ArticlePDF AvailableLiterature Review

Virulence Factors in Salmonella Typhimurium: The Sagacity of a Bacterium

Authors:

Abstract and Figures

Currently, Salmonella enterica Typhimurium (ST) is responsible for most cases of food poisoning in several countries. It is characterized as a non-specific zoonotic bacterium that can infect both humans and animals and although most of the infections caused by this microorganism cause only a self-limiting gastroenteritis, some ST strains have been shown to be invasive, crossing the intestinal wall and reaching the systemic circulation. This unusual pathogenicity ability is closely related to ST virulence factors. This review aims to portray the main virulence factors in Salmonella Typhimurium, in order to better understand the strategies that this pathogen uses to reach the systemic circulation and increase its infectivity in humans and animals. Thus, the most studied Salmonella pathogenicity islands in Salmonella Typhimurium were detailed as to the functions of their encoded virulence factors. In addition, available knowledge on virulence plasmid was also compiled, as well as the chromosome regions involved in the virulence of this bacterium.
This content is subject to copyright. Terms and conditions apply.
Vol.:(0123456789)
1 3
Current Microbiology
https://doi.org/10.1007/s00284-018-1510-4
REVIEW ARTICLE
Virulence Factors inSalmonella Typhimurium: The Sagacity
ofaBacterium
AnamariaM.P.dosSantos1· RafaelaG.Ferrari1,2· CarlosA.Conte‑Junior1,2,3
Received: 4 October 2017 / Accepted: 16 May 2018
© Springer Science+Business Media, LLC, part of Springer Nature 2018
Abstract
Currently, Salmonella enterica Typhimurium (ST) is responsible for most cases of food poisoning in several countries. It is
characterized as a non-specific zoonotic bacterium that can infect both humans and animals and although most of the infec-
tions caused by this microorganism cause only a self-limiting gastroenteritis, some ST strains have been shown to be invasive,
crossing the intestinal wall and reaching the systemic circulation. This unusual pathogenicity ability is closely related to
ST virulence factors. This review aims to portray the main virulence factors in Salmonella Typhimurium, in order to better
understand the strategies that this pathogen uses to reach the systemic circulation and increase its infectivity in humans and
animals. Thus, the most studied Salmonella pathogenicity islands in Salmonella Typhimurium were detailed as to the func-
tions of their encoded virulence factors. In addition, available knowledge on virulence plasmid was also compiled, as well
as the chromosome regions involved in the virulence of this bacterium.
Introduction
Salmonella spp. represents one of the principal causes of
food poisoning in several countries in the last 100years [72].
Approximately 16million cases of typhoid fever, 1.3bil-
lion cases of gastroenteritis, and 3million deaths involving
this bacterium are reported annually worldwide [83]. Due
to its endemicity, high morbidity and, especially, difficulty
in applying control and prevention measures, salmonellosis
is characterized as an important zoonosis in a public health
context [90].
The genus Salmonella, belonging to the Enterobac-
teriaceae family, includes two species, S. enterica and S.
bongori [90]. The first comprises six subspecies designated
by Roman numerals [25], and includes more than 2600
serotypes that differ from each other by their flagellar (H)
and somatic (O) structures [17]. S. enterica subspecies I
(enterica) is the most isolated subspecies in animals, and
is found in 99% of human isolates. S. bongori, on the other
hand, is often found in “cold-blooded” animals, such as rep-
tiles, amphibians, and fish, and accounts for less than 1% of
human isolates [83].
Diseases related to Salmonella enterica can be divided
into three groups: Typhoid fever, caused by S. Typhi; enteric
fever, caused by S. Paratyphi A, B, and C; and enterocolitis
or salmonellosis, caused by the other serotypes [90]. Accord-
ing to Connor and Schwartz [19], because their symptoms
are indistinguishable in a clinical analysis, both S. Typhi and
S. Paratyphi are classified as causative agents for typhoid
fever. These bacteria have human and some larger primates
as their only reservoirs [33]. Symptoms are characterized
by abdominal pain, headaches, fever, and diarrhea that may
appear within a week of incubation [25]. In contrast, sal-
monellosis affects both human and animals and is the main
cause of gastroenteritis in humans [1], resulting in 93.8mil-
lion cases, with 155,000 deaths each year [25].
Among these bacteria, the most frequently isolated
serotypes are S. enterica Typhimurium (ST) and S.
enterica Enteritidis (SE) [35]. In Brazil, ST was the most
isolated serotype up to the mid-1990s, being surpassed
after this by SE [1]. ST is a general pathogen, and can
be isolated from different animals, as well as different
foodstuffs. Thus, this serotype can be found in a variety
of foods, and is mainly present in poultry [70, 103], swine
* Carlos A. Conte-Junior
carlosconte@id.uff.br
1 Department ofFood Technology, Faculty ofVeterinary,
Molecular & Analytical Laboratory Center, Universidade
Federal Fluminense, Niterói, Brazil
2 Chemistry Institute, Food Science Program, Universidade
Federal doRio de Janeiro, RiodeJaneiro, Brazil
3 National Institute ofHealth Quality Control, Fundação
Oswaldo Cruz, RiodeJaneiro, Brazil
A.M.P.dos Santos et al.
1 3
[75, 95], and bovine meat [84]. However, some ST strains,
in addition to causing localized gastroenteritis, are capa-
ble of causing systemic infections, such as the case of the
multi-locus sequence type 313 (ST313) isolated in sub-
Saharan Africa [61], and the most commonly found ST19
[12]. These bacteria multiply in the small intestine and
extend beyond the intestinal wall, reaching the mesenteric
lymph nodes and from there, the liver and spleen, where
they settle and multiply [44].
It is believed that such an ability was acquired from
horizontal gene transfer (transduction, conjugation, or
transformation) over millions of years [33]. An example
for this reasoning is the fact that ST and S. Typhi share
about 90% of the genes present in their DNA [57]. This
similarity allows ST to be used as a model for the study
of typhoid fever in animals, since S. Typhi is human-
specific [88]. The remaining 10% include virulence fac-
tors, defined as structures, products, and strategies that
contribute to infection and that determine the degree of
pathogenicity of this serotype [57]. In ST, most of the
genes that encode such factors are located in the so-called
Salmonella Pathogenicity Islands (SPI) [76]. Although
less relevant, they may also be present in other parts of
the chromosome, such as the fimbriae and flagella. The
genes encoding these characteristics can be found in viru-
lence plasmids (pSLT), more precisely in the spv operon
[26].
In this context, this review aims to portray the main
virulence factors in Salmonella Typhimurium in order to
better understand what strategies this pathogen uses to
reach the systemic circulation and increase its infectivity
in humans and animals (Table1).
Overview oftheInfection Process
Infection caused by ST occurs mainly through the inges-
tion of contaminated animal or water food [11, 76]. Upon
reaching the stomach, this microorganism must cope with
the acidity of the medium, so its acid tolerance response
(ATR) is activated, which maintains the intracellular pH
higher than the extracellular pH, allowing the bacterium to
survive while in this environment [26]. Subsequently, ST
crosses the mucus layer present in the intestinal wall and
adheres to the epithelium [7, 48], where the infection will
occur. The interaction of ST with the epithelium results in
the appearance of the clinical diagnosis, characterized by
diarrhea, culminating in the loss of electrolytes and inducing
local inflammation of the intestine [46].
ST adhesion to the epithelium is achieved by host–recep-
tor interactions with many of the adhesion factors present
on the cell surface of this microorganism [76]. After this
stage, effector proteins are released into the enterocyte cyto-
plasm, causing changes in the cytoskeleton of the intestinal
epithelium [76]. These modifications lead to the formation
of membrane extensions, known as ruffles, similar to those
formed by phagocytic cells [25]. The ruffles encompass ST
and launch it into the cell [11]. With the engulfment of the
bacterium by enterocytes, intracellular phagosomal compart-
ments, termed Salmonella-Containing Vacuoles (SCV), are
formed [24, 31, 44, 46, 50, 67] (Fig.1). This compartment
is the only place in these host cells where ST the can survive
and multiply [48]. As a consequence of this invasion, tran-
scription activators are generated, leading to the production
of proinflammatory cytokines, such as interleukin (IL)-8,
which ultimately induce the inflammatory response [46].
With this induction, polymorphonuclear leucocytes (PMN)
Table 1 Components of T3SS-1
and T3SS-2, substructures of
its components (if any), and
structural proteins that comprise
those components
Components Substructures Protein compo-
nents T3SS-1
Protein compo-
nents T3SS-2
References
Needle complex Needle structure PrgI SsaG [23, 60]
PrgJ SsaI
Needle complex Inner membrane rings PrgK SsaJ [23, 60]
PrgH SsaD
Outer membrane rings InvG SsaC
Export apparatus SpaO SsaQ [23, 60]
SpaP SsaR
SpaQ SsaS
SpaR SsaT
SpaS SsaU
InvA SsaV
InvC SsaN
OrgA –
Translocon SipBCD SseCDB [23, 60]
Virulence Factors inSalmonella Typhimurium: The Sagacity ofaBacterium
1 3
migrate into the intestine and induce the production of anti-
microbial substances, where apparently ST is less affected
[52]. This generates a microenvironment in which ST has a
growth advantage over the resident microbiota, which is also
affected by the substances released by PMNs [7].
Jones etal. [56] noted that salmonellae adhere to and
preferentially invade M cells, which are modified cells of
the intestinal wall of the ileum present in Peyer’s plaques.
M cells are intimately associated with macrophages, which
reside in this tissue [53]. When the epithelial barrier is rup-
tured, ST can invade macrophages and dendritic cells and,
from these infected cells, gain systemic circulation [52].
During all stages of this infection, ST uses specific virulence
factors to conduct a more efficient invasion (Table2). In this
context, we will now report these factors and their roles in
infection caused by Salmonella Typhimurium.
Strategies andVirulence Factors
Type III Secretion System (T3SS)
Salmonella spp. possesses a molecular apparatus called the
Type III Secretion System (T3SS) [17, 44], which is respon-
sible for the ability of these microorganisms to inject effector
proteins into the cytosol of host cells and modulate host cells
signaling cascades for the benefit of bacteria [17, 57]. These
effectors, once within the cell, can alter cellular functions
such as cytoskeleton structure, membrane transport, signal
transduction, and cytokine expression [48]. These changes
allow for the invasion and permanence of the bacterium in
the infected cell [17]. T3SS comprises several components
(Table1), including more than twenty proteins [48], some of
them are homologous to those involved in flagellar assembly
[77], suggesting a evolutionary relation [66].
Salmonella enterica abstract two distinct T3SSs (T3SS-1
and T3SS-2), which are encoded by SPI-1 and SPI-2, respec-
tively, located in different Salmonella DNA clusters [46].
They have shown to function at different times during the
infection [46]. Thereby, T3SS-1 is active with the contact
of the bacterium with the host cell membrane and translo-
cates effector proteins into its cytoplasm, while T3SS-2 is
active within the phagosome and translocates effectors into
the vacuolar space [48]. This ability could be very important
for systemic increase of ST in its host, once T3SS-2 secreted
genes are required for bacterial growth in macrophages [16,
51, 82]. Most of the effectors translocated by T3SS-1 are
encoded on SPI-1, but, as we will show in this article, some
of them can be encoded on SPI- 5 [46], as well as some also
can be encoded in other parts of ST DNA [77]. Just like
T3SS-1, T3SS-2 translocates effectors which are encoded in
other parts of ST DNA [99], despites most of the effectors
Fig. 1 Pathogenesis of Salmonella Typhimurium. a Salmonella
adheres to the intestinal epithelial and M cells using many of adhe-
sions factors present on its cell surface. b, c Effector proteins are
released into enterocyte causing changes on its cytoskeleton and
forming structures in its surface known as ruffles. d Alternatively, the
bacterial cells can be directly taken by dendritic cell from the sub-
mucosa. e Once inside cytoplasm, Salmonella cells are located into
SCV (Salmonella Containing Vacuoles), where it multiplies. f The
SCV transcytose to the basolateral membrane and release to the sub-
mucosa. g Bacteria is internalized within phagocytes and then located
again into SCV; this figure was based on the one illustrated in the
article [89]
A.M.P.dos Santos et al.
1 3
translocated for this system is encoded on SPI-2 [51].
Although T3SS-1 and T3SS-2 have similar structures and
perform the same function (translocate effectors proteins),
these systems are not identical themselves, since both have
different structural proteins and, as said, secrete different
effector proteins [99]. Their structural proteins, however,
share some homology (Table1), which makes their mecha-
nisms not seem significantly different [71]. It is believed,
though, that they are not independent one of another, once
mutations in T3SS-2 causes a significant reduction in the
expression of some genes of T3SS-1, which harms the
ability of the bacterium to invade epithelial cells [22]. In
view of its homology, we prefer to use T3SS-1 of ST as
an archetype to prevent confusion and since T3SS-1 of ST
is, till now, the most well-characterized system [20, 59, 60,
66, 67]. The corresponding components of T3SS-2 can be
found in Table1.
Kubori etal. [61], after observing the structure of T3SS-1
apparatus under optical microscopy, has found a supramo-
lecular structure of ST T3SS-1 apparatus called the Nee-
dle Complex (NC) [48]. The NC is, by definition, a hollow
structure composed by two pairs of rings that are anchored
Table 2 Main virulence factors, DNA location, and main functions
a SPI Salmonella pathogenicity island
b SIF Salmonella-induced filaments
c SCV Salmonella-containing vacuoles
Virulence factor Location Function Reference
SipA SPI-1aCytoskeleton rearrangement [104]
Chemotaxis [76]
SipB SPI-1 Translocation of effector proteins [30]
Macrophage apoptosis impairment [18]
SipC SPI-1 Chemotaxis [76]
Cytoskeleton rearrangement
SptP SPI-1 Suppression of innate immunity [14]
trr genes SPI-2 Production of tetrathionate reductase [26]
SpiC SPI-2 Translocation of effector proteins [28]
Survival within SCVc[98]
SseB SPI-2 Formation of macromolecular structures which serves as a translocon [70]
SseC SPI-2 Formation of macromolecular structures which serves as a translocon [70]
SseD SPI-2 Formation of macromolecular structures which serves as a translocon [70]
SseF SPI-2 SCV perinuclear migration [68, 69]
Microtubule aggregation
SIF formationb
SseG SPI-2 SCV perinuclear migration [68, 69]
Microtubule aggregation
SIF formation
MisL SPI-3 Long-term persistence [24]
MgtCB SPI-3 Survival within macrophages [5]
MarT SPI-3 Activation of MisL expression [85]
SiiE SPI-4 Adhesion to the epithelium [34]
SopB SPI-5 Prevents apoptosis of epithelial cells [64]
SigE SPI-5 Chaperone [21]
SpvR pSLT Regulation of spv genes [38]
SpvB pSLT Prevents actin polymerization [38]
SpvC pSLT Inhibits MAP kinase and immune signaling [91]
Type I Fimbrae Chromosome Adhesion to the epithelium [3]
SifA Chromosome SIF formation [4]
SCV maintenance
SseJ Chromosome SIF formation [69]
SopE Chromosome Induce membrane ruffling in cell cultures [47]
SopE2 Chromosome Induce membrane ruffling in cell cultures [47]
Virulence Factors inSalmonella Typhimurium: The Sagacity ofaBacterium
1 3
to the inner and the outer membranes of the bacterium, and
a needle-like structure that extends beyond the surface of
the bacterium [59] (Fig.2). A NC is composed of at least
five proteins components: PrgI, J, K, H, and InvG [60], each
one of them is required for the type III secretion and the
invasion of epithelial cells [59]. PrgI was demonstrated to
be component of the needle portion of the NC, since ST prgI
mutant strain did not present needle substructures, although
it did exhibit apparently normal bases [67]. Beyond that, the
same study has showed that PrgI is essential for ST entry
into host cells, once prgI mutant was completely defective
in its ability to invade cultured intestinal cells [67]. PrgJ is
required for PrgI secretion as well as the formation of the
needle structure, and it is also the minor component of the
needle itself [60]. PrgH and PrgK, on the other hand, interact
with each other to form the inner membrane rings of the NC
[59]. Lastly, InvG, which is a member of the secretin fam-
ily of protein [67], forms the outer membrane rings of the
NC, which are believed to permit the passage of exported
substrates across the outer membrane [60]. Housed within
the inner membrane rings of the NC base, there is a Type III
Export Apparatus which is required to assembly of the NC
and the secretion of effector proteins [94]. This apparatus is
believed to be composed by eight proteins (SpaO, P, Q, R, S,
InvA, C, and OrgA), which are conserved among all T3SS
[55]. A few sets of other proteins (as SipBCD, explained
in the following section) comprise a Translocon, which is
believed to be involved in the translocation of effector pro-
teins into the host cytoplasm by forming a translocation pore
in the host cellular membrane [48, 60].
Salmonella Pathogenicity Islands
Salmonella pathogenicity islands (SPIs) are chromosomal
regions that carry virulence genes that act as a compact and
distinct genetic unit known as an operon [77]. These regions
display a different composition from the rest of the chromo-
some, with the characteristic presence of a greater amount
of G+C (guanine and cytosine) when compared with the
other parts of the DNA [79]. To date, five SPIs have been
described in ST, with SPI-1 and SPI-2 being the most rec-
ognized and studied [2, 17, 25, 44, 77]. Although the other
three SPIs have not been studied as closely as the first two
[26], studies have reported the involvement of these islands
in Salmonella invasion of and survival within host cells [11].
SPI-1 is the most well-characterized island among the
five ST SPIs [26, 33, 77]. Genes encoded by this region are
said to be essential in the invasion stage through the intesti-
nal epithelium [77], since salmonellae SPI-1 mutants have
shown an attenuation for systemic invasion when inoculated
orally in rats. However, this does not occur when inoculation
is performed by the parenteral route [31]. This suggests that
these genes are closely related to the host’s internalization
of the bacterium and crucial for ST to penetrate the intes-
tinal wall and consolidate the invasion [31]. SipABCD and
SptP are the major effector proteins encoded by these genes,
released into the host cell through T3SS-1.
SipA is said to be one of the first proteins responsible
for the induction of cytoskeletal rearrangement of epithelial
cells, which facilitates bacteria entry into the host cell [104].
This has been confirmed by invitro studies with ST sipA
mutants, in which it was demonstrated that such a microor-
ganism was less effective in inducing actin-cytoskeletal reor-
ganization in HeLa cells, in the same way that they were pre-
vented from being internalized by these cells [104]. As with
SipA, SipC is also responsible for inducing actin-cytoskel-
etal rearrangements [76] and, alongside SipB, is involved
in T3SS-1 translocation of effector molecules (e.g., SptP)
into the host cell [18]. Studies conducted with sipB, sipC
sipD mutants have shown the inability of these strains to
stimulate any response in the T3SS-1-dependent cell [18]. In
addition, SipA and SipC also activate a signal transduction
Fig. 2 T3SS-1 structure and its components. A Translocon, which
forms a translocation pore in the host cell membrane and translo-
cates effector proteins into the host cytoplasm, formed by SipB, C,
and D proteins; a needle structure, which extends beyond the surface
of the bacterium, formed by PrgI and PrgJ proteins; an outer mem-
brane rings, formed by InvG, located in outer membrane; and an
inner membrane rings, formed by PrgK and PrgH, located in the inner
membrane. In addition, an Export Apparatus, housed within the inner
membrane rings, formed by the proteins summarized in Table1; this
figure was based on the one illustrated in the article [31]
A.M.P.dos Santos et al.
1 3
cascade that leads to PMN migration [76], creating a micro-
environment favorable for ST growth [7]. The role of SipB
is still not fully elucidated [76]; however, Hersh etal. [53]
describe its ability to induce macrophage apoptosis from the
activation of caspase-1, and stated that Salmonella-induced
macrophage cytotoxicity is SipB-dependent.
Another protein encoded by genes located in SPI-1
is SptP. This effector is a phosphotyrosine phosphatase
which, when injected into epithelial cells, alters the actin
cytoskeleton [30]. Nevertheless, as demonstrated by Kaniga
etal. [58], SptP is not required for host cell invasion, since
sptP ST mutant strains enter cultured epithelial cells and
macrophages with the same efficiency as wild-type strains.
Likewise, sptP mutants did not display the same efficacy in
colonizing the liver of infected mice compared to wild-type
strains [58]. More recently, Choi etal. [14] reported the role
of SptP in suppressing the innate immunity of the host, dem-
onstrating that it is able to prevent mast cell degranulation,
which may be an important mechanism of ST virulence.
SPI-2 is a pathogenicity island that shows great impor-
tance in the maintenance of its permanence in the host cell.
This chromosome sequence contains about 40 genes [77],
grouped into four operons: ssa (type 3 secretion system), ssr
(secretion system regulation), ssc (chaperone secretion), and
sse (effector coding) [51].
SPI-2 is divided into two segments, one large and one
small [26]. The latter is characterized by the presence of trr
genes, which are involved in the production of tetrathion-
ate reductase, responsible for Salmonella-like tetrathionate
respiration [50] and seven open reading frames (ORFs) [26],
which, to date, have not demonstrated a significant role in
ST virulence [26]. On the other hand, the major follow-up
displays some importance in the ability of ST to survive
and multiply within SCV in host cells, either epithelial or
macrophages [82]. Fields etal. [27] reported that salmonel-
lae deficient in macrophage replication display attenuated
virulence, corroborating the idea that this capacity is essen-
tial for Salmonella pathogenesis. Similarly, Cirillo etal. [16]
have shown that SPI-2 mutants can colonize Peyer’s plaques,
but are unable to colonize the liver, spleen, or mesenteric
lymph nodes, which require invasion into macrophages. This
ability is given, mainly, by effector proteins SpiC, SseF, and
SseG that are translocated by T3SS-2.
Uchiya etal. [98] reported the first protein encoded
by SPI-2 and translocated via T3SS-2 to the macrophage
cytoplasm, SpiC. In their study, SpiC was classified as the
protein responsible for the inhibition of phagolysosome for-
mation (junction of lysosome-containing vacuoles with the
SCV). This mechanism enabled ST to be able to survive for
more than 24h within macrophages and dendritic cells [45].
Uchiya etal. [98] also demonstrated that spiC is of great
importance for ST virulence, since its deletion resulted in a
considerable attenuation of this characteristic. According to
Freeman etal. [28], SpiC is still required for the transloca-
tion of effector proteins inside host cells, such as SseB and
SseC.
Effector proteins SseB, SseC, and SseD have been shown
to be secreted by T3SS-2 when ST is cultured invitro under
SPI-2-inducing conditions [63]. These same proteins are
found on the cell surface of the bacterium and are necessary
for translocation of other effector proteins to the infected cell
[80]. They are believed to form a macromolecular structure
in the membrane that serves as a translocon [80]. In addi-
tion, SseC and SseD have also been shown to be essential for
virulence, since sseC and sseD mutants display a decrease
in this characteristic [63].
Hensel etal. [51], when analyzing sseF and sseG mutants,
noted that these genes play a role in systemic pathogenesis
and intracellular proliferation. In addition, Kuhle and Hensel
[68] demonstrated that sseF and sseG may be related to the
induction of Salmonella-induced filaments (SIF), since their
exclusion leads to a considerable decrease in the formation
of these structures when compared to wild-type strains. SIF
are structures that arise when, a few hours after internaliza-
tion, SCV begin to elongate into tubular filaments [8]. Their
formation involves the fusion of late endocytic compart-
ments [10] and, although this is clear, the mechanisms that
mediate this process are still unknown [8]. Studies, how-
ever, suggest that microtubules serve as a support for such a
formation [69]. Finally, Kuhle etal. [69] demonstrated that
SseF and SseG are co-localized with microtubules, present
in endosomal compartments that aggregate to SIF along
these same structures, further corroborating the hypothesis
postulated by Kuhle and Hensel [68].
As mentioned previously, the other SPIs has not been
described in as much detail as the first two, although some
information is available. SPI-3, for example, contains ten
ORFs organized into six transcriptional units [77]. The
proteins encoded by these genes, however, play roles with
no obvious relation to each other [6]. mgtCB, for exam-
ple, encodes proteins that are required for survival within
macrophages [5], thus contributing to its virulence. On the
other hand, MisL, a protein encoded in SPI-3, contributes to
intestinal colonization and is required for long-term intesti-
nal persistence and involved in host–pathogen interactions
during infection [24]. In addition, MarT has demonstrated
regulatory functions [26], in addition to activating MisL
expression [97].
SPI-4 contains six ORFs arranged in a single operon,
called siiABCDEF [26]. Morgan etal. [78] demonstrated
that, although not necessary for systemic infection, SPI-4
is required for intestinal colonization. Taking this into
consideration, Gerlach etal. [34] analyzed the role of SPI-
4-encoded proteins in the interaction with epithelial cells
and demonstrated that this island plays an important role in
adhesion to the intestinal epithelium. By infecting epithelial
Virulence Factors inSalmonella Typhimurium: The Sagacity ofaBacterium
1 3
cell lines (MDCK) with wild-type Salmonella and SPI-4-de-
ficient strains, Gerlach etal. [34] observed that a considera-
ble percentage (4.24%) of the wild-type Salmonella adhered
to the epithelium, whereas only 0.28% of SPI-4-deficient
Salmonella displayed this characteristic. Regarding the
SiiE protein encoded by the sii operon, some non-fimbrial
adhesin properties are displayed, evidenced by decreases
in adhesion when the siiE gene is deleted [34]. In order to
discover the individual role of each gene encoded by the sii
operon, Kiss etal. [62] tested the virulence of mutants carry-
ing non-polar deletions in individual sii genes with wild-type
Salmonella strains. The results demonstrated that, even at
different levels, wild-type strains were more virulent than
the six sii mutants, suggesting that they all display some sig-
nificance in Salmonella virulence. However, further studies
should be conducted to find out the exact role of these genes
in ST pathogenicity.
SPI-5 is composed of five genes (pipD, sigD/sopB,
sigE/pipC, pipB, and pipA) [64], which may be related to
the role of SPI-5 in enteropathogens [77]. Among the five
coded proteins, SopB (SigD in ST) is the best described and
known to be translocated by T3SS-1 into the host cell cyto-
plasm [46]. Galyov etal. [32] demonstrated that, although
sopB deletion did not affect pathogen invasiveness, SopB
plays an important role in the induction of fluid secretion by
enterocytes, as well as in PMN induction in the intestine. In
addition, SopB also presents inositol phosphate phosphatase
activity, which is directly related to the induction of diarrhea
[81]. Knodler etal. [65], on the other hand, demonstrated
that SopB is able to prevent apoptosis of cultured epithe-
lial cells from the activation of Akt, an important apoptosis
regulator in these cells, in a mechanism used by the host to
protect itself from a pathogen. SigE was characterized by
Hong and Miller [54] as a possible chaperone that acts on
the stability of SopB secretion. This hypothesis was later
confirmed by Darwin etal. [21]. Finally, mutational studies
were performed by Wood etal. [101] in order to discover the
role of the pipA, pipB, and pipD genes. None of the mutants
in these genes had their growth rate or ability to invade HeLa
cells affected. However, both inflammatory responses and
the rate of fluid secretion were reduced in pip mutants com-
pared to wild-type strains. This suggests that these genes
are directly linked to the Salmonella enteropathogenicity,
although they do not have an effect on the development of
systemic invasion [101].
Virulence Plasmids
Several Salmonella enterica carry virulence plasmids,
including S. Enteritidis, S. Dublin, S. Choleraesuis, and S.
Typhimurium [3941]. Such plasmids are differentiated
among serotypes according to their size [15]. In ST, the viru-
lence plasmid contains 95kb size [15] and is named pSLT
[92, 96]. It is believed that this structure was acquired hori-
zontally, since it is located in a region adjacent to the inser-
tion element and contains 46% less G+C than other parts
of the chromosome [77]. pSLT is equipped with a highly
conserved 8-kb sequence, with five ORFs, named the spv
operon (Salmonella plasmid of virulence) [40]. This small
part of the plasmid, however small, can restore virulence
in plasmid-cured salmonellae in a mouse model [41]. This
information led several research groups to further study the
role of spv genes in virulence, since strains lacking the spv
operon proved to be avirulent [39]. Gulig and Doyle [42],
for example, observed that spv genes were directly related to
the ability of ST to increase its multiplication rate within the
host during infection. Later, Guilloteau etal. [36] analyzed
the interactions of many wild-type and plasmid-cured Sal-
monella serotypes invitro murine and bovine macrophages
and invivo mice after infection. The presence of virulence
plasmids increased the lytic activity in macrophages infected
by ST, S. Dublin, and S. Choleraesuis, besides inducing
inflammatory responses in the peritoneum. Gulig etal. [43],
on the other hand, examined the relationship between spv
genes and the rate of ST growth in different cell populations,
including lymphocytes and neutrophils in addition to mac-
rophages. Genetic markers were used to measure the number
of cell divisions in cells infected with ST strains with or
without the virulence plasmid. It was noted that only the
macrophages showed relevance for ST multiplication [43],
since, with the quantitative decrease of these cells, the two
treatments (with and without the plasmid) showed the same
results for virulence [43]. Recently, Wu etal. [102], based
on analyses in zebrafish larvae, demonstrated the ability of
wild-type strains of ST to inhibit the autophagic activity of
macrophages and neutrophils, in addition to repressing the
function of macrophages and neutrophils when compared
to ST spv mutants. Thus, they demonstrated that the spv
genes are also involved in the suppression of innate immune
response of the host [102]. All these studies suggest the
important role that the spv operon plays in ST virulence,
although the mechanisms it applies are not yet fully under-
stood [102].
However, some mechanisms of effector proteins encoded
by spv operon have been elucidated. As mentioned previ-
ously, spv operon consists of five genes. The first one is spvR,
which is also the first gene to be transcribed [41]. The effec-
tor protein it encodes, SpvR, has as main function regulat-
ing the expression of the other four genes [40]. This protein
displays homology with MetR, a LysR member of positive
regulatory proteins [37]. SpvB is a cytotoxic protein [26],
which has the function of avoiding, above all, actin polym-
erization [38], being thus related to the intracellular phase
of infection. Additionally, Li etal. [74] demonstrated the
role of spvB in inhibiting autophagy in a zebrafish infection
model. This mechanism results in an increase of Salmonella
A.M.P.dos Santos et al.
1 3
virulence, since autophagy is a natural defense mechanism
of organisms infected by this pathogen [74]. SpvB was also
shown to be indispensable for virulence, since spvB mutants
are avirulent in mice [86]. SpvC is an anti-inflammatory
effector which plays an important role in the host’s proin-
flammatory response [26], since it inhibits MAP kinases
and, consequently, immune signaling [91]. Until recently,
SpvD still did not have its role defined; however, recently
Rolhion etal. [85] demonstrated the importance of this pro-
tein in the suppression of the immune response, interfering
with the transport of nuclear machinery mediated by NF-κβ.
Finally, SpvA still has no elucidated role in ST virulence,
requiring further studies [86].
Fimbriae andFlagella
Fimbriae are structures found on the cell surface of some
bacteria, which have been shown to play an important role in
the formation of biofilms, colonization, and the initial attack
of the bacterium on the host [57]. With the sequencing of the
ST genome, 13 operons were discovered [agf (csg), fim, lpf,
pef, bcf, stb, stc, std, stf, sth, sti, saf, and stj] with homology
to fimbrial biosynthesis genes [26, 76]. Since then, several
studies have been carried out to determine their action on the
virulence of this pathogen. Bäumler etal. [3] used a genetic
approach to investigate the role of three fimbrial operons
(fim, lpf, and pef) in ST adhesion to different epithelial cell
lines (HEp-2 and HeLa). These operons encode type I fim-
briae, long polar fimbriae, and plasmid-encoded fimbriae,
respectively [3]. The results showed a significant decrease in
adhesion to HEp-2 cells only for lpf ST mutants, whereas the
end deletion significantly decreased ST adhesion to HeLa.
These results suggest that the bacterial repertoire of fimbrial
adhesins determines which type of epithelial cell the bacte-
rium will adhere to during ST intestine infection [3]. Ween-
ing etal. [100] tested the role of colonization of ten fimbrial
operons (fim, pef, lbf, bcf, stb, stc, std, stf, sth, and agf) of
the bacteria in the intestine. Genetically resistant mice (e.g.,
CBA) were submitted to fimbrial operon mutation strains
and the results demonstrated that, from fimbrial operated
operons, only bcf, lpf, stb, stc, std, and sth contribute to the
ability of ST to be shed with the feces and to colonize the
cecum after 30 days of infection [100]. However, these same
data demonstrated that persistence in the intestine is com-
plex and involves the expression of more than 32 ST genes
[100]. Although we know the importance of these operons
in virulence, the exact mechanisms remain to be elucidated.
The flagella is a long helical filament coupled to rotat-
ing motors embedded within the outer membrane and cell
wall, which enables ST and the other bacteria that display
this feature to mobilize through the epithelial barrier after
ingestion [57]. It is characterized as a strong inflammation
inducer, mainly from the induction of Interleukin (IL)-8 and
activation of NF-κβ [17], a protein complex that plays the
role of a transcription factor and is involved in the immune
response to infection [73]. This induction is due to its chem-
otactic potential, which is also one of the main flagella char-
acteristics [13, 26].
Factors Present inOther Parts oftheChromosome
Some effector proteins are encoded outside the SPIs and
translocated to the host cell by T3SS. SopE (located in a
cryptic prophage) and SopE2, for example, although not
encoded in SPI-1, are translocated by T3SS-1 [77]. These,
together with the Sip proteins, play an important role in ST
invasion. Studies have demonstrated the ability of SopE
to induce membrane ruffling of cell cultures and that this
induction occurs from the GDP/GTP nucleotide exchange
in the Rho family, the G protein family responsible for the
intracellular dynamics of actin [49]. Because they are about
70% identical [47], SopE and SopE2 have similar roles in the
invasion [46]. However, Friebel etal. [29] demonstrated that
SopE and SopE2 activate different sets of RhoGTPases in
the host cell, suggesting that this small difference is respon-
sible for the existence of such similar proteins in the same
bacterium.
Other important effector proteins for ST virulence are
SifA and SseJ. These effectors are translocated into the host
cell cytoplasm by T3SS-2 [99]. They seem to have some
relation to each other [87] and their roles in ST virulence
are similar. Studies in sifA [4, 9] and sseJ [87] mutants dem-
onstrated an attenuation in their replication in macrophages
when compared with wild-type strains. In addition, a few
hours after internalization, sifA mutants appeared to dis-
play vacuolar membrane loss, with these organelles being
found loose in the host cell cytosol [4]. This demonstrates
the importance of this protein in both SCV maintenance and
intracellular replication. Another study reported that SifA is
also required for the formation of SIF [93] and, later, Kuhle
etal. [69] reached the same conclusion for SseJ, demonstrat-
ing that SseJ, like SseG and SseF, is located along with the
microtubules in host cells, which, as mentioned previously,
serve as support for the formation of SIF.
Conclusions
Salmonella Typhimurium is an intracellular pathogen, and
its ability to invade host cells and disseminate in the body is
closely related to its virulence genes. Further study of such
genes may allow for therapeutic measures against this micro-
organism to be developed, since the general knowledge of the
mechanisms used by ST in the invasion puts us one step ahead
in the fight against it. In light of what has been reported in this
review, it is notable that several advances have been achieved
Virulence Factors inSalmonella Typhimurium: The Sagacity ofaBacterium
1 3
in the research on the main factors of ST virulence. Among
these, it is perceived that SPIs play a crucial role in ST patho-
genicity, being, therefore, the most studied. The main studies
about the virulence of this pathogen are in the SPI-1 and SPI-
2. The discovery that both encode T3SSs was of great value,
since these apparatuses have the ability to inject not only effec-
tor proteins encoded in SPI-1 and 2, but also those encoded in
other parts of the chromosome into the cytoplasm of the host
cell. This fact further heightens the importance of these two
islands in maintaining virulence, since their absence eliminates
the ability of ST to inject T3SS-dependent effector proteins.
Despite this, much evidence also points out the role of the
other three SPIs, SPI-3, 4, and 5 on ST virulence. However,
much of the knowledge about these islands is still based on
empirical evidence, due to the absence of theoretical-scientific
basis, which can only be acquired with more research in the
area. An example of such structures are the SPI-4 sii genes and
Pip proteins in SPI-5, in which the exact mechanisms behind
the fact that their causes a certain attenuation in ST virulence
are not yet fully elucidated.
On the other hand, it has been shown that pSLT is
required for ST virulence. However, some mechanisms of
certain proteins produced in the spv operon, such as SpvA,
as well as their role in virulence, have not yet been fully elu-
cidated, requiring more comprehensive studies. Likewise, 13
operons encoding fimbriae are also involved in ST virulence,
contributing mainly to the adhesion of this bacterium to the
host cell. However, the exact mechanisms used by each gene
comprising this operon have not yet been discovered. Thus,
it can be concluded that, although many advances have been
made, there are still many challenges to be overcome regard-
ing ST virulence factors.
Acknowledgements We thank Virgínia P. Silveira for the design of
the figures.
Funding This study was funded by Fundação de Amparo à Pesquisa
do Estado do Rio de Janeiro (Process No. 232227, FAPERJ, Brazil).
Fundação de Amparo à Pesquisa do Estado do Rio de Janeiro (pro-
cess no. E-26/201.185/2014, FAPERJ, Brazil); Conselho Nacional de
Desenvolvimento Científico e Tecnológico (Process No. 311422/2016-
0, CNPq, Brazil), and Coordenação de Aperfeiçoamento de Pessoal
de Nível Superior (Process No. 125, CAPES/Embrapa 2014, CAPES,
Brazil).
Compliance with Ethical Standards
Conflict of interest The authors declare that the research was con-
ducted in the absence of any commercial or financial relationships that
could be construed as a potential conflict of interest.
References
1. Almeida F, Medeiros MIC, dos Prazeres Rodrigues D,
Allard MW, Falcão JP (2017) Molecular characterization
of Salmonella Typhimurium isolated in Brazil by CRISPR-
MVLST. J Microbiol Methods 133:55–61
2. Almeida F, Medeiros MIC, Rodrigues D, dos P, Falcão JP
(2015) Genotypic diversity, pathogenic potential and the resist-
ance profile of Salmonella Typhimurium strains isolated from
humans and food from 1983 to 2013 in Brazil. J Med Microbiol
64(11):1395–1407
3. Bäumler AJ, Tsolis RM, Heffron F (1996) Contribution of
fimbrial operons to attachment to and invasion of epithe-
lial cell lines by Salmonella typhimurium. Infect Immun
64(5):1862–1865
4. Beuzón CR, Méresse S, Unsworth KE, Ruíz-Albert J, Garvis S,
Waterman SR, Ryder TA, Boucrot E, Holden DW (2000) Salmo-
nella maintains the integrity of its intracellular vacuole through
the action of SifA. EMBO J 19(13):3235–3249
5. Blanc-Potard A-B, Groisman EA (1997) The Salmonella selC
locus contains a pathogenicity island mediating intramacrophage
survival. EMBO J 16(17):5376–5385
6. Blanc-Potard A-B, Solomon F, Kayser J, Groisman EA (1999)
The SPI-3 pathogenicity island of Salmonella enterica. J Bacte-
riol 181(3):998–1004
7. Broz P, Ohlson MB, Monack DM (2012) Innate immune
response to Salmonella typhimurium, a model enteric pathogen.
Gut Microbes 3(2):62–70
8. Brumell JH, Goosney DL, Finlay BB (2002) SifA, a Type III
secreted effector of Salmonella typhimurium, directs Salmonella-
induced filament (Sif) formation along microtubules. Traffic
3(6):407–415
9. Brumell JH, Rosenberger CM, Gotto GT, Marcus SL, Finlay BB
(2001) SifA permits survival and replication of Salmonella typh-
imurium in murine macrophages. Cell Microbiol 3(2):75–84
10. Brumell JH, Tang P, Mills SD, Finlay BB (2001) Characteri-
zation of Salmonella-induced filaments (Sifs) reveals a delayed
interaction between Salmonella-containing vacuoles and late
endocytic compartments. Traffic 2(9):643–653
11. Bueno SM, Riquelme S, Riedel CA, Kalergis AM (2012) Mecha-
nisms used by virulent Salmonella to impair dendritic cell func-
tion and evade adaptive immunity. Immunology 137(1):28–36
12. Carden S, Okoro C, Dougan G, Monack D (2015) Non-typhoidal
Salmonella Typhimurium ST313 isolates that cause bacteremia
in humans stimulate less inflammasome activation than ST19
isolates associated with gastroenteritis. Pathog Dis. https ://doi.
org/10.1093/femsp d/ftu02 3
13. Chilcott GS, Hughes KT (2000) Coupling of Flagellar gene
expression to Flagellar assembly in Salmonella enterica sero-
var Typhimurium and Escherichia coli. Microbiol Mol Biol Rev
64(4):694–708
14. Choi HW, Brooking-Dixon R, Neupane S, Lee C-J, Miao EA,
Staats HF, Abraham SN (2013) Salmonella Typhimurium
impedes innate immunity with a mast-cell-suppressing protein
tyrosine phosphatase, SptP. Immunity 39(6):1108–1120
15. Chu C, Hong S-F, Tsai C, Lin W-S, Liu T-P, Ou JT (1999) Com-
parative physical and genetic maps of the virulence plasmids of
Salmonella enterica Serovars Typhimurium, Enteritidis, Chol-
eraesuis, and Dublin. Infect Immun 67(5):2611–2614
16. Cirillo DM, Valdivia RH, Monack DM, Falkow S (1998) Mac-
rophage-dependent induction of the Salmonella pathogenicity
island 2 type III secretion system and its role in intracellular
survival. Mol Microbiol 30(1):175–188
17. Coburn B, Grassl GA, Finlay B (2007) Salmonella, the host and
disease: a brief review. Immunol Cell Biol 85:112–118
18. Collazo CM, Galán JE (1997) The invasion-associated type III
system of Salmonella typhimurium directs the translocation of
Sip proteins into the host cell. Mol Microbiol 24(4):747–756
19. Connor BA, Schwartz E (2005) Typhoid and paratyphoid fever
in travellers. Lancet Infect Dis 5(10):623–628
A.M.P.dos Santos et al.
1 3
20. Costa TRD, Felisberto-Rodrigues C, Meir A, Prevost MS,
Redzej A, Trokter M, Waksman G (2015) Secretion systems
in gram-negative bacteria: structural and mechanistic insights.
Nat Rev Microbiol 13(6):343–359
21. Darwin KH, Robinson LS, Miller VL (2001) SigE is a chaper-
one for the Salmonella enterica Serovar Typhimurium invasion
protein SigD. J Bacteriol 183(4):1452–1454
22. Deiwick J, Nikolaus T, Shea JE, Gleeson C, Holden DW, Hen-
sel M (1998) Mutations in Salmonella pathogenicity island 2
(SPI2) genes affecting transcription of SPI1 genes and resist-
ance to antimicrobial agents. J Bacteriol 180(18):4775–4780
23. Deng W, Marshall NC, Rowland JL, McCoy JM, Worrall LJ,
Santos AS, Strynadka NCJ, Finlay BB (2017) Assembly, struc-
ture, function and regulation of type III secretion systems. Nat
Rev Microbiol 15(6):323–337
24. Dorsey CW, Laarakker MC, Humphries AD, Weening EH,
Bäumler AJ (2005) Salmonella enterica serotype Typhimurium
MisL is an intestinal colonization factor that binds fibronectin.
Mol Microbiol 57(1):196–211
25. Eng S-K, Pusparajah P, Mutalib N-SA, Ser H-L, Chan K-G,
Lee L-H (2015) Salmonella: a review on pathogenesis epide-
miology and antibiotic resistance. Front Life Sci 8(3):284–293
26. Fàbrega A, Vila J (2013) Salmonella enterica Serovar Typh-
imurium skills to succeed in the host: virulence and regulation.
Clin Microbiol Rev 26(2):308–341
27. Fields PI, Swanson RV, Haidaris CG, Heffron F (1986)
Mutants of Salmonella typhimurium that cannot survive
within the macrophage are avirulent. Proc Natl Acad Sci USA
83(14):5189–5193
28. Freeman JA, Rappl C, Kuhle V, Hensel M, Miller SI (2002)
SpiC is required for translocation of Salmonella pathogenicity
island 2 effectors and secretion of translocon proteins SseB and
SseC. J Bacteriol 184(18):4971–4980
29. Friebel A, Ilchmann H, Aepfelbacher M, Ehrbar K, Machleidt
W, Hardt W-D (2001) SopE and SopE2 from Salmonella typh-
imurium activate different sets of RhoGTPases of the host cell.
J Biol Chem 276(36):34035–34040
30. Fu Y, Galán JE (1998) The Salmonella typhimurium tyrosine
phosphatase SptP is translocated into host cells and disrupts
the actin cytoskeleton. Mol Microbiol 27(2):359–368
31. Galán JE, Curtiss R (1989) Cloning and molecular charac-
terization of genes whose products allow Salmonella typhimu-
rium to penetrate tissue culture cells. Proc Natl Acad Sci USA
86(16):6383–6387
32. Galyov EE, Wood MW, Rosqvist R, Mullan PB, Watson PR,
Hedges S, Wallis TS (1997) A secreted effector protein of Sal-
monella dublin is translocated into eukaryotic cells and medi-
ates inflammation and fluid secretion in infected ileal mucosa.
Mol Microbiol 25(5):903–912
33. Garai P, Gnanadhas DP, Chakravortty D (2012) Salmonella
enterica serovars Typhimurium and Typhi as model organisms.
Virulence 3(4):377–388
34. Gerlach RG, Jäckel D, Stecher B, Wagner C, Lupas A, Hardt
W-D, Hensel M (2007) Salmonella pathogenicity Island 4
encodes a giant non-fimbrial adhesin and the cognate type 1
secretion system. Cell Microbiol 9(7):1834–1850
35. Ghilardi ÂCR, Tavechio AT, Fernandes SA (2006) Antimicro-
bial susceptibility, phage types, and pulsetypes of Salmonella
Typhimurium, in São Paulo, Brazil. Mem Inst Oswaldo Cruz
101(3):281–286
36. Guilloteau LA, Wallis TS, Gautier AV, MacIntyre S, Platt DJ,
Lax AJ (1996) The Salmonella virulence plasmid enhances
Salmonella-induced lysis of macrophages and influences
inflammatory responses. Infect Immun 64(8):3385–3393
37. Guiney DG, Fang FC, Krause M, Libby S (1994) Plasmid-medi-
ated virulence genes in non-typhoid Salmonella serovars. FEMS
Microbiol Lett 124(1):1–9
38. Guiney DG, Fierer J (2011) The role of the spv genes in Sal-
monella pathogenesis. Front Microbiol. https ://doi.org/10.3389/
fmicb .2011.00129
39. Gulig PA (1990) Virulence plasmids of Salmonella typhimurium
and other salmonellae. Microb Pathog 8(1):3–11
40. Gulig PA, Caldwell AL, Chiodo VA (1992) Identification, genetic
analysis and DNA sequence of a 7.8-kb virulence region of the
Salmonella typhimurium virulence plasmid. Mol Microbiol
6(10):1395–1411
41. Gulig PA, Danbara H, Guiney DG, Lax AJ, Norel F, Rhen M
(1993) Molecular analysis of spv virulence genes of the salmo-
nella virulence plasmids. Mol Microbiol 7(6):825–830
42. Gulig PA, Doyle TJ (1993) The Salmonella typhimurium viru-
lence plasmid increases the growth rate of salmonellae in mice.
Infect Immun 61(2):504–511
43. Gulig PA, Doyle TJ, Hughes JA, Matsui H (1998) Analysis of
host cells associated with the Spv-mediated increased intracel-
lular growth rate of Salmonella typhimurium in Mice. Infect
Immun 66(6):2471–2485
44. Haimovich B, Venkatesan MM (2006) Shigella and Salmonella:
death as a means of survival. Microbes Infect 8(2):568–577
45. Halici S, Zenk SF, Jantsch J, Hensel M (2008) Functional analy-
sis of the Salmonella pathogenicity island 2-mediated inhibi-
tion of antigen presentation in dendritic cells. Infect Immun
76(11):4924–4933
46. Hansen-Wester I, Hensel M (2001) Salmonella pathogenicity
islands encoding type III secretion systems. Microbes Infect
3(7):549–559
47. Hapfelmeier S, Ehrbar K, Stecher B, Barthel M, Kremer M,
Hardt W-D (2004) Role of the Salmonella pathogenicity island
1 effector proteins SipA, SopB, SopE, and SopE2 in Salmonella
enterica subspecies 1 Serovar Typhimurium Colitis in strepto-
mycin-pretreated mice. Infect Immun 72(2):795–809
48. Haraga A, Ohlson MB, Miller SI (2008) Salmonellae interplay
with host cells. Nat Rev Microbiol 6(1):53–66
49. Hardt W-D, Chen L-M, Schuebel KE, Bustelo XR, Galán JE
(1998) S. typhimurium encodes an activator of rho GTPases that
induces membrane ruffling and nuclear responses in host cells.
Cell 93(5):815–826
50. Hensel M, Nikolaus T, Egelseer C (1999) Molecular and func-
tional analysis indicates a mosaic structure of Salmonella patho-
genicity island 2. Mol Microbiol 31(2):489–498
51. Hensel M, Shea JE, Waterman SR, Mundy R, Nikolaus T, Banks
G, Vazquez-Torres A, Gleeson C, Fang FC, Holden DW (1998)
Genes encoding putative effector proteins of the type III secre-
tion system of Salmonella pathogenicity island 2 are required
for bacterial virulence and proliferation in macrophages. Mol
Microbiol 30(1):163–174
52. Herrero-Fresno A, Olsen JE (2017) Salmonella Typhimurium
metabolism affects virulence in the host—a mini-review. Food
Microbiol. https ://doi.org/10.1016/j.fm.2017.04.016
53. Hersh D, Monack DM, Smith MR, Ghori N, Falkow S, Zychlin-
sky A (1999) The Salmonella invasin SipB induces macrophage
apoptosis by binding to caspase-1. Proc Natl Acad Sci USA
96(5):2396–2401
54. Hong KH, Miller VL (1998) Identification of a novel Salmo-
nella invasion locus homologous to shigella ipgDE. J Bacteriol
180(7):1793–1802
55. Hueck CJ (1998) Type III protein secretion systems in bacte-
rial pathogens of animals and plants. Microbiol Mol Biol Rev
62(2):379–433
56. Jones BD, Ghori N, Falkow S (1994) Salmonella typhimurium
initiates murine infection by penetrating and destroying the
Virulence Factors inSalmonella Typhimurium: The Sagacity ofaBacterium
1 3
specialized epithelial M cells of the Peyer’s patches. J Exp Med
180(1):15–23
57. Jong HK de, Parry CM, Poll T van der, Wiersinga WJ (2012)
Host–pathogen interaction in invasive salmonellosis. PLoS Pat-
hog 8(10):e1002933
58. Kaniga K, Uralil J, Bliska JB, Galán JE (1996) A secreted pro-
tein tyrosine phosphatase with modular effector domains in the
bacterial pathogen Salmonella typhimurlum. Mol Microbiol
21(3):633–641
59. Kimbrough TG, Miller SI (2000) Contribution of Salmonella
typhimurium type III secretion components to needle complex
formation. Proc Natl Acad Sci USA 97(20):11008–11013
60. Kimbrough TG, Miller SI (2002) Assembly of the type III secre-
tion needle complex of Salmonella typhimurium. Microbes Infect
4(1):75–82
61. Kingsley RA, Msefula CL, Thomson NR etal (2009) Epidemic
multiple drug resistant Salmonella Typhimurium causing inva-
sive disease in sub-Saharan Africa have a distinct genotype.
Genome Res 19(12):2279–2287
62. Kiss T, Morgan E, Nagy G (2007) Contribution of SPI-4 genes
to the virulence of Salmonella enterica. FEMS Microbiol Lett
275(1):153–159
63. Klein JR, Jones BD (2001) Salmonella pathogenicity island
2-encoded proteins SseC and SseD are essential for virulence
and are substrates of the type III secretion system. Infect Immun
69(2):737–743
64. Knodler LA, Celli J, Hardt W-D, Vallance BA, Yip C, Finlay BB
(2002) Salmonella effectors within a single pathogenicity island
are differentially expressed and translocated by separate type III
secretion systems. Mol Microbiol 43(5):1089–1103
65. Knodler LA, Finlay BB, Steele-Mortimer O (2005) The
Salmonella effector protein SopB protects epithelial cells
from apoptosis by sustained activation of Akt. J Biol Chem
280(10):9058–9064
66. Kubori T, Matsushima Y, Nakamura D, Uralil J, Lara-Tejero M,
Sukhan A, Galán JE, Aizawa S-I (1998) Supramolecular struc-
ture of the Salmonella typhimurium type III protein secretion
system. Science 280(5363):602–605
67. Kubori T, Sukhan A, Aizawa S-I, Galán JE (2000) Molecular
characterization and assembly of the needle complex of the Sal-
monella typhimurium type III protein secretion system. Proc Natl
Acad Sci USA 97(18):10225–10230
68. Kuhle V, Hensel M (2002) SseF and SseG are translocated effec-
tors of the type III secretion system of Salmonella pathogenicity
island 2 that modulate aggregation of endosomal compartments.
Cell Microbiol 4(12):813–824
69. Kuhle V, Jäckel D, Hensel M (2004) Effector proteins encoded
by Salmonella pathogenicity island 2 interfere with the micro-
tubule cytoskeleton after translocation into host cells. Traffic
5(5):356–370
70. Lan Y, Wang S, Yin Y, Hoffmann WC, Zheng X (2008) Using
a surface plasmon resonance biosensor for rapid detection of
Salmonella Typhimurium in chicken carcass. J Bionic Eng
5(3):239–246
71. Lee CA (1997) Type III secretion systems: machines to deliver
bacterial proteins into eukaryotic cells? Trends Microbiol
5(4):148–156
72. Lee K-M, Runyon M, Herrman TJ, Phillips R, Hsieh J (2015)
Review of Salmonella detection and identification methods:
aspects of rapid emergency response and food safety. Food Con-
trol 47:264–276
73. Li Q, Verma IM (2002) NF-κB Regulation in the imunne system.
Nat Rev Immunol 2:725–734
74. Li Y, Wang T, Gao S, Xu G, Niu H, Huang R, Wu S (2016)
Salmonella plasmid virulence gene spvB enhances bacterial
virulence by inhibiting autophagy in a zebrafish infection model.
Fish Shellfish Immunol 49:252–259
75. Lo Fo Wong DMA, Hald T, van der Wolf PJ, Swanenburg M
(2002) Epidemiology and control measures for Salmonella in
pigs and pork. Livest Prod Sci 76(3):215–222
76. López FE, de las Mercedes Pescaretti M, Morero R, Delgado
MA (2012) Salmonella Typhimurium general virulence factors:
a battle of David against Goliath? Food Res Int 45(2):842–851
77. Marcus SL, Brumell JH, Pfeifer CG, Finlay BB (2000) Salmo-
nella pathogenicity islands: big virulence in small packages.
Microbes Infect 2(2):145–156
78. Morgan E, Campbell JD, Rowe SC, Bispham J, Stevens MP,
Bowen AJ, Barrow PA, Maskell DJ, Wallis TS (2004) Identifica-
tion of host-specific colonization factors of Salmonella enterica
serovar Typhimurium. Mol Microbiol 54(4):994–1010
79. Nieto PA, Pardo-Roa C, Salazar-Echegarai FJ, Tobar HE,
Coronado-Arrázola I, Riedel CA, Kalergis AM, Bueno SM
(2016) New insights about excisable pathogenicity islands in
Salmonella and their contribution to virulence. Microbes Infect
18(5):302–309
80. Nikolaus T, Deiwick J, Rappl C, Freeman JA, Schröder W, Miller
SI, Hensel M (2001) SseBCD proteins are secreted by the type
III secretion system of Salmonella pathogenicity island 2 and
function as a translocon. J Bacteriol 183(20):6036–6045
81. Norris FA, Wilson MP, Wallis TS, Galyov EE, Majerus PW
(1998) SopB, a protein required for virulence of Salmonella
dublin, is an inositol phosphate phosphatase. Proc Natl Acad
Sci USA 95(24):14057–14059
82. Ochman H, Soncini FC, Solomon F, Groisman EA (1996) Iden-
tification of a pathogenicity island required for Salmonella sur-
vival in host cells. Proc Natl Acad Sci USA 93(15):7800–7804
83. Pui CF, Wong WC, Chai LC, Tunung R, Jeyaletchumi P, Noor
Hidayah MS, Ubong A, Farinazleen MG, Cheah YK, Son R
(2011) Review article Salmonella: a foodborne pathogen. Int
Food Res J 18:465–473
84. Ridley A, Threlfall EJ (1998) Molecular epidemiology of anti-
biotic resistance genes in multiresistant epidemic Salmonella
typhimurium DT 104. Microb Drug Resist 4(2):113–118
85. Rolhion N, Furniss RCD, Grabe G, Ryan A, Liu M, Matthews
SA, Holden DW (2016) Inhibition of nuclear transport of NF-ĸB
p65 by the Salmonella type III secretion system effector SpvD.
PLoS Pathog 12(5):e1005653
86. Rotger R, Casadesús J (1999) The virulence plasmids of Salmo-
nella. Int Microbiol Off J Span Soc Microbiol 2(3):177–184
87. Ruiz-Albert J, Yu X-J, Beuzón CR, Blakey AN, Galyov EE,
Holden DW (2002) Complementary activities of SseJ and SifA
regulate dynamics of the Salmonella typhimurium vacuolar
membrane. Mol Microbiol 44(3):645–661
88. Sabbagh SC, Forest CG, Lepage C, Leclerc J-M, Daigle F (2010)
So similar, yet so different: uncovering distinctive features in
the genomes of Salmonella enterica serovars Typhimurium and
Typhi. FEMS Microbiol Lett 305(1):1–13
89. Sansonetti P (2002) Host–pathogen interactions: the seduction
of molecular cross talk. Gut 50(suppl 3):iii2–iii8
90. Shinohara NKS, Barros VB de, Jimenez SMC, Machado E, de
CL, Dutra, Filho RAF, De L JL (2008) Salmonella spp., impor-
tante agente patogênico veiculado em alimentos. Ciênc Amp
Saúde Coletiva 13(5):1675–1683
91. Silva C, Puente JL, Calva E (2017) Salmonella virulence plasmid:
pathogenesis and ecology. Pathog Dis. https ://doi.org/10.1093/
femsp d/ftx07 0
92. Smith HR, Humphreys GO, Grindley NDF, Grindley JN,
Anderson ES (1973) Molecular studies of an fi + plasmid from
strains of Salmonella typhimurium. Mol Gen Genet MGG
126(2):143–151
A.M.P.dos Santos et al.
1 3
93. Stein MA, Leung KY, Zwick M, Portillo FG, Finlay BB (1996)
Identification of a Salmonella virulence gene required for for-
mation of filamentous structures containing lysosomal mem-
brane glycoproteins within epithelial cells. Mol Microbiol
20(1):151–164
94. Sukhan A, Kubori T, Wilson J, Galán JE (2001) Genetic analy-
sis of assembly of the Salmonella enterica Serovar Typhimu-
rium type III secretion-associated needle complex. J Bacteriol
183(4):1159–1167
95. Suresh T, Hatha AAM, Sreenivasan D, Sangeetha N, Lashmana-
perumalsamy P (2006) Prevalence and antimicrobial resistance
of Salmonella enteritidis and other salmonellas in the eggs and
egg-storing trays from retails markets of Coimbatore, South
India. Food Microbiol 23(3):294–299
96. Tinge SA, Curtiss R (1990) Isolation of the replication and
partitioning regions of the Salmonella typhimurium virulence
plasmid and stabilization of heterologous replicons. J Bacteriol
172(9):5266–5277
97. Tükel Ç, Akçelik M, Jong MF de, Şimşek Ö, Tsolis RM, Bäumler
AJ (2007) MarT activates expression of the MisL autotransporter
protein of Salmonella enterica serotype Typhimurium. J Bacte-
riol 189(10):3922–3926
98. Uchiya K, Barbieri MA, Funato K, Shah AH, Stahl PD, Gro-
isman EA (1999) A Salmonella virulence protein that inhibits
cellular trafficking. EMBO J 18(14):3924–3933
99. Waterman SR, Holden DW (2003) Functions and effectors of the
Salmonella pathogenicity island 2 type III secretion system. Cell
Microbiol 5(8):501–511
100. Weening EH, Barker JD, Laarakker MC, Humphries AD, Tso-
lis RM, Bäumler AJ (2005) The Salmonella enterica Serotype
Typhimurium lpf, bcf, stb, stc, std, and sth Fimbrial Operons
Are Required for Intestinal Persistence in Mice. Infect Immun
73(6):3358–3366
101. Wood MW, Jones MA, Watson PR, Hedges S, Wallis TS, Galyov
EE (1998) Identification of a pathogenicity island required for
Salmonella enteropathogenicity. Mol Microbiol 29(3):883–891
102. Wu S, Wang L, Li J, Xu G, He M, Li Y, Huang R (2016) Sal-
monella spv locus suppresses host innate immune responses to
bacterial infection. Fish Shellfish Immunol 58:387–396
103. Yang L, Li Y (2005) Quantum dots as fluorescent labels for quan-
titative detection of Salmonella Typhimurium in chicken carcass
wash water. J Food Prot 68(6):1241–1245
104. Zhou D, Mooseker MS, Galán JE (1999) Role of the S. typh-
imurium actin-binding protein SipA in bacterial internalization.
Science 283(5410):2092–2095
... Salmonella enterica serovar Typhimurium (S. typhimurium) is a Gram-negative bacterium that can cause foodborne illness in humans [1,2]. It can cause a wide range of symptoms, including diarrhea, abdominal cramps, fever, and vomiting, resulting in a significant public health burden [3,4]. ...
... In most cases, the resulting condition is ultimately self-limited and usually manifests only mild symptoms [5]. However, under some circumstances, severe cases can occur in vulnerable human populations such as the elderly, young children, and immunocompromised individuals [2,6]. Due to the public health concerns for S. typhimurium illnesses, the government regulatory authorities and public health organizations cooperate to detect outbreaks, determine the sources of contamination, and, in turn, implement preventive measures to halt the further spread of the pathogen [7,8]. ...
Article
Full-text available
Studies have shown that the production of reactive oxygen species (ROS) is triggered by bactericidal antibiotics, which contributes significantly to the killing of bacterial cells and increasing mutations in surviving cells. In this study, we hypothesized that exposure of Salmonella to sublethal concentrations of hypochlorite (NaOCl), a commonly used sanitizer in household and food industries increases mutation rates, leading to the development of antibiotic resistance. We found that a sublethal concentration (20 ppm) of NaOCl increased the mutation rates of S. typhimurium 14028s significantly (p < 0.05), which was prevented by the ROS scavenger thiourea, supporting that the increased mutation was due to NaOCl-triggered ROS production. We further found that the exposure of S. typhimurium 14028s to the same sublethal concentration of NaOCl increases resistance to kanamycin among the 3 antibiotics evaluated. The results of this study suggest that when NaOCl applied as a sanitizer fails to kill Salmonella due to diluted local concentrations or presence of organic materials, it can cause an adverse outcome of developing antibiotic resistance of the pathogen.
... Some genes necessary for the invasion of intestinal epithelial cells and initiation of intestinal secretory and inflammatory responses are contained within Salmonella Pathogenicity Island 1 (SPI-1) [9]. Salmonella Pathogenicity Island 2 (SPI-2) is necessary for systemic infection and establishment beyond the intestinal epithelium and encodes genes essential for intracellular replication [9][10][11]. ...
... Some genes necessary for the invasion of intestinal epithelial cells and initiation of intestinal secretory and inflammatory responses are contained within Salmonella Pathogenicity Island 1 (SPI-1) [9]. Salmonella Pathogenicity Island 2 (SPI-2) is necessary for systemic infection and establishment beyond the intestinal epithelium and encodes genes essential for intracellular replication [9][10][11]. Intestinal epithelial surface adhesion is the initial Salmonella pathogenesis step and is central to its colonization. ...
Article
Full-text available
Foodborne Salmonella serovars are important facultative intracellular pathogens that cause gastroenteritis in humans. Four strains from three of the more predominant Salmonella serovars in poultry were studied: Typhimurium, Enteritidis, and Heidelberg. Gentamicin susceptibility was determined using an agar disc diffusion test and minimum inhibitory concentration (MIC) assays for S. Typhimurium ATCC 14028 and S. Heidelberg ARI-14. Both strains were susceptible to gentamicin in disc diffusion. The MIC of gentamicin was approximately 125 mg/ml for all strains tested. These strains' adhesion and invasion abilities were determined with two different cell lines, a human intestinal epithelial cell line (Caco-2) as well as a chicken macrophage cell line (HD11). Attachment percentages for each Salmonella strain were greater than the strain’s ability to invade cells. Similar attachment percentages to Caco-2 cells were observed for S. Typhimurium and S. Heidelberg. Attachment percentages were lower in HD11 cells than in Caco-2 cells, although Salmonella exhibited higher apparent HD11 invasion, likely from HD11 phagocytosis. Salmonella Enteritidis showed lower rates of adhesion and invasion in HD11 cells compared to Salmonella Typhimurium. Developing a better understanding of Salmonella virulence mechanisms is critical to reducing Salmonella infections.
... The pathogenicity of Salmonella spp, closely related to the expression of VFs, significantly affects the severity of human infections caused by this pathogen. 38 Our study detected a large number of VFs in all Salmonella strains, indicating the potential pathogenicity of these strains in our hospital. Additionally, a considerable portion of adhesionrelated genes contributed to biofilm formation, enhanced bacterial adhesion and colonization abilities, which are crucial for bacterial pathogenicity. ...
Article
Full-text available
Objective To investigate the clinical and molecular characteristics of Salmonella spp. causing bloodstream infections (BSIs) in our hospital. Methods We studied 22 clinical Salmonella isolates from BSIs and 16 from non-BSIs, performing antimicrobial susceptibility testing (AST) and whole genome sequencing (WGS). The analysis included serovars, antibiotic resistance genes (ARGs), virulence factors (VFs), sequence types (STs), plasmid replicons, and genetic relationships. We also assessed pathogenicity of the isolates causing BSIs through growth, biofilm formation, and anti-serum killing assays. Results WGS analysis identified 13 Salmonella serovars, with four responsible for BSIs. S. Enteritidis was the most prevalent serovar, involved in 19 (50.0%) cases. BSIs were caused by 17S. Enteritidis, two S. Typhimurium, two S. Munster and one S. Diguel. Of the 38 isolates, 27 (71.1%) exhibited high resistance to ampicillin, and 24 (63.2%) to ampicillin/sulbactam. Thirty-six types of ARGs were identified, with blaTEM-1B (n = 25, 65.8%) being the most frequent. Ten plasmid replicons were found; the combination of IncFIB(S)-IncFII(S)-IncX1 was the most common in S. Enteritidis (94.7%). Fifteen STs were identified, among which ST11 was the most prevalent and clonally disseminated, primarily responsible for BSIs. A total of 333 different VFs were detected, 177 of which were common across all strains. No significant differences were observed between the BSI and non-BSI isolates in terms of resistance rates, ARGs, plasmid replicons, and VFs, except for seven VFs. No strong pathogenicity was observed in the BSI-causing isolates. Conclusion BSIs were predominantly caused by clonally disseminated S. Enteritidis ST11, the majority of which carried multiple ARGs, VFs and plasmid replicons. This study provides the first data on clonally disseminated S. Enteritidis ST11 causing BSIs, highlighting the urgent need for enhanced infection control measures.
... The P-ring is critical for firmly holding the flagellum to the peptidoglycan layer of the cell wall, aiding the firm rotation of the flagella. The proper function of the P-ring relies on the presence of a functional flgA protein, underscoring the paramount role of flgA in maintaining bacterial movement, an essential physiological process for achieving its biological goals such as chemotaxis, colonization, and pathogenesis [7]. ...
Preprint
Full-text available
Salmonella Typhimurium is a typical Gram-Negative bacterium associated with food-borne illness threatening public health. This pathogen is reported to have caused millions of infections and thousands of hospitalizations yearly in the United States alone. Salmonella Typhimurium has a sophisticated array of virulent factors, including its flagella, that contribute significantly to its pathogenicity. Cannabidiol (CBD), the non-psychoactive component of the cannabis plant, has attracted the interest of many researchers recently because of its antimicrobial properties. With the growing interest in exploring CBD’s antimicrobial properties, we, in this study, have delved into unravelling how prolonged exposure of Salmonella Typhimurium to CBD influences the expression of the flgA gene. We hypothesized that the exposure of Salmonella Typhimurium to CBD may result in flgA upregulation, causing the bacteria to withstand various stresses and potentially enhance its survival rates and pathogenicity. We have employed several assays, such as mRNA expression, motility assays, pH induction assays, nutrient starvation interventions, and biofilm analysis, to explore the influence of CBD on the pathophysiology of Salmonella Typhimurium. Our findings reveal that CBD-resistant strains exhibit enhanced resilience to extreme pH levels and motility and survive poorly under nutrient starvation, with more robust biofilm formation compared to the susceptible strains. The results contribute to the broader comprehension of CBD resistance mechanisms in bacteria. This work provides insight into the potential therapeutic interventions for controlling CBD-resistant bacterial infections. This study also underscores the essence of exploring novel alternative antimicrobial agents to tackle the ever-persistence of bacterial antibiotic resistance.
... Lipopolysaccharide (LPS) is the main component of the outer membrane of Gram-negative bacteria and a key virulence factor of NTS strains [12]. LPS forms a barrier between the cell and the external environment, providing the bacteria resistance to antibiotics and detergents. ...
Article
Full-text available
Bacterial resistance to serum is a key virulence factor for the development of systemic infections. The amount of lipopolysaccharide (LPS) and the O-antigen chain length distribution on the outer membrane, predispose Salmonella to escape complement-mediated killing. In Salmonella enterica serovar Enteritidis (S. Enteritidis) a modal distribution of the LPS O-antigen length can be observed. It is characterized by the presence of distinct fractions: low molecular weight LPS, long LPS and very long LPS. In the present work, we investigated the effect of the O-antigen modal length composition of LPS molecules on the surface of S. Enteritidis cells on its ability to evade host complement responses. Therefore, we examined systematically, by using specific deletion mutants, roles of different O-antigen fractions in complement evasion. We developed a method to analyze the average LPS lengths and investigated the interaction of the bacteria and isolated LPS molecules with complement components. Additionally, we assessed the aspect of LPS O-antigen chain length distribution in S. Enteritidis virulence in vivo in the Galleria mellonella infection model. The obtained results of the measurements of the average LPS length confirmed that the method is suitable for measuring the average LPS length in bacterial cells as well as isolated LPS molecules and allows the comparison between strains. In contrast to earlier studies we have used much more precise methodology to assess the LPS molecules average length and modal distribution, also conducted more subtle analysis of complement system activation by lipopolysaccharides of various molecular mass. Data obtained in the complement activation assays clearly demonstrated that S. Enteritidis bacteria require LPS with long O-antigen to resist the complement system and to survive in the G. mellonella infection model.
Article
Full-text available
Salmonellosis is one of the important bacterial infectious diseases affecting the health of pigeons. Heretofore, the epidemiological characteristics of Salmonella in pigeon populations in China remain largely unclear. The present study investigated the antimicrobial resistance and genomic characteristics of Salmonella isolates in pigeons in different regions of China from 2022 to 2023. Thirty-two Salmonella isolates were collected and subjected to 24 different antimicrobial agents, representing nine categories. The results showed that these isolates were highly resistant to cefazolin (100%), gentamicin (100%), tobramycin (100%), and amikacin (100%). Three or more types of antimicrobial resistance were present in 90.62% of the isolates, indicating multidrug resistance. Furthermore, using whole genome sequencing technology, we analyzed the profiles of serotypes, multilocus sequence typing, virulence genes, antimicrobial resistance genes, and plasmid replicons and constructed phylogenetic genomics to determine the epidemiological correlation among these isolates. All strains belonged to Salmonella Typhimurium var. Copenhagen and exhibited five antimicrobial resistance genes and more than 150 Salmonella virulence genes. Moreover, each isolate contained both the IncFIB(S) and IncFII(S) plasmids. In addition, phylogenetic analysis showed that all isolates were very close to each other, and isolates from the same region clustered in the same branch. Overall, our findings provide the first evidence for the epidemiological characteristics of Salmonella in pigeons of China, highlighting the importance of preventing salmonellosis in pigeons.
Article
Full-text available
The bacterial envelope is an essential compartment involved in metabolism and metabolites transport, virulence, and stress defense. Its roles become more evident when homeostasis is challenged during host–pathogen interactions. In particular, the presence of free radical groups and excess copper in the periplasm causes noxious reactions, such as sulfhydryl group oxidation leading to enzymatic inactivation and protein denaturation. In response to this, canonical and accessory oxidoreductase systems are induced, performing quality control of thiol groups, and therefore contributing to restoring homeostasis and preserving survival under these conditions. Here, we examine recent advances in the characterization of the Dsb-like, Salmonella-specific Scs system. This system includes the ScsC/ScsB pair of Cu⁺-binding proteins with thiol-oxidoreductase activity, an alternative ScsB-partner, the membrane-linked ScsD, and a likely associated protein, ScsA, with a role in peroxide resistance. We discuss the acquisition of the scsABCD locus and its integration into a global regulatory pathway directing envelope response to Cu stress during the evolution of pathogens that also harbor the canonical Dsb systems. The evidence suggests that the canonical Dsb systems cannot satisfy the extra demands that the host-pathogen interface imposes to preserve functional thiol groups. This resulted in the acquisition of the Scs system by Salmonella. We propose that the ScsABCD complex evolved to connect Cu and redox stress responses in this pathogen as well as in other bacterial pathogens.
Article
Salmonella enterica is the leading cause of bacterial foodborne illness in the USA, with an estimated 95% of salmonellosis cases due to the consumption of contaminated food products. Salmonella can cause several different disease syndromes, with the most common being gastroenteritis, followed by bacteremia and typhoid fever. Among the over 2,600 currently identified serotypes/serovars, some are mostly host-restricted and host-adapted, while the majority of serotypes can infect a broader range of host species and are associated with causing both livestock and human disease. Salmonella serotypes and strains within serovars can vary considerably in the severity of disease that may result from infection, with some serovars that are more highly associated with invasive disease in humans, while others predominantly cause mild gastroenteritis. These observed clinical differences may be caused by the genetic make-up and diversity of the serovars. Salmonella virulence systems are very complex containing several virulence-associated genes with different functions that contribute to its pathogenicity. The different clinical syndromes are associated with unique groups of virulence genes, and strains often differ in the array of virulence traits they display. On the chromosome, virulence genes are often clustered in regions known as Salmonella pathogenicity islands (SPIs), which are scattered throughout different Salmonella genomes and encode factors essential for adhesion, invasion, survival, and replication within the host. Plasmids can also carry various genes that contribute to Salmonella pathogenicity. For example, strains from several serovars associated with significant human disease, including Choleraesuis, Dublin, Enteritidis, Newport, and Typhimurium, can carry virulence plasmids with genes contributing to attachment, immune system evasion, and other roles. The goal of this comprehensive review is to provide key information on the Salmonella virulence, including the contributions of genes encoded in SPIs and plasmids during Salmonella pathogenesis.
Article
Full-text available
Salmonella enterica serovar Typhimurium causes gastroenteritis and systemic infections in humans. For this bacterium the expression of a type III secretion system (T3SS) and effector proteins encoded in the Salmonella pathogenicity island-1 (SPI-1), is keystone for the virulence of this bacterium. Expression of these is controlled by a regulatory cascade starting with the transcriptional regulators HilD, HilC and RtsA that induce the expression of HilA, which then activates expression of the regulator InvF, a transcriptional regulator of the AraC/XylS family. InvF needs to interact with the chaperone SicA to activate transcription of SPI-1 genes including sicA, sopB, sptP, sopE, sopE2, and STM1239. InvF very likely acts as a classical activator; however, whether InvF interacts with the RNA polymerase alpha subunit RpoA has not been determined. Results from this study confirm the interaction between InvF with SicA and reveal that both proteins interact with the RNAP alpha subunit. Thus, our study further supports that the InvF/SicA complex acts as a classical activator. Additionally, we showed for the first time an interaction between a chaperone of T3SS effectors (SicA) and the RNAP.
Article
Full-text available
A current view on the role of the Salmonella virulence plasmid in the pathogenesis of animal and human hosts is discussed; including the possible relevance in secondary ecological niches. Various strategies towards further studies in this respect are proposed within the One Health Concept.
Article
Full-text available
Type III secretion systems (T3SSs) are protein transport nanomachines that are found in Gram-negative bacterial pathogens and symbionts. Resembling molecular syringes, T3SSs form channels that cross the bacterial envelope and the host cell membrane, which enable bacteria to inject numerous effector proteins into the host cell cytoplasm and establish trans-kingdom interactions with diverse hosts. Recent advances in cryo-electron microscopy and integrative imaging have provided unprecedented views of the architecture and structure of T3SSs. Furthermore, genetic and molecular analyses have elucidated the functions of many effectors and key regulators of T3SS assembly and secretion hierarchy, which is the sequential order by which the protein substrates are secreted. As essential virulence factors, T3SSs are attractive targets for vaccines and therapeutics. This Review summarizes our current knowledge of the structure and function of this important protein secretion machinery. A greater understanding of T3SSs should aid mechanism-based drug design and facilitate their manipulation for biotechnological applications.
Article
Full-text available
Salmonella enterica replicates in macrophages through the action of effector proteins translocated across the vacuolar membrane by a type III secretion system (T3SS). Here we show that the SPI-2 T3SS effector SpvD suppresses proinflammatory immune responses. SpvD prevented activation of an NF-ĸB-dependent promoter and caused nuclear accumulation of importin-α, which is required for nuclear import of p65. SpvD interacted specifically with the exportin Xpo2, which mediates nuclear-cytoplasmic recycling of importins. We propose that interaction between SpvD and Xpo2 disrupts the normal recycling of importin-α from the nucleus, leading to a defect in nuclear translocation of p65 and inhibition of activation of NF-ĸB regulated promoters. SpvD down-regulated pro-inflammatory responses and contributed to systemic growth of bacteria in mice. This work shows that a bacterial pathogen can manipulate host cell immune responses by interfering with the nuclear transport machinery.
Article
Full-text available
Pathogenicity islands (PAIs) are regions of the chromosome of pathogenic bacteria that harbor virulence genes, which were probably acquired by lateral gene transfer. Several PAIs can excise from the bacterial chromosome by site-specific recombination and in this review have been denominated "excisable PAIs". Here, the characteristic of some of the excisable PAIs from Salmonella enterica and the possible role and impact of the excision process on bacterial virulence is discussed. Understanding the role of PAI excision could provide important insights relative to the emergence, evolution and virulence of pathogenic enterobacteria.
Article
Full-text available
Salmonella Typhimurium is one of the leading serovars that causes salmonellosis worldwide. However, few studies have molecularly characterised S. Typhimurium strains in Brazil. In this study, we genotyped 92 S. Typhimurium strains isolated from humans(43) and food(49) between 1983-2013 in Brazil using PFGE, MLVA and ERIC-PCR. Moreover, we assessed the frequency of 12 virulence markers by PCR and the resistance profile against 12 antimicrobials. More than 85.8% of the strains studied carried 11 of the virulence markers or more. Thirty-three (25%) strains were multi-drug resistant (MDR). The 92 S. Typhimurium studied were grouped by PFGE in PFGE-A, PFGE-B1 and PFGE-B2; by MLVA in MLVA-A, MLVA-B1 and MLVA-B2 and, finally, were grouped by ERIC-PCR in ERIC-A and ERIC-B. The strains isolated from humans before the mid-1990s were allocated in all clusters. The strains isolated from humans after the mid-1990s were distributed in the PFGE-B1, MLVA-B1, MLVA-B2 and ERIC-A clusters. The strains isolated from food were distributed in all clusters, except in the PFGE-B2. All typing results suggest that the S. Typhimurium strains of human clinical origin isolated before the mid-1990s were genetically more diverse, which might indicate the selection of a more adapted S. Typhimurium subtype after S. Enteritidis became the most prevalent serovar in Brazil. Regarding strains isolated from food, the results suggest the current circulation of more than one subtype. Furthermore, the high frequency of virulence genes and the presence of MDR strains reinforces their potential hazard for humans and the risk of their presence in foods in Brazil.
Article
Salmonella enterica remains an important food borne pathogen in all regions of the world with S. Typhimurium as one of the most frequent serovars causing food borne disease. Since the majority of human cases are caused by food of animal origin, there has been a high interest in understanding how S. Typhimurium interacts with the animal host, mostly focusing on factors that allow it to breach host barriers and to manipulate host cells to the benefit of itself. Up to recently, such studies have ignored the metabolic factors that allow the bacteria to multiply in the host, but this is changing rapidly, and we are now beginning to understand that virulence and metabolism in the host are closely linked. The current review highlights which metabolic factors S. Typhimurium that are essential for Salmonella growth in the intestine, in cultured epithelial and macrophage-like cell lines, at systemic sites during invasive salmonellosis, and during long term asymptomatic colonization of the host. It also points to the limitations in our current knowledge, most notably that most studies have been carried out with few well-characterized laboratory strains, that we do not know how much the in vivo metabolism differs between serotypes, and that most results are based on challenge in the mouse model of infection. It will be very important to realize whether the current understanding of Salmonella metabolism in the host is true for all serotypes and all possible hosts.
Article
Salmonella enterica serovar typhimurium (S. typhimurium) is globally distributed and causes massive morbidity and mortality in humans and animals. S. typhimurium carries Salmonella plasmid virulence (spv) locus, which is highly conserved and closely related to bacterial pathogenicity, while its exact role in host immune responses during infection remains to be elucidated. To counteract the invaders, the host has evolved numerous strategies, among which the innate immunity and autophagy act as the first defense. Recently, zebrafish has been universally accepted as a valuable and powerful vertebrate model in analyzing bacteria-host interactions. To investigate whether spv locus enhances the virulence of Salmonella by exerting an effect on the host early defense, zebrafish larvae were employed in this study. LD50 of S. typhimurium to zebrafish larvae and bacterial dissemination were analyzed. Sudan black B and neutral red staining were performed to detect the responses of neutrophils and macrophages to Salmonella infection. Autophagy agonist Torin1 and inhibitor Chloroquine were used to interfere in autophagic flux, and the protein level of Lc3 and p62 were measured by western blotting. Results indicated that spv locus could decrease the LD50 of S. typhimurium to zebrafish larvae, accelerate the reproduction and dissemination of bacteria by inhibiting the function of neutrophils and macrophages. Moreover, spv locus restrained the formation of autophagosomes in the earlier stage of autophagy. These findings suggested the virulence of spv locus involving in suppressing host innate immune responses for the first time, which shed new light on the role of spv operon in Salmonella pathogenicity.
Article
Salmonella enterica serovar typhimurium (S. typhimurium) is a facultative intracellular pathogen that can cause gastroenteritis and systemic infection in a wide range of hosts. Salmonella plasmid virulence gene spvB is closely related to bacterial virulence in different cell and animal models, and the encoded protein acts as an intracellular toxin required for ADP-ribosyl transferase activity. However, until now there is no report about the pathogenecity of spvB gene on zebrafish. Due to the outstanding advantages of zebrafish in analyzing bacteria-host interactions, a S. typhimurium infected zebrafish model was set up here to study the effect of spvB on autophagy and intestinal pathogenesis in vivo. We found that spvB gene could decrease the LD50 of S. typhimurium, and the strain carrying spvB promoted bacterial proliferation and aggravated the intestinal damage manifested by the narrowed intestines, fallen microvilli, blurred epithelium cell structure and infiltration of inflammatory cells. Results demonstrated the enhanced virulence induced by spvB in zebrafish. In spvB-mutant strain infected zebrafish, the levels of Lc3 turnover and Beclin1 expression increased, and the double-membraned autophagosome structures were observed, suggesting that spvB can inhibit autophagy activity. In summary, our results indicate that S. typhimurium strain containing spvB displays more virulence, triggering an increase in bacterial survival and intestine injuries by suppressing autophagy for the first time. This model provides novel insights into the role of Salmonella plasmid virulence gene in bacterial pathogenesis, and can help to further elucidate the relationship between bacteria and host immune response.