF have yet 

g been identified. However, understanding these host-parasite interactions at 

§ the molecular level will be of considerable help in controlling the replication of 

< 

ph this bacterium, both environmentally in the protozoa and in human infection. 

in 

O 

g REFERENCES 

< 

Pi 

< Abu Kwaik, Y., Fields, B.S., and Engleberg, N.C. (1994). Protein expression by the 
" protozoan Hartmannella vermiformis upon contact with its bacterial parasite 

< Legionella pneumophila. Infect. Immun. 62, 1860-1866. 
z 

p Abu Kwaik, Y. (1996). The phagosome containing Legionella pneumophila within 

w the protozoan Hartmanella vermiformis is surrounded by the rough endoplas- 

tg mic reticulum. Appl. Environ. Microbiol. 62, 2022-2028. 

o Abu Kwaik, Y., Gao, L.-Y., Harb, O.S., and Stone, B.J. (1997). Transcriptional 

3 regulation of the macrophage-induced gene (gspA) of Legionella pneumophila 

< and phenotypic characterization of a null mutant. Mol. Microbiol. 24, 629- 

642. 

Abu Kwaik, Y. (1998a). Induced expression of the Legionella pneumophila gene en- 
coding a 20-kilodalton protein during intracellular infection. Infect. Immun. 
66, 203-212. 

Abu Kwaik, Y. (1998b). Fatal attraction of mammalian cells to Legionella pneu- 
mophila. Mol. Microbiol. 30, 689-696. 

Abu Kwaik, Y., Gao, L.-Y., Stone, B.J., Venkataraman, C, and Harb, O.S. (1998). 
Invasion of protozoa by Legionella pneumophila and its role in bacterial ecology 
and pathogenesis. Appl. Environ. Microbiol. 64, 3127-3133. 

Adeleke, A., Pruckler, J., Benson, R., Rowbotham, T., Halablab, M., and Fields, 
B.S. (1996). Legionella-\ike amoebal pathogens - phylogenetic status and pos- 
sible role in respiratory disease. Emerg. Infect. Dis. 2, 225-229. 



Alii, O.A.T., Gao, L.-Y., Pedersen, L.L., Zink, S., Radulic, M., Doric, M., and Abu 
Kwaik, Y. (2000). Temporal pore formation-mediated egress from macro- 
phages and alveolar epithelial cells by Legionella pneumophila. Infect. Immun. 
68, 6431-6440. 

Anderson, P. (1997). Kinase cascades regulating entry into apoptosis. Microbiol. 
Mol. Biol. Rev. 61, 33-46. 

Bachman, M.A. and Swanson, M.S. (2001). RpoS co-operates with other fac- 
tors to induce Legionella pneumophila virulence in the stationary phase. Mol. 
Microbiol. 40, 1201-1214. 

Barker, J., Brown, M.R.W., Collier, P.J., Farrell, I., and Gilbert, P. (1992). Re- 
lationships between Legionella pneumophila and Acanthamoebae polyphaga: 
physiological status and susceptibility to chemical inactivation. Appl. Environ. 
Microbiol. 58, 2420-2425. 

Barker, J., Lambert, P. A., and Brown, M.R.W. (1993). Influence of intra-amoebic § 

and other growth conditions on the surface properties of Legionella pneu- 
mophila. Infect. Immun. 61, 3503-3510. 

Barker, J., Scaife, H., and Brown, M.R.W. (1995). Intraphagocytic growth in- w 

duces an antibiotic-resistant phenotype of Legionella pneumophila. Antimi- ° 

crob. Agents Chemother. 39, 2684-2688. 3 

> 
Beckers, M.C., Yoshida, S., Morgan, K., Skamene, E., and Gros, P. (1995). Natural i 

resistance to infection with Legionella pneumophila: chromosomal localization g 

of the Lgnl susceptibility gene. Mamm. Genome 6, 540-545. £ 

Beckers, M.C., Ernst, E., Diez, E., Morissette, C., Gervais, F., Hunter, K., Hous- g 

o 
man, D., Yoshida, S., Skamene, E., and Gros, P. (1997). High-resolution £) 

linkage map of mouse chromosome 13 in the vicinity of the host resistance 2 

locus Lgnl. Genomics 39, 254-263. Jj 

Benin, A.L., Benson, R.F., and Besser, R.E. (2002). Trends in Legionnaires disease, § 



M 

M 
O 

a 



1980-1998: declining mortality and new patterns of diagnosis. Clin. Inf. Dis. C 

35, 1039-1046. 3 

Berger, K.H. and Isberg, R.R. (1993). Two distinct defects in intracellular growth q 

is; 

complemented by a single genetic locus in Legionella pneumophila. Mol. o 

Microbiol. 7, 7-19. 3 

Berger, K.H., Merriam, J., and Isberg, R.R. (1994). Altered intracellular targeting 

properties associated with mutations in the Legionella pneumophila dot A gene. 

Mol. Microbiol. 14, 809-822. 
Berk, S.G., Ting, R.S., Turner, G.W., and Ashburn, R.J. (1998). Production of 

respirable vesicles containing live Legionella pneumophila cells by two Acan- 

thamoeba spp. Appl. Environ. Microbiol. 64, 279-286. 
Birtles, R.J., Rowbotham, T.J., Raoult, D., and Harrison, T.G. (1996). Phylogenetic 

diversity of intra-amoebal legionellae as revealed by 16S rRNA gene sequence 

comparison. Microbiology 142, 3525-3530. 




Biurrun, A., Caballero, L., Pelaz, C, Leon, E., and Gago, A. (1999). Treatment of 
a Legionella pneumophila-colonized water distribution system using copper- 
silver ionization and continuous chlorination. Infect. Control Hosp. Epidemiol. 
20, 426-428. 

Bozue, J.A., and Johnson, W. (1996). Interaction of Legionella pneumophila with 
Acanthamoeba catellanii: uptake by coiling phagocytosis and inhibition of 
phago some -lyso some fusion. Infect. Immun. 64, 668-673. 

Brieland, J., Freeman, P., Kunkel, R., Chrisp, C, Hurley, M., Fantone, J., and 
Engleberg, N.C. (1994). Replicative Legionella pneumophila lung infection in 
intratracheally inoculated A/ J mice: a murine model of human Legionnaires' 
disease. Am. J. Pathol. 145, 1537-1546. 

Brieland, J.K., Remick, D.G., Freeman, P.T., Hurley, M.C., Fantone, J.C., and 

Engleberg, N.C. (1995). In vivo regulation of replicative Legionella pneu- 

% mophila lung infection by endogenous tumor necrosis factor alpha and nitric 

^ oxide. Infect. Immun. 63, 3253-3258. 

§ Brieland, J.K., Fantone, J.C., Remick, D.G., LeGendre, M., McClain, M., and 

< 

"^ Engleberg, N.C. (1997). The role of Legionella pneumophila-infected Hart- 

s manella vermiformis as an infectious particle in a murine model of Legion- 

p naires' disease. Infect. Immun. 65, 4892-4896. 

< Byrd, T.F. and Horwitz, M.A. (1989). Interferon gamma-activated human mono- 

< cytes downregulate transferrin receptors and inhibit the intracellular multi- 
" plication of Legionella pneumophila by limiting the availability of iron. J. Clin. 

< Invest. 83, 1457-1465. 
z 

p Byrd, T.F. and Horwitz, M.A. (2000). Aberrantly low transferrin receptor ex- 

w pression on human monocytes is associated with nonpermissiveness for 

tg Legionella pneumophila growth. J '. Infect. Dis. 181, 1394-1400. 

g Byrne, B. and Swanson, M.S. (1998). Expression of Legionella pneumophila viru- 

3 lence traits in response to growth conditions. Infect. Immun. 66, 3029-3034. 

< Cassatella, M.A. (1995). The production of cytokines by polymorphonuclear neu- 

trophils. Immunol. Today 16, 21-26. 

Christie, P.J. and Vogel, J. P. (2000). Bacterial type IV secretion: conjugation sys- 
tems adapted to deliver effector molecules to host cells. Trends Microbiol. 8, 
354-360. 

Cianciotto, N.P. and Fields, B.S. (1992). Legionella pneumophila mip gene potenti- 
ates intracellular infection of protozoa and human macrophages. Proc. Natl. 
Acad. Sci. USA 89, 5188-5191. 

Cirillo, J.D., Tompkins, L.S., and Falkow, S. (1994). Growth of Legionella pneu- 
mophila in Acanthamoeba castellanii enhances invasion. Infect. Immun. 62, 
3254-3261. 

Cirillo, J.D., Cirillo, S.L., Yan, L., Bermudez, L.E., Falkow, S., and Tompkins, 
L. S. (1999). Intracellular growth in Acanthamoeba castellanii affects monocyte 



M 

M 

o 

a 



entry mechanisms and enhances virulence of Legionella pneumophila. Infect. 
Immun. 67, 4427-4434. 

Cirillo, S.L., Lum, J., and Cirillo, J.D. (2000). Identification of novel loci involved 
in entry by Legionella pneumophila. Microbiology 146, 1345-1359. 

Coers, J., Monahan, C, and Roy, C.R. (1999). Modulation of phagosome biogene- 
sis by Legionella pneumophila creates an organelle permissive for intracellular 
growth. Nature Cell Biol. 1, 451-453. 

Coers, J., Kagan, J.C., Matthews, M., Nagai, H., Zuckman, D.M., and Roy, C.R. 
(2000). Identification of Icm protein complexes that play distinct roles in the 
biogenesis of an organelle permissive for Legionella pneumophila intracellular 
growth. Mol. Microbiol. 38, 719-736. 

Cohen, J.J. (1993). Mechanisms of apoptosis. Immunol. Today 14, 126-130. 

Deng, J.C., Tateda, K., Zeng, X., and Standiford, T.J. (2001). Transient transgenic 

expression of gamma interferon promotes Legionella pneumophila clearance § 

in immunocompetent hosts. Infect. Immun. 69, 6382-6390. 

Dietrich, W.F., Damron, D.M., Isberg, R.R., Lander, E.S., and Swanson, M.S. 

(1995). Lgnl, a gene that determines susceptibility to Legionella pneumophila, w 

maps to mouse chromosome 13. Genomics 26, 443-450. ° 

Dorn, B.R., Dunn, W.A. Jr., and Progulske-Fox, A. (2002). Bacterial interactions 3 

> 
with the autophagic pathway. Cell. Microbiol. 4, 1-10. i 

Dowling, J.N., Saha, A.K., and Glew, R.H. (1992). Virulence factors of the family g 

Legionellaceae. Microbiol. Rev. 56, 32-60. £ 

Dumenil, G. and Isberg, R.R. (2001). The Legionella pneumophila IcmR protein £ 

o 
exhibits chaperone activity for IcmQ by preventing its participation in high- £) 

molecular-weight complexes. Mol. Microbiol. 40, 1113-1127. 2 

Elliott, J. A. and Winn, W.C. Jr. (1986). Treatment of alveolar macrophages with Jj 

cytochalasin D inhibits uptake and subsequent growth of Legionella pneu- § 

mophila. Infect. Immun. 51, 31-36. C 

Enari, M., Sakahira, H., Yokoyama, H., Okawa, K., Iwamatsu, A., and Nagata, S. 5 

(1998). A caspase -activated DNase that degrades DNA during apoptosis, and q 

its inhibitor I CAD. Nature 391, 43-50. o 

Fields, B.S., Nerad, T.A., Sawyer, T.K., King, C.H., Barbaree, J.M., Martin, W.T., S 

Morrill, W.E., and Sanden, G.N. (1990). Characterization of an axenic strain 

of Hartmannella vermiformis obtained from an investigation of nosocomial 

legionellosis. J. Protozool. 37, 581-583. 
Fields, B.S. (1996). The molecular ecology of legionellae. Trends. Microbiol. 4, 

286-290. 
Fields, B.S., Benson, R.F., and Besser, R.E. (2002). Legionella and Legionnaires' 

disease: 25 years of investigation. Clin. Microbiol. Rev. 15, 506-526. 
Fliermans, C.B. (1996). Ecology of Legionella: from data to knowledge with a little 

wisdom. Microb. Ecol. 32, 203-228. 




Gagnon, E., Duclos, S., Rondeau, C, Chevet, E., Cameron, P.H., Steele-Mortimer, 
O., Paiement, J., Bergeron, J. J., and Desjardins, M. (2002). Endoplasmic 
reticulum-mediated phagocytosis is a mechanism of entry into macrophages. 
CellllO, 119-131. 

Gal-Mor, O., Zusman, T., and Segal, G. (2002). Analysis of DNA regulatory el- 
ements required for expression of the Legionella pneumophila icm and dot 
virulence genes. J. Bacteriol. 184, 3823-3833. 

Gao, L.-Y., Harb, O.S., and Abu Kwaik, Y. (1997). Utilization of similar mecha- 
nisms by Legionella pneumophila to parasitize two evolutionarily distant hosts, 
mammalian and protozoan cells. Infect. Immun. 65, 4738-4746. 

Gao, L.-Y., Harb, O.S., and Abu Kwaik, Y. (1998a). Identification of macrophage- 

specific infectivity loci (mil) of Legionella pneumophila that are not required 

for infectivity of protozoa. Infect. Immun. 66, 883-892. 

m Gao, L.-Y., Stone, B.J., Brieland, J.K., and Abu Kwaik, Y. (1998b). Different 

^ fates of Legionella pneumophila pmi and mil mutants within human-derived 

§ macrophages and alveolar epithelial cells. Microb. Pathog. 25, 291-306. 

< 

ph Gao, L.-Y. and Abu Kwaik, Y. (1999a). Activation of caspase-3 in Legionella 

s pneumophila-induced apoptosis in macrophages. Infect. Immun. 67, 4886- 

p 4894. 

< Gao, L.-Y. and Abu Kwaik, Y. (1999b). Apoptosis in macrophages and alveolar 

< epithelial cells during early stages of infection by Legionella pneumophila and 
" its role in cytopathogenicity. Infect. Immun. 67, 862-870. 

< Gao, L.-Y., Susa, M., Ticac, B., and Abu Kwaik, Y. (1999). Heterogeneity in in- 
z 

S tracellular replication and cytopathogenicity of Legionella pneumophila and 

m Legionella micdadei in mammalian and protozoan cells. Microb. Pathogen. 27, 

Eg 273-287. 

g Gao, L.-Y. and Abu Kwaik, Y. (2000). The mechanism of killing and exiting the 

3 protozoan host Acanthamoeba polyphaga by Legionella pneumophila. Environ. 

^ Microbiol. 2, 79-90. 

Gebran, S.J., Newton, C., Yamamoto, Y., Widen, R., Klein, T.W., and Friedman, 
H. (1994a). Macrophage permissiveness for Legionella pneumophila growth 
modulated by iron. Infect. Immun. 62, 564-568. 
Gebran, S.J., Yamamoto, Y., Newton, C, Klein, T.W., and Friedman, H. (1994b). 
Inhibition of Legionella pneumophila growth by gamma interferon in permis- 
sive A/J mouse macrophages: role of reactive oxygen species, nitric oxide, 
tryptophan, and iron(III). Infect. Immun. 62, 3197-3205. 
Gibson, F.C. Ill, Tzianabos, O.A., and Rodgers, F.G. (1993). Adherence of Le- 
gionella pneumophila to U-937 cells, guinea-pig alveolar macrophages, and 
MRC-5 cells by a novel, complement-independent binding mechanism. Can. 
J. Microbiol. 39, 718-722. 




M 

M 

o 

a 



Glavin, F.L., Winn, W.C., and Graighead, J.E. (1979). Ultrastructure of lung in 
Legionnaires' disease. Ann. Intern. Med. 90, 555-559. 

Hagele, S., Hacker, J., and Brand, B.C. (1998). Legionella pneumophila kills human 
phagocytes but not protozoan host cells by inducing apoptotic cell death. 
FEMS Microbiol. Lett. 169, 51-58. 

Hales, L.M. and Shuman, H.A. (1999). The Legionella pneumophila rpoS gene is 
required for growth within Acanthamoeba castellanii. J. Bacteriol. 181, 4879- 
4889. 

Hammer, B.K. and Swanson, M.S. (1999). Co-ordination of Legionella pneumophila 
virulence with entry into stationary phase by ppGpp. Mol. Microbiol. 33, 721- 
731. 

Hammer, B.K., Tateda, E.S., and Swanson, M.S. (2002). A two-component regula- 
tor induces the transmission phenotype of stationary-phase Legionella pneu- 
mophila. Mol. Microbiol. 44, 107-118. § 

Harb, O.S., Venkataraman, C, Haack, B.J., Gao, L.-Y., and Abu Kwaik, Y. (1998). 
Heterogeneity in the attachment and uptake mechanisms of the Legion- 
naires' disease bacterium, Legionella pneumophila, by protozoan hosts. Appl. w 

Environ. Microbiol. 64, 126-132. ° 

o 

Harb, O.S., Gao, L.-Y., and Abu Kwaik, Y. (2000). From protozoa to mammalian 3 

> 
cells: a new paradigm in the life cycle of intracellular bacterial pathogens. i 

Environ. Microbiol. 2, 251-265. g 

Heuner, K., Dietrich, C, Skriwan, C, Steinert, M., and Hacker, J. (2002). Influence £ 

of the alternative sigma(28) factor on virulence and flagellum expression of g 

o 
Legionella pneumophila. Infect. Immun. 70, 1604-1608. g 

Hilbi, H., Segal, G., and Shuman, H.A. (2001). I cm /dot-dependent upregulation 2 

of phagocytosis by Legionella pneumophila. Mol. Microbiol. 42, 603-617. % 

Horwitz, M.A. and Silverstein, S.C. (1980). Legionnaires' disease bacterium (Le- § 

gionella pneumophila) multiplies intracellularly in human monocytes. J. Clin. C 

Invest. 66, 441-450. 3 

Horwitz, M.A. and Silverstein, S.C. (1981a). Interaction of the Legionnaires' dis- q 

is; 

ease bacterium (Legionella pneumophila) with human phagocytes. II. Anti- o 

body promotes binding of L. pneumophila to monocytes but does not inhibit 
intracellular multiplication. J. Exp. Med. 153, 398-406. 

Horwitz, M.A. and Silverstein, S.C. (1981b). Interaction of the Legionnaires' 
disease bacterium (Legionella pneumophila) with human phagocytes. I. 
L. pneumophila resists killing by polymorphonuclear leukocytes, antibody, 
and complement. J. Exp. Med. 153, 386-397. 

Horwitz, M.A. (1983a). The Legionnaires' disease bacterium (Legionella pneu- 
mophila) inhibits phagosome-lysosome fusion in human monocytes. J. Exp. 
Med. 158, 2108-2126. 



m 




Horwitz, M.A. (1983b). Formation of a novel phagosome by the Legionnaires' 
disease bacterium (Legionella pneumophila) in human monocytes. J. Exp. 
Med. 158, 1319-1331. 
Horwitz, M.A. (1984). Phagocytosis of the Legionnaires' disease bacterium 
(Legionella pneumophila) occurs by a novel mechanism: engulfment within a 
pseudopod coil. Cell 36, 27-33. 
Horwitz, M.A. and Maxfield, F.R. (1984). Legionella pneumophila inhibits acidifi- 
cation of its phagosome in human monocytes. J. Cell Biol. 99, 1936-1943. 
Katz, S.M. and Hashemi, S. (1982). Electron microscopic examination of the 
inflammatory response to Legionella pneumophila in guinea pigs. Lab. Invest. 
46, 24-32. 
Katz, S.M. and Hashemi, S. (1983). Electron microscopic examination of the 
inflammatory response of guinea pig neutrophils and macrophages to Le- 
% gionella pneumophila. Adv. Exp. Med. Biol. 162, 327-333. 

^ Kaufmann, S.H.E. (1993). Immunity to intracellular bacteria. Annu. Rev. Im- 

g munol. 11, 129-163. 

g King, C.H., Fields, B.S., Shotts, E.B. Jr., and White, E.H. (1991). Effects of cytocha- 

& lasin D and methylamine on intracellular growth of Legionella pneumophila 

p in amoebae and human monocyte-like cells. Infect. Immun. 59, 758-763. 

< Kirby, J.E. and Isberg, R.R. (1998). Legionnaires' disease: the pore macrophage 

< and the legion of terror within. Trends Microbiol. 6, 256-258. 

™ Kirby, J.E., Vogel, J. P., Andrews, H.L., and Isberg, R.R. (1998). Evidence for pore- 

s' 

< forming ability by Legionella pneumophila. Mol. Microbiol. 27, 323-336. 
z 

p Komano, T., Yoshida, T., Narahara, K., and Furuya, N. (2000). The transfer region 

w of Incll plasmid R64: similarities between R64 tra and legionella icm/dot 

2 genes. Mol. Microbiol. 35, 1348-1359. 

g Kool, J.L., Carpenter, J.C., and Fields, B.S. (1999). Effect of monochloramine 

3 disinfection of municipal drinking water on risk of nosocomial Legionnaires' 

H-l 

< disease. Lancet 353, 272-277. 

Kusnetsov, J., Iivanainen, E., Elomaa, N., Zacheus, O., and Martikainen, P.J. 
(2001). Copper and silver ions more effective against legionellae than against 
mycobacteria in a hospital warm water system. Water Res. 35, 4217-4225. 

Mann, B.J., Torian, B.E., Vedvick, T.S., and Petri, W.A.J. (1991). Sequence of 
a cysteine-rich galactose-specific lectin of Entamoeba histolytica. Proc. Natl. 
Acad. Sci. USA 88, 3248-3252. 

Marrie, T.J., Raoult, D., La Scola, B., Birtles, R.J., and de Carolis, E. (2001). 
Legionella-\ike and other amoebal pathogens as agents of community- 
acquired pneumonia. Emerg. Infect. Dis. 7, 1026-1029. 

Matthews, M. and Roy, C.R. (2000). Identification and subcellular localization of 
the legionella pneumophila IcmX protein: a factor essential for establishment 



of a replicative organelle in eukaryotic host cells. Infect. Immun. 68, 3971- 
3982. 

Molmeret, M. and Abu Kwaik, Y. (2002). How does Legionella pneumophila exit 
the host cell? Trends Microbiol. 10, 258-260. 

Molmeret, M., Alii, O.A., Radulic, M., Susa, M., Doric, M., and Kwaik, Y.A. 
(2002a). The C-terminus of IcmT is essential for pore formation and for intra- 
cellular trafficking of Legionella pneumophila within Acanthamoebapolyphaga. 
Mol. Microbiol. 43, 1139-1150. 

Molmeret, M., Alii, O.A., Zink, S., Flieger, A., Cianciotto, N.P., and Kwaik, Y.A. 
(2002b). icmT is essential for pore formation-mediated egress of Legionella 
pneumophila from mammalian and protozoan cells. Infect. Immun. 70, 69- 
78. 

Muller, A., Hacker, J., and Brand, B. (1996). Evidence for apoptosis of human 

macro phage-like HL-60 cells by Legionella pneumophila infection. Infect. Im- § 

mun. 64, 4900-4906. g 

Muraca, P., Stout, J.E., and Yu, V.L. (1987). Comparative assessment of chlorine, £ 

heat, ozone, and UV light for killing Legionella pneumophila within a model w 

plumbing system. Appl. Environ. Microbiol. 53, 447-453. ° 

Nagai, H. and Roy, C.R. (2001). The DotA protein from Legionella pneumophila is S 

> 
secreted by a novel process that requires the Dot/Icm transporter. EMBO J. * 




o 
20, 5962-5970. £ 

Nagai, H., Kagan, J.C., Zhu, X., Kami, R.A., and Roy, C.R. (2002). A bacterial £ 

guanine nucleotide exchange factor activates ARF on Legionella phagosomes. g 

Science 295, 679-682. g 

Nash, T.W., Libby, D.M., and Horwitz, M.A. (1984). Interaction between the Le- S 

gionnaires' disease bacterium (Legionella pneumophila) and human alveolar Jj 

macrophages. Influence of antibody, lymphokines, and hydrocortisone. J. § 

Clin. Invest. 74, 771-782. tj 

Nash, T.W., Libby, D.M., and Horwitz, M.A. (1988). I FN -gamma-activated human 5 

alveolar macrophages inhibit the intracellular multiplication of Legionella q 

pneumophila.]. Immunol. 140, 3978-3981. o 

O'Brein, S.J. and Bhopal, R.S. (1993). Legionnaires' disease: the infective dose S 

paradox. Lancet 342, 5-6. 

Payne, N.R. and Horwitz, M.A. (1987). Phagocytosis of Legionella pneumophila 
is mediated by human monocyte complement receptors. J. Exp. Med. 166, 
1377-1389. 

Rechnitzer, C. and Blom, J. (1989). Engulfment of the Philadelphia strain of 
Legionella pneumophila within pseudopod coils in human phagocytes. Com- 
parison with the other Legionella strains and species. Acta Pathol. Microbiol. 
Immunol. Scand.[B]97, 105-114. 



Rodgers, F.G. and Gibson, F.C. Ill (1993). Opsonin-independent adherence and 

intracellular development of Legionella pneumophila within U-937 cells. Can. 

J. Microbiol. 39, 718-722. 
Rowbotham, T.J. (1980). Preliminary report on the pathogenicity of Legionella 

pneumophila for freshwater and soil amoebae. J. Clin. Pathol. 33, 1179-1183. 
Rowbotham, T.J. (1986). Current views on the relationships between amoebae, 

legionellae and man. Isr. J. Med. Sci. 22, 678-689. 
Roy, C.R. and Isberg, R.R. (1997). Topology of Legionella pneumophila DotA: an 

inner membrane protein required for replication in macrophages. Infect. 

Immun. 65, 571-578. 
Roy, C.R. and Tilney, L.G. (2002). The road less traveled: transport of Legionella 

to the endoplasmic reticulum. J. Cell Biol. 158, 415-419. 

Segal, G., Purcell, M., and Shuman, H.A. (1998). Host cell killing and bacterial 

% conjugation require overlapping sets of genes within a 22-kb region of the 

< 

^ Legionella pneumophila chromosome. Proc. Natl. Acad. Sci. USA 95, 1669- 

1674. 




< 

"^ Segal, G. and Shuman, H.A. (1998). Intracellular multiplication and human 



on 



s macrophage killing by Legionella pneumophila are inhibited by conjugal com- 

q ponents of IncQ plasmid RSF1010. Mol. Microbiol. 30, 197-208. 

< Segal, G., Russo, J. J., and Shuman, H.A. (1999). Relationships between a new 

< type IV secretion system and the icm/dot virulence system of Legionella pneu- 
5 mophila. Mol. Microbiol. 34, 799-809. 

< Segal, G. and Shuman, H.A. (1999). Legionella pneumophila utilizes the same 
z 

p genes to multiply within Acanthamoeba castellanii and human macrophages. 

^ Infect. Immun. 67, 2117-2124. 

tg Solomon, J.M. and Isberg, R.R. (2000). Growth of Legionella pneumophila in Dic- 

g tyostelium discoideum: a novel system for genetic analysis of host-pathogen 

3 interactions. Trends Microbiol. 8, 478-480. 

i-i 

< Solomon, J.M., Rupper, A., Cardelli, J.A., and Isberg, R.R. (2000). Intracellu- 

lar growth of Legionella pneumophila in Dictyostelium discoideum, a system 
for genetic analysis of host-pathogen interactions. Infect. Immun. 68, 2939- 
2947. 

Steinert, M., Emody, L., Amann, R., and Hacker, J. (1997). Resuscitation of 
viable but nonculturable Legionella pneumophila Philadelphia JR32 by Acan- 
thamoeba castellanii. Appl. Environ. Microbiol. 63, 2047-2053. 

Stone, B.J., Brier, A., and Kwaik, Y.A. (1999). The Legionella pneumophila prp locus; 
required during infection of macrophages and amoebae. Microb. Pathog. 27, 
369-376. 

Susa, M., Ticac, T., Rukavina, T., Doric, M., and Marre, R. (1998). Legionella 
pneumophila infection in intratracheally inoculated T cell depleted or non- 
depleted A/J mice. J. Immunol. 160, 316-321. 



Swanson, M.S. and Isberg, R.R. (1995a). Formation of the Legionella pneumophila 
replicative phagosome. Infect. Agents Dis. 2, 269-271. 

Swanson, M.S. and Isberg, R.R. (1995b). Association of Legionella pneumophila 
with the macrophage endoplasmic reticulum. Infect. Immun. 63, 3609-3620. 

Tateda, K., Moore, T.A., Deng, J.C., Newstead, M.W., Zeng, X., Matsukawa, 
A., Swanson, M.S., Yamaguchi, K., and Standiford, T.J. (2001a). Early re- 
cruitment of neutrophils determines subsequent T1/T2 host responses in a 
murine model of Legionella pneumophila pneumonia. J. Immunol. 166, 3355- 
3361. 

Tateda, K., Moore, T.A., Newstead, M.W., Tsai, W.C., Zeng, X., Deng, J.C., Chen, 
G., Reddy, R., Yamaguchi, K., and Standiford, T.J. (2001b). Chemokine- 
dependent neutrophil recruitment in a murine model of Legionella pneumo- 
nia: potential role of neutrophils as immunoregulatory cells. Infect. Immun. 
69, 2017-2024. 

Tilney, L.G., Harb, O.S., Connelly, P.S., Robinson, C.G., and Roy, C.R. (2001). 
How the parasitic bacterium Legionella pneumophila modifies its phagosome 



o 

M 

W 

o 

a 



and transforms it into rough ER: implications for conversion of plasma mem- w 



n 

o 



brane to the ER membrane. /. Cell Sci. 114, 4637-4650. S 



O 



Venkataraman, C, Haack, B.J., Bondada, S., and Abu Kwaik, Y. (1997). Identifi- 3 



> 



cation of a Gal/GalNAc lectin in the protozoan Hartmannella vermiformis as % 



o 



a potential receptor for attachment and invasion by the Legionnaires' disease g 

bacterium, Legionella pneumophila. J. Exp. Med. 186, 537-547. £ 

Venkataraman, C, Gao, L.-Y., Bondada, S., and Abu Kwaik, Y. (1998). Identifica- g 

o 
tion of putative cytoskeletal protein homologues in the protozoan Hartman- £) 

nella vermiformis as substrates for induced tyrosine phosphatase activity upon 2 

attachment to the Legionnaires' disease bacterium, Legionella pneumophila. % 

J. Exp. Med. 188, 505-514. § 

Vogel, J. P., Andrews, H.L., Wong, S.K., and Isberg, R.R. (1998). Conjugative S 

transfer by the virulence system of Legionella pneumophila. Science 279, 873- g 

876. q 

Watarai, M., Andrews, H.L., and Isberg, R.R. (2001a). Formation of a fibrous o 

structure on the surface of Legionella pneumophila associated with exposure S 

of DotH and DotO proteins after intracellular growth. Mol. Microbiol. 39, 
313-330. 

Watarai, M., Derre, I., Kirby, J., Growney, J.D., Dietrich, W.F., and Isberg, R.R. 
(2001b). Legionella pneumophila is internalized by a macropinocytotic uptake 
pathway controlled by the Dot/Icm system and the mouse Lgnl locus. J. Exp. 
Med. 194, 1081-1096. 

Weinbaum, D.L., Benner, R.R., Dowling, J.N., Alpern, A., Pasculle, A.W., and 
Donowitz, G.R. (1984). Interaction of Legionella micdadei with human mono- 
cytes. Infect. Immun. 46, 68-73. 




Winn, W.C. Jr. and Myerowitz, R.L. (1981). The pathology of the Legionella pneu- 
monias. A review of 74 cases and the literature. Hum. Pathol. 12, 401-422. 

Yamamoto, H., Ezaki, T., Ikedo, M., and Yabuuchi, E. (1991). Effects of biocidal 
treatments to inhibit the growth of legionellae and other microorganisms in 
cooling towers. Microbiol. Immunol. 35, 795-802. 

Yamamoto, Y., Klein, T.W., and Friedman, H. (1992). Genetic control of macro- 
phage susceptibility to infection by Legionella pneumophila. FEMS Microbiol. 
Immunol. 89, 137-146. 

Yamamoto, Y., Klein, T.W., and Friedman, H. (1996). Immunoregulatory role of 
nitric oxide in Legionella pneumophila-infected macrophages. Cell Immunol. 
171, 231-239. 

Zink, S.D., Pedersen, L, Cianciotto, N.P., and Abu-Kwaik, Y. (2002). The Dot/Icm 

type IV secretion system of Legionella pneumophila is essential for the indue - 

% tion of apoptosis in human macrophages. Infect. Immun. 70, 1657-1663. 

^ Zuckman, D.M., Hung, J.B., and Roy, C.R. (1999). Pore-forming activity is not 

§ sufficient for Legionella pneumophila phagosome trafficking and intracellular 

< 

g replication. Mol. Microbiol. 32, 990-1001. 

& Zusman, T., Gal-Mor, O., and Segal, G. (2002). Characterization of a Legionella 

p pneumophila relA insertion mutant and toles of RelA and RpoS in virulence 

< gene expression. J. Bacteriol. 184, 67-75. 

$ 

i— i 

pq 

i 

< 

% 
i— i 

Q 

H 

w 
Pi 
w 

i-i 
o 

w 

l-l 
1-1 

< 



CHAPTER 6 



Listeria monocytogenes invasion and 
intracellular growth 

Kendy K.Y. Wong and Nancy E. Freitag 



Studies focused on the Gram-positive, facultative intracellular bacterial path- 
ogen Listeria monocytogenes have provided valuable insights into many facets 
of biology, including cell-mediated immunity, cell physiology, and bacterial 
pathogenesis. The bacterium invades and replicates within a wide variety 
of cell types, and is capable of infecting an astonishing diversity of hosts, 
including mammals, fish, and insects (Gray and Killinger, 1966). The well- 
established use of murine and tissue culture models of infection, the ease of 
growing L. monocytogenes within the laboratory, and the existence of numer- 
ous genetic tools for the generation and analysis of bacterial mutants have 
helped to make L. monocytogenes a powerful model system for the exploration 
of the molecular basis of host-pathogen interactions. Because of its ubiquity 
within the environment and robust survival skills, this important foodborne 
pathogen remains a constant concern for public health departments and the 
food industry. 

Listeriae are noncapsulated, nonspore-forming, facultative anaerobic 
bacilli. They are 0.4 /xm by 1 to 1.5 /xm in size and are motile at 10° C to 25° C 
through the expression of polar flagellae (Farber and Peterkin, 1991; Lorber, 
1997; Vazquez- Boland et al., 2001b). There are six species in the Listeria 
genus: L. monocytogenes, I. ivanovii, L. innocua, L. seeligeri, L. welshimeri, and 
L. grayi (Collins et al., 1991; Sheehan et al., 1994). Only L. monocytogenes and 
I. ivanovii are considered pathogenic and are responsible for the disease 
known as listeriosis. L. ivanovii is mainly an animal pathogen, whereas 
L. monocytogenes has been reported to cause infections in humans, other ani- 
mals, birds, and fish (Gray and Killinger, 1966; McLauchlin, 1987; Alexander 
et al., 1992; Chand and Sadana, 1999; Ryser, 1999; Wesley, 1999). L. mono- 
cytogenes is the major listerial species pathogenic for humans; although rare, 
L. ivanovii has caused human infections, and at least one report of human 





listeriosis has been associated with L. seeligeri (Rocourt et al., 1986; Cummins 
et al., 1994; Lessing et al., 1994; Ramage et al., 1999). The Listeria spp. are 
widespread in nature and can be isolated from a variety of sources such as wa- 
ter, soil, food, and feces of animals (Schuchat et al., 1991; Fenlon, 1999). The 
natural habitat for Listeria organisms is believed to be decaying vegetation, but 
the two pathogenic species can invade and survive within a variety of hosts and 
host cell types. L. monocytogenes was first characterized in 1926 (Murray et al., 
1926), but it was not until the early 1980s that a clear link was established 
between listeriosis and the ingestion of food contaminated with the organ- 
ism (Schlech et al., 1983). Serious infections occur in immunocompromised 
individuals, neonates, the elderly, and pregnant women, usually through 
the ingestion of contaminated food such as dairy products and ready-to-eat 
deli meats (Gray and Killinger, 1966; Farber and Peterkin, 1991; Rocourt, 
^ 1996; Ryser, 1999). The infection of healthy individuals has been reported to 

w cause gastroenteritis, a condition thought to be underdocumented because 

* routine clinical testings do not search for listeriae (Hof, 2001). Although 

g I. monocytogenes does not usually pose a high risk for healthy individuals, 

< the mean mortality rate of human listeriosis is 20-30% (Farber and Peterkin, 

g 1991; Schuchat et al., 1991; Rocourt, 1996). The ability of I. monocytogenes to 

tolerate high salt concentrations and low pH' and to multiply at refrigerated 

o temperatures makes it a difficult and constant challenge for the food industry 

p\ (Lammerding and Doyle, 1990; Lou and Yousef, 1999). This chapter briefly 

>* reviews the current knowledge regarding L. monocytogenes invasion of host 

§ cells and the resulting course of infection following host cell entry. 



CLINICAL PRESENTATIONS OF LISTERIOSIS: AN OVERVIEW OF 
THE DIVERSITY OF CELL TYPES AND TISSUES THAT SUPPORT 

BACTERIAL REPLICATION 

Among the first cells encountered by I. monocytogenes following inges- 
tion of the bacterium by the host are the epithelial cells lining the gut. In 
the mouse, L. monocytogenes has been observed to translocate the intestine 
without causing gross lesions, and rapid listerial translocation occurs in rat 
ileal loop models of intestinal infection (Marco et al., 1992; Pron et al., 1998). 
On the basis of these findings, it has been suggested that bacterial replica- 
tion in the intestinal mucosa is probably not required for the establishment 
of systemic infection, although L. monocytogenes is able to proliferate at this 
location. I. monocytogenes can cross the intestinal barrier by means of M 
cells of Peyer's patches, and the Peyer's patches appear to be preferential 




multiplication sites for the bacteria, which can then invade neighboring cells 
by direct cell-to-cell spread (Havell et al., 1999; Daniels et al., 2000; Hof, 
2001). Dendritic cells are also early listerial targets in Peyer's patches (Pron 
et al., 2001). Gross intestinal lesions and gastroenteritis appear to occur only 
following the ingestion of large numbers of bacteria. The risk for gastroen- 
teritis also appears to increase in the presence of small amounts of mucosal 
damage within the intestine; such damage may happen periodically in hu- 
mans (MacDonald and Carter, 1980; Pron et al, 1998; Daniels et al, 2000). 
Once across the intestinal barrier, L. monocytogenes is translocated to the liver 
and spleen within minutes, where most of the organisms are then cleared 
(Conlan and North, 1991; Pron et al., 1998; Cousens and Wing, 2000; Daniels 
et al., 2000). In susceptible hosts, surviving L. monocytogenes multiplies, 
mainly in hepatocytes, and spreads through the liver parenchyma. Spleno- 
cytes are also target cells for L. monocytogenes, although they appear less 
permissive for bacterial replication than hepatocytes. In both organs, mi- g 

croabscess formation is observed and neutrophils are rapidly recruited to the g 

infection foci; however, L. monocytogenes-infected spleen cells appear to be § 

o 
less susceptible to lysis by neutrophils during early infection (Conlan and Q 

North, 1994; Rogers et al., 1996). L. monocytogenes has been demonstrated to § 

induce apoptosis in hepatocytes and Caco-2 cells (Rogers et al., 1996; Valenti ; - 

et al., 1999), and in dendritic cells apoptosis was recently shown to be trig- g 

gered by listeriolysin O (LLO; see Guzman et al., 1995). If I. monocytogenes > 

infection is not resolved within the liver and spleen, the bacteria enter into g 

circulation and disseminate to other tissues. £j 

L. monocytogenes appears to have a tropism for the placenta of pregnant 

women and for the central nervous system (CNS). The organism has the 3 

ability to cross both the feto-placental barrier and the blood-brain barrier S 

(Armstrong and Fung, 1993; Lorber, 1997; Greiffenberg et al., 1998; Parkassh g 

et al., 1998) . At the placenta, L. monocytogenes causes the formation of numer- » 

o 

ous microabscesses and the necrosis of placental villi (Parkassh et al., 1998; g 

Vazquez- Boland et al., 2001b). L. monocytogenes infects the developing fe- h 

tus and causes chorioamnionitis, which often leads to abortion or stillbirth. 
Alternatively, fetal listeriosis may lead to the birth of a baby with granu- 
lomatosis infantiseptica, characterized by microabscesses all over the body 
and internal organs (Klatt et al., 1986; Schuchat et al., 1991; Lorber, 1997). 
The mother may be asymptomatic or may only have mild flu-like symp- 
toms, and CNS infection rarely occurs in pregnant women. For the establish- 
ment of CNS infections, active actin-based intra-axonal movement may facil- 
itate L. monocytogenes spread (Otter and Blakemore, 1989; Antal et al., 2001). 



H 




Listeria-infected phagocytes and dendritic cells have also been shown to play a 
role in bacterial dissemination and in the initiation of CNS infection (Drevets 
et al., 1995, 2001; Guzman et al, 1995; Drevets, 1999; Pron et al, 2001). The 
growth of L. monocytogenes in brain tissues appears to be unrestricted, and the 
bacterium has been found to infect the epithelial cells of the choroids plexus, 
ependymal cells, microglial cells and neurons, and brain parenchymal tissue 
(Kirk, 1993; Schluter et al., 1999). 

In nonpregnant individuals, meningitis and encephalitis are common 
manifestations of I. monocytogenes disease. Clinical features often seen with 
listeriosis include fever, headache, nausea, vomiting, movement disorders, 
altered consciousness, and seizures. Bacteremia is also frequently observed in 
adult listeriosis (Lorber, 1997; Vazquez- Boland et al., 2001b). Some unusual 
forms of listerial infections include endocarditis, hepatitis, myocarditis, ar- 
^ teritis, pneumonia, sinusitis, conjunctivitis, ophthalmitis, joint infection, and 

w skin infection (Vazquez- Boland et al., 2001b and references therein). Febrile 

* gastroenteritis has also been described in recent years (Salamina et al., 1996; 

g Dalton et al., 1997; Miettinen et al, 1999; Aureli et al, 2000). This wide spec- 

< trum of I. monocytogenes disease manifestations reflects the diversity of cell 

g types and tissues that can harbor the bacterium. 

As L. monocytogenes can invade many types of cells and tissues within 

o a host, a number of tissue culture cell model systems have been developed 

^ to study bacteria-host cell interactions, including epithelial cells, such as 

!* Caco-2, Henle 407, Vero, HeLa, Chinese hamster ovary (CHO) cells, and 

§ potoroo rat kidney (PtK2) cells; endothelial cells, such as human umbilical 

vein endothelial cells (HUVEC) and human brain microvascular endothelial 

cells (HBMEC); fibroblast cells; macrophages (primary and macrophage cell 

lines such as J774); hepatocytes; and dendritic cells (Gaillard et al., 1987; 

Sun et al., 1990; Dramsi et al., 1995; Drevets et al., 1995; Guzman et al., 

1995; Mengaud et al, 1996; Greiffenberg et al, 1998; Parida et al, 1998; 

Kolb-Maurer et al., 2000; Suarez et al., 2001). The mechanisms governing 

the invasion of these various cell types by L. monocytogenes are the focus of 

the following section. 



INVASION OF HOST CELLS BY L MONOCYTOGENES 

The intracellular life cycle of I. monocytogenes has been extensively stud- 
ied. The overall process resulting in intracellular bacterial replication and cell- 
to-cell spread includes the following: (1) bacterial adherence or attachment 
to the host cells; (2) bacterial internalization by phagocytic cells, or bacterial 
induced-uptake into nonprofessional phagocytic cells; (3) bacterial escape 



Entry 

(inlA, inlB: Internalins InlA, InlB) 



Vacuole lysis f * 

(lily: Listeriolysin O/LLO, J\\ 
pic A: PlcA/PI-PLC)y^>. 

i 




Lysis of two-membrane 
vacuole 

(plcB: Lecithinase 
PlcB/PC-PLC) 



Intracellular v 

movement 

(act A: ActA) 



r; , x Cell to cell 
spread 

(act A : ActA) 



Figure 6.1. Overview of I. monocytogenes infection of host cells. The stippled regions 
represent host actin filaments. The major gene products that contribute to intracellular 
growth and cell-to-cell spread are indicated. Adapted from Tilney and Portnoy (1989) with 
permission from the Journal of Cell Biology. 



from the phagosome; (4) bacterial multiplication and actin polymerization- 
based movement within the host cell cytosol; and (5) spread of the bacte- 
ria to adjacent cells and escape from a double-membrane-bound vacuole 
(Fig. 6.1). 

Adherence and internalization mechanisms for nonprofessional 
phagocytic cells 

The first steps in I. monocytogenes infection following bacterial inges- 
tion are bacterial adherence and invasion of the gastrointestinal epithelium. 
Following adherence, the internalization of I. monocytogenes proceeds by 
means of a zipper-like mechanism in which the plasma membrane of the 
host cell closely enwraps the bacterium until it is completely engulfed into 
a vacoule. This process is strikingly different from the trigger mechanism 
induced by Salmonella and Shigella spp., which leads to the formation of dra- 
matic membrane ruffles in host cells during uptake (Swanson and Baer, 
1995; Dramsi and Cossart, 1998; Kuhn and Goebel, 2000). I. monocyto- 
genes has been found to enter polarized cells through the basolateral surface, 
and its internalization is increased following the disruption of intercellular 
junctions in the host cells (Gaillard and Finlay, 1996). Of the various liste- 
ria! ligands described to participate in bacterial attachment and invasion of 




3 

to 

> 

o 

% 
o 
n 

o 

hi 
% 

z 

$ 

o 

> 
o 
z 

H 

n 

M 
M 
M 

a 



to 

o 
to 
o 



ffi 




nonprofessional phagocytic host cells, the internalins (encoded by the inl 
genes) appear to be the major players and have been the best described. 

The inlAB gene products 

The inlAB locus was identified following the isolation of bacterial 
transposon-insertion mutants that were defective for invasion of Caco-2 cells 
(Gaillard et al., 1991). L. monocytogenes inlAB mutants were found to be 
severely impaired in host cell invasion, and the expression of InlA and InlB 
on the surface of latex beads or on the surface of normally noninvasive bac- 
teria (such as L. innocua and Enterococcus faecalis) was sufficient to induce 
internalization (Lecuit et al., 1997; Braun et al., 1998). 

I. monocytogenes InlA promotes entry into human intestinal epithelial 
SJ cell lines such as Caco-2 cells and into other cells expressing the E-cadherin 

n receptor ligand (Mengaud et al., 1996). E-cadherin is a 110-kDa transmem- 

* brane glycoprotein specific to epithelial tissues, such as those in the diges- 

g tive tract and in the cells lining the choroid plexus and placental chorionic 

< villi. The protein is predominantly expressed at the adherens junctions and 

g on the basolateral face of cells, where it mediates calcium-dependent cell- 

cell adhesion and helps maintain tissue cohesion. It has been suggested that 

o I. monocytogenes breaches the intestinal barrier through M cells at the Peyer's 

p\ patch surface to gain access to the basolaterally located E-cadherins. Alter- 

>* natively, the E-cadherin receptor might be transiently accessible when crypt 

§ cells migrate to the tips of intestinal villi for exfoliation (Mengaud et al., 1996; 

Vazquez- Boland et al., 2001b; Cossart, 2002). The cytoplasmic domain of E- 
cadherin interacts with proteins known as catenins, and these associations 
ultimately stimulate the cytoskeletal rearrangements within the host cell that 
result in L. monocytogenes uptake (Cossart and Lecuit, 1998; Kuhn and Goebel, 
1998; Vazquez-Boland et al., 2001b). 

InlA is an approximately 80-kDa protein that is targeted to the bacterial 
surface by an N-terminal signal sequence. The InlA C-terminus contains a 
hydrophobic stretch of 20 amino acids preceded by a cell wall anchor motif, 
LPXTG (Fig. 6.2). This motif is a signature of many Gram-positive surface pro- 
teins and is necessary for their covalent linkage to the peptidoglycan (Fischetti 
et al., 1990). A putative transamidase known as sortase, encoded by srtA, 
was recently identified and shown to be required for the cell wall anchor- 
ing of InlA (Bierne et al., 2002). Moreover, L. monocytogenes mutants lack- 
ing functional sortase were shown to be less invasive in vitro and atten- 
uated in a mouse model of infection (Garandeau et al., 2002). Similar 
to other members of the internalin family, the N-terminus of InlA has a 
series of leucine-rich repeats (LRRs) consisting of a string of 22 amino 



« 



signal 

peptide 



InlA 
(800 aa) 



i 



LPTTG 



15 LRRs 






E-Cadherin 
binding 



Cell wall 
anchor 



signal 
peptide 



InlB 
(630 aa) 



i 



8 LRRs 



GW GW GW 



gClq-R, 

MET binding 



Cell surface 
anchor 



Figure 6.2. Schematic illustration of InlA and InlB. LRR represents leucine-rich repeats, 
and GW represents the GW repeats present in InlB that anchor the protein to the bacterial 
cell surface. The LPTTG motif is involved in cell wall anchoring of InlA. The receptors for 
InlA and InlB are indicated below the LRR binding regions. 



acids containing leucine or isoleucine residues at seven fixed positions (at 
residues 3, 6, 9, 11, 16, 19, and 22; see Dramsi et al., 1997; Engelbrecht et al., 
1998a, 1998b; Raffelsbauer et al, 1998). There are 15 successive LRR units 
in InlA; each LRR is composed of a beta strand that alternates with a more 
flexible antiparallel helix, and the repeats are connected by coils. An intact 
LRR region is necessary and sufficient to induce bacterial entry, whereas the 
inter-repeat region has been suggested to play a role in the proper folding 
and stabilization of the LRRs (Lecuit et al., 1997). The InlA protein is thought 
to interact by means of the LRR regions with the extracellular domain of 
E-cadherin, whose cytoplasmic domain then leads to rearrangement of the 
host cell actin cytoskeleton for bacterial internalization. The proline residue 
at position 16 of E-cadherin has been shown to be critical for InlA interac- 
tion. Interestingly, the mouse and rat E-cadherins have glycine in place of 
proline at this position and do not facilitate InlA-dependent internalization 
of I. monocytogenes. The position 16 proline of E-cadherin thus contributes 
some measure of host specificity for I. monocytogenes (Lecuit et al., 1999). 
InlA has also been shown to contribute to the attachment and internalization 
of L. monocytogenes by macrophages (Sawyer et al., 1996). 

InlB is another L. monocytogenes surface protein that mediates bacterial 
invasion of a broader range of host cells, including CHO cells, Vero cells, hep- 
atocytes, and several endothelial, epithelial, and fibroblast cell lines (Dramsi 
et al., 1995; Lingnau et al, 1995; Ireton et al, 1996; Greiffenberg et al., 1998; 



3 

to 

> 

o 

% 
o 
n 

o 

hi 
% 

z 

$ 

o 

> 
o 
z 

H 

n 

M 
M 
M 

a 



to 

o 
to 
o 



ffi 




Parida et al., 1998). Two host receptors have been identified for InlB. One, 
gClq-R, is expressed in most tissues and is the receptor for the globular 
part of the complement component Clq (Braun et al., 2000; Galan, 2000). 
The gClq-R protein has no transmembrane domain and no cytoplasmic tail; 
thus a co-receptor may be required to mediate the effects of InlB-gClq-R 
interactions. The other receptor for InlB is Met, a 145-kDa tyrosine kinase 
that is also a high-affinity receptor for hepatocyte growth factor (Shen et al., 
2000). InlB-mediated listerial entry into host cells leads to the stimulation 
of phosphoinositide 3-kinase (PI-3 kinase; see Ireton et al., 1996). In ad- 
diton, the MEK-1, ERK-1, and ERK-2 protein kinases are activated during 
I. monocytogenes invasion of HeLa epithelial cells (Tang et al., 1998). 

InlB is a 71-kDa member of the internalin family and has a signal se- 
quence followed by 8 LRRs at the N-terminus. However, InlB does not possess 
^ the Gram-positive surface anchor LPXTG motif at its C-terminus. Instead, 

w a "GW" motif, containing repeats led by the sequence GW, functions as the 

* cell surface anchor domain (Fig. 6.2; also see Braun et al., 1997). InlB is 

g only loosely associated with lipoteichoic acid in the cell wall by means of the 

< GW motif, and some InlB is released into the supernatant (Jonquieres et al., 

g 1999). The GW motif also binds cellular glycosaminoglycans, and this inter- 

action appears to be required for efficient InlB-mediated bacterial invasion by 

o enhancing the internalization triggered by the N-terminal LRRs (Jonquieres 

>; etal.,2001). 

>* Besides inlAB, six additional genes encoding putative surface-anchored 

§ internalins have been identified in L. monocytogenes. Similar to InlA, these six 

members (InlC2, InlD, InlE, InlF, InlG, and InlH) all have signal sequences 
and various numbers of LRRs at the N-terminal region, plus the LPXTG 
motifs at the C-terminus (Dramsi et al., 1997; Engelbrecht et al., 1998a; 
Raffelsbauer et al., 1998). These internalins were recently shown to enhance 
InlA-mediated invasion of I. monocytogenes into Caco-2 cells and into HB- 
MEC. However, InlB-triggered bacterial internalization into these cell lines 
did not require any of the other internalin proteins (Bergmann et al., 2002). 
InlC is an unusual member of the internalin protein family. It is small 
(~30 kDa) and is secreted by L. monocytogenes instead of being anchored 
to the cell wall. The potential role of InlC in L. monocytogenes pathogenesis 
is still unclear. It has been suggested to be involved in the dissemination 
of L. monocytogenes infection; however, cell-to-cell spread is not affected in 
in/C mutant strains (Engelbrecht et al., 1996; Lingnau et al., 1996; Domann 
et al., 1997). Southern blot and genome analysis has indicated the presence 
of internalin homologues in the nonpathogenic L. innocua, and it is possible 
that these homologues might have functions that are distinct from those 
required for invasion (Gaillard et al., 1991; Glaser et al., 2001). 



Other possible L monocytogenes adherence and 
internalization factors 

In addition to the internalin protein family, a number of other L. monocy- 
togenes surface proteins have been implicated in bacterial adherence and inva- 
sion of host cells (Kuhn and Goebel, 1989; Alvarez- Dominguez et al., 1997b). 
The L. monocytogenes surface protein ActA, which mediates the polymeriza- 
tion of host actin filaments required for bacterial motility within the host 
cytosol, has been shown to recognize host cell heparan sulfate proteoglycans 
(HSPG) and may contribute to epithelial cell invasion (Alvarez- Dominguez 
et al., 1997b; Suarez et al., 2001). A major extracellular protein, p60 (en- 
coded by lap) , has been reported to be required for bacterial attachment and 
invasion of mouse fibroblasts. However, it is possible that the decrease in 
invasion of I. monocytogenes p60 mutants was due to septation defects asso- H 

ciated with the disruption of the p60 murein hydrolase activity (Bubert et al., h 

1992; Wuenscher et al., 1993). Bacterial adherence to Caco-2 cells seems to be £ 

independent of p60 and rather mediated by a 104-kDa surface protein known § 

as the Listeria adhesion protein (Lap; see Pandiripally et al., 1999). There is o 

also a fibronectin-binding protein that has been identified in L. monocytogenes g 

(Gilotetal, 1999, 2000). £ 

The completion of the I. monocytogenes genome sequence has revealed £ 

that, among the bacterial genomes currently sequenced, L. monocytogenes con- § 

tains the largest number of proteins with the cell wall anchor motif LPXTG 
(Glaser et al., 2001). It is possible that some of these putative surface proteins 



L monocytogenes entry into professional phagocytes 

Macrophages actively engulf I. monocytogenes, and several host cell com- 
ponents have been shown to participate in macrophage binding and uptake 



> 

on 



M 
M 



o 

> 
are involved in adherence and internalization. In addition to protein ligands, o 

other surface components, such as lipoteichoic acid, appear to be impor- * 

tant for L. monocytogenes adhesion to various cell lines. The D-alanylation of £ 

the lipoteichoic acids, catalyzed by products expressed from the dlt operon, 

was recently shown to be required for adhesion to murine macrophages and 

hepatocytes as well as human Caco-2 epithelial cells (Abachin et al., 2002). A o 

virulent I. monocytogenes strain possessing surface alpha-D-galactose residues g 

was shown to bind through lectin-like interactions to a human hepatocarci- ffi 

noma cell line, which expresses a well-characterized carbohydrate receptor 

for alpha-D-galactose at the surface (Cowart et al., 1990). Agglutination by 

lectins suggests the presence of other carbohydrate-binding ligands on the 

surface of I. monocytogenes (Cottin et al., 1990; Facinelli et al., 1998). 




of the bacterium. The complement component C3b is involved in mediating 
listerial interactions with macrophages by means of CR3, the complement 
receptor type 3 (Drevets and Campbell, 1991; Croize et al., 1993; Drevets 
et al., 1993). I. monocytogenes also binds Clq in a specific, saturable, and 
dose-dependent manner. Enhanced uptake of Clq-bound L. monocytogenes 
was mediated by Clq complement receptors expressed on the macrophage 
surface (Alvarez- Dominguez et al., 1993). Nonopsonic interactions may also 
be involved, as L. monocytogenes uptake is efficient in the absence of serum 
(Pierce et al., 1996). The type I and type II macrophage scavenger receptors 
can recognize and bind a wide range of polyanions, and these receptors have 
been shown to bind listerial lipoteichoic acids (Dunne et al., 1994; Greenberg 
et al., 1996; Ishiguro et al., 2001). N-acetylneuraminic acid (NAcNeu) has 
also been suggested to play a role in the attachment of I. monocytogenes to 
o murine macrophages (Maganti et al., 1998). Host nuclear factor /cB (NF-/cB) 

w is activated when I. monocytogenes adheres to the surface of macrophage -like 

* cell lines; long-lasting NF-/cB activation appears to require LLO and products 

g of the lecithinase operon (mpl-actA-plcB; see Hauf et al., 1994). 

< 

| LIFE AFTER INVASION: THE PHAGOSOME IS NO PLACE FOR 

% L MONOCYTOGENES TO LIVE 

o 

pi Inside activated macrophages, most of the internalized I. monocytogenes 

p* are killed within minutes after uptake (Davies, 1983; Raybourne and Bunning, 

§ 1994). Nevertheless, L. monocytogenes appears well adapted to establish and 

maintain a niche inside various types of nonbactericidal host cells. L. mono- 
cytogenes mediates escape from the host cell phagosome, multiplies within 
the cytosol, and moves to infect adjacent cells by means of the expression of 
a number of protein products that function together to enable the bacterium 
to exploit the host cell environment. 



L monocytogenes escape from the phagosome 

Host cell invasion ultimately provides access to a rich growth medium 
for L. monocytogenes: the host cytosol. Following host cell uptake of the bac- 
terium, L. monocytogenes is enclosed in a primary phagosomal vacuole that 
becomes quickly acidified. L. monocytogenes prevents further maturation of 
this phagosome and enhances its own survival by escaping into the cytosol 
after disruption of the phagosomal membrane (Alvarez- Dominguez et al., 
1997a). In mouse bone marrow-derived macrophages, bacterial-mediated 
perforation of the phagosome membrane is detectable within minutes 




(Beauregard et al., 1997). Complete disruption of the phagosomal membrane 
within infected Caco-2 cells is observed in approximately 30 minutes (Gaillard 
et al., 1987). The major factor contributing to phagosomal escape is LLO, first 
described following the isolation of bacterial transposon-insertion mutants 
that had a nonhemolytic phenotype on sheep blood agar plates (Gaillard et al., 
1986; Kathariou et al., 1987). LLO, encoded by hly, was the initial virulence 
factor to be identified as essential for I. monocytogenes virulence (Gaillard 
et al, 1986; Geoffroy et al, 1987; Kathariou et al, 1987; Mengaud et al, 1988; 
Cossart et al., 1989); bacterial mutants lacking LLO do not mediate phagoso- 
mal escape in most cell lines examined and are highly attenuated in mouse 
models of infection (Portnoy et al., 1992). The production of LLO by the non- 
pathogenic soil bacterium Bacillus subtilis is sufficient to mediate bacterial 
escape from macrophage phagosomes following uptake of B. subtilis into 
tissue culture cells (Bielecki et al., 1990). 

LLO is a 58-kDa cytolysin belonging to a large family of cholesterol- g 

dependent, pore-forming toxins that includes streptolysin O (SLO) and per- g 

fringolysin O (PFO), secreted by Streptococcus pyogenes and Clostridium per- § 

o 
fringens, respectively (Palmer, 2001). Similar to other family members, LLO Q 

has a conserved tryptophan-rich undecapeptide (ECTGLAWEWWR) at its C- § 

terminus. This region appears to be important for the cholesterol-dependent g 

membrane binding of these cytolysins (Iwamoto et al., 1990; Boulnois et al., g 

1991; Sekino-Suzuki et al., 1996; Palmer, 2001). LLO and its family mem- > 

bers mediate membrane disruption through the formation of oligomeric g 

ring-shaped structures that form membrane pores that are approximately £j 

20-30 nm in diameter (Palmer, 2001). The crystal structure of PFO has been 

solved, revealing four protein domains; this structure has provided valuable 3 

information regarding structure-function relationships within this family of S 

proteins (Rossjohn et al., 1997). A recent study demonstrated that two vari- g 

ants of LLO, comprising domains 1-3 or domain 4 alone, could be secreted » 



H 



o 



efficiently and reassembled into a functionally active protein when the do- g 

mains were expressed simultaneously (Dubail et al., 2001). Another study h 

demonstrated that domains 1-3 of LLO were essential for cytokine induc- 
tion, whereas domain 4 was important for binding to membrane cholesterol 
(Kohdaetal., 2002). 

LLO is unique among the family members because of its acidic pH 
optimum (Geoffroy et al., 1987). LLO is active at pH 4.5 to 6.5, an ideal range 
for the acidic phagocytic vacuole, the pH of which is approximately 5. The 
cytolysin PFO, which does not have an acidic pH optimum, is toxic to host 
cells when expressed by I. monocytogenes Ahly mutants, possibly as a result 
of PFO-mediated host cell plasma membrane damage (Jones and Portnoy, 




1994). Mutations that result in single amino acid substitutions within PFO 
can apparently convert the protein into a vacuole-specific hemolysin similar 
to LLO, but L. monocytogenes strains expressing these PFO variant molecules 
are still highly attenuated in mouse models of infection (Jones et al., 1996). A 
single leucine located at position 461 within LLO has been recently identified 
to confer the acidic pH activity (Glomski et al., 2002). The substitution of 
leucine 461 with a threonine found at a similar position within PFO increased 
the activity of LLO at a neutral pH. I. monocytogenes mutants expressing 
the hly L461T variant were cytotoxic to infected host cells (Glomski et al., 
2002). In addition, a PEST-like sequence in the N-terminus of LLO has been 
suggested to target LLO for rapid degradation within the host cytosol (Decatur 
and Portnoy, 2000). L. monocytogenes mutants producing LLO molecules that 
lack the PEST sequence accumulate the protein within the host cytosol, and 
u the mutants are extremely cytotoxic for host cells and highly attenuated for 

w virulence in mice. Moreover, the addition of the PEST sequence to PFO 

* can convert it into a protein variant that is nontoxic to host cells (Decatur 

g and Portnoy, 2000) . An independent study has shown that L. monocytogenes 

< strains that express an LLO variant with substitutions for all of the P, E, S, and 

g T residues is also highly attenuated for virulence in mice (Lety et al., 2001). It 

thus appears that multiple mechanisms exist to restrict LLO activity within 

o the host cytosol, thereby compartmentalizing the activity of the protein to the 

p\ phagosome and preventing LLO-mediated damage to the host cell plasma 

p* membrane. The existence of multiple mechanisms that serve to regulate LLO 

§ activity highlights how extraordinarily suited L. monocytogenes has become for 

life within the host cytosol. 

As previously discussed, LLO plays an important role in mediating the 
escape of I. monocytogenes from the phagosome of many cell types; however, 
its activity is not required for phagosome disruption within all cells. Non- 
hemolytic I. monocytogenes mutants lacking functional LLO are still capable 
of replicating within the cytosol of Henle 407 human epithelial cells, which 
suggests that other factors can contribute to phagosomal membrane disrup- 
tion (Portnoy et al., 1988). I. monocytogenes encodes two phospholipases: a 
33-kDa phosphatidylinositol-specific phospholipase C (PI-PLC) that is en- 
coded by plcA, and a broad-specificity phospholipase that is encoded by plcB 
(Camilli et al., 1991; Geoffroy et al., 1991; Leimeister-Wachter et al., 1991; 
Mengaud et al., 1991; Vazquez- Boland et al., 1992). 

PI-PLC can hydrolyze moieties that anchor many eukaryotic membrane 
proteins to the plasma membrane, and the secreted enzyme has an acidic 
pH optimum (Goldfine and Knob, 1992). PI-PLC is thought to be active in 
the phagocytic vacuole and to mediate lysis of the vacuolar membrane in 
conjunction with LLO (Goldfine and Knob, 1992). It has been suggested that 



the initial pores formed by LLO provide access to PI-PLC membrane sub- 
strates. Cooperation between the LLO hemolysin and the phospholipases was 
also suggested to increase the membrane affinity of LLO, thereby facilitating 
membrane lysis (Goldfine et al., 1995; Sibelius et al., 1996; Stachowiak and 
Bielecki, 2001). Recently, the combined activities of LLO and PI-PLC were 
shown to affect bacterial entry and vacuolar escape in J 774 cells by influenc- 
ing the translocation of protein kinase C and calcium signaling (Wadsworth 
and Goldfine, 2002). 

The 29-kDa lecithinase or phosphatidylcholine-phospholipase C (PC- 
PLC) encoded by plcB is another factor that contributes to the escape of 
L. monocytogenes from cell phagosomes. This enzyme has a wide pH optima 
and a broad substrate spectrum (Geoffroy et al., 1991; Goldfine et al., 1993). 
PlcB is synthesized as an inactive proenzyme pro- PlcB, a regulatory step that 
has probably evolved to prevent PlcB -mediated bacterial membrane damage. 
The N-terminal propeptide of PlcB is cleaved by a zinc-dependent metallo- g 

protease encoded by mpl (Poyart et al., 1993). Interestingly, L. monocytogenes g 

appears to sequester pools of inactive pro-PlcB during bacterial replication § 

o 
within the host cytosol, and active enzyme is then rapidly released when Q 

pH decreases below 7.0 (Marquis and Hager, 2000). PlcB is required for the § 

efficient escape of I. monocytogenes from the secondary double-membrane g 

vacuoles formed during the process of cell-to-cell spread (discussed in the g 

following section); however, the protein can functionally replace LLO to me- > 

diate escape from the primary phagosome in human epithelial cells (Marquis g 

etal., 1995). 

In addition to the factors mentioned herein, a previously unknown 

surface protein designated SvpA (surface virulence-associated protein) was 3 

recently suggested to promote phagosomal escape and thus intracellular S 

survival of L. monocytogenes (Borezee et al., 2001). A svpA mutant was shown g 

to be confined within the phagosomal vacuole and was attenuated in a » 



H 



> 
O 



o 



ffi 



mouse model of infection with decreased bacterial growth in the liver and g 

spleen (Borezee et al., 2001). L. monocytogenes therefore appears to possess a 
variety of factors that may work in concert to facilitate efficient escape from 
the phagosome so that the bacterium can reach the promised land: the host 
cytosol. 



THE GOOD LIFE: L MONOCYTOGENES AT HOME IN 
THE CYTOSOL 

The doubling time for L. monocytogenes within the host cytosol is 
very similar to the doubling time observed for bacterial cultures grown in 
rich broth media; therefore, the cytosol appears to be a very permissive 




environment for I. monocytogenes proliferation (Portnoy et al., 1988). Intra- 
cellular growth of L. monocytogenes does not require the induction of stress 
proteins, and several auxotrophic mutants of I. monocytogenes were able to 
replicate in the cytosol of the macrophage-like cell line J 774 and the human 
epithelial cell line Henle 407 with growth rates similar to those of wild-type 
strains (Marquis et al., 1993; Hanawa et al., 1995). The expression of three 
genes involved in purine and pyrimidine biosynthesis (purH, purD, and pyrE ) , 
and a gene encoding an ATP-dependent arginine transporter (arpj), is in- 
duced in bacteria located within host cells; however, L. monocytogenes strains 
carrying mutations in these genes were not affected for rates of intracellu- 
lar growth and were fully virulent in a mouse model of infection (Klarsfeld 
et al., 1994). Other in vivo-induced genes that encode metabolic enzymes and 
nutrient transport systems have been identified through the use of a variety of 
o genetic screens (Dubail et al., 2000; Gahan and Hill, 2000; Autret et al, 2001; 

w Wilson et al., 2001). It has been suggested that some nutrients, such as nU- 

* cleotides and certain amino acids, are not at limiting concentrations within 

g the cytosol but are nonetheless at low concentration, and I. monocytogenes 

< therefore induces the expression of genes required for the uptake and biosyn- 

g thesis of various metabolites to facilitate efficient intracellular proliferation 

o (Klarsfeld et al, 1994; Sheehan et al., 1994; Vazquez- Boland et al, 2001b). 

£ — 

o Recent evidence has implicated Hpt, a L. monocytogenes homologue of 

pi the mammalian glucose-6-phosphate (G-6-P) translocase, in facilitating rapid 

>* intracellular proliferation of bacteria through utilization of hexose pilos- 

is phates obtained from the host cytosol (Chico-Calero et al., 2002). The mam- 

malian G-6-P translocase transports G-6-P from the cytosol into the endoplas- 
mic reticulum. L. monocytogenes Hpt may function in an analogous manner to 
transport G-6-P from the cytosol into the bacterium; bacterial mutants lack- 
ing the G-6-P translocase were defective for intracellular proliferation and 
were attenuated for virulence in mice (Chico-Calero et al., 2002). 

The acquisition of iron is important for L. monocytogenes, as it is for all 
other living cells. Iron is an essential cofactor for a wide variety of enzymatic 
processes. Iron stimulates I. monocytogenes growth in vitro, and it increases 
bacterial proliferation in vivo (Sword, 1966; Stelma et al., 1987; Payne, 1993). 
Increased concentrations of iron have been shown to upregulate the expres- 
sion of inlAB and enhance L. monocytogenes invasion of Caco-2 cells (Conte 
et al., 1996). In macrophages, iron is required for antimicrobial responses in- 
volving the generation of reactive oxygen and nitrogen intermediates. How- 
ever, humans with iron overload are more susceptible to listeriosis; thus a 
balance of iron must be maintained to appropriately respond to L. mono- 
cytogenes infection (Schuchat et al., 1991; Fleming and Campbell, 1997). In 



mammalian hosts, iron is normally sequestered by serum transferrin and is 
also complexed by intracellular heme compounds; it is therefore not freely 
available to L. monocytogenes. Unlike many pathogens, L. monocytogenes does 
not appear to secrete the high-affinity iron-binding siderophores; however, 
the organism can use exogenous siderophores by means of an extracellu- 
lar ferric iron reductase. L. monocytogenes also possesses a citrate-inducible 
iron transport system and a cell-surface-associated trans ferrin-binding pro- 
tein (Adams et al., 1990; Hartford et al., 1993; Deneer et al., 1995; Barchini 
and Cowart, 1996; Coulanges et al., 1997). 

THE GRASS IS ALWAYS GREENER: SPREAD OF 
L MONOCYTOGENES TO ADJACENT HOST CELLS 

The cytosol may be a tasty smorgasbord for replicating L. monocytogenes, 
but it pays to move onto greener cell pastures. Each cell can only sustain a g 

finite amount of bacterial replication, and evidence suggests that I. mono- g 

cytogenes needs to spread to new host cells as a means of outrunning host § 

o 
cellular immune responses (Auerbuch et al., 2001). The necessity of cell-to- Q 

cell spread as a means of obtaining fresh supplies of nutrients and of avoiding § 

humoral and cellular immune responses was discussed in a recent review g 

(O'Riordan and Portnoy, 2002), in which the authors also noted that although g 

there are not many cytosolic bacterial pathogens, most of them are capable > 

of cell-to-cell spread. 




H 



O 

> 

O 



L monocytogenes-d'irected actin-based motility § 

L. monocytogenes cytosol motility and bacterial spread to adjacent cells S 

is accomplished by the exploitation of host actin polymerization machinery. g 

Observations of infected host cells by electron microscopy indicate that cy- » 

o 

tosolic L. monocytogenes is soon surrounded by a dense cloud of host-derived g 

F-actin, which eventually forms a comet-tail-like structure at one bacterial h 

pole (Tilney and Portnoy, 1989). Approximately 2h after internalization, 
I. monocytogenes begins to move in the host cytosol with a velocity of 0.1-0.4 
/xm/s, a rate approximately proportional to the rate of actin polymerization 
(Theriot et al., 1992). The L. monocytogenes surface protein ActA is absolutely 
necessary for actin-based motility within the host cytosol. 

ActA was identified following the isolation of I. monocytogenes mu- 
tants that lacked the ability to polymerize actin and spread to adjacent cells 
(Domann et al., 1992; Kocks et al., 1992). I. monocytogenes actA mutants in- 
vade host cells, escape from the phagosome, and replicate within the cytosol, 




3 



T 



L 



a 

< 






3 

'■5 
< 



I 

< 






oc 



oc 



3 



O 



? 



u 





1 


2 












1*1 










? 


► \ <i 










i 










© 


£ 


<m 








N 












<D 




< \ 








3 






00 






D- 






Tf 






CD 






^H 






Q- 












0> 






\<] 






> 






W 






c 






\ t- 






o 






\ " 












\ T 













-5V O 






2! 






?\3 






CL 






i*\ 






X 






00 \ 






CD 

a 






< \ 


00 




o 






^^\ 


T^ 




« 








1 1 






o 

D 



o 

a? 









s 



o 



p 

*- 
Q 

CJ 

N 

« 

CJ 

3 



ooqoqoqo 
oodd dddd 

in <5 in o ifi o K 

CO CO CJ CM i- i- 



« 







< 



so 

IS 




< 



so 




but the bacteria are unable to use actin polymerization as a motile force and 
thus accumulate as microcolonies within infected host cells. L. monocytogenes 
actA deletion mutants are 1000-fold less virulent in mice, and infection is re- 
stricted predominantly to the gut (Domann et al., 1992; Kocks et al., 1992; 
Manohar et al., 2001). ActA appears to be the only L. monocytogenes protein 
required to nucleate actin polymerization; actin-based motility and the for- 
mation of actin comet tails is observed in cell extracts for both streptococci 
and beads that have been asymmetrically coated with purified ActA (Smith 
etal., 1995; Cameron etal., 1999). L. innocua, a nonpathogenic Listeria species, 
does not polymerize actin when added to cell extracts, but the expression of 
the actA gene product on the L. innocua surface enables the bacteria to form 
actin tails and move (Kocks et al., 1995). 

The length of L. monocytogenes actin tails reflects the rates of bacterial 
movement (Theriot et al., 1992; Sechi et al., 1997), and the rate of move- 
ment is likely to affect the efficiency of cell-to-cell spread. Interestingly, there g 
seems to be an optimal window of ActA production required for normal g 

movement and cell-to-cell spread. A single additional copy of the actA gene § 

o 
integrated within the L. monocytogenes chromosome can reduce the efficiency Q 

of cell-to-cell spread, apparently because of increased ActA production (Lauer § 

et al., 2002). Conversely, recent research in our laboratory indicates that g 

a reduction in actA expression appears to proportionally reduce cell-to-cell g 

spread. A series of actA promoter mutants, demonstrated to express de- > 

creasing amounts of actA as measured through the use of actA-gus reporter g 

gene fusions, produced correspondingly smaller plaques in monolayers of £j 

mouse L2 fibroblast cells (Fig. 6.3; K. Wong and N. Freitag, unpublished 

data). A linear relationship was found to exist between actA expression and 3 

plaque size. Decreasing actA expression was also associated with decreasing S 

actin tail length as visualized by staining for filamentous actin in monolay- g 

ers of infected PtK2 cells in tissue culture (Fig. 6.3). These data support the » 

o 

o 

Figure 6.3. (facing page). Correlation of actA expression levels with L. monocytogenes * 

cell-to-cell spread. (A) Diagram of the actA promoter region. The dark arrow represents 

the transcript initiation site, SD indicates the ribosome binding site, and the PrfA binding 

site is indicated by the presence of the PrfA box. Targeted actA deletions introduced into 

the L. monocytogenes chromosome are shown, with the numbers representing the size of 

the deletion and the distance from the ATG start codon. (B) Actin staining of PtK2 

epithelial cells infected with L. monocytogenes wild type (NF-L476) or the actA promoter 

mutant strains. Filamentous actin is shown in green, and bacteria are shown in red. (C) 

Correlation between levels of actA expression and cell-to-cell spread ability of I. 

monocytogenes. See color section. 



H 




premise that an optimal concentration of ActA at the bacterial surface is 
critical for function (Moors et al., 1999). 

The mature ActA protein contains 610 amino acids after the cleavage of 
a 29-residue signal sequence from its N-terminus. ActA can be divided into 
three domains: an N-terminal domain (1-234) rich in cationic residues, a 
central domain (235-394) containing proline-rich repeats, and a C-terminal 
domain (395-610) containing a stretch of hydrophobic residues (585-606) 
that anchors the protein to the surface of L. monocytogenes. The N-terminal 
domain contains all the information necessary to mediate actin polymeriza- 
tion and support actin-based motility (Lasa et al., 1997). The cationic residues 
(129-153) within this domain appear to be critical for actin assembly; residues 
21-97 are responsible for the continuity of the assembly process; and residues 
117-121 contribute to tail formation (Lasa et al., 1997). 
u A key function of the ActA N-terminal domain is the recruitment of 

w the host Arp2/3 complex, which is critical for the nucleation step initiat- 

* ing actin filament formation (Welch et al., 1997). The N-terminus of ActA 

g shares homology with the Wiskott-Aldrich syndrome protein (WASP) fam- 

< ily in mammalian cells (Skoble et al., 2000; Boujemaa-Paterski et al., 2001), 

g which have been shown to stimulate actin polymerization through activation 

of Arp2/3 and to connect the cytoskeleton with signaling pathways that me- 

o diate actin rearrangement (Welch et al., 1997; Machesky and Insall, 1998; 

p\ Goldberg, 2001). ActA binds two actin monomers and three of the seven 

p* subunits of the Arp2/3 complex (Zalevsky et al., 2001). A mutational analy- 

§ sis of the ActA N-terminus indicates that the actin-binding region spans 40 

residues (31-72); a separate region of ActA (136-231) contributes to the actin 
cloud-to-tail transition that affects intracellular motility rates (Lauer et al., 
2001). 

The central domain of ActA contains four short proline-rich repeats 
(PRRs) with consensus sequence DFPPPPTDEEL. These PRRs are similar to 
PRRs present in mammalian cytoskeletal proteins, and they serve as binding 
sites for the host tetrameric vasodilator phospho-protein (VASP; see Reinhard 
etal., 1995; Niebuhretal., 1997; Dramsi and Cossart, 1998). VASP binds host 
profilin, an actin-sequestering protein that recruits actin monomers for the 
elongation of the actin filaments. Both VASP and profilin have been shown to 
co-localize at the interface between I. monocytogenes and the actin tails (The- 
riotetal., 1994; Chakrabortyetal., 1995). Interestingly, I. monocytogenes is still 
capable of actin recruitment and intracellular movement in the absence of 
the PRRs, VASP, or profilin, although the organism moves at a significantly 
lower rate (Lasa et al., 1995; Smith et al., 1996; Niebuhr et al., 1997; Loisel 




: s ActA dimer £7 actin monomer ^ profilin 
i tiitm actin filament^ capping protein ▼ cofilin 
W VASP ^ Arp2/3 complex J ct-actinin 

Figure 6.4. Model of actin nucleation and polymerization as directed by I. monocytogenes 
(based on previous models described by Cossart and Bierne, 2001 and Skoble et al., 2000). 
VASP binds to the four proline-rich repeats present in the central domain of ActA. The 
N-terminal domain of ActA contains three regions that contribute to actin binding and the 
activation of the Arp2/3 complex. 

et al., 1999). This suggests a stimulatory role for the ActA PRRs and VASP 
in actin-based motility. 

Several other host proteins are involved in L. monocytogenes-directed actin 
polymerization. Alpha-actinin is distributed over the actin tail and is respon- 
sible for stabilizing the tail by cross-linking actin filaments (Dabiri et al., 1990; 
Loisel et al., 1999) . Although alpha-actinin is not required for I. monocytogenes 
actin-based motility, the tails are less rigid and the bacteria tend to drift in its 
absence. Capping protein blocks actin assembly by binding to barbed ends of 
actin filaments. This protein was shown to be required in the reconstitution 
of L. monocytogenes motility in vitro. The absence of this protein or its pres- 
ence at low concentrations results in a fishbone-like web of actin filaments, 
which is inefficient at providing motile force (Marchand et al., 1995; Loisel 
et al., 1999; Pantaloni et al., 2000). Cofilin depolymerizes actin and may play 
a role in generating pools of actin monomers (Carlier et al., 1997); it is also 
required for reconstituting bacterial movement in vitro (Loisel et al., 1999). 
Taken together, a model depicting L. monocytogenes actin-based movement is 
shown in Fig. 6.4. 




3 

to 

> 

o 

% 
o 
n 

o 

hi 
% 

z 

$ 

on 

o 

> 
o 
z 

H 

n 

M 
(- 1 
M 

a 



to 

o 
to 
o 



ffi 




< 



L monocytogenes invasion of adjacent cells 

As I. monocytogenes moves through the cell and contacts the plasma 
membrane, it appears to push out the membrane and form a pseudopod-like 
(or Listeria-pod-\ike) structure that is somehow recognized or taken up by 
the neighboring cell. The bacterium is then enclosed in a double-membrane- 
bound secondary vacuole, and lysis of this compartment is facilitated by the 
phospholipase PC-PLC (Vazquez- Boland et al., 1992). The plcA gene product 
(PI-PLC) appears to function synergistically with PC-PLC, as I. monocyto- 
genes mutants lacking both enzymes have a more serious defect in cell-to-cell 
spread than either of the single mutants (Marquis et al., 1995). LLO is also 
required for escape from the double-membrane vacuoles (Gedde et al., 2000) . 
Through the coordinated activities of the hly, plcA, and plcB gene products, 
I. monocytogenes disrupts the double -membrane vacuole and makes a fresh 
eh home within the cytosol of the new cell victim. In this way, the bacterium 

g gains new pastures without exposure to the humoral immune responses of 

« the host. 

u 

< 

% COORDINATING THE BACTERIAL DANCE STEPS OF INVASION, 

t REPLICATION, AND CELL-TO-CELL SPREAD 

o 

p\ I. monocytogenes is similar to other bacterial pathogens in that many of 

w 

^ the gene products that contribute to infection are expressed preferentially 

§ within host cells. The expression of virulence determinants by I. monocyto- 

genes appears to be tightly regulated and can be triggered by specific host cell 
compartment environments, such as the phagosome or cytosol. The process 
of L. monocytogenes virulence gene regulation in response to host cell envi- 
ronments is a complex and fascinating story, with many of the players still 
waiting to be discovered. 

Most of the virulence genes identified thus far in L. monocytogenes are 
located within a 10-kb chromosomal region known as the Listeria pathogenic- 
ity island 1 (LIPI-1; see Vazquez- Boland et al., 2001a). In addition to the 
aforementioned virulence genes plcA, hly, mpl, actA, and plcB, a positive 
transcriptional regulator known as prfA is located within this cluster, and 
LIPI-1 is commonly referred to as the PrfA regulon (Fig. 6.5). The LIPI-1 
is only present in pathogenic Listeria species, although an apparently inacti- 
vated form is also present in I. seeligeri. This region is located between prs 
and Idh, two housekeeping genes coding for phosphoribosyl-pyrophosphate 
synthetase and lactate dehydrogenase, respectively. DNA sequences in the 
prs-ldh intergenic region of the nonpathogenic listeriae were shown to have 



+ + 



[prfA — ( plcA 





hly > mpl > act A / plcB 



lkb 



Figure 6.5. The L. monocytogenes LIPI— 1, also referred to as the PrfA regulon; 
+ designates promoters whose expression is increased by PrfA binding, and 
— designates promoters for which PrfA appears to decrease expression. 




homology with integration consensus sequences for a transposon, support- 
ing the hypothesis that horizontal gene transfer might be involved in the 
evolution of this gene cluster (Gouin et al., 1994; Cai and Wiedmann, 2001). 
PrfA is essential for L. monocytogenes virulence, and it is required for 
the expression of all the other genes within the LIPI-1. The prfA gene was 3 

first identified during the characterization of a spontaneous nonhemolytic ^ 

L. monocytogenes mutant that was found to contain a small deletion in an § 



H 



O 



open reading frame downstream of pic A. Complementation of the mutation Q 

with the complete open reading frame led to increased transcription from § 

genes located within LIPI-1 and demonstrated the pleiotropic effects of the g 

prfA gene product (Leimeister-Wachter et al., 1989, 1990). The expressions g 

of several other virulence genes, such as inlA, MB, hpt, and bsh (encoding a > 

bile salt hydrolase) , have also subsequently been shown to be completely or g 

partially dependent on PrfA (Kreft and Vazquez-Boland, 2001; Chico-Calero £j 
et al., 2002; Dussurget et al., 2002). Other I. monocytogenes genes, such as 

clpC, whose gene product contributes to stress responses, said flaA, encoding 3 

the flagellin subunit, appear to be negatively regulated by PrfA (Ripio et al., S 

1997, 1998). In contrast, there are virulence genes that are PrfA-independent, g 

such as svpA and the inlGHE operon (Lingnau et al., 1995; Ripio et al., 1997; * 



o 



Raffelsbauer et al., 1998; Borezee et al., 2001). § 

Although the expression of many L. monocytogenes virulence genes is h 

dependent on PrfA, the patterns of gene expression vary for individual genes 
during bacterial growth and during the course of host cell infection. inlA and 
MB expression is readily detectable during bacterial growth outside of host 
cells, whereas hly, pic A, and plcB have been reported to be induced within the 
phagosome (Bubert et al., 1999). The expression of mpl and actA is highly 
upregulated within the host cytosol, whereas actA expression alone appears 
downregulated within the secondary vacuoles that are formed following the 
spread of I. monocytogenes to adjacent cells (Bohne et al., 1994; Bubert et al., 
1999; Freitag and Jacobs, 1999). The patterns of virulence gene expression 




observed appear to correspond perfectly with the functional roles ascribed 
to the different products. However, it is unclear why plcB transcripts were 
detected in the phagosome in the absence of detectable actA mRNA, as the 
two genes are cotranscribed with no apparent intergenic promoter (Vazquez- 
Boland et al., 1992). It has been suggested that the actA mRNA region and 
the plcB mRNA region may differ in message stability. 

The 27-kDa PrfA protein is a member of the Crp/Fnr family of tran- 
scriptional activators, most members of which have been identified in Gram- 
negative bacteria (Lampidis et al., 1994; Ripio et al., 1997). In Escherichia coli, 
Crp (also known as Cap) and Fnr are involved in carbon utilization and anaer- 
obic growth responses, respectively (Spiro and Guest, 1990; Kolb et al., 1993). 
The PrfA sequence is only approximately 20% identical and 30% similar to 
that of Crp; however, the two proteins appear to share significant structural 
u similarities (based on the reported Crp structure and on sequence structure 

w predictions for PrfA). Both proteins contain an N-terminal /3-roll structure 

* and a C -terminal DNA-binding helix-turn -helix (HTH) motif. The activation 

g regions that are involved in the interaction of Crp with RNA polymerase may 

< also be conserved in PrfA. Perhaps most intriguing in regard to the similar- 

g ities between Crp and PrfA is the existence of mutationally activated forms 

of the proteins, known as Crp* and PrfA*. Crp requires the presence of a 

o cofactor, cAMP, for full activity, and two mutations within Crp have been 

^ described, G145S and A144T, that lead to cofactor-independent Crp activa- 

>* tion (Crp* mutants; see Garges and Adhya, 1988). Ripio et al. (1997) found 

§ that L. monocytogenes isolates that expressed high amounts of LLO or PC-PLC 

contained a single amino acid substitution within PrfA located in a region 
(G145S) similar to the mutations found to confer the Crp* phenotype to Crp. 
The PrfA* G145S substitution causes constitutively high expression of all 
PrfA-regulated genes, suggesting that, like Crp, a cofactor (which does not 
appear to be cAMP) may be required for PrfA activation (Lampidis et al., 
1994; Ripio et al., 1997; Goebel et al., 2000). 

Our laboratory has recently identified two additional mutations within 
PrfA that appear to confer a PrfA* phenotype (Shetron-Rama et al., 2003). 
L. monocytogenes PrfA E77K contains a glutamate -to -lysine substitution 
within the /3-roll structure of PrfA, whereas PrfA G155S contains a glycine- 
to-serine substitution near the location of the original PrfA* mutation 
(G145S). Both mutations confer high-level expression of PrfA-regulated 
genes, and strains containing either mutation appear hyper-invasive for ep- 
ithelial cell lines. Interestingly, the PrfA G155S mutant strain is approx- 
imately fivefold more virulent in a mouse model of infection than the 



Environmental signals that influence virulence gene expression 




wild-type strain, suggesting that constitutive activation of PrfA is not detri- 
mental to virulence. 

PrfA is a site-specific DNA-binding protein that recognizes a 14-bp palin- 
drome known as the "PrfA-box" within the -40 region of PrfA-dependent 
promoters. The binding affinity of PrfA for its target promoters is influenced 
by the nucleotide sequence of the PrfA box. PrfA binds with high affinity to 
the perfectly symmetrical PrfA box (TTAACA-NN-TGTTAA) shared by the 
divergently transcribed hly and plcA promoters. The PrfA boxes of the mpl 
and act A promoters contain single base pair asymmetries, and the inlA PrfA 
box has two base pair asymmetries. These PrfA box asymmetries have been 
postulated to cause PrfA to bind with lower affinity to these target promoters, 
and thus it has been speculated that the relative amount of PrfA present at 
any time within the cell determines the level of expression of the various 
virulence genes (Freitag et al., 1993; Goebel et al., 2000; Kreft and Vazquez- 
Boland, 2001). Recent studies have shown that although the expression of g 

some genes appears to follow the PrfA binding affinity rule (hly, mpl), other g 

PrfA-dependent promoters, such as the act A promoter, do not (Williams et § 

o 
al., 2000), and that even the increased expression of the mutationally acti- Q 

vated prfA* allele is not sufficient to raise levels of actA expression to those § 

observed for bacteria located within the host cell cytosol (Shetron-Rama et al., * 

2002; Greene and Freitag, 2003) . Our laboratory has isolated I. monocytogenes g 

mutants that express high levels of actA during growth in broth culture, and > 

these mutations map at least 40 kb outside of the PrfA regulon, indicating g 

the participation of additional factors in the induction of actA expression £j 

(Shetron-Rama et al., 2003). 



H 



O 



z 
n 

M 
M 
M 

a 



Multiple conditions have been reported to influence I. monocytogenes vir- » 



o 



ulence gene expression in culture, including temperature, available carbon g 

sources, nutrient deprivation, and iron (Kreft and Vazquez- Boland, 2001) . Vir- h 

ulence gene expression has been reported to be optimally activated at 37°C, 
the body temperature of mammalian hosts (Leimeister-Wachter et al., 1992), 
and the expression of the key regulatory protein PrfA appears to be under 
the control of a thermosensitive mechanism of translation in which optimal 
translation occurs at 37° C (Johansson et al., 2002). However, L. monocyto- 
genes virulence gene expression has been observed within infected insect cell 
lines maintained at room temperature (Mansfield et al., 2003 (in press)), 
indicating that the appropriate cell environment can stimulate bacterial 



gene expression even at low temperatures. Iron and available carbon sources 
influence the expression of I. monocytogenes virulence genes, and bacterial 
gene products involved in iron uptake or in the utilization of hexose sug- 
ars are upregulated by intracellular L. monocytogenes (Sword, 1966; Payne, 
1993; Klarsfeld et al, 1994; Borezee et al., 2001; Chico-Calero et al, 2002). 
Interestingly, the proximal prfAP2 promoter was recently demonstrated to be 
regulated by the stress-responsive alternative sigma factor a B ; however, the 
loss of a B function only modestly reduces the virulence of I. monocytogenes 
(Nadon et al., 2002) . As L. monocytogenes is a bacterium that survives in a mul- 
titude of environments, regulation of gene expression has probably evolved 
in order to respond to a variety of signals that tell the bacterium whether its lo- 
cation lies inside or outside of a multicellular organism. When the bacterium 
determines that it is indeed inside of a prospective host organism, additional 
^ signals may then serve to direct the bacterium to its replicative niche within 




H 



m the cytosol. 



a REFERENCES 

< 



« 



g Abachin, E., Poyart, C, Pellegrini, E., Milohanic, E., Fiedler, F., Berche, P., and 

Trieu-Cuot, P. (2002). Formation of D-alanyl-lipoteichoic acid is required for 

o adhesion and virulence of Listeria monocytogenes. Mol. Microbiol. 43, 1-14. 

pi Adams, T.J., Vartivarian, S., and Cowart, R.E. (1990). Iron acquisition systems of 

>i Listeria monocytogenes. Infect. Immun. 58, 2715-2718. 

g Alexander, A.V., Walker, R.L., Johnson, B.J., Charlton, B.R., and Woods, L.W. 

(1992). Bovine abortions attributable to Listeria ivanovii: four cases (1988- 
1990) J. Am. Vet. Med. Assoc. 200, 711-714. 
Alvarez-Dominguez, C, Carrasco-Marin, E., and Leyva-Cobian, F. (1993). Role 
of complement component Clq in phagocytosis of Listeria monocytogenes by 
murine macro phage-like cell lines. Infect. Immun. 61, 3664—3672. 
Alvarez-Dominguez, C, Roberts, R., and Stahl, P.D. (1997a). Internalized Listeria 
monocytogenes modulates intracellular trafficking and delays maturation of 
the phagosome./. Cell Sci. 110, 731-743. 
Alvarez-Dominguez, C, Vazquez-Boland, J.A., Carrasco-Marin, E., Lopez-Mato, 
P., and Leyva-Cobian, F. (1997b). Host cell heparan sulfate proteoglycans 
mediate attachment and entry of Listeria monocytogenes, and the listerial sur- 
face protein ActA is involved in heparan sulfate receptor recognition. Infect. 
Immun. 65, 78-88. 
Antal, E.A., Loberg, E.M., Bracht, P., Melby, K.K., and Maehlen, J. (2001). Evidence 
for intraaxonal spread of Listeria monocytogenes from the periphery to the 
central nervous system. Brain Pathol. 11, 432-438. 



Armstrong, R.W. and Fung, P.C. (1993). Brainstem encephalitis (rhomben- 
cephalitis) due to Listeria monocytogenes: case report and review. Clin. Infect. 
Dis. 16, 689-702. 

Auerbuch, V., Lenz, L.L., and Portnoy, D.A. (2001). Development of a competitive 
index assay to evaluate the virulence of Listeria monocytogenes actA mutants 
during primary and secondary infection of mice. Infect. Immun. 69, 5953- 
5957. 

Aureli, P., Fiorucci, G.C., Caroli, D., Marchiaro, G., Novara, O., Leone, L., and 
Salmaso, S. (2000). An outbreak of febrile gastroenteritis associated with corn 
contaminated by Listeria monocytogenes. N. Engl. J. Med. 342, 1236-1241. 

Autret, N., Dubail, I., Trieu-Cuot, P., Berche, P., and Charbit, A. (2001). Identi- 
fication of new genes involved in the virulence of Listeria monocytogenes by 
signature-tagged transposon mutagenesis. Infect. Immun. 69, 2054—2065. 

Barchini, E. and Cowart, R.E. (1996). Extracellular iron reductase activity produced 

by Listeria monocytogenes. Arch. Microbiol. 166, 51-57. » 

Beauregard, K.E., Lee, K.D., Collier, R.J., and Swanson, J. A. (1997). pH -dependent ^ 

perforation of macrophage phagosomes by listeriolysin O from Listeria mono- § 




H 



O 



cytogenes. J. Exp. Med. 186, 1159-1163. Q 

Bergmann, B., Raffelsbauer, D., Kuhn, M., Goetz, M., Horn, S., and Goebel, W. § 

(2002). InlA- but not InlB-mediated internalization of Listeria monocytogenes g 

by non-phagocytic mammalian cells needs the support of other internalins. g 

Mol. Microbiol. 43, 557-570. $ 

Bielecki, J., Youngman, P., Connelly, P., and Portnoy, D.A. (1990). Bacillus sub- § 

tills expressing a haemolysin gene from Listeria monocytogenes can grow in jj 

mammalian cells. Nature 345, 175-176. 

Bierne, H., Mazmanian, S.K., Trost, M., Pucciarelli, M.G., Liu, G., Dehoux, P., 3 

Jansch, L., Garcia-del Portillo, F., Schneewind, O., and Cossart, P. (2002). w 

Inactivation of the srtA gene in Listeria monocytogenes inhibits anchoring of 5 

surface proteins and affects virulence. Mol. Microbiol. 43, 869-881. » 

o 
Bohne, J., Sokolovic, Z., and Goebel, W. (1994). Transcriptional regulation of prfA g 

and PrfA-regulated virulence genes in Listeria monocytogenes. Mol. Microbiol. 

11, 1141-1150. 

Borezee, E., Pellegrini, E., Beretti, J.L., and Berche, P. (2001). SvpA, a novel sur- 
face virulence-associated protein required for intracellular survival of Listeria 
monocytogenes. Microbiology 147, 2913-2923. 

Boujemaa-Paterski, R., Gouin, E., Hansen, G., Samarin, S., Le Clainche, C, 
Didry, D., Dehoux, P., Cossart, P., Kocks, C, Carlier, M.F., and Pantalovi, 
D. (2001). Listeria protein ActA mimics WASP family proteins: it activates 
filament barbed end branching by Arp2/3 complex. Biochemistry 40, 11,390- 
11,404. 



ffi 




Boulnois, G.J., Paton, J.C., Mitchell, T.J., and Andrew, P.W. (1991). Structure 
and function of pneumolysin, the multifunctional, thiol-activated toxin of 
Streptococcus pneumoniae. Mol. Microbiol. 5, 2611-2616. 
Braun, L., Dramsi, S., Dehoux, P., Bierne, H., Lindahl, G., and Cossart, P. (1997). 
InlB : an invasion protein of Listeria monocytogenes with a novel type of surface 
association. Mol. Microbiol. 25, 285-294. 
Braun, L., Ohayon, H., and Cossart, P. (1998). The InlB protein of Listeria mono- 
cytogenes is sufficient to promote entry into mammalian cells. Mol. Microbiol. 
27, 1077-1087. 
Braun, L., Ghebrehiwet, B., and Cossart, P. (2000). gClq-R/p32, a Clq-binding 
protein, is a receptor for the InlB invasion protein of Listeria monocytogenes. 
EMBO J. 19, 1458-1466. 
Bubert, A., Kuhn, M., Goebel, W., and Kohler, S. (1992). Structural and functional 
^ properties of the p60 proteins from different Listeria species. J. Bacteriol. 174, 

S 8166-8171. 

* Bubert, A., Sokolovic, Z., Chun, S.K., Papatheodorou, L., Simm, A., and Goebel, 

g W. (1999). Differential expression of Listeria monocytogenes virulence genes 

< in mammalian host cells. Mol. Gen. Genet. 261, 323-336. 

g Cai, S. and Wiedmann, M. (2001). Characterization of the prfA virulence gene 

cluster insertion site in non-hemolytic Listeria spp. : probing the evolution of 

o the Listeria virulence gene island. Curr. Microbiol. 43, 271-277. 

pi Cameron, L.A., Footer, M.J. , van Oudenaarden, A., andTheriot, J.A. (1999). Motil- 

w 

^ ity of ActA protein-coated microspheres driven by actin polymerization. Proc. 

g Natl. Acad. Sci. USA 96, 4908-4913. 

Camilli, A., Goldfine, H., and Portnoy, D.A. (1991). Listeria monocytogenes mutants 
lacking phosphatidylinositol-specific phospholipase C are avirulent. J. Exp. 
Med. 173, 751-754. 

Carlier, M.F., Laurent, V., Santolini, J., Melki, R., Didry, D., Xia, G.X., Hong, 
Y., Chua, N.H., and Pantaloni, D. (1997). Actin depolymerizing factor 
(ADF/cofilin) enhances the rate of filament turnover: implication in actin- 
based motility. J. Cell Biol. 136, 1307-1322. 

Chakraborty, T., Ebel, F., Domann, E., Niebuhr, K., Gerstel, B., Pistor, S., Temm- 
Grove, C.J., Jockusch, B.M., Reinhard, M., and Walter, U. (1995). A focal ad- 
hesion factor directly linking intracellularly motile Listeria monocytogenes and 
Listeria ivanovii to the actin-based cytoskeleton of mammalian cells. EMBO 
J. 14, 1314-1321. 

Chand, P. and Sadana, J.R. (1999). Outbreak of Listeria ivanovii abortion in sheep 
in India. Vet. Rec. 145, 83-84. 

Chico-Calero, I., Suarez, M., Gonzalez-Zorn, B., Scortti, M., Slaghuis, J., Goebel, 
W., and Vazquez-Boland, J.A. (2002). Hpt, a bacterial homolog of the 



microsomal glucose-6-phosphate translocase, mediates rapid intracellular 

proliferation in Listeria. Proc. Natl. Acad. Sci. USA 99, 431-436. 
Collins, M.D., Wallbanks, S., Lane, D.J., Shah, J., Nietupski, R., Smida, J., Dorsch, 

M., and Stackebrandt, E. (1991). Phylogenetic analysis of the genus Listeria 

based on reverse transcriptase sequencing of 16S rRNA. Int. J. Syst. Bacteriol. 

41, 240-246. 
Conlan, J.W. and North, R.J. (1991). Neutrophil-mediated dissolution of infected 

host cells as a defense strategy against a facultative intracellular bacterium. 

J. Exp. Med. 174, 741-744. 
Conlan, J.W. and North, R.J. (1994). Neutrophils are essential for early anti- Listeria 

defense in the liver, but not in the spleen or peritoneal cavity, as revealed 

by a granulocyte -depleting monoclonal antibody. J. Exp. Med. 179, 259- 

268. 
Conte, M.P., Longhi, C, Polidoro, M., Petrone, G., Buonfiglio, V., Di Santo, S., 

Papi, E., Seganti, L., Visca, P., and Valenti, P. (1996). Iron availability affects £j 

entry of Listeria monocytogenes into the enterocytelike cell line Caco-2. Infect. ^ 



P. (1989). Listeriolysin O is essential for virulence of Listeria monocytogenes: 




H 



Immun. 64, 3925-3929. § 

o 
Cossart, P., Vicente, M.F., Mengaud, J., Baquero, F., Perez-Diaz, J.C., and Berche, Q 

H 

O 

to 

direct evidence obtained by gene complementation. Infect. Immun. 57, 3629- g 

z 



3636. 

Cossart, P. and Lecuit, M. (1998). Interactions of Listeria monocytogenes with mam- > 

malian cells during entry and actin-based movement: bacterial factors, eel- § 

lular ligands and signaling. EM BO J. 17, 3797-3806. jg 

Cossart, P. and Bierne, H. (2001). The use of host cell machinery in the patho- 

genesis of Listeria monocytogenes. Curr. Opin. Immunol. 13, 96-103. 3 

Cossart, P. (2002). Molecular and cellular basis of the infection by Listeria mono- w 

cytogenes: an overview. Int. J. Med. Microbiol. 291, 401-409. 3 

Cottin, J., Loiseau, O., Robert, R., Mahaza, C, Carbonnelle, B., and Senet, J.M. » 

o 

(1990). Surface Listeria monocytogenes carbohydrate-binding components re- g 

vealed by agglutination with neoglycoproteins. FEMS Microbiol. Lett. 56, 301- 

305. 

Coulanges, V., Andre, P., Ziegler, O., Buchheit, L., and Vidon, D.J. (1997). Uti- 
lization of iron-catecholamine complexes involving ferric reductase activity 
in Listeria monocytogenes. Infect. Immun. 65, 2778-2785. 

Cousens, L.P. and Wing, E.J. (2000). Innate defenses in the liver during Listeria 
infection. Immunol. Rev. 174, 150-159. 

Cowart, R.E., Lashmet, J., Mcintosh, M.E., and Adams, T.J. (1990). Adherence of 
a virulent strain of Listeria monocytogenes to the surface of a hepatocarcinoma 
cell line via lectin-substrate interaction. Arch. Microbiol. 153, 282-286. 



ffi 




Croize, J., Arvieux, J., Berche, P., and Colomb, M.G. (1993). Activation of the 
human complement alternative pathway by Listeria monocytogenes: evidence 
for direct binding and proteolysis of the C3 component on bacteria. Infect. 
Immun. 61, 5134-5139. 
Cummins, A.J., Fielding, A.K., and McLauchlin, J. (1994). Listeria ivanovii infec- 
tion in a patient with AIDS. J. Infect. 28, 89-91. 
Dabiri, G.A., Sanger, J.M., Portnoy, D.A., and Southwick, F.S. (1990). Listeria 
monocytogenes moves rapidly through the host-cell cytoplasm by inducing 
directional actin assembly. Proc. Natl. Acad. Sci. USA 87, 6068-6072. 
Dalton, C.B., Austin, C.C., Sobel, J., Hayes, P. S., Bibb, W.F., Graves, L.M., Swami- 
nathan, B., Proctor, M.E., and Griffin, P.M. (1997). An outbreak of gastroen- 
teritis and fever due to Listeria monocytogenes in milk. N. Engl. J. Med. 336, 
100-105. 
u Daniels, J. J., Autenrieth, I.B., and Goebel, W. (2000). Interaction of Listeria mono- 

w cytogenes with the intestinal epithelium. FEMS Microbiol. Lett. 190, 323-328. 

* Davies, W.A. (1983). Kinetics of killing Listeria monocytogenes by macrophages: 

g rapid killing accompanying phagocytosis. J. Reticuloendothel. Soc. 34, 131— 

< 141. 

g Decatur, A.L. and Portnoy, D.A. (2000). A PEST-like sequence in listeriolysin O 

essential for Listeria monocytogenes pathogenicity. Science 290, 992-995. 

o Deneer, H.G., Healey, V., and Boychuk, I. (1995). Reduction of exogenous ferric 

pi iron by a surface -associated ferric reductase of Listeria spp. Microbiology 141, 

" 1985-1992. 

§ Domann, E., Wehland, J., Rohde, M., Pistor, S., Hard, M., Goebel, W., Leimeister- 

Wachter, M., Wuenscher, M., and Chakraborty, T. (1992). A novel bacterial 
virulence gene in Listeria monocytogenes required for host cell microfilament 
interaction with homology to the proline-rich region of vinculin. EMBO J. 
11, 1981-1990. 
Domann, E., Zechel, S., Lingnau, A., Hain, T., Darji, A., Nichterlein, T., Wehland, 
J., and Chakraborty, T. (1997). Identification and characterization of a novel 
PrfA-regulated gene in Listeria monocytogenes whose product, IrpA, is highly 
homologous to internalin proteins, which contain leucine-rich repeats. Infect. 
Immun. 65, 101-109. 
Dramsi, S., Biswas, I., Maguin, E., Braun, L., Mastroeni, P., and Cossart, P. (1995). 
Entry of Listeria monocytogenes into hepatocytes requires expression of inIB, 
a surface protein of the internalin multigene family. Mol. Microbiol. 16, 251- 
261. 
Dramsi, S., Dehoux, P., Lebrun, M., Goossens, P.L., and Cossart, P. (1997). Iden- 
tification of four new members of the internalin multigene family of Listeria 
monocytogenes EGD. Infect. Immun. 65, 1615-1625. 



Dramsi, S. and Cossart, P. (1998). Intracellular pathogens and the actin cytoskele- 
ton. Annu. Rev. Cell Dev. Biol. 14, 137-166. 

Drevets, D.A. and Campbell, P.A. (1991). Roles of complement and complement 
receptor type 3 in phagocytosis of Listeria monocytogenes by inflammatory 
mouse peritoneal macrophages. Infect. Immun. 59, 2645-2652. 

Drevets, D.A., Leenen, P. J., and Campbell, P.A. (1993). Complement receptor type 
3(CDllb/CD18) involvement is essential for killing of Listeria monocytogenes 
by mouse macrophages. J. Immunol. 151, 5431-5439. 

Drevets, D.A., Sawyer, R.T., Potter, T.A., and Campbell, P.A. (1995). Listeria mono- 
cytogenes infects human endothelial cells by two distinct mechanisms. Infect. 
Immun. 63, 4268-4276. 

Drevets, D.A. (1999). Dissemination of Listeria monocytogenes by infected phago- 
cytes. Infect. Immun. 67, 3512-3517. 

Drevets, D.A., Jelinek, T.A., and Freitag, N.E. (2001). Listeria monocytogenes- 

infected phagocytes can initiate central nervous system infection in mice. £j 

Infect. Immun. 69, 1344-1350. £ 

Dubail, I., Berche, P., and Charbit, A. (2000). Listeriolysin O as a reporter to § 




H 



O 



identify constitutive and in vivo-inducible promoters in the pathogen Listeria Q 

monocytogenes. Infect. Immun. 68, 3242-3250. § 

Dubail, I., Autret, N., Beretti, J.L., Kayal, S., Berche, P., and Charbit, A. (2001). g 

Functional assembly of two membrane-binding domains in listeriolysin O, g 

the cytolysin of Listeria monocytogenes. Microbiology 147, 2679-2688. > 

Dunne, D.W., Resnick, D., Greenberg, J., Krieger, M., and Joiner, K.A. (1994). § 

The type I macrophage scavenger receptor binds to gram-positive bacteria £j 

o 
and recognizes lipoteichoic acid. Proc. Natl. Acad. Sci. USA 91, 1863-1867. 

Dussurget, O., Cabanes, D., Dehoux, P., Lecuit, M., Buchrieser, C, Glaser, P., 3 

and Cossart, P. (2002). Listeria monocytogenes bile salt hydrolase is a PrfA- w 

regulated virulence factor involved in the intestinal and hepatic phases of 5 

listeriosis. Mol. Microbiol. 45, 1095-1106. » 

o 

Engelbrecht, F., Chun, S.K., Ochs, C, Hess, J., Lottspeich, F., Goebel, W., and g 

Sokolovic, Z. (1996). A new PrfA-regulated gene of Listeria monocytogenes 
encoding a small, secreted protein which belongs to the family of internalins. 
Mol. Microbiol. 21, 823-837. 

Engelbrecht, F., Dickneite, C, Lampidis, R., Gotz, M., DasGupta, U., and Goebel, 
W. (1998a). Sequence comparison of the chromosomal regions encompass- 
ing the internalin C genes (inlC) of Listeria monocytogenes and L. ivanovii. 
Mol. Gen. Genet. 257, 186-197. 

Engelbrecht, F., Dominguez-Bernal, G., Hess, J., Dickneite, C, Greiffenberg, 
L., Lampidis, R., Raffelsbauer, D., Daniels, J. J., Kreft, J., Kaufmann, S.H., 
Vazquez-Boland, J.A., and Goebel W. (1998b). A novel PrfA-regulated 



x 




chromosomal locus, which is specific for Listeria ivanovii, encodes two small, 

secreted internalins and contributes to virulence in mice. Mol. Microbiol. 30, 

405-417. 
Facinelli, B., Giovanetti, E., Magi, G., Biavasco, F., and Varaldo, P.E. (1998). Lectin 

reactivity and virulence among strains of Listeria monocytogenes determined 

in vitro using the enterocyte-like cell line Caco-2. Microbiology 144, 109-118. 
Farber, J.M. and Peterkin, P.I. (1991). Listeria monocytogenes, a food-borne 

pathogen. Microbiol. Rev. 55, 476-511. 
Fenlon, D.R. (1999). Listeria monocytogenes in the natural environment. In Listeria, 

Listeriosis, and Food Safety, ed. E.T. Ryser and E.H. Marth, pp. 21-37. New 

York: Marcel Dekker. 
Fischetti, V.A., Pancholi, V., and Schneewind, O. (1990). Conservation of a 

hexapeptide sequence in the anchor region of surface proteins from gram- 
u positive cocci. Mol. Microbiol. 4, 1603-1605. 

w Fleming, S.D. and Campbell, P. A. (1997). Some macrophages kill Listeria mono- 

* cytogenes while others do not. Immunol. Rev. 158, 69-77. 

g Freitag, N.E., Rong, L., and Portnoy, D.A. (1993). Regulation of the prfA tran- 

< scriptional activator of Listeria monocytogenes: multiple promoter elements 

g contribute to intracellular growth and cell-to-cell spread. Infect. Immun. 61, 

2J 2537-2544. 

o Freitag, N.E. and Jacobs, K.E. (1999). Examination of Listeria monocytogenes intra- 

pi cellular gene expression by using the green fluorescent protein of Aequorea 

w 

;* victoria. Infect. Immun. 67, 1844-1852. 

§ Gahan, C.G. and Hill, C. (2000). The use of listeriolysin to identify in vivo induced 

genes in the gram-positive intracellular pathogen Listeria monocytogenes. Mol. 

Microbiol. 36, 498-507. 
Gaillard, J.L., Berche, P., and Sansonetti, P. (1986). Transposon mutagenesis as a 

tool to study the role of hemolysin in the virulence of Listeria monocytogenes. 

Infect. Immun. 52, 50-55. 
Gaillard, J.L., Berche, P., Mounier, J., Richard, S., and Sansonetti, P. (1987). In 

vitro model of penetration and intracellular growth of Listeria monocytogenes 

in the human enterocyte-like cell line Caco-2. Infect. Immun. 55, 2822-2829. 
Gaillard, J.L., Berche, P., Frehel, C, Gouin, E., and Cossart, P. (1991). Entry 

of L. monocytogenes into cells is mediated by internalin, a repeat protein 

reminiscent of surface antigens from gram-positive cocci. Cell 65, 1 127-1 141. 
Gaillard, J.L. and Finlay, B.B. (1996). Effect of cell polarization and differentiation 

on entry of Listeria monocytogenes into the enterocyte-like Caco-2 cell line. 

Infect. Immun. 64, 1299-1308. 
Galan, J.E. (2000). Alternative strategies for becoming an insider: lessons from 

the bacterial world. Cell 103, 363-366. 



Garandeau, C, Reglier-Poupet, H., Dubail, L, Beretti, J.L., Berche, P., and Charbit, 
A. (2002). The sortase SrtA of Listeria monocytogenes is involved in processing 
of internalin and in virulence. Infect. Immun. 70, 1382-1390. 

Garges, S. and Adhya, S. (1988). Cyclic AMP-induced conformational change of 
cyclic AMP receptor protein (CRP): intragenic suppressors of cyclic AMP- 
independent CRP mutations. J. Bacteriol. 170, 1417-1422. 

Gedde, M.M., Higgins, D.E., Tilney, L.G., and Portnoy, D.A. (2000). Role of lis- 
teriolysin O in cell-to-cell spread of Listeria monocytogenes. Infect. Immun. 68, 
999-1003. 

Geoffroy, C, Gaillard, J.L., Alouf, J.E., and Berche, P. (1987). Purification, char- 
acterization, and toxicity of the sulfhydryl-activated hemolysin listeriolysin O 
from Listeria monocytogenes. Infect. Immun. 55, 1641-1646. 

Geoffroy, C, Raveneau, J., Beretti, J.L., Lecroisey, A., Vazquez-Boland, J. A., Alouf, 
J.E., and Berche, P. (1991). Purification and characterization of an extracellu- 
lar 29-kilodalton phospholipase C from Listeria monocytogenes. Infect. Immun. £j 
59, 2382-2388. 



Gilot, P., Jossin, Y., and Content, J. (2000). Cloning, sequencing and characterisa- 




H 



> 

Gilot, P., Andre, P., and Content, J. (1999). Listeria monocytogenes possesses ad- § 

o 
hesins for fibronectin. Infect. Immun. 67, 6698-6701. Q 

H 

O 

tion of a Listeria monocytogenes gene encoding a fibronectin-binding protein. g 

z 



J. Med. Microbiol. 49, 887-896. 

Glaser, P., Frangeul, L., Buchrieser, C, Rusniok, C, Amend, A., Baquero, F., > 

Berche, P., Bloecker, H., Brandt, P., Chakraborty, T., Charbit, A., Chetovani, § 

F., Couvre, E., de Daruvar, A., Dehoux, P., Domann, E., Domingvez-Bernal, jj 

o 
G., Duchard, E., Durant, L., Dussurget, O., Entian, K.D., Fsihi, H., Portillo, 

F.G., Garrido, P., Gautier, L., Goebel, W., Gomez-Lopez, N., Hain, T., Hauf, 3 

J., Jackson, D., Jones, L.M., Kaerst, U., Kreft, J., Kuhn, M., Kunst, F., Kurapkat, w 

G., Madveno, E., Maitdurnam, A., Vicente, J.M., Ng, E., Nedjari, H., Nordsiek, 5 

G., Novella, S., de Pablos, B., Perez-Diaz, J.C., Purcell, R., Remmel, B., Rose, » 

o 
M., Schlueter, T., Simoes, N., Tierrez, A., Vazquez-Boland, J.A., Voss, H., g 

Wehland, J., and Cossart, P. (2001). Comparative genomics of Listeria species. 

Science 294, 849-852. 

Glomski, I. J., Gedde, M.M., Tsang, A.W., Swanson, J.A., and Portnoy, D.A. (2002). 
The Listeria monocytogenes hemolysin has an acidic pH optimum to compart- 
mentalize activity and prevent damage to infected host cells. J. Cell Biol. 156, 
1029-1038. 

Goebel, W., Kreft, J., and Bockmann, R. (2000). Regulation of virulence genes 
in pathogenic Listeria. In Gram-Positive Pathogens, ed. V.A. Fischetti, R.P. 
Novick, J.J. Ferretti, D.A. Portnoy, and J.I. Rood, pp. 449-506. Washington, 
DC: American Society for Microbiology. 



ffi 



Goldberg, M.B. (2001). Actin-based motility of intracellular microbial pathogens. 

Microbiol Mol Biol Rev. 65, 595-626. 
Goldfine, H. and Knob, C. (1992). Purification and characterization of Listeria 

monocytogenes phosphatidylinositol-specific phospholipase C. Infect. Immun. 

60, 4059-4067. 
Goldfine, H., Johnston, N.C., and Knob, C. (1993). Nonspecific phospholipase 

C of Listeria monocytogenes: activity on phospholipids in Triton X-100-mixed 

micelles and in biological membranes. J. Bacteriol. 175, 4298-4306. 
Goldfine, H., Knob, C., Alford, D., and Bentz, J. (1995). Membrane permeabiliza- 

tion by Listeria monocytogenes phosphatidylinositol-specific phospholipase C 

is independent of phospholipid hydrolysis and cooperative with listeriolysin 

O. Proc. Natl. Acad. Sci. USA 92, 2979-2983. 
Gouin, E., Mengaud, J., and Cossart, P. (1994). The virulence gene cluster of Lis- 
Q teria monocytogenes is also present in Listeria ivanovii, an animal pathogen, 




H 



w and Listeria seeligeri, a nonpathogenic species. Infect. Immun. 62, 3550- 

3553. 






g Gray, M.L. and Killinger, A.H. (1966). Listeria monocytogenes and listeric infec- 

< tions. Bacteriol. Rev. 30, 309-382. 

g Greenberg, J.W., Fischer, W., and Joiner, K.A. (1996). Influence of lipoteichoic 

acid structure on recognition by the macrophage scavenger receptor. Infect. 

o Immun. 64, 3318-3325. 

p\ Greene, S.L. and Freitag, N.E. (2003). Negative regulation of PrfA, the key acti- 

w 

^ vator of Listeria monocytogenes virulence gene expression, is dispensable for 

§ bacterial pathogenesis. Microbiol. 149, 111-120. 

Greiffenberg, L., Goebel, W., Kim, K.S., Weiglein, I., Bubert, A., Engelbrecht, 
F., Stins, M., and Kuhn, M. (1998). Interaction of Listeria monocytogenes with 
human brain microvascular endothelial cells: InlB -dependent invasion, long- 
term intracellular growth, and spread from macrophages to endothelial cells. 
Infect. Immun. 66, 5260-5267. 

Guzman, C.A., Rohde, M., Chakraborty, T., Domann, E., Hudel, M., Wehland, J., 
and Timmis, K.N. (1995). Interaction of Listeria monocytogenes with mouse 
dendritic cells. Infect. Immun. 63, 3665-3673. 

Hanawa, T., Yamamoto, T., and Kamiya, S. (1995). Listeria monocytogenes can 
grow in macrophages without the aid of proteins induced by environmental 
stresses. Infect. Immun. 63, 4595-4599. 

Hartford, T., O'Brien, S., Andrew, P.W., Jones, D., and Roberts, I.S. (1993). Uti- 
lization of transferrin-bound iron by Listeria monocytogenes. FEMS Microbiol 
Lett. 108, 311-318. 

Hauf, N., Goebel, W., Serfling, E., and Kuhn, M. (1994). Listeria monocytogenes 



infection enhances transcription factor NF -kappa B in P388Di macrophage- 

like cells. Infect. Immun. 62, 2740-2747. 
Havell, E.A., Beretich, G.R. Jr., and Carter, P.B. (1999). The mucosal phase of 

Listeria infection. Immunobiology 201, 164-177. 
Hof, H. (2001). Listeria monocytogenes: a causative agent of gastroenteritis? Eur. J. 

Clin. Microbiol. Infect. Dis. 20, 369-373. 
Ireton, K., Payrastre, B., Chap, H., Ogawa, W., Sakaue, H., Kasuga, M., and 

Cossart, P. (1996). A role for phosphoinositide 3 -kinase in bacterial invasion. 

Science 274, 780-782. 
Ishiguro, T., Naito, M., Yamamoto, T., Hasegawa, G., Gejyo, F., Mitsuyama, M., 

Suzuki, H., and Kodama, T. (2001). Role of macrophage scavenger receptors 

in response to Listeria monocytogenes infection in mice. Am. J. Pathol. 158, 

179-188. 
Iwamoto, M., Ohno-Iwashita, Y., and Ando, S. (1990). Effect of isolated C-terminal 

fragment of theta-toxin (perfringolysin O) on toxin assembly and membrane £j 

lysis. Eur. J. Biochem. 194, 25-31. fe 

Johansson, J., Mandin, P., Renzoni, A., Chiaruttini, C, Springer, M., and Cossart, § 




H 



O 



P. (2002). An RNA thermosensor controls expression of virulence genes in Q 

Listeria monocytogenes. Cell 110, 551-561. § 

Jones, S. and Portnoy, D.A. (1994). Characterization of Listeria monocytogenes g 

pathogenesis in a strain expressing perfringolysin O in place of listeriolysin g 

O. Infect. Immun. 62, 5608-5613. $ 

Jones, S., Preiter, K., and Portnoy, D.A. (1996). Conversion of an extracellular § 

cytolysin into a phagosome-specific lysin which supports the growth of an £j 

o 
intracellular pathogen. Mol. Microbiol. 21, 1219-1225. 

Jonquieres, R., Bierne, H., Fiedler, F., Gounon, P., and Cossart, P. (1999). Interac- 3 

tion between the protein InlB of Listeria monocytogenes and lipoteichoic acid: w 

a novel mechanism of protein association at the surface of gram-positive 5 

bacteria. Mol. Microbiol. 34, 902-914. » 

o 

Jonquieres, R., Pizarro-Cerda, J., and Cossart, P. (2001). Synergy between the N- g 

and C-terminal domains of InlB for efficient invasion of non-phagocytic cells 
by Listeria monocytogenes. Mol. Microbiol. 42, 955-965. 

Kathariou, S., Metz, P., Hof, H., and Goebel, W. (1987). Tn916-induced mutations 
in the hemolysin determinant affecting virulence of Listeria monocytogenes. 
J. Bacteriol. 169, 1291-1297. 

Kirk, J. (1993). Diagnostic ultrastructure of Listeria monocytogenes in human cen- 
tral nervous tissue. Ultrastruct. Pathol. 17, 583-592. 

Klarsfeld, A.D., Goossens, P.L., and Cossart, P. (1994). Five Listeria monocyto- 
genes genes preferentially expressed in infected mammalian cells: plcA, purH, 



x 




purD, pyrE and an arginine ABC transporter gene, arpj. Mol. Microbiol. 13, 

585-597. 
Klatt, E.C., Pavlova, Z., Teberg, A.J., and Yonekura, M.L. (1986). Epidemic peri- 
natal listeriosis at autopsy. Hum. Pathol. 17, 1278-1281. 
Kocks, C, Gouin, E., Tabouret, M., Berche, P., Ohayon, H., and Cossart, P. (1992). 
L. monocytogenes-induced actin assembly requires the actA gene product, a 
surface protein. Cell 68, 521-531. 
Kocks, C, Marchand, J.B., Gouin, E., d'Hauteville, H., Sansonetti, P. J., Carlier, 
M.F., and Cossart, P. (1995). The unrelated surface proteins ActA of Listeria 
monocytogenes and IcsA of Shigella flexneri are sufficient to confer actin-based 
motility on Listeria innocua and Escherichia coli respectively. Mol. Microbiol. 
18, 413-423. 
Kohda, C, Kawamura, I., Baba, H., Nomura, T., Ito, Y., Kimoto, T., Watanabe, I., 
^ and Mitsuyama, M. (2002). Dissociated linkage of cytokine -inducing activity 

m and cytotoxicity to different domains of listeriolysin O from Listeria mono- 

* cytogenes. Infect. Immun. 70, 1334-1341. 

g Kolb, A., Busby, S., Buc, H., Garges, S., and Adhya, S. (1993). Transcriptional 

< regulation by cAMP and its receptor protein. Anna. Rev. Biochem. 62, 749- 

| 795. 

Kolb-Maurer, A., Gentschev, I., Fries, H.W., Fiedler, F., Brocker, E.B., Kampgen, 

o E., and Goebel, W. (2000). Listeria monocytogenes-infected human dendritic 

^ cells: uptake and host cell response. Infect. Immun. 68, 3680-3688. 

^ Kreft, J. and Vazquez-Boland, J.A. (2001). Regulation of virulence genes in Listeria. 

g Int. J. Med. Microbiol. 291, 145-157. 

Kuhn, M. and Goebel, W. (1989). Identification of an extracellular protein of Lis- 
teria monocytogenes possibly involved in intracellular uptake by mammalian 
cells. Infect. Immun. 57, 55-61. 
Kuhn, M. and Goebel, W. (1998). Host cell signaling during Listeria monocytogenes 

infection. Trends Microbiol. 6, 11-15. 
Kuhn, M. and Goebel, W. (2000). Internalization of Listeria monocytogenes by 
nonprofessional and professional phagocytes. Subcell Biochem. 33, 411-436. 
Lammerding, A.M. and Doyle, M.P. (1990). Stability of Listeria monocytogenes by 
non-thermal processing conditions. In Foodborne Listeriosis, ed. A.J. Miller, 
J.L. Smith, and G.A. Somkuti, pp. 195-202. New York: Elsevier. 
Lampidis, R., Gross, R., Sokolovic, Z., Goebel, W., and Kreft, J. (1994). The viru- 
lence regulator protein of Listeria ivanovii is highly homologous to PrfA from 
Listeria monocytogenes and both belong to the Crp-Fnr family of transcription 
regulators. Mol. Microbiol. 13, 141-151. 
Lasa, I., David, V., Gouin, E., Marchand, J.B., and Cossart, P. (1995). The 
amino -terminal part of ActA is critical for the actin-based motility of Listeria 




monocytogenes; the central proline-rich region acts as a stimulator. Mol. Mi- 
crobiol 18, 425-436. 

Lasa, I., Gouin, E., Goethals, M., Vancompernolle, K., David, V., Vandekerckhove, 
J., and Cossart, P. (1997). Identification of two regions in the N-terminal 
domain of ActA involved in the actin comet tail formation by Listeria mono- 
cytogenes. EMBOJ. 16, 1531-1540. 

Lauer, P., Theriot, J.A., Skoble, J., Welch, M.D., and Portnoy, D.A. (2001). Sys- 
tematic mutational analysis of the amino -terminal domain of the Listeria 
monocytogenes ActA protein reveals novel functions in actin-based motility. 
Mol. Microbiol. 42, 1163-1177. 

Lauer, P., Chow, M.Y., Loessner, M.J., Portnoy, D.A., and Calendar, R. (2002). 
Construction, characterization, and use of two Listeria monocytogenes site- 
specific phage integration vectors. J. Bacteriol. 184, 4177-4186. 

Lecuit, M., Ohayon, H., Braun, L., Mengaud, J., and Cossart, P. (1997). Internalin 

of Listeria monocytogenes with an intact leucine-rich repeat region is sufficient £j 

to promote internalization. Infect. Immun. 65, 5309-5319. ^ 

Lecuit, M., Dramsi, S., Gottardi, C, Fedor-Chaiken, M., Gumbiner, B., and § 

o 
Cossart, P. (1999). A single amino acid in E-cadherin responsible for host Q 

specificity towards the human pathogen Listeria monocytogenes. EMBO J. 18, 

3956-3963. 

Leimeister-Wachter, M., Goebel, W., and Chakraborty, T. (1989). Mutations af- 
fecting hemolysin production in Listeria monocytogenes located outside the > 
listeriolysin gene. FEMS Microbiol. Lett. 53, 23-29. § 

Leimeister-Wachter, M., Haffner, C, Domann, E., Goebel, W., and Chakraborty, jj 

T. (1990). Identification of a gene that positively regulates expression of liste- 
riolysin, the major virulence factor of Listeria monocytogenes. Proc. Natl. Acad. 3 

Sci. USA 87, 8336-8340. 2 

Leimeister-Wachter, M., Domann, E., and Chakraborty, T. (1991). Detection of 5 

a gene encoding a phosphatidylinositol-specific phospholipase C that is co- » 

o 

ordinately expressed with listeriolysin in Listeria monocytogenes. Mol. Micro- g 

biol. 5, 361-366. 
Leimeister-Wachter, M., Domann, E., and Chakraborty, T. (1992). The expression 

of virulence genes in Listeria monocytogenes is thermoregulated. J. Bacteriol. 

174, 947-952. 
Lessing, M.P., Curtis, G.D., and Bowler, I.C. (1994). Listeria ivanovii infection. J. 

Infect. 29, 230-231. 
Lety, M.A., Frehel, C, Dubail, I., Beretti, J.L., Kayal, S., Berche, P., and Charbit, 

A. (2001). Identification of a PEST-like motif in listeriolysin O required for 

phagosomal escape and for virulence in Listeria monocytogenes. Mol. Microbiol. 

39, 1124-1139. 



H 



O 

hi 



X 




Lingnau, A., Domann, E., Hudel, M., Bock, M., Nichterlein, T., Wehland, J., and 
Chakraborty, T. (1995). Expression of the Listeria monocytogenes EGD inlA 
and inlB genes, whose products mediate bacterial entry into tissue culture 
cell lines, by PrfA-dependent and -independent mechanisms. Infect. Immun. 
63, 3896-3903. 
Lingnau, A., Chakraborty, T., Niebuhr, K., Domann, E., and Wehland, J. (1996). 
Identification and purification of novel internalin-related proteins in Listeria 
monocytogenes and Listeria ivanovii. Infect. Immun. 64, 1002-1006. 
Loisel, T.P., Boujemaa, R., Pantaloni, D., and Carlier, M.F. (1999). Reconstitution 
of actin-based motility of Listeria and Shigella using pure proteins. Nature 
401, 613-616. 
Lorber, B. (1997). Listeriosis. Clin. Infect. Dis. 24, 1-9. 

Lou, Y. andYousef, A.E. (1999). Characteristics of Listeria monocytogenes important 
u to food processors. In Listeria, Listeriosis, and Food safety, ed. E.T. Ryser and 

S E.H. Marth, pp. 131-224. New York: Marcel Dekker. 

* MacDonald, T.T. and Carter, P.B. (1980). Cell-mediated immunity to intestinal 

g infection. Infect. Immun. 28, 516-523. 

< Machesky, L.M. and Insall, R.H. (1998). Scarl and the related Wiskott-Aldrich 

g syndrome protein, WASP, regulate the actin cytoskeleton through the Arp2/3 

complex. Curr. Biol. 8, 1347-1356. 

o Maganti, S., Pierce, M.M., Hoffmaster, A., and Rodgers, F.G. (1998). The role 

p\ of sialic acid in opsonin-dependent and opsonin-independent adhesion of 

w 

^ Listeria monocytogenes to murine peritoneal macrophages. Infect. Immun. 66, 

g 620-626. 

Manohar, M., Baumann, D.O., Bos, N.A., and Cebra, J.J. (2001). Gut colonization 

of mice with actA-negative mutant of Listeria monocytogenes can stimulate a 

humoral mucosal immune response. Infect. Immun. 69, 3542-3549. 
Mansfield, B., Dionne, M., Schneider, D., and Freitag, N.E. (2003). Drosophila as 

a model host for Listeria monocytogenes infections. Cell. Microbiol., in press. 
Marchand, J.B., Moreau, P., Paoletti, A., Cossart, P., Carlier, M.F., and Pantaloni, 

D. (1995). Actin-based movement of Listeria monocytogenes: actin assembly 

results from the local maintenance of uncapped filament barbed ends at the 

bacterium surface. J. Cell Biol. 130, 331-343. 
Marco, A.J., Prats, N., Ramos, J.A., Briones, V., Blanco, M., Dominguez, L., and 

Domingo, M. (1992). A microbiological, histopathological and immunohisto- 

logical study of the intragastric inoculation of Listeria monocytogenes in mice. 

J. Comp. Pathol. 107, 1-9. 
Marquis, H., Bouwer, H.G., Hinrichs, D.J., and Portnoy, D.A. (1993). Intracyto- 

plasmic growth and virulence of Listeria monocytogenes auxotrophic mutants. 

Infect. Immun. 61, 3756-3760. 



Marquis, H., Doshi, V., and Portnoy, D.A. (1995). The broad-range phospholipase 
C and a metalloprotease mediate listeriolysin O-independent escape of Lis- 
teria monocytogenes from a primary vacuole in human epithelial cells. Infect. 
Immun. 63, 4531-4534. 

Marquis, H. and Hager, E.J. (2000). pH-regulated activation and release of a 
bacteria-associated phospholipase C during intracellular infection by Listeria 
monocytogenes. Mol. Microbiol. 35, 289-298. 

McLauchlin, J. (1987). Listeria monocytogenes, recent advances in the taxonomy 
and epidemiology of listeriosis in humans. J. Appl. Bacteriol. 63, 1—11. 

Mengaud, J., Vicente, M.F., Chenevert, J., Pereira, J.M., Geoffroy, C, Gicquel- 
Sanzey, B., Baquero, F., Perez-Diaz, J.C., and Cossart, P. (1988). Expression 
in Escherichia coli and sequence analysis of the listeriolysin O determinant 
of Listeria monocytogenes. Infect. Immun. 56, 766-772. 

Mengaud, J., Braun-Breton, C, and Cossart, P. (1991). Identification of 

phosphatidylinositol-specific phospholipase C activity in Listeria monocyto- £j 

genes: a novel type of virulence factor? Mol. Microbiol. 5, 367-372. fe 

Mengaud, J., Ohayon, H., Gounon, P., Mege, R.M., and Cossart, P. (1996). E- § 




H 



O 



cadherin is the receptor for internalin, a surface protein required for entry Q 

of L. monocytogenes into epithelial cells. Cell 84, 923-932. § 

Miettinen, M.K., Siitonen, A., Heiskanen, P., Haajanen, H., Bjorkroth, K.J., and g 

Korkeala, H.J. (1999). Molecular epidemiology of an outbreak of febrile gas- g 

troenteritis caused by Listeria monocytogenes in cold-smoked rainbow trout. > 

J. Clin. Microbiol. 37, 2358-2360. § 

Moors, M.A., Auerbuch, V., and Portnoy, D.A. (1999). Stability of the Listeria £j 

o 
monocytogenes Act A protein in mammalian cells is regulated by the N-end 

rule pathway. Cell. Microbiol. 1, 249-257. 3 

Murray, E.G.D., Webb, R.E., and Swann, M.B.R. (1926). A disease of rabbits 8 

characterized by a large mononuclear leucocytosis, caused by a hitherto un- 5 

described bacillus Bacterium monocytogenes (n. sp.). J. Pathol. Bacteriol. 29, » 

407-439. § 

Nadon, C.A., Bowen, B.M., Wiedmann, M., and Boor, K.J. (2002). Sigma B con- 
tributes to PrfA-mediated virulence in Listeria monocytogenes. Infect. Immun. 
70, 3948-3952. 

Niebuhr, K., Ebel, F., Frank, R., Reinhard, M., Domann, E., Carl, U.D., Walter, U., 
Gertler, F.B., Wehland, J., and Chakraborty, T. (1997). A novel proline-rich 
motif present in ActA of Listeria monocytogenes and cytoskeletal proteins is 
the ligand for the EVH1 domain, a protein module present in the Ena/VASP 
family. EMBOJ. 16, 5433-5444. 

O'Riordan, M. and Portnoy, D. (2002). The host cytosol: front-line or home front? 
Trends Microbiol. 10, 361. 



ffi 




Otter, A. and Blakemore, W.F. (1989). Observation on the presence of Listeria 

monocytogenes in axons. Acta Microbiol. Hung. 36, 125-131. 
Palmer, M. (2001). The family of thiol-activated, cholesterol-binding cytolysins. 

Toxicon 39, 1681-1689. 
Pandiripally, V.K., Westbrook, D.G., Sunki, G.R., and Bhunia, A.K. (1999). Sur- 
face protein pl04 is involved in adhesion of Listeria monocytogenes to human 
intestinal cell line, Caco-2. J. Med. Microbiol. 48, 117-124. 
Pantaloni, D., Boujemaa, R., Didry, D., Gounon, P., and Carlier, M.F. (2000). The 
Arp2/3 complex branches filament barbed ends: functional antagonism with 
capping proteins. Nat. Cell Biol. 2, 385-391. 
Parida, S.K., Domann, E., Rohde, M., Muller, S., Darji, A., Hain, T., Wehland, 
J., and Chakraborty, T. (1998). Internalin B is essential for adhesion and 
mediates the invasion of Listeria monocytogenes into human endothelial cells. 
o Mol. Microbiol. 28, 81-93. 

m Parkassh, V., Morotti, R.A., Joshi, V., Cartun, R., Rauch, C.A., and West, A.B. 

* (1998). Immunohistochemical detection of Listeria antigen in the placenta 

g in perinatal listeriosis. Int. J. Gynecol. Pathol. 17, 343-350. 

< Payne, S.M. (1993). Iron acquisition in microbial pathogenesis. Trends Microbiol. 

g 1, 66-69. 

Pierce, M.M., Gibson, R.E., and Rodgers, F.G. (1996). Opsonin-independent ad- 

o herence and phagocytosis of Listeria monocytogenes by murine peritoneal 

^ macrophages. J. Med. Microbiol. 45, 258-262. 

^ Portnoy, D.A., Jacks, P.S., and Hinrichs, D.J. (1988). Role of hemolysin for the 

§ intracellular growth of Listeria monocytogenes. J. Exp. Med. 167, 1459-1471. 

Portnoy, D.A., Chakraborty, T., Goebel, W., and Cossart, P. (1992). Molecular de- 
terminants of Listeria monocytogenes pathogenesis. Infect. Immun. 60, 1263- 

1267. 

Poyart, C., Abachin, E., Razafimanantsoa, I., and Berche, P. (1993). The zinc 
metalloprotease of Listeria monocytogenes is required for maturation of phos- 
phatidylcholine phospholipase C: direct evidence obtained by gene comple- 
mentation. Infect. Immun. 61, 1576-1580. 

Pron, B., Boumaila, C., Jaubert, F., Sarnacki, S., Monnet, J. P., Berche, P., and 
Gaillard, J.L. (1998). Comprehensive study of the intestinal stage of listeriosis 
in a rat ligated ileal loop system. Infect. Immun. 66, 747-755. 

Pron, B., Boumaila, C, Jaubert, F., Berche, P., Milon, G., Geissmann, F., and 
Gaillard, J.L. (2001). Dendritic cells are early cellular targets of Listeria mono- 
cytogenes after intestinal delivery and are involved in bacterial spread in the 
host. Cell. Microbiol. 3, 331-340. 

Raffelsbauer, D., Bubert, A., Engelbrecht, F., Scheinpflug, J., Simm, A., Hess, J., 
Kaufmann, S.H., and Goebel, W. (1998). The gene cluster inlC 2D E of Listeria 



monocytogenes contains additional new internalin genes and is important for 
virulence in mice. Mol. Gen. Genet. 260, 144-158. 

Ramage, C.P., Low, J.C., McLauchlin, J., and Donachie, W. (1999). Characteri- 
sation of Listeria ivanovii isolates from the UK using pulsed-field gel elec- 
trophoresis. FEMS Microbiol. Lett. 170, 349-353. 

Raybourne, R.B. and Bunning, V.K. (1994). Bacterium -host cell interactions at 
the cellular level: fluorescent labeling of bacteria and analysis of short-term 
bacterium-phagocyte interaction by flow cytometry. Infect. Immun. 62, 665- 
672. 

Reinhard, M., Jouvenal, K., Tripier, D., and Walter, U. (1995). Identification, pu- 
rification, and characterization of a zyxin-related protein that binds the focal 
adhesion and microfilament protein VASP (vasodilator-stimulated phospho- 
protein). Proc. Natl. Acad. Sci. USA 92, 7956-7960. 

Ripio, M.T., Dominguez-Bernal, G., Lara, M., Suarez, M., and Vazquez-Boland, 

J. A. (1997). A Glyl45Ser substitution in the transcriptional activator PrfA £j 

causes constitutive overexpression of virulence factors in Listeria monocyto- ^ 

genes. J. Bacteriol. 179, 1533-1540. § 



dence for expressional crosstalk between the central virulence regulator PrfA 




H 



O 



Ripio, M.T., Vazquez-Boland, J.A., Vega, Y., Nair, S., and Berche, P. (1998). Evi- Q 



o 
and the stress response mediator ClpC in Listeria monocytogenes. FEMS Mi- g 






crobiol. Lett. 158, 45-50. 
Rocourt, J., Hof, H., Schrettenbrunner, A., Malinverni, R., and Bille, J. (1986). > 

Acute purulent Listeria seeligeri meningitis in an immunocompetent adult. § 

Schweiz. Med. Wochenschr. 116, 248-251. jj 

Rocourt, J. (1996). Risk factors for listeriosis. Food Control 7, 195-202. 
Rogers, H.W., Callery, M.P., Deck, B., and Unanue, E.R. (1996). Listeria mono- 3 

cytogenes induces apoptosis of infected hepatocytes. J. Immunol. 156, 679- w 

684. 






Rossjohn, J., Feil, S.C., McKinstry, W.T., Tweten, R.K., and Parker, M.W. (1997). » 

o 

Structure of a cholesterol -binding, thiol-activated cytolysin and a model of its g 

membrane form. Cell 89, 685-692. h 

Ryser, E.T. (1999). Foodborne listeriosis. In Listeria, Listeriosis, and Food Safety, 

ed. E.T. Ryser and E.H. Marth, pp. 299-358. New York: Marcel Dekker. 
Salamina, G., Dalle Donne, E., Niccolini, A., Poda, G., Cesaroni, D., Bucci, M., 

Fini, R., Maldini, M., Schuchat, A., Swaminathan, B., Bibb, W., Rocourt, J., 

Binkin, N., and Salmaso, S. (1996). A foodborne outbreak of gastroenteritis 

involving Listeria monocytogenes. Epidemiol. Infect. 117, 429-436. 
Sawyer, R.T., Drevets, D.A., Campbell, P.A., and Potter, T.A. (1996). Internalin 

A can mediate phagocytosis of Listeria monocytogenes by mouse macrophage 

cell lines. J. Leukoc. Biol. 60, 603-610. 



Schlech, W.F., Lavigne, P.M., Bortolussi, R.A., Allen, A.C., Haldane, E.V., Wort, 
A.J., Hightower, A.W., Johnson, S.E., King, S.H., Nicholls, E.S., and Broome, 
C.V. (1983). Epidemic listeriosis -evidence for transmission by food. N. Engl. 
J. Med. 308, 203-206. 

Schluter, D., Buck, C, Reiter, S., Meyer, T., Hof, H., and Deckert-Schluter, M. 
(1999). Immune reactions to Listeria monocytogenes in the brain. Immunobi- 
ology 201, 188-195. 

Schuchat, A., Swaminathan, B., and Broome, C.V. (1991). Epidemiology of human 
listeriosis. Clin. Microbiol. Rev. 4, 169-183. 

Sechi, A.S., Wehland, J., and Small, J.V. (1997). The isolated comet tail pseu- 
dopodium of Listeria monocytogenes: a tail of two actin filament populations, 
long and axial and short and random. J. Cell Biol. 137, 155-167. 

Sekino-Suzuki, N., Nakamura, M., Mitsui, K.I., and Ohno-Iwashita, Y. (1996). 
Q Contribution of individual tryptophan residues to the structure and activ- 




H 



w ity of theta-toxin (perfringolysin O), a cholesterol -binding cytolysin. Eur. J. 






Biochem. 241, 941-947. 



g Sheehan, B., Kocks, C, Dramsi, S., Gouin, E., Klarsfeld, A.D., Mengaud, J., and 

< Cossart, P. (1994). Molecular and genetic determinants of the Listeria mono- 

5 cytogenes infectious process. Curr. Topics Microbiol. 192, 187-216. 

Shen, Y., Naujokas, M., Park, M., and Ireton, K. (2000). InIB -dependent internal- 

o ization of Listeria is mediated by the Met receptor tyrosine kinase. Cell 103, 

* 501-510. 

^ Shetron-Rama, L.M., Marquis, H., Bouwer, H.G., and Freitag, N.E. (2002). Intra- 

§ cellular induction of Listeria monocytogenes actA expression. Infect. Immun. 

70, 1087-1096. 
Shetron-Rama, L.M., Mueller, K., Bravo, J.M., Bouwer, H.G., and Freitag, N.E. 
(2003). Isolation of Listeria monocytogenes mutants with high level in vitro 
expression of host cyto sol-induced gene products. Mol. Microbiol., 48, 1537- 
1551. 
Sibelius, U., Chakraborty, T., Krogel, B., Wolf, J., Rose, F., Schmidt, R., Wehland, 
J., Seeger, W., and Grimminger, F. (1996). The listerial exotoxins lis- 
teriolysin and phosphatidylinositol-specific phospholipase C synergize to 
elicit endothelial cell phosphoinositide metabolism. J. Immunol. 157, 4055- 
4060. 
Skoble, J., Portnoy, D.A., and Welch, M.D. (2000). Three regions within ActA pro- 
mote Arp2/3 complex-mediated actin nucleation and Listeria monocytogenes 
motility./. Cell Biol. 150, 527-538. 
Smith, G.A., Portnoy, D.A., and Theriot, J.A. (1995). Asymmetric distribution of 
the Listeria monocytogenes ActA protein is required and sufficient to direct 
actin-based motility. Mol. Microbiol. 17, 945-951. 




Smith, G.A., Theriot, J.A., and Portnoy, D.A. (1996). The tandem repeat domain in 
the Listeria monocytogenes Act A protein controls the rate of actin-based motil- 
ity, the percentage of moving bacteria, and the localization of vasodilator- 
stimulated phosphoprotein and profilin. J. Cell Biol. 135, 647-660. 

Spiro, S. and Guest, J.R. (1990). FNR and its role in oxygen-regulated gene ex- 
pression in Escherichia coli. FEMS Microbiol. Rev. 6, 399-428. 

Stachowiak, R. and Bielecki, J. (2001). Contribution of hemolysin and phospho- 
lipase activity to cytolytic properties and viability of Listeria monocytogenes. 
Acta Microbiol. Pol. 50, 243-250. 

Stelma, G.N. Jr., Reyes, A.L., Peeler, J.T., Francis, D.W., Hunt, J.M., Spaulding, 
P.L., Johnson, C.H., and Lovett, J. (1987). Pathogenicity test for Listeria mono- 
cytogenes using immunocompromised mice. J. Clin. Microbiol. 25, 2085-2089. 

Suarez, M., Gonzalez-Zorn, B., Vega, Y., Chico-Calero, I., and Vazquez-Boland, 
J.A. (2001). A role for ActA in epithelial cell invasion by Listeria monocytogenes. 
Cell. Microbiol. 3, 853-864. g 

Sun, A.N., Camilli, A., and Portnoy, D.A. (1990). Isolation of Listeria monocyto- ^ 

genes small-plaque mutants defective for intracellular growth and cell-to-cell § 

o 
spread. Infect. Immun. 58, 3770-3778. Q 

Swanson, J.A. and Baer, S.C. (1995). Phagocytosis by zippers and triggers. Trends 
Cell Biol. 5, 89-93. 

Sword, C.P. (1966). Mechanisms of pathogenesis in Listeria monocytogenes infec- 
tion. I. Influence of iron. J. Bacteriol. 92, 536-542. > 

Tang, P., Sutherland, C.L., Gold, M.R., and Finlay, B.B. (1998). Listeria monocyto- § 

genes invasion of epithelial cells requires the MEK-l/ERK-2 mitogen-activated jj 

a 
protein kinase pathway. Infect. Immun. 66, 1106-1112. 

Theriot, J.A., Mitchison, T.J., Tilney, L.G., and Portnoy, D.A. (1992). The rate of * 

actin-based motility of intracellular Listeria monocytogenes equals the rate of w 

actin polymerization. Nature 357, 257-260. 3 

Theriot, J.A., Rosenblatt, J., Portnoy, D.A., Goldschmidt-Clermont, P. J., and » 

o 

Mitchison, T.J. (1994). Involvement of profilin in the actin-based motil- g 

ity of L. monocytogenes in cells and in cell-free extracts. Cell 76, 505- 

517. 
Tilney, L.G. and Portnoy, D.A. (1989). Actin filaments and the growth, movement, 

and spread of the intracellular bacterial parasite, Listeria monocytogenes. J. Cell 

Biol. 109, 1597-1608. 
Valenti, P., Greco, R., Pitari, G., Rossi, P., Ajello, M., Melino, G., and Antonini, G. 

(1999). Apoptosis of Caco-2 intestinal cells invaded by Listeria monocytogenes: 

protective effect of lactoferrin. Exp. Cell Res. 250, 197-202. 
Vazquez-Boland, J.A., Kocks, C., Dramsi, S., Ohayon, H., Geoffroy, C., Mengaud, 

J., and Cossart, P. (1992). Nucleotide sequence of the lecithinase operon of 



H 



O 

hi 



X 




Listeria monocytogenes and possible role of lecithinase in cell-to-cell spread. 

Infect. Immun. 60, 219-230. 
Vazquez-Boland, J.A., Dominguez-Bernal, G., Gonzalez-Zorn, B., Kreft, J., and 

Goebel, W. (2001a). Pathogenicity islands and virulence evolution in Listeria. 

Microbes Infect. 3, 571-584. 
Vazquez-Boland, J.A., Kuhn, M., Berche, P., Chakraborty, T., Dominguez-Bernal, 

G., Goebel, W., Gonzalez-Zorn, B., Wehland, J., and Kreft, J. (2001b). Listeria 

pathogenesis and molecular virulence determinants. Clin. Microbiol. Rev. 14, 

584-640. 
Wadsworth, S.J. and Goldfine, H. (2002). Mobilization of protein kinase C in 

macrophages induced by Listeria monocytogenes affects its internalization and 

escape from the phagosome. Infect. Immun. 70, 4650-4660. 
Welch, M.D., Iwamatsu, A., and Mitchison, T.J. (1997). Actin polymerization is 
^ induced by Arp2/3 protein complex at the surface of Listeria monocytogenes. 

S Nature 385, 265-269. 

* Wesley, I.V. (1999). Listeriosis in animals. In Listeria, Listeriosis, and Food Safety, 

g ed. E.T. Ryser and E.H. Marth, pp. 39-73. New York: Marcel Dekker. 

< Williams, J.R., Thayyullathil, C., and Freitag, N.E. (2000). Sequence variations 

g within PrfA DNA binding sites and effects on Listeria monocytogenes virulence 

gene expression. J. Bacteriol. 182, 837-841. 

o Wilson, R.L., Tvinnereim, A.R., Jones, B.D., and Harty, J.T. (2001). Identification 

pi of Listeria monocytogenes in in vivo-induced genes by fluorescence-activated 

w 

^ cell sorting. Infect. Immun. 69, 5016-5024. 

g Wuenscher, M.D., Kohler, S., Bubert, A., Gerike, U., and Goebel, W. (1993). The 

iap gene of Listeria monocytogenes is essential for cell viability, and its gene 

product, p60, has bacteriolytic activity. J. Bacteriol. 175, 3491-3501. 

Zalevsky, J., Grigorova, I., and Mullins, R.D. (2001). Activation of the Arp2/3 

complex by the Listeria ActA protein. ActA binds two actin monomers and 

three subunits of the Arp2/3 complex. J. Biol. Chem. 276, 3468-3475. 



CHAPTER 7 



A/, gonorrhoeae: The varying mechanism of 
pathogenesis in males and females 

Jennifer L. Edwards, Hillery A. Harvey, and Michael A. Apicella 



Neisseria gonorrhoeae, the gonococcus, is the causative agent of gonorrhea, 
one of the oldest human diseases on record. Biblical references to gonorrhea 
in Leviticus (15:1-15:19) showed that the infectious nature of the disease was 
recognized even at that time. Probably, the best description of gonorrhea in 
a man in the preantibiotic era can be found in the writings of Bos well, who 
described in detail each of his 19 episodes of infection (Ober, 1970). These 
descriptions also allude to the asymptomatic nature of the disease in women, 
as many of the contacts from whom he acquired the infection were without 
symptoms of disease. Today, it is estimated that greater than 1 million cases of 
N. gonorrhoeae infection occur in the United States, and 60 million cases 
are reported annually worldwide. Hence, N. gonorrhoeae infection remains 
prevalent in the general population, despite the fact that antibiotic therapy is 
readily available. The high incidence of this disease remains a major concern 
in lower socioeconomic groups in the United States; however, the highest 
incidence of infection and of complications resulting from infection occur 
in developing countries. In underdeveloped nations it has been shown that 
patients with gonorrhea are at much higher risk for contracting human im- 
munodeficiency virus (HIV). It is proposed that this increased susceptibility 
to HIV infection results from the inflammatory response generated by infect- 
ing gonococci with the subsequent disruption and shedding of the mucosal 
epithelium. 

The gonococcus was classically considered an extracellular pathogen that 
colonized the epithelial surfaces of the human genital tract. This was in spite 
of the fact that over the past three decades the pathogenesis of gonococci in 
eukaryotic cells has been studied extensively in patient exudates and in tissue 
and organ culture. Electron microscopic studies by Evans (1977) of biopsies 
of infected females show that N. gonorrhoeae colonize and invade epithelial 





1-1 
w 

u 

I— I 

Pn 

< 

< 

h-l 
w 

< 
u 



w 

s 

w 
1-1 
i-i 
i— i 

X 

CO 

Q 

$ 

Q 
w 

i-j 

Pi 
w 

Ph 
i— i 

w 




Figure 7.1. A cross section of a biopsy derived from the squamocolumnar junction of the 
uterine cervix of an infected female demonstrates the invasive nature of N. gonorrhoeae. 
Extensive cytoskeletal rearrangements (CR; e.g., filopodia, lamellipodia, and ruffles) of the 
mucosal epithelium are elicited by gonococcal infection and invasion and are visible on 
the apical cell surface. Gonococci (GC) are internalized in spacious vacuoles, which are 
denoted by arrows. 



cells of the cervical squamocolumnar junction (Fig. 7.1). Studies by Apicella 
and coworkers (1996) show that urethral epithelial cells in exudates from 
infected men contain numerous intracellular organisms (Fig. 7.2). It is now 
understood that the ability of this organism to gain entry into host cells is 
crucial to its capability to establish and sustain infection. 

The gonococcus can infect a number of different epithelial surfaces, 
thereby causing an array of clinical syndromes. Anorectal and pharyngeal 
infection usually results in an asymptomatic or a subclinical disease state. 
Conjunctival infection generally results in a severe conjunctivitis with corneal 
ulceration, although mild disease is occasionally reported. The gonococcus 
can also cause disseminated infection. However, disseminated gonococcal 
infection (DGI) occurs less frequently (1-3%) in both men and women. DGI 
results from gonococcal bacteremia and is commonly associated with un- 
treated asymptomatic gonococcal infection or, less frequently (less than 3%), 




Figure 7.2. An exfoliated urethral epithelial cell obtained from the exudate of a male with 
documented gonococcal urethritis. Gonococci (GC) are labeled with a gold-bead 
antibody-conjugate that is specific for lipooligosaccharide. Arrows denote the intimate 
association of N. gonorrhoeae with the apical surface of the urethral epithelial cell and 
pedestal formation. 



with complement (C) deficiency. The majority of DGIs are found in women, 
in which the onset of clinical symptoms is often associated with menses. 
Arthritis-dermatitis syndrome is the most common manifestation of DGI, 
but severe cases can lead to endocarditis and meningitis. 

In men, uncomplicated gonococcal infection is usually presented as an 
acute urethritis but can result in an acute epididymitis. Symptoms of disease 
primarily result as the consequence of the inflammatory response directed 
at the invasive gonococcus. However, experimental gonococcal infections in 
male human volunteers show that there is a delay between infection and 




o 

% 

o 
to 

m 

o 

B 

ft 

H 

X 
w 

< 
> 

o 

w 
n 

X 

> 

% 
i— i 

X 
o 

$ 

o 

o 

M 

2! 
M 

en 



> 
W 

> 

z 

o 

M 

> 

M 

00 




1-1 
w 

u 

I— I 

Pn 

< 

< 

w 

< 
u 



w 

s 

w 
1-1 
i-i 
i— i 

X 

CO 

Q 

$ 

Q 
w 

i-j 

Pi 
w 

Ph 
i— i 

w 




Figure 7.3. A 30-min experimental infection of a primary urethral epithelial cell with N. 
gonorrhoeae demonstrates filopodia extension induced by the gonococcus. The diplococcus 
arrangement, characteristic of N. gonorrhoeae, is visible at the upper left. Arrows denote 
filopodia engulfing bacteria. 



the onset of clinical symptoms (Schneider et al., 1995). During this time, 
gonococci cannot be cultured from the urethra. Microscopy studies of ure- 
thral exudates from men with gonorrhea demonstrate that N. gonorrhoeae 
is an intracellular pathogen (Apicella et al., 1996). These studies also re- 
veal that intracellular invasion by the gonococcus occurs through the inti- 
mate association between the gonococcus and the urethral cell membrane 
(Fig. 7.3), followed by internalization of gonococci within vacuoles. These ob- 
servations, in conjunction with data obtained from experimental infections 
of men, suggest that viable gonococci are later released from the epithelial 
cells back into the lumen of the urethra. The organism's ability to gain ac- 
cess to and to survive the intracellular environment and the mechanisms 
that allow its subsequent release to the extracellular environment are poorly 
understood. 

If left untreated, gonococcal infection of most men will resolve over a pe- 
riod of several weeks, but resolution may be mistaken for an asymptomatic 
carrier state. The occurrence of asymptomatic infection in men, however, is 
very low. Less than 3% of (infected) men exhibit asymptomatic gonococcal 
infection. In striking contrast to gonococcal infection of men, 50-80% of 
women exhibit asymptomatic genital tract infection, and 70-90% of women 



with DGI lack symptoms of genital tract involvement (Densen et al., 1982). 
Additionally, many women exhibit subclinical manifestations of N. gonor- 
rhoeae infection; consequently, medical treatment is often not sought, or 
gonococcal infection is undiagnosed. 

The high incidence of asymptomatic gonorrhea in women contributes 
substantially to the prevalence of N. gonorrhoeae in the general population 
and leads to acute and chronic complications in these women. Women with 
gonorrhea often exhibit multiple sites of N. gonorrhoeae colonization, which 
is usually attributed to contamination by gonococci-containing cervical secre- 
tions. Infection in women is further compounded by coinfection with other 
sexually transmittable organisms and a number of host physiological factors 
that, cooperatively, complicate disease (Bolan et al., 1999). Of the women 
who exhibit symptomatic gonorrhea, the severity of clinical presentation of 
disease is widely variable. Ascending gonococcal infection occurs in up to 
45% of infected women and can result in pelvic inflammatory disease (PID), g 

which can cause permanent fallopian tube scarring and blockage with sub- § 



o 



M 



sequent infertility and ectopic pregnancies. One in 10 women suffer from § 



o 



PID (Aral et al., 1991), of which N. gonorrhoeae is the etiological agent in 40% jj 

of all reported cases (Sweet et al., 1986). Greater than 25% of women with g 

PID will exhibit long-term complications of disease, 20% will suffer chronic < 

> 

pelvic pain, and 20% will become infertile. In contrast, men with a history of % 



* 



gonorrhea very rarely become infertile because of N. gonorrhoeae infection. o 



An additional 10% of women with PID will have an ectopic pregnancy, which m 

is the leading cause of death of women in their first trimester of pregnancy. jj 

Pregnancy is also considered to put a woman at an increased risk for acquir- £ 

ing DGI. The association of DGI with asymptomatic infection is reflected by o 

the disproportionate percentage (80%) of DGIs that occur in women. jg 

As with other bacterial pathogens, the widespread emergence of K 

antibiotic-resistant strains of N. gonorrhoeae has increased the urgency of 2 
vaccine development. The hope of a vaccine, however, requires a greater un- 



en 



in 



derstanding of both the human and the bacterial factors that contribute to a 
diseased state. Previous attempts at vaccine development have met with fail- s 

ure or with limited success. The propensity of gonococcal surface proteins to S 

> 



undergo both phase and antigenic variation has contributed to the ill success § 

of vaccine development. Recent analysis of the N. gonorrhoeae genome has * 

identified 100 gonococcal genes that exhibit phase variation (Snyder et al., > 

2001). Gonococcal infection was more recently shown to elicit only a mild 5 

humoral immune response, and no immunological memory has been ob- 
served. This suggests gonococcal subversion of the immune response by an 
as yet undefined mechanism. These difficulties may potentially be overcome 
by the use of genetically engineered hybrid proteins. 



N. GONORRHOEAE CONSTITUENTS IMPLICATED 
IN PATHOGENESIS 

Pili 

Pili (or fimbriae) are filamentous appendages that extend from the sur- 
face of a bacterial cell. N. gonorrhoeae express the type IV-A class of bacterial 
pili, which are characterized by the presence of a N-methyl-phenylalanine 
amino acid residue at the first position in the mature pilin protein. An individ- 
ual pilus fiber exists as a polymer composed of individual subunits (i.e., pilin) 
of approximately 18-22 kDa. Individual pilin subunits contain conserved and 
semivariable (SV) and hypervariable (HV) regions. Pilus assembly requires 
a repertoire of proteins that are highly conserved between bacterial species. 
Several levels of control regulate the expression of type IV pilus in the gono- 
coccus. Multiple copies of the pilin gene can be found, scattered, throughout 
the gonococcal genome. Usually only one copy (but sometimes two copies), 
< termed pilE (E for expressed), contains a promoter region and, thus, is ex- 

2 pressed. The remaining copies, termed pilS (S for silent), provide genetic 

E material for nonreciprocal recombination events, which leads to antigenic 

g variation between strains of gonococci. 

g Pili are also subject to phase variation that can result from alterna- 

tive cleavage of the N-terminal hydrophobic region of PilE to produce sol- 




i-i 

w 
u 



< 






% uble S-pilin or from the production of aberrant or excessively long (L-)pilin 



< 



K ( Forest and Tainer, 1997). Phase and antigenic variation provide mechanisms 



< 



^ of immune avoidance, which allows the gonococcus to proliferate within a 

S particular microenvironment of its human host. The importance of pili to 

W N. gonorrhoeae colonization of epithelial cells has been demonstrated by sev- 

g eral laboratories, and it is generally accepted that pili facilitate adherence 

^ by allowing the gonococcus to overcome electrostatic repulsion that occurs 

Q 

m . with the host cell. PilC, a pilus-associated protein, and which is also associated 

g with the bacterial outer membrane (Rahman et al., 1977) , may facilitate pilus- 

£ mediated adherence as a pilus tip-adhesin. Phase variation of PilC expression 

w occurs by frameshift mutations within a polyguanine region upstream of the 

two alleles (pilCl and pi\C2) encoding PilC (Dehio et al., 2000). 

The absence of pilus does not appear to influence adhesion of the gono- 
coccus to polymorphonuclear cells (PMNs), and it is suggested that the pre- 
sence of pili may confer resistance to phagocytosis by these cells (Dilworth 
et al., 1975). Pili also mediate genetic transformation and agglutination, 
and they are postulated to provide a mechanism by which bacteria are 
able to move across and colonize mucosal epithelia through twitching 
motility. The gonococcal pilus is covalently modified with an O-linked, 



N-acetylglucosamine(o;l-3) galactose (Galofl-3GlcNac) disaccharide, which 
(to date) makes it unique among all prokaryotic proteins (Parge et al., 
1995). A trisaccharide, galactose)/? 1-4) galactose^ l-3)2,4-diacetimido-2,4,6- 
trideoxyhexose (Gal/3 l-4Galal-3-DATDH; see Stimsonet al., 1995) is present 
in this same position of meningococcal pilin (Marceau et al., 1998). Removal 
of the meningococcal glycosyl residue confers a hyperadherent phenotype to 
mutant organisms (Marceau et al., 1998), suggesting that pilin glycosylation 
modulates pilin function and its interaction with epithelial cells. The role of 
phosphate (Forest et al., 1999) and phase -variable phosphorylcholine (Weiser 
et al., 1998) posttranslational modifications to pilus function are, currently, 
not known. 

The role of pili in gonococcal pathogenesis is implicated in vivo by ex- 
amination of gonococcal isolates from patients with gonorrhea. In addition, 
multiple in vitro studies have demonstrated the significance of pili in gono- 
coccal adherence to epithelial cells. Complement receptor type 3 (CR3; see g 
Edwards et al., 2001) and the C regulatory protein, CD46 (or membrane cofac- § 



Porin 




o 



M 



tor receptor; see Kallstrom et al., 1997), can serve as receptors for gonococcal § 



o 



pilus. CD46 is a human-specific transmembrane protein that is expressed by jj 

all nucleated cells. The association of pili with CD46 results in a rapid cal- g 

cium influx within the host cell (Kallstrom et al., 1998) . This suggests that pili < 

> 

may modulate host cell signaling mechanisms to aid gonococcal epithelial % 



* 



invasion. However, Tobiason and Seifert (2001) demonstrated that the pres- o 



ence of CD46 on several epithelial cell lines is inversely related to gonococcal m 

association with these cells. Additionally, CD46 does not appear to play a role jj 

in the CR3-mediated initial colonization of the uterine cervix (Edwards et al., K 

2002). Consequently, discernment of the role of CD46 in N. gonorrhoeae col- o 

onization will require further examination. jg 

H 
O 

o 

M 

% 
M 

en 



Porin (P.I; 32-39 kDa) accounts for greater than 60% of the total weight ~ 

of proteins present in the outer membrane of the gonococcus, making it the s 

most abundant outer membrane protein. Gonococcal nomenclature regard- S 

ing porin is based on homology to meningococcal PorB. A gonococcal homo- % 

logue of N. meningitidis PorA was not thought to exist; however, a gene show- * 

ing homology to the PorA gene sequence has recently been identified (Dehio > 

et al., 2000). Unlike pili, opacity-associated outer membrane proteins (Opa), 5 
and lipooligosaccharide (LOS), all of which exhibit phase or antigenic varia- 
tion or both, porin expression is stable within a given strain of N. gonorrhoeae. 
Antigenic variation between strains, however, has provided one system of 




N. gonorrhoeae serotyping. Two isotypes of porin are found in gonococci, P. I A 
and P. IB. Isotypes P.IA and P. IB can be further differentiated into serovars 
(classes 1-9) by their ability to react with a panel of monoclonal antibodies. 
For any N. gonorrhoeae, only one porin isotype is produced; however, anoma- 
lous P. I A/ P. IB hybrid porins have been clinically isolated on rare occasion 
(Cooke et al., 1998). Gonococci exhibiting a P.IA isotype are associated with 
an increased risk for DGI and are thought to exhibit increased resistance to 
killing by normal human serum (i.e., serum-resistance). 

Gonococcal porin resembles porins expressed by other Gram-negative 
bacteria in that it forms a water-filled channel in the bacterial outer mem- 
brane through which small solutes may cross. By analogy to Escherichia coli 
porins, gonococcal porin is suggested to exist in the bacterial outer mem- 
brane as a homotrimer, with each monomer forming a channel. Each porin 
S monomer possesses 16 membrane-spanning regions, resulting in the expo- 

S sure of eight antigenically variable loops in the extracellular environment. The 

< longest of these extracellular loops express the greatest antigenic variability; 

hj they are immunodominant and serve as docking sites for bactericidal anti- 

X bodies. Although porin is highly immunogenic, its intimate association with 

u 

g reduction-modifiable protein (Rmp or P. 1 1 1) and with LOS within the outer 

§ membrane makes it inaccessible to immunological attack in the presence of 

< 

^ LOS- or Rmp -directed blocking antibody. 

% In contrast to other Gram-negative porins, which exhibit cation selec- 

E tivity, gonococcal porins resemble mitochondrial porins in their anion se- 

g lectivity. Another distinguishing property of gonococcal and mitochondrial 

Jj porins is their ability to interact with the adenosine and guanosine triphos- 

e phates (ATP and GTP, respectively) as a means to regulate pore size, voltage- 

g dependent gating, and ion selectivity. Also unique among Gram-negative 

^ porins is the ability of gonococcal porin to translocate into eukaryotic host 

" cell membranes. Within the eukaryotic cell membrane, porin it is thought 

£ to form a voltage-gated channel that is modulated by the host cell ATP and 

I GTP. 

El The role of porin in potentiating disease by N. gonorrhoeae is thought to 

be multifactorial. Porin appears to modulate host cell function. Additional di- 
verse functions of porin suggest that it is an important virulence factor for the 
gonococcus. Studies using PMNs have indicated that porin changes the mem- 
brane potential of these cells upon PMN membrane insertion (Haines et al., 
1988). A change in membrane potential results in degranulation inhibition 
without altering the respiratory burst associated with oxidative killing. In con- 
trast, gonococcal porin increases formyl-methionine-leucine-phenylalanine 
(fM LP) -mediated hydrogen peroxide production in PMN (Haines etal., 1991). 



Further modulation of host cell function includes the ability of porin to in- 
hibit phagosome maturation and downregulation of immunologically impor- 
tant cell surface receptors, such as immunoglobulin G (IgG), Fc receptors II 
(FcyRII) and III (FcyRIII), and complement receptor 1 (CR1) and CR3, in 
professional phagocytic cells. 

Although porin appears to inhibit PMN actin polymerization in response 
to chemoattractants (Bjerknes et al., 1995), in epithelial cells porin acts as an 
actin-nucleating protein (Wen et al., 2000). In this respect porin may facil- 
itate the cytoskeletal rearrangements required for actin-mediated entry of 
the gonococcus into its target host cell. Porin also modulates apoptosis of 
epithelial cells by inducing a calcium influx and, consequently, calpain and 
caspase activity within these cells (Dehio et al., 2000). Recent data have sug- 
gested that the ability of porin to induce apoptosis in epithelial cells may 
play a role in the cytotoxicity observed in fallopian tube organ culture (FTOC) 
and in shedding of epithelial cells (Dehio et al., 2000), which occurs in vivo g 

during mucosal infection. Porin may also aid Opa-heparin sulfate proteo- § 




o 



M 



glycan (H SPG) -mediated invasion and Opa-independent invasion of Chang § 



o 



cells in the absence of phosphate (van Putten et al., 1998). In cooperation with jj 

gonococcus-bound iC3b and gonococcal pilus, porin also acts as an adhesin, g 

mediating adherence to the cervical epithelium (Edwards et al., 2002). 



< 
> 

3 



Opacity-associated outer membrane proteins g 



n 



Opa (P. II or class 5) proteins comprise a family of closely related in- jj 

tegral outer membrane proteins. Opa proteins were originally identified by £ 

their ability to confer opacity and color changes to colonies of N. gonorrhoeae o 

(grown on translucent agar) when viewed under diffused light with a stere- jg 

omicroscope. Subsequent studies revealed that opacity is the result of Opa- K 

LOS interactions that occur between adjacent bacteria. Opa proteins are heat 2 

modifiable, exhibiting altered electrophoretic mobility subsequent to heating 5 

at 1 00° C. Molecular mass (M r ) for individual Opa proteins ranges from 24 kDa ~ 

to 30 kDa at 37°C and from 30 kDa to 32 kDa at 100°C. Gel filtration suggests s 

that purified Opa proteins are trimers or tetramers; however, Opa isolation S 

has proven difficult because of its high (one third) content of hydrophobic % 

o 

amino acids and, thus, its insolubility in the absence of detergents. * 

Structural models predict eight membrane-spanning regions with four > 

membrane loops exposed on the extracellular face of the bacterium. The 5 
first three loops correspond to the SV, HV1, and HV2 regions; the fourth 
(C-terminal) loop is highly conserved. There are 11 or 12 opa loci that oc- 
cur throughout the chromosome of a given N. gonorrhoeae strain. Expression 




of an individual opa gene occurs independently of other opa genes; conse- 
quently, a single gonococcus may exhibit none, one, or more than one Opa 
protein simultaneously. Each opa locus is subject to antigenic and phase 
variation; therefore, a given strain of N. gonorrhoeae may be highly hetero- 
geneous with respect to Opa expression. Opa phase variation results from 
slipped-strand mispairing during replication in which alteration in the num- 
ber of pentameric (CTCTT) repeat elements causes translational frameshifts 
(Lemon and Sparling, 1999). 

A correlation has been made between the presence (and absence) of 
Opa proteins on N. gonorrhoeae clinical isolates with the site of isolation. 
In one human volunteer study, gonococci recovered from males infected with 
Opa - organisms had shifted to Opa + (Schneider etal., 1995). Clinical isolates 
obtained from men tend to express Opa proteins as do isolates recovered from 
3 human volunteer studies . Similarly, cervical isolates obtained from women at 

S the time of ovulation (i.e., midcycle) also exhibit Opa expression. Translucent, 

< Opa - organisms predominate in cervical isolates obtained during menses; in 

hj the fallopian tube; in genital, blood, and joint fluid obtained from patients with 

X DGI; and in asymptomatic men. Opa proteins facilitate adherence to PMNs 

u 

g and, consequently, may play a role in potentiating gonococcal urethritis. A 

§ recent study has proposed a novel role for Opa proteins. Williams et al. (1998) 

< 

>; suggested that Opa proteins may contribute to the intracellular survival of 

% gonococci by binding host cell pyruvate kinase in order to acquire intracellular 

ffi pyruvate, which is required for growth. 

< 

s 

w 

1-1 
1-1 

I— I 

X 

CO 



Lipooligosaccharide 



g LO S is an amphipathic glycolipid that constitutes a significant proportion 

^ of the N. gonorrhoeae outer membranes. In addition to differences in biosyn- 

" thesis genes, the major difference between LOS and lipopolysaccharide (LPS), 

£ which is prevalent among Gram-negative bacteria, is its lack of a repeating 

£ O-antigen. LOS is composed of a lipid A moiety, a biantennary or trianten- 

*L. nary core region composed of two L-glycero-D-manno-heptopyranose (hep- 

tose) and two 2-keto-3-deoxyoctulosonic acid (KDO) residues, and variable 

oligosaccharide side chains (Mandrell and Apicella, 1993; also see Fig. 7.4). 

The lipid A moiety anchors the LOS molecule within the outer membrane 

and consists of a dihexosamine backbone linked to four hydroxymeristic 

acids. These are substituted on the 3' position by lauric acid. KDO molecules 

are juxtaposed to the lipid A and one KDO serves as the site of the dihep- 

tose addition. The heptose molecules serve as docking sites for short (6-10 

sugar moieties) oligosaccharide addition(s). Additionally, the second heptose 




o 



NANA a 2 -> 3GlcNAc pl^ 3Gal pi-> 4Glcl -> 4Hepl->5KD0 2 -> Lipid A 

3 

T 
l 

GlcNAcl -^2Hep 
3 

T 

PEA 

Figure 7.4. A model of the sialylated LOS glycoform from N. gonorrhoeae strain 1291 is 
shown. This glycoform is represented as containing the sialyl-lactosamine found on the 
organism in the presence of CMP-NANA. Over 97% of gonococcal isolates express this 
lactosamine-containing glycoform that is important as a ligand to the human 
asialoglycoprotein receptor. 



(which is not directly liked to KDO) bears an N-acetylglucosamine residue. 

Oligosaccharide substitutions exhibit interstrain and intrastrain variability. g 

Interconversion of LOS oligosaccharides occurs spontaneously and is § 

dependent on the presence or absence of available substrates for, and the en- § 

zymes involved in, LOS biosynthesis. Several genes involved in LOS biosyn- £ 

thesis have been identified. The Igt locus, which encodes the genes for syn- g 

thesis of lacto-N-neotetraose (LNnT) and digalactoside moieties, has recently < 

been described. Several of these genes contain a polyguanine region that, 3 

in a manner similar to PilC phase variation, can result in frameshift mu- o 

tations with concurrent loss (or gain) of certain oligosaccharide moieties m 

available for LOS assembly. Phosphate, phosphoethanolamine or pyrophos- jj 

phoethanolamine, and O -acetyl substitutions of the core region along with the £ 

addition of sialic acid to terminal galactose residues confer further variation o 

to the LOS structure. jg 

LOS oligosaccharide side chains terminate in epitopes that mimic oligo- K 



in 



saccharide moieties of mammalian glycosphingolipids (including P , i, and £ 

paragloboside), regardless of oligosaccharide chain length. This form of mo- 
lecular mimicry not only provides the bacterium with a method of immune 

avoidance but also allows the bacterium to use host-derived molecules that g 

> 

normally associate with the mimicked structure. The presence of a terminal S 

LNnT epitope (Gal^l-4GlcNac^l-3Galygl-4Glc) on LOS, which mimics hu- % 

o 

man paragloboside, allows the gonococcus to invade the urethral epithelium * 

of men by adherence to the asialoglycoprotein receptor (ASGP-R; see Harvey > 

et al., 2001), and this may facilitate disease transmission by adherence to the 5 

ASGP-R on human sperm (Harvey et al., 2000). 

An analysis of N. gonorrhoeae strains demonstrates the predominance 
of the LNnT epitope among gonococci (Campagnari et al., 1990). These 




observations were confirmed in situ in that the LNnT epitope is selected 
for in men in human volunteer studies and in men with naturally acquired 
gonococcal urethritis. Schneider et al. (1995) further demonstrated that se- 
lection of the paragloboside epitope significantly enhances the ability of the 
gonococcus to cause disease in human volunteers. The predominance of a 
paragloboside moiety may enhance gonococcal survival in two ways: first, by 
avoidance of immune recognition by molecular mimicry of the LNnT, and 
second, by an increased serum-resistance by sialylation of the terminal galac- 
tose residue of the LNnT epitope. Thus, a lower infectious dose is required to 
establish disease, because a greater proportion of the inoculum would survive 
and proliferate. 

The prevalence of the paragloboside moiety among N. gonorrhoeae is con- 
sistent with the finding that most clinically isolated gonococci initially exhibit 
S serum-resistance; however, this property is lost with subsequent subculture. 

S This unstable form of serum-resistance is attributed to sialylation of the ter- 

< minal lactosamine of the LNnT moiety present on LOS. LOS and LPS stimu- 

hj late cytokine production, which is associated with cytotoxicity. Stimulation of 

X cytokine production by gonococcal LOS may allow this bacterium to access 

u 

g subepithelial tissues while increased serum-resistance and the predominance 

§ of human-like epitopes on the bacterium surface enhances survival. Sialyla- 

< 

^ tion of the gonococcus surface inhibits its ability to be phagocytosed by PMNs 

% and may increase the survival of those gonococci that are phagocytosed. Sia- 

E lylation of the gonococcus, however, significantly inhibits its ability to invade 

g urethral epithelial cells (Harvey et al., 2001) and to initiate disease in human 

3 volunteer studies (Schneider et al., 1996). 

e Studies that explored the importance of LOS sialylation during infection 

g and that used human volunteers and in vitro tissue culture models led to the 

^ hypothesis that although LOS sialylation may protect extracellular organisms 

" from complement-mediated killing, these organisms become desialylated 

£ prior to internalization by host cells. Immunoelectron microscopy studies 

£ of urethral exudates from men with gonococcal urethritis show that in vivo 

El sialylation of gonococcal LOS Gal/? 1-4-GlcNAc residue occurs during human 

infection and that although the majority of the LOS of intracellular gonococci 

is sialylated, approximately 10% of LNnT-terminal LOS remain unsialylated 

(Apicella et al., 1990). Studies by other groups later showed evidence that 

gonococci with sialylated LOS are less invasive for immortalized tissue culture 

epithelial cells than unsialylated gonococci (van Putten and Robertson, 1995) . 

Conversion (by phase variation) of LOS to epitopes unable to be sialylated 

allows the gonococcus to adapt to environmental changes associated with 




o 



progressive disease. LOS significantly contributes to the pathogenicity of the 
gonococcus in that it not only functions as an adhesin but also facilitates 
disease transmission, modulates bacterial invasion, and modulates disease 
progression. 

Additional virulence factors 

Through coevolution with its sole human host, the gonococcus has devel- 
oped an impressive repertoire of gene products that allow it to exert redundant 
mechanisms by which it is able to infect, colonize, and proliferate within the 
hostile and benign microenvironments of the human body. Completion of 
the N. gonorrhoeae (strain FA1090) genome project and continued research 
will undoubtedly result in the identification of (as yet unknown) additional 
factors that contribute to gonococcal virulence. Several researchers have de- 
scribed factors that may contribute to virulence; however, their actual role in g 
human disease is less well defined. § 

Many mucosal pathogens escape immunological killing by their abil- § 

ity to produce, and secrete, proteases that inactivate IgAl and its secretory £ 

counterpart (slgAl), the principal antibody isotype of the mucosa and mu- g 

cosal secretions. Cleavage of these antibodies within their hinge region yields < 

Fab a monomeric fragments that are separate from their effector Fc a counter- 3 

parts; consequently, they are rendered inactive. Serine, cysteine, and metallo o 

IgA proteases have been identified among those mucosal pathogens that pro- m 

duce these endopeptidases. N. gonorrhoeae, N. meningitidis, and Haemophilus jj 

influenzae produce serine proteases that cleave the IgA hinge region between £ 

proline and serine residues (i.e., type 1 activity) or between proline and thre- o 

onine residues (i.e., type 2 activity). All N. gonorrhoeae and N. meningitidis jg 

strains examined to date have been found to constitutively express either a K 

type 1 or a type 2 IgAl protease. In gonococci the type of IgAl protease pro- 2 

duced correlates with auxotype and the serovar of porin produced. Gonococci 



en 



that possess a nutritional requirement for arginine, hypoxanthine, and uracil « 

(i.e., AHU auxotype) and a P. I A class 1 or 2 porin serovar produce a type 1 s 

IgAl protease. Type 2 IgAl protease activity is prominent in gonococci that S 

are excluded from the AHU/P.IA1/2 serotype. % 

o 

The role of neisserial IgAl protease activity in disease processes is unde- * 

termined. It is generally presumed that inactivation of potentially bactericidal > 

IgAl by a mucosal pathogen potentiates colonization and, therefore, potenti- 5 
ates disease caused by IgA protease-producing microbes. However, infection 
with N. gonorrhoeae IgAl null mutants in human volunteer studies resulted in 




infection and disease indistinguishable from that observed with the parental 
wild-type strain. Analysis of cervical secretions yielded similar findings in 
that the level of IgAl protease activity in samples obtained from women with 
gonococcal cervicitis were comparable with those obtained from women who 
were not infected. These observations suggest that IgAl protease does not 
significantly influence colonization of the (uro) genital epithelia. 

In epithelial cells, cleavage of lysosome-associated membrane proteins 1 
(LAM PI) with concurrent phagosome modulation has been proposed as an 
alternative function for IgAl protease activity. However, LAMP1 of profes- 
sional phagocytes are resistant to (gonococcal) IgAl protease activity. Within 
professional phagocytes LAM PI are heavily glycoslyated; consequently, it is 
thought the hinge region is protected from protease cleavage. These studies 
suggest that IgAl protease activity is limited to epithelial cells. This apparent 
S conundrum pertaining to gonococcal IgAl protease activity has to be exam- 

S ined further before a role for IgAl protease activity in gonococcal pathogen- 

< esis can be resolved. 

hj The ability of pathogenic organisms to sequester iron from their host 

X animal is well established to contribute significantly to a diseased state. The 

u 

g gonococcus is no exception, and gonococci possess multiple mechanisms by 

§ which they are able to scavenge iron (Schryvers and Stojiljkovic, 1999) . Neisse- 

< 

^ via do not produce siderophores; however, they do produce siderophore recep- 

% tors that are homologous to siderophore receptors of other bacterial species. 

E Expression of siderophore receptors allows N. gonorrhoeae to use siderophores 

g produced by other bacterial species with which they may share an ecological 

Jj niche. In addition to siderophore receptors, most pathogenic Neisseria also 

E produce receptors for transferrin (Tf), lactoferrin (Lf), hemoglobin (Hb), 

g heme, and haptoglobin-hemoglobin (Hp-Hb). Lf is found in higher concen- 

^ tration on mucosal surfaces than is Tf and, therefore, was presumed to be the 

" preferred iron source for pathogenic Neisseria spp. However, some gonococci 

£ do not produce receptors for Lf, and mutants defective in Lf acquisition retain 

£ their virulence in human volunteer studies (Cornelissen et al., 1998). 

|z; 

El In contrast, however, gonococci deficient in Tf acquisition are avirulent 

in human volunteer studies (Cornelissen et al., 1998), suggesting that Tf is 
required for gonococcal colonization of (at least) the male urethral epithelium. 
The ability of HmbR (the heme and Hb receptor), HpuAB (the Hb and Hp- 
Hb receptor), and TonB mutants to grow in the presence of heme as a sole 
iron source has presented the idea that heme may passively diffuse across 
the gonococcal outer membranes as a result of its hydrophobic character. 
Alternatively, porin may facilitate heme uptake (Schryvers and Stojiljkovic, 
1999). However, recent data indicate that PilQ may serve a dual function, 




allowing pilin transport (with subsequent pilus assembly) out of, and heme 
passage into, the bacterial cell body (Chen et al., 2002). 

The ability of gonococci to use heme, Hb, and Hp-Hb has been postu- 
lated to be responsible for the increased risk observed in women to develop 
PID and DGI during menses. However, fluctuations in hormone levels and 
associated changes to the female reproductive tract may also be responsible 
for the increased risk of complicated gonococcal infection with menses. En- 
vironmental fluctuations may allow gonococci of a particular isotype to gain 
access to subepithelial tissues or ascend the female genital tract and, thereby, 
initiate DGI or PID, respectively. 

Interaction of the gonococcus with the lutropin receptor (LHr) is sug- 
gested to confer a contact-inducible phenotype to this bacterium that in- 
creases its invasive character, thereby allowing it to invade epithelia by ad- 
herence to the LHr (Spence et al., 1997). Identification of a LHr-gonococcus £ 
interaction involved the use of several immortal cell lines. Because LHr is 2 
upregulated in immortal cells and because ligand binding downregulates § 
LHr surface expression, the significance of a LHr-gonococcus interaction in § 
vivo within the lower female genital tract (e.g., the ectocervix and endocervix > 
and the distal endometrium) remains to be determined. Gorby et al. (1991) g 

demonstrated that gonococci bind to the LHr on nonciliated cells in FTOC, < 

> 

suggesting that the LHr may be paramount to colonization of the fallop- % 

ian epithelia. The presence of the LHr on human uterus, placenta, decidua, o 

fetal membrane, and fallopian tube tissues, in conjunction with increased m 

expression of this receptor during menses, suggests that a gonococcus-LHr jj 

interaction could facilitate ascending infection in women. £ 

s 
Although speculative, a gonococcus-LHr interaction occurring on de- o 

cidua and placental membranes could potentially result in severe compli- jg 

cations and may, in part, contribute to the increased risk of spontaneous K 

abortion associated with N. gonorrhoeae infection. The inhibitory effect of 2 

human chorionic gonadotropin (hCG), a ligand for LHr, on the observed 



en 



in 



LHr-gonococcus interaction suggests that gonococci possess an unidenti- 
fied surface molecule that mimics hCG; however, the gonococcal LHr ligand s 

has yet to be identified. The interaction of the gonococcus with the LHr ap- S 

> 



pears to be a significant factor in colonization of fallopian tube tissue and % 

may pose a serious risk to a developing fetus in pregnant women. * 

Several gonococcal gene products appear to be induced under conditions > 

of limited oxygen. Because the oxygen tension is presumed to be low in 5 

some sites of gonococcal infection (and under certain circumstances), the 
contribution of these gene products, such as AniA (or Panl; see Householder 
et al., 1999), Pan2, and Pan3 (Clark et al., 1987) to progressive infection and 




h-l 



disease cannot be overlooked. Currently, however, there is little or no evidence 
to confirm or negate the contribution of these anaerobically induced proteins 
to the pathogenesis of N. gonorrhoeae in males or females. 

COMPLEMENT AND PATHOGENIC NEISSERIA 

Coevolution of the human C system with potentially pathogenic micro- 
organisms has allowed several human pathogens to adapt mechanisms by 
which they are able to avoid C'-mediated eradication (Vogel and Frosch, 1999) . 
Methods evolved by microorganisms that allow them to escape the potentially 
lethal effects of C activation include the following: (1) development of an an- 
tiphagocytic capsule; (2) the development of surface structures that inhibit 
C3-convertase activity; (3) evasion of the lytic activity of the membrane at- 
tack complex (MAC), mediated by cell surface structures; (4) mimicry of C' 
3 components; and (5) exploitation of C' components to facilitate infection. 

< The pathogenic Neisseria, that is, N. meningitidis and N. gonorrhoeae, are strict 

hj human pathogens that remain prevalent in the general population, in part 

X because of their exquisite capability to avoid host defense mechanisms. In 

§ addition to the generalized defense mechanisms of phase and antigenic varia- 

§ tion, pathogenic Neisseria have evolved highly specialized immune avoidance 

£ mechanisms that allow them to persist within their (sole) human host. 

% N. meningitidis (the meningococcus) produces a capsule composed pri- 

E marily, or entirely, of sialic acid. The presence of an extracellular capsule 

g prohibits the interaction between cell-wall-deposited opsonins and host cells 

Jj by providing a physical barrier impermeable to the phagocyte; that is, it is 

e antiphagocytic. Meningococci of serogroup B possess a sialic acid capsule 

g that is structurally similar to neural cell adhesion molecules (NCAM). This 

^ NCAM-like capsule confers further protection to serogroup B meningococci 

" by rendering these bacteria nonimmunogenic. Sialic acid is widely distributed 

£ upon host cells and thus is recognized as an indigenous (or nonactivating) 

£ surface. In the absence of an effective antibody response (as already noted), 

*L. direct recognition of a potential pathogen becomes critical to bacterial clear- 

ance. Within the alternative pathway (AP) of the C system, the presence of 
sialic acid on a target surface favors binding of factor H (f H) (with C inactiva- 
tion) over binding of factor B (f B), which would otherwise result in activation 
of the C system. 

Although N. gonorrhoeae does not possess a capsule, it does share with 
the meningococcus the ability to sialylate its LOS. Gonococci that bear sialic 
acid moieties on their LOS have been found in cell-free systems to be resis- 
tant to C'-mediated killing; that is, they are serum-resistant. Serum-resistance 



conferred by sialylated LOS is attributed to the ability of LOS to bind a specific 
site within short consensus repeats (SCRs) 16-20 of fH (Ram et al., 1998). 
This results in C inactivation through the cofactor activity of f H in factor I (fl)- 
mediated inactivation of C3b. Because gonococci are not capable of producing 
cytosine S'-monophosphate N-acetylneuraminic acid (CMP-NANA, the pre- 
cursor for sialic acid synthesis), sialylation of a gonococcal LOS requires the 
use of an exogenous (i.e., host derived) substrate. In contrast, N. meningitidis 
is capable of synthesizing its own sialic acid and, therefore, incorporates en- 
dogenous sialic acid into its capsular and LOS structures. Significantly less 
f H is deposited upon capsulated meningococci when compared with strains 
that lack a capsule (Estabrook et al., 1997), suggesting that other factors may 
play a role in fH deposition on encapsulated organisms. An additional role 
for sialic acid has also been elucidated. The presence of sialic acid moieties 
can lead to a decrease in C9 deposition (Jarvis, 1995) upon the gonococcal 
surface, thereby prohibiting MAC assembly and, thus, its associated ability g 

to lyse these cells. § 




o 



M 



Serum resistance of gonococci is also attributed to the expression of one § 



o 



of two isotypes ofporin, PI. A and Pl.B. Isotype PI. A is found more frequently jj 

on gonococci in people with DGI. Expression of PI. A by the gonococcus g 

is attributed to increased serum resistance in comparison with gonococcal < 



> 



strains that express the Pl.B isotype. Although porin is intimately associated % 



* 



with gonococcal LOS, binding ofporin to f H occurs independently of LOS, o 



at a site distinct from that defined for sialic acid (Ram et al., 1998). Within m 

the classical pathway (CP) of the C system, C4bp functions in an analogous jj 

manner to f H of the AP. C4bp does not bind to the surface of sialylated £ 

gonococci; however, C4bp can bind gonococci when it is not sialylated (Ram o 

et al., 2001). Binding of C4bp occurs on gonococci that express a Pl.B isotype. jg 

Several studies have indicated that gonococcal killing is mediated primarily K 

through the action of the CP; consequently, the ability of certain gonococcal 2 
strains to directly bind C4bp, which allows fl-mediated inactivation of C4b 



en 



in 



deposited on the gonococcal surface, may be critical to their survival within 

their human host. g 

> 

In vitro N. gonorrhoeae infection studies and an examination of clini- S 

> 



cally isolated N. gonorrhoeae reveal that inactive C3b (i.e., iC3b, in com- § 

parison with active C3b) is predominant on the surface of this bacterium. * 

iC3b opsonization allows the gonococcus to invade primary cervical epithe- > 

lial cells in a process mediated by the action of gonococcal porin and pilus 5 

(Edwards et al., 2002), and which occurs independently of the sialylation sta- 
tus of the bacterium (Edwards and Apicella, 2002). Because endocytosis that 
is mediated by CR3 occurs independently of a proinflammatory response, 



internalization of the gonococcus within cervical epithelial cells may con- 
tribute to the asymptomatic nature associated with gonococcal infection in 
women. 

The membrane-associated fl co-factor, CD46, may serve as a receptor 
for gonococcal pilus on some cell lines (Kallstrom et al., 1997). Although the 
physiological relevance of a CD46-pilus interaction on the urogenital epithe- 
lium remains controversial, pilus binding to CD46 present on PMNs could 
facilitate fl -mediated inactivation of C3b or C4b on an opsonized bacterium. 
Another family of surface molecules expressed by the pathogenic Neisse- 
ria is the Opa proteins. Opa proteins serve as ligands for members of the car- 
cinoembryonic antigen-related family of cell adhesion molecules (CEACAM 
or CD66). CEACAM molecules can serve as coreceptors for the leukocyte 
integrins, including CR3 and CR4. Interaction of CEACAM with CD18 (the 
S beta subunit shared by CR3 and CR4) is thought to inhibit the respiratory 

S burst that occurs with ligand binding and phagocytosis by PMNs. This may 

< promote the intracellular survival of PMN phagocytosed Neisseria. 

^ An additional role for Opa proteins relies on their ability to bind heparin. 

X Gonococcal binding of heparin by means of their Opa proteins increases the 

u 

g serum-resistance observed by these bacteria (Chen et al., 1995). Although 

§ the mechanism by which serum-resistance is conferred has yet to be defined, 

the ability of heparin to augment the regulatory role of vitronectin in the inhi- 




< 






rt bition of MAC assembly provides one explanation for the observed increased 

ffi serum-resistance. 



g Despite the redundant ability of the pathogenic Neisseria to evade C- 

Jj mediated killing, the C system is still capable of eliciting an efficient an- 

e timicrobial defense. This becomes most evident in individuals with various 

g C deficiencies that are prone to recurrent bacterial infections, most notably 

^ with Neisseria spp. Individuals deficient in C components of the MAC exhibit 

" an increased incidence of neisserial infections. In particular, deficiencies of 

£ C5, C6, C7, and C8 lead to a much greater risk for recurrent, and commonly 

£ systemic, infection with Neisseria spp. Despite the increased risk for systemic 

*L. neisserial infections, a decreased mortality associated with infection is also 

observed and is thought to be attributable to a decrease in endotoxin released 
in the absence of a functional MAC. 

People who are deficient in C9 are also at a much greater risk for Neisseria 
spp. infection. This has led to the suggestion that the bactericidal action of 
normal human serum is critical to controlling Neisseria spp. infection. Sim- 
ilarly, greater than 50% of people who lack the AP components fB, factor D 
(fD), and properdin exhibit recurrent neisserial infections. Studies focusing 
on neisserial serum resistance have primarily involved cell-free systems in 



ular host cell target and microenvironment. 



Infection of the male urethra 




which serum-resistance has been measured as the ability of these organisms 
to survive the lytic action of pooled human serum. However, mucosal mem- 
branes serve as the primary site for neisserial infection. Increasing evidence 
suggests that AP C' components are produced by epithelial cells, albeit at a 
much lower level than is observed in human serum. Therefore, the ability 
of the pathogenic Neisseria to survive within their human host may reside in 
their ability to evade AP ([/-mediated killing at the level of the mucosal ep- 
ithelium with a subsequent transient carrier-like state. Progressive infection 
with dissemination into areas of the human body where specific antibody 
and C serum concentrations would far exceed those observed at the mucosal 
epithelium may lead to C -mediated Neisseria eradication. 

THE GONOCOCCUS AS AN INTRACELLULAR PATHOGEN ^ 

Despite the historic prevalence of N. gonorrhoeae within the human pop- g 

ulation and the considerable severity and frequency with which adverse com- § 

plications accompany disease (most notably in women), it has only been in § 

recent years that we have begun to understand gonococcal pathogenesis as it > 

occurs within its (sole) human host. Discernment of gonococcal pathogene- g 

sis has been hindered by the lack of a good animal model by which to study < 

> 

the pathology of N. gonorrhoeae infection. Although several animal models % 

have been described, no animal model exists that mirrors the full spectrum o 

of human disease caused by the gonococcus. Consequently, researchers have m 

relied on the use of human volunteers, tissue and organ cultures, and im- jj 

mortalized or malignant tissue culture cell lines by which to study gonococcal £ 

pathogenesis. Microscopic analyses of clinical biopsies and male urethral ex- o 

udates have allowed extrapolation of successful gonococcus infection as it jg 

occurs in vivo with the realization that the gonococcus exists as an intracel- K 

lular pathogen. However, more recent studies indicate that the mechanisms 2 
used by the gonococcus to obtain an intracellular status vary with the partiC- 



en 



in 



> 
t- 1 
W 

> 

Z 

o 



The development of a primary, human, male urethral epithelial cell cul- * 

ture system (Harvey et al., 1997) has greatly enhanced analysis of N. gonor- > 

rhoeae infection of the urethral epithelium. Within the urethral epithelium of 5 

men, infection is thought to occur as a sequential process that is initiated by 
attachment to and an intimate association with the targeted host cell that re- 
sults in pedestal formation of the urethral cell membrane beneath an adherent 




1-1 
w 

u 

I— I 

Pn 

< 

< 

w 

< 
u 



w 

s 

w 
1-1 
i-i 
i— i 

X 

CO 

Q 

$ 

Q 
w 

i-j 

Pi 
w 

Ph 
i— i 

w 




Figure 7.5. Polystyrene beads coated with LOS isolated from N. gonorrhoeae bind to the 
ASGP-R on human sperm as denoted by the arrow. This LOS glycoform harbors a 
terminal lactosamine as shown in Fig. 7.4. The LOS-ASGP-R interaction between bacteria 
and sperm may facilitate disease transmission. 



gonococcus. Gonococci are internalized within vacuoles by actin-dependent 
(Giardina et al., 1998) and clathrin-dependent (Harvey et al., 1997) processes, 
where they replicate prior to their release to the extracellular milieu. Epithelial 
cells are subsequently shed. The intimate association observed between the 
gonococcus and the urethral cell surface is mediated by the ASGP-R (Harvey 
et al., 2001). This receptor binds to the terminal galactose of the gonococcus 
LOS. Consequently, LOS sialylation greatly impairs the ability of gonococci 
to associate with the male urethral epithelium. However, the presence or 
absence of sialic acid on gonocococal LOS or the presence of the ASGP-R on 
the cervical epithelium does not appear to contribute to lower female genital 
tract colonization. The ASGP-R is also present on human sperm (Harvey 
et al., 2000), suggesting that an LOS-ASGP-R interaction may also facilitate 
disease transmission (Fig. 7.5). 



Studies in fallopian tube explants 

Similar processes to those observed with infection of the male urethra 
have been demonstrated to occur in FTOC. In this model, however, gono- 
coccal adherence occurs selectively on the nonciliated cells of fallopian tube 
tissue (McGee et al., 1981). Adherence is mediated by the LHr in a contact- 
inducible manner (Gorby et al., 1991). Following endocytosis of gonococci, 
bacteria-containing vacuoles are trancytosed to the basolateral surface of the 
infected epithelial cell, where they are released to the extracellular space. In 
this manner invasive bacteria are able to access the subepithelial tissues. 
Although the gonococcus selectively invades nonciliated cells, it is the ad- 
jacent ciliated cells that exhibit gonococcus -mediated cytotoxicity and ulti- 
mately results in the complete loss of ciliary action. The release of LOS and 
peptidoglycan by the gonococcus is thought to facilitate cytotoxicty of cili- ^ 

ated cells. Cytotoxicity occurs either directly or indirectly by the induction o 

of increased production of the inflammatory cytokine, tumor necrosis factor § 

(TNF). 



Infection studies of the uterine cervix 




w 

m 

Analysis of clinical N. gonorrhoeae isolates obtained from the fallopian § 

tube as well as FTOC infection studies suggest that gonococcal pili also facili- • 

tate fallopian tube colonization. Additionally, gonococci that exhibit a translu- S 

cent phenotype (Tr), and, therefore, lack Opa outer membrane proteins, are > 

more invasive in FTOC. The prevalence of Tr gonococci in clinical isolates g 

o 
from the fallopian tube and in genital, blood, and joint fluid of DGI patients g 

supports this observation. Expression of the LHr increases in an ascending g 

manner from the endometrium to the fallopian tubes. Conversely, surface z 

level expression of CR3 decreases progressively from the ectocervix to the g 

upper female genital tract. Although it is currently not known if the expres- * 

'■d 

sion of CR3 within the cervical epithelia exhibits cyclic variation, synthesis of in 

C3 and surface expression of the LHr is upregulated at the time of menses. g 

In a concerted action these molecules may contribute to an increased risk of 2 
complicated gonococcal infection. 



in 



> 
t- 1 
w 

> 

Z 

o 



In contrast to the pathology of N. gonorrhoeae infection observed for the * 

male urethra and the upper female genital tract, lower genital tract (i.e., the > 

uterine cervix) infection does not provoke a proinflammatory response. Re- " 

cent data suggest that this subclinical or asymptomatic condition may be 
the result of the ability of the gonococcus to subvert the AP of the C sys- 
tem. Primary cervical cells produce all of the alternative pathway C proteins 





1-1 
w 

u 

I— I 

< 
< 

w 

< 
u 



w 

s 

w 
1-1 
i-i 
i— i 

X 

CO 

Q 

Q 
w 

i-j 

Pi 
w 

Ph 
i— i 

w 



Figure 7.6. Gonococci (denoted by arrows) elicit membrane ruffles upon their association 
with the cervical epithelium. Membrane ruffling is the result of extensive host cytoskeletal 
rearrangements and results in bacterial internalization within macropinosomes. 
Membrane ruffles induced on ectocervical cells (left) exhibit a long ribbon-like 
morphology; ruffles formed by endocervical cells (right) are spherical in nature. 

required for C activation and its inactivation. C protein C3 is activated upon 
N. gonorrhoeae infection to form C3b, which is deposited upon the lipid A 
portion of gonococcal LOS and is rapidly inactivated to form iC3b (Edwards 
and Apicella, 2002). The presence of iC3b on the gonococcal surface allows 
binding to CR3 in a cooperative manner with gonococcal pilus and porin 
(Edwards et al., 2002). Several studies demonstrate that ligand binding to 
the I -(or inserted) domain of CR3 does not invoke a proinflammatory re- 
sponse in immune cells. Therefore, the cooperative binding of iC3b, pilus, 
and porin to this region of CR3 may contribute to the asymptomatic nature 
of gonococcal cervicitis. Once the gonococcus adheres to CR3, a complex 
signal transduction cascade is initiated that results in extensive cervical cell 
cytoskeletal rearrangement. Large protrusions of the cervical cell membrane, 
termed membrane ruffles (Fig. 7.6), engulf the bacterium, resulting in gono- 
cocci internalization within macropinosomes (Fig. 7.7) in an actin-dependent 
manner (Edwards et al., 2000). 



The interaction of the gonococcus with human neutrophils 

The gonococcus, being a strict human pathogen, exquisitely senses its 
particular microenvironment (within its sole human host) and alters its phys- 
iological response to that environment in such a manner as to promote its 
own survival. This is evident by the finding that the interaction of the gono- 
coccus with PMNs, which express CR3, occurs independently of a CR3 asso- 
ciation (Farrell and Rest, 1990). The gonococcus-PMN association requires 




Figure 7.7. A cross section of polarized cervical cells experimentally infected with N. 
gonorrhoeae reveals that gonococci (GC) reside within spacious macropinosomes (denoted 
by arrows). Gonococci have been labeled with a gold-bead antibody-conjugate to 
gonococcal porin. A large membrane protrusion, indicative of a membrane ruffle (MR), 
can be seen on the right. 

the presence of Opa outer membrane proteins; however, pilin proteins are 
not required. In contrast, gonococcal adherence to primary cervical and male 
urethral epithelial cells occurs independently of Opa proteins but requires 
gonococcal pilin. 

Opa proteins are thought to confer distinct cellular trophisms to individ- 
ual gonococci by the ability of different Opa proteins to differentially recog- 
nize host cell surface molecules. Two broad classes of Opa proteins exist that 
are represented by Opaso (i.e., Opa proteins that recognize host cell HSPG) 
and Opa52 (i.e., Opa proteins that recognize CEACAM). The interaction of 
Opa proteins with HSPG is dependent on the presence of vitronectin or fi- 
bronectin (van Putten et al., 1998), which functions as a bridging molecule 
between the gonococcus and its target receptor(s). Association with an in- 
tegrin coreceptor (a v /3i or a w Ps for vitronectin-mediated adherence or a w P\ 




o 

% 

o 
to 

m 

o 

B 

ft 

H 

X 
w 

< 
> 

o 

w 
n 

X 

> 

% 
i— i 

X 
o 

$ 

o 

o 

z 

M 
in 



> 
t- 1 
w 

> 

Z 

o 

M 

> 

M 

00 



for fibronectin-mediated adherence) triggers a signaling cascade within the 
target cell that is dependent on the activation of protein kinase C (PKC). 

A second, distinct, interaction between Opaso and HSPG is demonstrated 
in Chang conjunctiva epithelial cells. Interaction of Opaso with HSPG initi- 
ates a signaling cascade that activates phosphatidylcholine-dependent phos- 
pholipase C (PC-PLC). Generation of diacylglycerol (DAG) in turn activates 
acidic spingomyelinase (ASM), which results in ceramide production from 
spingomyelin. Ceramide is speculated to modulate the cytoskeletal rearrange- 
ments required for endocytosis of the cell-associated gonococcus. 

Variable CEACAM isoforms are differentially expressed by a variety of 
different cell types, including PMNs and epithelial cells. Activation of the host 
cell Src tyrosine kinases, Hck and Fgr, results in the subsequent activation 
of Rac, which results in epithelial cytoskeletal rearrangement and internal- 
S ization of the gonococcus. Although a single, primary receptor has not been 

S identified that modulates the gonococcus-PMN interaction, the presence of 

< the C E AC AM , integrin, and H S PG receptors on PM Ns suggests that any or all 

hj of these molecules may play a role in gonococcus phagocytosis. Additionally, 

X binding of specific antibody to the gonococcus may facilitate phagocytosis by 

u 




Q 

< 



means of Fey receptors. 



s A COMPREHENSIVE MODEL OF N. GONORRHOEAE 

< PATHOGENESIS IN MEN AND WOMEN 



x 

g The male urethra and the female uterine cervix are the primary sites 

Jj of the majority of gonococcal infections. These sites are illustrative of the 

E varying environments the organism must face in its pathogenesis, as these 

g are entirely different epithelial surfaces that are derived from distinct em- 

^ bryological origins. The urethra in the male and female is derived from the 

" initial division of the internal cloaca to form the urogenital sinus. The in- 

£ termediate section of this sinus becomes the pars pelvina, which ultimately 

£ becomes the female urethra, a portion of the vestibule and the distal vagina 

*L. in the female, and the prostatic urethra in the male. The caudalmost portion 

of the urogenital sinus becomes the membranous and penile portions of the 
urethra in the male and a portion of the vestibule in the female. In contrast, 
the uterine cervix is derived from the paramesonephric ducts. It is, therefore, 
not surprising that the surface receptors interacting with the organism on ep- 
ithelial surfaces of these structures are different. The gonococcus has adapted 
very well to these different conditions; consequently, it is not surprising that 
distinct pathogenic mechanisms occur with N. gonorrhoeae infection in men 
and women. 




N. gonorrhoeae possesses an impressive array of mechanisms that allow 
it to gain entry into host cells and to evade the host immune response. The 
importance of virulence factors such as pilus and Opa adhesins in gonococ- 
cal association with host cells is well established both in vitro and in vivo. 
In addition, although gonococcal LOS is implicated in the inflammatory re- 
sponse associated with gonococcal infection, clinical studies support in vitro 
evidence that gonococcal LOS plays a role in gonococcal adherence to and 
invasion of host cells. LOS and LPS play a role in invasion of host cells by 
other pathogenic bacteria. All of these virulence factors, which are located 
on the surface of the gonococcus, are capable of phase or antigenic variation 
or both. Because an association between the gonococcus and the host cell 
surface is a prerequisite for entry, it is believed that pili, Opa, LOS, and porin 
play an important role in the initiation of disease. Multiple studies show 
that the process of adherence to, followed by the invasion of, host cells in- 
volves the interaction of a repertoire of bacterial and host constituents. Given g 
that gonorrhea is a sexually transmitted disease and reflecting upon what is § 
known about gonococcal pathogenesis, assembling a model in the context of § 
the male and female urogenital tracts may promote an understanding of this > 
process at the level of human disease (Fig. 7.8). g 

Gonococci enter and survive within cervical epithelial cells. Because the < 

intracellular environment is a source of CMP-NANA, the model shown in % 

i— i 

Fig. 7.8 presumes that gonococci with sialylated LOS are released from o 

cervical epithelial cells to the extracellular environment. LOS sialylation is m 

mediated by gonococcus-encoded sialyltransferase, but with host-derived jj 

CM P-NANA. Cooperativity between the gonococcus and other microbes that £ 

produce sialidases in the vaginal microflora results in desialylation of gono- o 

coccal LOS. Loss of the sialic acid moiety on gonococcal LOS, while in res- jg 

idence within the lower female genital tract, may prime the gonococcus for K 

infection of the male urethra (where the presence of a sialic acid moiety 2 
would inhibit urethral interaction with ASGP-R). Upon transfer to the male 



o 



en 



in 



partner, these desialylated gonococci can interact with the ASGP-R through 

the now available lactosamine residue on the nonreducine end of the LOS g 

> 
oligosaccharide. This initiates clathrin-mediated endocytosis of the organism, S 

> 



facilitating entry into the urethral epithelial cell. § 

Experiments examining the mechanism of gonococcal internalization * 

involving LOS and the ASGP-R show that, although the number of ASGP- > 

Rs at the epithelial cell surface increases after 4 h following infection with 5 

terminal lactosamine-expressing N. gonorrhoeae, no difference in the levels 
of total cellular ASGP-R protein or ASGP-R mRNA could be detected. This 
suggests cellular ASGP-R traffics to the cell surface during infection rather 



than being newly synthesized. Considering these findings together with what 
is known about the ASGP-R, one can propose a model in which extracellular 
gonococci bind ASGP-R and are internalized in clathrin-coated vesicles. 

Clathrin-coated vesicles, containing receptor-ligand complexes, fuse 
with other clathrin-coated vesicles. In this early endosome, the clathrin coat 
is shed in an ATP-dependent process, and a pH drop results in uncoupling 
of the receptor-ligand complex. The ASGP-R is then recycled back to the cell 
surface where it is again available for binding ligand. In addition to the recy- 
cling back of ASGP-R to the cell surface in the absence of ligand, evidence is 
presented that a large fraction of ASGP-R is recycled to the cell surface prior 
to ligand dissociation, a process termed diacytosis, which may play a role in 
potentiating gonococcal urethritis. 

Signaling mechanisms that are required for recruitment of cellular 
ASGP-R to the cell surface, whether or not fusion of gonococci-containing 
vesicles occurs in vivo, and the intracellular fate of the gonococcus remain to g 

be determined. Human volunteer studies, in vitro invasion assays, and mi- § 



o 



M 



croscopy studies suggest that gonococci survive intracellularly. Lysosomes are § 



o 



considered the endpoint of ASGP-R-mediated endocytosis, but it is not known jj 

if gonococci internalized by means of ASGP-R escape lysosomal degradation. g 

Pathogenic Neisseria species secrete an IgA protease, which cleaves host cell < 

> 

LAM PI. LAM PI is involved in maturation of the endosome; consequently, % 



2 



the ability of the gonococcus to cleave this molecule is believed to contribute o 



to gonococcal survival in the intracellular environment. m 

° n 

However, recent studies with a gonococcal IgAl protease mutant showed jj 

that, compared to the wild type, this mutant was not compromised in its ability £ 

o 

> 

Figure 7.8. (facing page). The different pathogenic interactions of the gonococcus in men d 

and women. In women, infection of the cervical epithelial cells CEC occurs by membrane o 



ruffling, which is an actin- dependent process. Once within the epithelial cell, the § 

organism acquires the substrate for sialylation and sialylates its LOS (•). The £ 

inflammatory response is minimal, and as cells are shed they are released into the vaginal 2 

environment, which is rich in sialidases (\) produced by endogenous microflora that > 

results in a desialylation of the LOS (•). Organisms passed to men are predominately ^ 

> 

desialylated and capable of entering urethral epithelial cells (UEC) by the interaction of 2 

the terminal lactosamine on the LOS and the asialoglycoprotein receptor on the urethral JJ 

cell by a clathrin-dependent receptor-mediated endocytosis. Once within the cell, the > 

organisms become resialylated. As cells are shed, organisms are released and " 

phagocytosed by PMNs. Infected epithelial cells and PMNs are transferred to the female 

in the ejaculate. Uptake by the cervical cell through CR3 is unaffected by the sialylation 

state of the organism. Nonsialylated gonococci can also bind to human sperm and may aid 

in infection spread. See color section. 




to cause urethritis in human volunteers. Whether the organism can prolifer- 
ate within the host cell is still a matter for study, although exfoliated urethral 
epithelial cells in exudates from infected men show numerous organisms. 
This suggests that gonococcus replication very likely occurs within host (ep- 
ithelial) cells. Within urethral epithelial cells, host-derived CMP-NANA is 
again available as a substrate for sialylation of gonococcal LOS. By an as yet 
undefined mechanism, gonococci traffic to the surface of the epithelial cell 
and are released from the surface. Once on the surface, these sialylated gono- 
cocci cannot reenter urethral epithelial cells or interact with the ASGP-R on 
sperm until the bacteria have divided in this CMP-NANA-free environment 
and the terminal lactosamine of their LOS becomes free of sialic acid. 

In males, gonococcal infection likely results in inflammatory cytokine 
release by urethral epithelial cells. Few studies have addressed the question 
S of epithelial cell signal transduction resulting from gonococcal adherence 

S and invasion of these cells. However, in vitro studies using primary and im- 

< mortal epithelial cells demonstrate that infection with N. gonorrhoeae elicits 

hj the production of the cytokines, TNF-a, interleukin (IL)-6, and IL-1/3 and the 

X chemokine, IL-8, which triggers PMN influx and significantly contributes to 

u 

g local inflammation within the male urethra. It is also shown that, in exper- 

§ imental gonococcal infection in men, the cytokines, IL-6, IL-8, and TNF-a, 

< 

^ are released at the local site of infection, the urethral mucosa. This is in con- 

% trast with findings observed in women with gonococcal cervicitis where local 

E levels of IL-1, IL-6, and IL-8 are not elevated (Hedges et al., 1998). Therefore, 

g the release of inflammatory cytokines by the cervical epithelium in response 

3 to N. gonorrhoeae infection remains under question. 

e In both males and females, gonococcal infection usually results in shed- 

g ding of urethral and cervical epithelial cells, respectively. PMNs containing 

^ gonococci are also found in exudates from males with gonorrhea. Similar 

" to the urethral epithelial cells, PMNs could also serve as a source of inflam- 

£ matory cytokines and CMP-NANA and may release viable gonococci to the 

£ extracellular environment for disease transmission or reentry into urethral 

^ epithelial cells following desialylation by host sialidases. An Opa-CEACAM 

interaction on PMNs may allow survival of some intracellular gonococci, be- 
cause engagement of CE AC AM initiates a priming signal within these cells 
that results in activation of adhesion receptors without the release of inflam- 
matory mediators or the induction of a respiratory burst. 

The majority of gonococci that are transmitted from males to their female 
partners have sialylated LOS. However, the presence or absence of sialic acid 
on gonococcal LOS does not influence the association of the gonococcus 
with the cervical epithelium (Edwards and Apicella, 2002). In this model of 



cervical invasion, AP C components are produced by the cervical epithelium 
and subsequently released to the extracellular milieu. Upon transmission of 
gonococci to the lower female genital tract, C3 deposition occurs on the lipid A 
determinant of gonococcal LOS. Expression of an LOS (as opposed to an LPS) 
structure may increase fH activity and, consequently, the conversion of C3b 
to iC3b on the gonococcus surface. Similarly, the resemblance of gonococcal 
LOS to human paraglobosides and glycosphingolipids (some of which serve 
as synthetic precursors of cervical mucins) may favor C inactivation. The 
interaction of gonococcal pilin with the I -domain may initially serve as an 
anchor to CR3 that allows this bacterium to overcome repulsive forces that 
occur with the host cell membrane and, thus, position the gonococcus at the 
cervical cell membrane where the concentration of C components may be 
expected to be sufficient to allow an efficient interaction with CR3. 

Porin is very closely associated with LOS on the gonococcal surface, and 
its role in CR3-mediated endocytosis may be multifactorial; however, the di- g 

rect association of porin with the I-domain of CR3 is crucial to cervical cell § 




o 



M 



invasion. The juxtaposition of porin to LOS within the outer gonococcal mem- § 



o 



brane may spatially favor porin and iC3b adherence to the CR3-I-domain. > 

Additionally, pilin- and porin-mediated binding of the gonococcus to CR3 g 

may elicit a biphasic calcium flux in cervical epithelial cells that increases < 



> 



the affinity and avidity of iC3b-, porin-, and pilin-mediated CR3 adherence, % 

as well as the release of additional effector proteins that might augment the o 

eonococcus-CR3 interaction. m 

° n 

Membrane ruffling in professional phagocytic cells is attributed to the jj 

activity of the Rho family of GTPases. Although ruffling has not been pre- £ 

viously observed in response to CR3 activation, similar responses probably o 

occur within the cervical epithelium in that the inhibition of Rho activation jg 

also inhibits invasion of cervical epithelial cells. Jones et al. (1998) recently K 

reported that fMLP- and Fey coreceptor-induced stimulation of CR3 results 2 
in the activation ofp21 -activating kinase (PAK1). Activation of PAK1 can elicit 



en 



in 



lamellipodia and ruffle formation, as well as vinculin accumulation, through 

a signaling cascade involving the PAK-interacting guanine nucleotide ex- s 

change factor (PIX) and Rac. The ability of porin to inhibit fMLP-induced S 

> 



signaling events in neutrophils suggests that these molecules may share § 

effector proteins; however, it is unknown if binding of gonococcal porin to * 

CR3 elicits membrane ruffling in primary cervical cells by augmenting signal > 

transduction of PIX and PAK1. 

Binding of CR3 triggers a cascade of events that may involve the activation 
of PIX and PAK1 and that results in ezrin and vinculin focal complex forma- 
tion and the cytoskeletal changes required for ruffle formation. Engagement 



t- 1 

00 




of CR3 by the gonococcus is sufficient to elicit membrane ruffling; however, 
other fluid phase molecules may facilitate gonococcal internalization. Once 
membrane ruffles are formed, endocytosis ensues, allowing gonococcal inter- 
nalization within macropinosom.es. These ideas are consistent with the wide 
variability associated with the incubation period, the clinical manifestations, 
and the high prevalence of asymptomatic gonococcal cervicitis in women. 
Clearly, coevolution with its human host has allowed the gonococcus to care- 
fully orchestrate a mechanism of CR3 adherence in which constituents of its 
outer membrane, that is, porin and pilus, function cooperatively with host 
cell factors (i.e., AP proteins) to ensure its increased survival. 

The deposition of iC3b on the gonococcal lipid A core may confer a 
survival advantage to the gonococcus by displacing the LOS oligosaccharide 
chain such that it is unable to bind to the lectin-binding domain of CR3. 
S iC3b-mediated adherence to CR3 occurs independently of a proinflamma- 

S tory response; however, binding of the lectin domain stimulates a respiratory 

< burst through the generation of a second signal within the eukaryotic (host) 

hj cell. Consequently, a CR3 association mediated through the lectin domain 

X of this receptor may result in decreased survival of the invasive organism. 

u 

g Additionally, the presence of iC3b on the lipid A core may obstruct the asso- 

§ ciation of the lipid A core region with CD 18. It, therefore, appears that C3b 

< 

^ deposition on lipid A facilitates gonococcal infection and may promote bac- 

% terial survival in that any or all of these phenomena would favor adherence to 

E the I -domain of CR3 and subsequent bacterial internalization in the absence 

g of a respiratory burst. iC3b-mediated epithelial adherence, in the absence of 

Jj a respiratory burst, may allow the gonococcus to attain a carrier-like state 

e within the cervical epithelium. Ascent to the upper female reproductive tract 

g may subsequently occur as the result of hormonal changes that alter host 

^ cell epithelia, molecules available for gonococcal use, or the expression of 

" gonococcal virulence factors. These ideas are supported by the increased risk 

£ to women for PID and DGI that is associated with menses and with the cor- 

£ relation made between the presence or absence of Opa proteins and the site 

*L. of gonococcus isolation. (Transparent gonococci are predominantly isolated 

from the cervix at the time of menses and from the fallopian tubes; from 

asymptomatic men; and from blood, joint, and genital fluids of DGI patients. 

Opa-expressing gonococci are predominantly isolated from men and from 

the cervix at the time of ovulation.) 

Surface level expression of CR3 decreases progressively from the ecto- 
cervix to the upper female genital tract. Conversely, expression of the LHr 
increases in an ascending manner from the endometrium to the fallopian 
tubes (where it serves as a receptor for gonococcal adherence; see Gorby et al., 



1991). Although it is currently not known if the expression of CR3 within the 
cervical epithelia exhibits cyclic variation, synthesis of C3 and surface ex- 
pression of the LHr is upregulated at the time of menses. In a concerted 
action, these molecules may contribute to an increased risk of complicated 
gonococcal infection. 

The level of sialic acid found within the cervical milieu exhibits cyclic 
variation and is inversely related to C3 production. Additionally, although 
sialidase activity is present within cervical mucus throughout the menses 
cycle, cyclic variation also exists with the ability of this enzyme to use endoge- 
nous or exogenous substrates. With regard to the increased susceptibility to 
ascending gonococcal infection correlated with menses, the decreased preva- 
lence of sialic acid within the cervical milieu at the time of menses might 
suggest that sialylated bacteria are impaired in their ability to ascend the fe- 
male genital tract, although no discernable difference exists in the ability of 
sialylated or unsialylated gonococci to interact with cervical epithelia. How- g 

ever, removal of sialic acid from the mucosal surfaces of the female genital § 



REFERENCES 




o 



M 



tract may allow gonococci of a particular isotype to gain access to subepithelial § 



o 



tissues or to ascend the female genital tract and thereby initiate DGI or PID, jj 

respectively, and may have important implications with regard to disease g 

transmission. 



< 
> 

o 

M 

n 

Apicella, M.A., Ketterer, M., Lee, F.K.N., Zhou, D., Rice, P.A., and Blake. M.S. jji 

2; 
(1996). The pathogenesis of gonococcal urethritis in men: confocal and im- 5 

munoelectron microscopic analysis of urethral exudates from men infected o 

with Neisseria gonorrhoeae. J. Infect. Dis. 173, 636-646. jg 

Apicella, M.A., Mandrell, R.E., Shero, M., Wilson, M.E., Griffiss, J.M., Brooks, K 

G.F., Lammel, C, Breen, J.F., and Rice, P.A. (1990). Modification by sialic 2 

acid of Neisseria gonorrhoeae lipooligosaccharide epitope expression inhuman 



en 



in 



urethral exudates: an immunoelectron microscopic analysis. J. Infect. Dis. 

162, 506-512. S 

> 

Aral, S.O., Mosher, W.D., and Cates, W. (1991). Self-reported pelvic inflammatory w 

disease in the United States. JAMA 266, 2570-2573. % 

o 
Bjerknes, R., Guttormsen, H.-K., Solberg, CO., and Wetzler, L.M. (1995). Neis- ** 

serial porins inhibit human neutrophil actin polymerization, degranulation, > 

opsonin receptor expression, and phagocytosis but prime the neutrophils to " 

increase their oxidative burst. Infect. Immun. 63, 160-167. 

Bolan, G., Ehrhardt, A.A., and Wasserheit, J.N. (1999). Gender perspectives and 

STDs. In Sexually Transmitted Diseases, ed. K.K. Holmes, P. -A. Mardh, P.F. 



Sparling, S.M. Lemon, W.E. Stamm, P. Piot, and J.N. Wasserheit, 3rd ed., 
pp. 117-127. New York: McGraw-Hill. 

Campagnari, A. A., Spinola, S.M., Lesse, A.J., Kwaik, Y.A., Mandrell, R.E., and 
Apicella, M.A. (1990). Lipooligosaccharide epitopes shared among gram- 
negative non-enteric mucosal pathogens. Microb. Pathog. 8, 353-362. 

Chen, C.J., Thomas, C.E., McLean, D.S., Rouquette-Loughlin, C, Shafer, W.M., 
and Sparling, P.F. (2002). PilQ point mutation results in hemoglobin utiliza- 
tion in the absence of HpuA/B. In Abstracts of the 13th International Pathogenic 
Neisseria Conference, ed. D.A. Caugant and E. Wedege, p. 103. Oslo: Nordberg 
Aksidenstrykkeri AS. 

Chen, T.C., Swanson, J., Wilson, J., and Belland, R.J. (1995). Heparin protects 
Opa + Neisseria gonorrhoeae from the bactericidal action of normal human 
serum. Infect. Immun. 63, 1790-1795. 
< Clark, V.L., Campbell, L.A., Palermo, D.A., Evans, T.M., and Klimpel, K.W. (1987). 

5 Induction and repression of outer membrane proteins by anaerobic growth 




< of Neisseria gonorrhoeae. Infect. Immun. 55, 1359-1364. 

m Cooke, S.J., Jolley, K., Ison, C.A., Young, H., and Heckels, J.E. (1998). Naturally 

x occurring isolates of Neisseria gonorrhoeae, which display anomalous serovar 

u 

§ properties, express PIA/PIB hybrid porins, deletions in PIB or novel PIA 

g molecules. FEMS Microbiol. Lett. 162, 75-82. 

< 

^ Cornelissen, C.N., Kelley, M., Hobbs, M.M., Anderson, J.E., Cannon, J.G., 

% Cohen, M.S., and Sparling, P.F. (1998). The transferrin receptor expressed 

E by gonococcal strain FA1090 is required for experimental infection of human 

£ male volunteers. Mol. Microbiol. 27, 611-616. 

3 Dehio, C, Gray-Owen, S.D., and Meyer, T.F. (2000). Host cell invasion by 

E pathogenic Neisseriae. In Subcellular Biochemistry, Vol. 33: Bacterial Invasion 

g into Eukaryotic Cells, ed. T.A. Oelschlaeger and J. Hacker, pp. 61-96. New 

^ York: Plenum. 

" Densen, P., MacKeen, L.A., and Clark, R.A. (1982). Dissemination of gonococcal 

£ infection is associated with delayed stimulation of complement-dependent 

£ neutrophil chemotaxis in vitro. Infect. Immun. 38, 563-572. 

E. Dilworth, J.A., Hendley, J.O., and Mandell, G.L. (1975). Attachment and ingestion 

of gonococci by human neutrophils. Infect. Immun. 11, 512-516. 
Edwards, J.L. and Apicella, M.A. (2002). The role of lipooligosaccharide in 

Neisseria gonorrhoeae pathogenesis of cervical epithelia: lipid A serves as a 

C3 acceptor molecule. Cell. Microbiol. 4, 585-598. 
Edwards, J.L., Brown, E.J., Ault, K.A., and Apicella, M.A. (2001). The role of 

complement receptor 3 (CR3) in Neisseria gonorrhoeae infection of human 

cervical epithelia. Cell. Microbiol. 3, 611-622. 
Edwards, J.L., Brown, E.J., Uk-Nham, S., Cannon, J.G., Blake, M.S., and Apicella, 

M.A. (2002). A co-operative interaction between Neisseria gonorrhoeae and 



complement receptor 3 mediates infection of primary cervical epithelial cells. 
Cell. Microbiol. 4, 571-584. 

Edwards, J.L., Shao, J.Q., Ault, K.A., and Apicella, M.A. (2000). Neisseria gonor- 
rhoeae elicits membrane ruffling and cytoskeletal rearrangements upon in- 
fection of primary human endocervical and ectocervical cells. Infect. Immun. 
68, 5354-5363. 

Estabrook, M.M., Griffiss, J.M., and Jarvis, G.A. (1997). Sialylation of Neisseria 
meningitidis lipooligosaccharide inhibits serum bactericidal activity by mask- 
ing lacto-N-neotetraose. Infect. Immun. 65, 4436-4444. 

Evans, B.A. (1977). Ultrastructure study of cervical gonorrhea. J. Infect. Dis. 136, 
248-255. 

Farrell, C.F. and Rest, R.F. (1990). Up-regulation of human neutrophil recep- 
tors for Neisseria gonorrhoeae expressing PII outer membrane proteins. Infect 
Immun. 58, 2777-2784. * 

Forest, K.T., Dunham, S.A., Koomey, M., andTainer, J. A. (1999). Crystallographic g 

structure reveals phosphorylated pilin from Neisseria: phosphoserine sites § 

modify type IV pilus surface chemistry and fiber morphology. Mol. Microbiol. § 

31, 743-752. £ 

Forest, K.T. and Tainer, J. A. (1997). Type-4 pilus -structure: outside to inside and g 

top to bottom - a minireview. Gene 192, 165-169. < 

> 

Giardina, P.C., Williams, R., LubarofF, D., and Apicella, M.A. (1998). Neisseria % 

gonorrhoeae induces focal polymerization of actin in primary human urethral o 

epithelium. Infect. Immun. 66, 3416-3419. m 

Gorby, G.L., Clemens, CM., Barley, L.R., and McGee, Z.A. (1991). Effect of hu- jji 

man chorionic gonadotropin (hCG) on Neisseria gonorrhoeae invasion of and s 

IgA secretion by human fallopian tube mucosa. Microb. Pathogen. 10, 373- o 

384. jg 

Haines, K.A., Reibman, J., Tang, X., Blake, M.S., and Weissmann, G. (1991). Ef- K 

fects of protein I of Neisseria gonorrhoeae on neutrophil activation: generation 2 

M 

en 



of diacylglycerol from phosphatidylcholine via a specific phospholipase C is 
associated with exocytosis. J. Cell Biol. 114, 433-442. 



on 



Haines, K.A., Yeh, L., Blake, M.S., Cristello, P., Korchak, H., and Weissmann, S 

> 

G. (1988). Protein I, a translocatable ion channel from Neisseria gonorrhoeae, w 



> 



selectively inhibits exocytosis from human neutrophils without inhibiting jzj 

2 ~ generation. J. Biol. Chem. 263, 945-951. g 

Harvey, H.A., Jennings, M.P., Campbell, C.A., Williams, R., and Apicella, M.A. > 

(2001). Receptor-mediated endocytosis of Neisseria gonorrhoeae into primary " 

human urethral epithelial cells: the role of the asialoglycoprotein receptor. 
Mol. Microbiol. 42, 659-672. 

Harvey, H.A., Porat, N., Campbell, C.A., Jennings, M.P., Gibson, B.W., Phillips, 
N.J., Apicella, M.A., and Blake, M.S. (2000). Gonococcal lipooligosaccharide 



is a ligand for the asialoglycoprotein receptor on human sperm. Mol. Micro- 
biol. 36, 1059-1070. 

Harvey, H.A., Ketterer, M.R., Preston, A., Lubaroff, D., Williams, R., and Apicella, 
M.A. (1997). Ultrastructure analysis of primary human urethral epithelial cell 
cultures infected with Neisseria gonorrhoeae. Infect. Immun. 65, 2420-2427. 

Hedges, S.R., Sibley, D.A., Mayo, M.S., Hook, E.W. Ill, and Russell, M.W. (1998). 
Cytokine and antibody responses in women infected with Neisseria gonor- 
rhoeae: effects of concomitant infections. J. Infect. Dis. 178, 742-751. 

Householder, T.C., Belli, W.A., Lissenden, S., Cole, J.A., and Clark, V.L. (1999). 
cis- and trans-acting elements involved in regulation of aniA, the gene encod- 
ing the major anaerobically induced outer membrane protein in Neisseria 
gonorrhoeae.]. Bacteriol. 181, 541-551. 

Jarvis, G.A. (1995). Recognition and control of neisserial infection by antibody 
and complement. Trends Microbiol. 3, 198-201. 
3 Jones, S.L., Knaus, U.G., Bokoch, G.M., and Brown, E.J. (1998). Two signal- 




i-i 



< ing mechanisms for activation of a M p 2 avidity in polymorphonuclear neu- 

m trophils. J. Biol. Chem. 273, 10,556-10,566. 

x Kallstrom, H., Islam, Md.S., Berggren, P.-O., and Jonsson, A.-B. (1998). Cell sig- 

u 

§ naling by the type IV pili of pathogenic Neisseria. J. Biol. Chem. 273, 21,777— 

g 21,782. 

< 

^ Kallstrom, H., Liszewski, M.K., Atkinson, J. P., and Jonsson, A.-B. (1997). Mem- 

% brane cofactor protein (MCP or CD46) is a cellular pilus receptor for patho- 

E genie Neisseria. Mol. Microbiol. 25, 639-647. 
< 



g Lemon, S.M. and Sparling, P.F. (1999). Pathogenesis of sexually transmitted viral 

^ and bacterial infections. In Sexually Transmitted Diseases, ed. K.K. Holmes, 

£ P.-A. Mardh, P.F. Sparling, S.M. Lemon, W.E. Stamm, P. Piot, and J.N. 

g Wasserheit, pp. 433-449. New York: McGraw-Hill. 

^ Mandrell, R.E. and Apicella, M.A. (1993). Lipo-oligosaccharides (LOS) of mucosal 

" pathogens: molecular mimicry and host -modification of LOS. Immunobiology 

£ 187, 382-402. 

£ Marceau, M., Forest, K., Beretti, J.-L., Tainer, J., and Nassif, X. (1998). Conse- 

|z; 

£, quences of the loss of O-linked glycosylation of meningococcal type IV pilin 

on piliation and pilus-mediated adhesion. Mol. Microbiol. 27, 705-715. 

McGee, Z.A., Johnson, A.P., and Taylor-Robinson, D. (1981). Pathogenic mecha- 
nisms of Neisseria gonorrhoeae: observations on damage to human fallopian 
tubes in organ culture by gonococci of colony type 1 or type 4. J. Infect. Dis. 
143, 413-422. 

Ober, W.B. (1970). BoswelPs clap. JAMA 212, 91-95. 

Parge, H.E., Forest, K.T., Hickey, M.J., Christensen, D., Getzoff, E.D., and Tainer, 
J.A. (1995). Structure of the fibre-forming protein pilin at 2.6 A resolution. 
Nature 378, 32-38. 



van Putten, J. P.M., Duensing, T.D., and Carlson, J. (1998). Gonococcal invasion 
of epithelial cells driven by P. I A, a bacterial ion channel with GTP binding 
properties. J. Exp. Med. 188, 941-952. 

van Putten, J. P.M., Duensing, T.D., and Cole, R.L. (1998). Entry of Opa + gonococci 
into Hep-2 cells requires concerted action of glycosaminoglycans, fibronectin 
and integrin receptors. Mol. Microbiol. 29, 369-379. 

van Putten, J. P.M. and Robertson, B.D. (1995). Molecular mechanisms and im- 
plications for infection of lipopolysaccharide variation in Neisseria. Mol. Mi- 
crobiol. 16, 847-853. 

Rahman, M., Kallstrom, H., Normark, S., and Jonsson, A.-B. (1997). PilC of 
pathogenic Neisseria is associated with the bacterial cell surface. Mol. Mi- 
crobiol. 25, 11-25. 

Ram, S., Cullinane, M., Blom, A.M., Gulati, S., McQuillen, D.P., Boden, R., 
Monks, B.G., O'Connell, C, Elkins, C, Pangburn, M.K., Dahlback, B., and 
Rice, P. A. (2001). C4bp binding to porin mediates stable serum resistance of g 

Neisseria gonorrhoeae. Intl. Immunopharmacol. 1, 423-432. § 



1422-1427. 



o 



w 



Ram, S., McQuillen, D.P., Gulati, S., Elkins, C, Pangburn, M.K., and Rice, P.A. g 



o 



(1998). Binding of complement factor H to loop 5 of porin protein 1A: a jj 

molecular mechanism of serum resistance of nonsialylated Neisseria gonor- g 

rhoeae. J. Exp. Med. 187, 743-752. < 

> 

Ram, S., Sharma, A.K., Simpson, S.D., Gulati, S., McQuillen, D.P., Pangburn, 3 

M.K., and Rice, P.A. (1998). A novel sialic acid binding site on factor H o 

mediates serum resistance of sialylated Neisseria gonorrhoeae. J. Exp. Med. m 

187, 743-752. £ 

Schneider, H., Cross, A.S., Kuschner, R.A., Taylor, D.N., Sandoff, J.C., Boslego, s 

J.W., and Deal, CD. (1995). Experimental human gonococcal urethritis: 250 o 

Neisseria gonorrhoeae MSllmkC are infective. J. Infect. Dis. 172, 180-185. jg 

Schneider, H., Schmidt, K.A., Skillman, D.R., Van De Verg, L, Warren, R.L., K 

Wylie, H.J., Sandoff, T.C., Deal, CD., and Cross, A.S. (1996). Sialylation 2 
lessens the infectivity of Neisseria gonorrhoeae MSllmkC. J. Infect. Dis. 173, 



en 



in 



Schryvers, A.B. and Stojiljkovic, I. (1999). Iron acquisition systems in the patho- £ 

genie Neisseria. Mol. Microbiol. 32, 1117-1123. 5 

Snyder, L.A.S., Butcher, S.A., and Saunders, N.J. (2001). Comparative whole- jj 

a 
genome analyses reveal over 100 putative phase variable genes in the *n 

pathogenic Neisseria spp. Microbiology 147, 2321-2332. > 

Spence, J.M., Chen, J.C-R., and Clark, V.L. (1997). A proposed role for the lutropin " 

receptor in contact-inducible gonococcal invasion of HeclB cells. Infect. Im- 

mun. 65, 3736-3742. 
Stimson, E., Virji, M., Makepeace, K., Dell, A., Morris, H.R., Payne, G., Saunders, 

J.R., Jennings, M.P., Barker, S., Panico, M., Bcench, I., and Moxon, E.R. 



(1995). Meningococcal pilin: a glycoprotein substituted with digalaactosyl 
2,4-diacetamido-2,4,6-trideoxyhexose. Mol. Microbiol. 17, 1201-1214. 

Sweet, R.L., Blankfort-Doyle, M., Robbie, M.O., and Schacter, J. (1986). The occur- 
rence of chlamydial and gonococcal salpingitis during the menstrual cycle. 
JAMA 255, 2062-2064. 

Tobiason, D.M. and Seifert, H.S. (2001). Inverse relationship between pilus- 
mediated gonococcal adherence and surface expression of the pilus receptor, 
CD46. Microbiology 147, 2333-2340. 

Vogel, U. and Frosch, M. (1999). Mechanisms of neisserial serum resistance. Mol. 
Microbiol. 32, 1133-1139. 

Weiser, J.N., Goldberg, J.B., Pan, N., Wilson, L, and Virji, M. (1998). The phos- 
phorylcholine epitope undergoes phase variation on a 43-kilodalton protein 
in Pseudomonas aeruginosa and on pili of Neisseria meningitidis and Neisseria 
gonorrhoeae. Infect. Immun. 66, 4263-4267. 
S Wen, K.-K., Giardina, P.C., Blake, M.S., Edwards, J.L., Apicella, M.A., and Ruben- 

< stein, P. A. (2000). Interaction of the gonococcal porin P. IB with G-and F-actin. 

Biochemistry 39, 8638-8647. 




i-i 



< 

w 

x Williams, J.M., Chen, G.-C., Zhu, L., and Rest, R.F. (1998). Using the yeast two- 

u 

§ hybrid system to identify human epithelial cell proteins that bind gonococcal 

§ Opa proteins: intracellular gonococci bind pyruvate kinase via their Opa pro- 

< 

^ teins and require host pyruvate for growth. Mol. Microbiol. 27, 171-186. 

< 
X 

< 

S 

w 

i-i 
i-i 
i— i 

X 

CO 

Q 

$ 

Q 
w 

i-i 

Pi 
w 

Ph 
i— i 

!zi 



CHAPTER 8 

Group A streptococcal invasion of host cells 

Harry S. Courtney and Andreas Podbielski 



Group A streptococci (GAS), Streptococcus pyogenes, are beta-hemolytic, Gram- 
positive, pyogenic cocci that usually grow in chains. S. pyogenes is classified as 
group Abased on serological reactions with its C carbohydrate, which consists 
of polymers ofrhamnose substituted with N-acetylglucos amine (Lancefield, 
1933). For epidemiological purposes, GAS are further classified into more 
than 100 different types based on serological reactions with the variable do- 
mains of M proteins, or, more recently, based on 5' emm gene sequences 
(Beall et al., 1996). Other typing schemes based on serological reactions with 
serum opacity factor, T proteins, and R proteins are also used (Johnson and 
Kaplan, 1993). 

S. pyogenes is almost exclusively associated with humans and commonly 
causes a variety of diseases, including pharyngotonsillitis, impetigo, scarlet 
fever, and more severe infections, such as puerperal sepsis, myositis, necro- 
tizing fasciitis, and toxic shock syndrome. Among several of the nonsuppu- 
rative complications of group A streptococcal infections are acute rheumatic 
fever and acute glomerulonephritis, which are usually preceded by infections 
of the throat and skin, respectively. These sequelae are thought to be due to 
autoimmune T- and B-cell responses induced by streptococcal products. Ac- 
cumulating evidence also suggests that group A streptococcal infections may 
lead to other autoimmune diseases, such as obsessive compulsive disorders, 
or they may exacerbate others such as guttate psoriasis (reviewed by Cun- 
ningham, 2000) . 

ADHESION: PRELUDE TO INVASION? 

To establish these infections, the streptococcus must first attach to the ep- 
ithelium of the host. This attachment is accomplished by specific interactions 





CO 

1-1 
w 
I— I 
PQ 
Q 
O 
Ph 

CO 

< 
W 

Q 

< 
Q 

< 
w 

O 

u 

CO 

Pi 
< 

X 




Figure 8.1. An electron micrograph of streptococcal adhesion to a human buccal 
epithelial cell. 



between surface structures of the streptococcus (adhesins) and receptors on 
host cells (reviewed by Courtney et al., 2002). Attachment of streptococci to 
a human buccal epithelial cell is shown in Fig. 8.1. Note that the epithelial 
cell has a ruffled surface that is due to convoluted ridges that cover its sur- 
face and that appear as projections in electron micrographs of thin sections. 
Similar ridges are found on the surfaces of human pharyngeal and tonsil- 
lar epithelial cells (Stenfors et al., 2000). Streptococcal attachment to these 
cells occurs primarily by interactions with receptors on these ridges, and ad- 
hesion is often mediated by multiple interactions resulting in high-avidity 
binding. 

Early adhesion experiments demonstrated that streptococci did not at- 
tach to all buccal cells but instead attached primarily to those cells coated 
with fibronectin (Abraham et al., 1983; Simpson and Beachey, 1983). These 
findings coupled with additional research confirmed that fibronectin was a 
receptor for GAS (Courtney et al., 1986). Lipoteichoic acid (LTA) was the first 
streptococcal adhesin to be identified that reacted with fibronectin (Beachey 
and Ofek, 1976; Courtney et al., 1986). Subsequently, at least 24 other ad- 
hesins have been described, many of which react with fibronectin or other 



Table 8.1. Putative streptococcal adhesins-invasins and receptors 



Adhesin-Invasin Receptor 



References 



C5a peptidase 


Fn 


C -carbohydrate 


• 


Collagen-binding 


collagen 


protein 




Fba(FbaA) 


Fn 


FbaB 


Fn 


FBP54 


Fn, Fgn 


Galactose-binding 


} 


protein 




G3PDH(orSDH) 


Fn, Fgn, 30-kDa 




protein 


Hyaluronic acid 


CD44 



Lbp 




Iaminin 


Lsp 




Iaminin 


Lipoteichoic 


acid 


Fn 

scavenger receptor, 
CD14 


M protein 




CD46, galactose, 
Fn, Iaminin 
fucose/fucosylated 
glycoprotein 
sialic acid-containing 
receptors 


PFBP 




Fn 


Protein Fl/Sfbl 


Fn 



integrins 



Cheng et al., 2002 
Botta, 1981 
Podbielskietal., 1999; 

Visai et al., 1995 
Teraoetal.,2001 
Teraoetal.,2002 
Courtney et al., 1996 
Gerlachetal., 1994 

Pancholi & Fischetti, 

1992, 1997 
Ashbaughetal., 2000; 

Cywes et al., 2000; 

Schrager et al., 1998; 

Wessels and Bronze, 1994 
Teraoetal.,2002 
Eisner et al., 2002 
Beachey and Ofek, 1976; 

Courtney et al., 1986 
Dunne etal., 1994 

Courtney et al., 1994; 

Okadaetal., 1994, 1995; 

Perez-Casal et al., 1995; 

Berkower et al., 1999; 

Waostromauo Tylewska, 

1982; Wang and Stinson, 

1994 

Ryan etal., 2001 
Rocha & Fischetti, 1999 
Hanski & Caparon, 1992; 

Selaetal., 1993; 

Molinari et al., 1997; 

Talayetal., 1993 
Ozeri et al., 1998; 

Okadaetal., 1998 

(cont.) 



a 
o 

> 

in 
H 
JO 
M 

H 
O 

n 

o 
n 
n 

> 

M 

< 
> 

I— I 
O 

Z 

o 

X 
o 

on 
H 

n 

M 
M 
M 

on 



Table 8.1. (cont.) 





Adhesin-Invasin 


Receptor 




References 






caveolin-1 




Rohde et al., 2003 






Fn-collagen interactions 


Dinklaetal., 2003 




Protein F2 


Fn 




Jaffe et al., 1996 




Pullulanase 


thyro globulin, mucin, fetuin 


Hytonenetal., 2003 




R28 


• 




Stalhammar-Carlemalm 
et al., 1999 




Sell, (SclA) 


• 




Rasmussen et al., 2000; 


3 


Scl2 (SclB) 


• 




Rasmussen & Bjorck, 
2001; Whatmore, 2001; 


l— i 








Lukomski et al., 2001 


CO 

1-1 


SpeB 


integrin 




Stockbauer et al., 1999; 


pq 

I— i 
M 




laminin, mucin, 


fetuin, 


Hytonenetal., 2001 


Q 
O 




thyrogobulin 






CO 

< 


Sfbx 


Fn 




Jengetal., 2003 


Q 


SOF 


Fn 




Kreikemeyer et al., 1995; 


4h 
< 

Q 








Rakonjacetal., 1995 


Z 

< 


Vn-binding protein 


Vn 




Valentin -Wiegand et al., 


pq 

z 








1988 




28-kDa protein 


Fn 




Courtney et al., 1992 


o 
u 


• 


cytokeratin 




Tamura & Nittayajarn, 2000 


CO 


• 


heparin sulfate 




Duensing et al., 1999 



< 



Notes: Fn = fibronectin, Fgn = fibrinogen, Vn = vitronectin, PFBP = S. pyogenes 
Fn-binding protein, FBP54 = Fn-binding protein 54, G3PDH = glyceraldehyde- 
3-phosphate-dehydrogenase, SOF = serum opacity factor, Scl = streptococcal 
collagen-like protein, Sfbx = streptococcal Fn-binding protein x, Lbp = laminin- 
binding protein, Fba = Fn-binding protein a, FbaB = Fn-bindin protein B, and 
Lsp = lipoprotein of S. pyogenes, ? = Unknown 



components of the extracellular matrix (Table 8.1). It has been proposed that 
adhesion of streptococci is a multistep process initiated by LTA, which facili- 
tates the binding of a second adhesin, such as M protein or protein F (Hasty 
et al., 1992). The secondary adhesin provides host-cell specificity and leads 
to high-affinity attachment. The list of adhesins in Table 8.1 emphasizes the 
fact that GAS utilize multiple adhesins to mediate attachment to host cells. 
There are conflicting reports on some of these adhesins. 




Figure 8.2. Scanning electron micrograph of GAS interacting with HEp-2 cells. M type 49 
S. pyogenes (A and B) or its nra mutant (C-E) were added to monolayers of HEp-2 cells. 
The attachment and internalization of streptococci are shown after 1 h (A) and 3 h (B) of 
infection. The microvilli were in close contact with bacteria (A) and no significant 
morphological changes were apparent in HEp-2 cells incubated with wild-type 
streptococci (B). The nra mutant exhibited a similar interaction pattern after 1 h of 
infection (C), whereas drastic changes were observed after 3 h of infection (D and E). The 
nra mutant induced large invaginations and was found to enter one such invagination and 
exit through another. No such alterations were noted with the wild-type parent. 
(Reproduced with permission from Molinari et al., 2001.) 




o 
o 

a 
> 

in 
H 
JO 
M 

H 
O 

n 

o 
n 
n 

> 

M 

< 

> 
00 
I— I 

o 
Z 
o 

X 
o 

on 
H 

n 

M 
M 
M 

on 



Some investigators found a role for M protein-CD46 interactions or 
hyaluronate-CD44 interactions in streptococcal adhesion to keratinocytes 
(Table 8.1), whereas others reported that inactivation of a variety of viru- 
lence factors, including M protein and capsule, did not decrease adhesion 
or colonization (Darmstadt et al., 2000; Jadoun et al., 2002; Ji et al., 1998). 



Although there is, as yet, no entirely satisfactory explanation for all of these 

discrepancies, it is important to note that the findings for one serotype of 

GAS cannot necessarily be extrapolated to all other serotypes. Moreover, not 

all serotypes of GAS have the same array of adhesins and not all adhesins will 

be expressed simultaneously (Mclver et al., 1995) . Which adhesin or adhesins 

that will be used by a particular serotype of GAS will depend on its repertoire 

of adhesin genes, on environmental signals, and on the receptors expressed 

by a particular type of host cell. 

Some, but not all, of the adhesins listed in Table 8.1 will initiate invasion 

of the host cells by streptococci. Whether or not a host cell is invaded will 

depend on the physiological status of the host cell and on the nature of the 

adhesin-receptor interactions. For example, the adhesion depicted in Fig. 8.1 

probably will not lead to invasion because most of the cells in the superficial 

5 layer of the buccal epithelium are dead and ready to be desquamated. HOW- 

CO 

w ever, GAS have been found to invade human tonsillar cells in vivo (Stenfors 

pq 

g et al., 2000) and to invade many types of tissue culture cells (Table 8.1). An 

example of adhesion that leads to internalization of streptococci is shown in 




o 

Ph 

CO 

3 Fig. 8.2. 

< 

< PATHWAYS FOR INFECTION 

h There are several possible pathways for infection that streptococci may 

o follow once they have attached to the epithelium (Fig. 8.3). One outcome is 

^ the persistence of the streptococci on the host epithelia. Under such circum- 

g stances, the bacteria will eventually face the activation of nonspecific, innate, 

ffi and specific host defense mechanisms. Such effects, as well as a potential 

induction of bacterial "apoptosis," can result in the complete clearance of 
the bacteria from the infected host; thus the self-limiting nature of the in- 
fection. Alternatively, the individual may become an asymptomatic carrier 
of the streptococci. Although the phenotype of streptococci within carriers is 
not known, certain virulence factors are probably downregulated during car- 
riage, resulting in less damage to the host and less activation of host defense 
mechanisms (Podbielski et al., 2003). 

A second pathway is the invasion of tissues and the metastatic spread 
of the streptococci. This may be accomplished by the release of degradative 
factors and toxins that can destroy both the intercellular substances and eu- 
karyotic cells, resulting in the spreading of the bacteria to new anatomical 
sites in the host. At the new sites, the bacteria are exposed to more favor- 
able conditions and can multiply rapidly and continue to spread to adjacent 
tissues. 




d 

O 
tj 



U 
U 

o 
u 

o 

-M 

Dh 

CU 

!m 
H- 1 
C/3 

<-w 

O 

C/} 

• •— i 

cu 
d 

cu 
&D 

o 

Dh 

a 

-M 

o 

CU 

u 

d 

-M 

w 
w 

J-H 

a 

Dh 

!-h 

03 

3 

cu 
u 



d 

o3 

d 

o 

•i-H 

w 

03 

d 

o 

••— i 

CO 

cu 

03 

O 
-M 

u 

o3 

Dh 



-M 
O 

"cu 

o 




O (D 



00 

CU 

J-H 

a 

•i-H 




A third pathway that can occur during the infectious process is the inter- 
nalization of the adherent bacteria by the host cells. In the intracellular space, 
the bacteria will survive provided they can evade or inhibit the activation of 
phagolysosomes. Thus, the intracellular space can be regarded as a sanctu- 
ary where the streptococci can avoid the defenses of the host or antibiotics. 
Indeed, the internalization of streptococci by host cells has been proposed as 
an explanation for recurrent infections after antibiotic treatment (Osterlund 
et al., 1997). Once inside the host cell, the streptococci probably downreg- 
ulate degradative and lytic factors while upregulating "persistence" factors. 
In response to a signal, as yet unknown, GAS will escape from their intra- 
cellular sanctuary by inducing an apoptotic response in host cells. Another 
form of invasion has been described for S. agalactiae, which can transverse 
a polarized epithelium without altering its structural integrity (Kallman and 

S Kihlstrom, 1997). However, no such transcytosis has been noted with GAS. 

w There are varying amounts of supporting data for each of these pathways . 

pq 

o That GAS initiate adhesion and persist at a local site for some time before 

S being eliminated has been observed since the days of ancient medicine and is 

£ regarded as a textbook standard (Bisno and Stevens, 2000). Similar observa- 

* tions have long been made on the invasion of host tissues and the metastatic 

§ spread of streptococci throughout the body of the host. 

< 

* However, the concept of an intracellular location and the intracellular 
h persistence of GAS is relatively new and does not have extensive, supporting, 
o clinical data. In fact, there are only three studies in which the intracellular 

u 

w presence of S. pyogenes was directly demonstrated in tissue specimens from 

g humans (Osterlund and Engstrand, 1997; Osterlund et al., 1997; Podbielski 

ffi et al., 2003). There are a number of studies in which clinical isolates of GAS 

were tested for their in vitro capability to enter host cells, and even more groups 
have demonstrated the ability of GAS to be internalized by cultured cell lines 
(see Cleary and Cue, 2000, for review). The epidemiologic and experimental 
investigations have focused, so far, on the effect of an intracellular location 
on the persistence of GAS in upper respiratory tract infections. There is no 
evidence that persistent and recurrent GAS infections of the skin, especially 
of the lower legs, are associated with an intracellular location of the bacteria. 

EXTRACELLULAR MECHANISMS OF TISSUE INVASION 

As just described, streptococci can invade tissues and can be internalized 
by cells within these tissues (LaPenta et al., 1994). Tissue invasion and inter- 
nalization are not mutually exclusive and may occur simultaneously. We first 
discuss the extracellular mechanisms utilized by GAS to invade tissues, and 




then we discuss the mechanisms required for internalization. One way that 
streptococci could invade tissues is to blast their way through the epithelial 
barrier by secreting cytolytic toxins that induce apoptosis of the host cells, 
thereby clearing a pathway to deeper tissues. Adhesion of streptococci to the 
host cell would help to effectively deliver the toxin(s) to the desired target 
(Ofek et al., 1990). Just such a mechanism has been described by Madden 
et al. (2001). These investigators found that the M -protein-mediated at- 
tachment of streptococci to keratinocytes was required for the insertion of 
streptolysin O (SLO) into cholesterol-rich domains in the cell membranes. 
The pore formed by SLO facilitated the delivery of streptococcal NAD- 
glycohydrolase (SPN) into the host cell, resulting in the lysis of host cells. 
There are several pathways whereby SPN may contribute to virulence. SPN 
can cleave /3-NAD, resulting in a decrease in the intracellular pools of NAD 

and an increase in nicotinamide, a vasoactive compound. SPN also forms £ 

o 

cyclic ADP-ribose and has ADP-ribosylating activities (Stevens et al., 2000). g 

Expression of both SLO and SPN were required for lysis of host cells. Interest- £ 

ingly, spn is located adjacent to slo, and these genes appear to be cotranscribed £ 

from a common promoter (Madden et al., 2001). o 

Other factors, in addition to SLO and SPN, are involved in the lysis of ° 

host cells. Sierig et al. (2003) reported that both SLO and SLS (streptolysin fc 

I— i 

S) contributed to the lysis of keratinocytes. These investigators also found % 

> 

that expression of SLO, but not SLS, impaired the killing of nonencapsulated S 

S. pyogenes by neutrophils. The reason why SLO did not impair the killing of * 

encapsulated streptococci is not known. ^ 

Another form of extracellular invasion, termed paracellular translocation, ™ 

was recently described by Cywes and Wessels (2001). In this case, adhesion 2 

to human keratinocytes was mediated by interactions between the hyaluronic w 

capsule of S. pyogenes and CD44, an integrin expressed on several cell types 
(for an integrin review, see Hynes, 2002) . The adhesion of encapsulated strep- 
tococci induced membrane ruffling and disruption of intercellular junctions. 
In some instances, the cell membranes curled back and lifted away from adja- 
cent cells, exposing a passage for streptococci around the cells and into deeper 
spaces. The soluble form of hyaluronate also induced similar responses in 
host cells. It was suggested that the binding of hyaluronate to CD44 acti- 
vated Racl, a member of the Rho family of GTPases, to phosphorylate a yet 
unidentified protein that recruited ezrin, an actin linker protein. These inter- 
actions triggered the cytoskeletal movements resulting in alterations of cell 
morphology and intercellular junctions. 

In addition to the membrane effects described herein, this activation 
of the cytoskeleton also caused the formation of lamellipodia. Encapsulated 




streptococci were found to primarily associate with these lamellipodia, but 
such interactions did not lead to internalization of the streptococci. Interest- 
ingly, an acapsular mutant did not induce such changes, and these strep- 
tococci bound primarily to membrane sites devoid of the lamellipodia. The 
fates of these two different phenotypes of streptococci were also different. 
Instead of remaining extracellularly, the nonencapsulated streptococci were 
internalized by the keratinocytes. It was concluded that the hyaluronic acid 
capsule contributes to translocation by two mechanisms (Cywes and Wessels, 
2001). In one, the hyaluronic acid capsule disrupts intercellular junctions and 
thereby facilitates the passage of streptococci into deeper tissues. In the sec- 
ond, the capsule prevents streptococci from being internalized and trapped 
within epithelial cells. As a corollary, the acapsular mutant failed to be effi- 
ciently translocated, because it became internalized and trapped within ep- 
S ithelial cells. It is interesting to note that the acapsular mutant was more 

w cytolytic than the encapsulated wild type, presumably because the acapsular 

pq 

o mutant was internalized more efficiently. 

S Another intriguing hypothesis concerning the role of ezrin and the strep- 

£ tococcal inhibitor of complement (SIC) in the fate of streptococci that invade 

* human tissues was recently proposed by Hoe et al. (2002). Inactivation of 

§ the sic gene enhanced adhesion to host cells, suggesting that the SIC protein 

< 

x interfered with adhesion in some manner. SIC was found to rapidly enter 

h epithelial cells and to react with ezrin and moesin. The mode of entry is not 

o known, but perhaps the cytolysin-mediated transfer mechanism described 

u 

w by Madden et al. (2001) may have a role. Because ezrin, radixin, and moesin 

g (ERM) are involved in the formation of lamellipodia, filopodia, and microvilli 

<3 _ 

ffi through linkages with actin filaments (Bretscher et al., 2002), it was sug- 

gested that the binding of SIC to ERM alters actin-mediated formation of 
microvilli at the cell surface. The implication is that such disruptions may 
also prevent receptor clustering at lamellipodia, which may be required for 
efficient adhesion. 

SIC was also found to impair phagocytosis of the Ml serotype by neu- 
trophils. It was proposed that the ability of SIC to impede adhesion provided 
a survival advantage within the host (Hoe et al., 2002). It is also possible that 
SIC could have a role in the dissemination of streptococci. Adhesion may 
occur while SIC is downregulated, and upregulation of SIC may release the 
streptococci and enhance their ability to spread to other tissues within the 
host or to other hosts. In this regard, it would be interesting to determine if 
the application of SIC to animals that are colonized with S. pyogenes would 
attenuate or enhance the virulence of the streptococci. It should be noted that 
sic has been found only in Ml, M12, M55, and M57 serotypes (Akessonetal., 
1996; Hartas and Sriprakash, 1999). 



Although the effects of hyaluronic acid and SIC on adhesion were dif- 
ferent, their effect on internalization of the streptococci was the same, that 
is, inhibition of uptake. In each case, the ERM proteins were found to be 
intimately involved. It is not yet clear how hyaluronate recruits actin to help 
in the formation of lamellipodia yet prevents actin from participating in the 
uptake of streptococci. The effects of hyaluronic acid and SIC on epithelial 
cells are similar to those seen in cells in which the gene for moesin was in- 
activated (Speck et al., 2003), which also resulted in the loss of polarity and 
intercellular junctions. In addition, these moesin-deficient cells were found 
to detach and migrate. This finding suggests an intriguing possibility - that 
the invasion of host cells by GAS could result in the detachment of a portion 
of these cells and their migration, resulting in the spread of the internalized 
streptococci to other tissues. 

MECHANISMS FOR INTERNALIZATION OF STREPTOCOCCI BY 
HOST CELLS 



o 
o 

a 
> 

in 
H 
JO 
M 



Several different mechanisms have been described for the internalization o 



o 
n 



of S. pyogenes by host cells . One of the most common mechanisms is that me- ° 

diated by adhesins that interact with extracellular matrix proteins (Table 8.1). fc 

Our discussion is limited primarily to adhesins that are known to be involved % 

> 
in the internalization of the streptococci. Future research will undoubtedly S 

reveal additional adhesins that contribute to the uptake of streptococci by * 

host cells. ^ 

o 

in 
H 

n 

Protein F/Sfbl p 

Protein Fl/Sfbl is a large adhesin that is anchored to the bacterial cell 
surface by an LPXXG motif. Although the first description of this protein used 
the designation Sfbl (Talay et al., 1991), most publications have used protein 
Fl, which is the designation that is used in the following sections. Protein Fl 
promotes adhesion and invasion of host cells by interactions with fibronectin. 
The binding of fibronectin to protein Fl involves two binding domains - a 
repeating peptide domain that is immediately preceded by an upper binding 
domain (UBD) . The effect of these domains on binding is sequentially exerted 
(Talay et al., 2000). The second binding step appears to trigger a conforma- 
tional change in fibronectin that exposes its cell attachment domain (RGD), 
which is cryptic in soluble fibronectin (Tomasini-Johansson et al., 2001). The 
exposed RGD domain of fibronectin can then interact with integrins on the 
surface of the host cell and trigger internalization of the bacterium. Protein- 
Fl -mediated invasion of host cells may also require interactions between the 



UBD of protein Fl and the collagen-binding domain of fibronectin for the 
optimal uptake of streptococci (Dinkla et al. 2003). Recently, an alternative 
internalization mechanism involving protein Fl and caveolin-1 on the mem- 
branes of host cells and the formation of omega-shaped cavities was identified 
(RohdeetaL, 2003). 

Although the in vitro data on the involvement of protein Fl in the in- 
ternalization process are very convincing, the epidemiological data are not 
as strong. Protein Fl is not an absolute requirement, because up to 48% 
of the GAS strains do not carry a prtFl gene (Natanson et al., 1995). Some 
investigators found an association between the expression of protein Fl and 
internalization (Neeman et al., 1998), whereas others found no such associ- 
ation (Brandt et al, 2001). 



2 M proteins 

w 

I— I 
pq 

g M proteins are alpha-helical, coiled-coil proteins that are one of the major 

w virulence factors of the organism and contribute to the ability of streptococci 

£ to evade phagocytosis and to attach to host cells (reviewed by Fischetti, 1989; 

5 Cunningham, 2000). A number of M proteins are adhesins (Table 8.1) and 

§ at least two of the M proteins, Ml and M3, bind fibronectin (Schmidt et al., 

< 

g 1993; Cue et al., 1998) . An M type 1 S. pyogenes bound fibronectin and invaded 

h human lung cell cultures, whereas an M -negative mutant had a reduced ca- 

o pacity to bind fibronectin and to invade these cells (Cue et al., 1998) . Laminin 

u 




CO 

>1 



was also found to promote invasion by the same M type 1 S. pyogenes, indicat- 
% ing that a single strain can utilize multiple pathways that lead to invasion of 

ffi host cells. It was suggested that invasion is mediated by interactions between 

M protein and either fibronectin or laminin. The bound fibronectin then re- 
acts with integrins on the cell surface that induce actin polymerization and 
uptake of streptococci by a zipper-like mechanism (Dombek et al., 1999). The 
M3 protein also mediates adhesion to HEp-2 cells and HaCat keratinocytes 
(Berkower et al., 1999). The expression of the M3 protein led to an increase in 
the invasion of HaCat cells, but the M3-protein-mediated invasion of HEp-2 
was dependent on the presence of serum. Although it was not determined 
what component of serum was responsible for this invasion, it is likely that 
fibronectin will play a role similar to that found for Ml protein. 

It is clear that certain M proteins that bind fibronectin can initiate adhe- 
sion and participate in the invasion process. However, most M proteins do 
not bind fibronectin. Do these M proteins contribute to invasion of host cells? 
Only a few such types of M proteins have been investigated for their role in 
the invasion process. The most widely investigated is the type M6 protein, 




but the results and conclusions of different investigators varied consider- 
ably. Fluckiger et al. (1998) found that inactivation of emm 6 had no effect on 
streptococcal adhesion but dramatically reduced invasion of Detroit 565 pha- 
ryngeal cells. However, a more common finding is that protein F is the major 
component responsible for invasion and that the M6 protein contributes to 
a lesser degree (Jadoun et al., 1997). 

One of the ways that the M6 protein can contribute to invasion is to in- 
crease the numbers of adherent streptococci, which can lead to an increase in 
the numbers of streptococci that invade the host cells by a protein- F-mediated 
mechanism. This possibility is supported by the findings of Okahashi et al. 
(2003), who reported that inactivation of emm 6 reduced adhesion to mouse 
osteoblasts by ~90% but only reduced invasion of these cells by ~ 35%. In 
contrast, inactivation of prtf had no effect on adhesion but reduced invasion 

by ~90%. In addition, the introduction and expression of emm6 in S. gor- £ 

o 

donii led to an increase in adhesion but not invasion of HEp-2 tissue culture g 

cells (von Hunolstein et al., 2000). However, the introduction and expression £ 

of emm 6 in Enterococcus faecalis led to an increase in invasion of host cells £ 

(Okadaetal, 1998). § 

From the foregoing paragraphs, it appears that the background in which ° 

M proteins are presented can have a dramatic impact on its role in the inva- fc 

I— i 

sion of host cells. To evaluate the roles of several different M proteins in the % 

> 

same genetic background, Berkower et al. (1999) introduced emm3, emm6, S 

and emm 18 into an isogenic M -negative and protein- Fl -negative mutant and * 

tested the recombinant strains in adhesion and invasion assays. Their results ^ 

indicated that M3, M18, and M6 proteins do mediate adhesion to HEp-2 cells 3 

and to HaCat keratinocytes. These findings are similar to those of Court- 2 

ney et al. (1997a), who found that introduction of emm 18 and emmS into an w 

M -negative strain restored adhesion to HEp-2 cells. However, the role of M 
proteins in the invasion of host cells appears to depend on the type of host 
cells and on the serotype of the M protein. Berkower et al. (1999) reported 
that the expression of the M6 protein, but not M3 and M18 proteins, restored 
the ability to invade HEp-2 cells. In contrast, expression of both M6 and M3 
proteins, but not the M18 protein, led to invasion of HaCat keratinocytes. 
The effect of serum on invasion was dependent on the M type. M6- and M 18- 
protein-mediated invasion of HEp-2 cells was unaffected by serum, whereas 
invasion by streptococci expressing the M3 protein was enhanced by serum. 
These data indicate that the ability of M proteins to act as adhesins and in- 
vasins can be dependent on M type; that is, it will depend on the variable 
sequences in the N-terminus of the molecule (Fig. 8.4). Furthermore, strep- 
tococci expressing different M proteins can bind to different receptors on the 




GO 

1-1 
w 
I— I 
PQ 
Q 
O 
Ph 

CO 

< 
W 

Q 

< 
Q 

< 
w 

O 

u 

CO 

Pi 
< 

X 



M-type-dependent adhesion 

Hypervariable Domain 

Variable Domain 



M-type-independent adhesion 



Conserved Domain 




Cell wall spanning domain 
LPXTG anchor motif 



Figure 8.4. Schematic of M proteins and their role in adhesion and invasion of host cells. 
M protein is the primary virulence factor of S. pyogenes and confers the abilities to resist 
phagocytosis in blood and to attach to host cells. M protein is an a -helical, coiled-coil 
protein that emanates from the surface of the bacteria. The number and sequence of the A 
and B repeats will vary depending on the M type. The C repeats are conserved among 
different M types. Adhesion to host cells can be mediated by either the variable domain or 
the conserved domain, depending on the receptors that are expressed by host cells. Host 
cells expressing CD46 can react with the C-repeat domain and mediate adhesion that is 
independent of the M type. Adhesion and invasion mediated by the variable domain will 
depend on the type of M protein expressed by the streptococcus and on the kinds of 
receptors expressed by the targeted tissues. The binding of fibronectin to M proteins (and 
therefore the internalization of streptococci by host cells) is mediated by the variable 
domain of M proteins. (Reproduced with permission from Courtney et al., 2002.) 

same host cell. Berkower et al. (1999) found that recombinant strains express- 
ing the M6 protein did not inhibit adhesion-invasion of HaCat keratinocytes 
by a recombinant strain expressing the M18 protein, implying that the M6 
and M18 proteins bind to different receptors on these keratinocytes. 

Frick et al. (2000) described another property of M proteins that promotes 
adhesion and invasion. They found a 19 amino acid consensus sequence in 
certain M and M-like proteins that promotes interbacterial aggregation and 
the attachment of microcolonies to host cells. Such attachment also enhanced 
invasion of host cells. A synthetic peptide copying this aggregative sequence 
not only inhibited adhesion and invasion of host cells but also dramatically 
increased the survival of mice challenged with M type 1 S. pyogenes. 



FbaA 

FbaA, originally termed OrfX (Podbielski et al., 1996), is a 38-kDa protein 
with an LPXTG cell-anchoring motif (Terao et al., 2001). FbaA is expressed 
by M serotypes 1, 2, 4, 22, 28, and 49, but not by other M serotypes or by 



C5a peptidase 




streptococcal groups B, C, and D. It has three or four proline-rich repeat do- 
mains that are highly homologous to the fibronectin-binding repeats found 
in FbpA. Similar to many other fibronectin-binding proteins, expression of 
FbaA promoted the streptococcal invasion of HEp-2 tissue culture cells. Al- 
though it was not determined how the binding of fibronectin to FbaA led 
to invasion, it is likely that fibronectin served as a cross-linking agent as 
described herein. A comparison of invasion of a multigene activator (Mga)- 
negative mutant and a FbaA-negative mutant with the parental Ml strain 
suggested that other surface proteins under the control of Mga may also be 
involved. This finding is not surprising, because others have reported that 
the Ml protein is under control of Mga and that it mediates adhesion and 
invasion of host cells (see earlier paragraphs) . The FbaA-negative mutant was 
also less virulent in a mouse model than its wild-type parent, suggesting that 

FbaA is a virulence factor. £ 

o 

a 

> 

in 
H 
JO 
W 

The scpA gene encodes the C5a peptidase, which degrades the chemoat- o 

tractive complement factor C5a and thereby inhibits migration of neutrophils ° 

into an infected area. The C5a peptidase also binds fibronectin and is involved fc 

in adhesion and invasion of host cells by group B streptococci (Cheng et al., % 

> 

2002) . Whether it functions as an invasin in GAS remains to be demonstrated. S 

The scpa gene is cotranscribed with the JbaA gene. Transcription attenuation * 

in GAS was first demonstrated in the scpa promoter (Pritchard and Cleary, ^ 

1996) and later in the downstream JbaA promoter (Podbielski et al., 1996). 3 

There are no known stimuli that derepress this transcription attenuation. 2 

M 

FbaB and other RGD-containing proteins 

FbaB is a newly described 79-kDa, fibronectin-binding protein found only 
in M 3 and M 18 serotypes of S. pyogenes (Terao et al., 2002) . FbaB has a repeated 
peptide that binds fibronectin and has homology to the fibronectin-binding 
repeats of other streptococcal proteins. In addition, the N-terminal domain of 
FbaB has ^45% homology with an N-terminal region of protein Fl, protein 
F2, and Cpa. Inactivation of the gene for FbaB resulted in decreased adhesion 
to and invasion of HEp-2 epithelial cells. The FbaB-negative mutant was 
less virulent in mice-challenged IP, indicating that FbaB is another virulence 
factor. Relative to its virulence, the FbaB protein was expressed mainly by type 
M3 and M 18 S. pyogenes isolated from toxic-shock-like syndrome patients. No 
expression of FbaB was found in pharyngeal isolates of M types 3 and 18. 




CO 

h-l 



These findings suggest that FbaB may contribute to the invasive potential of 
GAS. 

One interesting structural feature of this protein is that, in addition to 
its fibronectin-binding repeat peptide, FbaB also contains the RGD peptide. 
As already described, many of the adhesins interact with fibronectin, which 
then interacts, by means of its RGD domain, with integrins on host cell mem- 
branes. Thus, a streptococcal surface protein that contains the RGD motif may 
be able to bind directly to integrins without the need for a cross-linking agent. 
For example, a variant of SpeB that contained the RGD sequence is expressed 
by a highly virulent clinical isolate of M type 1 S. pyogenes (Stockbauer et al., 
1999). This speB variant, called mSpeB2, was found to interact with human 
integrins a Y p$ and ot\\hP?> and to mediate adhesion to host cells. However, it 
is not known if the RGD sequence in FbaB or mSpeB2 has a direct role in 
the invasion of host cells. 
w Another RGD-containing protein of GAS is the Mac protein (Lei et al., 

pq 

o 2002). This protein was so named because of its homology to the human 

S Mac-1 leukocyte integrin, but it has also been termed IdeS for its IgG de- 

£ grading activity (von Pawel-Rammingen et al., 2002). The streptococcal Mac 

* protein is secreted and is involved in the resistance of streptococci to phago- 

§ cytosis. Variants of Mac that contained the RGD sequence were found to bind 

< 

x to different integrins and to block uptake of streptococci by neutrophils, but 

h it is not known if the secreted Mac interferes with adhesion and invasion of 

o host cells other than professional phagocytes. 

u 

CO 

B 

< Alternative mechanisms for internalization 

m 

Invasion of host cells can also be triggered by interactions that do not 
involve integrins or fibronectin. Such alternative pathways could be serotype 
specific and involve the formation of large invaginations (Molinari et al., 
2000). For the majority of strains, the responsible mechanisms remain to be 
elucidated. One potential bacterial receptor candidate is the Lsp (for lipopro- 
tein of S. pyogenes) surface protein. 

Lsp is a recently described surface molecule involved in adhesion and 
invasion of host cells by GAS ( Eisner etal., 2002). Lsp is a member of the Lral- 
lipoprotein family that is expressed by a variety of streptococci. Lsp was found 
to bind directly to laminin, suggesting that laminin-Lsp interactions may 
have a role in adhesion and invasion. Lsp-negative mutants neither attached 
as well nor invaded host cells as well as the parental M49 strain. However, 
the Lsp-negative mutant also demonstrated reduced binding of fibronectin. 



Table 8.2. Responses of host cells to group A streptococci 







Response of 




M type 


Host cells 


host cells 


References 


M6 


keratinocytes 


IL-la, IL-1/3, 
IL-6, IL-8 


Wang et al., 1997 






prostagland in E 2 


Ruiz et al., 1998 


M24 


HEp-2 cells 


IL-6 


Courtney et al., 1997b 


M? 


HEp-2 cells 


NF-/cB 


Medina et al., 2002 


M49, M52 


keratinocytes 


p defensin 2 


Dinulos et al., 2003 


M22, M54, 


mouse 


cathelicidin 


Dorschner et al., 2001 


M59 


keratinocytes 






Ml 


A-549, HEp-2 cells 


apoptosis 


Tsaietal., 1999 


M3 


human blood cells 


CDlllb, ROS, I 
CD62L 


Saetre et al., 2000 


Ml 


human mononuclear 


IL-1/3, IL-6, TNFcx, 


Miettinenetal., 1998 




cells 


IL-12, 
IL-18, IFN-y 




M6 


Detroit 562, FaDu 


phosphorylation 


Pancholi & Fischetti, 




cells 




1997 


M6 


HEp-2 cells 


caspase 9, apoptosis 


Nakagawa et al., 2001 


M49 


HEp-2 cells, 


annexin V, caspases 


Kreikemeyer et al., 




keratinocytes 


2,3,7 


2003 



Because Lsp does not react directly with fibronectin, this suggested that, in 
addition to Lsp, the expression of other surface proteins may have been altered 
in the Lsp-negative mutant. Thus, the role of Lsp in streptococcal adhesion 
and invasion remains to be clarified. 




o 
o 

a 
> 

in 
H 
JO 
M 

H 
O 

n 

o 
n 
n 

> 

M 

< 
> 

I— I 

o 
Z 
o 

X 
o 

on 
H 

n 

M 
M 
M 

on 



RESPONSES OF THE HOST TO STREPTOCOCCAL ADHESION 
AND INVASION 

The attachment of GAS to host cells is known to stimulate responses from 
host cells (Table 8.2) . For example, the adhesion of an M type 24 S. pyogenes to 
HEp-2 cells resulted in the release of IL-6, whereas an isogenic, M24-protein- 
negative mutant did not attach to these cells and failed to stimulate the release 
of IL-6 (Courtney et al., 1997). Similarly, the adhesion of a type M6 S. pyogenes 




to cultures of HaCat keratinocytes caused a dramatic increase in the release 
of IL-la, IL-1/3, IL-6, IL-8, and prostaglandin E 2 (Wang et al, 1997; Ruiz 
et al., 1998). In contrast, an M6 protein-negative mutant either caused no 
such increase or did so at later time points. 

Wang et al. (1997) reported that adherent streptococci caused injury to 
keratinocytes that coincided with the release of cytokines. Subsequent re- 
search by Ruiz et al. (1998) indicated that both SLO expression and adhesion 
were required for the stimulation of IL-6, IL-8, and prostaglandin E 2 expres- 
sion in HaCat keratinocytes. The role of SPN in the SLO-mediated stimulation 
of the cytokines was not determined. Interestingly, expression of SLO had 
no impact on the stimulation of IL-1 by adherent streptococci (Ruiz et al., 
1998), suggesting that other streptococcal products may stimulate the IL-1 
response. Indeed, there are a number of streptococcal products including 
S LTA, DNA, and a variety of streptococcal pyrogenic exotoxins that can evoke 

w cytokine responses in the host. These responses are thought to be a central 

pq 

o cause of morbidity and mortality associated with toxic shock syndrome and 

S necrotizing fasciitis. 

£ Streptococci can also stimulate phosphorylation of proteins in host cells. 

* SDH (streptococcal dehydrogenase) was identified as one streptococcal pro- 

§ tein that can initiate phosphorylation events in host cells (Pancholi and 

< 

x Fischetti, 1992, 1997). SDH is a 35.8-kDa glyceraldhyde-3-phosphate dehy- 

h drogenase that is expressed on the surface of streptococci. It also has ADP- 

o ribosylating activity and binds to a variety of mammalian proteins. SDH was 

u 

w found to react with 32-kDa membrane proteins in Detroit 562 and FaDu 

g tissue culture cells derived from human pharyngeal carcinomas. The inter- 

ffi action of SDH and M type 6 S. pyogenes with these pharyngeal cells stimulated 

the phosphorylation of histone H3. SDH did not induce phosphorylation in 
other tissue culture cells such as Chang cells, Chinese hamster ovarian cells, 
and human epitheloid carcinoma cells. It was postulated that the induction of 
tyrosine kinases and protein kinase C may regulate actin polymerization and 
invasion of host cells by streptococci. The findings that genistein, a tyrosine- 
kinase inhibitor, and staurosporine, a serine/threonine-kinase inhibitor, had 
no inhibitory effect on adhesion, yet almost completely blocked invasion, 
clearly indicate that phosphorylation of host proteins is required for invasion 
(Pancholi and Fischetti, 1997). 

The adhesion of streptococci induces other host cell responses, including 
the production of IL-la andTNF-a (Table 8.2). Darmstadt etal. (1999) demon- 
strated that, in addition to their proinflammatory role, these cytokines inhib- 
ited the adhesion of S. pyogenes to keratinocytes. The mechanism whereby 
IL-la and TNF-a blocked adhesion is not known. 




The interaction of streptococci with host cells also enhanced the expres- 
sion of the transcription factor known as nuclear factor kappa B (NF-/cB; 
Medina et al., 2002; Okahashi et al., 2003). Adhesion induced an initial spike 
of transcription, whereas internalization of the streptococci resulted in a sus- 
tained NF-a:B response that was blocked by cytochalisin D. LTA is one of the 
streptococcal components that can contribute to this response. LTA induced 
the expression of NF-/cB and COX 2 expression in a human pulmonary cell 
line, and this expression was blocked by genistein (Lin et al., 2002), again 
indicating a role for tyrosine kinases in host cell responses to streptococci 
and their products. 

Bacterial interactions with host cells can also induce the expression of 
antibacterial peptides, which is part of the innate immune defenses. There 
are presently four main groups of antibacterial peptides: a-defensins, f5- 

defensins, 0-defensins, and cathelicidins. GAS stimulated the production £ 

o 

of cathelicidin in vivo in mouse keratinocytes (Dorschner et al., 2001). Cathe- g 

licidin in the \-iuM range inhibited the growth of S. pyogenes, indicating that £ 

it has the potential to prevent or mitigate streptococcal infections. To eval- £ 

uate this potential, Nizet et al. (2001) compared the invasiveness of GAS in o 

Cnlp -null mice and wild-type mice by using a subcutaneous model of in- ° 

fection. Cnlp encodes the antimicrobial peptide CRAMP, the mouse form fc 

of cathelicidin. The lesions induced by S. pyogenes in wild-type mice were % 

> 

almost completely resolved by day 8 but were still readily apparent in Cnlp- S 

null mice. These investigators also showed that a CRAMP-resistant strain of z 

S. pyogenes caused greater necrosis than wild-type S. pyogenes in their sub- ^ 

cutaneous model of infection. These data indicate that cathelicidins are an 3 

important part of the innate immune response to streptococcal infections of 2 

the skin. w 

GAS also induced, albeit poorly, the production of /3-defensin 2 in hu- 
man keratinocytes (Dinulos et al., 2003). That S. pyogenes is a poor stimulator 
of /?-defensins is somewhat unexpected, because S. pyogenes induces the ex- 
pression of NF-/cB (Medina et al., 2002), which can upregulate the expression 
of defensins. However, it was suggested that this lack of stimulation of f5- 
defensins could contribute to the virulence of S. pyogenes in skin infections, 
because ^-defensin 2 is a potent killer of S. pyogenes. 



RESPONSES OF GAS TO INTERACTIONS WITH HOST CELLS 

The interaction of streptococci with host cells not only induces a response 
in the host cell but also induces a response in the streptococci. Broudy et al. 
(2001, 2002) reported that coculturing a type M6 S. pyogenes with Detroit 562 




pharyngeal cells induced the expression of a phage-encoded streptococcal 
pyrogenic exotoxin C (SpeC) and DNase (called Spdl). SpeC and Spdl are 
coexpressed from the same transcript. The coculturing of S. pyogenes strain 
CS112 with pharyngeal cells also resulted in the induction of lysogenic bacte- 
riophage particles. These investigators went on to show that a factor, released 
by the pharyngeal cells, was responsible for these responses. The soluble fac- 
tor was <10 kDa and resistant to heat and proteolytic degradation. 

The expression of a variety of virulence factors is upregulated when GAS 
invade the host. The bacterial factors regulated in response to invasion versus 
those regulated in response to internalization with host cells have to be dis- 
tinguished from each other. Investigators have long known that expression 
of M protein increases when the bacteria enter the bloodstream. Gryllos et al. 
(2001) found that introduction of GAS into the pharynx of baboons induced 
S expression of the hyaluronic acid capsule. A similar increase in capsule pro- 

w duction was found when the streptococci were introduced into the blood and 

pq 

o the peritoneal cavity of mice. However, it is not entirely clear if this induction 

S of virulence factors streptococci is in response to signals from the host or is 

£ due to the growth phase of the streptococci. The expression of the capsule 

* and a variety of other virulence factors are upregulated during the exponential 

§ growth phase and downregulated during the stationary phase (M elver and 

i Scott, 1997; Gryllos et al., 2001). 

h Only recently, some data on bacterial virulence gene expression were 

o collected from GAS during their internalization or intracellular persistence. 

u 

^ During such behavior, it seems favorable for the bacteria to suppress the 

g expression of secreted lytic factors. This suppression is at least partially con- 

ffi trolled by the negative global regulator Nra (Podbielski et al., 1999) , a member 

of the RALP (RofA-like proteins) regulator family (Fogg et al., 1997; Granok 
et al., 2000) . Inactivation of Nra in an M49 strain greatly increased the cytolysis 
of HEp-2 cells. Utilizing whole genome array hybridization, Nra was shown 
to regulate between 120 and 290 genes depending on the growth phase and 
filtering levels used (Podbielski et al., unpublished results). Approximately 
two thirds of the genes are negatively regulated by Nra and include the cap- 
sule synthesis operon has, the genes for surface factors such as protein F2, 
Cpa, ScpA, SOF/Sfbll, SfbX, SclA, and SclB; and genes for secreted proteins 
such as the CAMP factor, SLS, SpeB, and SpeA. Consequently, inactivation 
of nra resulted in an increase in the expression of these genes. However, 
the expression of SLO was unaffected. It is interesting to note that the Nra- 
negative mutant induced large invaginations in the membranes of HEp-2 
cells, and chains of streptococci were found to enter one such invagination 
and resurface through another (Fig. 8.2). No such phenomenon was seen 
with the parent. 



Another predominantly negative regulator is encoded by the Fas-regulon. 
It comprises two histidine kinases, FasB/FasC, the response regulator, FasA, 
and an effector RNA, FasX. Judged by its sequence, its temporal expression 
pattern with a maximum at the end of the exponential growth phase, and the 
type of dependent genes (positively regulated secreted factors such as SLS, 
streptokinase, and SpeB; negatively regulated surface proteins-adhesins such 
as the M-related protein, SCPA, and FBP54), the Fas-regulon resembles the 
Staphylococcus aureus Agr-regulon. However, unlike the Agr-regulon, the Fas- 
regulon has not been demonstrated to be involved in a cell-density-dependent 
type of regulation (Kreikemeyer et al., 2001). A recent whole genome array 
hybridization analysis revealed that 18 to 90 genes are influenced by a func- 
tional Fas-regulon. This fact, as well as the potential convergence of several 
signaling pathways in the two histidine kinase sensors of the Fas-regulon, 

indicate that there is probably more than growth-phase-dependent regulation § 

o 
that is exerted by this regulon. g 

By using various Fas-regulon mutants and complemented strains, re- ^ 

searchers recently showed that the Fas-regulon exerts a negative control of £ 

bacterial fibronectin binding and adhesion to keratinocytes and H Ep-2 epithe- o 

Hal cells - albeit at different time points of the growth curve (keratinocytes, ° 

exponential phase; HEp-2 cells, stationary phase). Inactivation of the Fas- £ 

I— i 

regulon not only enhanced adhesion but also increased internalization of % 

> 

the streptococci. Once internalized, the mutants exhibited a strongly reduced ~ 

induction of apoptotic pathways as measured by annexin V, and caspase-2, * 

-3, and -7 expression (Kreikemeyer et al., unpublished data). The GAS iso- % 

late used for this analysis was a serotype M49 strain. Utilizing a serotype 3 

M6 strain, Nagakawa et al. (2001) found a somewhat different pattern of 2 

apoptosis induction that relied on caspase-9 activation and subsequent se- w 

lective degradation of mitochondrial functions. Because the M6 strain also 
carried a typical Fas-regulon, there could more than one way that GAS induce 
apoptosis. 

GAS exhibit the highest tendency for internalization once they enter 
the stationary phase. The Fas-regulon and the nra gene are maximally tran- 
scribed during the transition from exponential to stationary growth phase. 
This temporal expression profile fits well into the concept of FasBCAX and 
Nra (and other RALPs, Beckert et al., 2001) as central regulators that keep 
the delicate balance between extracellular aggressiveness and intracellular 
"tameness" that would be necessary for the bacteria to best survive in these 
different environments. 

Voyich et al. (2003) used a gene array encompassing the whole genome 
of an Ml serotype of GAS to probe gene expression during phagocytosis by 
human neutrophils. A two-component regulatory system, Ihb-Irr regulon, 



was identified that was expressed during phagocytosis. The Ihb-Irr system 
played a fundamental role in the survival of GAS within phagocytes. In addi- 
tion, the in gene was highly expressed by multiple serotypes of GAS during 
infections of the human pharynx, indicating that it regulates gene expression 
in response to interactions with the host in vivo. 

INTRASPECIES AND INTERSPECIES SIGNALING BETWEEN 
STREPTOCOCCI 

Cell-to-cell signaling among bacteria is a relatively new discovery and 
has been referred to as quorum sensing to indicate that thresholds of signals 
are achieved only when a sufficient density of bacteria is reached. These sig- 
naling peptides have also been called pheromones because of their role in 
2 regulating competence and conjugation. For many Gram-positive bacteria, 

w this signal will consist of small, linear, or cyclic peptides (reviewed by Dunny 

pq 

g and Leonard, 1997). Upton et al. (2001) described a lantibiotic peptide, sail- 
er varicin A (SalA), that is secreted by streptococci and whose expression is 
£ autoregulated. The secreted form of this peptide, called SalAl in GAS, not 
5 only upregulated expression of salAl in S. pyogenes but also upregulated the 

§ expression of salA in S. salivarius. Similarly, SalA upregulates expression of 

< 

g salAl in S. pyogenes. In addition to autoregulation, SalA/ SalAl also induces 

h the expression of an immunity factor that is required for resistance to the 

o antibiotic effect of SalA. Thus, strains of streptococci that do not produce 

u 




CO 

>1 



SalA/SalAl are sensitive to its antibiotic effect. Hence, strains of streptococci 
% that express SalA/SalAl could modulate the bacterial population in the host. 

X Although the effect of adhesion on secretion of SalA has not been determined, 

adhesion may enhance secretion as a result of an increase in the density of 
the bacterial population. In addition, adherent bacteria can multiply and form 
microcolonies that are very dense. 

Another factor that seems to be involved in interbacterial signaling with 
effects on the GAS virulence in animals, as well as on a putative natural com- 
petence of the bacteria, is the Sil/Blp-like system (Hidalgo-Grass et al., 2002; 
Smoot et al., 2002). The silA-E genes encode a regulatory secreted peptide, 
a potential processing machinery, a two-component regulatory system, and 
another protein with an unknown function. The sil locus resembles the S. au- 
reus agr and S. pneumoniae com quorum-sensing regulons. Inactivation of silC 
reduced the invasiveness of an emml4 strain of S. pyogenes in a mouse model. 
Coinjection of the silC mutant with an avirulent emm 14- negative mutant re- 
stored virulence and invasiveness. It was suggested that the emm 14-negative 
mutant complemented the defect of the silC mutant by secreting a signaling 



peptide that activated silA/B, which then activated other genes required for 
invasion. So far, sil has been identified in only two serotype strains, M14 and 
M18. It apparently resides on a 13.8-kb genomic island that is not contained 
in two other genomes of S. pyogenes. Exactly how the various components of 
the sil locus interact with each other, and with other genes involved in the 
infectious process, remains to be elucidated. Sil may be a target for novel 
therapies because of its involvement in the regulation of factors required for 
invasion. 




ASSAY FOR INTERNALIZATION 

The most widely used assay for internalization of bacteria by host cells is 
the antibiotic protection assay. This assay is based on the assumptions that 

internalized bacteria will not be killed by the addition of antibiotics to the % 

o 

medium, and that only those bacteria on the surface will be killed. A measure g 

of both adhesion and internalization can be obtained by determining the £ 

total number of CFU before and after treatment with antibiotics. However, £ 

investigators utilizing this assay should be aware of a potential problem that o 

has been described by Molinari et al. (2001). These investigators found a low ° 

rate of internalization of streptococci using the antibiotic protection assay, £ 

I— i 

but double-staining, immunofluorescent, microscopic analysis indicated that % 

> 

the internalization rate was actually high. It was suggested that the host S 

cell membrane is damaged during invasion, allowing antibiotics to enter * 

host cells and kill the bacteria. Thus, investigators need to ensure that the ^ 

integrity of the host cell membrane is maintained during the assay to obtain 3 

valid results. The antibiotic protection assay also cannot distinguish between 2 

the rate of uptake of bacteria and the rate of intracellular multiplication, or w 
between adherent bacteria and those that have been internalized and released 
by apoptosis. 



FUTURE CONSIDERATIONS 

It is clear that many types of eukaryotic cells can internalize streptococci 
and that GAS can invade host cells in vivo. Some of the molecular mechanisms 
involved in these processes have been identified. Additional research has to 
be done to sort out the adhesins and mechanisms of invasion that are truly 
involved in streptococcal infections from those that have a role only under in 
vitro conditions. Much progress has been made in these directions, but there 
are many remaining questions. How long can GAS persist intracellularly in 
humans? Do intracellular GAS multiply or are they dormant? Do L-forms 



of GAS have a role in intracellular persistence? Do intracellular GAS have 
the means to initiate their efflux from the cell without apoptosis? If so, are 
there specific stimuli for the bacterial switch between persistence and efflux? 
What are the genes involved in persistence versus those involved in tissue 
invasion? 

A hint to the answers of some of these questions has come from re- 
search on bacterial regulation, which has identified a number of genes that 
are regulated in response to internalization. Whole genome gene arrays in 
concert with real-time reverse transcription polymerase chain reaction (RT- 
PCR) analyses hold great promise in identifying genes required for inva- 
sion of tissues, metastatic spreading, internalization, persistence, and resis- 
tance to innate defense mechanisms. These approaches are only possible 
as a result of the genome sequencing work of Ferretti et al. (2001), Smoot 
5 et al. (2002), and Beres et al. (2002). Currently being conducted are gene 

w array analyses of the genomes of Ml, M3, M18, and M49, and real-time 

pq 

Q RT-PCR analyses of fas and nra at three different phases of growth (Pod- 

S bielski et al, unpublished work). Undoubtedly, these types of investigative 

£ approaches will generate a wealth of information, a better understanding 




Q 



3 of streptococcal virulence, and novel directions to pursue for preventive 



< 

§ therapies. 

< 

% 

% REFERENCES 

o 
u 



CO 



Abraham, S.N., Beachey, E.H., and Simpson, W.A. (1983). Adherence of Strepto- 

<£ coccus pyogenes, Escherichia coli, and Pseudomonas aeruginosa to fibronectin- 

< 

X coated and uncoated epithelial cells. Infect. Immun. 41, 1261-1268. 

Akesson, P., Sjoholm, A.G., and Bjorck, L. (1996). Protein SIC, a novel extracel- 
lular protein of Streptococcus pyogenes interfering with complement function. 
J. Biol. Chem. 271, 1081-1088. 

Ashbaugh, C, Moser, T., Shearer, M., White, G., Kennedy, R., and Wessels, M. 
(2000). Bacterial determinants of persistent throat colonization and the asso- 
ciated immune response in a primate model of human group A streptococcal 
pharyngeal infection. Cell. Microbiol. 2, 283-292. 

Beachey, E.H. and Ofek, I. (1976). Epithelial cell binding of group A streptococci 
by lipoteichoic acid on fimbriae denuded of M protein. J. Exp. Med. 143, 
759-771. 

Beall, B., Facklam, R., and Thompson, T. (1996). Sequencing emm-specific PCR 
products for routine and accurate typing of group A streptococci. J. Clin. 
Microbiol. 34, 953-958. 

Beckert, S., Kreikmeyer, B., and Podbielski, A. (2001). Group A streptococcal rofA 



gene is involved in the control of several virulence genes and eucaryotic cell 
attachment and internalization. Infect. Immun. 69, 534-537. 

Beres, S., Sylva, G., Barbian, K., Lei, B., Hoff, J., Mammarella, N., Liu, M., Smoot, 
J., Porcella, S., Parkins, L., Campbell, D., Smith, T., McCormick, J., Leung, D., 
Schlievert, P., and Musser, J.M. (2002). Genome sequence of a serotype M3 
strain of group A streptococcus: phage-encoded toxins, the high-virulence 
phenotype, and clone emergence. Proc. Natl Acad. Sci. USA 99, 10,078- 
10,083. 

Berkower, C, Ravins, M., Moses, A., and Hanski, E. (1999). Expression of different 
group A streptococcal M proteins in an isogenic background demonstrates 
diversity in adherence to and invasion of eukaryotic cells. Mol. Microbiol. 31, 
1463-1475. 

Bisno, A. and Stevens, D. (2000). Streptococcus pyogenes (including streptococcal 

shock and necrotizing fasciitis). In Principles and Practice of Infectious Dis- £ 




o 



> 

on 
O 

z 

o 



on 



eases, ed, Mandell, G., Bennet, J., and Dolin, R. pp. 2102-2117. Philadelphia, g 

Churchill Livingstone. ^ 

Botta, G. (1981). Surface components in adhesion of group A streptococci to £ 

pharyngeal epithelial cells. Curr. Microbiol. 6, 101-104. o 

Brandt, C, Allerberger, F., Spellerberg, B., Holland, R., Lutticken, R., and Haase, ° 

G. (2001). Characteristics of consecutive Streptococcus pyogenes isolates from £ 

patients with pharyngitis and bacteriological treatment failure: special refer- % 

ence to prtFl and sic/drs. J. Infect. Dis. 183, 670-674. 

Bretscher, A., Edwards, K., and Fehon, R.G. (2002). ERM proteins and merlin: 

integrators at the cell cortex. Nat. Rev. Mol. Cell. Biol. 3, 586-599. ffi 

Broudy, T., Pancholi, V., and Fischetti, V. (2001). Induction of lysogenic bac- ^ 

teriophage and phage-associated toxin from group A streptococci during w 

co-culture with human pharyngeal cells. Infect. Immun. 69, 1440-1443. 

Broudy, T., Pancholi, V., and Fischetti, V. (2002). The in vitro interaction of Strep- 
tococcus pyogenes with human pharyngeal cells induces a phage-encoded ex- 
tracellular DNase. Infect. Immun. 70, 2805-2811. 

Cheng, Q., Stafslien, D., Purushothaman, S., and Cleary, P. (2002). The group 
B streptococcal C5a peptidase is both a specific protease and invasin. Infect. 
Immun. 70, 2408-2413. 

Cleary, P. and Cue, D. (2000). High frequency invasion of mammalian cells by f3 
hemolytic streptococci. Subcell. Biochem. 33, 137-166. 

Courtney, H.S., Bronze, M.S., Dale, J.B., and Hasty, D.L. (1994). Analysis of the 
role of M24 protein in group A streptococcal adhesion and colonization by 
use of ^-interposon mutagenesis. Infect. Immun. 62, 4868-4873. 

Courtney, H.S., Dale, J.B., and Hasty, D.L. (1996). Differential effects of the 
streptococcal fibronectin-binding protein, FBP54, on adhesion of group A 



streptococci to human buccal cells and HEp-2 tissue culture cells. Infect. 
Immun. 64, 2415-2419. 

Courtney, H.S., Hasty, D.L., and Dale, J.B. (2002). Molecular mechanisms of 
adhesion, colonization, and invasion of group A streptococci. Ann. Med. 34, 
77-87. 

Courtney, H.S., Hasty, D.L., Dale, J.B., and Poirier, T.P. (1992). A 28 kilodalton 
fibronectin-binding protein of group A streptococci. Curr. Microbiol. 25, 245- 
250. 

Courtney, H.S., Liu, S., Dale, J.B., and Hasty, D.L. (1997a). Conversion of M 
serotype 24 of Streptococcus pyogenes to M serotypes 5 and 18: effect on re- 
sistance to phagocytosis and adhesion to host cells. Infect. Immun. 65, 2472- 
2474. 

Courtney, H.S., Ofek, I., Hasty, and D.L. (1997b). M protein mediated adhesion of 
5 M type 24 Streptococcus pyogenes stimulates release of interleukin-6 by HEp-2 




CO 
1-1 



w tissue culture cells. FEMS Microbiol. Lett. 151, 65-70. 



PQ 



Q Courtney, H.S., Ofek, I., Simpson, W.A., Hasty, D.L., and Beachey, E.H. (1986). 

Binding of Streptococcus pyogenes to soluble and insoluble fibronectin. Infect. 



CO 

< 



£ Immun. 53, 454-459. 

5 Cue, D., Dombek, P., Lam, H., and Cleary, P. (1998). Streptococcus pyogenes 



< 

Q 



§ serotype M 1 encodes multiple pathways for entry into human epithelial cells. 



< 



x Infect. Immun. 66, 4593-4601. 

h Cunningham, M. (2000). Pathogenesis of group A streptococcal infections. Clin. 

§ Microbiol. Rev. 13, 470-511. 

u 

w Cywes, C, Stamenkovic, I., and Wessels, M.R. (2000). CD44 as a receptor for 

g colonization of the pharynx by group A streptococci. J. Clin. Invest. 106, 995- 

E 1002. 

Cywes, C. and Wessels, M. (2001). Group A streptococcus tissue invasion by 
CD-44-mediated cell signalling. Nature 414, 648-652. 

Darmstadt, G.L., Fleckman, P., and Rubens, C.E. (1999). Tumor necrosis factor- 
alpha and interleukin-1 alpha decrease the adherence of Streptococcus pyogenes 
to cultured keratinocytes. J. Infect. Dis. 180, 1718-1721. 

Darmstadt, G.L., Mentele, L., Podbielski, A., and Rubens, C.E. (2000). Role of 
group A streptococcal virulence factors in adherence to keratinocytes. Infect. 
Immun. 68, 1215-1221. 

Dinkla, K., Rohde, M., Jansen, W., Carapetis, J., Chhatwal, G., andTalay, S. (2003). 
Streptococcus pyogenes recruits collagen via surface-bound fibronectin: a novel 
colonization and immune evasion mechanism. Mol. Microbiol. 47, 861-869. 

Dinulos, J., Mentele, L., Fredericks, L., Dale, B., and Darmstadt, G. (2003). Ker- 
atinocyte expression of human j3 defensin 2 following bacterial infection: 
role in cutaneous host defense. Clin. Diagnost. Lab. Immunol. 10, 161-166. 



Dombek, P., Cue, D., Sedgewick, J., Ruschkowski, S., Finlay, B., and Cleary, P. 
(1999). High frequency intracellular invasion of epithelial cells by serotype 
Ml group A streptococci: Ml protein-mediated invasion and cytoskeletal 
rearrangements. Mol. Microbiol. 31, 859-70. 

Dorschner, R., Pestonjamasp, V., Tamakuwala, S., Ohtake, T., Rudisill, J., Nizet, 
V., Agerberth, B., Gudmundsson, G., and Gallo, R. (2001). Cutaneous in- 
jury induces the release of cathelicidin anti-microbial peptides active against 
group A streptococcus. J. Invest. Dermatol. 117, 91-97. 

Duensing, T., Wing, J., and van Putten, J. (1999). Sulfated polysaccharide -directed 
recruitment of mammalian host proteins: a novel strategy in microbial patho- 
genesis. Infect. Immun. 67, 4463-4468. 

Dunne, D.W., Resnick, D., Greenberg, D.J., Kreiger, J.M., and Joiner, K.A. (1994). 
The type 1 macrophage scavenger receptor binds to Gram-positive bacteria 
and recognizes lipoteichoic acid. Proc. Natl Acad. Sci. USA 91, 1863-1867. £ 



bacteria. Anna. Rev. Microbiol. 51, 527-564. 




o 



Dunny, G. and Leonard, B. (1997). Cell-cell communication in gram-positive g 



> 

in 
H 



Eisner, A., Kreikmeyer, B., Braukn-Kiewnick, A., Spellerberg, B., Buttaro, B., and £ 






Podbielski, A. (2002). Involvement of Lsp, a member of the Lral -lipoprotein o 



n 

o 



family in Streptococcus pyogenes, in eukaryotic cell adhesion and internaliza- ° 

tion. Infect. Immun. 70, 4859-4869. fc 

I— i 

Ferretti, J., McShan, W., Ajdic, D., Savic, D., Savic, G., Lyon, K., Primeaux, C, % 
Sezate, S. Suvorov, A., Kenton, S., Lai, H.S., Lin, S.P., Qian, Y., Jia, H.G., 
Najar, F.Z., Ren, Q., Zhu, N., Song, L., White, J., Yuan, X., Clifton, S.W., 

Roe, B.A., and McLaughlin, R. (2001). Complete genome sequence of an Ml ffi 



> 

o 
Z 

o 



strain of Streptococcus pyogenes. Proc. Natl Acad. Sci. USA 98, 4658-4663. ^ 

Fischetti, V.A. (1989). Streptococcal M protein: molecular design and biological w 

behaviour. Clin. Microbiol. 2, 285-314. °° 

Fluckiger, U., Jones, K., and Fischetti, V. (1998). Immunoglobulins to group 
A streptococcal surface molecules decrease adherence to and invasion of 
human pharyngeal cells. Infect. Immun. 66, 974-979. 

Fogg, G.C. and Caparon, M.G. (1997). Constitutive expression of fibronectin bind- 
ing in Streptococcus pyogenes as a result of anaerobic activation of rofA. J. 
Bacteriol. 179, 6172-6180. 

Frick, I., Morgelin, M., and Bjorck, L. (2000). Virulent aggregates of Streptococ- 
cus pyogenes are generated by homophillic protein-protein interactions. Mol. 
Microbiol. 37, 1232-1247. 

Gerlach, D., Schalen, C, Tigyi, Z., Nilsson, B., Forsgren, A., and Naidu, A.S. 
(1994). Identification of a novel lectin in Streptococcus pyogenes and its possi- 
ble role in bacterial adherence to pharyngeal cells. Curr. Microbiol. 28, 331— 
338. 



Granok, A., Parsonage, D., Ross, R., and Caparon, M. (2000). The RofA binding 
site in Streptococcus pyogenes is utilized in multiple transcriptional pathways. 
J. Bacteriol. 182, 1529-1540. 

Gryllos, I., Cywes, C, Shearer, M., Cary, M., Kennedy, R., and Wessels, M. (2001). 
Regulation of capsule gene expression by group A streptococcus during pha- 
ryngeal colonization and invasive infection. Mol. Microbiol. 42, 61-74. 

Hanski, E. and Caparon, M. (1992). Protein F, a fibronectin-binding protein, is 
an adhesin of the group A streptococcus, Streptococcus pyogenes. Proc. Natl 
Acad. Sci. USA 89, 6172-6176. 

Hartas, J. and Sriprakash, K. (1999). Streptococcus pyogenes strains containing 
emml2 and emm55 possess a novel gene coding for distantly related SIC 
protein. Microb. Pathogen. 26, 25-33. 

Hasty, D.L., Ofek, I., Courtney, H.S., and Doyle, R. (1992). Multiple adhesins of 
5 streptococci. Infect. Immun. 60, 2147-2152. 

w Hidalgo-Grass, C, Ravins, M., Dan-Goor, M., Jaffe, J., Moses, A., and Hanski, E. 




PQ 



Q (2002). A locus of group A streptococci involved in invasive disease and DNA 

S transfer. Mol. Microbiol. 46, 87-99. 

< 

£ Hoe, N., Ireland, R., DeLeo, F., Gowen, B., Dorward, D., Voyich, J., Liu, M., 

^ Burns, E. Jr., Culnan, D., Bretscher, A., and Musser, J.M. (2002). Insight into 

§ the molecular basis of pathogen abundance: group A streptococcus inhibitor 

< 

5h of complement inhibits bacterial adherence and internalization into human 

h cells. Proc. Natl Acad. Sci. USA 99, 7646-7651. 

pi 

o Hynes, R.O. (2002). Integrins: bidirectional, allosteric signaling machines. Cell 

^ 110, 673-687. 

g Hytonen, J., Haataja, S., and Finne, J. (2003). Streptococcus pyogenes glycoprotein- 

< 

ffi binding strepadhesin activity is mediated by a surface-associated 

carbohydrate -degrading enzyme, pullanase. Infect. Immun. 71, 784-793. 
Hytonen, J., Haataja, S., Gerlach, D., Podbielski, A., and Finne, J. (2001). The speB 

virulence factor of Streptococcus pyogenes, a multifunctional secreted and cell 

surface molecule with strepadhesin, laminin-binding and cysteine protease 

activity. Mol. Microbiol. 39, 512-519. 
Jadoun, J., Burstein, E., Hanski, E., and Sela, S. (1997). Proteins M6 and Fl are 

required for efficient invasion of group A streptococci into cultured epithelial 

cells. Adv. Exp. Med. 418, 511-515. 
Jadoun, J., Eyal, O., and Sela, S. (2002). Role of CsrR, hyaluronic acid, and SpeB 

in the internalization of Streptococcus pyogenes M type 3 strain by epithelial 

cells. Infect. Immun. 70, 462-469. 
Jaffe, J., Natanson-Yaron, S., Caparon, M., and Hanski, E. (1996). Protein F2, a 

novel fibronectin-binding protein from Streptococcus pyogenes, possesses two 

binding domains. Mol. Microbiol. 21, 373-384. 



Jeng, A., Sakota, V., Li, Z., Datta, V., Beall, B., and Nizet, V. (2003). Molecular 
genetic analysis of a group A streptococcus operon encoding serum opacity 
factor and a novel fibronectin-binding protein, SfbX. J. Bacteriol. 185, 1208- 
1217. 

Ji, Y., Schnitzler, N., DeMaster, E., and Cleary, P. (1998). Impact of M49, Mrp, 
Enn, and C5a peptidase proteins of colonization of the mouse oral mucosa 
by Streptococcus pyogenes. Infect. Immun. 66, 5399-5405. 

Johnson, D. and Kaplan, E. (1993). A review of the correlation of T-agglutination 
patterns and M -protein typing and opacity factor production in the identifi- 
cation of group A streptococci. J. Med. Microbiol. 38, 311-315. 

Kallman, J. and Kihlstrolm, E. (1997). Penetration of group B streptococci through 
polarized Madin-Darby canine kidney cells. Fed. Res. 42, 799-804. 

Kreikemeyer, B., Boyle, M., Buttaro, B., Heinemann, M., and Podbielski, A. 

(2001). Group A streptococcal growth-phase associated virulence factor reg- £ 

o 
ulation by a novel operon (Fas) with homologies to two-component -type reg- g 

ulators requires a small RNA molecule. Mol. Microbiol. 39, 392-406. ^ 

Kreikemeyer, B., Mclver, K., and Podbielski, A. (2003). Virulence factor regula- £ 

tion and regulatory networks in Streptococcus pyogenes and their impact on o 

pathogen-host interactions. Trends Microbiol. 11, 224-232. n 

Kreikemeyer, B., Talay, S.R., and Chhatwal, G.S. (1995). Characterization of a £ 

novel fibronectin-binding surface protein in group A streptococci. Mol. Mi- % 

crobiol. 17, 137-145. 
Lancefield, R.C. (1933). A serological differentiation in human and other groups 

of hemolytic streptococci. J. Exp. Med. 57, 571-595. ffi 




> 

on 
O 

z 

o 



LaPenta, D., Rubens, C, Chi, E., and Cleary, P. (1994). Group A streptococci ^ 

efficiently invade human respiratory epithelial cells. Proc. Natl Acad. Sci. w 

USA 91, 12,115-12,119. " 

Lei, B., Deleo, F., Reid, S., Voyich, J., Magoun, L., Liu, M., Braughton, 
K., Ricklefs, S., Hoe, N., Cole, R.L., Leong, J.M., Musser, J.M. (2002). 
Opsonophagocytosis-inhibiting Mac protein of group A streptococcus: iden- 
tification and characteristics of two genetic complexes. Infect. Immun. 70, 
6880-6890. 

Lin, C, Kuan, I., Wang, C, Lee, H., Lee, W., Sheu, J., Hsiao, G., Wu, C, and 
Kuo, H. (2002). Lipoteichoic acid-induced cyclooxygenase-2 expression re- 
quires activation of p44/42 and p38 mitogen-activated protein kinase signal 
pathways. Eur. J. Pharmacol. 450, 1-9. 

Lukomski, S., Nakashima, K., Abdi, I., Cipriano, V., Shelvin, B., Graviss, E., and 
Musser, J. (2001). Identification of characterization of a second extracellular 
collagen-like protein made by group A streptococcus: control of production 
at the level of transcription. Infect. Immun. 69, 1729-1738. 



Madden, J., Ruiz, N., and Caparon, M. (2001). Cytolysin-mediated translocation 
(CMT): a functional equivalent of type III secretion in gram-positive bacteria. 
Cell 104, 143-152. 

Mclver, K.S., Heath, A.S., and Scott, J.R. (1995). Regulation of virulence by envi- 
ronmental signals in group A streptococci: influence of osmolarity, tempera- 
ture, gas exchange, and iron limitation on emm transcription. Infect. Immun. 
63, 4540-4542. 

Mclver, K.S. and Scott, J.R. (1997). Role of mga in growth phase regulation of 
virulence genes of the group A streptococcus. J. Bacteriol. 179, 5178-5187. 

Medina, E., Anders, D., and Chhatwal, G.S. (2002). Induction of NF-kappaB nu- 
clear translocation in human respiratory epithelial cells by group A strepto- 
cocci. Microbiol. Pathog. 33, 307-313. 

Miettinen, M., Matikainen, S., Vuopio-Varjkila, J., Pirhonen, J., Varkila, K., Ku- 
5 rimoto, M., and Julkunen, I. (1998). Lactobacilli and streptococci induce 

w interleukin-12 (IL-12), IL-18, and gamma interferon production in human 




PQ 



Q periphereal blood mononuclear cells. Infect. Immun. 66, 6058-6062. 

Molinari, G., Rohde, M., Guzman, C., and Chhatwal, G. (2000). Two distinct path- 



< 



£ ways for the invasion of streptococci in non-phagocytic cells. Cell. Microbiol. 

5 2, 145-154. 



Q 



§ Molinari, G., Rohde, M., Talay, S.R., Chhatwal, G.S., Beckert, S., and Podbielski, 



< 

^ A. (2001). The role played by the group A streptococcal negative regulator 

h Nra on bacterial interactions with epithelial cells. Mol. Microbiol. 40, 99-114. 

o Molinari, G., Talay, S.R., Valentin-Weigand, P., Rohde, M., and Chhatwal, G. 

u 

^ (1997). The fibronectin-binding protein of Streptococcus pyogenes, Sfbl, is in- 

g volved in internalization of group A streptococci by epithelial cells. Infect. 

< 

ffi Immun. 65, 1357-1363. 

Nakagawa, I., Nakata, M., Kawabata, S., and Hamada, S. (2001). Cytochrome 
C-mediated caspase-9 activation triggers apoptosis in Streptococcus pyogenes 
infected epithelial cells. Cell. Microbiol. 3, 359-405. 

Natanson, S., Sela, S., Moses, A., Musser, J., Caparon, M., and Hanski, E. (1995). 
Distribution of fibronectin-binding proteins among group a streptococci of 
different M types. J. Infect. Dis. 171, 871-878. 

Neeman, R., Keller, N., Barzilai, A., Korenman, Z., and Sela, S. (1998). Prevalence 
of internalization-associated gene, prtFl, among persisting group-A strepto- 
coccus strains isolated from asymptomatic carriers. Lancet 352, 1974-1977. 

Nizet, V., Ohtake, T., Lauth, X., Trowbridge, J., Rudisill, J., Dorschner R., Peston- 
jamasp, V., Piraino, J., Huttner, K., and Gallo, R.L. (2001). Innate antimicro- 
bial peptide protects the skin from invasive bacterial infections. Nature 414, 
454-457. 



Ofek, I., Zafiri, I., Goldhar, J., and Eisenstein, B. (1990). Inability of toxin in- 
hibitors to neutralize enhanced toxicity caused by bacteria adherent to tissue 
culture cells. Infect. Immun. 58, 3737-3742. 

Okada, N., Liszewski, M., Atkinson, J., and Caparon, M. (1995). Membrane cofac- 
tor protein (CD46) is a keratinocyte receptor for the M protein of the group 
A streptococcus. Proc. Natl Acad. Sci. USA 92, 2489-2493. 

Okada, N., Pentland, A., Falk, P., and Caparon, M. (1994). M protein and pro- 
tein F act as important determinants of cell-specific tropism of Streptococcus 
pyogenes in skin tissue. J. Clin. Invest. 94, 965-977. 

Okada, N., Tatsuno, I., Hanski, E., Caparon, M., and Sasakawa, C. (1998). Strepto- 
coccus pyogenes protein F promotes invasion of HeLa cells. Microbiology 144, 
3079-3086. 

Okahashi, N., Sakurai, A., Nakagawa, I., Fujiwara, T., Kawabata, S., Amano, A., 

and Hamada, S. (2003). Infection by Streptococcus pyogenes induces the recep- £ 



Immun. 71, 948-955. 




o 



tor activator of NF-kB ligand expression in mouse osteoblastic cells. Infect. g 



> 

in 
H 



Osterlund, A. and Engstrand, L. (1997). An intracellular sanctuary for Streptococcus £ 






pyogenes in human tonsillar epithelium - studies of asymptomatic carriers o 



> 

on 
O 

z 

o 



on 



n 
and in vitro cultured biopsies. Acta Otolaryngol. 117, 883-888. n 

Osterlund, A., Popa, R., Nikkila, T., Scheynius, A., and Engstrand, L. (1997). £ 

Intracellular reservoir of Streptococcus pyogenes in vivo: a possible explanation % 

for recurrent pharyngotonsillitis. Laryngoscope 107, 640-646. 
Ozeri, V., Rosenshine, I., Mosher, D., Fassler, R., and Hanski, E. (1998). Roles of 

integrins and fibronectin in the entry of Streptococcus pyogenes into cells via ffi 

protein Fl. Mol. Microbiol. 30, 625-637. g 

Pancholi, V. and Fischetti, V. (1992). A major surface protein on group A strep- w 

tococci glyceraldehyde- 3 -phosphate -dehydrogenase with multiple binding 

activity. J. Exp. Med. 176, 415-426. 
Pancholi, V. and Fischetti, V. (1997). Regulation of the phosphorylation of human 

pharyngeal cell proteins by group A streptococcal surface dehydrogenase: 

signal transduction between streptococci and pharyngeal cells. J. Exp. Med. 

186, 1633-1643. 
Perez-Casal, J., Okada, N., Caparon, M., and Scott, J. (1995). Role of the conserved 

C-repeat region of the M protein of Streptococcus pyogenes. Mol. Microbiol. 15, 

907-916. 
Podbielski, A., Beckert, S., Schattke, R., Leithauser, F., Lestin, F., Gobler, B., 

and Kreikemeyer, B. (2003). Epidemiology and virulence gene expression of 

intracellular group A streptococci in tonsils of recurrently infected adults. 

Int. J. Med. Microbiol. 293, 179-190. 



Podbielski, A., Woischnik, M., Leonard, B., and Schmidt, K. (1999). Characteri- 
zation of nra, a global negative regulator gene in group A streptococci. Mol. 
Microbiol. 31, 1051-1064. 

Podbielski, A, Woischnik, M., Pohl, B., and Schmidt, K.H. (1996). What is the size 
of the group A streptococcal vir regulon? The Mga regulator affects expression 
of secreted and surface virulence factors. Med. Microbiol. Immunol. 185, 171- 
181. 

Pritchard, K. and Cleary, P. (1996). Differential expression of genes in the vir 
regulon of Streptococcus pyogenes is controlled by transcription termination. 
Mol. Gen. Genet. 250, 207-213. 

Rakonjac, J.V., Robbins, J.C., and Fischetti, V. (1995). DNA sequence of the serum 
opacity factor of group A streptococci: identification of a fibronectin-binding 
repeat domain. Infect. Immun. 63, 622-631. 
5 Rasmussen, M.andBjorck, L. (2001). Unique regulation ofSclB- a novel collagen- 

w like surface protein of Streptococcus pyogenes. Mol. Microbiol. 40, 1427- 

S 1438. 

o 

w Rasmussen, M., Eden, A., and Bjorck, L. (2000). SclA, a novel collagen-like surface 

£ protein of Streptococcus pyogenes. Infect. Immun. 68. 6370-6377. 

^ Rocha, C. and Fischetti, V. (1999). Identification and characterization of a novel 

§ fibronectin-binding protein on the surface of group A streptococci. Infect. 

g Immun. 67, 2720-2728. 

h Rodhe, M., Muller, E., Chhatwal, G., and Talay, S. (2003). Host cell caveolae act 

o as an entry port for group A streptococci. Cell. Microbiol. 5, 323-342. 

u 




CO 

>1 



Ruiz, N., Wang, B., Pentland, A., and Caparon, M. (1998). Streptolysin O 

gj and adherence synergistically modulate proinflammatory responses of ker- 

< 

X atinocytes to group A streptococci. Mol. Microbiol. 27, 337-346. 

Ryan, P., Pancholi, V., and Fischetti, V. (2001). Group A streptococci bind to 

mucin and human pharyngeal cells through sialic acid-containing receptors. 

Infect. Immun. 69, 7402-7412. 
Saetre, T., Hoiby, E., Kahler, H., and Lyberg, T. (2000). Changed expression of 

leukocyte adhesion molecules and increased production of reactive oxygen 

species caused by Streptococcus pyogenes in human whole blood. APMIS 108, 

573-580. 
Schmidt, K.H., Mann, K., Cooney, J., and Kohler, W. (1993). Multiple binding 

of type 3 streptococcal M protein to fibrinogen, albumin, and fibronectin. 

FEMS Immun. Med. Microbiol. 7, 135-144. 
Schrager, H.M., Alberti, S., Cywes, S., Dougherty, G.J., and Wessels, M.R. (1998). 

Hyaluronate acid capsule modulates M protein mediated adherence and acts 

as a ligand for attachment of group A streptococci to CD44 on human ker- 

atinocytes. /. Clin. Invest. 101, 1708-1716. 




Sela, S., Aviv, A., Tovi, A., Burstein, I., Caparon, M., and Hanski, E. (1993). Protein 
F: an adhesin of Streptococcus pyogenes binds fibronectin via two distinct 
domains. Mol. Microbiol 10, 1049-1055. 

Sierig, G., Cywes, C, Wessels, M., and Ashbaugh, C. (2003). Cytotoxic effects of 
streptolysin O and streptolysin S enhance the virulence of poorly encapsu- 
lated group A streptococci. Infect. Immun. 71, 446-455. 

Simpson, W.A. and Beachey, E.H. (1983). Adherence of group A streptococci to 
fibronectin on oral epithelial cells. Infect. Immun. 39, 275-279. 

Smoot, J., Barbian, K., Van Gompel, J., Smoot, L., Chaussee, M., Sylva, G., Sturde- 
vant D., Ricklefs, S., Porcella, S.F., Parkins, CD., Beres, S.B., Campbell, D.S., 
Smith, T.M., Zhang, Q., Kapur, V., Daly, J. A., Veasy, L.G., and Musser, J.M. 
(2002). Genome sequences and comparative microarray analysis of serotype 
M18 group A streptococcus strains associated with acute rheumatic fever 

outbreaks. Proc. Natl Acad. Sci. USA 99, 4668-4673. g 

o 

Speck, O., Hughes, S., Noren, N., Kulikauskas, R., and Fehon, R. (2003). Moesin g 

functions antagonistically to the Rho pathway to maintain epithelial integrity. £ 

Nature All, 83-87. B 

Stalhammar-Carlemalm, M., Areschoug, T., Larsson, C, and Lindahl, G. (1999). o 

The R28 protein of Streptococcus pyogenes is related to several group B strepto- ° 

coccal surface proteins, confers protective immunity and promotes binding £ 

to human epithelial cells. Mol. Microbiol. 33, 208-219. % 

Stenfors, L., Bye, H., Raisanen, S., and Myklebust, R. (2000). Bacterial penetra- 
tion into tonsillar surface epithelium during infectious mononucleosis. J. 
Laryngol. Otol. 114, 848-852. £ 

Stevens, D., Salmi, D., Mclndoo, E., and Bryant, A. (2000). Molecular epidemiol- ^ 

ogy of nga and NAD glycohydralase/ADP-ribosyltransferase activity among w 

Streptococcus pyogenes causing toxic shock syndrome. J. Infect. Dis. 182, 1117- 
1128. 

Stockbauer, K.E., Magoun, L., Liu, M., Burns, E.H. Jr., Gubbam S., Renish, S., 
Pan, X., Bodary, S.C., Baker, E., Coburn, J., Leong, J.M., and Musser, J.M. 
(1999). A natural variant of the cysteine protease virulence factor of group A 
streptococcus with an arginine -glycine -aspartic acid (RGD) motif preferen- 
tially binds human integrins avf3?> and cdIb/33. Proc. Natl Acad. Sci. USA 96, 
242-247. 

Talay, S.R., Valentin-Weigand, P., Jerlstrom, P.G., Timmis, K.N., and Chhatwal, 
G.S. (1993). Fibronectin-binding protein of Streptococcus pyogenes: sequence 
of the binding domain involved in adherence of streptococci to epithelial 
cells. Infect. Immun. 60, 3837-3844. 

Talay, S.R., Zock, A., Rohde, M., Molinari, G., Oggioni, M., Pozzi, G., Guzman, 
C.A., and Chhatwal, G.S. (2000). Co-operative binding of human fibronectin 



> 

on 
O 

z 

o 



on 



to Sfbl protein triggers streptococcal invasion into respiratory epithelial cells. 
Cell. Microbiol. 2, 521-535. 

Talay, S., Ehrenfeld, E., Chhatwal, G., and Timmis, K. (1991). Expression of the 
fibronectin-binding components of Streptococcus pyogenes in Escherichia coli 
demonstrates that they are proteins. Mol. Microbiol. 5, 1727-1734. 

Tamura, G. and Nittayajarn, A. (2000). Group B streptococci and other gram- 
positive cocci bind to cytokeratin 8. Infect. Immun. 68, 2129-2134. 

Terao, Y., Kawabata, S., Kunitomo, E., Murakami, J., Nakagawa, I., and Hamada, 
S. (2001). Fba, a novel fibronectin-binding protein from Streptococcus pyo- 
genes, promotes bacterial entry into epithelial cells, and the^iba gene is posi- 
tively transcribed under the Mga regulator. Mol. Microbiol. 42, 75-86. 

Terao, Y., Kawabata, S., Nakata, ML, Nakagawa, I., and Hamada, S. (2002). Molec- 
ular characterization of a novel fibronectin-binding protein of Streptococcus 
5 pyogenes strains isolated from toxic shock-like syndrome patients. J. Biol. 




CO 
1-1 



n Chem. 277, 47,428-47,435. 



PQ 



Q Terao, Y., Kawabata, S., Kunitomo, E., Nakagawa, I., and Hamada, S. (2002). 

Novel laminin-binding protein of Streptococcus pyogenes, Lbp, is involved in 



CO 

< 

£ adhesion to epithelial cells. Infect. Immun. 70, 993-997. 

^ Tomasini-Johansson, B., Kaufman, N., Ensenberger, M., Ozeri, V., Hanski, E., 

§ and Mosher, D. (2001). A 49-residue peptide from adhesin Fl of Streptococcus 

< 

5h pyogenes inhibits fibronectin-matrix assembly. J. Biol. Chem. 276, 23,430- 

h 23,439. 

pi 

o Tsai, P., Lin, Y., Kuo C., Lei, H., and Wu, J. (1999). Group A streptococcus induces 

u 



CO 

>1 



apoptosis in human epithelial cells. Infect. Immun. 67, 4334-4339. 

gj Upton, M., Tagg, J.R., Wescombe, P., and Jenkinson, H.F. (2001). Intra- and inter- 

< 

X species signalling between Streptococcus salivarius and Streptococcus pyogenes 

mediated by SalA and SalAl lantibiotic peptides. J. Bacteriol. 183, 3931-3938. 
Valentin-Wiegand, P., Grulich-Henn, J., Chhatwal, G., Muller-Berhasus, G., 

Blobel, H., and Preissner, K. (1988). Mediation of adherence of streptococci 

to human endothelial cells by complement S protein (vitronectin). Infect. 

Immun. 56, 2851-2855. 
Visai, L., Bozzini, S., Raucci, G., Toniolo, A., and Speziale, P. (1995). Isolation 

and characterization of a novel collagen-binding protein from Streptococcus 

pyogenes strain 6414. J. Biol. Chem. 270, 347-353. 
von Hunolstein, C, Greco, R., Ajello, M., Orefici, G., and Valenti, P. (2000). Strep- 
tococcus pyogenes internalization by Hela cells is not mediated by M6 protein. 

In Streptococci and Streptococcal Diseases - Entering the New Millennium, ed. 

D. Martin and J. Tagg, pp. 681-683, Wellington, New Zealand: Securacopy. 
Von Pawel-Rammingen, U., Johansson, B., and Bjorck, L. (2002). IdeS, a novel 

streptococcal cysteine proteinase with unique specificity for immunoglobulin 

G. EMBOJ. 21, 1607-1615. 



Voyich, J., Sturdevant, D., Braughton, K., Kobayashi, S., Lei, B., Virtaneva, K., 
Dorward, D., Musser, J., and Deleo, F. (2003). Genome-wide protective re- 
sponse used by group A streptococcus to evade destruction by human poly- 
morphonuclear leukocytes. Proc. Natl Acad. Sci. USA 100, 1996-2001. 

Wadstrom, T. and Tylewska, S. (1982). Glycoconjugates as possible receptors for 
Streptococcus pyogenes. Curr. Microbiol. 7, 343-346. 

Wang, B., Ruiz, N., Pentland, A., and Caparon, M. (1997). Keratinocyte proin- 
flammatory responses to adherent and nonadherent group A streptococci. 
Infect. Immun. 65, 2119-2126. 

Wang, J. and Stinson, M. (1994). Streptococcal M6 protein binds to fucose- 
containing glycoproteins on cultured human epithelial cells. Infect. Immun. 
62, 1268-1274. 

Wessels, M.R. and Bronze, M.S. (1994). Critical role of the group A streptococcal 

capsule in pharyngeal colonization and infection in mice. Proc. Natl Acad. £ 

Sci. USA91, 12,238-12,242. § 

Whatmore, A.M. (2001). Streptococcus pyogenes sclB encodes a putative hypervari- ^ 

able surface protein with a collagen-like repetitive structure. Microbiology 1 47, £ 

419-429. 




H 
O 

n 

o 
n 
n 

> 

M 

< 
> 

I— I 
O 

Z 

o 

X 
o 

on 
H 

n 

M 
M 
M 

on 



CHAPTER 9 



Invasion of oral epithelial cells by 
Actinobacillus actinomycetemcomitans 

Diane Hutchins Meyer, Joan E. Lippmann, and Paula Fives-Taylor 



ACTINOBACILLUS ACTINOMYCETEMCOMITANS: ADHERENCE 
MECHANISMS REQUIRED FOR INVASION 

Colony phase variation 

A. actinomycetemcomitans produces three distinct colonial morphologies 
on solid medium. A rough colony phenotype is typically generated by organ- 
isms upon isolation from the gingiva. These are small (~0.5-l mm in diam- 
eter), translucent circular colonies with rough surfaces and irregular edges 
(Fig. 9.1). An internal star-shaped or crossed cigar morphology that embeds 
the agar is a distinguishing characteristic that gives A. actinomycetemcomitans 
its name (Zambon, 1985). In liquid culture, the rough colony phenotype cells 
form aggregates on the vessel walls, resulting in a clear medium (Fig. 9.1). 
Repeated subculture on agar of rough phenotypic isolates yields two distinct 
colonial variants; one is smooth surfaced and transparent, and the other is 
smooth surfaced and opaque (Slots, 1982; Scannapieco et al., 1987; Rosan 
et al., 1988; Inouye et al., 1990). The transparent smooth- surfaced variants 
appear to be an intermediate between the transparent rough-surfaced and 
opaque smooth-surfaced types (Inouye et al., 1990). In broth, the smooth- 
surfaced opaque type grows as a turbid homogeneous suspension, whereas 
the smooth-surfaced transparent type aggregates and adheres to the vessel 
walls (Inouye et al., 1990). In general, isolates undergo a rough-to-smooth 
variant transition soon after culture in vitro. In contrast, a smooth-to-rough 
variant transition that appears to be associated with nutritional requirements 
occurs only rarely during in vitro culture (Inouye et al., 1990; Meyer et al., 
1991; Meyer, unpublished observation). 





o 

1-1 

% 

w 

> 

I— I 

Ph 



< 

Q 

< 

< 
p^ 



O 



Pi 
w 

!* 

w 

CO 

i— i 

X 
u 

H 

X 
w 
Z 
< 





Figure 9.1. Micrographs of A. actinomycetemcomitans phenotypes. Rough (A and C) and 
smooth (B and D) colonial variants cultured in agar (top panel) and broth (lower panel). A 
unique characteristic of the rough colony morphology is the lovely star form that is 
embedded in the agar (A). In broth culture, the rough phenotype forms tight aggregates 
(C). 

Surface-associated molecules and organelles 

Bacterial colonial variation is indicative of the differential expression 
of cell surface components (Braun, 1965). A. actinomycetemcomitans rough 
colony variants express 43- and 20-kDa outer membrane proteins, rough 
colony protein A and B, respectively, that are not expressed in smooth colony 
variants (Haase et al., 1999). These proteins are encoded by genes that have 
homology to genes known to encode fimbriae-associated proteins. In that 
regard, A. actinomycetemcomitans rough colony variants are heavily fimbri- 
ated, whereas smooth colony variants have few or no fimbriae (Scannapieco 
et al., 1987; Rosan et al., 1988; Inouye et al., 1990; Meyer and Fives-Taylor, 
1994) . In accordance, rough colony variants adhere better than smooth colony 
variants in vitro to epithelial cells (Meyer and Fives-Taylor, 1994). In contrast, 
smooth colony variants invade epithelial cells in vitro significantly better than 
do rough variant colonies (Meyer et al., 1991). A. actinomycetemcomitans strain 
SUNY 465, the invasion prototype, exhibits a smooth -to -rough variant shift 
after anaerobic growth on agar that is accompanied by cell fimbriation and in- 
creased adherence to KB oral epithelial cells (Meyer and Fives-Taylor, 1994). 
Although the role of the phenotypic variation is not known, it has been pro- 
posed that it may play some role in the episodic nature of periodontitis (Meyer 
etal., 1991). 

A. actinomycetemcomitans fimbriae are peritrichous, ~2 /xm in length 
and 5 nm in diameter, and frequently occur in bundles (Holt et al., 1980; 





Figure 9.2. Scanning electron micrographs of vesicles (blebs) associated with the A. 
actinomycetemcomitans surface. Vesicles may vary in number and occur in different forms 
or shapes and sizes. The vesicles associated with the organism in panel A (35,000x) are 
relatively short and peritrichous (arrows), whereas those in panel B (70,000x) are 
extremely long and essentially wrap around the organism (arrows). 

Scannapieco et al., 1983, 1987; Preus et al., 1988; Rosan et al, 1988). They 
are composed of a 54-kDa fimbrial subunit (Inouye et al., 1990). Adhesion 
of A. actinomycetemcomitans-fimbrisLted strains to both buccal epithelial cells 
and the Gin-1 fibroblast cell line is inhibited by antiserum to the 54-kDa pro- 
tein, indicating a role for it in adhesion. Fimbrial-associated protein (fap), an 
A. actinomycetemcomitans attachment factor, is expressed in fimbriated, but 
not in nonfimbriated, strains (Ishihara et al., 1997). Flp, a 6.5-kDa protein 
which is a component of A. actinomycetemcomitans fimbriae, exhibits some 
amino acid sequence similarity to type IV pilin (Inoue et al., 1998). Clearly, 
there is a correlation between A. actinomycetemcomitans fimbriation and ad- 
hesion. However, A. actinomycetemcomitans cells devoid of fimbriae exhibit 
adhesiveness as well, indicating that nonfimbrial components also function 
in adhesion (Inouye et al., 1990; Meyer and Fives-Taylor, 1994). 

Membraneous vesicles (blebs) are a prominent feature of the surface of 
A. actinomycetemcomitans. These structures, which are thought to be predom- 
inantly lipopoly saccharide in nature, are extensions of the outer membrane 
that either remain attached to or bud off from the cell surface (Fig. 9.2). 
Large numbers of vesicles are released into the external environment dur- 
ing culture (Holt et al., 1980). The formation and morphology of vesicles are 
altered by growth conditions (Meyer and Fives-Taylor, 1993). Cells grown 
on agar have vesicles characterized by thick fibrils with knob-like ends. A. 




< 
> 

on 
O 

z 

o 



M 

M 

i— i 

H 

X 

M 

M 
I— I 

> 

n 

M 
M 
M 

oo 

CO 
> 

o 

> 

n 

H 
H 
H 

> 
O 

S 
n 

hi 

s 

n 

o 

S 



actinomycetemcomitans vesicles contain endotoxin, bone resorption activity, 
actinobacillin (a bacteriocin) , and leukotoxin (Hammond et al., 1981, 1987; 
Nowotny etal., 1982; Stevens etal., 1987; Lucia etal., 2002). Highly leukotoxic 
A. actinomycetemcomitans strains have an abundance of vesicles, whereas low- 
leukotoxic or nonleukotoxic strains have few or no vesicles (Lai et al., 1981). 
The leukotoxicity of vesicles associated with highly leukotoxic strains is 5 
to 10 times greater than that of minimally leukotoxic strains (Kato et al., 
2001). 

A. actinomycetemcomitans vesicles also demonstrate adhesiveness. The 
addition of vesicles to weakly adherent or nonadherent strains significantly 
increases the ability of those strains to attach to epithelial cells (Meyer and 
Fives-Taylor, 1993). The adhesive nature of the vesicles prompted the hy- 
pothesis that these organelles function in A. actinomycetemcomitans as deliv- 
g ery vehicles for the toxic materials that they harbor (Meyer and Fives-Taylor, 

5 1993). Scanning electron microscopy of the process of invasion of A. actino- 




H 



£ mycetemcomitans into epithelial cells revealed that bacteria in contact with, 

E and in the process of being internalized by, the cells have surface-associated 

g vesicles. In contrast, bacteria not in contact with epithelial cells do not pos- 

* sess vesicles, suggesting that the internalization process may be a trigger 

5 for vesicle formation (Meyer et al., 1996). The role of vesicles in the A. acti- 

| nomycetemcomitans internalization process is not known, but clearly further 

2 investigation is warranted. 

S The surface of A. actinomycetemcomitans may also be associated with an 

w extracellular amorphous material (ExAmMat) that can embed groups of cells 

< in a matrix (Holt et al., 1980) . Whereas an early study reported that cells grown 

rt in liquid medium lacked amorphous material, other reports indicate that Ex- 

n AmMat can occur on cells grown in liquid culture (Wilson et al., 1985; Meyer 

co and Fives-Taylor, 1994). Expression of ExAmMat has some association with 

£ growth in tryptone-based medium (Wilson et al., 1985). Therefore, similar 

g to fimbriae and vesicles, culture conditions can modify expression of ExAm- 

w Mat. It has been determined that ExAmMat is proteinacious, most likely a 

$ glycoprotein, and has adhesive properties (Meyer and Fives-Taylor, 1993). 

ExAmMat is not affixed firmly to the cell surface, as the washing of cells 

with phosphate-buffered saline removes ExAmMat and results in reduced 

adhesion to epithelial cells (Meyer and Fives-Taylor, 1993). Moreover, weakly 

adherent A. actinomycetemcomitans strains adhere strongly to epithelial cells 

after suspension in ExAmMat, a process termed conveyed adhesion. The 

surfaces of A. actinomycetemcomitans suspended in ExAmMat are associated 

with large amounts of material, indicating that the conveyed adhesion is the 

result of a direct transfer of ExAmMat onto the bacterial surface (Fives-Taylor 

etal., 1995). 



Specific adhesion to epithelial cells 

Most A. actinomycetemcomitans strains that have been tested to date ad- 
here to epithelial cells strongly; however, the adhesiveness of strains does 
vary (Meyer and Fives-Taylor, 1994). The rapid process reaches saturation 
levels within lh of infection (Mintz and Fives-Taylor, 1994). Adhesion is 
affected by growth conditions (Meyer and Fives-Taylor, 1994), which likely 
determine the expression of specific adhesins. Cell surface entities that me- 
diate adherence include fimbriae (Rosan et al., 1988; Meyer and Fives-Taylor, 
1994), ExAmMat (Meyer and Fives-Taylor, 1994), and vesicles (Meyer and 
Fives-Taylor, 1993). Trypsin and protease treatment of smooth, nonfimbri- 
ated strains reduces adhesion of A. actinomycetemcomitans to epithelial cells, 
indicating that these nonfimbriae types of adhesins are proteinacious (Mintz 
and Fives-Taylor, 1994). 

Recently, a surface-associated protein determined to be an autotrans- $ 

Vi 

porter was shown to be involved in A. actinomycetemcomitans adhesion to o 
epithelial cells (Rose et al., 2003). The adhesin, called Aae for adhesion to o 
epithelial cells, is encoded by a gene (aae) that is homologous to autotrans- 2 
porter genes of Haemophilus influenzae and Neisseria species (St. Geme et £ 
al., 1994; St. Geme and Cutter, 2000). A unique feature of aae is a 135-base q 
repeat sequence that varies in number from strain to strain (Rose et al., 2003). « 
Four alleles have been identified to date. Lactoferrin in human milk whey £ 
was shown to decrease epithelial cell binding of A. actinomycetemcomitans 
strain 29523, but not of SUNY 465, a strain having an allele with fewer re- 
peats (Rose et al., 2003). On the basis of these results it was suggested that > 
the repeats may play a role in the binding of lactoferricin, the peptide on the g 
N -terminus of lactoferrin that interacts with and causes damaee to the cells' t» 



n 

M 
M 
M 

CO 



outer membrane (Yamauchi et al., 1993). Fewer copies of the repeats would 

effectively reduce the opportunity for binding. S 



> 
n 

H 
H 
H 



In summary, the adhesion of A. actinomycetemcomitans to epithelial cells q 

is multifactorial, with several adhesins and mechanisms playing a role. § 

n 

hi 

a 

HOST CELL SIGNALING PATHWAYS MODULATED IN RESPONSE g 

TO THE BACTERIAL CHALLENGE § 

Several lines of evidence suggest that A. actinomycetemcomitans infec- § 

tion and subsequent internalization is associated with activation of host cell 
signaling pathways. An active metabolic state and novel protein synthesis by 
both A. actinomycetemcomitans and the epithelial cell are required for invasion 
(Sreenivasan et al., 1993). In addition, entry is associated with an elevation of 
intracellular Ca 2+ levels (Fives-Taylor et al., 1996), a process associated with 




o 

1-1 

% 

w 

> 

I— I 

Ph 



< 

Q 

< 

< 
p^ 



O 



Pi 
w 

!* 

w 

CO 

i— i 

X 
u 

H 

X 
w 
Z 
< 




Figure 9.3. Real-time video microscopy frames of the interaction of A. 
actinomycetemcomitans strain SUNY 465 with KB epithelial cells. KB microvilli (arrow) 
become highly active and seek out bacteria (arrowhead) soon after infection (A). A 
microvillus makes contact with a bacterium (B). The microvillus seizes the A. 
actinomycetemcomitans and draws it toward and eventually into the KB cell (C). 



signaling events (Berridge, 1995) and characteristic of invasion by some bac- 
teria (Baldwin etal., 1991; Pace etal., 1993; Izutsuetal., 1996). It has also been 
shown that staurosporin, a wide-spectrum protein kinase inhibitor, decreases 
A. actinomycetemcomitans internalization, whereas genistein, a specific in- 
hibitor of tyrosine protein kinase, increases internalization (Fives-Taylor and 
Meyer, 1998). These results led to the hypothesis that staurosporin may be 
modulating a signaling pathway involved in the A. actinomycetemcomitans 
internalization process, whereas genistein is modulating one involved in its 
egression from the host cell. 

In related studies it was shown that an infection of KB cells produces 
changes in both its SDS-PAGE tyrosine-phosphorylated protein profile and 
immunofluorescence microscopy phosphotyrosine-labeling pattern (Fives- 
Taylor and Meyer, 1998). The fact that cross talk or signaling is an absolute 
requirement for A. actinomycetemcomitans invasion was ascertained by use of 5 

real-time video immunofluorescence microscopy (Fig. 9.3). Uninfected KB & 

cells and those subjected to nonviable A. actinomycetemcomitans are abso- z 

o 

lutely quiescent. In contrast, infection with viable A. actinomycetemcomitans * 

arouses the KB cells; their microvilli become highly active — seeking out, se- > 

questering, drawing in, and eventually engulfing the bacterium (Lippmann 3 

and Fives-Taylor, 1999). x 

M 
l— l 

> 
t- 1 

PHYSICAL MECHANISM OF INTERNALIZATION AND g 



INTRACELLULAR LOCATION, FATE OF BACTERIA, 




M 
M 

oo 

CO 

AND CELL-TO-CELL SPREAD > 

A. actinomycetemcomitans penetration of the gingival epithelium was § 

to 

demonstrated in early clinical studies (Saglie et al., 1986; Christersson et al., % 

1987). Those in vivo studies revealed that A. actinomycetemcomitans occurs £ 

in very specific intracellular locations and exhibits a very distinctive penetra- > 
tion pattern (Saglie et al., 1982, 1986, 1988). It has also been established that 



3 

o 



intracellular A. actinomycetemcomitans is present in vivo in human buccal § 

epithelial cells, where it occurs in clusters (Rudney et al., 2001). A. actino- g 

mycetemcomitans invasion of epithelial cells, both in an oral cell line (Meyer g 



n 



et al., 1991) and in human primary gingival cells (Fives -Taylor et al., 1995), o 

si 

has been demonstrated by use of in vitro studies . The overall A. actinomycetem- h 

comitans invasion process is dynamic and complex: it involves attachment to § 

the host cell, entry in a vacuole, escape from the vacuole into the cytoplasm, 
intracellular spread, and cell-to-cell spread (Meyer at al., 1996). Entry of the 
epithelial cell is initiated when interaction of A. actinomycetemcomitans with 
the epithelial cell triggers effacement of microvilli and formation of craters 



on the epithelial cell surface (Sreenivasan et al., 1993; Meyer et al., 1996). The 
majority of A. actinomycetemcomitans strains use an actin-dependent mecha- 
nism for invasion, but the mode of entry of a few strains is actin independent 
(Brissette and Fives-Taylor, 1998). The invasion process in strains that uti- 
lize the actin-independent mechanism is not well studied. Thus, a role for 
the two different mechanisms is not known; nor is it known whether actin 
dependence or independence is the only characteristic that separates these 
two invasion processes. 

The process of invasion described here is that of the A. actinomycetem- 
comitans invasion prototype, strain SUNY 465, a strain that utilizes the 
actin-dependent mechanism. Attachment of A. actinomycetemcomitans to 
the epithelial cell promotes rearrangement of actin from the periphery of 
the epithelial cell to a focal point beneath the organism at the point of entry 
g (Fives-Taylor et al., 1995). A. actinomycetemcomitans enters the epithelial cell 

5 through ruffled, lip-rimmed apertures (Fives-Taylor et al., 1996; Meyer et al., 




H 



g 1996). To date, two pathways believed to be associated with the entry of A. 

E actinomycetemcomitans into epithelial cells have been identified; the transfer- 

al rin receptor is implicated in one pathway, and integrins are implicated in the 

* other (Meyer et al., 1997a). 

5 Subsequent to the initial interaction and attachment, A. actinomycetem- 

| comitans enters the host cell in a membrane-bound vacuole by receptor- 

2 mediated endocytosis. In the classical endocytic pathway, macromolecules 

S are taken into early endosomes and delivered to lysosomes. Internalized or- 

e's ganisms are known to use a variety of means to avoid lysosomal degradative 

< enzymes; these include blockage of delivery to lysosomes, inhibition of en- 

rt dosome acidification, and acid activation of virulence factors that modify the 

n lysosome. A. actinomycetemcomitans trafficking within the vacuole, a process 

co that does not require endosomal acidification, is as follows. Within 30 min of 

£ infection, 40% of internalized A. actinomycetemcomitans are in the early en- 

g dosome. By 60 min, the number in the early endosome is greatly reduced and 

w A. actinomycetemcomitans are also present in the late endosome. By 2h, es- 

$ sentially all A. actinomycetemcomitans are associated with the late endosome, 

and by 3-4 h the vacuoles are devoid of bacteria and A. actinomycetemcomi- 
tans are present in both the cytoplasm and cell culture medium. Thus, the A. 
actinomycetemcomitans organisms avoid degradation by lysosomal enzymes 
by escaping from these organelles 3-4 h postinfection (Lippmann and Fives- 
Taylor, 2000) . 

The mechanism(s) used by A. actinomycetemcomitans for lysis of the host 
vacuole is not known. Preliminary studies suggest a role for leukotoxin in 
lysis of the vacuole (Meyer, unpublished results). A. actinomycetemcomitans 
possesses phospholipase C (PLC; see Meyer et al., 1997b), a molecule used 



by certain enteric pathogens for vacuole lysis (Camilli et al., 1991; Smith 
et al., 1995); thus, PLC is another possibility. 

Clearly, intracellular A. actinomycetemcomitans are not quiescent. A short 
time after escape from the vacuole into the cytoplasm, A. actinomycetemcomi- 
tans interact in a highly specific manner with host cell microtubules and 
utilize a microtubule -mediated mechanism for intracellular spread, as well 
as for spread to neighboring epithelial cells (Meyer et al., 1999). The spread 
to neighboring cells is by means of A. actinomycetemcomitans-induced inter- 
cellular protrusions, which are extensions of the host cell membrane that 
extend from one epithelial cell to another (Meyer et al., 1996). Bacteria can be 
seen within these protrusions by scanning, transmission, and fluorescent mi- 
croscopy (Meyer et al., 1996). The cell-to-cell spread involves movement and 
transfer through protrusions (Meyer et al., 1996), not engulfment of protru- 
sions as is the case with both Shigella and Listeria (Tilney and Portnoy, 1989; 5 
Kadurugamuwa et al., 1991). In addition to microtubules, the protrusions & 

contain microfilaments that may be involved in the structural framework z 

o 

of protrusions rather than mechanistically in the movement process (Meyer * 

etal, 1999). Sj 

Whereas the precise means by which A. actinomycetemcomitans usurps 3 

host cell microtubules for movement is not known, in vitro studies show that x 

A. actinomycetemcomitans localizes exclusively with the plus ends of micro- « 

tubules of taxol-induced microtubule asters (Fig. 9.4), indicating a specific A. n 

actinomycetemcomitans-microtubule interaction (Rose et al., 1998, 1999) . Fur- P 

thermore, both immunofluorescence microscopy and bactELISA indicate the 3 

presence of a kinesin-like entity on the surface of A. actinomycetemcomitans, Q 

thus implicating motor proteins in the movement along the microtubules § 

to 

(Meyer et al., 2000) . It is hypothesized that the kinesin entity on the A. actino- % 

mycetemcomitans surface mediates the interaction with microtubules and that £ 

the bacterium is transported in a manner similar to that of organelles and > 
vesicles, the usual cargo of microtubules. Whereas certain bacteria, such as 



3 

o 



Shigella and Listeria, usurp host cell actin to move within cells and to spread § 

to adjacent cells (Tilney and Portnoy, 1989; Kadurugamuwa et al., 1991), g 

this provides the first evidence that host cell dispersion of an intracellular S 



n 



pathogen involves usurpation of the host cell microtubule transport system. o 

Two genes have been implicated in A. actinomycetemcomitans invasion. h 

One is homologous to apaH, a gene that encodes RGD, a sequence known § 

to bind integrins (Saarela et al., 1999). A. actinomycetemcomitans DNA that 
contains the apaH gene confers on noninvasive Escherichia coli the ability 
to invade epithelial cells (Meyer et al., 1995; Saarela et al., 1999). It was also 
determined that insertional inactivation of apaH in A. actinomycetemcomitans 
substantially reduced its invasion (Lippmann and Fives-Taylor, unpublished 




o 

1-1 

% 

w 

> 

I— I 

Ph 



< 

Q 

< 

< 
p^ 



O 



Pi 
w 

!* 

w 

CO 

i— i 

X 
u 

H 

X 
w 
Z 
< 





Figure 9.4. Microtubule asters and the A. actinomycetemcomitans kinesin-like entity. 
Immunofluorescent microscopy of the interaction of A. actinomycetemcomitans strain 
SUNY 465 (arrows) with taxol-induced asters (panel A) and the kinesin-like motor protein 
(arrowheads) associated with the A. actinomycetemcomitans surface (panel B). The A. 
actinomycetemcomitans (green) bind specifically to the plus ends (periphery) of the asters 
(orange) (panel A). Binding of A. actinomycetemcomitans to the microtubules is believed to 
be mediated by the kinesin-like entities (yellow) on the A. actinomycetemcomitans surface 
(orange) (panel B). See color section. 

observation). Furthermore, RGD peptides inhibit A. actinomycetemcomitans 
invasion, whereas RAD peptides have no effect on invasion (Saarela et al., 
1999). The apaH gene is a homolog of invA, ialA, and ygdP, genes that 
are associated with invasion by Rickettsia prowazekii (Gaywee et al., 2002), 
Bartonella bacilliformis (Mitchell and Minnick, 1995), and E. coli Kl (Bessman 
et al., 2001), respectively. These genes produce proteins that are members 




Figure 9.5. Schematic representation of A. actinomycetemcomitans invasion of epithelial 
cells. Aae and ApaH have been identified as molecules on the surface of A. 
actinomycetemcomitans that can mediate its contact with epithelial cells. ApaH interacts 
with host cell integrins; the receptor to which Aae binds is not known. The transferrin 
receptor is also implicated as a receptor for A. actinomycetemcomitans. Entry associated 
with actin rearrangement and a Ca ++ flux occurs in a membrane-bound vacuole. The 
mechanism of escape from the vacuole is unclear, but leukotoxin (Ltx) and phospholipase 
C (PLC) are implicated. Once in the cytoplasm, A. actinomycetemcomitans can move within 
the host cell and spread to adjacent cells by usurping host cell microtubules. Kinesin-like 
motor proteins on the A. actinomycetemcomitans surface are implicated in mediating its 
interaction with microtubules. See color section. 

of the Nudix family of hydrolases, which catalyze the dinucleoside polyphos- 
phates, a class of signaling nucleotides (Conyers and Bessman, 1999; Saarela 
et al., 1999; Bessman et al., 2001). It has also been reported that A. actino- 
mycetemcomitans invasion involves genes that share some homology to spa 
genes, genes that are involved in the export of proteins from cells (Laing- 
Gibbard et al., 1998). 

In summary, it is clear that the process by which A. actinomycetemcomi- 
tans enters and escapes from host cells is very complex (Fig. 9.5). Future 
studies should reveal even more dynamic interactions and lead to targeting 
of mechanisms suitable for therapeutic intervention. 



285 



< 
> 

on 
O 

z 

o 



M 

M 

i— i 
H 

X 

M 

M 
I— I 

> 

n 

M 
M 
M 

oo 

CO 
> 

o 

> 

n 

H 
H 
H 

> 
O 

S 
n 

hi 

s 

n 

o 

£ 




CONSEQUENCES OF INVASION WITH REGARD TO INNATE 
HOST RESPONSE 

The first line of defense of the host against intrusive bacteria is phagocyte 
recruitment (chemotaxis) to the region. A number of steps are involved in 
this process: binding of chemotactic signaling factors, upregulation of adhe- 
sion receptors, binding to the endothelium, and movement of phagocytes to 
the underlying tissues. Whereas invasion per se is a means by which bac- 
teria can ultimately escape this response, it initially represents a significant 
challenge to intrusive organisms. Thus, the ability to disrupt chemotaxis pro- 
motes survival of pathogenic organisms. A. actinomycetemcomitans secretes a 
low-molecular-weight protein that can inhibit polymorphonuclear leukocyte 
chemotaxis (Van Dyke et al., 1982; Ashkenazi et al., 1992). It is also known 
that A. actinomycetemcomitans capsular-like serotype-specific polysaccharide 

2 antigen (SPA) plays an important role in its ability to resist phagocytosis and 

<? 

h killing by polymorphonuclear lymphocytes (Yamaguchi et al., 1995). Expres- 

> sion of chemoattractant protein 1 (MCP-1) and neutrophil chemotactic factor 

<j I L-8 mRNA is increased in monocytes after stimulation with S PA (Yamaguchi 

% etal, 1996). 

Polymorphonuclear leukocytes can also kill bacteria by fusing with lyso- 

1 somes from which they acquire potent antibacterial agents. Bacteria able to 
5j inhibit the fusion or ward off the antibactericidal action are protected. A. acti- 
ph nomycetemcomitans can inhibit the production by polymorphonuclear leuko- 
^ cytes of some of these compounds, and it is resistant to others. A heat-stable 
& protein in A. actinomycetemcomitans inhibits the production of hydrogen 
SL peroxide by polymorphonuclear leukocytes (Ashkenazi etal., 1992), and many 
w strains are intrinsically resistant to high concentrations of hydrogen peroxide 

2 (Miyasaki et al., 1984). In addition, A. actinomycetemcomitans is resistant to a 



CO 



g number of defensins, which are cationic peptides that occur in neutrophils 

5 (Miyasaki etal, 1990). 

x A. actinomycetemcomitans induces apoptotic cell death in murine macro- 



< 



5 phages (J 774.1 cells) in vitro. Entry of A. actinomycetemcomitans into the 

macrophages is an absolute requirement for the cytotoxicity, that is, nuclear 
morphology changes and an increase in the proportion of fragmented DNA 
(Kato et al., 1995). Studies suggest that protein kinase C signaling regu- 
lates the apoptosis (Nonaka et al., 1997). With the use of LR-9 cells, CD14- 
defective mutants of J774.1 cells, it was determined that CD14 molecules 
likely participate in the phagocytosis of A. actinomycetemcomitans, as well as 
in the regulation of the apoptotic events (Muro et al., 1997). Both caspase-1 
and caspase-3 appear to play a role in the A. actinomycetemcomitans-induced 



apoptosis in macrophages (Nonaka et al., 2001). In general, infected macro- 
phages kill bacteria within phagosomes by means of nitric oxide (NO). In 
this regard, it has been reported that NO affords A. actinomycetemcomitans- 
infected murine macrophages partial protection from apoptosis by decreasing 
caspase activity (Nakashima et al., 2002). An interesting caveat that requires 
further investigation is the finding that A. actinomycetemcomitans LPS stimu- 
lates the production of NO, a process that involves activation of protein kinase 
C and protein tyrosine kinase, as well as the regulatory control of cytokines 
(Sosroseno et al., 2002). 



CORRELATION AMONG INVASION, PATHOGENICITY, 
AND CLINICAL PRESENTATION 

A lack of an adequate animal model in which to study invasion precludes, 



REFERENCES 

Ashkenazi, M., White, R.R., and Dennison, D.K. (1992). Neutrophil modulation 
by Actinobacillus actinomycetemcomitans. I. Chemotaxis, surface receptor ex- 
pression and F-actin polymerization. J. Periodontal Res. 27, 264—273. 




< 
> 

on 



at this juncture, the ability to establish a correlation among A. actinomycetem- 8 

comitans invasion, pathogenicity, and clinical presentation. On the basis of o 



►n 



the collective in vivo and in vitro observations made to date, the hypothesis g 

is put forth that invasion of epithelial cells per se and the dynamic process £ 

of intercellular and intracellular spread are the means by which A. actino- q 

mycetemcomitans migrates to gingival and connective tissue and initiates the ™ 

destruction associated with periodontal disease. The demonstration of A. acti- £ 

nomycetemcomitans in buccal epithelial cells prompted the hypothesis that * 

bacteria inside exfoliated cells would be afforded a protected environment „ 

for exchange between various oral niches both within and between human > 

subjects (Rudney et al., 2001). The rough-to-smooth and smooth-to-rough g 

variant shifts, with concomitant shifts in adhesive-to-invasive and invasive- g 

to-adhesive capabilities, respectively, may account in part for the episodic £ 

nature of periodontal disease. S 

it, 

A total understanding of the A. actinomycetemcomitans invasion process q 

and its role in pathogenicity (Fives-Taylor etal., 1999) is all the more important § 

today in light of increasing evidence that indicates that there is a link between 5 

periodontal disease and systemic disorders (DeStefano et al., 1993; Beck h 

et al, 1996; Meyer and Fives-Taylor, 1998; Teng et al, 2002). J5 

o 



2 




Baldwin, T.J., Ward, W., Aitken, A., Knutton, S., and Williams, P.H. (1991). El- 
evation of intracellular free calcium levels in HEp-2 cells infected with en- 
teropathogenic Escherichia coli. Infect. Immun. 59, 1599-1604. 
Beck, J., Garcia, R., Heiss, G., Vokonas, P.S., and Offenbacher, S. (1996). Peri- 
odontal disease and cardiovascular disease. J. Periodontol. 67, 1123-1137. 
Berridge, M.J. (1995). Calcium signaling and cell proliferation. Bioessays 17, 491- 

500. 
Bessman, M.J., Walsh, J.D., Dunn, C.A., Swaminathan, J., Weldon, J.E., and 
Shen, J. (2001). The gene ygdP, associated with the invasiveness of Es- 
cherichia coli Kl, designates a nudix hydrolase, Orfl76, active on adenosine 
(S'J-pentaphospho-fS'J-adenosine (Ap 5 A). J. Biol. Chem. 276, 37,834-37,838. 
Braun, W. (1965). Bacterial genetics. Philadelphia: W.B. Saunders. 
Brissette, C.A. and Fives-Taylor, P.M. (1998). Actinohacillus actinomycetemcomi- 
g tans may utilize either actin-dependent or actin-independent mechanisms of 

5 invasion. Oral Microbiol. Immunol. 13, 137-142. 

g Camilli, A., Goldfine, H., and Portnoy, D.A. (1991). Listeria monocytogenes mutants 

E lacking phosphatidylinositol-specific phospholipase C are avirulent. J. Exp. 

g Med. 173, 751-754. 

* Christersson, L.A., Albini, B., Zambon, J. J., Wikesjo, U.M., and Genco, R.J. 

§ (1987). Tissue localization of Actinohacillus actinomycetemcomitans in human 

| periodontitis. I. Light, immunofluorescence and electron microscopic stud- 

2 ies. J. Periodontol. 58, 529-539. 

S Conyers, G.B. and Bessman, M.J. (1999). The gene, ialA, associated with invasion 

w of human erythrocytes by Bartonella bacilliformis, designates a nudix hydro- 

< lase active on dinucleoside 5' -polyphosphate. J. Biol. Chem. 274, 1203-1206. 

tf DeStefano, F., Anda, R.F., Kami, S., Williamson, D.F., and Russell, CM. (1993). 

n Dental disease and risk of coronary heart disease and mortality. BMJ 306, 

688-691. 

g 

£ Fives-Taylor, P.M. and Meyer, D.H. (1998). The complex, multistep process of 

invasion of epithelial cells by the periodontopathogen, Actinohacillus actino- 

w mycetemcomitans, In The 2nd Indiana Conference: Microbial Pathogenesis, Cur- 

$ rent and Emerging Issues, ed. D.J. LeBlanc, M.S. Lantz, and L.M. Switalski, pp. 

3-16 Indianapolis: Indiana University. 
Fives-Taylor, P.M., Hutchins Meyer, D., Mintz, K.P., and Brissette, C. (1999). Acti- 
nohacillus actinomycetemcomitans: a causative agent of destructive periodontal 
disease. Periodontology 20, 136-167. 
Fives-Taylor, P., Meyer, D., and Mintz, K. (1995). Characteristics of Actinohacillus 
actinomycetemcomitans invasion of and adhesion to cultured epithelial cells. 
Adv. Dent. Res. 9, 55-62. 



Fives-Taylor, P., Meyer, D., and Mintz, K. (1996). Virulence factors ofthe periodon- 
topathogen Actinohacillus actinomycetemcomitans. J. Periodontal. 67, 291-297. 

Gaywee, J., Xu, W., Radulovic, S., Bessman, M.J., and Azad, A.F. (2002). The 
Rickettsia prowazekii invasion gene homolog (invA) encodes a nudix hydrolase 
active on adenosine (5') -pentaphospho-(5') -adenosine. Mol. Cell Proteomics 1, 
179-183. 

Haase, E.M., Zmuda, J.L., and Scannapieco, F.A. (1999). Identification and molec- 
ular analysis of rough-colony-specific outer membrane proteins of Acti- 
nohacillus actinomycetemcomitans. Infect. Immun. 67, 2901-2908. 

Hammond, B.F., Darkes, M., Lai, C, and Tsai, C.C. (1981). Isolation and charac- 
terization of membrane vesicles of Actinohacillus actinomycetemcomitans. J. 
Dent. Res. 60, 333. 

Hammond B.F., Lillard, S.E., and Stevens, R.H. (1987). A bacteriocin of Acti- 
nohacillus actinomycetemcomitans. Infect. Immun. 55, 686-691. 5 

Holt, S.C., Tanner, A.C., and Socranzky, S.S. (1980). Morphology and ultra struc- % 

o 
ture of oral strains of Actinohacillus actinomycetemcomitans and Haemophilus % 

o 
aphrophilus. Infect. Immun. 30, 588-600. * 

Inoue, T., Tanimoto, I., Ohta, H., Kato, K., Murayama, Y., and Fukui, K. (1998). > 

Molecular characterization of low-molecular-weight component protein, Flp, 3 

in Actinohacillus actinomycetemcomitans fimbriae. Microhiol. Immunol. 42, a 

253-258. 



epithelial cells and Porphyromonas gingivalis. FEMS Microhiol. Lett. 144, 145- 
150. 







> 

Inouye, T., Ohta, H., Kokeguchi, S., Fukui, K., and Kato, K. (1990). Colonial n 



M 



variation and fimbriation of Actinohacillus actinomycetemcomitans. FEMS Mi- 
crohiol. Lett. 57, 13-17. " 
Ishihara, K., Honma, K., Miura, T., Kato, T., and Okuda, K. (1997). Cloning Q 
and sequence analysis of the fimbriae associated protein (fap) gene from § 

to 

Actinohacillus actinomycetemcomitans. Microh. Pathog. 23, 63-69. % 

I- . 

Izutsu, K.T., Belton, CM., Chan, A., Fatherazi, S., Kanter, I. P., Park, Y., and h 

Lamont, R.J. (1996). Involvement of calcium in interactions between gingival > 

o 

S 
^< 

Kadurugamuwa, J.L., Rohde, M., Wehland, J., and Timmis, K.N. (1991). Intra- Q 

cellular spread of Shigella flexneri through a monolayer mediated by mem- S 

branous protrusions and associated with reorganization of the cytoskeletal o 

protein vinculin. Infect. Immun. 59, 3463-3471. h 

Kato S., Muro, M., Akifusa, S., Hanada, N., Semba, I., Fujii, T., Kowashi, Y., S 

and Nishihara, T. (1995). Evidence for apoptosis of murine macrophages 
by Actinohacillus actinomycetemcomitans infection. Infect. Immun. 63, 3914- 
3919. 




Kato, S., Kowashi, Y., and Demuth, D.R. (2001). Outer membrane-like vesicles 

secreted by Actinohacillus actinomycetemcomitans. Microh. Pathog. 32, 1-13. 
Lai, C.H., Listgarten, M.A., and Hammond, B.F. (1981). Comparative ultrastruc- 
ture of leukotoxic and non-leukotoxic strains of Actinohacillus actinomycetem- 
comitans. J. Periodontal Res. 16, 379-389. 
Laing-Gibbard, L.P., Lepine, G., and Ellen, R.P. (1998). DNA fragments of Acti- 
nohacillus actinomycetemcomitans involved in invasion of KB cells. J. Dent. 
Res. 77SI-B, 770. 
Lippmann, J.E. and Fives-Taylor, P.M. (2000). Co -localization of intracellular A. 
actinomycetemcomitans with vacuoles containing early endosomal and late 
endosomal proteins. J. Dent. Res. 79SI, 255. 
Lucia, L.F., Farias, F.F., Eustaquio, C.J., Auxiliadora, M., Carvalho, R., Alviano, 
C.S., and Farias L.M. (2002). Bacteriocin production by Actinohacillus acti- 

g nomycetemcomitans isolated from the oral cavity of humans with periodontal 

5 disease, periodontally healthy subjects and marmosets. Res. Microhiol. 153, 

t 45-52. 

E Meyer, D.H. and Fives-Taylor, P.M. (1993). Evidence that extracellular compo- 

g nents function in adherence of Actinohacillus actinomycetemcomitans to ep- 

* ithelial cells. Infect. Immun. 61, 4933-4936. 

§ Meyer, D.H. and Fives-Taylor, P.M. (1994). Characteristics of adherence of Acti- 

| nohacillus actinomycetemcomitans to epithelial cells. Infect. Immun. 62, 928- 

S 935. 

S Meyer, D.H. and Fives-Taylor, P.M. (1998). Oral pathogens: from dental plaque 

w to cardiac disease. Curr. Opin. Microhiol. 1, 88-95. 

< Meyer, D.H., Lippmann, J.E., and Fives-Taylor, P.M. (1996). Invasion of epithelial 

« cells by Actinohacillus actinomycetemcomitans: a dynamic, multistep process. 

£ Infect. Immun. 64, 2988-2997. 

w Meyer, D.H., Rose, J.E., Lippmann, J.E., and Fives-Taylor, P.M. (1999). Micro- 

£ tubules are associated with intracellular movement and spread of the pe- 

riodontopathogen Actinohacillus actinomycetemcomitans. Infect. Immun. 67, 

n 6518-6525. 

$ Meyer, D.H., Wei, J., and Fives-Taylor, P.M. (1995). Cloning of a DNA fragment 

associated with Actinohacillus actinomycetemcomitans invasion. J. Dent. Res. 
74SI, 200. 
Meyer, D.H., Mintz, K.P., and Fives-Taylor, P.M. (1997a). Models of invasion of 
enteric and periodontal pathogens into epithelial cells: a comparative analy- 
sis. Crit. Rev. Oral Biol. Med. 8, 389-409. 
Meyer, D.H., Sreenivasan, P.K., and Fives-Taylor, P.M. (1991). Evidence for inva- 
sion of a human oral cell line by Actinohacillus actinomycetemcomitans. Infect. 
Immun. 59, 2719-2726. 




Meyer, D.H., Fives-Taylor, P.M., and Rose, J.E. (2000). Actinobacillus actino- 
mycetemcomitans displays an entity that binds antibody to a kinesin-like mi- 
crotubule motor protein. J. Dent. Res. 79SI, 256. 

Meyer, D.H., Mackie, T.N., and Fives-Taylor, P.M. (1997b). Actinobacillus actino- 
mycetemcomitans exhibits phospholipase C-B (PC-PLC) activity. J. Dent. Res. 
76SI, 26. 

Mintz, K.P. and Fives-Taylor, P.M. (1994). Adhesion of Actinobacillus actino- 
mycetemcomitans to a human oral cell line. Infect. Immun. 62, 3672-3678. 

Mitchell, S.J. and Minnick, M.F. (1995). Characterization of a two-gene locus from 
Bartonella bacilliformis associated with the ability to invade human erythro- 
cytes. Infect. Immun. 63, 1552-1562. 

Miyasaki, K.T., Bodeau, A.L., Ganz, T., Selsted, M.E., and Lehrer, R.I. (1990). In 

vitro sensitivity of oral, gram-negative, facultative bacteria to the bactericidal 

activity of human neutrophil defensins. Infect. Immun. 58, 3934-3940. 5 

Miyasaki, K.T., Wilson, M.E., Reynolds, H.S., and Genco, R.J. (1984). Resistance 8j 

o 

of Actinobacillus actinomycetemcomitans and differential susceptibility of oral s 

o 

Haemophilus species to the bactericidal effects of hydrogen peroxide. Infect. * 

Immun. 46, 644-648. > 

Muro, M., Koseki, T., Akifusa, S., Kato, S., Kowashi, Y., Ohsaki, Y., Yamato, 3 

l-H 

Y., Nishijima, M., and Nishihara, T. (1997). Role of CD14 molecules in in- a 

ternalization of Actinobacillus actinomycetemcomitans by macrophages and C 

subsequent induction of apoptosis. Infect. Immun. 65, 1147-1151. n 

Nakashima K., Tomioka, J., Kato, S., Nishihara, T., and Kowashi, Y. (2002). Ni- £ 

trie oxide-mediated protection of A. actinomycetemcomitans-infected murine *< 

macrophages against apoptosis. Nitric Oxide 6, 61-68. Q 

Nonaka K., Ishisaki, A., Muro, M., Kato, S., Oido, M., Nakashima, K., Kowashi, Y., § 

to 

and Nishihara, T. (1997). Possible involvement of protein kinase C in apop- % 

I- . 

totic cell death of macrophages infected with Actinobacillus actinomycetem- g 

comitans. FEMS Microbiol. Lett. 159, 247-254. £ 

n 

Nonaka K., Ishisaki, A., Okahashi, N., Koseki, T., Kato, S., Muro, M., Nakashima, 2 

K., Nishihara, T., and Kowashi, Y. (2001). Involvement of caspases in apop- § 

totic cell death of murine macrophages infected with Actinobacillus actino- Q 

mycetemcomitans. J. Periodontal Res. 36, 40-47. > 

Nowotny, A., Behling, U.H., Hammond, B., Lai, C.H., Listgarten, M., Pham, o 

P.H., and Sanavi, F. (1982). Release of toxic microvesicles by Actinobacillus h 

actinomycetemcomitans. Infect. Immun. 37, 151-154. § 

Pace, J., Hayman, M.J., and Galan, J.E. 1993. Signal transduction and invasion of 
epithelial cells by Salmonella typhimurium. Cell 72, 505-514. 

Preus, H.R., Namork, E., and Olsen, I. (1988). Fimbriation of Actinobacillus acti- 
nomycetemcomitans. Oral Microbiol. Immunol. 3, 93-94. 



Rosan, B., Slots, J., Lamont, R.J., Listgarten, M.A., and Nelson, G.M. (1988). 

Actinobacillus actinomycetemcomitans fimbriae. Oral Microbiol. Immunol. 3, 

58-63. 
Rose, J.E., Meyer, D.H., and Fives-Taylor, P.M. (1998). Detection of bacteria- 

microtubule interactions in a cell-free extract. Meth. Cell Sci. 19, 325- 

330. 
Rose, J.E., Meyer, D.H., and Fives-Taylor, P.M. (1999). Actinobacillus actino- 
mycetemcomitans binds specifically to the plus ends of microtubules. J. Dent. 

Res. 78SI, 133. 
Rose, J.E., Meyer, D.H., and Fives-Taylor, P.M. (2003). Aae, an autotransporter 

involved in adhesion of Actinobacillus actinomycetemcomitans to epithelial 

cells. Infect. Immun. 71, 2386-2393. 
Rudney, J.D., Chen, R., and Sedgewick, G.J. (2001). Intracellular Actinobacillus 
g actinomycetemcomitans and Porphyromonas gingivalis in buccal epithelial cells 

5 collected from human subjects. Infect. Immun. 69, 2700-2707. 




H 



£ Saarela, M., Lippmann, J.E., Meyer, D.H., and Fives-Taylor, P.M. (1999). Acti- 



> 



E nobacillus actinomycetemcomitans apaH is implicated in invasion of epithelial 

g cells. J. Dent. Res. 78SI, 259. 



< 



* Saglie, F.R., Marfany, A., and Camargo, P. (1988). Intragingival occurrence of 

§ Actinobacillus actinomycetemcomitans and Bacteroides gingivalis in active de- 

ll structive periodontal lesions. J. Periodontol. 59, 259-265. 

2 Saglie, F.R., Smith, C.T., Newman, M.G., Carranza, F.A. Jr., Pertuiset, J.H., 

S Cheng, L., Auil, E., and Nisengard, R.J. (1986). The presence of bacteria in 

w the oral epithelium in periodontal disease. II. Immunohistochemical identi- 

< fication of bacteria. J. Periodontol. 57, 492-500. 

pj Saglie, F.R., Carranza, F.A. Jr., Newman, M.G., Cheng, L., and Lewin, K.J. (1982). 

n Identification of tissue-invading bacteria in human periodontal disease. J. 

Periodontal Res. 17, 452-455. 
£ Scannapieco, F.A., Kornman, K.S., and Coykendall, A.L. (1983). Observation of 

fimbriae and flagella in dispersed subgingival dental plaque and fresh bac- 

w terial isolates from periodontal disease. J. Periodontal Res. 18, 620-633. 

$ Scannapieco, F.A., Millar, S.J., Reynolds, H.S., Zambon, J.J., and Levine, M.J. 

(1987). Effect of anaerobiosis on the surface ultrastructure and surface pro- 
teins of Actinobacillus actinomycetemcomitans (Haemophilus actinomycetem- 
comitans). Infect. Immun. 55, 2320-2323. 
Slots, J. (1982). Selective medium for isolation of Actinobacillus actinomycetem- 
comitans. J. Clin. Microbiol. 15, 606-609. 
Smith, G.A., Marquis, H., Jones, S., Johnston, N.C., Portnoy, D.A., and Goldfine, 
H. (1995). The two distinct phospholipases C of Listeria monocytogenes have 
overlapping roles in escape from a vacuole and cell-to-cell spread. Infect. 
Immun. 63, 4231-4237. 



Sosroseno, W., Barid, I., Herminajeng, E., and Susilowati, H. (2002). Nitric ox- 
ide production by a murine macrophage cell line (RAW 264.7) stimulated 
with lipopolysaccharide from Actinohacillus actinomycetemcomitans. Oral Mi- 
crobiol. Immunol. 17, 72-78. 

Sreenivasan, P.K., Meyer, D.H., and Fives-Taylor, P.M. (1993). Requirements for 
invasion of epithelial cells by Actinohacillus actinomycetemcomitans. Infect. 
Immun. 61, 1239-1245. 

Stevens, R.H., Lillard, S.E., and Hammond, B.F. (1987). Purification and bio- 
chemical properties of a bacteriocin from Actinohacillus actinomycetemcomi- 
tans. Infect. Immun. 55, 692-697. 

St. Geme, J.W. Ill and Cutter, D. (2000). The Haemophilus influenzae Hia adhesin 
is an autotransporter protein that remains uncleaved at the C terminus and 
fully cell associated. J. Bacteriol. 182, 6005-6013. 

St. Geme, J.W. Ill, de la Morena, M.L., and Falkow, S. (1994). A Haemophilus in- 5 

fluenzae IgA protease-like protein promotes intimate interaction with human ^ 

epithelial cells. Mol. Microhiol. 14, 217-233. 3 

o 
Teng, Y.T., Taylor, G.W., Scannapieco, F., Kinane, D.F., Curtis, M., Beck, J.D., and * 

Kogon, S. (2002). Periodontal health and systemic disorders. J. Can. Dent. > 

Assoc. 68, 188-192. 5 

l-H 

Tilney, L.G. and Portnoy, D.A. (1989). Actin filaments and the growth, movement, a 

and spread of the intracellular bacterial parasite, Listeria monocytogenes. J. Cell C 

Biol. 109, 1597-1608. n 

M 

Van Dyke, T.E., Bartholomew, E., Genco, R.J., Slots, J., and Levine, M.J. (1982). £ 

Inhibition of neutrophil chemotaxis by soluble bacterial products. J. Peri- *< 

odontol. 53, 502-508. Q 

Wilson, M., Kamin, S., and Harvey, W. (1985). Bone resorbing activity of purified § 

to 

capsular material from Actinohacillus actinomycetemcomitans. J. Periodontal % 

Res. 20, 484-491. B 

Yamaguchi, N., Kawasaki, M., Yamashita, Y., Nakashima, K., and Koga, T. (1995). > 

n 

Role of capsular polysaccharide -like serotype-specific antigen in resistance 2 

of Actinohacillus actinomycetemcomitans to phagocytosis by human polymor- § 

phonuclear leukocytes. Infect. Immun. 63, 4589-4594. Q 

Yamaguchi, N., Yamashita, Y., Ikeda, D., and Koga, T. (1996). Actinohacillus actino- S 

mycetemcomitans serotype b-specific polysaccharide antigen stimulates pro- o 

duction of chemotactic factors and inflammatory cytokines by human mono- h 

cytes. Infect. Immun. 64, 2563-2570. § 

Yamauchi, K., Tomita, M., Giehl, T.J., and Ellison, R.T. III. (1993). Antibacte- 
rial activity of lactoferrin and a pepsin-derived lactoferrin peptide fragment. 
Infect. Immun. 61, 719-728. 

Zambon, J.J. (1985). Actinohacillus actinomycetemcomitans in human periodontal 
disease. J. Clin. Periodontol. 12, 1-20. 



CHAPTER 10 

Invasion by Porphyromonas gingivalis 

Ozlem Yilmaz and Richard J. Lamont 



Porphyromonas gingivalis cells are Gram-negative, anaerobic, nonmotile short 
rods that produce black pigmented colonies on blood agar. The taxonomy of 
the species dates back to 1921 when Oliver and Wherry isolated an organism 
from a variety of oral and nonoral sites that they were to designate Bacterium 
melaninogenicum. This heterogeneous grouping was later subdivided into 
nonfermenters, weak fermenters, and strong fermenters. After a number of 
status changes within the genus Bacteroides, asaccharolytic oral isolates were 
assigned to the taxon P. gingivalis. The primary ecological niche of P. gingivalis 
is in the subgingival crevice, the gap between the surfaces of the tooth and the 
gingiva (gum); however, the organism can be found elsewhere in the mouth, 
including supragingival (above the gum) tooth surfaces, the tongue, tonsils, 
and buccal (cheek) mucosa. Although the species has been associated with 
odontogenic abscesses and nonoral infections (discussed later), the primary 
pathogenic potential of P. gingivalis is in periodontal disease. The periodon- 
tal tissues include the gingiva, periodontal ligament, and alveolar bone, and 
they constitute the supporting tissues of the teeth. Chronic destruction of the 
periodontium, such as occurs in periodontal diseases, can eventually lead to 
exfoliation of teeth and is the most common cause of tooth loss in adults. 
Periodontal diseases vary in severity and age of onset, and P. gingivalis is 
associated, either alone or in combination with other bacteria, with the most 
severe manifestations. However, the frequent occurrence of the organism in 
healthy adults and in young children indicates that a complex interplay be- 
tween host and pathogen exists, and that disruption of this ecological balance 
is required for disease to ensue. 

In the gingival crevice the area of contact between the gingiva and the 
tooth is known as the junctional epithelium, and it is characterized by a 
lack of keratinization, limited differentiation, and a relatively permeable 





structure. In destructive periodontal disease there is migration of the junc- 
tional epithelium, resulting in enlargement of the crevice into a deeper pe- 
riodontal pocket that contains inflammatory cells such as neutrophils and 
T cells. The gingiva itself also contains immune cells, including B cells, 
T cells, and dendritic cells. The microbiota of the gingival area in both health 
and disease is complex, with at least 500 species of bacteria present in the 
gingival crevice. Colonization and persistence by periodontal pathogens re- 
quire, therefore, successful encounters with antecedent bacterial species and 
with host eukaryotic cells. Consistent with these constraints, P. gingivalis can 
bind to, invade, and survive inside a variety of host cells, including epithelial 
and endothelial cells. 



P. GINGIVALIS ADHESION 

H 
% 

§ An early event in the process of bacterial internalization is adhesion to 

^ the host cell surface. Attachment of bacteria is often required to trigger the 

§ signaling pathways within the host cells that ultimately induce the mem- 

= brane and cytoskeletal rearrangements that bring the bacteria into the cell. 

* P. gingivalis is endowed with a multiplicity of adhesins that mediate adhesion 

% to epithelial cells, endothelial cells, fibroblasts, and erythrocytes, and to com- 

< ponents of the extracellular matrix, namely laminin, elastin, fibronectin, type 

~ I collagen, thrombospondin, and vitronectin (Sojar et al., 1995, 1999; Kontani 

§ et al., 1997; Nakamura et al., 1999; Dorn et al., 2000; Lamont and Jenkinson, 

.g 2000). Adhesive activity has been demonstrated for fimbriae, outer mem- 

brane proteins, and proteases. These molecules may function collectively, 
and their activities are integrated and controlled at the transcriptional and 
posttranslational levels (reviewed in Lamont and Jenkinson, 1998, 2000). 

The major fimbriae of P. gingivalis are distinct from other Gram-negative 
fimbriae and do not appear to belong to any of the existing classifications. 
The fimbriae are composed of an ^43-kDa fimbrillin (FimA) monomer that 
possesses a number of binding domains for individual substrate recognition. 
The functional domain of FimA for epithelial cells has been localized to a re- 
gion spanning amino acid residues 49-90, although the boundaries of the 
interactive site are not known precisely (Sojar et al., 1999) . Fimbriae-mediated 
binding is important for subsequent invasion, as FimA-deficient mutants are 
attenuated in their ability to internalize, and both purified fimbrillin and an- 
tibodies to fimbriae can block invasion (Njoroge et al., 1997; Weinberg et al., 
1997). Furthermore, fimbrillin-coated microspheres are efficiently taken 
up by epithelial cells (Nakagawa et al., 2002). A number of epithelial cell 
molecules have been found to function as cognate receptors for fimbrillin, 
including a 48-kDa surface protein, cytokeratins, and integrins (Weinberg 



et al., 1997; Sojar et al., 2002; Yilmaz et al., 2002). With regard to stimulation 
of invasion, fimbriae-integrin binding may be the most significant interac- 
tion, as integrin antibodies can inhibit P. gingivalis invasion of epithelial cells 
(Yilmaz et al., 2002). Moreover, as integrins are initiators of signal trans- 
duction pathways, the engagement of an integrin receptor by P. gingivalis 
fimbriae may be a means by which the organism begins to seize control of 
host cell signaling machinery. P. gingivalis fimbriae are also involved in the 
adhesion-dependent invasion of endothelial cells (Deshpande et al., 1998; 
Khlgatian et al., 2002), although other adhesins may work in concert (Dorn 
etal.,2000). 

The Jim A gene is monocistronic, although immediately downstream are 
four genes whose products may be associated with the mature fimbriae 
(Watanabe et al., 1996). The fimA upstream region contains a functionally 
active sigma-70-like promoter consensus sequence along with a potential UP 5 

element (Xie and Lamont, 1999). AT-rich sequences upstream of the RNA & 

polymerase binding sites are involved in positive regulation of transcriptional z 

activity. Environmental cues to which the fimA promoter responds include ^ 

temperature, hemin concentration, and salivary molecules (Xie et al., 1997), a 

which are parameters with relevance to conditions in the oral cavity. The^mA 53 

gene can also be positively autoregulated by the FimA protein (Xie et al., 2 

2000). In addition, expression of FimA decreases following association of % 

P. gingivalis with epithelial cells (Wang et al., 2002). Thus, after completing £ 

i-i 

the task of inducing invasion, fimbriae may no longer be required by the § 

intracellular P. gingivalis cells. As fimbrillin has a number of immunostim- g 

ulatory properties, such as induction of cytokines and chemokines (Ogawa & 

et al., 1994), reduced FimA expression may aid in immune avoidance by the 
organism. 

Although the primary function of proteinases secreted by the asaccha- 
rolytic P. gingivalis is the provision of nutrients, proteinases are also involved 
both directly and indirectly in adhesion. Several distinct proteinases are pro- 
duced by P. gingivalis (reviewed in Potempa et al., 1995; Kuramitsu, 1998; 
Curtis et al., 1999), and direct enzyme-substrate interactions may be able 
to effectuate adhesion, although such binding is unlikely to persist for ex- 
tended periods. Of greater importance, the C-terminal coding regions of 
the Arg-X and Lys-X specific proteases RgpA and Kgp contain extensive re- 
gions with hemagglutinin activity (Barkocy-Gallagher et al., 1996; Lamont and 
Jenkinson, 1998; Curtis et al., 1999). These regions, therefore, can be pre- 
dicted to bind directly to human cell surface receptors. Proteinases can also 
contribute to adherence through the partial degradation of substrates result- 
ing in the subsequent exposure of epitopes for adhesin recognition. For exam- 
ple, hydrolysis of fibronectin or other matrix proteins by the Arg-X specific 



proteases RgpA and RgpB displays C-terminal Arg residues that mediate 
fimbriae -dependent binding (Kontani et al., 1996). Other means by which 
RgpA and RgpB can contribute to adhesion are through processing the leader 
peptide from the flmbrillin precursor (Nakayama et al., 1996), and by upreg- 
ulating transcription of the fimA gene (Tokuda et al., 1996; Xie et al., 2000). 
In addition to the hemagglutinin-associated activities of the RgpA and 
Kgp proteinases, several additional hemagglutinin (hag) genes are present 
in the P. gingivalis genome. The hagA gene encodes a large protein of over 
230 kDa containing four contiguous direct 440-456 aa residue repeat blocks 
(Han et al., 1996). Each repeat block may represent a functional hemagglu- 
tinin domain, and similar hagA-hke sequences are found at multiple sites in 
the chromosome. The hagB and hagC genes are at distinct chromosomal loci, 
although the HagB and C proteins (~40 kDa) are very similar (Progulske- 
g Fox et al., 1995). A minimal peptide motif PVQNLT has been shown to be 




o 



§ associated with hemagglutinating activity and is found within the proteinase- 



i 



Q 



hemagglutinin sequences (Shibita et al., 1999) and at multiple chromosomal 



§ sites. 

< 



g Thus, P. gingivalis possesses a variety of adhesins with differing recep- 

* tor specificities and affinities that can potentially impinge to varying degrees 



^ upon diverse receptor-dependent host cell biochemical pathways. This may 

< allow the organism to utilize more than one pathway for internalization. For 

= example, although fimbriae-deficient mutants show reduced uptake into ep- 

§ ithelial cells, a low level of invasion remains (Weinberg et al., 1997). Nonfim- 

g brial dependent uptake, although less efficient than FimA-mediated uptake, 

may result from attachment by other surface adhesins. 



UPTAKE OF P. GINGIVALIS BY HOST CELLS 

P. gingivalis can invade epithelial cells, endothelial cells, and dendritic 
cells (reviewed in Lamont and Jenkinson, 1998; Lamont and Yilmaz, 2002). 
Interestingly, although the overall mechanistic basis is similar in these cell 
systems, the signal transduction pathways activated by the organism and the 
intracellular locations and trafficking of the bacteria differ according to cell 
type. However, in all cases invasion is an active, bacterially driven process that 
has a significant impact on the phenotypic properties and fate of the host cell. 



Primary gingival epithelial cells 

The study of primary cultures of gingival epithelial cells (GEC) provided 
the first evidence for intracellular invasion by P. gingivalis in 1992 (Lamont 




et al., 1992). GEC are cultured from basal epithelial cells extracted from gin- 
gival explants, and they can be maintained in culture for several generations. 
Immunohistochemical staining has shown that the cells are nondifferenti- 
ated and noncornified, which are characteristics that are in common with 
the junctional epithelium. Thus, although not derived from the junctional 
epithelium, GEC demonstrate similar properties and thus provide a relevant 
ex vivo model for the events that occur at the base of the gingival crevice. 

P. gingivalis lacks the components and effectors of the type III secretion 
machinery that are important in the invasive processes of other organisms 
and that are discussed elsewhere in this volume. However, when in contact 
with GEC, P. gingivalis is induced to secrete a novel set of extracellular proteins 
(Park and Lamont, 1998). Some functional equivalence to type III effectors 
is implied by the finding that one of the secreted proteins of P. gingivalis 
is a homologue of a phosphoserine phosphatase (Lamont and Yilmaz, 2002), 5 

although the functionality of this molecule, either intracellularly or extra- & 

cellularly, remains to be established. Nevertheless, despite the absence of a z 

classical type III secretion apparatus, P. gingivalis invasion of GEC is swift ^ 

and profuse (Belton et al., 1999). a 

Fluorescent microscopic imaging has shown that invasion is complete 53 

within 15 min, and that all GEC exposed to P. gingivalis take up large num- 2 

bers of organisms: this is invasion on a scale unsurpassed by any of the % 

enteropathogens. Once inside the cells, the bacteria are not confined to a £ 

i-i 
membrane-bound vacuole and congregate in the perinuclear region (Belton § 

et al., 1999). P. gingivalis cells remain viable and capable of intracellular repli- g 

cation (Lamont et al., 1995). Interestingly, despite the burden of large num- & 

bers of intracellular bacteria, GEC do not undergo necrotic or apoptotic cell 
death (Nakhjiri et al., 2001); however, the cells contract and there is conden- 
sation of the actin cytoskeleton after prolonged cohabitation with P. gingivalis 
(Belton et al., 1999). 

The signaling events required for uptake into GEC are thought to ensue 
primarily from fimbriae-integrin interactions, although other signal trans- 
duction pathways may be operational (Yilmaz et al., 2002). Integrins have 
both a structural role, in linking extracellular matrix proteins with the cellular 
actin cytoskeleton in order to regulate cell shape and tissue architecture, and 
a signaling role, through intracellular signals generated by integrin-receptor 
coupling that regulate cell migration, gene expression, growth, survival, and 
inflammatory responses. Integrins are heterodimeric transmembrane recep- 
tors with no catalytic activity; therefore, the signals initiated by integrin- 
ligand interactions are transduced into cells through the activation of a num- 
ber of specialized cytoplasmic proteins. Enhanced tyrosine phosphorylation 





1." -J 

i - 

4 


r 


• 






1 
< » 

* 'V. 


■ V 


™ - 

•m 

/ i 

• 


• 


t r \ 


t* 




< 


f 

■ 




Figure 10.1. (facing page). Immunofluorescence microscopy (400x) of P. gingivalis 
invasion of GEC. P. gingivalis (green) induces recruitment of paxillin (red), A, and FAK 
(red), B, to the cell peripheries. GEC in C and D are stained with antibodies to paxillin or 
FAK, respectively, in the absence of P. gingivalis. See color section. 




of these proteins plays an essential role in the perpetuation of the intracel- 
lular signals created by integrins into specific targets in cells. Among those 
proteins, FAK (focal adhesion kinase) and paxillin have emerged as key signal- 
transducing components (Turner, 2000). Phosphorylation of paxillin is asso- 
ciated with coordinate formation of focal adhesions and stress fibers. Paxillin 
thereby provides a platform for efficient propagation of signals from one 
component to the next. Invasion by fimbriated P. gingivalis promotes a sig- 
nificant amount of tyrosine phosphorylation of paxillin in GEC during early 
infection (between 5 and 20 min; Yilmaz et al., 2002). 

Recent immunofluorescent studies (Yilmaz et al., 2003) have visualized 
the subcellular distribution of paxillin during P. gingivalis invasion (Fig. 10.1). 
Immediately after infection by P. gingivalis, paxillin aggregates at the plasma 
membrane and forms microspikes or lamellipodial-like extensions at the 
edges of cells. These results suggest that P. gingivalis recruits paxillin to the 5 

plasma membrane and promotes formation of focal adhesion complexes & 

that may potentiate efficient bacterial uptake in GEC. In addition to the early z 

redistribution of paxillin, after 24 h there is colocalization of paxillin with ^ 

P. gingivalis in the perinuclear space. Immunofluorescence also indicates that a 

FAK is activated and recruited to the membrane of GEC during early periods 53 

of P. gingivalis exposure (Fig. 10.1), and later it relocates to the perinuclear 2 

area with P. gingivalis. Paxillin and FAK phosphorylations, therefore, appear ^ 

to be important in both early and late events of P. gingivalis invasion and 
intracellular trafficking. 

Integrin signaling also modulates actin cytoskeleton and microtubule § 

dynamics. Sensitivity of the invasion process to the inhibitors cytochalasin & 

D and nocodazole provides indirect evidence that both actin microfilament 
and microtubule rearrangements are required for P. gingivalis entry into the 
host cell (Lamont et al., 1995). Indeed, immunofluorescent microscopy has 
shown that actin is assembled into filament-rich microspikes at the periphery 
of the GEC during P. gingivalis invasion. Later, both the actin microfilament 
and microtubules are dramatically depolymerized and nucleated. 

Signaling pathways downstream of integrin focal adhesions often funnel 
through the MAP kinase family of signal transduction mediators. Consistent 



> 

H 



with this, invasion by P. gingivalis results in activation of JNK, a stress - 
activated protein kinase of the MAP kinase family (Watanabe et al., 2001). 
Also consonant with integrin activation, P. gingivalis invasion is associated 
with a transient increase in the intracellular Ca 2+ concentration. The cal- 
cium ion increase results, at least in part, from release of calcium from 
a thapsigargin-sensitive intracellular store (Izutsu et al., 1996). However, 
P. gingivalis is capable of multiple independent interventions on GEC sig- 
nal transduction pathways as, in contrast to JNK, internalized P. gingivalis 
cause dephosphorylation of ERK1/2 (Watanabe et al., 2001). In addition, P. 
gingivalis does not induce activation and nuclear translocation of the eukary- 
otic transcriptional activator NF-/cB. Such an ability to selectively activate and 
suppress different components of related signaling pathways signifies a de- 
gree of versatility and sophistication of P. gingivalis in its dealings with GEC, 
g pointing toward a long evolutionary association between the two cell types. 




o 



§ The interactions between P. gingivalis and GEC are represented schematically 

* in Fig. 10.2. 



Q 
Pi 
< 
X 

3 KB (HeLa) cells 

Q 

Z 



Jj Another cell type used to study of P. gingivalis invasion is the KB cell. Long 

s thought to be derived from the oral epithelium, KB cells are now recognized 

£ to be in fact HeLa cells that contaminated the original cell culture. Invasion 

3 of these transformed epithelial cells by P. gingivalis is somewhat less efficient 

N 

° than GEC, with values less than 0.1% of the initial inoculum generally re- 

ported (Duncan et al., 1993; Sandros et al., 1994; Njoroge et al., 1997). This 
may be a consequence of the alterations in signal transduction pathways and 
surface protein expression that accompany transformation. Initial adherence 
to KB cells can be mediated by both cysteine proteases (Chen et al., 2001) 
and FimA (Njoroge et al., 1997). In contrast to GEC, engulfment of bacteria 
then occurs by classic receptor-mediated endocytosis (Sandros et al., 1996), 
and bacteria can be found both free in the cytoplasm and contained within 
membrane-bound vacuoles (Njoroge et al., 1997). Features in common with 
GEC include the accumulation of P. gingivalis cells in the perinuclear region 
(Houalet-Jeanne et al., 2001) and subsequent bacterial replication (Madianos 
etal., 1996). 



Endothelial cells 

Invasion of bovine and human heart and aortic endothelial cells by P. 
gingivalis has been established (Deshpande et al., 1998; Dorn et al., 1999). 



P. gingivalis 



Ca gated chamiel(s) 



CM- 




inlcgnn 



secreted pntfems 

■ *-6-j phosphatase 



9 



<....> 



stores 



MAP-kinase family r-" 







NM 




f 

Ca++ 




proteases 

phosphatase 



Disruption of nuclear transcription factor activity and 
modulation of gene expression (e.g., IL-8, Bcl-2) 

Figure 10.2. Model of currently understood P. gingivalis interactions with primary gingival 
epithelial cells. P. gingivalis cells bind through adhesins such as fimbriae to integrins on 
gingival cells. FAK and paxillin are recruited, and microtubules and microfilaments are 
rearranged to facilitate invagination of the membrane that results in the engulfment of 
bacterial cells. P. gingivalis rapidly locate in the perinuclear area where they replicate. 
Calcium ions are released from intracellular stores, which may regulate calcium-gated 
pores in the cytoplasmic membrane. Other signaling molecules such as the MAP-kinase 
family can be phosphorylated/dephosphorylated or degraded. Gene expression in the 
epithelial cells is ultimately affected. Abbreviations: Ca = calcium; CM = cytoplasmic 
membrane; IL-8 = interleukin-8; MF = actin microfilaments; MT = tubulin 

microtubules; NM = nuclear membrane; P = phosphate; ► is for a pathway with 

potential intermediate steps; >• is for translocation; -► is for release; «■- — > is for 

reversible association. 



< 
> 

on 
O 

z 

w 
*< 

o 

S3 

a 

o 

£ 
o 

o 

2 

t-H 

H 



Invasion requires remodeling of the microfilament and microtubule cyto- 
skeleton through protein phosphorylation-dependent signaling (Deshpande 
et al., 1998). Once inside the cells the bacteria are present in multimem- 
branous vacuoles that resemble autophagosom.es (Dorn et al., 2001). These 
vacuoles are positive for the early endosomal marker Rab5, and they rapidly 
acquire HsGsa7p, required for the formation of the autophagosome. The 
bacteria then traffic to a late autophagosome that contains both the rough 
endoplasmic reticulum protein BiP and the lysosomal protein LGP120, but 
that lacks cathepsin L and late endosomal markers (Dorn et al., 2001). In en- 
dothelial cells, therefore, P. gingivalis evades the endocytic pathway leading to 




O 



Q 
Pi 
< 
X 
u 
i— i 

Pi 
Q 

< 

N 
< 

1-1 

I— I 

w 
i-i 
N 
:Q 




Figure 10.3. Transmission electron microcopy (12,500x) of P. gingivalis captured by a 
dendritic cell and located within multivesiculated compartments (arrowheads). (Image 
provided by C. W. Cutler.) 

lysosomes, and instead it traffics to the autophagosome, where development 
of this organelle is impeded. Invasion of endothelial cells by P. gingivalis could 
allow access to cells of vascular walls, where the induction of autophagocytic 
pathways may alter the properties of the cells. The persistence of P. gingivalis 
could then exacerbate the immune response along the vasculature. Invasion 
of vascular endothelial cells may also provide a portal for bacterial entry into 
the bloodstream with subsequent systemic spread. 



Dendritic cells 

Dendritic cells are antigen-presenting cells that can activate lymphocytes, 
including, distinctively naive T cells (Banchereau and Steinman, 1998). Den- 
dritic cells increase in number at sites of diseased oral epithelium that contain 
intragingival bacteria. P. gingivalis can internalize within cultured dendritic 
cells (Fig. 10.3), and invasion is associated with sensitization and activation 
of the dendritic cells (Cutler et al., 1999). This process has parallels with 



contact hypersensitivity responses, and hence such reactions may play a role 
in periodontal diseases. The mechanistic basis for invasion in these cells has 
yet to be investigated. 




CONSEQUENCES OF P. GINGIVALIS INVASION 
Epithelial cells 

The presence of large numbers of intracellular P. gingivalis would, at first 
glance, appear to represent an insult that could be terminal for the epithe- 
lial cells. Indeed, after several hours of infection, GEC begin to shrink and 
round up, features indicative of apoptotic cell death. However, the cells do 
not detach from the substratum and remain capable of trypan blue exclusion, 
calcein hydrolysis, and maintenance of physiologic intracellular calcium ion 5 

concentrations (Bel ton et al., 1999). An examination of molecular markers & 

o 
of apoptosis showed that at early time points of P. gingivalis invasion there is 2 

td 

an increase in proapoptotic molecules such as Bax. However, after extended ^ 

incubation the Bax levels decline, and there is an increase in the expres- a 

sion of the antiapoptotic molecule Bcl-2 (Nakhjiri et al., 2001). Furthermore, 2j 

P. gingivalis is even capable of blocking apoptosis induced by the human § 

topoisomerase I inhibitor, camptothecin. These results can be interpreted in % 

terms in which the initial response of GEC to the P. gingivalis onslaught is one £ 

1-1 

of activation of programmed cell death. P. gingivalis, however, having located § 

in a nutritionally rich, immune-privileged site, blocks the cell death pathways g 

in order to maintain its intracellular lifestyle. Furthermore, the proteases of & 

P. gingivalis may protect the organism from /3-defensins, which are intra- 
cellular antimicrobial peptides produced by epithelial cells (Devine et al., 
1999). Unconstrained replication of P. gingivalis may be prevented, how- 
ever, by calprotectin, an S100 calcium-binding protein with broad-spectrum 
antimicrobial activity produced inside epithelial cells (Nisapakultorn et al., 
2001). 

The interaction between P. gingivalis and GEC is more than a life or death 
struggle for supremacy and can result in more subtle phenotypic changes in 
the host. Transcription and secretion of interleukin (IL)-8 (a potent neutrophil 
chemokine) by GEC is inhibited following P. gingivalis invasion. Moreover, 
P. gingivalis can antagonize IL-8 secretion following stimulation of epithelial 
cells by common plaque commensals (Darveau et al., 1998). Reduced ex- 
pression of epithelial cell intercellular adhesion molecule (ICAM)-l may also 
contribute to downregulation of the innate host response (Madianos et al., 
1997). Regulation of matrix metalloproteinase (MMP) production by gingival 



epithelial cells is disrupted following contact with P. gingivalis (Fravalo et al., 
1996), thus interfering with extracellular matrix repair and reorganization. 

Endothelial cells 

Although the ultimate metabolic fate of endothelial cells and their in- 
ternalized organisms remains to be determined, it is becoming apparent 
that many important properties of endothelial cells are modulated by P. gin- 
givalis infection. Invasion by P. gingivalis leads to upregulation of ICAM-1 
and vascular cell adhesion molecule (VCAM)-l, along with P- and E-selectins 
(Khlgatian et al., 2002). The effect appears to be mediated through the major 
fimbriae, as it can be blocked with fimbrial antibodies and does not occur with 
a fimbriae-deficient mutant. Increased expression of these molecules can be 
£ predicted to elevate the levels of leukocytes recruited to sites of P. gingivalis 

^ infection. Levels of IL-8 and MCP-1 are also modulated in response to P. 

ha gingivalis. In these cases, however, whole cells of P. gingivalis abolish normal 

Q 

pj IL-8 and MCP-1 responses, whereas isolated outer membrane components 

u and fimbrillin can stimulate production (Nassar et al., 2002). The inhibitory 

Q effect of whole cells is not invasion dependent, as it occurs in the absence 

< of detectable internalized P. gingivalis. Indeed, under certain conditions, en- 

N 

s dothelial cells can be induced to secrete MCP-1 by internal P. gingivalis (Kang 

i-i 

£ and Kuramitsu, 2002). Thus endothelial cell responses to P. gingivalis are 

w complex and multithreaded, possibly a means for facilitating the long-term 

: o association of the organism with the host cell. 




H 



RELEVANCE TO HEALTH AND DISEASE 
Periodontal diseases 

The pathogenesis of periodontal diseases involves multiple bacteria with 
a range of virulence factors interfacing with a variety of host cells and im- 
mune effector molecules. Within this framework, P. gingivalis invasion could 
play a number of important roles in the disease process. An intracellular 
location will shelter the bacteria from the ravages of the immune system, 
and it may allow the organisms to increase in number to exceed a threshold 
required to initiate disease. The phenotypic changes in epithelial cells that are 
related to P. gingivalis invasion also have the potential to impinge on several 
aspects of pathogenesis (Lamont and Yilmaz, 2002). 

The gene encoding the neutrophil chemokine IL-8 is transcriptionally 
downregulated by P. gingivalis even in the presence of otherwise stimulatory 




organisms such as Fusobacterium nucleatum (Darveau et al., 1998; Huang 
et al., 1998, 2001). In clinically healthy tissue, IL-8 forms a concentration 
gradient that increases from the gingival tissue toward the surface (Tonetti et 
al., 1994) and will thus direct neutrophils to sites of bacterial accumulation. 
Hence, low level expression of IL-8 is considered to be important in ensuring 
gingival health by controlling bacteria and preventing neutrophil-mediated 
damage in the tissues. Inhibition of IL-8 accumulation by P. gingivalis at 
sites of bacterial invasion could impede innate host defense at the bacteria- 
epithelia interface, as the host would no longer be able to detect the presence 
of bacteria and direct neutrophils for their removal. The ensuing overgrowth 
of bacteria would then contribute to a burst of disease activity. 

Nonetheless, host polymorphonuclear neutrophils and other defense 
mechanisms do eventually become mobilized, as evidenced by the inflamma- 
tory nature of P. gingivalis-as sociated periodontal diseases. The overgrowth of 5 
subgingival plaque bacteria, or of P. gingivalis itself, that ensues after initial & 
immune suppression may therefore reactivate the immune response. Indeed, z 
epithelial cells can be induced to secrete IL-8 by P. gingivalis, depending on ^ 
the conditions of stimulation (Huang et al., 2001). Moreover, the encounter a 
with different host cells as the infection progresses may also stimulate the im- 53 
mune response. For example, P. gingivalis invasion of dendritic cells results 2 
in maturation, increased costimulatory molecule expression, and stimula- % 

tory activity for T cells (Cutler et al., 1999) . The migration and proliferation of £ 

i-i 

P. gingivalis-speciFic effector T cells could be one means by which the im- j* 

mune system gears up in periodontal disease (Saglie et al., 1987; Cutler et al., $ 

1999). 

In addition to effects on immune modulators, invasion by P. gingivalis can 
impinge on the activity of MMP enzymes. MMPs are members of a family of 
zinc-dependent endopeptidases with a broad spectrum of proteolytic activity 
that collectively can degrade all of the components of the extracellular matrix 
(DeCarlo et al., 1998). MMPs are required for epithelial cell migration and to 
detach the cells from the underlying matrix, and they are also involved in tis- 
sue remodeling by facilitating the removal of damaged tissue. These activities 
are important in maintenance of the integrity of the epithelial layer. How- 
ever, destruction of the extracellular matrix, which is a feature of periodontal 
lesions, may be caused by elevated MMP activity (Tonetti et al., 1994; DeCarlo 
et al., 1997, 1998), and a higher percentage of tissues from periodontitis sites 
have mRNA for MMPs as compared with tissue from healthy sites (Tonetti 
et al., 1994). Control of MMP activity is, therefore, important in sustaining 
gingival health. Invasion of P. gingivalis disrupts the expression of MMPs by 
gingival epithelial cells (Fravalo et al., 1996), which could contribute both to 



H 




tissue destruction and to failure to repair a periodontal lesion. These activities 
are distinct from, but probably complementary to, the direct action of pro- 
teolytic enzymes that will be delivered in close proximity to their substrates 
during the adhesion and entry process. Indeed, P. gingivalis proteinases can 
activate and upregulate the transcription of MMP enzymes (DeCarlo et al., 
1997, 1998). Furthermore, P. gingivalis proteinases can degrade IL-8 and other 
cytokines along with occludin, cadherins, catenins, and integrins (Fletcher 
et al., 1997; Darveau et al., 1998; Yun et al., 1999; Zhang et al, 1999; Katz 
et al., 2000), which are proteins that are important in maintaining the barrier 
function of the epithelium. 

Systemic diseases 

g Although tissue destruction in periodontal diseases is limited to the sup- 

§ porting structures of the teeth, epidemiological evidence is emerging for an 

^ association between periodontal infections and serious systemic diseases, 

§ including coronary artery disease (Scannapieco and Genco, 1999). Several 

= observations provide a credible, though still preliminary, basis for a causal 

* link between infections with periodontal organisms such as P. gingivalis 

% and heart disease. P. gingivalis has been detected in carotid and coronary 

< atheromas (Chiu, 1999; Haraszthy et al., 2000), and the organism can induce 

~ platelet aggregation, which is associated with thrombus formation ( Herzberg 

§ et al., 1994). Furthermore, P. gingivalis infection accelerates the progression 

.g of atherosclerosis in a heterozygous apolipoprotein E -deficient murine model 

(Li et al., 2002). Although common dental procedures, even vigorous tooth 
brushing, can lead to the presence of oral bacteria in the bloodstream, it is 
also possible that tissue and cell invasion by P. gingivalis in the highly vascu- 
larized gingiva may be a means by which these bacteria can gain access to the 
circulating blood and establish infections at remote sites. Once located at sites 
such as the heart vessel walls, the invasion of endothelial cells (Deshpande 
et al., 1998; Dorn et al., 1999) could constitute a chronic insult to arte- 
rial walls that may increase susceptibility to tissue damage. In addition, 
modulation of factors that regulate recruitment of leukocytes (ICAM-1, 
VCAM-1, MCP-1, and selectins) could enhance arthrosclerosis progression 
(Kuramitsu et al., 2001; Khlgatian et al., 2002; Nassar et al., 2002). For exam- 
ple, monocytes migrating through the endothelial layer into the subendothe- 
lial spaces may produce foam cells (Kang and Kuramitsu, 2002), a feature of 
both early and late artherosclerotic lesions. P. gingivalis can also directly in- 
duce foam cell production in a murine macrophage cell line (Kuramitsu et al., 
2001). 




It is likely that we are only beginning to uncover the full range of con- 
sequences of the interactions between invasive oral bacterial and host cells. 
It is also important to consider that P. gingivalis is present in the mouths 
of healthy individuals and can be found in high numbers inside epithelial 
cells (Rudney et al., 2001). An intracellular location may initially serve as an 
integral part of the interactions that maintain a balance between host and 
pathogen. Disruption of this balance, either by the action of the internal bac- 
teria or by other pathogen or host factors, may be required to activate the 
disease process. 

REFERENCES 

Banchereau, J. and Steinman, R.M. (1998). Dendritic cells and the control of 

immunity. Nature 392, 245-252. 2 

Barkocy-Gallagher, G.A., Han, N., Patti, J.M., Whitlock, J., Progulske-Fox, A., and ^ 

o 
Lantz, M.S. (1996). Analysis of the prtP gene encoding porphypain, a cysteine s 

td 

proteinase of Porphyromonas gingivalis. J. Bacteriol. 178, 2734-2741. * 

Belton, CM., Izutsu, K.T., Goodwin, P.C., Park, Y., and Lamont, R.J. (1999). Fluo- 8 

rescence image analysis of the association between Porphyromonas gingivalis 2j 

and gingival epithelial cells. Cell. Microbiol. 1, 215-224. 2 

Chen, T., Nakayama, K., Belliveau, L., and Duncan, M.J. (2001). Porphyromonas % 

gingivalis gingipains and adhesion to epithelial cells. Infect. Immun. 69, 3048- £ 

3056. § 

Chiu, B. (1999). Multiple infections in carotid atherosclerotic plaques. Am. Heart J 

J. 138, S534-536. § 

Curtis, M.A., Kuramitsu, H.K., Lantz, M., Macrina, F.L., Nakayama, K., Potempa, 
J., Reynolds, E.C., and Aduse-Opoku, J. (1999). Molecular genetics and 
nomenclature of proteases of Porphyromonas gingivalis. J. Periodont. Res. 34, 
464-472. 

Cutler, C.W., Jotwani, R., Palucka, K.A., Davoust, J., Bell, D., and Banchereau, 
J. (1999). Evidence and a novel hypothesis for the role of dendritic cells and 
Porphyromonas gingivalis in adult periodontitis. J. Periodont. Res. 34, 406-412. 

Darveau, R.P., Belton, CM., Reife, R.A., and Lamont, R.J. (1998). Local chemokine 
paralysis: a novel pathogenic mechanism of Porphyromonas gingivalis. Infect. 
Immun. 66, 1660-1665. 

DeCarlo, A.A., Grenett, H.E., Harber, G.J., Windsor, L.J., Bodden, M.K., Birkedal- 
Hansen, B., and Birkedal-Hansen, H. (1998). Induction of matrix metallo- 
proteinases and a collagen-degrading phenotype in fibroblasts and epithelial 
cells by secreted Porphyromonas gingivalis proteinase. J. Periodont. Res. 33, 
408-420. 



DeCarlo, A.A., Windsor, L.J., Bodden, M.K., Harber, G.J., Birkedal-H arisen, B., 
and Birkedal-H ansen, H. (1997). Activation and novel processing of matrix 
metalloproteinases by thiol-proteinase from the oral anaerobe Porphyromonas 
gingivalis. J. Dent. Res. 76, 1260-1270. 

Deshpande, R.G., Khan, M.B., and Genco, C.A. (1998). Invasion of aortic and heart 
endothelial cells by Porphyromonas gingivalis. Infect. Immun. 66, 5337-5343. 

Devine, D.A., Marsh, P.D., Percival, R.S., Rangarajan, M., and Curtis, M.A. (1999). 
Modulation of antibacterial peptide activity by products of Porphyromonas 
gingivalis and Prevotella spp. Microbiology 145, 965-971. 

Dorn, B.R., Burks, J.N., Seifert, K.N., and Progulske-Fox, A. (2000). Invasion of 
endothelial and epithelial cells by strains of Porphyromonas gingivalis. FEMS 
Microbiol. Lett. 187, 139-144. 

Dorn, B.R., Dunn, W.A. Jr., and Progulske-Fox, A. (1999). Invasion of human 
g coronary artery cells by periodontal pathogens. Infect. Immun. 67, 5792-5798. 




o 



2 Dorn, B.R., Dunn, W.A. Jr., and Progulske-Fox, A. (2001). Porphyromonas gingi- 

< 

valis traffics to autophagosomes in human coronary artery endothelial cells. 

g Infect. Immun. 69, 5698-5708. 

< 

g Duncan, M.J., Nakao, S., Skobe, Z., and Xie, H. (1993). Interactions of Porphy- 

rt romonas gingivalis with epithelial cells. Infect. Immun. 61, 2260-2265. 

5j Fletcher, J., Reddi, K., Poole, S., Nair, S., Henderson, B., Tabona, P., and 

< Wilson, M. (1997). Interactions between periodontopathogenic bacteria and 

g cytokines. J. Periodont. Res. 32, 200-205. 

§ Fravalo, P., Menard, C, and Bonnaure- Mallet, M. (1996). Effect of Porphyromonas 

.g gingivalis on epithelial cell MMP-9 type IV collagenase production. Infect. 

Immun. 64, 4940-4945. 
Han, N., Whitlock, J., and Progulske-Fox, A. (1996). The hemagglutinin gene A 
(hagA) of Porphyromonas gingivalis 381 contains four large, contiguous, direct 
repeats. Infect. Immun. 64, 4000-4007. 
Haraszthy, V.I., Zambon, J. J., Trevisan, M., Zeid, M., and Genco, R.J. (2000). 
Identification of periodontal pathogens in atheromatous plaques. J. Peri- 
odontol. 71, 1554-1560. 
Herzberg, M.C., MacFarlane, G.D., Liu, P., and Erickson, P.R. (1994). The platelet 
as an inflammatory cell in periodontal diseases: the interactions with Por- 
phyromonas gingivalis. In Molecular Pathogenesis of Periodontal Disease, ed. R. 
Genco, S. Hamada, T. Lehner, J. McGhee, and S. Mergenhagen, pp. 247-256. 
Washington, DC: ASM Press. 
Houalet- Jeanne, S., Pellen-Mussi, P., Tricot-Doleux, S., Apiou, J., and Bonnaure- 
Mallet, M. (2001). Assessment of internalization and viability of Porphyro- 
monas gingivalis in KB epithelial cells by confocal microscopy. Infect. Immun. 
69, 7146-7151. 




Huang, G.T., Haake, S.K., Kim, J.W., and Park, N. (1998). Differential expression 
of interleukin-8 and intercellular adhesion molecule-1 by human gingival 
epithelial cells in response to Actinohacillus actinomycetemcomitans or Por- 
phyromonas gingivalis infection. Oral Microbiol. Immunol. 13, 301-309. 

Huang, G.T., Kim, D., Lee, J.K., Kuramitsu, H.K., and Haake, S.K. (2001). 
Interleukin-8 and intercellular adhesion molecule 1 regulation in oral epithe- 
lial cells by selected periodontal bacteria: multiple effects of Porphyromonas 
gingivalis via antagonistic mechanisms. Infect. Immun. 69, 1364-1372. 

Izutsu, K.T., Belton, CM., Chan, A., Fatherazi, S., Kanter, J. P., Park, Y., and 
Lamont, R.J. (1996). Involvement of calcium in interactions between gingival 
epithelial cells and Porphyromonas gingivalis. FEMS Microbiol. Lett. 144, 145- 
150. 

Kang, I.C. and Kuramitsu, H.K. (2002). Induction of monocyte chemoattractant 

protein-1 by Porphyromonas gingivalis in human endothelial cells. FEMS Im- 5 

munol. Med. Microbiol. 34, 311-317. s 

Katz, J., Sambandam, V., Wu, J.H., Michalek, S.M., and Balkovetz, D.F. (2000). * 

Characterization of Porphyromonas gmgiva/is-induced degradation of epithe- * 

lial cell junctional complexes. Infect. Immun. 68, 1441-1449. a 

Khlgatian, M., Nassar, H., Chou, H.H., Gibson, F.C., and Genco, C.A. (2002). 23 

Fimbria-dependent activation of cell adhesion molecule expression in Por- § 

phyromonas gingivalis-inf ected endothelial cells. Infect. Immun. 70, 257-267. i? 

Kontani, M., Kimura, S., Nakagawa, I., and Hamada, S. (1997). Adherence of ^ 

Porphyromonas gingivalis to matrix proteins via a fimbrial cryptic receptor ^ 

exposed by its own arginine-specific protease. Mol. Microbiol. 24, 1179-1 187. 5 

Kontani, M., Ono, H., Shibata, H., Okamura, Y., Tanaka, T., Fujiwara, T., Kimura, & 

S., and Hamada, S. (1996). Cysteine protease of Porphyromonas gingivalis 
381 enhances binding of fimbriae to cultured human fibroblasts and matrix 
proteins. Infect. Immun. 64, 756-762. 

Kuramitsu, H.K. (1998). Proteases of Porphyromonas gingivalis: what don't they 
do? Oral Microbiol. Immunol. 13, 263-270. 

Kuramitsu, H.K., Qi, M., Kang, I.C, and Chen, W. (2001). Role for periodontal 
bacteria in cardiovascular diseases. Ann. Periodontol. 6, 41-47. 

Lamont, R.J., Chan, A., Belton, CM., Izutsu, K.T., Vasel, D., and Weinberg, A. 
(1995). Porphyromonas gingivalis invasion of gingival epithelial cells. Infect. 
Immun. 63, 3878-3885. 

Lamont, R.J. and Jenkinson, H.F. (1998). Life below the gum line: pathogenic 
mechanisms of Porphyromonas gingivalis. Microbiol. Mol. Biol. Rev. 62, 1244- 
1263. 

Lamont, R.J. and Jenkinson, H.F. (2000). Subgingival colonization by Porphy- 
romonas gingivalis. Oral Microbiol. Immunol. 15, 341-349. 




Lamont, R.J. and Yilmaz, O. (2002). In or out: the invasiveness of oral bacteria. 
Periodontology 2000 30, 61-69. 

Lamont, R.J., Oda, D., Persson, R.E., and Persson, G.R. (1992). Interaction of 
Porphyromonas gingivalis with gingival epithelial cells maintained in culture. 
Oral Microbiol. Immunol. 7, 364—367. 

Li, L., Messas, E., Batista, E.L. Jr., Levine, R.A., and Amar, S. (2002). Porphy- 
romonas gingivalis infection accelerates the progression of atherosclerosis in 
a heterozygous apolipoprotein E -deficient murine model. Circulation 105, 
861-867. 

Madianos, P.N., Papapanou, P.N., Nannmark, U., Dahlen, G., and Sandros, J. 
(1996). Porphyromonas gingivalis FDC381 multiplies and persists within hu- 
man oral epithelial cells in vitro. Infect. Immun. 64, 660-664. 

Madianos, P.N., Papapanou, P.N., and Sandros, J. (1997). Porphyromonas gingiva- 

g lis infection of oral epithelium inhibits neutrophil transepithelial migration. 

§ Infect. Immun. 65, 3983-3990. 

. Nakagawa, I., Amano, A., Kuboniwa, M., Nakamura, T., Kawabata, S., and 

§ Hamada, S. (2002). Functional differences among FimA variants of Porphy- 

< 

g romonas gingivalis and their effects on adhesion to and invasion of human 

rt epithelial cells. Infect. Immun. 70, 277-285. 

5j Nakamura, T., Amano, A., Nakagawa, I., and Hamada, S. (1999). Specific inter- 

< actions between Porphyromonas gingivalis fimbriae and human extracellular 

£ matrix proteins. FEMS Microbiol. Lett. 175, 267-272. 

§ Nakayama, K., Yoshimura, F., Kadowaki, T., and Yamamoto, K. (1996). Involve- 

.g ment of arginine-specific cysteine proteinase (Arg-gingipain) in fimbriation 

of Porphyromonas gingivalis. J. Bacteriol. 178, 2818-2824. 
Nakhjiri, S.F., Park, Y., Yilmaz, O., Chung, W.O., Watanabe, K., El-Sabaeny, A., 
Park, K., and Lamont, R.J. (2001). Inhibition of epithelial cell apoptosis by 
Porphyromonas gingivalis. FEMS Microbiol. Lett. 200, 145-149. 
Nassar H, Chou, H.H., Khlgatian, M., Gibson, F.C., Van Dyke, T.E., and Genco, 
C.A. (2002). Role for fimbriae and lysine-specific cysteine proteinase gingi- 
pain K in expression of interleukin-8 and monocyte chemoattractant protein 
in Porphyromonas gingivalis-infected endothelial cells. Infect. Immun. 70, 268- 
276. 
Nisapakultorn, K., Ross, K.F., and Herzberg, M.C. (2001). Calprotectin expres- 
sion in vitro by oral epithelial cells confers resistance to infection by Porphy- 
romonas gingivalis. Infect. Immun. 69, 4242-4247. 
Njoroge, T., Genco, R.J., So jar, H.T., and Genco, C.A. (1997). A role for fimbriae 
in Porphyromanas gingivalis invasion of oral epithelial cells. Infect. Immun. 
65, 1980-1984. 
Ogawa, T., Ogo, H., Uchida, H., and Hamada, S. (1994). Humoral and cellular 



immune responses to the fimbriae of Porphyromonas gingivalis and their 

synthetic peptides. J. Med. Microbiol. 40, 397-402. 
Park, Y. and Lamont, R.J. (1998). Contact-dependent protein secretion in Porphy- 
romonas gingivalis. Infect. Immun. 66, 4777-4782. 
Potempa, J., Pavloff, N., and Travis, J. (1995). Porphyromonas gingivalis: a pro- 

teinase/gene accounting audit. Trends Microbiol. 3, 430-434. 
Progulske-Fox, A., Tumwasorn, S., Lepine, G., Whitlock, J., Savett, D., Ferretti, 

J. J., and Banas, J.A. (1995). The cloning, expression and sequence analysis 

of second Porphyromonas gingivalis gene that encodes for a protein involved 

in hemagglutination. Oral Microbiol. Immunol. 10, 311-318. 
Rudney, J.D., Chen, R., and Sedge wick, G.J. (2001). Intracellular Actinobacillus 

actinomycetemcomitans and Porphyromonas gingivalis in buccal epithelial cells 

collected from human subjects. Infect. Immun. 69, 2700-2707 '. 
Saglie, F.R., Pertuiset, J.H., Smith, C.T., Nestor, M.G., Carranza, F.A., Newman, 5 

M.G., Rezende, M.T., and Nisengard, R. (1987). The presence of bacteria in % 




o 



the oral epithelium in periodontal disease. III. Correlation with Langerhans 2 






cells. /. Periodontol. 58, 417-422. 
Sandros, J., Madianos, P.N., and Papapanou, P.N. (1996). Cellular events concur- m 



a 



> 

H 



rent with Porphyromonas gingivalis invasion of oral epithelium in vitro. Eur. 2j 

J. Oral Sci. 104, 363-371. g 

Sandros, J., Papapanou, P.N., Nannmark, U., and Dahlen, G. (1994). Porphy- £ 

romonas gingivalis invades human pocket epithelium in vitro. J. Periodont. 
Res. 29, 62-69. 

Scannapieco, F.A. and Genco, R.J. (1999). Association of periodontal infections jg 

with atherosclerotic and pulmonary diseases. J. Periodont. Res. 34, 340- & 

345. 

Shibita, Y., Hayakawa, M., Takiguchi, H., Shiroza, T., and Abiko, Y. (1999). Deter- 
mination and characterization of the hemagglutinin-associated short motifs 
found in Porphyromonas gingivalis multiple gene products. J. Biol. Chem. 274, 
5012-5020. 

Sojar, H.T., Han, Y., Hamada, N., Sharma, A., and Genco, R.J. (1999). Role of the 
amino -terminal region of Porphyromonas gingivalis fimbriae in adherence to 
epithelial cells. Infect. Immun. 67, 6173-6176. 

Sojar, H.T., Lee, J.-Y., and Genco, R.J. (1995). Fibronectin binding domain of 
P. gingivalis fimbriae. Biochem. Biophys. Res. Commun. 216, 785-792. 

Sojar, H.T., Sharma, A., and Genco, R.J. (2002). Porphyromonas gingivalis fimbriae 
bind to cytokeratin of epithelial cells. Infect. Immun. 70, 96-101. 

Tokuda, M., Duncan, ML, Cho, M.I., and Kuramitsu, H.K. (1996). Role of Por- 
phyromonas gingivalis protease activity in colonization of oral surfaces. Infect. 
Immun. 64, 4067-4073. 



Tonetti, M.S., Imboden, M.A., Gerber, L., Lang, N.P., Laissue, J., and Mueller, 
C. (1994). Localized expression of mRNA for phagocyte -specific chemotactic 
cytokines in human periodontal infections. Infect. Immun. 62, 4005-4014. 

Turner, C.E. (2000). Paxillin and focal adhesion signaling. Nat. Cell Biol. 2, E231- 
E236. 

Wang, T., Zhang, Y., Chen, W., Park, Y., Lamont, R.J., and Hackett, M. (2002). 
Reconstructed protein arrays from 3D HPLC/tandem mass spectrometry 
and 2D gels: complementary approaches to Porphyromonas gingivalis protein 
expression. The Analyst 127 , 1450-1456. 

Watanabe, K., Onoe, T., Ozeki, M., Shimizu, Y., Sakayori, T., Nakamura, H., and 
Yoshimura, F. (1996). Sequence and product analyses of the four genes down- 
stream from the fimbrillin gene (fimA) of the oral anaerobe Porphyromonas 
gingivalis. Microbiol. Immunol. 40, 725-734. 
g Watanabe, K., Yilmaz, O., Nakhjiri, S.F., Belton, CM., and Lamont, R.J. (2001). 




o 



2 Association of mitogen-activated protein kinase pathways with gingival ep- 

ithelial cell responses to Porphyromonas gingivalis infection. Infect. Immun. 



g 69, 6731-6737. 

< 

- Weinberg, A., Belton, C.A., Park, Y., and Lamont, R.J. (1997). Role of fimbriae in 

rt Porphyromonas gingivalis invasion of gingival epithelial cells. Infect. Immun. 

* 65, 313-316. 

< Xie, H., Cai, S., and Lamont, R.J. (1997). Environmental regulation of fimbrial 

~ gene expression in Porphyromonas gingivalis. Infect. Immun. 65, 2265-2271. 

g Xie, H., Chung, W., Park, Y., and Lamont, R.J. (2000). Regulation of the Porphy- 

.g romonas gingivalis fimA (fimbrillin) gene. Infect Immun. 68, 6574-6579. 

Xie, H. and Lamont, R.J. (1999). Promoter architecture of the Porphyromonas 

gingivalis fimbrillin gene. Infect. Immun. 67, 3227-3235. 
Yilmaz, O., Watanabe, K., and Lamont, R.J. (2002). Involvement of integrins in 
fimbriae -mediated binding and invasion by Porphyromonas gingivalis. Cell. 
Microbiol. 4, 305-314. 
Yilmaz, O., Young, P. A., Lamont, R.J., and Kenny, G.E. (2003). Gingival epithelial 
cell signaling and cytoskeletal responses to Porphyromonas gingivalis invasion. 
Microbiology 149, 2417-2426. 
Yun, P.L., DeCarlo, A.A., and Hunter, N. (1999). Modulation of major histo- 
compatibility complex protein expression by human gamma interferon me- 
diated by cysteine proteinase-adhesin polyproteins of Porphyromonas gingi- 
valis. Infect. Immun. 67, 2986-2995. 
Zhang, J., Dong, H., Kashket, S., and Duncan, M.J. (1999). IL-8 degradation by 
Porphyromonas gingivalis proteases. Microb. Pathog. 26, 275-280. 



Index 




Acanthamoeba castellanii, 135, 142-143, 145 
Acanthamoeba polyphaga, 126, 135, 145-146 
Actinobacillus actinomycetemcomitans, 275-285 

actinobacillin, 278 

adhesion, 275-278 

adhesion, Aae adhesin and, 279, 285 

adhesion, conveyed, 278 

adhesion, specific to epithelial cells, 279 

autotransporter genes, 279 

bone resorption activity, 278 

calcium, 279, 285 

CD14, 286 

cell-cell spread, 283 

colony phase variation, 275 

egression, from host cell, 281 

endocytic pathway, classical, 282 

endotoxin, 278 

epithelial cells (micrograph), 280 

extracellular amorphous material 
(ExAmMat), 278 

fimbriae, 276-277, 279 

IL-8 and, 286 

integrins, 282 

internalization process, 278, 281-283 

intracellular-location mechanisms, 283 

invasion, 283-285 

invasion gene homologs, E. coli, 284 

invasion, host response to, 286 

invasion/pathogenicity correlation, 287 

lactoferrin, 279 

leukotoxin, 278, 282, 285 



Listeria, 283 

motor proteins/kinesin-like entity, 

283-285 
nitric oxide (NO), 287 
periodontitis, episodic nature of, 276, 287 
phenotype /variant shifts, 287 
phenotypes (micrographs of), 276 
phospholipase C (PLC), 282, 285 
RGD, 283-284 

rough colony phenotypes, 275-276 
serotype-specific polysaccharide antigen 

(SPA), 286 
Shigella, 283 

signaling pathways of host cell, 279-281 
smooth colony phenotypes, 275-276 
staurosporin, 281 
surface-associated molecules /organelles, 

276 
transferrin receptor, 282, 285 
type-IV pilin, 277 
vesicles, 277-278 
ampicillin, 2 
apoptosis 

bundle-forming pili (BFP), 104 

cancer, xii 

dendritic cells and, 163 

dot/icm genes, 125-126 

Escherichia coli, 103-105 

gingival epithelial cells (GEC), 299 

Group A streptococci, 244, 246-247, 259, 

261 




X 
w 

Q 



apoptosis (cont.) 

Legionella pneumophila, 144-145, 
149-150 

Legionella pneumophila (figure), 140 

Listeria monocytogenes, 163 

macrophages, 74-75, 125-126 

necrosis, distinguished from, 146 

Neisseria gonorrhoeae, 211 

porin, 211 

Porphyromonas gingivalis, 305 

REPEC, 104 

Shigella, 25-26 

Yersinia, inhibited by, 72-73 

Yersinia outer proteins (Yop), 74-75 
arp2/3 complex, 43, 45 

Escherichia coli, 95, 97-98 

Listeria monocytogenes, 178-179 

profilin, 48 

Shigella, 38, 45-47 

See also N-WASP; VirG-N-WASP 
complex 
arthritis-dermatitis syndrome, 205 
asialoglycoprotein receptor (ASGP-R), 213, 

222, 227-230 
attaching/effacing (A/E) lesions, 90-91, 
93-94, 99, 102-103 

bacteria 

minimizing damage to host, xi-xii 

number on human body, xi 
bacteroides-epithelium development 

interactions, xii 
blood-brain barrier, 163-164 
Boswell, 203 

bubonic plague, 59, 65. See also Y. pestis 
bundle-forming pili (BFP) 

apoptosis, 104 

Escherichia coli, 88-91, 93, 104 

Clq, 170 

C3, 170, 218-220, 223-224, 231-233 

C4, 219-220 

C5 through C8, 220 

C9, 219-220 

calcium 

Actinohacillus actinomycetemcomitans, 279, 
285 



attaching /effacing (A/E) lesions, 102-103 

Escherichia coli, 102-103 

neutrophils, 72 

Porphyromonas gingivalis, 302, 305 

YopH, 72 
calprotectin, 305 
cancer, xii 
CD4, 147-149 
CD8, 147-149 
CD14, 286 
CD18, 17, 220, 232 
CD44, 33-35, 241, 243, 247 
CD46, 209, 220, 241, 243, 252 
CD62, 255 
CD66, 220 
CD111, 255 
Cdc42, 43 

EPEC, 105-106 

GTPase activating proteins, 14 

IpaC, 37, 38 

N-WASP, 46-47 

Salmonella, 11-14 

Shigella, 37-38, 46-47 

sopB, 14 

sopE, 13-14 

sopE/E2, 14-15 

SptP, 14-15 

VirG, 46-47 

Yersinia, 67-70 
Cdkl, 99 

cephalosporins, 2 
chlorine, 127-128 
complementation 

complement receptors, 170 

See also specific components, receptors 
conjunctivitis, 164 
CR1, 211 
CR3, 170, 209, 211, 219-220, 223-224, 231, 

233 
CR4, 220 
cystic fibrosis transmembrane regulator 

(CFTR), 102 
cytoskeletal structures. See specific molecules, 
species 

defect in organelle trafficking (dot) genes. See 
dot/icm genes 



dendritic cells 

apoptosis of, 163 

CNS infection, 163-164 

Listeria monocytogenes, 163-164 

Porphyromonas gingivalis, 296, 304-305, 
307 

Salmonella, 17 

sampling of bacteria, 17 
Dictyostelium discoideum, 135 
dot/km genes, 125, 131-132, 135-139, 
143-144, 150 

DNA regulatory elements of, 143-144 

macrophages and, 125-126 

macropinocytosis, 125-126 

See also type IV secretion systems 

E-cadherin, 166-167 
encephalitis, 164 
endocarditis, 164, 205 
enterohemorrhagic E. coli (EH EC), 87 

adherence mechanisms of, 88-90 

EHEC factor for adherence (efal), 90 

hemolytic uremic syndrome (HUS), 88 

initial attachment of, 88-90 

intimin, 89-90 

IrgA homologue adhesion (Iha), 90 

kidney cells, 88 

LifA, 90 

as noninvasive pathogen, 88 

North America/Europe, 88 

Shiga-like toxin (SLT), 88, 104, 107 

Shigella, 40 

VirA, 40 

See also Escherichia coli 
enteroinvasive E. coli (EIEC) 

shigellosis, 25 

VirG, 44 
enteropathogenic E. coli (E PEC), 33, 87 

arp2/3 complex (figure), 95 

ATP release, 101-102 

attaching/effacing (A/E) lesions, 90-91, 99 

attachment/adhesion, 88-90 

breast-milk antibodies, 108 

bundle-forming pili (BFP), 88-89, 91, 93 

Cdc42, 105-106 

Cdkl, 99 

chloride secretion, 101-102 



cystic fibrosis transmembrane regulator 

(CFTR), 102 
EPEC adherence factor plasmid (EAF), 

88-90 
EspF, 101-102, 104 
filopodia, 105 
flagella of, 89 
focal adhesions, 99 
infant mortality and, 87 
intimin, 89-90, 99, 105-106 
locus for enterocyte effacement (LEE), 

90-91 
mitosis, inhibition of, 99-100 
N-WASP (figure), 95 
needles, 33 

as noninvasive pathogen, 88 
pedestal formation, 91-92, 95, 99 
Shigella, 33, 40 
signaling events, induced in host cell 

(schematic), 97 
tight junctions, disruption of, 100-101 
Tir protein. See Tir proteins 
type III secretion systems. See type III 

secretion systems 
VirA, 40 

Western world/developing world, 87 
Escherichia coli, 87 

Actinohacillus actinomycetemcomitans, 284 
adaptive immunity, 107 
apoptosis, 103-105 
arp2/3 complex, 97-98 
attaching/effacing (A/E) lesions, 90-91, 

93-94, 102-103 
bundle-forming pili (BFP), 89-90, 93, 104, 

107-108 
calcium, 102-103 
chloride secretion, 102, 106 
Crp/Fnr family, 182 
galanin-1 receptors, 101 
immune response, modification of, 

106-107 
inflammatory response, 106-107 
initial attachment of, 89-94 
inte grins, 89 
intimin, 89-90, 103, 107 
IpaC, 37 
LEE region, 93-94 




o 

w 
X 




X 
w 

Q 



Escherichia coli (cont.) 
lymphostatin, 108 
M cells, 98-99 
mitochondrial membrane, disruption of, 

105 
mitogen-activated protein (MAP) kinase 

cascades, 103 
myosin light chain (MLC) phosphorylation, 

100 
N-WASP, 97-98 
Neisseria gonorrhoeae porins, 210 
nucleolin, 89 
occludin, 100 
pedestal formation, 94-98 
phagocytosis, inhibition of, 98-99 
REPEC, 99-100, 104 
resistance to antibiotics, 108-109 
Rho family/ GTPases, 97, 105 
Tir protein. See Tir proteins 
type III secretion systems. See type III 

secretion systems 
urinary tract, 87 
See also specific serotypes 

fallopian tubes, 223 
fetal-placental barrier, 163-164 
fimbriae, 279 
flagella, 6-8 

needle complex, 6-8 

SPI-1 genes, 18 

type III secretion systems, 30 
focal adhesion kinases (FAK) 

phagocytosis, 67 

Porphyromonas gingivalis, 301 

YopH, 69, 72 

gene arrays, 262 

genetic screens, 174 

Group A streptococci (GAS), 239-252 

adhesins/invasins (table/micrographs), 
240-243 

adhesion, 239-240, 244, 247-257 

adhesion-invasion-persistence model 
(diagram), 245 

antibacterial peptides, 257 

antibiotic protection assay, 261 

apoptosis, 246-247, 259, 261 



apoptosis, bacterial, 244 

asymptomatic carriers, 244 

autoimmune T and B cell responses, 239 

C5a peptidase, 241, 253 

CD111, 255 

CD44, 241, 243, 247 

CD46, 241, 243, 252 

CD62, 255 

COX 2, 257 

diseases caused by, 239 

ezrin-radixin-moesin (ERM), 248-249 

F/Sfbl protein, 249-250 

Fas regulon, 259 

FbaA, 241, 252-253 

FbaB, 241, 253-254 

host responses to, 255-257 

IFN-gamma, 255 

Ihb-Irr regulon, 259-260 

IL-1, 255-256 

IL-6, 255-256 

IL-8, 255-256 

IL-12, 255 

IL-18, 255 

infection pathways of, 244-246 

internalization, 248-255, 261 

internalization, as an infection pathway, 

246 
internalization of, 243 
invasion, 255-257 
lamellipodia, 247-248 
Lsp, 254-255 
M proteins, 239, 241-243, 247, 250-253, 

255,258 
Ml, 262 
M3, 253, 262 
M6, 255-257, 259 
M14, 261 

M18, 253, 261-262 
M24, 255 

M49, 258-259, 262 
metastatic spread of, 244, 246 
negative global regulator (Nra), 243, 

258-259, 262 
NF-kappaB, 257 

paracellular translocation, 247-248 
persistence, 244, 246, 258 
pheromones, 260 



quorum sensing, 260 

resistance, recurrence after treatments, 

246 
responses to interactions with host cells, 

257-260 
RGD containing proteins, 253-254 
Rho family/GTPases, 247 
S. agalactiae, 246 
salivaricin A (SalA), 260 
self-limiting infections, 244 
signaling between streptococci, 260-261 
Sil/Blp-like system, 260-261 
stationary phase, 259 
streptococcal dehydrogenase (SDH), 256 
streptococcal inhibitor of complement 

(SIC), 248-249 
streptococcal NAD-glycohydrolase (SPN), 

247 
streptolysin O (SLO), 171, 247, 256, 258 
streptolysin S (SLS), 247 
tissue invasion, 244, 246-249 
TNF-alpha, 255-256 
transcytosis invasion, 246 
typing schemes for, 239 
GTPase activating proteins (GAPs), 11 
Cdc42, 14 

nucleotide cycling of, 1 1 
Salmonella, 14 
Salmonella typhimurium, 70 
Yersinia, 14, 70 
YopE, 14 

YopE, structural similarity with, 70 
GTPases, 105 
guanine nucleotide exchange factors (GEFs), 

69,71 
gut microbiota, protective effect of, 2 

Haemophilus influenzae 

autotransporter genes, 279 

IgAl, 215 
Hartmanella vermiformis, 123, 129-130, 135 
hemolytic uremic syndrome (HUS), 88 
hepatitis, 164 

human chorionic gonadotropin, 217 
human immunodeficiency virus (H IV) , 

203 
human sperm, 213, 222 



IFN-gamma, 147-149, 255 

IgA, 215-216, 229 

IgG, 211 

IL-1, 25-26, 148, 230, 255-256 

IL-2, 148 

IL-4, 148 

IL-6, 73, 148-149, 230, 255-256 

IL-8, 25-26, 230, 255-256, 286, 305-308 

IL-10, 76, 148 

IL-12, 147-149, 255 

IL-18, 25-26, 255 

immune system, low-level stimulation by 

bacteria, xii 
insulin receptor, 38 
intercellular adhesion molecule (ICAM)-l, 

305 
intracellular multiplication (icm) genes. See 

dot/icm genes 
iron 

lactoferrin, 279 

Listeria monocytogenes, 174-175, 184 

Neisseria gonorrhoeae, lib-Ill 

transferrin receptor, 282, 285 

Yersinia, 64 

joint infection, 164 

jun amino-terminal kinase (JNK), 73 

lactoferrin, 279 
Legionella 

protozoan hosts for, 127 

species of, 123, 127 

as ubiquitous, 126 
Legionella-\ike amoebal pathogens (LLAP), 

127 
Legionella micdadei, 131-132 
Legionella pneumophila, 123 

A/J mouse model, 148 

Acanthamoeba castellanii, 142-143, 145 

Acanthamoeba polyphaga, 123, 126, 135, 
145-146 

acidity, 128 

adherence/attachment, 128-131, 147 

ADP ribosylation factor- 1 (ARF1), 
137-141 

alveolar epithelial cells, 132, 147 

alveolar macrophages, 123, 147 




o 

w 
X 




X 
w 

Q 



Legionella pneumophila (cont.) 
alveolitis, 147 

amino acid depletion/limitation, 142-143 
amoebae, 123, 128. See also specific 

organisms 
antibiotics, 141 
antibodies, 147 

apoptosis, 140, 144-145, 149-150 
autophagy machinery of host, 136 
biocides, 128, 141 
bronchitis, 147 
CD4, 147-149 
CD8, 147-149 
chemical disinfection, 128 
coiling phagocytosis, 126, 130-131 
complement, 147 
complement component C3, 147 
complement-mediated killing, resistance 

to, 147 
cytolysis model (figure), 140 
D. discodeum, 135 

dot/km genes, 131-139, 143-144, 150 
egress, 138-139, 150 
endocytic pathway, evasion of, 132-135 
endoplasmic reticulum (ER), 136 
entry, 129-130 
environmental life of, within host (figure), 

125 
flagellin, 142 
freeze thawing, 128 
gene regulation in, 141-142 
growth phase, 141-142 
Hartmanella vermiformis, 123, 129-130, 

135 
IFN-gamma, 147-149 
IL-1, 148 
IL-2, 148 
IL-4, 148 
IL-6, 148-149 
IL-10, 148 
IL-12, 147-149 
immune response to, 147 
infective phase, 141 
inhalation of, 123 
integrins, 129 
intracellular survival and replication, 

131-132 
invasion and trafficking pore, 138-139 



lectins, 129 

Legionnaires' disease, 123 

Ignl locus, 131 

low infectious dose, 129 

macrophages, 123, 132, 147 

mitochondria, 125, 131-132, 136, 140 

monocytes, 147 

motility, 141-142 

MRC-5 cells, 147 

murine A/J model of, 147-148 

necrosis, induction of, 145-146 

neutrophils, 148 

nitric oxide (NO), 148-149 

opsonization, 148 

osmolarity, 128, 141 

oxidizing agents, 128 

peripheral blood monocytes, 132 

phagocytosis, 130-131, 136-137 

phases, in killing of mammalian cells, 

146 
Philadelphia in 1976, 126 
Pontiac fever, 123 
pores, 137-138, 140-141, 145. See also type 

IV secretion systems 
postexponential phase, 142-143 
protozoa species, 123 
protozoan receptor, 129 
release of intracellular bacteria {rib), 

138-139, 145 
replicative phase, 141 
resistance, 127-128 
rough endoplasmic reticulum (RER), 

125-126, 132, 134-137, 140 
sodium sensitivity, 141-142 
sonication, 128 
temperature, 128 

transmission between individuals, 129 
transmission in humans, 123 
transmission phenotype, 142 
tumor necrosis factor alpha (TNF-alpha), 

147-148 
type IV secretion system. See type IV 

secretion system 
U937 macrophages, 126, 139, 147 
uptake by monocytes and macrophages, 

130-131 
uptake, genetic aspects of, 131 
vesicles, 125, 131-132, 136-137 



virulence, 127-128, 141-142 

water, 123, 128 
legionellae 

biocides, 127-128 

chlorine, 127-128 

copper-silver ions, 127 

ecology of, 126-129 

eradication strategies, 127-128 

L. micdadei, 132 

Legionnaires' disease, 131 

monochloramine, 127 

opsonization, 131 

species of, 123 

UV irradiation, 127-128 

water overheating, 127-128 
Legionnaires' disease, 123 

amoebae, 128 

aveolar cells, 130, 147 

CD4, 147-148 

CD8, 147-148 

infectious particle proposal, 128 

L. micdadei, 131 

See also Legionella pneumophila 
lipooligosaccharide (LOS) 

ASPG-R interaction (figure), 222 

interconversion of oligosaccharides, 213 

lipopolysaccharide (LPS), contrasted, 
212 

Neisseria gonorrhoeae, 210-212, 218-219, 
222-224, 227, 230-232 

Neisseria meningitidis, 209, 212-215, 
218-219, 230-231 

opacity-associated outer membrane 
proteins (Opa), 211 

sialylated glycoform of (figure), 213 
Listeria 

Actinobacillus actinomycetemcomitans, 283 

cell-cell spread, 283 

microtubule transport systems, 283 

species in genus, 161-162 
Listeria grayi, 161 
Listeria innocua, 161, 168, 177 
Listeria ivanovii, 161 
Listeria monocytogenes, 161-181, 183 

acidity, 162 

actA, 169, 175-177, 179, 181, 183 

actin-based motility, 175-179 

actin (figure), 179 



adherence/internalization for 

nonprofessional phagocytic cells, 

165-170 
alpha-actinin, 179 
alpha- D- galactose residues, 169 
apoptosis, 163 
arp2/3 complex, 178-179 
blood-brain barrier, 163-164 
Clq, 170 

capping protein, 179 
cell heparan sulfate preoglycans (HSPG), 

169 
cofilin, 179 

complement component C3b, 170 
contaminated food and, 162 
coordinating invasion, replication, spread, 

180-183 
Crp and PrfA compared, 182 
Crp/Fnr family, 182 
in cytosol, 173-175 
dendritic cells, 163-164 
E-cadherin, 166-167 
encephalitis, 164 
entry into professional phagocytes, 

169-170 
fetal-placental barrier, 163-164 
fibronectin-binding protein, 169 
G-6-P, 174 

gene expression, induced in host, 174 
Hpt, 174 

infections, unusual forms of, 164 
internalins (inl genes), 166-168, 174, 

181 
invasion of adjacent host cells, 180 
invasion of host cells by, 164-170 
iron, 174-175, 184 
lectins, 169 

limiting nutrients, 174 
lipoteichoic acids and, 169 
Listeria adhesion protein (Lap), 169, 

180-182. See also prfA 
liver, 163 
listeriolysin (LLO), 163, 170-173, 

180-182 
M cells, 162, 166 
macrophages, inside of, 170-173 
major extracellular protein, 60, 169 
meningitis, 164 




o 

w 
X 




X 
w 

Q 



Listeria monocytogenes (cont.) 
mortality rates, 162 

N-acetylneuraminic acid (NA cNeu), 170 
overview of infection (figure), 165 
perfringolysin O (PFO), 171-172 
phagosome, escape from, 170-173 
phosphatidylcholine-phospholipase C 

(PC-PLC), 173, 180, 182 
phosphatidylinositol-specific phospholipase 

C (PI-PLC), 172-173, 180 
prfA, 177, 180-181, 183-184. See also 

Listeria pathogenicity island 1 
profilin, 178 
pseudopods, 180 
purH, purD, pyrE, arpj, 174 
salt concentrations, 162 
spleen, 163 

spread to adjacent cells, 175-180 
streptolysin O (SLO), 171 
surface virulence-associated protein (SvpA), 

173 
temperature, 162, 183 
tissue culture model systems, 164 
uptake, and nonopsonic interactions, 170 
uptake, Salmonella compared, 165 
uptake, Shigella compared, 165 
vasodilator phosphoprotein (VASP), 

178-179 
virulence gene expression, 183-184 
virulence gene regulation, 180-183 
Wiskott-Aldrich syndrome protein (WASP), 

178 
Listeria seeligeri, 161, 180, 182 
Listeria welshimeri, 161 
listeriosis, 162-164 

M cells, 59 

Escherichia coli, 98-99 

Listeria monocytogenes, 162, 166 

Salmonella, 1-2, 17 

Shigella, 25, 27 

Yersinia enter ocolitica, 64-65 

Yersinia pseudotuberculosis, 64-65 

See also Peyer's patches 
M proteins, 239, 241-243, 247, 250-253, 258. 

See also specific proteins 
M3, 262 



M6, 255-257, 259 

M14, 261 

M18, 261-262 

M24, 255 

M49, 258-259, 262 

macrophages 

alveolar, 147 

apoptosis, 74-75, 125-126 

dot/icm genes, 125-126 

Legionella pneumophila, 132, 147 

Legionnaires' disease, 123 

Listeria monocytogenes, 170-173 

Salmonella, 17 

U937, 126, 139, 147 

used to access host sites, 17 

Yersinia outer proteins (Yop), 74-75 
macropinocytosis, 9, 125-126 
major extracellular protein, 60, 169 
MENA, 48 
meningitis, 164, 205 
menses, 217, 232 

microtubule dynamic instability, 39-41 
mitochondria, xi, 125, 131-132, 136, 140 
mitogen-activated protein (MAP) kinase 
cascades 

Escherichia coli, 103 

Salmonella, 103 
monocyte chemoattractant protein 1 (MCP-1) 

Yersinia, 75 

YopH, 75 
Mycobacterium tuberculosis, xi 
myocarditis, 164 

necrosis, 144-145. See also tumor necrosis 

factor-alpha 
needle complexes, 8, 33, 35 

Escherichia coli, 33 

flagella, evolutionary relationship, 6-8 

length of, in Shigella flexneri, 32 

Salmonella, 32 

See also type III secretion systems 
Neisseria gonorrhoeae, 203, 229 

anorectal infection, 204 

antibiotic resistant strains of, 207 

antigen variation, 207-210, 212-213, 227 

apoptosis, 211 

arthritis-dermatitis syndrome, 205 



Neisseria gonorrhoeae (cont.) 

asialoglycoprotein receptor (ASGP-R), 213, 

222, 227-230 
asymptomatic, in men, 206 
asymptomatic, in women, 206-207, 220 
asymptomatic, with disseminated 

gonoccocal infection (DGI), 207 
autotransporter genes, 279 
Boswell, 203 

C3, 218-220, 223-224, 231-233 
C4, 219-220 
C5 through C8, 220 
C9, 219-220 
CD18, 220, 232 
CD46, 209, 220 
CD66, 220 

cervical epithelial cells, 227, 230 
cervical squamo-columnar junction 

(figure), 204 
classical pathway (CP) of C system, 219 
clinical syndromes caused by, 204-207 
coinfection with other STD organisms, 207 
complement/pathogenic Neisseria, 218-221 
comprehensive model of pathogenesis, 

226-233 
conjunctival infection, 204 
CR1, 211 

CR3, 209, 211, 219-220, 223-224, 231, 233 
CR4, 220 

cytoskeletal rearrangements (figure), 204 
delay, between infection and symptoms, 

205-206 
diacytosis, 229 

diplococcus arrangements (figure), 206 
disseminated gonococcal infection (DGI), 

204-205, 207, 210, 212, 217, 219, 223, 

232-233 
embryological origins of urethra /uterine 

cervix, 226 
endocarditis, 205 
fallopian tubes, 223 
filopodia extension (figure), 206 
genome project for, 215 
human chorionic gonadotropin (hCG), 217 
human immunodeficiency virus (HIV), 203 
human sperm, 213, 222 
IgA, 215-216, 229 



IgG,211 
IL-1, 230 
IL-6, 230 
IL-8, 230 

immune response, subversion of, 207 
infection rates worldwide, 203 
infertility in men from, 207 
internalization within vacuoles, 206 
intracellular invasion by, 206 
intracellular pathogen idea, 221 
iron, 216-217 
lipooligosaccharide (LOS). See 

lipooligosaccharide 
lipopolysaccharide (LPS), 212 
LNnT, 213-214 
lutropin receptor (LHr), 217, 223, 

232-233 
membrane ruffling (figure), 224-225 
meningitis, 205 
menses, 217, 232 
models for, 221 

neutrophils, interactions with, 224-226 
oligosaccharides, interconversion of, 213 
opacity-associated outer membrane 

proteins (Opa), 209, 211-212, 220, 

225-227, 230, 232 
opsonization, 219 
oxygen limitations, 217-218 
PAK1, 231 
pathogenic interactions, men versus 

women, 229 
PC-PLC, 226 

pedestal formation (figure), 205 
pelvic inflammatory disease (PID), 207, 

217, 232-233 
pharyngeal infection, 204 
phase variation, 207-210, 212-213, 227 
pili (fimbriae), 208-209, 219-220, 223, 

227 
pili (fimbriae), and resistance to 

phagocytosis, 208 
pili (fimbriae) , facilitation of adherence by, 

208 
pili (fimbriae) , glycosylation and, 209 
porins, 209-211, 219 
porins, E. coli porins, 210 
porins, nomenclature, 209 




o 

w 
X 




X 
w 

Q 



porins, PMN membrane potential changes, 
210 

pregnancies and, 207, 217 

reduction-modifiable protein (Rmp), 210 

Rho family/ GTPases, 231 

TNF-alpha, 223, 230 

translucent phenotype, 223 

untreated in men, consequences of, 206 

urethra, 204-206, 221-222, 226, 230 

urethra, released back into lumen of, 206 

uterine cervix, 223-224, 226 
Neisseria meningitidis, 215, 218 

capsule produced by, 218 

lipooligosaccharide (LOS), 218-219 

neural cell adhesion molecules (NCAM), 
218 

PorA, 209 
neural cell adhesion molecules (NCAM), 218 
neural Wiskott-Aldrich syndrome protein 
(N-WASP), 43 

Cdc42, 46-47 

Escherichia coli, 95, 97-98 

functional domains of (figure), 45 

profilin, 48 

Shigella, 43, 46-47 

See also Arp2/3 complex, 
VirG-N-WASP-Arp2/3 complex, WASP 
family 
nitric oxide (NO), 148-149, 287 

ophthalmitis, 164 
opsonization 

Legionella, 131 

Neisseria gonorrhoeae, 219 
oxygen, 217-218 

pathogenicity island (PAI) 

genetic structure of (figure), 29 

S.jlexneri, 27-29 
Penrose-Hameroff model, xii 
perfringolysin O (PFO), 171-172 
Peyer's patches, 59, 64-65, 162, 166. See also 

M cells 
pheromones, 260 

phosphatidylcholine-phospholipase C 
(PC-PLC), 173 



phosphatidylinositol-specific phopholipase C 

(PI-PLC), 172-173 
phospholipase C (PLC), 285 
pinocytosis, 9 
pneumonia, 164 
Pontiac fever, 123 
Porphyromonas gingivalis, 295-304 

adhesion, 296-298 

antecedent bacterial species, interactions 
with, 296 

apoptosis, 299, 305 

autophagocytic pathways and, 303-304 

B-cells, 296 

calcium, 302, 305 

coronary artery disease and, 308 

dendritic cells, 296, 304-305, 307 

ecological niche of, 295 

endothelial cells, 302-304, 306, 308 

fimbriae, 297-299 

focal adhesion kinase (FAK), 301 

gingival epithelial cells (GEC), 298-299, 
301-302, 305-306 

heart and aortic cells, 302-304 

hemagglutinin activity, 298 

IL-8, 305-308 

immune system reactivation, in response to 
overgrowth, 307 

integrins, 299-302 

intercellular adhesion molecule (ICAM)-l, 
305-306, 308 

intracellular replication, 299 

invasion, 301-302, 305-306 

JNK, 302 

junctional epithelium, 295-296, 299 

KB cells, 302 

major fimbriae, 296-297, 306 

matrix metalloproteinase (MMP), 305, 
307-308 

MCP-1, 306, 308 

model of GEC interactions (figure), 303 

neutrophils, 296 

paxillin, 301 

perinuclear region, 299 

periodontal disease, 295, 306-308 

proteinases secreted by, 297-298, 308 

rough endoplasmic reticulum (RER), 303 



subgingival crevice, 295 

supragingival tooth surfaces, 295 

systemic diseases, 308-309 

T-cells, 296, 307 

taxonomy of, 295 

type III secretion systems, functional 
equivalence of, 299 

uptake, 298-302, 305 

uptake, non-fimbrial dependent, 298 
pregnancies, 163-164, 217 
profilin, 43 

arp2/3 complex, 48 

Listeria monocytogenes, 178 

MENA, 48 

N-WASP,48 

Shigella, 44, 48 

VASP, 48 

WAVE/Scar, 48 
Pseudomonas, 33, 142 

quantum computations, xii 
quorum sensing, 260 

Racl-IRSp53-WAVE2 complex, 38 

Ralstonia solanacearum, 33 

Real Time RT-PCR analyses, 262 

reduction-modifiable protein (Rmp), 210 

resistance 

ampicillin, 2 

cephalosporins, 2 

E. coli, 108-109 

Group A streptococci, 246 

Legionella pneumophila, 127-128 

Neisseria gonorrhoeae, 207 

Salmonella, 2 
Rho family/GTPases 

actin-based cytoskeleton structures and, 11 

bacterial uptake by epithelial cells, 13 

Escherichia coli, 97 

Group A streptococci (GAS), 247 

Neisseria gonorrhoeae, 231 

Salmonella, 11-13 

Shigella, 33, 41 

SPI-1 genes, 13 

Yersinia, 67, 70-71 

Yop proteins (schematic), 68-69 



YopE, 67, 77 
YopT, 71, 77 

Salmonella, 1 
ampicillin, 2 
CD18, 17 
Cdc42, 11-13 
cephalosporins, 2 
complementation, 3 
dendritic cells, 17 
drug resistance, 2 
entry, 3-8. See also Salmonella pathogenicity 

island- 1 
flagellar export system, 30 
GTPase activating proteins (GAPS), 14 
infection, animal models of, 8-9 
infection, cell culture models of, 9-10 
infections through, 2 
internalization, 8-10 
inv mutant, 3 
invasion-dependent vs independent 

phenotypes, 17-18 
invasion gene expression, 5 
invasion-gene expression, environmental 

signals affecting, 5 
invasion, genes regulating, 13-16 
invasion mechanisms, 1-11 
invasion phenotype, 3 
M cells, 1-2, 17 

macrophages, used to access sites, 17 
mitogen-activated protein (MAP) kinase 

cascades, 103 
needle, 32 
Rho GTPase, 11-13 

rough endoplasmic reticulum (RER), 137 
Salmonella pathogenicity islands. See 

Salmonella pathogenicity islands 
Shigella, 18, 31-33, 165 
Shigella Jlexneri, 30, 32 
Sip proteins, 8, 18, 33 
SipA, 15-16 
SipC, 15-16 
SirA/BarA of, 142 
sopB, 14-15, 39 
sopE, 13-15 
sopE/E2, 14-15 




o 

w 
X 




X 
w 

Q 



Salmonella (cont.) 

SPI-1 lacking, dissemination to systemic 
organs, 17 

SptP, 14-15 

symptoms, 2-3 

treatment for, 2 

type III secretion systems. See type III 
secretion systems 

uptake compared, Listeria monocytogenes, 
165 

uptake compared, Shigella, 165 

Yersinia, 14, 18, 31-32 

YopM, 76 

See also specific species 
Salmonella choleraesuis, 2 
Salmonella enterica, 59, 64 
Salmonella pathogenicity island- 1 (SPI-1), 1, 
3-6, 8, 13, 15, 18 

flagella, 18 

gene expression, regulation of, 5-6 

historical/evolutionary perspective, 18 

pathogenesis, role in, 16-18 

proteins delivered by, 8 

regulation of, 3-5 

Rho GTPase, 13 

Shigella, 18 

Yersinia, 18 

See also type III secretion systems 
Salmonella pathogenicity island-2 (SPI-2), 6 
Salmonella typhimurium 

GAP domains, 70 

VirG, 42 

YopE, 70 

See also Salmonella 
Shiga-like toxin (SLT), 88, 104, 107 
Shigella, 25-45 

Actinohacillus actinomycetemcomitans, 
283 

apoptosis, 25-26 

arp2/3 complex, 38, 45-47 

CD44, 33-35 

Cdc42, 37-38, 46-47 

cell-cell spread, 41, 283 

cytoskeletal rearrangements (figure), 35 

enterohemorrhagic E. coli (EH EC), 40 

enteropathogenic E. coli (EPEC), 32-33 



epithelial cell infection (schematic), 26 

IL-1, 25-26 

IL-8, 25-26 

IL-18, 25-26 

induction of host cellular signaling, 33-38 

inflammatory response and, 25-27 

insulin receptor, 38 

Ipa proteins, 33-39 

large plasmid invasion-associated genes, 

27-29 
M cells, 25, 27 
macrophages, entry into, 37 
microtubule dynamic instability, 39-41 
microtubule transport systems, 283 
needle structures, 33, 35 
neural Wiskott-Aldrich syndrome protein 

(N-WASP), 43, 45-47 
pathogenicity island, 27-29 
polarized epithelial cells, basolateral entry 

into, 27 
profilin, 44, 48 

Racl-IRSp53-WAVE2 complex, 38 
Rho GTPases, 33, 41 
Salmonella, 18, 31-33 
Salmonella pathogenicity island-1 (SPI-1), 

18 
Salmonella, uptake mechanisms compared, 

165 
Sip proteins, 18 
SopA, 44 
SopB, 39 
type III secretion systems. See type III 

secretion systems 
uptake, Listeria monocytogenes compared, 

165 
vasodilator stimulating phosphoprotein 

(VASP), 44 
VirA, 39-41 
VirG, 41, 44, 46-47 
VirG ligands, 44-45 
VirG model (figure), 43 
VirG-N-WASP complex, 45 
Yersinia, 18, 31-33 
YopM, 76 
Shigella boydii, 25 
Shigella dysenteriae, 25 



Shigella flexneri, 25, 30, 32 
needle, length of, 32 
pathogenicity island of, 27-29 
type III secretion systems. See type III 
secretion systems 

Shigella sonnei, 25 

shigellosis, 25 

sinusitis, 164 

skin infection, 164 

Staphylococcus aureus agr, 260 

Streptococcus agalactiae, 246 

Streptococcus pneumoniae, 260 

Streptococcus pyogenes. See Group A 
streptococci 

T-cells, 248, 307 
temperature 

L. monocytogenes, 183 

Legionella pneumophila, 128 

legionellae, 127-128 
Tir protein 

enteropathogenic E. coli (EPEC), 91, 99, 
105-106 

Escherichia coli, 89-90, 93-95, 97-98, 
107 

focal adhesions, 97 

REPEC99 

tyrosine phosphorylation of, 97 
TMpred, 37 

toll-like receptor (TLR) system, 73, 75 
transferrin receptor, 282 
tumor necrosis factor- alpha (TNF -alpha) 

Group A streptococci (GAS), 255-256 

Legionella pneumophila, 147-148 

Neisseria gonorrhoeae, 223, 230 

Yersinia, 76 

Yop P/J, 73 
type III secretion systems, 1, 6-8, 59 

ATP release and, 101-102 

bacteria, widely distributed among, 3 

basal part of, 32 

conditions for inducing, 5-6 

conserved morphological features of, 
32-33 

enteropathogenic E. coli (EPEC), 33, 91, 
99-102 



Escherichia coli, 90-94, 103 
flagellar export system, 30 
functional equivalence in Porphyromonas 

gingivalis, 299 
genes regulating invasion, 13-16 
invasion-dependent vs independent 

phenotypes, 17-18 
IpgD, 38-39 
proteins, 3, 31-32 
purified components from, 32 
S. flexneri, 27-29, 32 
Salmonella, 3-8, 13, 15, 17-18, 31-33 
Salmonella enterica, 59, 64 
schematics of, 4-5, 7, 30-31 
Shigella, 18, 29-33, 38-39, 48-49, 64 
Shigella flexneri, 27-29, 32, 59 
SPI-1 lacking, in Salmonella, 17 
temperature, 62-63 
Yersinia, 18, 31-32, 59, 62-64 
Yersinia enter ocolitica, 64 
Yersinia, Ysc-Yop system, 77 
See also needle complexes, salmonella 

pathogenicity islands, specific proteins 
type IV secretion system, 125, 132, 136, 138, 

146, 150 
See also dot/icm genes 

ubiquitin-like protein proteinases, 74 
uropathogenic E. coli (U PEC), 87 

vasodilator stimulating phosphoprotein 
(VASP), 43 

L. monocytogenes, 178-179 

profilin, 48 

Shigella, 44 
Vibrio cholerae, 42 
VirG-N-WASP complex, 45, 47 

WAVE family, 45 

functional domains of (figure) , 45 

profilin, 48 
Wiskott-Aldrich syndrome protein (WASP), 
45 

functional domains of (figure) , 45 

Listeria monocytogenes, 178 

Shigella VirG, 46 




o 

w 
X 




X 
w 

Q 



YadA, 63 
Yersinia, 59-69 

apoptosis, inhibition of, 72-73 

Cdc42, 67, 70 

GTPase activating proteins (GAPs), 14, 

70 
guanine nucleotide exchange factors 

(GEFs), 69, 71 
high pathogenicity island (HPI), 64 
infection, routes of (schematic), 60, 62 
inflammatory response, inhibited, 72-73 
integrins, 76 

interleukin factors. See specific factors 
invasin (Inv) protein, 76 
iron, 64 

jun amino-terminal kinase (JNK), 73 
LcrV, 76, 77 
monocyte chemoattractant protein 1 

(MCP-1), 75 
phagocytosis, inhibition of, 67 
Pla protease, 63-64 
Rho family/ GTPases, 67, 70-71 
Salmonella, 14, 18, 31-32 
Salmonella pathogenicity island- 1 (SPI-1). 

See Salmonella pathogenicity island- 1 
Shigella, 18, 31-33 

specific Yersinia chaperone (Syc), 63 
temperature, 62-63 
toll-like receptor (TLR) system, 73 
tumor necrosis factor-alpha, 76 
type III secretion systems. See type III 

secretion systems 
ubiquitin-like protein proteinases, 74 
virulence system of, 61-64 
Yop proteins. See Yersinia outer proteins 
Yersinia enter ocolitica, 59, 64-65 



Yersinia outer proteins (Yop), 33, 59-61, 

67-69. See also specific proteins 
Yersinia pestis, 59, 65 
Yersinia pseudotuberculosis, 42, 59, 64-65 
Yop P/J, 73-75 

IL-6, 73 

TLR system, 73 

tumor necrosis factor-alpha (TNF-alpha), 73 

ubiquitin-like protein proteinases, 74 
YopE, 14,61,67-70, 77 

deletion of, 67 

GAPs, 14, 70 

Rho family/ GTPases, 67, 77 

Salmonella typhimurium, 70 
YopH, 61, 67, 72, 75-77 

calcium, 72 

deletion of, 67 

focal adhesion kinases (FAK), 69, 72 

integrins, 72 

monocyte chemoattractant protein 1 
(MCP-1), 75 

neutrophils, 72 
Yop J, 61 
YopM, 76-77 
YopO, 61, 77 
YopO/YpkA, 67, 71 
YopP, 61, 76 
YopP/J, 61, 77 

apoptosis, in macrophages, 74-75 

TLR system, 73 
YopT,61,67, 71, 77 

deletion of, 67 

guanine nucleotide exchange factors 
(GEFs), 71 

Rho family/ GTPases, 71, 77 
YpkA, 61 




GDP 



GDP 



GEF 



GTP 



GEF 



YopE 



YopE 
YopT 
YpkA 



Inhibition of 
phagocytosis 



GTP ^ y ^» 

Rho YopT 
Cdc42 
' Rac -^ 

YpkA v "\ 



Actin 



Rh 



Filopodia 

and 

microspikes 



FR 




-A 



W GTP w ^\ 

ho Rac Cdc42 



Lamellipodia 

and 

membrane ruffles 



Stress fiber 
formation 



Actin-based 
Cytoskeletal rearrangements 



B 



PI|3,4.5|Pj 

PKB/ 



PDK1 





FKHR 



YopH 



YopH 



Actin-based 
Cytoskeletal rearrangements 



YopH 



Inhibition of phagocytosis 
Downregulation of inflammation 



r 



Yersinia 




LPS, 
or other bacterial signals 



? 
Casp-8 

I 



"\ 



<P 



Apaf-1 
Casp-9 

T 

V. Casp-3/7 

\ 

APOPTOSIS 



J 



YopP/J 





YopP/J 



YopP/J 
X-S^Q + O 



De-sumoylation ? 
De-ubiquitinylation 



YopP/J 



YopP/J 



Downregulation of inflammation 
Induction of macrophage apoptosis 




Proinflammatory cytokine production 
Production of survival genes 



Figure 3.4. Molecular mechanism of the Yop effectors in the host cell. 





Figure 4.2. A cluster of pedestal structures induced by EPEC on the surface of a HeLa cell. 




t 




CO 



■B 

O 

bo 






Oh • 

0> - 

<+- 

o 

c 
o 

Cfl 

ta 



5/J 



V5 

• — 
cfl 

2 
s* o 

^ E 

? tS 

Cfl 

u 



M 

e 

• — 

£ 

'Sic 

<u o 

5 V 

(X g 



c« 



o 

J.* 



a. 
ri 



cd 



o 

s 

• I— I 
CU 

■s 

ti 

3 

u 

I 

■i-H 

.2 
13 

.S 



CU 

§ 

-St 



sx 
Cfl 

■3 

U 






C+H 

o 

Cfl 

•|-H 

Cfl 
>s 

O 

s 

(U 

Oh 
CD 

V 

cu 

-a 



£* M 



o 

H 

5b _ 

o 



CD 

O 



< 



^t 



cu 
J-l 

a 

•i-H 



CU 

u 

-M 
Cfl 

O 



ex 

Cfl 

cu 



o 

10 tfcj 



Cfl 
Cfl 

cu 

U 

cu 



A 









a> 

N 
'</> 

0> 

3 
CT 
JO 

Q. 

CO 

> 

c 
o 

CO 
CO 

0) 

1_ 

Q. 

X 
<1) 






9 



<\ 



I 

-r 
<1 



IT) 
SC 

<1 



>C 



SC 

< 






H 

'— 
Z 


► \ 

00 ' 

1 

IT) 

< 


< 

1 


00 

iH 
1 00 
1 w* 
\< 
W 
\ *- 
» ^— 

\ *^ 

^©1 O 

w ~- 
-7 \ < 

^♦\ 

00 \ 

< \ 
3\ 

< 

1 


00 
\<3 


- 



o 

(M 



O 
O 



O 
CO 



O 
CD 



O 



O 
CM 



CD 
N 
W 

CD 

3 
O" 
CD 



O 
O 
LO 
CO 

S69 



O 

o 
o 

CO 



o 
o 

IX) 
CM 



o 
o 
o 

CM 



o 

o 
in 



o 
o 
o 



o 
o 

LO 



o 
o 



u 



QO/AuAjpesnO josuun 



SC 

T 



-1 



L 



SC 



SC 



T 



O 
-D 






« 




00 



< 



z 







< 



in 
00 

1 

< 



C/l 

=5 

cj 

■ 

u 
00 



8P 

o 
o 



C/3 

I 

o 

■i-H 

CO 
CO 

B 

cs 
O 

a 

o 

■5 

o 
U 

so 

I 

tin 



- 












'..■.■■.■■.■■.■» 

^-••••:-:ivi 

q''-ai^.:,, 

CO 





c 



8 

Z 

H 

H-3 



4 



Z 

< 
z 






x 

■ 

t 




CO 

U 



Z 



SB 

S3 

< 

Z 

z 



t t -I 






■■■■L.V- r 






fe£ 

E 



I 

o 




QJ 

s 

O 



w 

u 
u 

o 
u 

o 

o 

QJ 

+-> 
<-M 

o 

O 
t) 



u 

■I— t 

0£ 

O 

Oh 

J-l 

• i— i 

fS 

00 
QJ 

S-H 
•i-H 





Figure 9.4. Microtubule asters and the A. actinomycetemcomitans kinesin-like entity. 




Figure 9.5. Schematic representation of A. actinomycetemcomitans invasion of epithelial 
cells. 







u 

m 
O 

o 

o 

•1— I 

CO 

> 



3 

"Sb 

o 

Pi 

o 

o 



u 

<D 

u 
en 
QJ 

O 

o 

3 



CD 

•I— C