Shigella and enteroinvasive E coli pathogenicity factors

background image

MiniReview

Shigella spp. and enteroinvasive Escherichia coli pathogenicity factors

Claude Parsot

*

Unite´ de Pathoge´nie Microbienne Mole´culaire, INSERM U389, Institut Pasteur, 25 rue du Dr Roux, 75724 Paris Cedex 15, France

Received 23 June 2005; accepted 17 August 2005

First published online 15 September 2005

Edited by I. Henderson

Abstract

Bacteria of Shigella spp. (S. boydii, S. dysenteriae, S. flexneri and S. sonnei) and enteroinvasive Escherichia coli (EIEC) are

responsible for shigellosis in humans, a disease characterized by the destruction of the colonic mucosa that is induced upon bacterial
invasion. Shigella spp. and EIEC strains contain a virulence plasmid of

220 kb that encodes determinants for entry into epithelial

cells and dissemination from cell to cell. This review presents the current model on mechanisms of invasion of the colonic epithelium
by these bacteria and focuses on their pathogenicity factors, particularly the virulence plasmid-encoded type III secretion system.
Ó 2005 Federation of European Microbiological Societies. Published by Elsevier B.V. All rights reserved.

Keywords: Invasion; Secretion; Chaperone; Regulation; Pathogenesis; Virulence; Plasmid; Type III secretion

1. Introduction

Bacteria of Shigella spp. and enteroinvasive Esche-

richia coli (EIEC) are responsible for shigellosis in hu-
mans. The burden of this disease was estimated to 150
million cases and 1 million deaths per year in the devel-
oping world

[1]

. Shigellosis is characterized by the

destruction of the colonic epithelium provoked by the
inflammatory response that is induced upon invasion
of the mucosa by bacteria. Shigella is divided in four
groups (or species), S. boydii, S. dysenteriae, S. flexneri
and S. sonnei. However, these bacteria are so closely re-
lated to each other and to E. coli strains that, in fact,
they all belong to the species E. coli

[2]

. Most studies

have been conducted on S. flexneri and, notwithstand-
ing strain specificities, results probably apply to other
Shigella spp. and EIEC. As compared to commensal
and other pathogenic E. coli strains, characteristic fea-
tures of Shigella spp. and EIEC strains are the presence
of a virulence plasmid (VP) of

220 kb and their ability

to induce their entry into epithelial cells and disseminate
from cell to cell. This review briefly presents the model
on mechanisms of invasion of the epithelium by bacteria
of Shigella spp. and focuses on their pathogenicity fac-
tors, particularly the VP-encoded type III secretion
(TTS) system. This system consists of a TTS apparatus
(TTSA) that spans the bacterial envelope, translocators
that transit through the TTSA and insert into the host
cell membrane to form a pore (translocon), effectors that
transit through the TTSA and the translocon and are in-
jected within the eukaryotic cell, chaperones that bind to
translocators and some effectors prior to their transit
through the TTSA, and transcription activators. Due
to space limitations, priority for references was given
to recent publications in which readers will find refer-
ences to older, but not less interesting, works.

2. Pathogenesis

The current model on mechanisms of pathogenesis

induced by bacteria belonging to Shigella spp. are

0378-1097/$22.00

Ó 2005 Federation of European Microbiological Societies. Published by Elsevier B.V. All rights reserved.

doi:10.1016/j.femsle.2005.08.046

*

Tel.: +33 1 45 68 83 00; fax: +33 1 45 68 89 53.
E-mail address:

cparsot@pasteur.fr

.

www.fems-microbiology.org

FEMS Microbiology Letters 252 (2005) 11–18

background image

derived from in vitro and in vivo studies using various
cell types (epithelial cells, macrophages, monocytes,
fibroblasts and red blood cells) and animal models of
infection, such as the cornea in guinea pigs, ligated ileal
loops in rabbits and lungs in mice, in which bacteria in-
duce an inflammatory response leading to destruction of
the corresponding epithelium. In the colonic mucosa,
bacteria are proposed to cross the epithelial layer by
invading M cells overlaying lymphoid follicles, which al-
lows them to reach the basolateral pole of epithelial cells
where they induce their uptake (

Fig. 1

). Entry into epi-

thelial cells involves rearrangements of the cell cytoskel-
eton that extend beyond the zone of contact between the
bacterium and the cell membrane, leading to membrane
ruffling and engulfment of the bacterium within a vacu-
ole

[3]

. Once internalized by epithelial cells, bacteria rap-

idly lyse the membrane of the entry vacuole and gain
access to the cell cytoplasm where they multiply with a
generation time of

40 min. By inducing actin polymer-

ization at one of their poles, intracellular bacteria move
within the cytoplasm of infected cells. This movement
generates the formation of protrusions that contain
one bacterium at their tip and are engulfed by adjacent
epithelial cells, thereby allowing bacteria to disseminate
from cell to cell without being exposed to the external
milieu. Peptidoglycan fragments released by intracellu-
lar bacteria are detected by the Nod1 pathway, leading
to phosphorylation and degradation of IjB, transloca-
tion of NF-jB to the nucleus and activation of NF-jB
regulated genes

[4,5]

. Analysis of the transcriptome of

infected epithelial cells showed, in particular, increased
expression of the gene encoding IL-8, a potent chemoat-
tractant for neutrophils

[6]

. Thus, epithelial cells actively

participate in the detection and signaling of invasive
bacteria to host defences. Bacteria released from M cells
(after their initial uptake) or epithelial cells (after intra-

cellular multiplication) interact with macrophages, es-
cape

from

the

phagocytic

vacuole

and

induce

apoptosis of infected cells. Apoptotic macrophages re-
lease pro-inflammatory cytokines, including IL-1 and
IL-18, which, together with IL-8 released from infected
epithelial cells, leads to recruitment of polymorphonu-
clear cells (PMN) at the site of infection. Transmigration
of PMN destabilises the epithelial barrier and facilitates
further invasion by luminal bacteria. Interactions of
bacteria with host cells, innate and adaptive immune re-
sponses induced upon infection and vaccine develop-
ments are presented in recent reviews

[7–9]

.

3. Chromosomal pathogenicity islands and black holes

The genomic sequence of two S. flexneri 2a strains,

CP301 and 2457T, was recently determined

[10,11]

.

Genomes of these strains are very similar in size
(4 Mbp), sequence and organization. As compared to
E. coli K-12 strain MG1655, major differences are (i)
the number of insertion sequences (IS), with 400 com-
plete or partial IS in CP301 vs. 40 in MG1655; (ii) the
number of genes that are inactivated or missing, 900
genes of MG1655 being absent or inactivated by frame-
shift mutations or insertion of IS in 2457T; and (iii) the
presence of 195 genes specific to S. flexneri. Most spe-
cific genes are carried by islands that either encode an
integrase, are located at tRNA sites, or are associated
with IS or prophage remnants. A number of these genes
encode proteins predicted to be membrane-associated or
secreted or exhibiting similarities to adhesins. Genes car-
ried by islands involved in pathogenesis include sit genes
encoding an iron uptake system

[12]

, sigA encoding an

extracellular protease

[13]

, gtr genes encoding proteins

involved in glucosylation of the O antigen

[14]

, shiA

Fig. 1. Model of pathogenesis induced by Shigella spp. Bacteria cross the epithelium barrier by entering into M cells (1). They are delivered to
resident macrophages, in which they induce apoptosis (2), and reach the basolateral pole of epithelial cells (3), in which they induce their entry (4).
Movement of intracellular bacteria (5) leads to the formation of protrusions and dissemination of bacteria within the epithelium (6). Release of
cytokines and chemokines, including IL-1 by apoptotic macrophages (A) and IL-8 by infected enterocytes (B), promotes recruitment of monocytes
that migrate through the epithelial barrier (C), facilitating entry of luminal bacteria into epithelial cells (D) and increasing invasion of the epithelium
(E).

12

C. Parsot / FEMS Microbiology Letters 252 (2005) 11–18

background image

encoding a membrane-associated protein of unknown
function

[15]

and stx genes encoding the Shiga toxin

(in S. dysenteriae strains). Many chromosomal genes
that are not specific of Shigella spp. are required for
virulence, including those encoding the transcription ter-
mination factor Rho, the DNA topoisomerase I, nucle-
oid-associated proteins H-NS and FIS

[16]

, tRNA

modifying enzymes

[17]

, the superoxide dismutase SodB,

the protease DegP, the outer membrane protein OmpC,
the periplasmic disulfide bond catalyst DsbA and pro-
teins involved in lipopolysaccharide biosynthesis or cell
division.

4. Overview of the VP

Sequence analysis of pWR100 and pCP301 from S.

flexneri 5 and 2a strains, respectively, indicated that
the VP is composed of a mosaic of

100 genes and

numerous IS, these latter representing one-third of the
VP

[10,18,19]

. The 30-kb region (designated the entry

region) essential for entry into epithelial cells carries
genes encoding components of a TTSA, translocators
and some effectors, chaperones and transcription activa-
tors (

Fig. 2

). Genes encoding other effectors are scat-

tered on the VP. The TTS system comprises 50 genes
exhibiting a similarly low G+C content (34% G+C),
indicating that it was acquired by lateral transfer. Two
gene clusters exhibit a G+C content of 42%, one encod-
ing IcsA/VirG (involved in movement of intracellular
bacteria), a truncated PapC, UshA (a periplasmic
UDP-sugar hydrolase) and PhoN1 (a periplasmic acid
phosphatase) and the other encoding Orf185 and
Orf186 (of unknown function), VirK (involved in pro-
duction or localization of IcsA) and MsbB2 (an acyl

transferase modifying the lipid A). The organization
and G+C content of these two regions suggest that they
were acquired from the same source, different from the
entry region, and that they initially encoded Pap pili.
The sepA gene encoding a secreted protease has a
G+C content of 49%, suggesting that it came from a
different source. The VP contains several gene clusters
involved in plasmid replication, partition and post seg-
regation killing and an incomplete transfer region. The
presence of two seemingly intact and one incomplete
partition systems exhibiting G+C contents that are dif-
ferent from one another and different from that of the
replication region suggests that the VP carries elements
from four different plasmids

[18]

. pWR100 and pCP301

differ by the presence or absence 16 fragments, most of
which correspond to IS and deletions involving IS or di-
rect repeats. This suggests that the common ancestor of
these two VP contained more intact genes and less IS.
Genes detected on one VP are not present on all VP:
impCAB genes of pCP301 are absent on pWR100 and
sepA and several other genes of pCP301 and pWR100
were not detected on the VP of a S. flexneri 6 strain

[20]

. Many S. flexneri strains carry two small plasmids,

one of which encodes the Cld protein controlling the
number of O antigen repeats, and often more than
one large plasmid. For example, a 165-kb plasmid
exhibiting 99.7% identity with plasmid R27 from Salmo-
nella enterica serovar Typhi was present in strain 2457T
used for sequencing

[11]

.

5. Expression of VP genes

The regulation of expression of VP genes has been

extensively studied

[21]

. Transcription of genes of the

Fig. 2. Genetic map of the entry region and electron microscopy analysis of the TTSA. Top panel: schematic genetic map of the 30-kb entry region
(in two parts); genes encoding transcription activators are in purple, effectors in red, chaperones in blue, components of the needle complex in green
and innermembrane proteins in yellow. Bottom panel: structural analysis of the TTSA; (A) negative staining of one TTS apparatus on osmotically
shocked bacteria; (B) deduced projection density map of images as shown in A; (C) negative staining of one purified needle complex; (D) average
image of the major class of needle complexes (as shown in C); (E) surface representation of the volume of the needle complex, assuming cylindrical
symmetry (adapted from

[32,34]

).

C. Parsot / FEMS Microbiology Letters 252 (2005) 11–18

13

background image

entry region is regulated by temperature and under
the control of two VP-encoded proteins, VirF, a mem-
ber of the AraC family of transcription activators, and
VirB, a member of the ParB family of partition pro-
teins. Binding of H-NS to virF and virB promoters
prevents transcription of these genes at 30

°C

[22]

.

At 37

°C, changes in DNA conformation lead to in-

creased transcription of virF and activation of the virB
promoter by VirF and production of VirB

[23]

. Acti-

vation of VirB-dependent promoters, including those
of the entry region, might involve oligomerization of
VirB and displacement of H-NS

[24]

. In addition to

genes of the entry region, VirB controls transcription
of

15 VP genes

[25]

, most of which are part of the

TTS system. Inactivation of orf81 encoding a tran-
scription activator of the AraC family did not affect
expression of any VP gene

[25]

. Since orf81 is inacti-

vated or absent in other S. flexneri strains, this gene
might be a remnant of a regulatory circuit that is
no longer used.

The TTSA assembled by bacteria growing in broth at

37

°C is only weakly active and is activated upon contact

of bacteria with epithelial cells (and deregulated in some
mutants). TTSA activation (or deregulation) leads to in-
creased transcription of 12 VP genes encoding TTSA
substrates

[25–28]

. Increased transcription of these genes

in conditions of secretion is controlled by MxiE, a tran-
scription activator of the AraC family produced by tran-
scriptional frameshifting from two overlapping ORF

[29]

. The cis-acting element required for increased tran-

scription of MxiE-controlled promoters is the 17-bp
MxiE box (GTATCGTTTTTTTANAG) located be-
tween positions

49 and 33 with respect to the tran-

scription start site

[30]

. There are 9 MxiE boxes on the

VP and MxiE boxes are present in the 5

0

region of chro-

mosomal ipaH genes encoding TTSA substrates

[25]

. To

be active, MxiE requires IpgC, the chaperone of the
translocators IpaB and IpaC, acting as a co-activator

[27]

. In conditions of non-secretion, IpaB and IpaC

are associated independently with IpgC, which titrates
IpgC. In addition, MxiE is associated with the TTSA
substrate OspD1 and its chaperone Spa15, which pre-
vents MxiE from being activated by free IpgC

[31]

.

Upon TTSA activation, transit of IpaB and IpaC liber-
ates IpgC, but this is not sufficient to activate MxiE as
long as OspD1 is present in the cytoplasm. Following
transit of OspD1 through the TTSA, MxiE is liberated
and an interaction between MxiE and IpgC would
allow MxiE to bind to and activate transcription at
MxiE-controlled promoters. On the basis of their
expression profiles, TTSA substrates can be classified
into three categories (

Fig. 3

): (i) those that are expressed

independently of the TTSA activity; (ii) those that are
expressed in conditions of non-secretion and induced
in conditions of secretion; and (iii) those that are
expressed only in conditions of secretion

[25]

.

6. The TTS apparatus

The TTSA encoded by mxi and spa genes of the entry

region has been visualized by transmission electron
microscopy on osmotically shocked bacteria

[32,33]

and scanning electron microscopy on intact bacteria

[14]

. The core of the TTSA, designated the needle com-

plex, consists in a needle, composed of MxiH and MxiI,
and a platform, composed of MxiG, MxiJ, MxiD and
presumably MxiM

[33,34]

(

Fig. 2

). The platform can

be assembled in the absence of MxiH and MxiI, but is
not functional, i.e., unable to secrete TTSA substrates.
In the needle, MxiH subunits are arranged in a helical
fashion to form a 500-A

˚ cylindrical structure with an

outer diameter of 70 A

˚ and a central channel of 20 A˚

[35]

. Inner membrane proteins MxiA, Spa9, Spa24,

Spa29 and Spa40 are not present in the needle complex,
the purification of which involves extraction with deter-
gents. Incorporation of MxiH and MxiI into the needle
completes the TTSA and requires Spa47, the ATPase
providing energy for transit of substrates through the
TTSA. Steady-state amounts of MxiH and MxiI are re-
duced in secretion defective mutants, indicating that, in
contrast to other TTSA substrates stored when the
TTSA is not active, needle components are degraded
when they are not delivered to the TTSA. MxiK and
MxiN, both of which interact with Spa47, are required
for transit of needle components, but not Ipa proteins
(when the TTSA is artificially assembled by overexpres-
sion of MxiH or MxiI in mxiK or mxiN mutants)

[36]

.

Inactivation of spa32 led to assembly of needles that
were two times longer than those of the wild-type strain
and inability to secrete Ipa proteins upon TTSA activa-
tion

[37,38]

, suggesting that Spa32 is required to control

the needle length and switch the TTSA substrate speci-
ficity from needle components to other substrates.
Spa32, Spa33 and MxiC transit through the TTSA.
Analysis of a spa33 mutant indicated that Spa33 is re-
quired for IpaB and IpaC secretion, but the involvement
of Spa33 in completion of the needle complex is not
known

[39]

. Sequence similarities suggest that, like

YopN of pathogenic Yersinia spp., MxiC might be in-
volved in control of TTSA activity.

The TTSA is activated upon contact of bacteria with

epithelial cells or exposure to the dye Congo red and is

Fig. 3. Repertoires of TTS effector genes controlled by VirB and/or
MxiE. Transcription of genes controlled by VirB is independent upon
the TTSA activity, whereas transcription of genes controlled by MxiE
is activated (or increased) in condition of secretion (adapted from

[25]

).

14

C. Parsot / FEMS Microbiology Letters 252 (2005) 11–18

background image

deregulated, i.e., constitutively active, by inactivation of
ipaB or ipaD. The mechanism by which IpaB and IpaD
control TTSA activity is not known, one possibility
being that these proteins are required to form a complex
that plugs the TTSA. It was recently shown that choles-
terol-containing membranes trigger TTSA activity

[40]

.

Genes whose expression is regulated by TTSA activity
are induced upon entry into epithelial cells and no
longer transcribed in bacteria growing in the intracellu-
lar compartment, suggesting that TTSA activity is
turned on during entry and off in the cell cytoplasm

[26]

. The involvement of the TTS system in lysis of the

cell membrane that surrounds bacteria in protrusions
during cell-to-cell spread suggests that the TTSA is re-
activated in protrusions

[41,42]

.

7. TTS chaperones, translocators and effectors

Due to regulation of TTSA activity by external sig-

nals, TTSA substrates are stored in the bacterial cyto-
plasm prior to their transit through the TTSA.
Dedicated chaperones stabilize translocators and some
effectors and/or maintain them in a secretion-competent
state while they are stored in the bacterial cytoplasm

[43]

. Four chaperones are encoded by the VP, (i) IpgA,

the chaperone for IcsB

[44]

; (ii) IpgE, the chaperone for

IpgD

[45]

; (iii) Spa15, the chaperone for IpaA, IpgB1,

OspB, OspD1, OspC2 and OspC3

[46,47]

; (i) and IpgC,

the chaperone for IpaB and IpaC

[48]

. IpgC and Spa15

are also involved in the regulatory mechanism by which
TTSA activity controls transcription of a set of genes
encoding effectors

[27,31]

.

Translocators are proposed to be IpaB and IpaC as

the sequence of these proteins exhibits two hydrophobic
regions and purified IpaB and IpaC interact with lipid
membranes

[49,50]

and both proteins are essential for

contact hemolytic activity (which reflects the ability of
bacteria to insert a pore of a defined size in the mem-
brane of erythrocytes) and associate with the membrane
of erythrocytes

[32]

. IpaB interacts with numerous part-

ners, probably in a sequential order: it binds to IpgC in
the bacterial cytoplasm and probably to IpaD within the
TTSA, forms trimers

[50]

and associates with IpaC in

the external milieu. It also interacts with CDC44 on
the cell surface

[51]

, cholesterol in the cell membrane

[52]

and caspase-1 in the cytoplasm of infected macro-

phages

[53]

. Although not part of the translocon, IpaD

is probably required for insertion of IpaB and IpaC in
cell membranes

[54]

. IpaB and IpaC are required for ly-

sis of the membrane of the vacuole surrounding bacteria
internalized by macrophages and epithelial cells and the
membrane of protrusions during cell-to-cell spread

[41,42]

.

Effectors are defined as proteins that, via the TTSA

and the translocon, are injected into the cell, where they

affect cellular functions. Binding of IpaB to caspase-1
and a direct role of IpaC in promoting cellular signaling
cascades that lead to internalization of bacteria in epi-
thelial cells

[55]

suggest that translocators are also en-

dowed with effector functions. The repertoire of
effectors, identified by N-terminal sequencing of pro-
teins secreted by a TTSA-deregulated strain and se-
quence analysis of the virulence plasmid and, more
recently, the chromosome consists in

25 proteins en-

coded by the VP and 3–5 proteins encoded by the chro-
mosome (depending on strains). These latter belong to
the IpaH family and, so far, there is no evidence that
the chromosome encodes other TTSA substrates. Mem-
bers of the IpaH family are characterized by a variable
N-terminal domain containing 7–10 repeats of a LRR
motif and a constant C-terminal domain. In addition
to IpaH proteins, several VP-encoded effectors belong
to multigene families, with three ospC genes, three ospD
genes, two ospE genes and two ipgB genes

[18]

.

The roles of effectors are starting to be elucidated, but

their exact function, i.e., enzymatic or binding activities,
remains elusive in most cases. IpgD specifically dephos-
phorylates

phosphatidyl-inositol

4,5-biphosphate

(PtdIns(4,5)P2) into phosphatidyl-inositol 5-phosphate
upon entry of bacteria within epithelial cells, which
uncouples the plasma membrane from the cellular cyto-
skeleton

[56]

. IpaA binds to the focal adhesion protein

vinculin, which promotes depolymerization of actin fila-
ments

[57]

. IpgB1, directly or indirectly, stimulates activ-

ities of Rho GTPases Rac1 and Cdc42, contributing to
the formation of membrane ruffles during entry

[58]

.

VirA binds ab-tubulin dimers and destabilizes microtu-
bules, which also leads to Rac1 activation and forma-
tion of membrane ruffles

[59]

. IpgD, IpaA, IpgB1 and

VirA facilitate the formation of entry structures, indicat-
ing that several effectors act synergistically to promote
internalization of bacteria. IcsB was shown to interact
with IcsA (VirG) and this interaction was proposed to
prevent association of IcsA with the autophagy protein
Atg5

[60]

.

8. Other VP genes involved in pathogenicity

In addition of the TTS system, the VP carries several

genes involved in pathogenicity. icsA (virG) encodes the
outer membrane protein directly responsible for pro-
moting the movement of intracellular bacteria. IcsA
binds the Neural Wiskott-Aldrich syndrome protein
(N-WASP), leading to the formation of a complex with
Arp2/3 that induces actin polymerization

[61,62]

. IcsA is

asymmetrically distributed on the bacterial surface and
enriched at the old pole of the bacterium

[63]

. A certain

proportion of IcsA is released into the culture medium
following cleavage by the VP-encoded outer membrane
protease IcsP (SopA)

[64,65]

. Inactivation of virK led

C. Parsot / FEMS Microbiology Letters 252 (2005) 11–18

15

background image

to a decreased production of IcsA by an unknown mech-
anism

[66]

. The sepA gene encodes a secreted serine pro-

tease of the IgA protease family that, like IcsA, is
transported across the outer membrane by an autotrans-
porter C-terminal domain. Inactivation of sepA in S.
flexneri 5 led to an attenuated phenotype upon infection
of rabbit ileal loops

[67]

. The absence of sepA in other S.

flexneri strains

[20]

suggests that a protein functionally

equivalent to SepA might be encoded by the chromo-
some or the VP in some strains. The VP msbB2 gene
and the chromosomal msbB1 gene encode acyl transfer-
ases that catalyze the acyl–oxyacyl linkage of a myristate
on the hydroxy-myristate attached to the 3

0

position of

glucosamide disaccharide in lipid A. This modification
contributes to lipid A toxicity. Both MsbB1 and MsbB2
are required for the modification of all lipid A molecules
and mutants in either msb1 or msbB2 are attenuated

[68]

.

9. Concluding remarks

Invasion of the colonic epithelium by bacteria in-

volves various interactions between bacteria, both extra-
and intracellular, and different cell types. Moreover,
these interactions, their outcomes and their conse-
quences are evolving during infection, as cell popula-
tions are changing, in nature, relative numbers and
states of activation. Likewise, external signals influence
repertoires of bacterial genes that are expressed by ex-
tra- and intracellular bacteria. Thus, infection is far
more complex than what is studied using homogenous
populations interacting in vitro. Nevertheless, in vitro
models of infection, imperfect as they are, have been
essential in deciphering some events that take place dur-
ing infection. In many cases, in vitro results have been
validated using animal models, mostly infection of li-
gated loops in rabbits and lungs in mice. However, we
still have much to learn on the functions of bacterial vir-
ulence factors and the responses of infected cells.

Comparison of S. flexneri 2a and E. coli K-12 gen-

omes indicates that many genes are inactivated or miss-
ing in S. flexneri strains. Inactivation of these genes
suggests that their products were detrimental to invasive
bacteria. Indeed, introduction of some E. coli K-12
genes, such as ompT and cadA, into S. flexneri interfered
with some virulence traits

[69,70]

. Acquisition, or assem-

bly, of the VP allowed an ancestral E. coli strain to
explore a new environment, but exploiting this ‘‘oppor-
tunity’’ required numerous adjustments of the chromo-
some. Understanding why these genes have been
inactivated might shed light on metabolic constraints
imposed by an intracellular lifestyle. On the other hand,
we start to have ideas on functions of genes carried by
pathogenicity islands and the VP. However, as exempli-
fied by stx genes present only in S. dysenteriae 1 strains

or by orf81 present on pWR100 and inactivated on
pCP301, genes specific to particular strains or isolates
might not be of general relevance to the invasive behav-
iour. Ongoing sequencing projects for S. dysenteriae and
S. sonnei strains should lead to a better evaluation of
genes that are conserved, or inactivated, between mem-
bers of Shigella spp. and, likewise, comparison of several
VP should help to understand the history of the VP. The
TTS system is central to the pathogenicity of Shigella
spp. and EIEC strains, being essential for entry of bac-
teria in epithelial cells and killing of infected macro-
phages and dendritic cells. The large repertoire of
effectors suggests that the TTS system is used for addi-
tional functions, possibly to modulate cell responses in-
duced by invading bacteria

[71]

. The differential

regulation of genes encoding TTS effectors suggests that
different effectors might be required at different times
following contact of bacteria with host cells. Elucidating
the function of these effectors and their role(s) at various
stages of infection is a challenging prospect.

References

[1] Kotloff, K.L., Winickoff, J.P., Ivanoff, B., Clemens, J.D., Swerd-

low, D.L., Sansonetti, P.J., Adak, G.K. and Levine, M.M. (1999)
Global burden of Shigella infections: implications for vaccine
development and implementation of control strategies. Bull.
World Health Organization 77, 651–666.

[2] Lan, R.T. and Reeves, P.R. (2002) Escherichia coli in disguise:

molecular origins of Shigella. Microb. Infect. 4, 1125–1132.

[3] Nhieu, G.T.V., Bourdet-Sicard, R., Dumenil, G., Blocker, A. and

Sansonetti, P.J. (2000) Bacterial signals and cell responses during
Shigella entry into epithelial cells. Cell Microbiol. 2, 187–193.

[4] Girardin, S.E., Tournebize, R. and Mavris, M., et al. (2001)

CARD4/Nod1 mediates NF-kappa B and JNK activation by
invasive Shigella flexneri. EMBO Rep. 2, 736–742.

[5] Girardin, S.E., Boneca, I.G. and Carneiro, L.A.M., et al. (2003)

Nod1 detects a unique muropeptide from Gram-negative bacterial
peptidoglycan. Science 300, 1584–1587.

[6] Pedron, T., Thibault, C. and Sansonetti, P.J. (2003) The invasive

phenotype of Shigella flexneri directs a distinct gene expression
pattern in the human intestinal epithelial cell line Caco-2. J. Biol.
Chem. 278, 33878–33886.

[7] Phalipon, A. and Sansonetti, P.J. (2003) Shigellosis: Innate

mechanisms of inflammatory destruction of the intestinal epithe-
lium, adaptive immune response, and vaccine development. Crit.
Rev. Immunol. 23, 371–401.

[8] Sansonetti, P.J. (2004) War and peace at mucosal surfaces. Nat.

Rev. Immunol. 4, 953–964.

[9] Philpott, D.J., Girardin, S.E. and Sansonetti, P.J. (2001) Innate

immune responses of epithelial cells following infection with
bacterial pathogens. Curr. Opin. Immunol. 13, 410–416.

[10] Jin, Q., Yuan, Z.H. and Xu, J.G., et al. (2002) Genome sequence

of Shigella flexneri 2a: insights into pathogenicity through
comparison with genomes of Escherichia coli K12 and O157.
Nucleic Acids Res. 30, 4432–4441.

[11] Wei, J., Goldberg, M.B. and Burland, V., et al. (2003) Complete

genome sequence and comparative genomics of Shigella flexneri
serotype 2a strain 2457T. Infect. Immun. 71, 4223.

[12] Runyen-Janecky, L.J., Reeves, S.A., Gonzales, E.G. and Payne,

S.M. (2003) Contribution of the Shigella flexneri Sit, Iuc, and Feo

16

C. Parsot / FEMS Microbiology Letters 252 (2005) 11–18

background image

iron acquisition systems to iron acquisition in vitro and in
cultured cells. Infect. Immun. 71, 1919–1928.

[13] Al-Hasani, K., Henderson, I.R., Sakellaris, H., Rajakumar, K.,

Grant, T., Nataro, J.P., Robins-Browne, R. and Adler, B. (2000)
The sigA gene which is borne on the she pathogenicity island of
Shigella flexneri 2a encodes an exported cytopathic protease
involved in intestinal fluid accumulation. Infect. Immun. 68,
2457–2463.

[14] West, N.P., Sansonetti, P. and Mounier, J., et al. (2005)

Optimization of virulence functions through glucosylation of
Shigella LPS. Science 307, 1313–1317.

[15] Ingersoll, M.A., Moss, J.E., Weinrauch, Y., Fisher, P.E., Grois-

man, E.A. and Zychlinsky, A. (2003) The ShiA protein encoded
by the Shigella flexneri SHI-2 pathogenicity island attenuates
inflammation. Cell Microbiol. 5, 797–807.

[16] Prosseda, G., Falconi, M., Giangrossi, M., Gualerzi, C.O.,

Micheli, G. and Colonna, B. (2004) The virF promoter in Shigella:
more than just a curved DNA stretch. Mol. Microbiol. 51, 523–
537.

[17] Durand, J.M.B. and Bjork, G.R. (2003) Putrescine or a combi-

nation of methionine and arginine restores virulence gene
expression in a tRNA modification-deficient mutant of Shigella
flexneri: a possible role in adaptation of virulence. Mol. Micro-
biol. 47, 519–527.

[18] Buchrieser, C., Glaser, P., Rusniok, C., Nedjari, H., dÕHauteville,

H., Kunst, F., Sansonetti, P. and Parsot, C. (2000) The virulence
plasmid pWR100 and the repertoire of proteins secreted by the
type III secretion apparatus of Shigella flexneri. Mol. Microbiol.
38, 760–771.

[19] 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.

[20] Lan, R.T., Stevenson, G. and Reeves, P.R. (2003) Comparison of

two major forms of the Shigella virulence plasmid pINV: positive
selection is a major force driving the divergence. Infect. Immun.
71, 6298–6306.

[21] Dorman, C.J. and Porter, M.E. (1998) The Shigella virulence gene

regulatory cascade: a paradigm of bacterial gene control mech-
anisms. Mol. Microbiol. 29, 677–684.

[22] Beloin, C. and Dorman, C.J. (2003) An extended role for the

nucleoid structuring protein H-NS in the virulence gene regula-
tory cascade of Shigella flexneri. Mol. Microbiol. 47, 825–838.

[23] Dorman, C.J., McKenna, S. and Beloin, C. (2001) Regulation of

virulence gene expression in Shigella flexneri, a facultative
intracellular pathogen. Int. J. Med. Microbiol. 291, 89–96.

[24] McKenna, S., Beloin, C. and Dorman, C.J. (2003) In vitro DNA-

binding properties of VirB, the Shigella flexneri virulence regu-
latory protein. FEBS Lett. 545, 183–187.

[25] Le Gall, T., Mavris, M., Martino, M.C., Bernardini, M.L.,

Denamur, E. and Parsot, C. (2005) Analysis of virulence plasmid
gene expression defines three classes of effectors in the type III
secretion system of Shigella flexneri. Microbiology 151, 951–962.

[26] Demers, B., Sansonetti, P.J. and Parsot, C. (1998) Induction of

type III secretion in Shigella flexneri is associated with differential
control of transcription of genes encoding secreted proteins.
EMBO J. 17, 2894–2903.

[27] Mavris, M., Page, A.L., Tournebize, R., Demers, B., Sansonetti,

P. and Parsot, C. (2002) Regulation of transcription by the
activity of the Shigella flexneri type III secretion apparatus. Mol.
Microbiol. 43, 1543–1553.

[28] Kane, C.D., Schuch, R., Day, W.A. and Maurelli, A.T. (2002)

MxiE regulates intracellular expression of factors secreted by the
Shigella flexneri 2a type III secretion system. J. Bacteriol. 184,
4409–4419.

[29] Penno, C., Sansonetti, P.J. and Parsot, C. (2005) Frameshifting by

transcriptional slippage is involved in production of MxiE, the

transcription activator regulated by the activity of the type III
secretion apparatus in Shigella flexneri. Mol. Microbiol. 56, 204–214.

[30] Mavris, M., Sansonetti, P.J. and Parsot, C. (2002) Identification

of the cis-acting site involved in activation of promoters regulated
by activity of the type III secretion apparatus in Shigella flexneri.
J. Bacteriol. 184, 6751–6759.

[31] Parsot, C., Ageron, E., Penno, C., Mavris, M., Jamoussi, K.,

dÕHauteville, H., Sansonetti, P. and Demers, B. (2005) A secreted
anti-activator, OspD1, and its chaperone, Spa15, are involved in
the control of transcription by the type IIII secretion apparatus
activity in Shigella flexneri. Mol. Microbiol. 56, 1627–1635.

[32] Blocker, A., Gounon, P., Larquet, E., Niebuhr, K., Cabiaux, V.,

Parsot, C. and Sansonetti, P. (1999) The tripartite type III
secretion of Shigella flexneri inserts IpaB and IpaC into host
membranes. J. Cell. Biol. 147, 683–693.

[33] Tamano, K., Aizawa, S., Katayama, E., Nonaka, T., Imajoh-

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. EMBO J. 19, 3876–3887.

[34] Blocker, A., Jouihri, N., Larquet, E., Gounon, P., Ebel, F.,

Parsot, C., Sansonetti, P. and Allaoui, A. (2001) Structure and
composition of the Shigella flexneri Ôneedle complexÕ, a part of its
type III secretion. Mol. Microbiol. 39, 652–663.

[35] Cordes, F.S., Komoriya, K., Larquet, E., Yang, S.X., Egelman,

E.H., Blocker, A. and Lea, S.M. (2003) Helical structure of the
needle of the type III secretion system of Shigella flexneri. J. Biol.
Chem. 278, 17103–17107.

[36] Jouihri, N., Sory, M.P., Page, A.L., Gounon, P., Parsot, C. and

Allaoui, A. (2003) MxiK and MxiN interact with the Spa47
ATPase and are required for transit of the needle components
MxiH and Mxil, but not of Ipa proteins, through the type III
secretion apparatus of Shigella flexneri. Mol. Microbiol. 49, 755–
767.

[37] Tamano, K., Katayama, E., Toyotome, T. and Sasakawa, C.

(2002) Shigella Spa32 is an essential secretory protein for
functional type III secretion machinery and uniformity of its
needle length. J. Bacteriol. 184, 1244–1252.

[38] Magdalena, J., Hachani, A., Chamekh, M., Jouihri, N., Gounon,

P., Blocker, A. and Allaoui, A. (2002) Spa32 regulates a switch in
substrate specificity of the type III secreton of Shigella flexneri
from needle components to Ipa proteins. J. Bacteriol. 184, 3433–
3441.

[39] Schuch, R. and Maurelli, A.T. (2001) Spa33, a cell surface-

associated subunit of the Mxi-Spa type III secretory pathway of
Shigella flexneri, regulates Ipa protein traffic. Infect. Immun. 69,
2180–2189.

[40] van der Goot, F.G., van Nhieu, G.T., Allaoui, A., Sansonetti, P.

and Lafont, F. (2004) Rafts can trigger contact-mediated secre-
tion of bacterial effectors via a lipid-based mechanism. J. Biol.
Chem. 279, 47792–47798.

[41] Page, A.L., Ohayon, H., Sansonetti, P.J. and Parsot, C. (1999)

The secreted IpaB and IpaC invasins and their cytoplasmic
chaperone IpgC are required for intercellular dissemination of
Shigella flexneri. Cell Microbiol. 1, 183–193.

[42] Schuch, R., Sandlin, R.C. and Maurelli, A.T. (1999) A system for

identifying post-invasion functions of invasion genes: require-
ments for the Mxi-Spa type III secretion pathway of Shigella
flexneri in intercellular dissemination. Mol. Microbiol. 34, 675–
689.

[43] Parsot, C., Hamiaux, C. and Page, A.L. (2003) The various and

varying roles of specific chaperones in type III secretion systems.
Curr. Opin. Microbiol. 6, 7–14.

[44] Ogawa, M., Suzuki, T., Tatsuno, I., Abe, H. and Sasakawa, C.

(2003) IcsB, secreted via the type III secretion system, is
chaperoned by IpgA and required at the post-invasion stage of
Shigella pathogenicity. Mol. Microbiol. 48, 913–931.

C. Parsot / FEMS Microbiology Letters 252 (2005) 11–18

17

background image

[45] 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 flexneri, is chaperoned by IpgE
and implicated in entry focus formation. Mol. Microbiol. 38, 8–
19.

[46] Page, A.L., Sansonetti, P. and Parsot, C. (2002) Spa15 of Shigella

flexneri, a third type of chaperone in the type III secretion
pathway. Mol. Microbiol. 43, 1533–1542.

[47] van Eerde, A., Hamiaux, C., Perez, J., Parsot, C. and Dijkstra,

B.W. (2004) Structure of Spa15, a type III secretion chaperone
from Shigella flexneri with broad specificity. EMBO Rep. 5, 477–
483.

[48] Page, A.L., Fromont-Racine, M., Sansonetti, P., Legrain, P. and

Parsot, C. (2001) Characterization of the interaction partners of
secreted proteins and chaperones of Shigella flexneri. Mol.
Microbiol. 42, 1133–1145.

[49] De Geyter, C., Wattiez, R., Sansonetti, P., Falmagne, P.,

Ruysschaert, J.M., Parsot, C. and Cabiaux, V. (2000) Character-
ization of the interaction of IpaB and IpaD, proteins required for
entry of Shigella flexneri into epithelial cells, with a lipid
membrane. Eur. J. Biochem. 267, 5769–5776.

[50] Hume, P.J., McGhie, E.J., Hayward, R.D. and Koronakis, V.

(2003) The purified Shigella IpaB and Salmonella SipB translo-
cators share biochemical properties and membrane topology.
Mol. Microbiol. 49, 425–439.

[51] Skoudy, A., Mounier, J., Aruffo, A., Ohayon, H., Gounon, P.,

Sansonetti, P. and Nhieu, G.T.V. (2000) CD44 binds to the
Shigella IpaB protein and participates in bacterial invasion of
epithelial cells. Cell Microbiol. 2, 19–33.

[52] Hayward, R.D., Cain, R.J., McGhie, E.J., Phillips, N., Garner,

M.J. and Koronakis, V. (2005) Cholesterol binding by the
bacterial type III translocon is essential for virulence effector
delivery into mammalian cells. Mol. Microbiol. 56, 590–603.

[53] Hilbi, H., Moss, J.E. and Hersh, D., et al. (1998) Shigella-induced

apoptosis is dependent on Caspase-1 which binds to IpaB. J. Biol.
Chem. 273, 32895–32900.

[54] Picking, W.L., Nishioka, H., Hearn, P.D., Baxter, M.A.,

Harrington, A.T., Blocker, A. and Picking, W.D. (2005) IpaD
of Shigella flexneri is independently required for regulation of Ipa
protein secretion and efficient insertion of IpaB and IpaC into
host membranes. Infect. Immun. 73, 1432–1440.

[55] Van Nhieu, G.T., Caron, E., Hall, A. and Sansonetti, P.J. (1999)

IpaC induces actin polymerization and filopodia formation during
Shigella entry into epithelial cells. EMBO J. 18, 3249–3262.

[56] Niebuhr, K., Giuriato, S. and Pedron, T., et al. (2002) Conver-

sion of PtdIns(4,5)P-2 into PtdIns(5)P by the S. flexneri effector
IpgD reorganizes host cell morphology. EMBO J. 21, 5069–5078.

[57] Bourdet-Sicard, R., Rudiger, M., Jockusch, B.M., Gounon, P.,

Sansonetti, P.J. and Van Nhieu, G.T. (1999) Binding of the
Shigella protein IpaA to vinculin induces F-actin depolymeriza-
tion. EMBO J. 18, 5853–5862.

[58] Ohya, K., Handa, Y., Ogawa, M., Suzuki, M. and Sasakawa, C.

(2005) IpgB1 is a novel Shigella effector protein involved in
bacterial invasion of host cells: its activity promotes membrane
ruffing via Rac1 and Cdc42 activation. J. Biol. Chem. 280, 24022–
24034.

[59] Yoshida, S., Katayama, E., Kuwae, A., Mimuro, H., Suzuki, T. and

Sasakawa, C. (2002) Shigella deliver an effector protein to trigger
host microtubule destabilization, which promotes Rac1 activity and
efficient bacterial internalization. EMBO J. 21, 2923–2935.

[60] Ogawa, M., Yoshimori, T., Suzuki, T., Sagara, H., Mizushima,

N. and Sasakawa, C. (2005) Escape of intracellular Shigella from
autophagy. Science 307, 727–731.

[61] 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.

[62] Egile, C., Loisel, T.P., Laurent, V., Li, R., Pantaloni, D.,

Sansonetti, P.J. and Carlier, M.F. (1999) Activation of the
CDC42 effector N-WASP by the Shigella flexneri IcsA protein
promotes actin nucleation by Arp2/3 complex and bacterial actin-
based motility. J. Cell. Biol. 146, 1319–1332.

[63] 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. Micro-
biol. 41, 861–872.

[64] Egile, C., dÕHauteville, H., Parsot, C. and Sansonetti, P.J. (1997)

SopA, the outer membrane protease responsible for polar locali-
zation of IcsA in Shigella flexneri. Mol. Microbiol. 23, 1063–1073.

[65] Shere, K.D., Sallustio, S., Manessis, A., DAversa, T.G. and

Goldberg, M.B. (1997) Disruption of IcsP, the major Shigella
protease that cleaves IcsA, accelerates actin-based motility. Mol.
Microbiol. 25, 451–462.

[66] Nakata, N., Sasakawa, C., Okada, N., Tobe, T., Fukuda, I.,

Suzuki, T., Komatsu, K. and Yoshikawa, M. (1992) Identification
and characterization of VirK, a virulence associated large plasmid
gene essential for intercellular spreading of Shigella flexneri. Mol.
Microbiol. 6, 2387–2395.

[67] Benjelloun-Touimi, Z., Sansonetti, P.J. and Parsot, C. (1995)

SepA, the major extracellular protein of Shigella flexneri: auton-
omous secretion and involvement in tissue invasion. Mol.
Microbiol. 17, 123–135.

[68] dÕHauteville, H., Khan, S. and Maskell, D.J., et al. (2002) Two

msbB genes encoding maximal acylation of lipid A are required
for invasive Shigella flexneri to mediate inflammatory rupture and
destruction of the intestinal epithelium. J. Immunol. 168, 5240–
5251.

[69] Maurelli, A.T., Fernandez, R.E., Bloch, C.A., Rode, C.K. and

Fasano, A. (1998) ‘‘Black holes’’ and bacterial pathogenicity: a
large genomic deletion that enhances the virulence of Shigella spp.
and enteroinvasive Escherichia coli. Proc. Natl. Acad. Sci. USA
95, 3943–3948.

[70] Nakata, N., Tobe, T., Fukuda, I., Suzuki, T., Komatsu, K.,

Yoshikawa, M. and Sasakawa, C. (1993) The absence of a surface
protease, OmpT, determines the intercellular spreading ability of
Shigella – the relationship between the ompT and kcpA loci. Mol.
Microbiol. 9, 459–468.

[71] Kim, D.W., Lenzen, G., Page, A.-L., Legrain, P., Sansonetti, P.J.

and Parsot, C. (in press) The Shigella flexneri effector OspG
interferes with innate immune responses by targeting ubiquitin-
conjugating enzymes. Proc. Natl. Acad. Sci. USA 102.

18

C. Parsot / FEMS Microbiology Letters 252 (2005) 11–18


Document Outline


Wyszukiwarka

Podobne podstrony:
Identification of enteroinvasive E coli and Shigella strains
A protocol for polymerase chain reaction detection of Enterococcus faecalis and Enterococcus faec
Shigella flexneri infection pathogenesis and vaccine development
Legg Perthes disease in three siblings, two heterozygous and one homozygous for the factor V Leiden
Candida glabrata review of epidemiology and pathogenesis
Variations in Risk and Treatment Factors Among Adolescents Engaging in Different Types of Deliberate
The Bloodborne Pathogens Standard and Disinfection
Rapid and efficient purification and refolding of a (His) tagged recombinant protein produced in E c
Retrospective Analysis of Social Factors and Nonsuicidal Self Injury Among Young Adults
Resilience and Risk Factors Associated with Experiencing Childhood Sexual Abuse
Phylogeny of the enterobacteriaceae based on genes encoding elongation factor Tu
Tea consumption and cardiovascular risk in the SU VI MAX study Are life style factors important
Shigellosis from symptoms to molecular pathogenesis
Shigellosis challenges and management issues
Modification of Intestinal Microbiota and Its Consequences for Innate Immune Response in the Pathoge
Charlie And The Chocolate Factory Verucasalt Danny Elfman
Developmental protective and risk factors in bpd (using aai)
Pathogenesis and antifungal drug resistance of C glabrata

więcej podobnych podstron