2,946
Views
5
CrossRef citations to date
0
Altmetric
Review

Immune evasion and persistence in enteric bacterial pathogens

Article: 2163839 | Received 15 Aug 2022, Accepted 27 Dec 2022, Published online: 08 Jan 2023

ABSTRACT

The major function of the mammalian immune system is to prevent and control infections caused by enteropathogens that collectively have altered human destiny. In fact, as the gastrointestinal tissues are the major interface of mammals with the environment, up to 70% of the human immune system is dedicated to patrolling them The defenses are multi-tiered and include the endogenous microflora that mediate colonization resistance as well as physical barriers intended to compartmentalize infections. The gastrointestinal tract and associated lymphoid tissue are also protected by sophisticated interleaved arrays of active innate and adaptive immune defenses. Remarkably, some bacterial enteropathogens have acquired an arsenal of virulence factors with which they neutralize all these formidable barriers to infection, causing disease ranging from mild self-limiting gastroenteritis to in some cases devastating human disease.

This article is part of the following collections:
Enteric Bacterial Infections

Introduction

The mammalian gut microbiome of healthy individuals is one of the most complex microbial communities ever discoveredCitation1. In the case of humans, it is composed of trillions of bacteria from over 1000 different species, which are essential to the host in denying pathogens a foothold.Citation2 By housing these microbes, the host not only gains nutrients but also a formidable array of defenses against pathogens collectively termed colonization resistance. In addition to housing the endogenous microflora, the gastrointestinal tract and the mucosal tissue provide physical barriers intended to shield the bloodstream and systemic organs from enteropathogens. The dissemination of enteric pathogens to the systemic circulation was traditionally considered passive on the part of microbes but is increasingly being appreciated as an active, microbe-influenced or directed process. The innate immune system discriminates between self and foreign by recognizing certain pathogen-associated molecular patterns (PAMPs) that do not belong, including lipopolysaccharide (LPS), peptidoglycan, flagellin and microbial nucleic acids.Citation3 The responsible germline encoded receptors include Toll-like receptors that detect microbial components on the cell surface and Nod-like receptors that serve a similar function within the cytosol.Citation3 Following the detection of PAMPs, the pro-inflammatory MAPK and NF-κB signaling pathways are stimulated and inflammasomes are assembled and activated, which are multiprotein signaling complexes that coordinate antimicrobial defenses.Citation4 In addition to recognizing PAMPS, the innate immune system can also detect certain conserved pathogen induced processes to distinguish between pathogens and less threatening commensal microbes.Citation4 These danger signals activate specific arms of the innate immune system, enabling the host to modulate inflammation according to the threat.Citation4 Enteropathogens must strike a careful balance between inducing inflammation in a way that enhances colonization early in infection and suppressing it at later times to prevent premature clearance and to in some cases allow systemic dissemination. There is a shared underlying logic through which enteropathogens surmount the considerable host barriers to infection. All the pathogens discussed here overcome colonization resistance, modulate the proinflammatory MAPK family and NF-κB, both activating them at times and suppressing them or their effects at others to skew the innate immune response to the pathogen’s benefit. They also manipulate cell death pathways in a manner that promotes virulence, triggering either an inflammatory cell death and/or inducing non-inflammatory apoptosis. Finally, they evade adaptive immunity. These aspects of infectious disease along with the emerging theme of virulence genes with allelic variants that dictate how the pathogens harboring them interact with the immune system are discussed. The following subset of prevalent bacterial enteropathogens: Salmonella, Shigella, Campylobacter, Yersinia and Vibrio are included.

Salmonella

Of the enteric bacterial pathogens, perhaps the best studied is Salmonella. The genus is composed of pathogens of diverse animals as well as humans, responsible for both acute and chronic infections. There are two species: Salmonella bongori and S. enterica. There are several subspecies and over 2,000 serovars. These serovars are typically classified into one of the two groups, those causing typhoid fever in humans and ones that are typically responsible for gastroenteritis. The typhoidal group includes S. Typhi, S. Paratyphi and S. Sendai, all three of which are human restricted. There are over 14 million cases of typhoid and paratyphoid fever per year world-wide.Citation5 Some of the more common non-typhoidal serovars that cause disease in humans are S. Typhimurium, S. Dublin and S. Enteritidis. The Centers for Disease Control estimates that there are 1.3 million infections with these serovars in the U.S. alone and invasive disease with them, mostly Typhimurium and Enteritidis, is on the rise and is especially relevant in sub-Saharan Africa, with a high case fatality rate.Citation6 This is typically seen in immunocompromised hosts, but also occasionally in the otherwise healthy. This section provides an overview of how Salmonella interacts with the immune system in terms of overcoming colonization resistance, manipulating inflammation, overcoming physical barriers, modulating host cell death pathways and what allelic variation contributes to these aspects of its virulence. For a more comprehensive description of all Salmonella virulence factors, further reading is suggested.Citation7–10

Animal models of disease

There are a variety of animal models for studying different aspects of Salmonella pathogenesis. S. enterica infection in these systems produces intestinal inflammation and diarrhea, or acute systemic disease that resembles typhoid fever in humans or chronic systemic disease. The disease symptoms depend on both the serotype of the Salmonella strain used and host susceptibility.Citation11 Infection of cows with S. Typhimurium produces enterocolitis that includes intestinal inflammation and diarrhea.Citation12 Both acute and chronic infections of humans with S. Typhi can be modeled in mice. S. Typhimurium infection of SLC11A1 (NRAMP1) mutant mice with S. Typhimurium causes acute systemic disease. Typhimurium literally means Typhi of mice. Infection of wild-type mice with S. Typhimurium produces a chronic, systemic disease with long-term carriage as sometimes occurs with S. Typhi in humans. Intestinal inflammation and diarrhea are not observed with non-typhoidal Salmonella infections of mice unless they are pre-treated with streptomycin.Citation13

Type III secretion systems

Salmonella possesses two independent type III secretion systems encoded by Salmonella pathogenicity island-1 and Salmonella pathogenicity island-2. S. Typhimurium utilizes Salmonella pathogenicity island-1 in the gastrointestinal stage of disease to invade enterocytes and invoke acute inflammation that enhances its growth in the lumen of the gut. Salmonella was traditionally thought to only deploy Salmonella pathogenicity island-2 in the systemic phase of disease, to facilitate intracellular survival and proliferation; however, more recent studies revealed that there is a fair degree of temporal and functional overlap between the two systems.Citation14–17

Overcoming colonization resistance

An inflammatory response is intended to be protective; however, in the case of S. Typhimurium infections, it is deliberately induced and enhanced by the pathogen and is crucial for its colonization of the intestinal tract.Citation18,Citation19 The inflammation depletes the resident gut microflora of the host enabling Salmonella to overcome colonization resistance, by reducing competition for essential nutrients. It also provides Salmonella with new carbon sources and electron acceptors that facilitate its outgrowth. It generates reactive oxygen species that produce a new respiratory electron acceptor in tetrathionate that Salmonella uses to support its growth on ethanolamine, which cannot be used by competing bacteria.Citation20,Citation21 Salmonella can use the Salmonella pathogenicity island-1 associated effector SopE to generate an energetically more favorable respiratory electron acceptor in host-derived nitrate.Citation22 S. enterica also overcomes colonization resistance with the deployment of a type VI secretion system encoded with the conserved Salmonella pathogenicity island-6. Salmonella uses its type 6 secretion system as an antibacterial weapon and it is required for colonization of the mouse gut.Citation23 For more information on how Salmonella interacts with the gut microbiota, the reader is directed to another recent review.Citation24

Overcoming physical barriers

Infections with non-typhoidal Salmonella are usually self-limiting when confined to the gastrointestinal tract but are fatal 20% of the time with hospitalization after entering the blood.Citation6 To reduce fatality, this is the most logical step to therapeutically interdict. Salmonella can breach the mucosal barrier, spread to the blood and subsequently internal organs through multiple pathways. The relevant contributions of the various pathways are unclear and in need of additional study. The various strategies used by enteropathogens to breach the gastrointestinal epithelium included in this review are summarized in .

Figure 1. Strategies used by four enteropathogens to breach the gastrointestinal epithelium. A) Salmonella and likely Yersinia can traverse the physical barrier within CX3CR1+ phagocytes that send dendrites through the paracellular space without disrupting the tight junctions. SrfH seems to regulate this process and Yersinia may require invasin for it to be efficiently utilized. B) Campylobacter cleaves E-cadherin, occludin and claudin-8 with HtrA to destabilize tight junctions and pass in between epithelial cells. C) Salmonella and Campylobacter can also pass through enterocytes. Salmonella requires both SPI-1 and SPI-2 effectors for this while Campylobacter enhances it with sialylated lipooligosaccharide cell structures. D) Salmonella and Shigella can recruit polymorphonuclear neutrophils to migrate through the paracellular space to gain access to the submucosa independently of the M cells. E) Finally, Salmonella, Shigella and Yersinia can pass through the M cells. This figure was generated with Biorender.

Figure 1. Strategies used by four enteropathogens to breach the gastrointestinal epithelium. A) Salmonella and likely Yersinia can traverse the physical barrier within CX3CR1+ phagocytes that send dendrites through the paracellular space without disrupting the tight junctions. SrfH seems to regulate this process and Yersinia may require invasin for it to be efficiently utilized. B) Campylobacter cleaves E-cadherin, occludin and claudin-8 with HtrA to destabilize tight junctions and pass in between epithelial cells. C) Salmonella and Campylobacter can also pass through enterocytes. Salmonella requires both SPI-1 and SPI-2 effectors for this while Campylobacter enhances it with sialylated lipooligosaccharide cell structures. D) Salmonella and Shigella can recruit polymorphonuclear neutrophils to migrate through the paracellular space to gain access to the submucosa independently of the M cells. E) Finally, Salmonella, Shigella and Yersinia can pass through the M cells. This figure was generated with Biorender.

S. Typhimurium rapidly and preferentially invades the M cells of murine ligated intestinal loops. The destruction of the M cells allows the bacteria to invade adjacent enterocytes.Citation25 Salmonella infection of the epithelium induces the basolateral secretion of IL-8 that recruits polymorphonuclear neutrophils from the gut associated microvasculature. Tight junctions are integral to the mucosal barrier’s ability to promote intestinal homeostasis, regulating water and solute flow through the paracellular space and preventing the invasion of pathogens.Citation26 The type III effector SipA stimulates the synthesis of the chemoattractant hepoxilin A3 (HXA3), which is secreted apically by the epithelium. The bioactive eicosanoid gradient across the epithelium promotes the transmigration of polymorphonuclear neutrophils from the lamina propria to the luminal side of the barrier through the paracellular space.Citation27,Citation28 Interestingly, the ability of SipA to induce HXA3 synthesis is confined to the N-terminal domain, while the C-terminal region is responsible for the polymerization of actin, which was shown to be important for invasion into the epithelial cells via ruffling.Citation29–31 The two domains are not just functionally distinct but are also physically separated by host caspase-3 cleavage, which may have interesting evolutionary implications.Citation31 Salmonella further disrupts the tight junctions of the epithelium with the type III effector SpvB.Citation32

The pathways exploited by Salmonella and likely other enteropathogens to spread from the gut to the bloodstream are summarized in . Many enteropathogens disseminate to the blood through the lymphatic system. The underlying phagocytes can carry microbes to the regional nodes where secondary immune responses are generated. Microbes that survive within them can eventually drain from the lymph nodes into the systemic circulation inside of them and thereby passively spread throughout the body. Some reports, however, have challenged this classic model for the extraintestinal dissemination of gut pathogens. Salmonella and Yersinia were found to spread systemically in lymphotoxin ß-receptor knockout mice, which completely lack Peyer’s patches and in congenic control mice similarly.Citation13,Citation33 Moreover, some studies have provided evidence that the mesenteric lymph nodes act as a firewall, largely containing oral infections, allowing the generation of a local immune response while shielding the host from systemic, microbial dissemination.Citation33,Citation34 While the availability of migratory dendritic cells is the rate-limiting step in mesenteric lymph node colonization, modulation of dendritic cell numbers or migratory properties within the lymphatic system in one such study had no effect on bacteremia.Citation34 In this report, the authors showed that while in vivo FLT-3 L-induced expansion of dendritic cell numbers, as well as stimulation of dendritic cell migration by TLR agonists, results in increased numbers of S. Typhimurium bacteria reaching the mesenteric lymph nodes, there was no increase in microbial dissemination to deeper tissue.Citation34 The authors also showed that in CCR7-deficient mice, very few bacteria reach the nodes, but there was no obvious defect in hepatosplenic colonization.Citation34 The classic view and this study could in part be explained by the observation of another group of extracellular Salmonella autonomously traveling to the mesenteric lymph nodes.Citation35 Extracellular Salmonella draining through the thoracic duct into the bloodstream would make it unnecessary for intracellular Salmonella to modulate the surface expression of host proteins, such as sphingosine-1-phosphate receptor-1. This receptor is up-regulated by the host in response to infection presumably to trap infected cells within the nodes to fight the infection more effectively and guard against sepsis.Citation36

Figure 2. Pathways for Salmonella and likely other gut pathogens to spread systemically. A) Reverse transmigration. Infected phagocytes traverse the blood vascular endothelium in the basal to apical direction following uptake of gut bacteria. The tight junctions of the epithelium and the endothelium remain intact. B) Salmonella can pass through the epithelial cells and may be able to trigger lamina propria phagocytes to reverse transmigrate or alternatively drain through the lymphatics to the blood. C) Salmonella can invade the endothelium and manipulate ß-catenin/Wnt signaling, rendering the blood vessels permeable. D) Salmonella can pass through the M cells and drain through the lymphatics to the blood or perhaps trigger the reverse transmigration of infected phagocytes through the high endothelial venules associated with the Peyer’s patches and/or mesenteric lymph nodes. E) S. Typhi rapidly destroys epithelial cells and then drains through the lymphatics to the blood and/or triggers the reverse transmigration of infected phagocytes. This illustration was generated with Biorender.

Figure 2. Pathways for Salmonella and likely other gut pathogens to spread systemically. A) Reverse transmigration. Infected phagocytes traverse the blood vascular endothelium in the basal to apical direction following uptake of gut bacteria. The tight junctions of the epithelium and the endothelium remain intact. B) Salmonella can pass through the epithelial cells and may be able to trigger lamina propria phagocytes to reverse transmigrate or alternatively drain through the lymphatics to the blood. C) Salmonella can invade the endothelium and manipulate ß-catenin/Wnt signaling, rendering the blood vessels permeable. D) Salmonella can pass through the M cells and drain through the lymphatics to the blood or perhaps trigger the reverse transmigration of infected phagocytes through the high endothelial venules associated with the Peyer’s patches and/or mesenteric lymph nodes. E) S. Typhi rapidly destroys epithelial cells and then drains through the lymphatics to the blood and/or triggers the reverse transmigration of infected phagocytes. This illustration was generated with Biorender.

Vasquez-Torres et al. elegantly defined a pathway of extraintestinal dissemination available to enteropathogens that is independent of the lymphatic system. The authors demonstrated that transgenic mice deficient in CD18, an integrin that mediates leukocyte transmigration, was 100-fold more resistant to the extraintestinal dissemination of a Salmonella mutant defective in invading the gut epithelium.Citation37 In this alternative pathway, luminal bacteria are internalized by CX3CR1+ lamina phagocytes dispersed throughout the lamina propria that send dendrites into the lumen of the gut as a component of antigen sampling.Citation37 Salmonella and perhaps other enteropathogens can then trigger the infected phagocytes to traverse the blood vascular endothelium in the basal to apical direction without disrupting the tight junctions of the endothelium.Citation15–17,Citation37,Citation38 Traversing the blood vascular endothelium in such a fashion without disrupting the tight junctions is termed reverse transmigration when it occurs with uninfected cells, which typically does not occur with infected ones.Citation15 With Salmonella-infected phagocytes, the process requires the type III effector SpvC.Citation15

In a parallel pathway, in the murine model of gastroenteritis, following Salmonella pathogenicity island-1 mediated apical invasion of enterocytes, Salmonella pathogenicity island-2 type III effectors facilitate the trafficking of the bacteria to the basal side of the epithelium. The bacteria then exocytose to the lamina propria where they are briefly extracellular before being internalized by phagocytes.Citation39 The authors proposed that phagocytes sample the contents of the gastrointestinal epithelium as part of a constitutive antigen sampling pathway that Salmonella exploits.Citation39 It would be interesting and important to assess in a different model of disease if Salmonella colonized the lamina propria and perhaps accessed deeper tissue. This could occur by infected phagocytes draining through the portal vein or the thoracic duct or reverse transmigrating through the gut associated blood microvasculature.

In another pathway, in the murine model of typhoid fever, Salmonella can access the blood after increasing the permeability of the blood vasculature associated with the gut by manipulating ß-catenin/Wnt signaling with Salmonella pathogenicity island-2.Citation40 When transgenic mice expressing a degradation resistant ß-catenin allele within endothelial cells were infected with Salmonella, no extracellular dissemination was observed. ß-catenin/Wnt signaling, however, affects many diverse cellular processes.Citation41

There are clearly myriad ways in which enteropathogens can disseminate extra-intestinally and an important challenge for the field is assessing the relative contributions of each. A limitation in many of these studies is that when one pathway is blocked and no defect is observed, more of the inoculum may simply go through the other pathways if they are not normally saturated. This may mask the contribution of the one that is unavailable under certain conditions. The interpretation of these studies is further complicated by the use of mouse and bacterial strains of different genetic backgrounds as well as the use of different time points.

It is interesting to consider that there may be pathways to the blood from the gut available to enteropathogens that remain to be discovered. When CD18 mutant mice are infected with Salmonella pathogenicity island-1 mutant bacteria that are deficient in invasion, Salmonella is still able to colonize hepatosplenic tissue in significant numbers.Citation37 This suggests the existence of an additional pathway to the bloodstream that does not require reverse transmigration or Salmonella pathogenicity island-1 mediated invasion of the epithelium or endothelium. Two possibilities are the invasins PagN and Rck, which can facilitate invasion of epithelial cells independently of Salmonella pathogenicity island-1.Citation42,Citation43 A unifying observation is that Salmonella pathogenicity island-2 and the anti-inflammatory effectors SpvC and SpvD are required for any systemic dissemination in the murine model of Salmonella infection, even though they are dispensable for intracellular survival.Citation15,Citation44–46 They also appear to be required to cause bacteremia in humans with non-typhoidal serovars.Citation15,Citation44,Citation47–49

Activation and suppression of inflammation

The pathogen must strike a careful balance between inducing inflammation initially and later dampening it to prevent the infection from being prematurely cleared and to allow dissemination to deeper tissue. Non-typhoidal serovars typically trigger inflammatory responses, whereas the typhoidal ones suppress it. Salmonella, like other enteric bacterial pathogens, possesses a large repertoire of secreted effectors, which either induce or down regulate the inflammatory response of infected hosts. S. Typhimurium must bypass the negative regulatory mechanisms of the host intended to prevent constitutive activation of innate immune receptors in the gut. The mechanisms presumably exist to prevent chronic inflammation resulting from exposure to common bacterial products from harmless members of the microflora. This is presumably to minimize the chances of autoimmune disorders such as Crohn’s disease or inflammatory bowel disease.

The features of the more prominent pro-inflammatory effectors of the five genera of enteric bacterial pathogens included in this review are summarized in . The three Salmonella pathogenicity island-1 effectors SopE, SopE2 and SopB circumvent host inflammation attenuating mechanisms by activating the Rho family GTPases RAC1 and CDC4 and a non-canonical signaling complex downstream of Toll-like receptors.Citation50,Citation53,Citation54,Citation76,Citation77 This ultimately leads to MAPK and NF-κB signaling and the production of pro-inflammatory cytokines. Importantly, inhibiting the downstream signaling hub attenuates the growth of S. Typhimurium in the gut but increases the number of bacteria that translocate to systemic sites.Citation53

Table 1. Major pro-inflammatory effectors.

The effectors SopA and SopD amplify the inflammatory responses induced by SopE, SopE2 and SopB. SopA is a ubiquitin ligase that activates TRIM56 and TRIP65 to induce interferon genes and inflammation.Citation78 SopD enhances inflammation without engaging innate immune receptors. This effector interdicts an anti-inflammatory pathway downstream of Toll-like receptors that the host uses to limit tissue damage after fighting an infection that depends on RAB8 and AKT.Citation79,Citation80

Salmonella possesses mechanisms to not only induce inflammation to promote colonization and replication but also the ability to attenuate it. The functions of the more prominent effectors of the five enteropathogens discussed in this review that reign in the inflammatory response elicited by other effectors are described in . Preserving host homeostasis may seem counterintuitive at first. They may have evolved to promote a long-standing association with the host. They also may dampen inflammation in a spatiotemporally regulated fashion to allow intracellular pathogens capable of causing systemic illness the opportunity to spread to deeper tissue after successfully overcoming colonization resistance. Deleting some type III effectors results in diminished intestinal disease but enhances spread to systemic sitesCitation84,Citation111,Citation112 These effectors antagonize the actions of the pro-inflammatory ones by directly countering their effects or alternatively by activating anti-inflammatory pathways.

Table 2. Major anti-inflammatory effectors.

SptP is a member of the first group, which serves as a GTPase activating protein for CDC42, RAC1 and RhoA. It directly counters the effectors SopE and SopE2 that are guanine nucleotides exchange factors for these GTPases.Citation81 A subset of effectors composed of PipA, GtgA and GogA form a family of related proteases that cleave RELA and RELB, which are transcription factors for NF-κB.Citation84 The importance of limiting the inflammatory response in this fashion is indicated by the presence of at least one member of this set of genes in all characterized Salmonella isolates.Citation84

The effectors SseK1 and SseK3 also inhibit NF-κB signaling albeit with a different mechanism. These two effectors are arginine glycosyltransferase that modifies the death domains of proteins in the NF-κB signaling pathway.Citation87,Citation88 The S. Typhimurium effector protein AvrA possesses acetyltransferase activity toward MAPKKs that attenuate inflammation by inhibiting JNK and NF-κB signaling.Citation85,Citation86

Some Salmonella serovars carry plasmids, which share a highly conserved locus called the spv (Salmonella plasmid virulence) operon. In addition to acting on tight junctions, SpvB depolymerizes actin and was recently reported to inhibit NF-κB activity by downregulating IKKβ.Citation89 SpvC is a phosphothreonine lyase that dephosphorylates the MAP kinases Erk1/2, p38 and JNK.Citation90 SpvC is not required for survival within macrophages or the gastrointestinal tract but is necessary for non-typhoidal Salmonella to cause bacteremia in mice and humans.Citation15,Citation44,Citation49 SpvD is a cysteine hydrolase with a serovar-specific polymorphism that determines the degree to which it inhibits nuclear transport of NF-κB p65.Citation91,Citation113

An often-overlooked component of an inflammatory response is the inhibition of the movement of infected cells. Salmonella neutralizes this host checkpoint intended to prevent the spread of pathogen within a host in part with SpvC. This allows it to reach privileged sites of infection with reverse transmigration and likely with enhanced migration through the lymphatic system as well. It will be important to test if SpvC stimulates the migration of infected phagocytes through lymphatic vessels toward the nodes and/or reverse transmigration through the high endothelial venules associated with the Peyer’s patches and mesenteric lymph nodes.

A recent report established SopD as the newest member of an emerging class of bifunctional virulence factors that remarkably possess seemingly opposed activities, which raises interesting questions about the regulation of their activities and/or targets. SopD both induces and limits inflammation. It inhibits Rab8 with GTPase activity. This stimulates inflammation, but it can also activate it by displacing Rab8 from its cognate guanosine dissociation inhibitor, which is anti-inflammatory.Citation56 Both SopD and SopB activate a RAB8-dependent PI3K–PKB–mTOR pathway that is downstream of Toll-like receptors, which results in the production of the anti-inflammatory effector, IL-10.Citation56,Citation114–116

The effector SteE/SarA/GogC was recently shown in a pair of reports to mimic a cytokine receptor’s intracellular domain and reprogram a non-canonical downstream effector to induce the expression of anti-inflammatory rather than pro-inflammatory genes.Citation82,Citation83 SteE specifically targets signal transducer and activator of transcription 3 (STAT3), which the host uses among other things to recover after launching an inflammatory response.Citation117,Citation118 Interestingly, the mechanism does not involve the Janus kinases but rather the host kinase GSK3 that phosphorylates SteE.Citation82,Citation83

Modulation of host cell death

Salmonella modulates three different host cell death pathways, either inducing them or suppressing them, including apoptotic, necroptotic and pyroptotic ones to promote persistence.Citation119 Ultimately, Salmonella-infected epithelial cells undergo pyroptosis, which releases proinflammatory cytokines and releases the bacteria. It actively invades macrophages next to reduce the likelihood that it is internalized and killed by more microbicidal neutrophils. Several Salmonella pathogenicity island-1 and Salmonella pathogenicity island-2 secreted type III effectors have been implicated in inducing either rapid or delayed apoptosis of macrophages.Citation120–122 The two events were speculated to be either temporally or anatomically regulated within hosts and are genetically separable.Citation121 After inducing cell death, apoptotic cells harboring Salmonella are engulfed by new macrophages, creating a fresh intracellular niche for the pathogen, allowing continued persistence and avoidance of extracellular host defenses. For more information on the manipulation of host cell death pathways by Salmonella, another review is suggested.Citation119

Evasion of adaptive immune responses

In addition to subverting innate immunity, Salmonella also delays an adaptive immune response. S. Typhimurium can kill CD8α+ dendritic cells in mesenteric lymph nodes and can inhibit T cell proliferation by a direct, contact-dependent immunosuppressive effect, which inhibits the ability of T cells to produce cytokines and proliferate.Citation123,Citation124 CD4+ T cells specific for Salmonella have been identified shortly following infection; however, they do not play a role in battling the infection until several weeks later and this delay requires Salmonella pathogenicity island-2. Several Salmonella pathogenicity island-2 effectors interfere with antigen presentation by dendritic cells and in the case of SteD ubiquitinates MHC-II molecules.Citation8,Citation125,Citation126 In addition to downregulating antigen presentation by dendritic cells, Salmonella pathogenicity island-2 actively induces the apoptosis of antigen-specific CD4+ T cells.Citation127 For additional information on the evasion of adaptive immunity by Salmonella, additional reading is suggested.Citation8,Citation125

Persistent infections

One of the most interesting disease manifestations of host adapted strains of Salmonella in mammals is that of an asymptomatic, chronic infection. The bacteria overcome the formidable defenses of the gastrointestinal tract and lymphatic system to ultimately reside within macrophages at systemic sites for in some cases the lifetime of the host. This is a highly advantageous niche for the bacteria that are generally free of endogenous flora and rich in nutrients, and moreover the pathogen can be intermittently shed. Biofilm formation on the surface of gallstones is associated with the establishment and perpetuation of such infections with S. Typhi.Citation128 These infections are largely asymptomatic but can be a contributing factor in the development of gallbladder cancer.Citation128 The carriers are also a public health concern as the pools of bacteria may allow the generation of new genotypes, and they serve as reservoirs for dissemination to new hosts.Citation129

Once within macrophages at systemic sites, some S. Typhimurium cells enter a non-growing but transcriptionally and translationally active state that allows for long-term, chronic infections that are resistant to antibiotic treatment. These cells trigger a non-inflammatory M2 polarization of the infected macrophages with the Salmonella pathogenicity island-2 effector SteE, which antagonizes TNF-mediated pathogen restriction.Citation130,Citation131 In addition to manipulating the immune status of these cells, Salmonella also metabolically reprograms them to create a favorable environment for bacterial persistence. Among other changes, these cells are induced to express a high level of PPARS, a eukaryotic transcription factor involved in fatty acid metabolism, by increasing the availability of glucose.Citation132 For additional information on persistent Salmonella infections, other reviews are recommended.Citation129,Citation133,Citation134

A growing problem with Salmonella is the multi-drug resistance persistence strategy that is employed by all enteropathogens. The public health relevance of the typhoidal serovars is increasing with the rise of such resistance, seen in >60% of strains.Citation135 In addition to the persister state described above, strains of typhoidal Salmonella that have acquired specific resistance mechanisms to chloramphenicol, ampicillin, and trimethoprim have caused numerous outbreaks.Citation135 As a result of the widespread dissemination of such strains, chloramphenicol was withdrawn as the first-line drug for typhoid fever and replaced with fluoroquinolones and third generation cephalosporins.Citation136 However, outbreaks of typhoid fever caused by strains resistant to nalidixic acid and ciprofloxacin have become endemic in the Indian subcontinent and have also been reported in the US and UK among other developed countries, reflecting the emergence of a global problem.Citation137 The presence of a plasmid-borne integron in ciprofloxacin-resistant S. typhi may soon produce widespread instances of nearly intractable typhoid fever.Citation137

In a genome-wide screen for S. Typhimurium genes that promote chronic disease in a mouse model of long-term systemic infections, dozens of genes were implicated in facilitating chronic infections.Citation138 These genes are contained within Salmonella pathogenicity island-1, 2, 3, 4, 5 and 6.Citation138 Fimbriae and genes within the integrated phages GIFSY-1 and GIFSY-2 among others were also identified.Citation138 Many of the genes are of unknown or putative function. A Salmonella pathogenicity island-2 associated effector identified of particular interest in causing persistent infections is SrfH/SseI.Citation138 Some alleles of SrfH can deamidate the heterotrimeric G protein Gαi2, resulting in persistent activation through non-polarized activation of it.Citation139 This results in the loss of directed dendritic cell migration, perhaps through increased adhesion.Citation139,Citation140 This promotes the long-term colonization of wild-type mice, which do not typically succumb to infection as C57Bl/6 mice do.Citation140 This is perhaps due to a reduction in the directed migration of phagocytes along T cell chemoattractive gradients in systemic organs.Citation140 It also may suppress the de-adhesion and migration of infected cells within the lymphatic system and of antigen sampling dendritic cells in the reverse transmigration pathway.Citation16,Citation17,Citation140,Citation141

Role of allelic variants in infection dynamics

A fair amount of attention has been devoted to understanding what differences in gene content allow enteric microbes to cause disease versus non-pathogenic relatives. Additionally, studies have identified and characterized what such differences contribute to microbes from the same genera causing different types of disease, often differing in severity. However, very little has been invested in identifying and characterizing how changes between various alleles of the same genes influence virulence. Understanding how differences among naturally occurring alleles of virulence genes affect how the pathogens harboring them interact with the immune system and persist within hosts is an area in need of more study.

SrfH and SpvD were the first two of over a dozen genes for which non-synonymous differences among naturally occurring alleles were determined to influence whether Salmonella escapes confinement within the gastrointestinal tract.Citation16,Citation142 In a bioinformatics study with non-typhoidal Salmonella, either a dominant or a recessive allele of 22 genes was associated with invasive versus gastrointestinal disease.Citation142 Most of these genes play roles in mediating adhesion or inflammation. From the 22 effectors for which known allelic variants are associated with a particular phase of disease, a subset of three: SspH2, SlrP and SrfH were shown to inhibit dendritic cell chemotaxis toward CCL19 in a microfluidic device.Citation143 It is interesting to consider that anti-inflammatory effects have been attributed to all three effectors in some reports.Citation92,Citation144,Citation145

SrfH is particularly interesting because of the potential of some alleles, as with SopD, to possess not just bifunctional, but seemingly opposed activities. In strains primarily associated with gastrointestinal disease in humans, the catalytic triad in the C-terminus appears to suppress deadhesion, which can be at least partially alleviated with certain polymorphisms in the N-terminus found in more invasive strains and/or with ones within the catalytic site of the carboxyl terminus.Citation16,Citation17,Citation139–142 Intriguingly, hyper-invasive sequence type 313 isolates, which are responsible for epidemic multiple drug resistant invasive disease with S. Typhimurium in Sub-Sahara Africa, possess a disrupted SrfH allele.Citation141 This pseudogene contains the catalytic triad of the carboxyl terminus necessary for Gαi2 deamidation but lacks any of the compensatory polymorphisms seen in other highly invasive strains that seem to mask the effect of the carboxyl region on adhesion.Citation140–142 The different effects on adhesion and motility and extraintestinal dissemination reported for SrfH in various reports are likely attributable to the use of strains possessing different alleles of this intriguing effector.Citation16,Citation17,Citation140

Shigella

Shigella is responsible for an estimated 125 million diarrheal episodes annually that result in 160,000 deaths and is especially prevalent among young children.Citation146 The resulting disease, termed Shigellosis, is characterized by watery diarrhea, abdominal cramps and fever. The Shigella genus is composed of four major species: Shigella dysenteriae, Shigella flexneri, Shigella boydii, and Shigella sonnei. S. dysenteriae and S. flexneri are associated with poverty and poor hygiene, while S. sonnei is more common in affluent regions. Considering the low infective dose, on the order of 10 to 100 organisms, associated disease severity and increasing antimicrobial resistance, preventing shigellosis is a major public health issue. Shigella promotes its pathogenesis with chromosomal pathogenicity islands and a large virulence plasmid that harbors a type III secretion system (Mxi-Spa) and numerous effectors, including several invasion plasmid antigens (Ipas). These genomic features are key to its manipulation and neutralization of innate and adaptive host defenses. This section provides an overview of how Shigella interacts with the immune system in terms of overcoming colonization resistance, manipulating inflammation, evading adaptive immunity and modulating host cell death pathways. For a comprehensive discussion of all Shigella virulence factors, the reader is referred to other reviews.Citation96,Citation147–149

Animal models of shigellosis

Bacillary dysentery is a human-specific disease that follows the triggering of a strong inflammatory response after Shigella invades the colonic epithelium. A drawback in Shigella pathogenesis studies has been the lack of an appropriate animal model that mimics the natural course of infection and immune response in humans. Adult mice are refractory to oral infection; however, murine models were developed with intraperitoneal, as well as oral infection that involves a zinc deficient diet and pre-treatment with antibiotics.Citation150,Citation151 There is also an infant rabbit model available as is a guinea pig model that requires the intrarectal administration of bacteria.Citation152,Citation153

Overcoming colonization resistance

Shigella is inadvertently ingested through the fecal-oral route with contaminated food or water. The low infectious dose of Shigella indicates that it possesses potent mechanisms for overcoming colonization resistance. Shigella withstands gastric acid better than some of the other enteric pathogens, due primarily to its glutamate decarboxylase system.Citation154 Shigella disrupts the production of bactericidal peptides at the surface of the gastrointestinal epithelium, which is an important component of innate immunity.Citation155 Shigella resists colicins with WzzB, an O-antigen chain length regulator and moreover secretes colicins of its own.Citation156 Shigella enterotoxin 1 (ShET1) and Shigella enterotoxin 2 (ShET2) induce fluid secretion in the jejunum to establish infection. It produces the watery diarrhea seen early in shigellosis. There is no homology between ShET1 and ShET2. ShET1 is encoded by set1A and set1B genes contained within a chromosomal pathogenicity island only present in S. flexneri 2a isolates and are believed to form a holo-AB-type toxin complex.Citation157,Citation158 ShET2 is a 63 kDa protein encoded by ospD3 that induces IL-8 secretion by epithelial cells and is found in all serotypes.Citation159 ShET2 but not ShET1 is secreted by the type III secretion system.110 S. sonnei can additionally kill members of the endogenous microflora with a type VI secretion system and because of it outcompetes S. flexneri, which lacks it, in mixed infections of mice.Citation160 This secretion system may explain the increasing prevalence of dysentery caused by S. sonnei at the expense of S. flexneri, seen with global development.

Overcoming physical barriers

Shigella targets the M cells for transit through the epithelium. Following its release from the underlying macrophages, the bacteria can induce its internalization into enterocytes by triggering the reorganization of the cytoskeleton resulting in a macropinocytic event and subsequently spreads intra- and intercellularly. Recently, a bimodal model was proposed for the invasion of the colonic epithelium, following studies on a human in vitro M cell model. In this system, Shigella either rapidly transited the M cells or alternatively, perhaps to evade some of the immune surveillance functions of these cells, spreads laterally from the M cells to neighboring enterocytes.Citation161

Upon internalization within host cells, Shigella lyses the phagosome it is initially contained within to reach the cytosol, its preferred replicative niche, sheltered from many components of the immune system. Shigella initially re-programs enterocyte gene expression toward a strong pro-inflammatory profile that promotes polymorphonuclear neutrophil infiltration.Citation27 The integrity of the epithelium is destroyed by the polymorphonuclear neutrophils recruited by Shigella infection, which allow more luminal bacteria access to the submucosa independently of the M cells. This massive inflammation and destruction of the epithelium is necessary for the development of watery diarrhea and greatly benefits Shigella initially but the polymorphonuclear neutrophils ultimately resolve the infection and thus Shigella down-regulates inflammation in the final stage of infection.

Modulation of host cell death

Shigella can quickly kill submucosal macrophages by triggering pyroptosis, which frees the bacteria to invade the basal face of the colonic epithelium, but also invokes an intense inflammatory response.Citation162 Intriguingly, Shigella can also trigger a non-inflammatory macrophage apoptosis with IpaD.Citation163 Shigella secretes IpgD, which is homologous to Salmonella SopB into infected epithelial cell cytosol, where it dephosphorylates phosphatidylinositol 4,5-bisphosphate (PIP2) into phosphatidylinositol 5-phosphate (PI5P).Citation164 IpgD promotes host cell survival because PIP5 among other things activates AKT1.Citation165 The existence of both parallel pathways may be attributable to Shigella striking a delicate balance between infectiousness and immune evasion.

There are three forms of cell death regulated by intestinal epithelial cells that are important components of the innate immune response to intracellular bacteria. There is crosstalk among the three pathways to deny pathogens an intracellular niche even if they can suppress one of them. Various inflammasome complexes can initiate a lytic inflammatory cell death termed pyroptosis that results from caspase-1/caspase-4-mediated generation of gasdermin pores, whereas caspase-8/caspase-9 activation induces apoptosis. A lytic necroptotic death follows phosphorylation and trimerization of MLKL that causes membrane destabilization. In murine epithelial cells, loss of caspase-8, which prevents apoptosis, will lead to necroptotic cell death.Citation166

Regardless of the host cell death pathway redundancies, Shigella can prevent all forms of regulated epithelial cell death. Shigella uses the T3SS effector OspC3 early in infection to prevent caspase-4-dependent pyroptotic death.Citation167 Later, OspC1 interferes with the induction of apoptosis by preventing caspase-8 activation. The host senses caspase-8 inhibition and as a fail-safe induces necroptosis as a backup cell death pathway. It attempts to do this with RIPK/RIPK3 interaction and subsequent phosphorylation of MLKL, but Shigella deploys OspD3 to disrupt this pathway by cleaving RIPK1 and RIPK3.Citation168 Collectively, the actions of these effectors prevent the host from denying Shigella an intracellular niche. For more information on the modulation of host cell death by Shigella, a recent review is suggested.Citation169

Overcoming innate defenses

Once free in the epithelial cytosol, Shigella inhibits ER to Golgi traffic with VirA and IpaJ, which mainly inactivate Rab1 and ARF6, respectively, to escape autophagic host defense.Citation170,Citation171 Spreading intercellularly within the colonic epithelium is important for renewing Shigella’s preferred replicative niche and for evading immunological detection. IcsA (VirG) is important for actin-based motility and spread within the epithelium. Actin-based motility mediated by this autotransporter results in the membrane protrusions that allow the invasion of one cell in an epithelium from an adjacent one.Citation172 IcsB prevents sequestration by autophagy that would be triggered by the effector IcsA by camouflaging its Atg5 binding site.Citation173

The bacteria neutralize the innate defense against invading microbes of rapid mucosal epithelium turnover and exfoliation by hijacking integrin-linked kinase with the virulence factor OspE3 to stabilize focal adhesions and block cell detachment.Citation174 OspE cognate genes are conserved among many enteropathogens, including enteropathogenic E. coli, enterohaemorrhagic E. coli, Citrobacter rodentium and Salmonella.Citation149 Shigella deploys IpaB to further preserve its cytosolic niche by inhibiting cell division. Shigella initially re-programs enterocyte gene expression toward a strong pro-inflammatory profile that promotes polymorphonuclear neutrophil infiltration.Citation175

Activation and suppression of inflammation

Shigella skillfully skews the immune response to infection in opposing ways in a temporal fashion. Initially, some type III effectors are secreted to phosphorylate and activate MAPK pathways.Citation149 The ensuant increased apical secretion of the chemoattractant IL-8 results in polymorphonuclear neutrophil transmigration across the epithelium, destabilizing the barrier and allowing luminal Shigella access to the submucosa independently of the M cells. The massive inflammation promotes early infection, but the bacteria subsequently secrete effectors that downregulate proinflammatory signals, perhaps to balance the severity of the inflammation to a level beneficial for the bacteria and to render the host partially susceptible to re-infection. IpgB1 and IpgB2 are guanine nucleotide exchange factors (GEF) that target Rac1 and RhoA, respectively, which results in membrane ruffling.Citation57 Interestingly, IpgB2 can activate NOD1 signaling with ensuant NF-κB activation independently of its GEF activity.Citation176

In the instance of OspF, similar to Salmonella SopB, it may both increase and dampen inflammation with the same effector at different times during the infection. The S. flexneri type III effector protein OspF is homologous to Salmonella SpvC and Pseudomonas syringae HopAI.Citation177 This family of enzymes permanently deactivates the MAPK signaling pathways with phosphothreonine lyase activity.Citation90 Additionally, in the guinea pig model of Shigellosis, OspF diminishes the phosphorylation of HP1γ, a member of the heterochromatin protein 1 family of epigenetic regulators that modulate the amplitude of the innate immune response.Citation178 OspF, in addition to attenuating inflammation, interestingly has also been attributed pro-inflammatory roles.Citation179 This is yet another example of the same effector having opposite effects on the same host process at different times. It will be important to identify all the targets of OspF and to determine how these targets and/or the activities of this effector are regulated.

OspB

OspB is a type III effector that stimulates inflammation by activating p38, ERK1/2, and phospholipase A2 (PLA2). PLA2 stimulates the secretion of chemoattractant and IL-8.Citation58 PLA2 activation also mediates eicosanoid generation, which enhances inflammation.Citation180 OspB also modifies the intracellular niche to the advantage of Shigella by manipulating the mechanistic target of rapamycin complex 1, which is a master regulator of cell growth and proliferation.Citation181

IpaH family

The IpaH family of type III effectors is composed of 12 members that reside on either the chromosome or the large plasmid that target NF-κB signaling.Citation182 These virulence factors possess N-terminal leucine-rich repeats (LRR) that are recognized as a pathogen-associated molecular pattern by the infected host cell and conserved E3 ubiquitin ligase activity in their C-terminal regions.Citation182,Citation183 While the members of this family of Shigella effectors share homology, their LRR regions, which are the substrate recognition sites, differ as does their subcellular localization. This implies that they have different host targets and make distinct contributions to disease. A variety of pathogens have homologs of the IpaH family, including Edwardsiella, Bradyrhizobium, Rhizobium, and some Pseudomonas species and also Salmonella (SspH1, SspH2, SspH3 and SlrP).Citation182,Citation184

Shigella IpaH7.8 is an interesting member of the IpaH family of ubiquitin ligases, which stimulates the proteasomal degradation of human but not murine gasdermin D, thereby blocking pyroptosis in Shigella-infected human cells. It appears to function differently in mouse cells.Citation94IpaH1.4 and IpaH2.5 are homologous E3 ligases that irreversibly inactivate the catalytic subunit of the linear ubiquitin chain assembly complex, HOIP. Normally, the association of the linear ubiquitin chain assembly complex with activated cytokine or pattern recognition receptors results in the ligation of ubiquitin chains to RIPK1 and IKKγ. This event is necessary for the activation of NF-κB. Thus, these two effectors are anti-inflammatory because they attenuate NF-κB nuclear translocation in response to pathogen-associated molecular patterns, TNF as well as IL-1β.Citation95

IpaH9.8 dampens NF-κB signaling by targeting NEMO/IKKγ and also ubiquitinates a subset of guanylate-binding proteins which are members of the dynamin superfamily of GTPases important for protecting cells from intracellular pathogens that escape into the cytosol.Citation96–98 IpaH9.8 was also reported to bind to a splicing factor whose depletion from cells lowers the levels of pro-inflammatory cytokines.Citation185

Among other targets for the ubiquitin ligase IpaH4.5 is TANK-binding kinase 1 (TBK1). TBK1 is involved in the regulation of IFN regulatory factors and activates interferon regulatory factor 3 in response to infection. IpaH4.5 promotes the K48-linked polyubiquitylation of TBK1 resulting in its proteasomal degradation and thereby attenuates the host antibacterial response.Citation99 IpaH0722 inhibits NF-κB activity by targeting TRAF2 for degradation, which lies between PKC and NF-κB.Citation100

The OspC family, OspI, OspZ and OspG, also has anti-inflammatory effects. The OspC type III effectors were recently shown to block interferon signaling independently of inhibiting cell death by blocking calmodulin kinase II and downstream JAK/STAT signaling and preventing STAT1 phosphorylation.Citation186 OspG is a serine/threonine kinase homologous to EPEC NleH that prevents the ubiquitination of phospho-IκBα that is necessary for NF-κB activity.Citation101,Citation102 OspI deamidates UBC13 to blocks TRAF6-dependent NF-κB signaling.Citation103 The nuclear localization of the NF-κB subunit p65 is blocked by OspZ.Citation60

Evading adaptive immune responses

Shigella targets lymphocytes, including activated human peripheral blood B cells, CD4+ T, and CD8+ T lymphocytes as well as switched memory B cells with type III effectors without triggering internalization.Citation187 This direct targeting of lymphocytes independent of invasion may contribute to an inefficient priming of the adaptive immune response. Shigella infection of human B and T lymphocytes impairs T cell chemotaxis and the ability of these cells to scan antigen presenting cells. This may be attributable to the targeting of intracellular vesicular trafficking by IpaJ and VirA.Citation188

B cells were shown to undergo apoptosis independently of both invasion and type III effector injection following the activation of TLR2-dependent signaling by the type III secretion system needle-tip protein, which was termed the ‘kiss-and-run’ strategy.Citation189,Citation190 This may delay the adaptive immune response of mucosal sIgAs and systemic IgGs, which can mediate immunity to Shigella. Immunity to Shigella infection is ultimately due to an immune response against the LPS O-antigen and Ipa proteins among other molecules.Citation190 Shigella evades host immunity and renders a recovered host at least partially susceptible to reinfection by varying the composition of the O-antigens across strains and serotypes.Citation190 In addition to TBK-1, the ubiquitin ligase IpaH 4.5 also induces the degradation of the proteasome regulatory particle non-ATPase 13 (RPN13). This suppresses proteasome-catalyzed peptide splicing that reduces antigen cross-presentation to CD8+ T cells in vitro and in mice.Citation191

The O-antigen of Shigella LPS may function as an immunological decoy in multiple ways that dampens the adaptive immune response. First, LPS is serotype specific. Second, the O-antigen is a carbohydrate that is a thymus independent type 1 antigen, which activates B cells without helper T-cell augmentation. This prevents class switching or a memory B-cell response to protect against re-infection.Citation192 O-antigen specific antibodies are produced when LPS molecules are liberated from damaged bacteria at a low concentration, which can provide immunity to Shigella infection; however, they are short-lived and defeated with serotype conversion. The high concentration of O-antigen found on the surface of the bacteria leads to non-specific polyclonal B cell activation and the production of antibodies that do not confer protection. Thus, the acquisition of the SHI-O pathogenicity island may confer upon Shigella the ability to evade host immunity in more ways than one through LPS modification. To prevent an antibody response of any kind to its LPS, Shigella also induces apoptotic B-cell death, mediated by IpaD.Citation189

Persistence

Shigella typically causes a short-term infection; however, long-term carriage is possible.Citation193 During long-term infection, Shigella acquires drug resistance and undergoes large chromosomal structural variations and rearrangements mediated by insertion sequences.Citation193 Outbreaks of Shigella sonnei and Shigella flexneri with extensive multi-drug resistance including the ability to withstand fluoroquinolones and macrolides are on the rise, which curtails treatment options for severe infections.Citation194,Citation195

Role of allelic variants in infection dynamics

OspZ reinforces the emerging theme in infectious disease of different alleles of the same gene fulfilling different roles in virulence and, in some cases, remarkably having opposed effects on the same process. A naturally occurring truncated OspZ allele found in some S. flexneri strains plays a role in intensifying the inflammatory response. In other strains, the full-length OspZ protein attenuates inflammation similar to its EPEC NleE homolog.Citation60

Campylobacter

Campylobacter enteritis is primarily caused by Campylobacter jejuni but also Campylobacter coli and occasionally by other species such as Campylobacter lari, Campylobacter upsaliensis and Campylobacter fetus. Campylobacter jejuni is a commensal bacterium in many animals that is especially common in poultry flocks that can cause diarrheal disease in humans when a potentially low dose of as little as 800 CFU is consumed with contaminated food or water. Diarrheal disease caused by this organism is a world-wide major public health issue. In 2017, C. jejuni surpassed Salmonella as the leading cause of food-borne bacterial diarrheal disease in the United States as it is in many European countries.Citation196 Campylobacter activates and evades host immune responses in ways that directly benefit the pathogen and either directly or indirectly harm the host. Interestingly, numerous postinfectious disorders are associated with resolved campylobacteriosis including the development of Guillain-Barré syndrome, reactive arthritis, and irritable bowel syndrome. This may be attributable to the dysregulation or misdirection of the host immune response.Citation197 This section emphasizes how Campylobacter overcomes innate barriers to infection and evades the adaptive immune system. For a more detailed discussion of Campylobacter pathogenesis and the host response, the reader is directed to additional reviews.Citation197–200

Animal models of disease

Small animal models of C. jejuni enteritis and bloody diarrhea have been somewhat difficult to establish as wild-type mice are colonized by the pathogen inefficiently and inconsistently. Ferret and chick models have been used, but they require genetic manipulation of the bacteria to mimic the enteritis seen in humans, limiting their utility to understanding aspects of disease beyond colonization.Citation201–204 Infection of severe combined immunodeficient mice with a defined limited microflora or the introduction of microbes into antibiotic treated mice fed a zinc or protein deficient diet allows intestinal and systemic inflammation and has been used as models of human disease.Citation205,Citation206

Overcoming colonization resistance

Following ingestion, C. jejuni colonizes the mucous layers of the epithelium of the lower intestinal tract. The intestinal microflora contributes metabolites that stimulate the expression of C. jejuni colonization factors and promote growth.Citation207 Some Campylobacter strains enhance colonization by evading innate defenses with a type 6 contact dependent secretion system that delivers effector proteins into both prokaryotic and eukaryotic cells.Citation208 These effectors mediate resistance to oxidative stress and provide a competitive advantage over the microflora and were also found to be important for in vivo survival.Citation208

Overcoming physical barriers

After penetrating the mucus and reaching the apical face of the epithelium, C. jejuni can pass either between or through individual intestinal epithelial cells to reach the basolateral side. Whether the bacteria transmigrate through the gastrointestinal epithelium exclusively paracellularly or transcellularly or both is unclear.183 C. jejuni secretes a serine protease, HtrA, which cleaves E-cadherin, occludin and claudin-8 to destabilize tight junctions, which enables the bacteria to transmigrate using a paracellular mechanism.Citation209–211 Interestingly, HtrA proteins of other enteric pathogens including Shigella, Helicobacter and EPEC also cleave E-cadherin.Citation212 Following transmigration, the bacteria adhere to the fibronectin-integrin complex that connects the epithelium with the underlying tissue with the adhesin CadF/FlpA and then invades the basal face of the epithelium.Citation209

There is also evidence that Campylobacter invades the apical face of the epithelium and traffics within a Campylobacter containing vacuole to reach the basolateral side of the colonocytes and then exocytoses to the underlying tissue through a mechanism that is poorly understood. In one study, C. jejuni was found to invade epithelial cells and translocate through them following the sialylation of C. jejuni lipooligosaccharide structures, which generates human nerve ganglioside mimics.Citation213 This may be of clinical relevance as ganglioside-like lipooligosaccharide-expressing isolates often cause severe gastroenteritis.Citation213

A variety of studies provided conflicting results on the ability of Campylobacter to transmigrate through an epithelial monolayer with some reporting transcellular and other paracellular routes while others observed both.Citation214–221 Most of these studies observed a time delayed drop in transepithelial electrical resistance.Citation214–220 The most significant delay was observed in a three-dimensional in vitro cell culture assay that is likely more physiologically relevant than the more conventional two-dimensional ones.Citation221 The differences among the various studies are likely attributable to different time points being used as well as different strains being employed. One study used 10 fresh Campylobacter isolates and found 10-fold differences among them in their ability to transmigrate through an epithelium.Citation215 Some isolates favored transcellular penetration of the in vitro monolayer, while others primarily passed through it paracellularly while others used a combination.Citation215

Following transit through the epithelium via paracellular and/or transcellular routes, several Campylobacter species, including C. jejuni, C. coli, C. fetus, and C. upsaliensis, can cause sepsis in both the immunocompromised and the otherwise healthy.Citation222 In addition to the lamina propria, Campylobacter has been recovered from the mesenteric lymph nodes, blood and hepatosplenic tissue of infected animals.Citation223–227 C. jejuni presumably benefits from colonizing the submucosa because it prevents their premature expulsion from the gastrointestinal tract by peristaltic forces and also may provide the bacteria with access to more nutrients. The pathogen can also contact basal receptors such as fibronectin and may be more protected from antibiotics as compared to the gut lumen. Also, by traversing the epithelium, Campylobacter can induce inflammatory diarrhea that enhances transmission to new hosts.

It is interesting to consider how and why Campylobacter sometimes reaches systemic sites and what the future may hold for the evolution of Campylobacter virulence. In one intriguing study, an unidentified extracellular Campylobacter factor(s) promoted the translocation of non-pathogenic bacteria not only through the gut epithelium but also to the blood, spleen and liver of mice.Citation220 This may have implications for understanding and treating autoimmune disorders associated with Campylobacter infections. The passage of live non-pathogenic bacteria to systemic sites suggests a constitutive host directed antigen sampling pathway stimulated by a virulence factor as previously proposed for Salmonella.Citation37

It is not clear why Campylobacter in some instances enters the bloodstream and colonizes systemic organs. In the case of Salmonella, it diverged from Escherichia coli, its nearest non-pathogenic relative to the acquisition of Salmonella pathogenicity island-1 and became a gastrointestinal pathogen.Citation228 It is hypothesized to have become capable of systemic virulence when it later acquired Salmonella pathogenicity island-2 associated genes.Citation228 Two to five percent of systemic infections with typhoidal serovars of Salmonella become asymptomatic within internal organs, with intermittent shedding, potentially for the lifetime of the host.Citation128,Citation229 Thus, the driving force in its ability to reach the bloodstream and the internal organs that filter the blood may have been to allow a longer-term association with the host. Perhaps while still mostly a short-term gastrointestinal pathogen, some strains of Campylobacter are following a similar evolutionary trajectory to that taken by some serovars of Salmonella.

Activation and suppression of inflammation

Campylobacter exhibits numerous traits that facilitate colonization by directly subverting innate immune defenses that modulate inflammation. It possesses several phase-variable loci that modify cell surface structures, such as the capsular polysaccharides or lipooligosaccharides. Base deletion, insertion and single-nucleotide mutations lead to further inter- and intra-strain diversity of cell surface structures that allow subdivision into serotype. Collectively, this variation in enzymes and transferases changes resistance and immunogenicity and helps the bacteria evade the host’s immune system.Citation230

C. jejuni eludes flagellum dependent TLR5 recognition and ensuant inflammation by carrying mutations in the eight amino acid epitope common to many bacterial pathogens’ flagellin. Interestingly, introducing this variant into Salmonella destabilizes the filament and eliminates motility. Campylobacter overcomes this by stabilizing the filament with new contacts not seen in other flagellar filaments.Citation231

Intracellular C. jejuni can be detected with NOD1 and NOD2, which activate antibacterial defenses that are effective in limiting colonization by the microbe. Intriguingly, C. jejuni transitions from helical to coccoid peptidoglycan under times of stress, such as when they are within host cells that result in less NOD1 and NOD2-directed cell activation and inflammatory signaling.Citation232 Campylobacter lipoproteins and LPS, however, are sensed by other Toll-like receptors, which activate NF-κB, resulting in the production and secretion of numerous pro-inflammatory cytokines including IL-8, which recruits neutrophils to the site of infection.Citation62 The involvement of bactericidal neutrophils and other polymorphonuclear leukocytes can damage the intestinal epithelium and produce characteristic bloody diarrhea, which promotes bacterial dissemination to new hosts.

Campylobacter lacks numerous genes common to other enteric bacteria, notably including certain toxins and a canonical type III secretion system. The Cia proteins (Campylobacter invasion antigen) are introduced into the cytosol of epithelial cells with the flagellum export apparatus, which is ancestrally related to type III secretion systems.Citation233 The secretion of these proteins is host cell contact dependent. Roles in promoting adhesion, invasion and intracellular survival have been attributed to them and thus they are essential for colonization.Citation234

CiaC enhances invasion by mediating host cell cytoskeletal rearrangements that contribute to membrane ruffling necessary for internalization.Citation233 Campylobacter remains confined within a Campylobacter-containing vesicle inside of the epithelium that it prevents from fusing with lysosomes.Citation62 Once internalized into epithelial cells, CiaD activates the MAPKs to enhance the inflammatory response that includes additional IL-8 secretion from epithelial cells, which potently recruits neutrophils.Citation62 At least one of the ∼18 secreted Cia proteins attenuates cytokine production rather than enhancing it C. jejuni in fact activates a PI3K/Akt-dependent pathway in human intestinal epithelial cells that is anti-inflammatory.Citation104 By comparison with other enteropathogens, there is likely much more to be discovered in this area.

Some C. jejuni strains produce a genotoxin inside of host cells termed cytolethal distending toxin (CDT). CDT exhibits features of a type I deoxyribonuclease that is an important virulence factor. It arrests the cell cycle in G2/M transition phase. Cell division is inhibited but the cytoplasm continues to grow and distend, resulting in cells up to five times their normal size that remain viable for extended periods of time, before ultimately undergoing cell death.Citation235 CDT was more recently proposed to be a unique and particularly potent virulence factor that acts as a tri-perditious toxin that impairs host defenses by disrupting epithelial barriers, suppressing acquired immunity and by promoting pro-inflammatory responses.Citation63,Citation64 Additional mechanisms may damage host DNA as some C. jejuni strains that lack CDT can still damage DNA and cause disease. A fascinating example is the recent discovery of C. jejuni-derived outer membrane vesicles that contain guide-free Cas9, which transcriptionally reprograms epithelial cells by activating DNA damage and NF-κB signaling and cell death pathways.Citation65,Citation66

Persistence

Antimicrobial resistance represents a persistence strategy utilized by many enteric pathogens including Campylobacter. The prevalence of antimicrobial resistance in Campylobacter is increasing in humans and livestock. Like other bacteria, Campylobacter has developed various resistance mechanisms including ones that confer resistance to several classes of anti-microbial agents, which are of particular concern. One example of many is a mutant allele of cfr that methylates 23s rRNA in a manner that provides the bacteria with resistance to four different classes of anti-microbials.Citation236 This gene is plasmid-borne and thus readily transferable through horizontal gene transfer. Other genetic factors of particular concern are the multiple drug resistance genomic islands of Campylobacter that are transferable between Campylobacter organisms via natural transformation and multi-drug efflux systems.Citation236

Host nutrition plays an important role in determining if Campylobacter infection is acute or chronic. Campylobacter infection is more persistent and tends to be endemic in developing regions, whereas in contrast, in more developed areas, it more often presents as an acute, inflammatory illness.Citation197 In the latter, the diet consists of a relatively higher proportion of processed foods that are mostly devoid of fiber. The dietary differences may result in people in less developed areas experiencing less intense inflammation because of a relative increase in colonic microbial fermentation due a fiber rich diet that results in greater short-chain fatty acid production which is anti-inflammatory.Citation237 Additionally, postinfectious disorders associated with excessive inflammation occur more frequently in the developed world.Citation197

Role of allelic variants in infection dynamics

Campylobacter allelic variants that determine the degree of enteritis and the potential for extraintestinal dissemination have been identified with a variety of approaches.Citation238,Citation239 In one such study, amino acid substitutions within the porA gene were identified that are associated with extraintestinal disease. Lys189 is conserved among extraintestinal strains, and Asn189 is found in porA alleles associated within intestinal disease.Citation238 This single amino acid change may alter the topology of the membrane domain in a way that allows the strains harboring this allele to evade recognition by the immune system. It may also facilitate ion transport to enhance bacterial metabolism.Citation238 The discrepancies regarding the role of porA in causing disease may be related to the specific alleles that were present in the strains being studied.Citation238

Yersinia

In the genus Yersinia, three species are well-described human pathogens. Yersinia enterocolitica is ingested with contaminated food and can cause acute enteritis, enterocolitis, mesenteric lymphadenitis, and terminal ileitis and occasionally sepsis. It is the third most common foodborne zoonotic disease reported in European countries.Citation240 Yersinia pseudotuberculosis is also inadvertently ingested with contaminated food and usually causes self-limiting mesenteric adenitis, ileitis and diarrhea but can also invade the bloodstream. Yersinia pestis, on the other hand, is a zoonosis transmitted to humans by fleas that causes the fatal systemic disease plague. Yersinia is primarily extracellular pathogens that mostly reside within the lymphoid tissue. Pathogenic species resist phagocytosis by professional phagocytes with a type III secretion system and suite of immunomodulatory type III effectors referred to as Yops encoded on a virulence plasmid. This section provides an overview of how Yersinia manipulates inflammation and a discussion of literature pertaining to the mechanisms underlying its persistence within hosts as well as its modes of extraintestinal dissemination. For a more detailed discussion of Yersinia pathogenesis and the immunomodulatory Yops, other reviews are recommended.Citation241–243

Animal models of disease

Y. enterocolitica and Y. pseudotuberculosis pathogenesis are typically studied in rodents, including BALB/c, C57Bl6 mouse strains and rabbits.Citation244 Numerous mouse strains are highly susceptible to infection and the ity locus (Nramp) does not confer protection as it does in the case of S. Typhimurium and other pathogens.Citation245 In murine models, the bacteria can cause gastroenteritis, adenolymphitis and septicemia with hepatosplenic tissue colonization. They can also cause a persistent infection and can be used to model Yersinia-associated reactive arthritis.Citation246,Citation247

Overcoming colonization resistance

Yersinia binds to the mucus with the adhesin YadA and thins and degrades it with enzymes.Citation248 Yersinia appears to scavenge iron better than commensal bacteria, giving it an advantage, although it appears to be sensitive to colicins, which are bactericidal proteins that are continuously released by the resident microflora.Citation248 Yersinia enterocolitica has genes for tetrathionate reduction and may use it as an electron acceptor in the inflamed intestine, similar to Salmonella.Citation20 Yersinia can bind to the epithelium with YadA and invasin.Citation249,Citation250 Overall, research in overcoming colonization resistance by Yersinia is lacking compared to other enteropathogens and in need of additional studies.

Overcoming physical barriers

After penetrating the mucus, Y. pseudotuberculosis and Y. enterocolitica bind the ß1 integrins of M cells with the outer membrane protein invasin, which triggers an internalization event.Citation251 Yersinia is transported within a vacuole in the cytosol of epithelial cells to the basal side of the epithelium from which it exits and is conventionally believed to resist phagocytosis by the underlying professional phagocytes. The infection can proceed to the regional lymph nodes and is believed in some cases to drain through the thoracic duct into the bloodstream. However, there is evidence of Yersinia being transported to the mesenteric lymph nodes by CD103+ dendritic cells in a CCR7-dependent manner.Citation252

While enteropathogenic Yersinia reside within lymphatic tissue, the sequential ordered spread of these pathogens through the lymphatic system to the blood has been challenged in a pair of studies.Citation33,Citation252 In one, the authors observed that Y. pseudotuberculosis disseminated to the blood and subsequently internal organs following replication in the intestine of mice independently of the lymphatic system.Citation33 The authors reported that Y. pseudotuberculosis colonized the hepatosplenic tissues of mice genetically devoid of Peyer’s patches identically to congenic control mice, as was also observed with Salmonella, indicating the availability of alternative routes of extraintestinal dissemination for enteropathogens.Citation13,Citation15,Citation17,Citation33,Citation39,Citation40

In another more recent study, Y. enterocolitica was likely transported by reverse transmigrating CX3CR1+ phagocytes associated with the lamina propria to the spleen independently of Peyer’s patches. The authors observed that the association of the bacteria with these host cells was dependent on invasin but independent of YadA.Citation252 This study used a 1-hour time point; however, in the former one by Barnes et al., the rapid colonization of hepatosplenic tissue following oral inoculation was resolved. A temporally distinct delayed translocation (out to 1 week) that did not involve the lymphatic system resulted in successful colonization of the spleen and liver. Whether or not transport from the lamina propria to the spleen and liver by CX3CR1+ phagocytes involves reverse transmigration needs to be determined and a time course is needed.

Barnes et. al. addressed the possibility that blocking one pathway of extraintestinal dissemination may merely result in more bacteria disseminating through other ones. In an experiment in which none of the pathways were blocked, clonal analysis of uniquely tagged Y. pseudotuberculosis strains revealed that bacteria that successfully colonized the hepatosplenic tissue were not derived from the Peyer’s patches or the mesenteric lymph nodes. Rather, they translocated directly from a replicating pool in the lumen of the intestine out to 1 week post-oral infection.Citation33 As is the case with Salmonella, more study is needed to clarify the routes of extraintestinal dissemination exploited by Yersinia.

Activation and suppression of inflammation

Yersinia actively suppresses microbial internalization by microbicidal phagocytes with YopE and YopT. YopE stimulates the hydrolysis of GTP associated with the Rho GTPases Rac1, RhoA, and Cdc42, to render them in their inactive GDP-bound form.Citation67,Citation68 YopT, on the other hand, is a cysteine protease that disrupts RhoA activity by disassociating RhoA from the plasma membrane by cleaving the prenyl group.Citation70–72 By interfering with RhoA GTPase activity, YopE and YopT activate Pyrin.Citation69 Pyrin is an inflammasome NLR highly expressed by phagocytes that is activated when RhoA GTPase activity is modulated, which is a shared tool through which extracellular pathogens prevent their phagocytosis.Citation253

The sensitivity of Yersinia to pyrin-mediated inflammation is evidenced by Yersinia secreting YopM into phagocytes, which blocks pyrin activation and is essential for virulence.Citation106 YopM is a leucine-rich repeat effector protein that is harbored by all strains of Yersinia that are pathogenic to humans. Its loss reduces colonization, increases host inflammation and often allows the host to survive.Citation106 Macrophages infected with a Yersinia yopM deletion strain have an increased level of activation, IL-1β secretion, and host cell death compared to those infected with wild-type strains.Citation106,Citation107 YopM inhibits caspase-1 activation by subverting host kinases that are negative regulators of pyrin, which down regulates the pro-inflammatory cytokines IL-1β/IL-18.Citation105 This allows Yersinia Yops to block phagocytosis with little consequence. Macrophages infected with yopM deletion mutants initiate a strong pyrin inflammatory response, demonstrating the importance of YopM to infection and mice infected with them are more likely to survive than those infected with wild-type Yersinia.Citation106 On the other hand, mice that lack pyrin are susceptible to ΔyopM Yersinia.Citation106 Thus, pyrin inflammasome activation limits Yersinia persistence, but YopM counters it.

YopK/YopQ is a type III effector that partially blocks inflammasome activation indirectly by regulating the translocation of other Yops into host cells.Citation108 YopK prevents hyper translocation of the T3SS pore-forming proteins YopB and YopD, but this does not prevent entry of low basal levels of these proteins, which is associated with inflammasome activation.Citation108,Citation109 Indeed, the magnitude of inflammasome activation appears to be a function of the relative amount of YopB and YopD introduced into the host cell, as increasing levels of YopD in the host cytosol correlate with larger caspase-1 oligomers.Citation109 YopK affects the size of pores introduced into the host cell membrane. YopK interacts with both YopB and the host protein RACK1 but it is not known how YopK regulates translocon pore structure.Citation107 Both canonical and noncanonical inflammasomes are activated when cells are infected with a yopK mutant.Citation107 Enteropathogenic Yersinia yopK mutants can colonize the Peyer patches, but mice infected with these mutants mount an unusually early inflammatory response that prevents microbial dissemination to the spleen and liver.Citation254 Studies of YopK establish the importance of bacterial virulence factor regulation in dampening type III secretion system-induced innate immune signaling. However, while all organisms that harbor a type III secretion system have YopB and YopD orthologs, no YopK orthologs are known to exist outside Yersinia.

Modulation of host cell death

The type and mechanisms of host cell death induced by Yersinia infection are unclear with reports of necrosis, pyroptosis and apoptosis. YopJ/YopP inactivates the MAPKs with its acetyltransferase activity, which effectively blocks NF-κB and MAPK signaling. Importantly, when NF-κB/MAPK signaling is uninhibited, it allows the expression of proinflammatory and pro-survival genes, enforcing cell survival during infection. In fact, YopJ is indispensable for Yersinia-induced rapid killing of macrophages.Citation255,Citation256 The targeting of immune cells by the type III secretion system might be expected to render these cells ineffectual in responding to Yersinia infection. However, this is not the case; instead, YopJ interdiction of MAPK and NF-κB signaling trips a host pattern-of-pathogenesis alarm that triggers a unique cell death response in infected macrophages. This type of cell death was proposed by one group to involve PANoptosis.Citation257,Citation258 PANoptosis is an inflammatory programmed cell death pathway that shares features of necrosis, pyroptosis and apoptosis and features crosstalk between the respective signaling complexes.Citation257,Citation258 More work will be required to clarify the type of cell death induced by Yersinia and to delineate the underlying mechanism.

Role of allelic variation in infection dynamics

An emerging theme in infectious disease is for pathogens to possess more than one allele of the same gene that in some instances differ by only a single nucleotide polymorphism that changes infection dynamics. Despite the conservation of yopJ in the human-pathogenic Yersinia lineage, the role of YopJ-induced cell death in virulence is not clear with some studies reporting that it contributes to virulence with others finding that it attenuates it. A recent report found that a conservative substitution at codon 177 that encodes a leucine in Y. enterocolitica, codes for a phenylalanine in nearly all of the more pathogenic Y. pseudotuberculosis and Y. pestis strains. The change, despite being conservative, renders YopJ less cytotoxic to macrophages, yet more virulent.Citation259 Yersinia pseudotuberculosis engineered to express YopJ Leu177 in a mildly attenuated ksgA mutant background was completely avirulent in mice.Citation259 YopJ is one of the numerous genes in enteropathogens for which different naturally occurring alleles have recently been discovered that dictate how the pathogens harboring them interact with the immune system.

Persistent infections

Low levels of Yersinia species are frequently found in the human ileum that are not often associated with disease. Perhaps surprisingly, 7.6% of people analyzed by a highly sensitive method were asymptomatic Yersinia carriers.Citation260 This was a small study and a larger scale one to assess what percentage of the population asymptomatically carries various gut pathogens would be interesting. Such infections are a public health concern because they may lead to autoimmune disorders and can serve as reservoirs in which the pathogens can evolve and disseminate to more susceptible hosts.Citation261 Y. pseudotuberculosis can cause persistent infection in mice, where it remains associated with the lymphoid follicles of the cecum.Citation247 In this model, infecting with a low dose of bacteria leads to asymptomatic infection in 20–30% of mice with carriage for as long as 115 days. Even if no signs of disease are present, Y. pseudotuberculosis persistence is associated with an immune response in which the bacterial foci are surrounded by polymorphonuclear neutrophils and this is accompanied by prolonged fecal shedding.Citation247

Yersinia, Salmonella, and Campylobacter have all been reported to infect and affect the ileocecal area in humans, suggesting that the cecum is a beneficial niche for bacterial persistence.Citation247 The mechanisms enabling Y. pseudotuberculosis to persist in cecal tissue in the presence of immune cells for a prolonged time are largely unknown. YopH and YopE contribute to Y. pseudotuberculosis persistence in the cecum, likely by enabling initial colonization in the presence of phagocytic cells. Later, the bacteria become persistent with a novel expression profile, suggesting substantial transcriptional reprogramming. Genes on the virulence plasmid are highly up-regulated initially that in the chronic stage are down-regulated, perhaps to allow persistence in the face of an ongoing immune response. In association with the latter stage, genes involved in anaerobic growth, motility, protection against acidic and oxidative stress as well as genes indicating envelope perturbation are upregulated, suggesting adaptation to the harsh environment in cecal tissue.Citation262 Along with considerable transcriptional reprogramming, the copy number of the Y. pseudotuberculosis virulence plasmid changes during infection. An early increase in copy number is essential for virulence. A later attenuation in gene dosage may contribute to persistence by reducing the chance that the immune system responds to the Yops and is able to ultimately clear the infection.Citation263,Citation264

Persistent Yersinia infection mimics and some have proposed plays a role in Crohn’s disease. Both yersiniosis and Crohn’s disease are characterized by persistent reactivity against the gut microbiota.Citation265 Like yersiniosis, the inflammation characteristic of Crohn’s disease is initially focused around the follicle-associated epithelium and then may progress into deeper, more diffuse ulcerations. Lymphatic vessels have been proposed to play key roles in both the dissemination of yersiniosis and Crohn’s disease.Citation266

Vibrio

Vibrio cholerae is one of the most consequential pathogens in human history. It is responsible for many pandemics of severe diarrheal disease that has produced case fatality rates of 50% or more, with death often occurring within hours of the onset of symptoms.Citation267 Vibrios are natural pathogens of both human and aquatic animals that are commonly found in both freshwater and marine environments.Citation268,Citation269 Of the greater than 100 species, about 12 are pathogenic to humans. The most common Vibrio species that cause disease in humans are Vibrio cholerae, Vibrio parahaemolyticus, Vibrio vulnificus and Vibrio alginolyticus. V. cholerae is typically contracted through the consumption of contaminated seafood or water and is the etiological agent of the severe diarrheal disease known simply as cholera. Other Vibrio species, such as V. parahaemolyticus and V. vulnificus, are responsible for a group of infections termed Vibriosis that have different clinical manifestations. The specific disease caused varies by species, route of infection and host susceptibility and includes nothing more than gastroenteritis but they can also cause primary septicemia. Additionally, they can colonize skin wounds, which may lead to secondary septicemia.

Cholera remains an important public health challenge, despite the development of effective vaccines, with an estimated 5 million cases globally on an annual basis with approximately 100,000 fatalities and has been endemic in Asia for centuries, which continues to this day.Citation270 The public health threat may be heightened in coming years by increasing temperatures associated with climate change as Vibrio thrives in warm water.Citation269 The numerous Vibrio species are diverse but share several notable features including, among others, genomes that are composed of two circular chromosomes that have been extensively shaped by horizontal evolution.

The V. cholerae species is subdivided into over 200 serogroups with two prominent ones, O1 and O139, responsible for epidemic cholera. The O1 serogroup is composed of two biotypes, classical and El Tor. The classical and El Tor biotypes differ in many phenotypic traits. The classical biotype caused the first six cholera pandemics, while the El Tor biotype is responsible for the ongoing seventh pandemic.Citation268 In recent years, there has been an increase in the number of infections with other serotypes of V. cholerae with most causing mild gastroenteritis, ear and wound infections or bacteremia in the immunocompromised; however, extraintestinal infections including fatal bacteremia in otherwise healthy patients do occur.Citation271 This section discusses how Vibrios overcome colonization resistance and their intriguing interactions with the resident gut microflora. It also describes the effects of Vibrios on the inflammatory response. For comprehensive reviews of Vibrio pathogenesis, additional reading is recommended.Citation268,Citation272–274

Animal models of disease

The most frequently used model of V. cholerae pathogenesis is the infant mouse model. Typically, following oral inoculation, within 16 hours, a diarrheal illness manifests, similar to human infection that requires cholera toxin.Citation275 The scarcity of biological materials present in this model for transcriptomic studies and genome-wide functional screens often results in the utilization of an infant rabbit model.Citation275 Drosophila melanogaster, Caenorhabditis elegans and Danio rerio models have also been developed.Citation276–278

Overcoming colonization resistance

V. cholerae is the best understood of the Vibrio species that are pathogenic to humans. It is a non-invasive pathogen whose preferred niche is the mucosal surface of the intestine. The host microbiome has long been suspected to play a role in susceptibility to V. cholerae infection as the household contacts of cholera patients are known to be at elevated risk of infection. Moreover, in animal models of infection, the normal microbiota must be disrupted with antibiotics to render them susceptible to V. cholerae.

V. cholerae employs multiple strategies for overcoming the formidable innate defense of endogenous microflora that results in a multi-log reduction of resident microbes during infection. First, the pathogen has a faster growth rate than members of the commensal microflora giving it a competitive advantage. Secondly, V. cholera deploys a type VI secretion system in response to environmental cues found in the intestine that kills adjacent bacteria but shields the pathogen from its own bactericidal effectors.Citation279 A third feature of V. cholerae pathogenesis that dramatically reduces the density of members of the resident microflora is the profuse watery diarrhea that is induced and the destruction of the mucus layer. A fourth feature of V. cholerae infection that benefits the pathogen at the expense of colonic microbes is the activation of the innate immune system that is ineffective in responding to the initial infection to which the microflora is sensitive. This involves NF-κB and MAPK activation as well as Toll-like receptor signaling and the release of bactericidal proteins. The ensuant immune response includes the generation of reactive oxygen species such as dual oxidase 2 and nitric oxide synthase, which V. cholerae counters with specific inducible resistance mechanisms that the microflora lacks.Citation280

The characteristic profuse, watery diarrhea associated with V. Cholerae infection is primarily attributable to its primary virulence factor, cholera toxin. The cholera toxin B subunit binds to a carbohydrate on the surface of host cells that triggers the internalization of the A subunit, which mediates cAMP-mediated activation of anion secretion and fluid loss. V. cholerae secretes various other enzymes and toxins that promote its persistence within the gastrointestinal tract. This includes Hap, a secreted Zn2+-dependent metalloproteinase that has mucinolytic activity that promotes mucin penetration and the spreading of infection along the gastrointestinal tract.Citation281 The toxin zonula occludens disrupts epithelial integrity by interacting with tight junction proteins.Citation282

Cell curvature is another important feature of V. cholera pathogenesis that aids in persistence. A periskeletal element, CvrA, inserts more peptidoglycan in the outer face of the bacteria than the inner face. This asymmetrical patterning of peptidoglycan produces the curvature, which is characteristic of all V. cholera isolates studied to date and enhances host colonization.Citation283

Vibrio tightly regulates and coordinates the expression of its arsenal of virulence factors with the transcriptional regulator ToxT. ToxT is regulated by environmental signals and up-regulates cholera toxin, the toxin co-regulated pilus (Tcp) and a type IV pilus, which are essential for persistence. The TCP facilitates intestinal colonization by promoting microcolony formation by mediating interaction between adjacent bacteria through direct pilus to pilus contacts, tethering them together as well as to enterocytes and is essential.Citation284

Several recent studies utilized metagenomics and gut microbiome transplantation into gnotobiotic mice to reveal intriguing aspects of the complex interactions that occur between V. cholerae and the host gut microbiome. Some species that can be found in the commensal microflora of humans promote the virulence of this pathogen while others suppress it. In two such studies, interpersonal differences in the human gut microbiota were shown to predict susceptibility to V. cholerae infection, with depleted levels of Bacteroidetes microbes rendering one vulnerable and species within the genera Prevotella and Bifidobacterium being protective.Citation285,Citation286 Blautia obeumcan, a commensal species in the gastrointestinal tract of healthy humans, was shown in another to suppress V. cholerae colonization by degrading taurocholate, a host-produced virulence-inducing molecule.Citation287 Ruminococcus obeum was demonstrated to restrict the expansion of V. cholerae in the gut of mice by producing a signaling molecule that suppresses several of these pathogen’s colonization factors that are regulated by quorum sensing.Citation288 Intriguingly, Paracoccus aminovorans forms a multi-species biofilm with V. cholerae that facilitates its colonization of the gut.Citation289 For more information on how Vibrio interacts with the gut resident microflora, readers are referred to other reviews.Citation267,Citation290

Activation and suppression of inflammation

V. cholerae directly engages the innate immune defenses of the colonic epithelium, both activating it and suppressing it to its advantage. The pathogen disrupts its integrity with cholera toxin and further activates the inflammatory response to its own advantage with the pore-forming toxin cytolysin, encoded by the hlyA gene. The cytolysin is a ß-barrel pore forming toxin that induces mitochondria-dependent apoptosis.Citation73 It results in inflammation in part by stimulating the NF-κB and MAPK pathways.Citation74 Vibrio cholerae secretes Hap, a Zn2+-dependent protease that has mucinolytic activity to neutralize the intestinal mucus layer. It produces another toxin, zonula occludens (ZOT), to disrupt epithelial integrity by manipulating the tight junction proteins occludin and zonula occludens 1 protein (ZO1).Citation281,Citation291 V. cholerae blunts the inflammatory response of the epithelium provoked by cytolysin with virulence factors that increase levels of microRNA miR-146a, which has immunomodulatory effects, that are delivered to host cells with outer membrane vesicles.Citation75 By releasing these outer membrane vesicles with cargo that both activate and dampen the inflammatory response of the epithelium, the pathogen can evade innate immune defenses allowing severe disease and persistence. An intriguing observation was that V. Cholera appears to combat neutrophil extracellular traps with DNases.Citation110

Non-O1/non-O139 serogroup strains

Although they have similar clinical outcomes, pathogenic non-O1/non-O139 serogroup strains achieve them with different virulence factors than epidemic associated O1 and O139 serogroup ones. The former possesses a more diverse collection of virulence factors than the latter, which are mostly clonal.Citation292 Most clinical isolates belonging to non-O1/non-O139 strains lack cholera toxin and the toxin co-regulated pilus but achieve the same effect through less characterized mechanisms. Some of them encode a type III secretion system that helps them colonize the epithelium, disrupt its homeostasis and engage innate immune signaling pathways. The type III secretion system of these strains exerts cytotoxic effects on host cells via osmotic lysis following cortical membrane reorganization and disassembly of epithelial junctions, rather than by apoptotic or necrotic mechanisms.Citation292 It is most similar in sequence with the chromosomally encoded Yersinia Ysc type III secretion system and the two systems from these enteric pathogens show a fair degree of synteny. A subset of V. cholerae non-O1/non-O139 serogroup strains which do not cause epidemic disease but do provoke diarrhea encode a second type III secretion system and lack the canonical virulence factors.Citation273 Many V. parahaemolyticus strains express both type III secretion systems, one from each chromosome.Citation293 Animal models of infection reveal that strains that encode both systems use the second one to promote colonization and disease.Citation273,Citation294 A second type III secretion system can also be harbored by V. mimicus and V. anguillarum that are pathogenic to humans.Citation295

Persistence

V. cholera usually produces a short-term infection that is typically treated with rehydration therapy; however, it is an enteropathogen notorious for multi-drug resistance, which is problematic in severe or prolonged cases of illness. A fascinating recent discovery was that CvrA mRNA is regulated by a sRNA that is in turn induced by cell-wall targeting antibiotics so that cell shape is optimized for resistance.Citation296 Most drug resistance, however, is due to the genome plasticity of this pathogen. Vibrio frequently acquires extrachromosomal mobile genetic elements from not just closely but also distantly related species of bacteria that often allow it to resist antimicrobial agents.Citation297 In fact, over time, V. cholera has acquired mobile genetic elements that confer resistance to all commonly used antibiotics.Citation297

Discussion

While considerable progress has been made in understanding infectious disease, many open questions and challenges remain. For example, we need to determine how to modulate the endogenous microflora to manipulate the complex interactions of enteropathogens with it to suppress infection. We also need to enhance our understanding of the spatiotemporal regulatory inputs that are integrated by pathogens to strike the delicate balance between the induction and suppression of inflammation by enteropathogens to optimize infectiousness and persistence. An exciting emerging trend is the differences being discovered among allelic variants of virulence genes that dictate how the enteropathogens harboring them interact with the immune system. We will need to determine how widespread this is and much follow-up on metagenomic and bioinformatic studies will be required to fully understand the allelic differences observed. These experiments may provide us with unique insight into the molecular mechanisms underlying the subversion of the immune system by pathogens and new tools for fighting them. Perhaps predictive models for what genetic changes are likely to cause outbreaks can be developed.

There are dozens of enteropathogenic secreted effectors with anti-inflammatory properties in addition to SpvC, many of which are required for disseminated infections.Citation87,Citation125 The requirement of many of these genes for lethal disease has historically been attributed to them blunting the innate immune system primarily by subverting killing mechanisms; however, enigmatically, many of them are dispensable for both intracellular survival and the inhibition of host cell death.Citation84,Citation87,Citation125 A potentially important, although often overlooked, component of a pro-inflammatory response is the production and release of cytokines that potently inhibit the movement of infected cells. The field will need to determine which effectors with anti-inflammatory properties are required for virulence because in addition to or perhaps in some cases instead of attenuating more traditional host anti-microbial mechanisms, they manipulate the ability of infected cells to control their movement. For example, macrophage migration inhibitory factor, the release of which is an integral component of an inflammatory response, was initially described as a soluble factor that potently inhibits the migration of phagocytes.Citation298

Chronic infections and the long-term sequelae associated with them are on the rise and just beginning to be understood. Their incidence is becoming more and more of a public health issue with the alarming increase in multi-drug resistance. Outbreaks of acute disease caused by multi-drug resistant bacterial enteropathogens have become so commonplace that some infections with them are essentially untreatable. A systematic network analysis of Salmonella typhimurium metabolism in the murine model of typhoid fever determined that nearly all S. typhimurium enzymes are non-essential, due to extensive metabolic redundancies.Citation299 Not surprisingly, most microbial targets considered druggable are already interdicted with existing drugs.Citation299 Thus, a final remaining challenge is the need to develop new drugs for treating infections caused by enteropathogens, preferably that the bacteria will be unable to quickly evolve ways to overcome.

Disclosure statement

No potential conflict of interest was reported by the authors.

Additional information

Funding

Research reported in this publication was supported by the National Institute Of Allergy And Infectious Diseases of the National Institutes of Health under Award Number R15AI174177. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health .

References

  • Vighi G, Marcucci F, Sensi L, Di Cara G, Frati F. Allergy and the gastrointestinal system. Clin Exp Immunol. 2008;153(Supplement_1):3–38. doi:10.1111/j.1365-2249.2008.03713.x.
  • Gill SR, Pop M, Deboy RT, Eckburg PB, Turnbaugh PJ, Samuel BS, Gordon JI, Relman DA, Fraser-Liggett CM, Nelson KE, et al. Metagenomic analysis of the human distal gut microbiome. Science. 2006;312(5778):1355–1359. doi:10.1126/science.1124234.
  • Pandey S, Kawai T, Akira S. Microbial sensing by Toll-like receptors and intracellular nucleic acid sensors. Cold Spring Harb Perspect Biol. 2015;7(1):a016246. doi:10.1101/cshperspect.a016246.
  • Brewer SM, Brubaker SW, Monack DM. Host inflammasome defense mechanisms and bacterial pathogen evasion strategies. Curr Opin Immunol. 2019;60:63–70. doi:10.1016/j.coi.2019.05.001.
  • Stanaway JD, Reiner RC, Blacker BF, Goldberg EM, Khalil IA, Troeger CE, Andrews JR, Bhutta ZA, Crump JA, Im J, et al. The global burden of typhoid and paratyphoid fevers: a systematic analysis for the global burden of disease study 2017. Lancet Infect Dis. 2019;19(4):369–381. doi:10.1016/S1473-3099(18)30685-6.
  • Ao TT, Feasey NA, Gordon MA, Keddy KH, Angulo FJ, Crump JA. Global burden of invasive nontyphoidal Salmonella disease. Emerg Infect Dis. 2015;2010:21.
  • Azimi T, Zamirnasta M, Sani MA, Soltan Dallal MM, Nasser A. Molecular mechanisms of salmonella effector proteins: a comprehensive review. Infect Drug Resist. 2020;13:11–26. doi:10.2147/IDR.S230604.
  • Bernal-Bayard J, Ramos-Morales F. Molecular mechanisms used by salmonella to evade the immune system. Curr Issues Mol Biol. 2018;25:133–168. doi:10.21775/cimb.025.133.
  • Jennings E, Thurston TLM, Holden DW. Salmonella SPI-2 type III secretion system effectors: molecular mechanisms and physiological consequences. Cell Host Microbe. 2017;22(2):217–231. doi:10.1016/j.chom.2017.07.009.
  • Lou L, Zhang P, Piao R, Wang Y. Salmonella pathogenicity island 1 (SPI-1) and its complex regulatory network. Front Cell Infect Microbiol. 2019;9:270. doi:10.3389/fcimb.2019.00270.
  • Uzzau S, Brown DJ, Wallis T, Rubino S, Leori G, Bernard S, Casadesús J, PLATT DJ, Olsen JE. Host adapted serotypes of Salmonella enterica. Epidemiol Infect. 2000;125(2):229–255. doi:10.1017/S0950268899004379.
  • Tsolis RM, Kingsley RA, Townsend SM, Ficht TA, Adams LG, Baumler AJ. Of mice, calves, and men. Comparison of the mouse typhoid model with other Salmonella infections. Adv Exp Med Biol. 1999;473:261–274.
  • Barthel M, Hapfelmeier S, Quintanilla-Martinez L, Kremer M, Rohde M, Hogardt M, Pfeffer K, Russmann H, Hardt W-D. Pretreatment of mice with streptomycin provides a salmonella enterica serovar typhimurium colitis model that allows analysis of both pathogen and host. Infect Immun. 2003;71(5):2839–2858. doi:10.1128/IAI.71.5.2839-2858.2003.
  • Brown NF, Vallance BA, Coombes BK, Valdez Y, Coburn BA, Finlay BB, Isberg R. Salmonella pathogenicity island 2 Is expressed prior to penetrating the intestine. PLoS Pathog. 2005;1(3):e32. doi:10.1371/journal.ppat.0010032.
  • Gopinath A, Allen TA, Bridgwater CJ, Young CM, Worley MJ, Cascales E. The Salmonella type III effector SpvC triggers the reverse transmigration of infected cells into the bloodstream. PLoS One. 2019;14(12):e0226126. doi:10.1371/journal.pone.0226126.
  • Thornbrough JM, Worley MJ, Zhou D. A naturally occurring single nucleotide polymorphism in the Salmonella SPI-2 type III effector srfH/sseI controls early extraintestinal dissemination. PLoS One. 2012;7(9):e45245. doi:10.1371/journal.pone.0045245.
  • Worley MJ, Nieman GS, Geddes K, Heffron F. Salmonella typhimurium disseminates within its host by manipulating the motility of infected cells. Proc Natl Acad Sci U S A. 2006;103(47):17915–17920. doi:10.1073/pnas.0604054103.
  • Turner JR. Intestinal mucosal barrier function in health and disease. Nat Rev Immunol. 2009;9(11):799–809. doi:10.1038/nri2653.
  • Louis P, Hold GL, Flint HJ. The gut microbiota, bacterial metabolites and colorectal cancer. Nat Rev Microbiol. 2014;12(10):661–672. doi:10.1038/nrmicro3344.
  • Winter SE, Thiennimitr P, Winter MG, Butler BP, Huseby DL, Crawford RW, Russell JM, Bevins CL, Adams LG, Tsolis RM, et al. Gut inflammation provides a respiratory electron acceptor for Salmonella. Nature. 2010;467(7314):426–429. doi:10.1038/nature09415.
  • Thiennimitr P, Winter SE, Winter MG, Xavier MN, Tolstikov V, Huseby DL, Sterzenbach T, Tsolis RM, Roth JR, Bäumler AJ, et al. Intestinal inflammation allows Salmonella to use ethanolamine to compete with the microbiota. Proc Natl Acad Sci U S A. 2011;108(42):17480–17485. doi:10.1073/pnas.1107857108.
  • Lopez CA, Winter SE, Rivera-Chavez F, Xavier MN, Poon V, Nuccio SP, et al. Phage-mediated acquisition of a type III secreted effector protein boosts growth of salmonella by nitrate respiration. mBio. 2012;3.
  • Sana TG, Flaugnatti N, Lugo KA, Lam LH, Jacobson A, Baylot V, Durand E, Journet L, Cascales E, Monack DM, et al. Salmonella Typhimurium utilizes a T6SS-mediated antibacterial weapon to establish in the host gut. Proc Natl Acad Sci U S A. 2016;113(34):E5044–51. doi:10.1073/pnas.1608858113.
  • Rogers AWL, Tsolis RM, Baumler AJ. Salmonella versus the Microbiome. Microbiol Mol Biol Rev. 2021;85(1).
  • Jones BD, Ghori N, Falkow S. Salmonella typhimurium initiates murine infection by penetrating and destroying the specialized epithelial M cells of the Peyer’s patches. J Exp Med. 1994;180(1):15–23. doi:10.1084/jem.180.1.15.
  • Patel S, McCormick BA. Mucosal inflammatory response to salmonella Typhimurium infection. Front Immunol. 2014;5:311. doi:10.3389/fimmu.2014.00311.
  • McCormick BA, Parkos CA, Colgan SP, Carnes DK, Madara JL. Apical secretion of a pathogen-elicited epithelial chemoattractant activity in response to surface colonization of intestinal epithelia by Salmonella typhimurium. J Immunol. 1998;160:455–466.
  • Mrsny RJ, Gewirtz AT, Siccardi D, Savidge T, Hurley BP, Madara JL, McCormick BA. Identification of hepoxilin A 3 in inflammatory events: a required role in neutrophil migration across intestinal epithelia. Proc Natl Acad Sci U S A. 2004;101(19):7421–7426. doi:10.1073/pnas.0400832101.
  • Wall DM, Nadeau WJ, Pazos MA, Shi HN, Galyov EE, McCormick BA. Identification of the Salmonella enterica serotype Typhimurium SipA domain responsible for inducing neutrophil recruitment across the intestinal epithelium. Cell Microbiol. 2007;9(9):2299–2313. doi:10.1111/j.1462-5822.2007.00960.x.
  • Zhou D, Mooseker MS, Galan JE. Role of the S. typhimurium actin-binding protein sipa in bacterial internalization. Science. 1999;283(5410):2092–2095. doi:10.1126/science.283.5410.2092.
  • Srikanth CV, Wall DM, Maldonado-Contreras A, Shi H, Zhou D, Demma Z, Mumy KL, McCormick BA. Salmonella pathogenesis and processing of secreted effectors by caspase-3. Science. 2010;330(6002):390–393. doi:10.1126/science.1194598.
  • Sun L, Yang S, Deng Q, Dong K, Li Y, Wu S, Huang R. Salmonella effector SpvB disrupts intestinal epithelial barrier integrity for bacterial translocation. Front Cell Infect Microbiol. 2020;10:606541. doi:10.3389/fcimb.2020.606541.
  • Barnes PD, Bergman MA, Mecsas J, Isberg RR. Yersinia pseudotuberculosis disseminates directly from a replicating bacterial pool in the intestine. J Exp Med. 2006;203(6):1591–1601. doi:10.1084/jem.20060905.
  • Voedisch S, Koenecke C, David S, Herbrand H, Forster R, Rhen M, Pabst O. Mesenteric lymph nodes confine dendritic cell-mediated dissemination of Salmonella enterica serovar typhimurium and limit systemic disease in mice. Infect Immun. 2009;77(8):3170–3180. doi:10.1128/IAI.00272-09.
  • Bravo-Blas A, Utriainen L, Clay SL, Kastele V, Cerovic V, Cunningham AF, Henderson IR, Wall DM, Milling SWF. Salmonella enterica serovar typhimurium travels to mesenteric lymph nodes both with host cells and autonomously. J Immunol. 2019;202(1):260–267. doi:10.4049/jimmunol.1701254.
  • Cyster JG, Schwab SR. Sphingosine-1-phosphate and lymphocyte egress from lymphoid organs. Annu Rev Immunol. 2012;30(1):69–94. doi:10.1146/annurev-immunol-020711-075011.
  • Vazquez-Torres A, Jones-Carson J, Baumler AJ, Falkow S, Valdivia R, Brown W, Le M, Berggren R, Parks WT, Fang FC, et al. Extraintestinal dissemination of Salmonella by CD18-expressing phagocytes. Nature. 1999;401(6755):804–808. doi:10.1038/44593.
  • Bianchi G, D’Amico G, Sozzani S, Mantovani A, Allavena P. Transendothelial migration and reverse transmigration of in vitro cultured human dendritic cells. Methods Mol Med. 2001;64:325–330. doi:10.1385/1-59259-150-7:325.
  • Muller AJ, Kaiser P, Dittmar KE, Weber TC, Haueter S, Endt K, Songhet P, Zellweger C, Kremer M, Fehling H-J, et al. Salmonella gut invasion involves TTSS-2-dependent epithelial traversal, basolateral exit, and uptake by epithelium-sampling lamina propria phagocytes. Cell Host Microbe. 2012;11(1):19–32. doi:10.1016/j.chom.2011.11.013.
  • Spadoni I, Zagato E, Bertocchi A, Paolinelli R, Hot E, Di Sabatino A, Caprioli F, Bottiglieri L, Oldani A, Viale G, et al. A gut-vascular barrier controls the systemic dissemination of bacteria. Science. 2015;350(6262):830–834. doi:10.1126/science.aad0135.
  • Silva-Garcia O, Valdez-Alarcon JJ, Baizabal-Aguirre VM. Wnt/beta-catenin signaling as a molecular target by pathogenic bacteria. Front Immunol. 2019;10:2135. doi:10.3389/fimmu.2019.02135.
  • Lambert MA, Smith SG. The PagN protein of Salmonella enterica serovar Typhimurium is an adhesin and invasin. BMC Microbiol. 2008;8(1):142. doi:10.1186/1471-2180-8-142.
  • Mambu J, Virlogeux-Payant I, Holbert S, Grepinet O, Velge P, Wiedemann A. An updated view on the Rck invasin of salmonella: still much to discover. Front Cell Infect Microbiol. 2017;7:500. doi:10.3389/fcimb.2017.00500.
  • Chen D, Burford WB, Pham G, Zhang L, Alto LT, Ertelt JM, Winter MG, Winter SE, Way SS, Alto NM, et al. Systematic reconstruction of an effector-gene network reveals determinants of Salmonella cellular and tissue tropism. Cell Host Microbe. 2021;29(10):1531–1544.e9. doi:10.1016/j.chom.2021.08.012.
  • Fierer J. Extra-intestinal salmonella infections: the significance of spv genes. Clin Infect Dis. 2001;32(3):519–520. doi:10.1086/318505.
  • Zuo L, Zhou L, Wu C, Wang Y, Li Y, Huang R, Wu S. Salmonella spvC gene inhibits pyroptosis and intestinal inflammation to aggravate systemic infection in mice. Front Microbiol. 2020;11:562491. doi:10.3389/fmicb.2020.562491.
  • Mazurkiewicz P, Thomas J, Thompson JA, Liu M, Arbibe L, Sansonetti P, Holden DW. SpvC is a Salmonella effector with phosphothreonine lyase activity on host mitogen-activated protein kinases. Mol Microbiol. 2008;67(6):1371–1383. doi:10.1111/j.1365-2958.2008.06134.x.
  • Montenegro MA, Morelli G, Helmuth R. Heteroduplex analysis of salmonella virulence plasmids and their prevalence in isolates of defined sources. Microb Pathog. 1991;11(6):391–397. doi:10.1016/0882-4010(91)90035-9.
  • Fierer J, Krause M, Tauxe R, Guiney D. Salmonella typhimurium bacteremia: association with the virulence plasmid. J Infect Dis. 1992;166(3):639–642. doi:10.1093/infdis/166.3.639.
  • Patel JC, Galan JE. Differential activation and function of Rho GTPases during Salmonella-host cell interactions. J Cell Biol. 2006;175(3):453–463. doi:10.1083/jcb.200605144.
  • Norris FA, Wilson MP, Wallis TS, Galyov EE, Majerus PW. SopB, a protein required for virulence of salmonella Dublin, is an inositol phosphate phosphatase. Proc Natl Acad Sci U S A. 1998;95(24):14057–14059. doi:10.1073/pnas.95.24.14057.
  • Friebel A, Ilchmann H, Aepfelbacher M, Ehrbar K, Machleidt W, Hardt W-D. SopE and SopE2 from salmonella typhimurium activate different sets of RhoGTPases of the Host Cell. J Biol Chem. 2001;276(36):34035–34040. doi:10.1074/jbc.M100609200.
  • Sun H, Kamanova J, Lara-Tejero M, Galán JE. Salmonella stimulates pro-inflammatory signalling through p21-activated kinases bypassing innate immune receptors. Nature Microbiol. 2018;3(10):1122–1130. doi:10.1038/s41564-018-0246-z.
  • Hardt W-D, Chen L-M, Schuebel KE, Bustelo XR, Galán JE. S. typhimurium encodes an activator of Rho GTPases that induces membrane ruffling and nuclear responses in host cells. Cell. 1998;93(5):815–826. doi:10.1016/s0092-8674(00)81442-7.
  • Kamanova J, Sun H, Lara-Tejero M, Galán JE, Baumler AJ. The salmonella effector protein SopA modulates innate immune responses by targeting TRIM E3 ligase family members. PLoS Pathog. 2016;12(4):e1005552. doi:10.1371/journal.ppat.1005552.
  • Lian H, Jiang K, Tong M, Chen Z, Liu X, Galan JE, Gao X. The Salmonella effector protein SopD targets Rab8 to positively and negatively modulate the inflammatory response. Nat Microbiol. 2021;6(5):658–671. doi:10.1038/s41564-021-00866-3.
  • Parsons JT, Horwitz AR, Schwartz MA. Cell adhesion: integrating cytoskeletal dynamics and cellular tension. Nat Rev Mol Cell Biol. 2010;11(9):633–643. doi:10.1038/nrm2957.
  • Six DA, Dennis EA. The expanding superfamily of phospholipase A(2) enzymes: classification and characterization. Biochim Biophys Acta. 2000;1488(1–2):1–19. doi:10.1016/S1388-1981(00)00105-0.
  • Ambrosi C, Pompili M, Scribano D, Limongi D, Petrucca A, Cannavacciuolo S, Schippa S, Zagaglia C, Grossi M, Nicoletti M, et al. The shigella flexneri OspB effector: an early immunomodulator. Int J Med Microbiol. 2015;305(1):75–84. doi:10.1016/j.ijmm.2014.11.004.
  • Newton HJ, Pearson JS, Badea L, Kelly M, Lucas M, Holloway G, Wagstaff KM, Dunstone MA, Sloan J, Whisstock JC, et al. The type III effectors NleE and NleB from Enteropathogenic E. coli and OspZ from shigella block nuclear translocation of NF-κB p65. PLoS Pathog. 2010;6(5):e1000898. doi:10.1371/journal.ppat.1000898.
  • Zurawski DV, Mitsuhata C, Mumy KL, McCormick BA, Maurelli AT. OspF and OspC1 Are Shigella flexneri Type III secretion system effectors that are required for postinvasion aspects of virulence. Infect Immun. 2006;74(10):5964–5976. doi:10.1128/IAI.00594-06.
  • Samuelson DR, Eucker TP, Bell JA, Dybas L, Mansfield LS, Konkel ME. The Campylobacter jejuniCiaD effector protein activates MAP kinase signaling pathways and is required for the development of disease. Cell Com Signal. 2013;11(1):79. doi:10.1186/1478-811X-11-79.
  • Scuron MD, Boesze-Battaglia K, Dlakic M, Shenker BJ. The cytolethal distending toxin contributes to microbial virulence and disease pathogenesis by acting as a Tri-perditious toxin. Front Cell Infect Microbiol. 2016;6:168. doi:10.3389/fcimb.2016.00168.
  • Hickey TE, McVeigh AL, Scott DA, Michielutti RE, Bixby A, Carroll SA, Bourgeois AL, Guerry P. Campylobacter jejuni cytolethal distending toxin mediates release of interleukin-8 from intestinal epithelial cells. Infect Immun. 2000;68(12):6535–6541. doi:10.1128/IAI.68.12.6535-6541.2000.
  • Saha C, Horst-Kreft D, Kross I, Van Der Spek PJ, Louwen R, Van Baarlen P. Campylobacter jejuni Cas9 Modulates the Transcriptome in Caco-2 intestinal epithelial cells. Genes. 2020;11(10):1193. doi:10.3390/genes11101193.
  • Saha C, Mohanraju P, Stubbs A, Dugar G, Hoogstrate Y, Kremers GJ, van Cappellen WA, Horst-Kreft D, Laffeber C, Lebbink JHG, et al. Guide-free Cas9 from pathogenic Campylobacter jejuni bacteria causes severe damage to DNA. Sci Adv. 2020;6(25):eaaz4849. doi:10.1126/sciadv.aaz4849.
  • Black DS, Bliska JB. The RhoGAP activity of the Yersinia pseudotuberculosis cytotoxin YopE is required for antiphagocytic function and virulence. Mol Microbiol. 2000;37(3):515–527. doi:10.1046/j.1365-2958.2000.02021.x.
  • Von Pawel-Rammingen U, Telepnev MV, Schmidt G, Aktories K, Wolf-Watz H, Rosqvist R. GAP activity of the Yersinia YopE cytotoxin specifically targets the Rho pathway: a mechanism for disruption of actin microfilament structure. Mol Microbiol. 2000;36(3):737–748. doi:10.1046/j.1365-2958.2000.01898.x.
  • Medici NP, Rashid M, Bliska JB. Characterization of Pyrin dephosphorylation and inflammasome activation in macrophages as triggered by the Yersinia effectors YopE and YopT. Infect Immun. 2019;87(3):1–13.
  • Aepfelbacher M, Trasak C, Wilharm G, Wiedemann A, Trülzsch K, Krauss K, Gierschik P, Heesemann J. Characterization of YopT effects on Rho GTPases in Yersinia enterocolitica-infected Cells. J Biol Chem. 2003;278(35):33217–33223. doi:10.1074/jbc.M303349200.
  • Zumbihl R, Aepfelbacher M, Andor A, Jacobi CA, Ruckdeschel K, Rouot B, Heesemann J. The cytotoxin YopT of Yersinia enterocolitica induces modification and cellular redistribution of the small GTP-binding protein RhoA. J Biol Chem. 1999;274(41):29289–29293. doi:10.1074/jbc.274.41.29289.
  • Shao F, Vacratsis PO, Bao Z, Bowers KE, Fierke CA, Dixon JE. Biochemical characterization of the Yersinia YopT protease: cleavage site and recognition elements in Rho GTPases. Proc Natl Acad Sci U S A. 2003;100(3):904–909. doi:10.1073/pnas.252770599.
  • Rai AK, Chattopadhyay K. Revisiting the membrane interaction mechanism of a membrane-damaging β-barrel pore-forming toxin V ibrio cholerae cytolysin. Mol Microbiol. 2015;97(6):1051–1062. doi:10.1111/mmi.13084.
  • Khilwani B, Mukhopadhaya A, Chattopadhyay K. Transmembrane oligomeric form of Vibrio cholerae cytolysin triggers TLR2/TLR6–dependent proinflammatory responses in monocytes and macrophages. Biochem J. 2015;466(1):147–161. doi:10.1042/BJ20140718.
  • Bitar A, Aung KM, Wai SN, Hammarström M-L. Vibrio cholerae derived outer membrane vesicles modulate the inflammatory response of human intestinal epithelial cells by inducing microRNA-146a. Sci Rep. 2019;9(1). doi:10.1038/s41598-019-43691-9.
  • Hobbie S, Chen LM, Davis RJ, Galan JE. Involvement of mitogen-activated protein kinase pathways in the nuclear responses and cytokine production induced by Salmonella typhimurium in cultured intestinal epithelial cells. J Immunol. 1997;159:5550–5559.
  • Bruno VM, Hannemann S, Lara-Tejero M, Flavell RA, Kleinstein SH, Galán JE, Isberg RR. Salmonella Typhimurium Type III secretion effectors stimulate innate immune responses in cultured epithelial cells. PLoS Pathog. 2009;5(8):e1000538. doi:10.1371/journal.ppat.1000538.
  • Galan JE. Salmonella Typhimurium and inflammation: a pathogen-centric affair. Nat Rev Microbiol. 2021;19(11):716–725. doi:10.1038/s41579-021-00561-4.
  • Kelly D, Conway S, Aminov R. Commensal gut bacteria: mechanisms of immune modulation. Trends Immunol. 2005;26(6):326–333. doi:10.1016/j.it.2005.04.008.
  • Shibolet O, Podolsky DK. TLRs in the Gut. IV. negative regulation of Toll-like receptors and intestinal homeostasis: addition by subtraction. Am J Physiol Gastrointest Liver Physiol. 2007;292:G1469–73. doi:10.1152/ajpgi.00531.2006.
  • Fu Y, Galán JE. A Salmonella protein antagonizes Rac-1 and Cdc42 to mediate host-cell recovery after bacterial invasion. Nature. 1999;401(6750):293–297. doi:10.1038/45829.
  • Gibbs KD, Washington EJ, Jaslow SL, Bourgeois JS, Foster MW, Guo R, Brennan RG, Ko DC. The salmonella secreted effector SarA/SteE Mimics Cytokine Receptor Signaling to Activate STAT3. Cell Host Microbe. 2020;27(1):129–39 e4. doi:10.1016/j.chom.2019.11.012.
  • Panagi I, Jennings E, Zeng J, Gunster RA, Stones CD, Mak H, Jin E, Stapels DAC, Subari NZ, Pham THM, et al. Salmonella effector SteE converts the mammalian serine/threonine kinase GSK3 into a tyrosine kinase to direct macrophage polarization. Cell Host Microbe. 2020;27(1):41–53 e6. doi:10.1016/j.chom.2019.11.002.
  • Sun H, Kamanova J, Lara-Tejero M, Galán JE, Philpott DJ. A family of salmonella type III secretion effector proteins selectively targets the NF-κB signaling pathway to preserve host homeostasis. PLoS Pathog. 2016;12(3):e1005484. doi:10.1371/journal.ppat.1005484.
  • Du F, Galán JE, Stebbins CE. Selective inhibition of type III secretion activated signaling by the salmonella effector AvrA. PLoS Pathog. 2009;5(9):e1000595. doi:10.1371/journal.ppat.1000595.
  • Jones RM, Wu H, Wentworth C, Luo L, Collier-Hyams L, Neish AS. Salmonella AvrA Coordinates Suppression of Host Immune and Apoptotic Defenses via JNK Pathway Blockade. Cell Host Microbe. 2008;3(4):233–244. doi:10.1016/j.chom.2008.02.016.
  • Gunster RA, Matthews SA, Holden DW, Thurston TLM. SseK1 and SseK3 Type III Secretion System Effectors Inhibit NF-kappaB Signaling and Necroptotic Cell Death in Salmonella-Infected Macrophages. Infect Immun. 2017;85(3).
  • Newson JPM, Scott NE, Yeuk Wah Chung I, Wong Fok Lung T, Giogha C, Gan J, Wang N, Strugnell RA, Brown NF, Cygler M, et al. Salmonella effectors SseK1 and SseK3 Target Death Domain Proteins in the TNF and TRAIL Signaling Pathways. Mol Cell Proteomics. 2019;18(6):1138–1156. doi:10.1074/mcp.RA118.001093.
  • Yang S, Deng Q, Sun L, Zhu Y, Dong K, Wu S, Huang R, Li Y. Salmonella Effector SpvB Inhibits NF-kappaB Activity via KEAP1-mediated downregulation of IKKbeta. Front Cell Infect Microbiol. 2021;11:641412. doi:10.3389/fcimb.2021.641412.
  • Li H, Xu H, Zhou Y, Zhang J, Long C, Li S, Chen S, Zhou J-M, Shao F. The phosphothreonine lyase activity of a bacterial type III effector family. Science. 2007;315(5814):1000–1003. doi:10.1126/science.1138960.
  • Rolhion N, Furniss RC, Grabe G, Ryan A, Liu M, Matthews SA, Holden DW. Inhibition of Nuclear Transport of NF-kB p65 by the salmonella type III secretion system effector SpvD. PLoS Pathog. 2016;12(5):e1005653. doi:10.1371/journal.ppat.1005653.
  • Sontag RL, Nakayasu ES, Brown RN, Niemann GS, Sydor MA, Sanchez O, et al. Identification of novel host interactors of effectors secreted by salmonella and citrobacter. mSystems. 2016;1(4).
  • Pilar AVC, Reid-Yu SA, Cooper CA, Mulder DT, Coombes BK, Schneider DS. GogB is an anti-inflammatory effector that limits tissue damage during salmonella infection through interaction with human FBXO22 and Skp1. PLoS Pathog. 2012;8(6):e1002773. doi:10.1371/journal.ppat.1002773.
  • Luchetti G, Roncaioli JL, Chavez RA, Schubert AF, Kofoed EM, Reja R, Cheung TK, Liang Y, Webster JD, Lehoux I, et al. Shigella ubiquitin ligase IpaH7.8 targets gasdermin D for degradation to prevent pyroptosis and enable infection. Cell Host Microbe. 2021;29(10):1521–30.e10. doi:10.1016/j.chom.2021.08.010.
  • De Jong MF, Liu Z, Chen D, Alto NM. Shigella flexneri suppresses NF-κB activation by inhibiting linear ubiquitin chain ligation. Nature Microbiol. 2016;1(7):16084. doi:10.1038/nmicrobiol.2016.84.
  • Nasser A, Mosadegh M, Azimi T, Shariati A. Molecular mechanisms of Shigella effector proteins: a common pathogen among diarrheic pediatric population. Mol Cellular Pediatrics. 2022;9(1). doi:10.1186/s40348-022-00145-z.
  • Ji C, Du S, Li P, Zhu Q, Yang X, Long C, Yu J, Shao F, Xiao J, et al. Structural mechanism for guanylate-binding proteins (GBPs) targeting by the Shigella E3 ligase IpaH9.8. PLoS Pathog. 2019;15(6):e1007876. doi:10.1371/journal.ppat.1007876.
  • Ashida H, Kim M, Schmidt-Supprian M, Ma A, Ogawa M, Sasakawa C. A bacterial E3 ubiquitin ligase IpaH9.8 targets NEMO/IKKγ to dampen the host NF-κB-mediated inflammatory response. Nat Cell Biol. 2010;12(1):66–73. doi:10.1038/ncb2006.
  • Zheng Z, Wei C, Guan K, Yuan Y, Zhang Y, Ma S, Cao Y, Wang F, Zhong H, He X, et al. Bacterial E3 Ubiquitin Ligase IpaH4.5 of shigella flexneri targets TBK1 to dampen the host antibacterial response. J Immunol. 2016;196(3):1199–1208. doi:10.4049/jimmunol.1501045.
  • Ashida H, Nakano H, Sasakawa C, Tran Van Nhieu G. Shigella IpaH0722 E3 Ubiquitin Ligase Effector Targets TRAF2 to Inhibit PKC–NF-κB activity in invaded epithelial cells. PLoS Pathog. 2013;9(6):e1003409. doi:10.1371/journal.ppat.1003409.
  • Kim DW, Lenzen G, Page AL, Legrain P, Sansonetti PJ, Parsot C. The Shigella flexneri effector OspG interferes with innate immune responses by targeting ubiquitin-conjugating enzymes. Proc Natl Acad Sci U S A. 2005;102(39):14046–14051. doi:10.1073/pnas.0504466102.
  • Zhou Y, Dong N, Hu L, Shao F, Kwaik YA. The shigella type three secretion system effector OSPG directly and specifically binds to host ubiquitin for activation. PLoS ONE. 2013;8(2):e57558. doi:10.1371/journal.pone.0057558.
  • Sanada T, Kim M, Mimuro H, Suzuki M, Ogawa M, Oyama A, Ashida H, Kobayashi T, Koyama T, Nagai S, et al. The Shigella flexneri effector OspI deamidates UBC13 to dampen the inflammatory response. Nature. 2012;483(7391):623–626. doi:10.1038/nature10894.
  • Li YP, Vegge CS, Brondsted L, Madsen M, Ingmer H, Bang DD. Campylobacter jejuni induces an anti-inflammatory response in human intestinal epithelial cells through activation of phosphatidylinositol 3-kinase/Akt pathway. Vet Microbiol. 2011;148(1):75–83. doi:10.1016/j.vetmic.2010.08.009.
  • Ratner D, Orning MPA, Proulx MK, Wang D, Gavrilin MA, Wewers MD, Alnemri ES, Johnson PF, Lee B, Mecsas J, et al. The Yersinia pestis effector YopM Inhibits Pyrin Inflammasome Activation. PLoS Pathog. 2016;12(12):e1006035. doi:10.1371/journal.ppat.1006035.
  • Chung LK, Park YH, Zheng Y, Brodsky IE, Hearing P, Kastner DL, Chae JJ, Bliska JB. The Yersinia Virulence Factor YopM Hijacks Host Kinases to Inhibit Type III Effector-Triggered Activation of the Pyrin Inflammasome. Cell Host Microbe. 2016;20(3):296–306. doi:10.1016/j.chom.2016.07.018.
  • Brodsky IE, Palm NW, Sadanand S, Ryndak MB, Sutterwala FS, Flavell RA, Bliska JB, Medzhitov R. A Yersinia effector protein promotes virulence by preventing inflammasome recognition of the type III secretion system. Cell Host Microbe. 2010;7(5):376–387. doi:10.1016/j.chom.2010.04.009.
  • Dewoody R, Merritt PM, Marketon MM. YopK controls both rate and fidelity of Yop translocation. Mol Microbiol. 2013;87(2):301–317. doi:10.1111/mmi.12099.
  • Zwack EE, Snyder AG, Wynosky-Dolfi MA, Ruthel G, Philip NH, Marketon MM, Francis MS, Bliska JB, Brodsky IE, et al. Inflammasome activation in response to the Yersinia type III secretion system requires hyperinjection of translocon proteins YopB and YopD. mBio. 2015;6(1):e02095–14. doi:10.1128/mBio.02095-14.
  • Bishop AL, Patimalla B, Camilli A, Payne SM. Vibrio cholerae-Induced Inflammation in the Neonatal Mouse Cholera Model. Infect Immun. 2014;82(6):2434–2447. doi:10.1128/IAI.00054-14.
  • Haneda T, Ishii Y, Shimizu H, Ohshima K, Iida N, Danbara H, Okada N. Salmonella type III effector SpvC, a phosphothreonine lyase, contributes to reduction in inflammatory response during intestinal phase of infection. Cell Microbiol. 2012;14(4):485–499. doi:10.1111/j.1462-5822.2011.01733.x.
  • Lu R, Wu S, Liu X, Xia Y, Zhang Y-G, Sun J, Bereswill S. Chronic Effects of a Salmonella Type III secretion effector protein AvrA In Vivo. PLoS ONE. 2010;5(5):e10505. doi:10.1371/journal.pone.0010505.
  • Grabe GJ, Zhang Y, Przydacz M, Rolhion N, Yang Y, Pruneda JN, et al. The Salmonella Effector SpvD Is a cysteine hydrolase with a serovar-specific polymorphism influencing catalytic activity, suppression of immune responses, and bacterial virulence. J Biol Chem. 2016;291(50):25853–25863. doi:10.1074/jbc.M116.752782.
  • Wall AA, Luo L, Hung Y, Tong SJ, Condon ND, Blumenthal A, Sweet MJ, Stow JL. Small GTPase Rab8a-recruited phosphatidylinositol 3-Kinase γ regulates signaling and cytokine outputs from endosomal toll-like receptors. J Biol Chem. 2017;292(11):4411–4422. doi:10.1074/jbc.M116.766337.
  • Luo L, Wall AA, Tong SJ, Hung Y, Xiao Z, Tarique AA, Sly PD, Fantino E, Marzolo M-P, Stow JL, et al. TLR crosstalk activates LRP1 to recruit Rab8a and PI3Kγ for suppression of inflammatory responses. Cell Rep. 2018;24(11):3033–3044. doi:10.1016/j.celrep.2018.08.028.
  • Tong SJ, Wall AA, Hung Y, Luo L, Stow JL. Guanine nucleotide exchange factors activate Rab8a for Toll-like receptor signalling. Small GTPases. 2021;12(1):27–43. doi:10.1080/21541248.2019.1587278.
  • Leppkes M, Neurath MF, Herrmann M, Becker C. Immune deficiency vs. immune excess in inflammatory bowel diseases- STAT3 as a rheo-STAT of intestinal homeostasis. J Leukoc Biol. 2016;99(1):57–66. doi:10.1189/jlb.5MR0515-221R.
  • Hillmer EJ, Zhang H, Li HS, Watowich SS. STAT3 signaling in immunity. Cytokine Growth Factor Rev. 2016;31:1–15. doi:10.1016/j.cytogfr.2016.05.001.
  • Wemyss MA, Pearson JS. Host Cell Death Responses to Non-typhoidal Salmonella Infection. Front Immunol. 2019;10:1758. doi:10.3389/fimmu.2019.01758.
  • Chen LM, Kaniga K, Galan JE. Salmonella spp. are cytotoxic for cultured macrophages. Mol Microbiol. 1996;21(5):1101–1115. doi:10.1046/j.1365-2958.1996.471410.x.
  • van der Velden AW, Lindgren SW, Worley MJ, Heffron F, O’Brien AD. Salmonella Pathogenicity Island 1-independent induction of apoptosis in infected macrophages by salmonella enterica serotype typhimurium. Infect Immun. 2000;68(10):5702–5709. doi:10.1128/IAI.68.10.5702-5709.2000.
  • Hersh D, Monack DM, Smith MR, Ghori N, Falkow S, Zychlinsky A. The Salmonella invasin SipB induces macrophage apoptosis by binding to caspase-1. Proc Natl Acad Sci U S A. 1999;96(5):2396–2401. doi:10.1073/pnas.96.5.2396.
  • Sundquist M, Wick MJ. Salmonella induces death of CD8 + dendritic cells but not CD11cintCD11b+ inflammatory cells in vivo via MyD88 and TNFR1. J Leukoc Biol. 2009;85(2):225–234. doi:10.1189/jlb.0708413.
  • van der Velden AW, Copass MK, Starnbach MN. Salmonella inhibit T cell proliferation by a direct, contact-dependent immunosuppressive effect. Proc Natl Acad Sci U S A. 2005;102(49):17769–17774. doi:10.1073/pnas.0504382102.
  • Wang M, Qazi IH, Wang L, Zhou G, Han H. Salmonella virulence and immune escape. Microorganisms. 2020;9(1):8. doi:10.3390/microorganisms9010008.
  • Bayer-Santos E, Durkin CH, Rigano LA, Kupz A, Alix E, Cerny O, Jennings E, Liu M, Ryan AS, Lapaque N, et al. The salmonella effector SteD mediates MARCH8-dependent ubiquitination of MHC II molecules and inhibits T Cell activation. Cell Host Microbe. 2016;20(5):584–595. doi:10.1016/j.chom.2016.10.007.
  • Srinivasan A, Nanton M, Griffin A, McSorley SJ. Culling of Activated CD4 T Cells during typhoid is driven by salmonella virulence genes. J Immunol. 2009;182(12):7838–7845. doi:10.4049/jimmunol.0900382.
  • Gunn JS, Marshall JM, Baker S, Dongol S, Charles RC, Ryan ET. Salmonella chronic carriage: epidemiology, diagnosis, and gallbladder persistence. Trends Microbiol. 2014;22(11):648–655. doi:10.1016/j.tim.2014.06.007.
  • Gonzalez-Escobedo G, Marshall JM, Gunn JS. Chronic and acute infection of the gall bladder by Salmonella Typhi: understanding the carrier state. Nat Rev Microbiol. 2011;9(1):9–14. doi:10.1038/nrmicro2490.
  • Stapels DAC, Hill PWS, Westermann AJ, Fisher RA, Thurston TL, Saliba AE, Blommestein I, Vogel J, Helaine S. Salmonella persisters undermine host immune defenses during antibiotic treatment. Science. 2018;362(6419):1156–1160. doi:10.1126/science.aat7148.
  • Pham THM, Brewer SM, Thurston T, Massis LM, Honeycutt J, Lugo K, Jacobson AR, Vilches-Moure JG, Hamblin M, Helaine S, et al. Salmonella-driven polarization of granuloma macrophages antagonizes TNF-mediated pathogen restriction during persistent infection. Cell Host Microbe. 2020;27(1):54–67 e5. doi:10.1016/j.chom.2019.11.011.
  • Eisele A, Nicholas, Ruby T, Jacobson A, Manzanillo S, Paolo, Cox S, Jeffery, Lam L, Mukundan L, Chawla A, Monack D. Salmonella require the fatty acid regulator PPARδ for the establishment of a metabolic environment essential for long-term persistence. Cell Host Microbe. 2013;14(2):171–182. doi:10.1016/j.chom.2013.07.010.
  • Foster N, Tang Y, Berchieri A, Geng S, Jiao X, Barrow P. Revisiting Persistent Salmonella Infection and the Carrier State: what Do We Know? Pathogens. 2021;10(10):1299. doi:10.3390/pathogens10101299.
  • JB A, Hill J, JB C, JP A. Host restriction, pathogenesis and chronic carriage of typhoidal Salmonella. FEMS Microbiol Rev. 2021;45(5).
  • Mutai WC, Muigai AWT, Waiyaki P, Kariuki S. Multi-drug resistant Salmonella enterica serovar Typhi isolates with reduced susceptibility to ciprofloxacin in Kenya. BMC Microbiol. 2018;18(1):18. doi:10.1186/s12866-018-1161-4.
  • Bhan MK, Bahl R, Bhatnagar S. Typhoid and paratyphoid fever. Lancet. 2005;366(9487):749–762. doi:10.1016/S0140-6736(05)67181-4.
  • Gaind R, Paglietti B, Murgia M, Dawar R, Uzzau S, Cappuccinelli P, Deb M, Aggarwal P, Rubino S. Molecular characterization of ciprofloxacin-resistant Salmonella enterica serovar Typhi and Paratyphi A causing enteric fever in India. J Antimicrob Chemother. 2006;58(6):1139–1144. doi:10.1093/jac/dkl391.
  • Lawley TD, Chan K, Thompson LJ, Kim CC, Govoni GR, Monack DM, Isburg R. Genome-wide screen for Salmonella genes required for long-term systemic infection of the mouse. PLoS Pathog. 2006;2(2):e11. doi:10.1371/journal.ppat.0020011.
  • Brink T, Leiss V, Siegert P, Jehle D, Ebner JK, Schwan C, Shymanets A, Wiese S, Nürnberg B, Hensel M, et al. Salmonella Typhimurium effector SseI inhibits chemotaxis and increases host cell survival by deamidation of heterotrimeric Gi proteins. PLoS Pathog. 2018;14(8):e1007248. doi:10.1371/journal.ppat.1007248.
  • McLaughlin LM, Govoni GR, Gerke C, Gopinath S, Peng K, Laidlaw G, Chien Y-H, Jeong H-W, Li Z, Brown MD, et al. The Salmonella SPI2 effector SseI mediates long-term systemic infection by modulating host cell migration. PLoS Pathog. 2009;5(11):e1000671. doi:10.1371/journal.ppat.1000671.
  • Carden SE, Walker GT, Honeycutt J, Lugo K, Pham T, Jacobson A, Bouley D, Idoyaga J, Tsolis RM, Monack D, et al. Pseudogenization of the Secreted Effector Gene sseI Confers Rapid Systemic Dissemination of S. Typhimurium ST313 within Migratory Dendritic Cells. Cell Host Microbe. 2017;21(2):182–194. doi:10.1016/j.chom.2017.01.009.
  • Rakov AV, Mastriani E, Liu SL, Schifferli DM. Association of Salmonella virulence factor alleles with intestinal and invasive serovars. BMC Genomics. 2019;20(1):429. doi:10.1186/s12864-019-5809-8.
  • McLaughlin LM, Xu H, Carden SE, Fisher S, Reyes M, Heilshorn SC, Monack DM. A microfluidic-based genetic screen to identify microbial virulence factors that inhibit dendritic cell migration. Integr Biol (Camb). 2014;6(4):438–449. doi:10.1039/C3IB40177D.
  • Shappo MOE, Li Q, Lin Z, Hu M, Ren J, Xu Z, Pan Z, Jiao X. SspH2 as anti-inflammatory candidate effector and its contribution in Salmonella Enteritidis virulence. Microb Pathog. 2020;142:104041. doi:10.1016/j.micpath.2020.104041.
  • Bernal-Bayard J, Ramos-Morales F. Salmonella Type III Secretion Effector SlrP Is an E3 Ubiquitin Ligase for Mammalian Thioredoxin. J Biol Chem. 2009;284(40):27587–27595. doi:10.1074/jbc.M109.010363.
  • Troeger C, Forouzanfar M, Rao PC, Khalil I, Brown A, Reiner RC, Fullman N, Thompson RL, Abajobir A, Ahmed M, et al. Estimates of global, regional, and national morbidity, mortality, and aetiologies of diarrhoeal diseases: a systematic analysis for the Global Burden of Disease Study 2015. Lancet Infect Dis. 2017;17(9):909–948. doi:10.1016/S1473-3099(17)30276-1.
  • Kotloff KL, Riddle MS, Platts-Mills JA, Pavlinac P, Zaidi AKM. Shigellosis. Lancet. 2018;391(10122):801–812. doi:10.1016/S0140-6736(17)33296-8.
  • Baker S, The HC. Recent insights into Shigella. Curr Opin Infect Dis. 2018;31(5):449–454. doi:10.1097/QCO.0000000000000475.
  • Mattock E, Blocker AJ. How Do the Virulence Factors of Shigella Work Together to Cause Disease? Front Cell Infect Microbiol. 2017;7:64. doi:10.3389/fcimb.2017.00064.
  • Yang JY, Lee SN, Chang SY, Ko HJ, Ryu S, Kweon MN. A mouse model of shigellosis by intraperitoneal infection. J Infect Dis. 2014;209(2):203–215. doi:10.1093/infdis/jit399.
  • Qsm PH, Ledwaba SE, Bolick DT, Giallourou N, Yum LK, Costa DVS, Oriá RB, Barry EM, Swann JR, Lima AÂM, et al. A murine model of diarrhea, growth impairment and metabolic disturbances with Shigella flexneri infection and the role of zinc deficiency. Gut Microbes. 2019;10(5):615–630. doi:10.1080/19490976.2018.1564430.
  • Yum LK, Agaisse H. Mechanisms of bacillary dysentery: lessons learnt from infant rabbits. Gut Microbes. 2020;11(3):597–602. doi:10.1080/19490976.2019.1667726.
  • Shim DH, Suzuki T, Chang SY, Park SM, Sansonetti PJ, Sasakawa C, Kweon M-N. New animal model of shigellosis in the Guinea pig: its usefulness for protective efficacy studies. J Immunol. 2007;178(4):2476–2482. doi:10.4049/jimmunol.178.4.2476.
  • Niu C, Yang J, Liu H, Cui Y, Xu H, Wang R, Liu X, Feng E, Wang D, Pan C, et al. Role of the virulence plasmid in acid resistance of Shigella flexneri. Sci Rep. 2017;7(1):46465. doi:10.1038/srep46465.
  • Islam D, Bandholtz L, Nilsson J, Wigzell H, Christensson B, Agerberth B, Gudmundsson GH. Downregulation of bactericidal peptides in enteric infections: a novel immune escape mechanism with bacterial DNA as a potential regulator. Nat Med. 2001;7(2):180–185. doi:10.1038/84627.
  • Tran ENH, Papadopoulos M, Morona R. Relationship between O-antigen chain length and resistance to colicin E2 in Shigella flexneri. Microbiology. 2014;160(3):589–601. doi:10.1099/mic.0.074955-0.
  • Vargas M, Gascon J, De Anta MT J, Vila J. Prevalence of Shigella Enterotoxins 1 and 2 among Shigella Strains Isolated from Patients with Traveler’s Diarrhea. J Clin Microbiol. 1999;37(11):3608–3611. doi:10.1128/JCM.37.11.3608-3611.1999.
  • Yavzori M, Cohen D, Orr N. Prevalence of the genes for shigella enterotoxins 1 and 2 among clinical isolates of shigella in Israel. Epidemiol Infect. 2002;128(3):533–535. doi:10.1017/S0950268802006866.
  • Farfan MJ, Toro CS, Barry EM, Nataro JP. Shigella enterotoxin-2 is a type III effector that participates in Shigella -induced interleukin 8 secretion by epithelial cells. FEMS Immunol Med Microbiol. 2011;61(3):332–339. doi:10.1111/j.1574-695X.2011.00778.x.
  • Anderson MC, Vonaesch P, Saffarian A, Marteyn BS, Sansonetti PJ. Shigella sonnei Encodes a Functional T6SS Used for Interbacterial Competition and Niche Occupancy. Cell Host Microbe. 2017;21(6):769–76.e3. doi:10.1016/j.chom.2017.05.004.
  • Rey C, Chang -Y-Y, Latour-Lambert P, Varet H, Proux C, Legendre R, Coppée J-Y, Enninga J. Transcytosis subversion by M cell-to-enterocyte spread promotes Shigella flexneri and Listeria monocytogenes intracellular bacterial dissemination. PLoS Pathog. 2020;16(4):e1008446. doi:10.1371/journal.ppat.1008446.
  • Schroeder GN, Hilbi H. Molecular Pathogenesis of Shigella spp.: controlling Host Cell Signaling, Invasion, and Death by Type III Secretion. Clin Microbiol Rev. 2008;21(1):134–156. doi:10.1128/CMR.00032-07.
  • Arizmendi O, Picking WD, Picking WL, Payne SM. Macrophage Apoptosis Triggered by IpaD from Shigella flexneri. Infect Immun. 2016;84(6):1857–1865. doi:10.1128/IAI.01483-15.
  • Niebuhr K, Giuriato S, Pedron T, Philpott DJ, Gaits F, Sable J, Sheetz MP, Parsot C, Sansonetti PJ, Payrastre B, et al. Conversion of PtdIns(4,5)P(2) into PtdIns(5)P by the S.flexneri effector IpgD reorganizes host cell morphology. EMBO J. 2002;21(19):5069–5078. doi:10.1093/emboj/cdf522.
  • Pendaries C, Tronchère H, Arbibe L, Mounier J, Gozani O, Cantley L, Fry MJ, Gaits-Iacovoni F, Sansonetti PJ, Payrastre B, et al. PtdIns(5)P activates the host cell PI3-kinase/Akt pathway during Shigella flexneri infection. EMBO J. 2006;25(5):1024–1034. doi:10.1038/sj.emboj.7601001.
  • Günther C, Martini E, Wittkopf N, Amann K, Weigmann B, Neumann H, Waldner MJ, Hedrick SM, Tenzer S, Neurath MF, et al. Caspase-8 regulates TNF-α-induced epithelial necroptosis and terminal ileitis. Nature. 2011;477(7364):335–339. doi:10.1038/nature10400.
  • Kobayashi T, Ogawa M, Sanada T, Mimuro H, Kim M, Ashida H, Akakura R, Yoshida M, Kawalec M, Reichhart J-M, et al. The Shigella OspC3 Effector Inhibits Caspase-4, Antagonizes Inflammatory Cell Death, and Promotes Epithelial Infection. Cell Host Microbe. 2013;13(5):570–583. doi:10.1016/j.chom.2013.04.012.
  • Ashida H, Sasakawa C, Suzuki T. A unique bacterial tactic to circumvent the cell death crosstalk induced by blockade of caspase‐8. EMBO J. 2020;39.
  • Ashida H, Suzuki T, Sasakawa C. Shigella infection and host cell death: a double-edged sword for the host and pathogen survival. Curr Opin Microbiol. 2021;59:1–7. doi:10.1016/j.mib.2020.07.007.
  • Dong N, Zhu Y, Lu Q, Hu L, Zheng Y, Shao F. Structurally Distinct Bacterial TBC-like GAPs Link Arf GTPase to Rab1 Inactivation to Counteract Host Defenses. Cell. 2012;150(5):1029–1041. doi:10.1016/j.cell.2012.06.050.
  • Burnaevskiy N, Peng T, Reddick E,L, Hang C, Howard, Alto M,N. Myristoylome Profiling Reveals a Concerted Mechanism of ARF GTPase Deacylation by the Bacterial Protease IpaJ. Mol Cell. 2015;58(1):110–122. doi:10.1016/j.molcel.2015.01.040.
  • Bernardini ML, Mounier J, d’Hauteville H, Coquis-Rondon M, Sansonetti PJ. Identification of icsA, a plasmid locus of Shigella flexneri that governs bacterial intra- and intercellular spread through interaction with F-actin. Proc Natl Acad Sci U S A. 1989;86(10):3867–3871. doi:10.1073/pnas.86.10.3867.
  • Ogawa M, Yoshimori T, Suzuki T, Sagara H, Mizushima N, Sasakawa C. Escape of Intracellular Shigella from Autophagy. Science. 2005;307(5710):727–731. doi:10.1126/science.1106036.
  • Miura M, Terajima J, Izumiya H, Mitobe J, Komano T, Watanabe H. OspE2 of Shigella sonnei Is Required for the Maintenance of Cell Architecture of Bacterium-Infected Cells. Infect Immun. 2006;74(5):2587–2595. doi:10.1128/IAI.74.5.2587-2595.2006.
  • McCormick BA, Siber AM, Maurelli AT, Barbieri JT. Requirement of the Shigella flexneri Virulence Plasmid in the Ability To Induce Trafficking of Neutrophils across Polarized Monolayers of the Intestinal Epithelium. Infect Immun. 1998;66(9):4237–4243. doi:10.1128/IAI.66.9.4237-4243.1998.
  • Fukazawa A, Alonso C, Kurachi K, Gupta S, Lesser CF, McCormick BA, Reinecker H-C. GEF-H1 mediated control of NOD1 dependent NF-kappaB activation by Shigella effectors. PLoS Pathog. 2008;4(11):e1000228. doi:10.1371/journal.ppat.1000228.
  • Phalipon A, Sansonetti PJ. Shigella’ s ways of manipulating the host intestinal innate and adaptive immune system: a tool box for survival? Immunol Cell Biol. 2007;85(2):119–129. doi:10.1038/sj.icb7100025.
  • Harouz H, Rachez C, Meijer BM, Marteyn B, Donnadieu F, Cammas F, Muchardt C, Sansonetti P, Arbibe L. Shigella flexneri targets the HP 1γ subcode through the phosphothreonine lyase O sp F. EMBO J. 2014;33(22):2606–2622. doi:10.15252/embj.201489244.
  • Reiterer V, Grossniklaus L, Tschon T, Kasper CA, Sorg I, Arrieumerlou C. Shigella flexneri type III secreted effector OspF reveals new crosstalks of proinflammatory signaling pathways during bacterial infection. Cell Signal. 2011;23(7):1188–1196. doi:10.1016/j.cellsig.2011.03.006.
  • Linkous A, Yazlovitskaya E. Cytosolic phospholipase A2 as a mediator of disease pathogenesis. Cell Microbiol. 2010;12(10):1369–1377. doi:10.1111/j.1462-5822.2010.01505.x.
  • Wood TE, Westervelt KA, Yoon JM, Eshleman HD, Levy R, Burnes H, et al. The Shigella Spp. Type III Effector Protein OspB Is a Cysteine Protease mBio. 2022;13:e0127022.
  • Ashida H, Sasakawa C. Shigella IpaH Family Effectors as a Versatile Model for Studying Pathogenic Bacteria. Front Cell Infect Microbiol. 2015;5:100. doi:10.3389/fcimb.2015.00100.
  • Rohde JR, Breitkreutz A, Chenal A, Sansonetti PJ, Parsot C. Type III Secretion Effectors of the IpaH Family Are E3 Ubiquitin Ligases. Cell Host Microbe. 2007;1(1):77–83. doi:10.1016/j.chom.2007.02.002.
  • Bullones-Bolaños A, Bernal-Bayard J, Ramos-Morales F. The NEL Family of Bacterial E3 Ubiquitin Ligases. Int J Mol Sci. 2022;23(14):7725. doi:10.3390/ijms23147725.
  • Okuda J, Toyotome T, Kataoka N, Ohno M, Abe H, Shimura Y, Seyedarabi A, Pickersgill R, Sasakawa C. Shigella effector IpaH9.8 binds to a splicing factor U2AF(35) to modulate host immune responses. Biochem Biophys Res Commun. 2005;333(2):531–539. doi:10.1016/j.bbrc.2005.05.145.
  • Alphonse N, Wanford JJ, Voak AA, Gay J, Venkhaya S, Burroughs O, Mathew S, Lee T, Evans SL, Zhao W, et al. A family of conserved bacterial virulence factors dampens interferon responses by blocking calcium signaling. Cell. 2022;185(13):2354–69.e17. doi:10.1016/j.cell.2022.04.028.
  • Pinaud L, Samassa F, Porat Z, Ferrari ML, Belotserkovsky I, Parsot C, Sansonetti PJ, Campbell-Valois F-X, Phalipon A. Injection of T3SS effectors not resulting in invasion is the main targeting mechanism of Shigella toward human lymphocytes. Proc Natl Acad Sci U S A. 2017;114(37):9954–9959. doi:10.1073/pnas.1707098114.
  • Samassa F, Ferrari ML, Husson J, Mikhailova A, Porat Z, Sidaner F, Brunner K, Teo T-H, Frigimelica E, Tinevez J-Y, et al. Shigella impairs human T lymphocyte responsiveness by hijacking actin cytoskeleton dynamics and T cell receptor vesicular trafficking. Cell Microbiol. 2020;22(5):e13166. doi:10.1111/cmi.13166.
  • Nothelfer K, Arena ET, Pinaud L, Neunlist M, Mozeleski B, Belotserkovsky I, Parsot C, Dinadayala P, Burger-Kentischer A, Raqib R, et al. B lymphocytes undergo TLR2-dependent apoptosis upon Shigella infection. J Exp Med. 2014;211(6):1215–1229. doi:10.1084/jem.20130914.
  • Brunner K, Samassa F, Sansonetti PJ, Phalipon A. Shigella -mediated immunosuppression in the human gut: subversion extends from innate to adaptive immune responses. Hum Vaccin Immunother. 2019;15(6):1317–1325. doi:10.1080/21645515.2019.1594132.
  • Otsubo R, Mimuro H, Ashida H, Hamazaki J, Murata S, Sasakawa C. Shigella effector IpaH4.5 targets 19S regulatory particle subunit RPN13 in the 26S proteasome to dampen cytotoxic T lymphocyte activation. Cell Microbiol. 2019;21(3):e12974. doi:10.1111/cmi.12974.
  • Konradt C, Frigimelica E, Nothelfer K, Puhar A, Salgado-Pabon W, Bartolo V, Scott-Algara D, Rodrigues C, Sansonetti P, Phalipon A, et al. The Shigella flexneri Type Three Secretion System Effector IpgD Inhibits T Cell Migration by Manipulating Host Phosphoinositide Metabolism. Cell Host Microbe. 2011;9(4):263–272. doi:10.1016/j.chom.2011.03.010.
  • Bengtsson RJ, Dallman TJ, Allen H, De Silva PM, Stenhouse G, Pulford CV, et al. Accessory Genome Dynamics and Structural Variation of Shigella from Persistent Infections. mBio. 2021;12(2).
  • Bardsley M, Jenkins C, Mitchell HD, Mikhail AFW, Baker KS, Foster K, et al. Persistent Transmission of Shigellosis in England Is Associated with a Recently Emerged Multidrug-Resistant Strain of Shigella sonnei. J Clin Microbiol. 2020;58(4).
  • Charles H, Prochazka M, Thorley K, Crewdson A, Greig DR, Jenkins C, Painset A, Fifer H, Browning L, Cabrey P, et al. Outbreak of sexually transmitted, extensively drug-resistant Shigella sonnei in the UK, 2021–22: a descriptive epidemiological study. Lancet Infect Dis. 2022;22(10):1503–1510. doi:10.1016/S1473-3099(22)00370-X.
  • Marder EP, Cieslak PR, Cronquist AB, Dunn J, Lathrop S, Rabatsky-Ehr T, et al. Incidence and Trends of Infections with Pathogens Transmitted Commonly Through Food and the Effect of Increasing Use of Culture-Independent Diagnostic Tests on Surveillance — foodborne Diseases Active Surveillance Network, 10 U.S. Sites. MMWR Morbidity and Mortality Weekly Report 2017. Vol. 66. 397–403. 2013–2016.
  • Callahan SM, Dolislager CG, Johnson JG, Ottemann KM. The Host Cellular Immune Response to Infection by Campylobacter Spp. and Its Role in Disease. Infect Immun. 2021;89(8):e0011621. doi:10.1128/IAI.00116-21.
  • Kreling V, Falcone FH, Kehrenberg C, Hensel A. Campylobacter sp.: pathogenicity factors and prevention methods—new molecular targets for innovative antivirulence drugs? Appl Microbiol Biotechnol. 2020;104(24):10409–10436. doi:10.1007/s00253-020-10974-5.
  • Elmi A, Nasher F, Dorrell N, Wren B, Gundogdu O. Revisiting Campylobacter jejuni Virulence and Fitness Factors: role in Sensing, Adapting, and Competing. Front Cell Infect Microbiol. 2020;10:607704. doi:10.3389/fcimb.2020.607704.
  • Burnham PM, Hendrixson DR. Campylobacter jejuni: collective components promoting a successful enteric lifestyle. Nat Rev Microbiol. 2018;16(9):551–565. doi:10.1038/s41579-018-0037-9.
  • Babakhani FK, Bradley GA, Joens LA. Newborn piglet model for campylobacteriosis. Infect Immun. 1993;61(8):3466–3475. doi:10.1128/iai.61.8.3466-3475.1993.
  • Fox JG, Ackerman JI, Taylor N, Claps M, Murphy JC. Campylobacter jejuni infection in the ferret: an animal model of human campylobacteriosis. Am J Vet Res. 1987;48:85–90.
  • Askoura M, Stintzi A. Using Galleria mellonella as an Infection Model for Campylobacter jejuni Pathogenesis. Methods Mol Biol. 2017;1512:163–169.
  • Davis L, DiRita V. Experimental Chick Colonization by Campylobacter jejuni. Curr Protoc Microbiol. 2008;11(1). Chapter 8:Unit 8A 3. doi:10.1002/9780471729259.mc08a03s11.
  • Giallourou N, Medlock GL, Bolick DT, Medeiros PH, Ledwaba SE, Kolling GL, Tung K, Guerry P, Swann JR, Guerrant RL, et al. A novel mouse model of Campylobacter jejuni enteropathy and diarrhea. PLoS Pathog. 2018;14(3):e1007083. doi:10.1371/journal.ppat.1007083.
  • Chang C, Miller JF. Campylobacter jejuni Colonization of Mice with Limited Enteric Flora. Infect Immun. 2006;74(9):5261–5271. doi:10.1128/IAI.01094-05.
  • Hofreuter D. Defining the metabolic requirements for the growth and colonization capacity of Campylobacter jejuni. Front Cell Infect Microbiol. 2014;4:137. doi:10.3389/fcimb.2014.00137.
  • Liaw J, Hong G, Davies C, Elmi A, Sima F, Stratakos A, Stef L, Pet I, Hachani A, Corcionivoschi N, et al. The Campylobacter jejuni Type VI Secretion System Enhances the Oxidative Stress Response and Host Colonization. Front Microbiol. 2019;10:2864. doi:10.3389/fmicb.2019.02864.
  • Hatayama S, Shimohata T, Amano S, Kido J, Nguyen AQ, Sato Y, Kanda Y, Tentaku A, Fukushima S, Nakahashi M, et al. Cellular Tight Junctions Prevent Effective Campylobacter jejuni Invasion and Inflammatory Barrier Disruption Promoting Bacterial Invasion from Lateral Membrane in Polarized Intestinal Epithelial Cells. Front Cell Infect Microbiol. 2018;8:15. doi:10.3389/fcimb.2018.00015.
  • Sharafutdinov I, Esmaeili DS, Harrer A, Tegtmeyer N, Sticht H, Backert S. Campylobacter jejuni Serine Protease HtrA Cleaves the Tight Junction Component Claudin-8. Front Cell Infect Microbiol. 2020;10:590186. doi:10.3389/fcimb.2020.590186.
  • Harrer A, Bücker R, Boehm M, Zarzecka U, Tegtmeyer N, Sticht H, Chang F-Y, Lin J-C. Campylobacter jejuni enters gut epithelial cells and impairs intestinal barrier function through cleavage of occludin by serine protease HtrA. Gut Pathog. 2019;11:11. doi:10.1186/s13099-019-0285-x.
  • Hoy B, Geppert T, Boehm M, Reisen F, Plattner P, Gadermaier G, Sewald N, Ferreira F, Briza P, Schneider G, et al. Distinct Roles of Secreted HtrA Proteases from Gram-negative Pathogens in Cleaving the Junctional Protein and Tumor Suppressor E-cadherin. J Biol Chem. 2012;287(13):10115–10120. doi:10.1074/jbc.C111.333419.
  • Louwen R, Nieuwenhuis EES, Van Marrewijk L, Horst-Kreft D, De Ruiter L, Heikema AP, van Wamel WJB, Wagenaar JA, Endtz HP, Samsom J, et al. Campylobacter jejuni Translocation across Intestinal Epithelial Cells Is Facilitated by Ganglioside-Like Lipooligosaccharide Structures. Infect Immun. 2012;80(9):3307–3318. doi:10.1128/IAI.06270-11.
  • Grant CC, Konkel ME, Cieplak W, Tompkins LS. Role of flagella in adherence, internalization, and translocation of Campylobacter jejuni in nonpolarized and polarized epithelial cell cultures. Infect Immun. 1993;61(5):1764–1771. doi:10.1128/iai.61.5.1764-1771.1993.
  • Harvey P, Battle T, Leach S. Different invasion phenotypes of Campylobacter isolates in Caco-2 cell monolayers. J Med Microbiol. 1999;48(5):461–469. doi:10.1099/00222615-48-5-461.
  • Konkel ME, Mead DJ, Hayes SF, Cieplak W Jr. Translocation of Campylobacter jejuni across human polarized epithelial cell monolayer cultures. J Infect Dis. 1992;166(2):308–315. doi:10.1093/infdis/166.2.308.
  • Boehm M, Hoy B, Rohde M, Tegtmeyer N, Bæk KT, Oyarzabal OA. Immune response to and histopathology of Campylobacter jejuni infection in ferrets (Mustela putorius furo). Comp Med. 2009;59:363–371.
  • Beltinger J, Buono JD, Skelly MM, Thornley J, Spiller RC, Stack WA, Hawkey CJ. Disruption of colonic barrier function and induction of mediator release by strains of Campylobacter jejuni that invade epithelial cells. World J Gastroenterol. 2008;14(48):7345. doi:10.3748/wjg.14.7345.
  • Van Deun K, Pasmans F, Van Immerseel F, Ducatelle R, Haesebrouck F. Butyrate protects Caco-2 cells from Campylobacter jejuni invasion and translocation. Brit J Nutrition. 2008;100(3):480–484. doi:10.1017/S0007114508921693.
  • Kalischuk LD, Inglis GD, Buret AG. Campylobacter jejuni induces transcellular translocation of commensal bacteria via lipid rafts. Gut Pathog. 2009;1(1):2. doi:10.1186/1757-4749-1-2.
  • Alzheimer M, Svensson SL, König F, Schweinlin M, Metzger M, Walles H, Sharma CM. A three-dimensional intestinal tissue model reveals factors and small regulatory RNAs important for colonization with Campylobacter jejuni. PLoS Pathog. 2020;16(2):e1008304. doi:10.1371/journal.ppat.1008304.
  • Backert S, Boehm M, Wessler S, Tegtmeyer N. Transmigration route of Campylobacter jejuni across polarized intestinal epithelial cells: paracellular, transcellular or both? Cell Com Signal. 2013;11(1):72. doi:10.1186/1478-811X-11-72.
  • Everest PH, Goossens H, Sibbons P, Lloyd DR, Knutton S, Leece R, Ketley JM, Williams PH. Pathological changes in the rabbit ileal loop model caused by Campylobacter jejuni from human colitis. J Med Microbiol. 1993;38(5):316–321. doi:10.1099/00222615-38-5-316.
  • Nemelka KW, Brown AW, Wallace SM, Jones E, Asher LV, Pattarini D, Applebee L, Gilliland TC, Guerry P, Baqar S, et al. Immune response to and histopathology of Campylobacter jejuni infection in ferrets (Mustela putorius furo). Comp Med. 2009;59(4):363–371.
  • Vuckovic D, Abram M, Doric M. Primary Campylobacter jejuni infection in different mice strains. Microb Pathog. 1998;24(4):263–268. doi:10.1006/mpat.1997.0194.
  • Watson KG, Holden DW. Dynamics of growth and dissemination of Salmonella in vivo. Cell Microbiol. 2010;12(10):1389–1397. doi:10.1111/j.1462-5822.2010.01511.x.
  • Lamb-Rosteski JM, Kalischuk LD, Inglis GD, Buret AG. Epidermal Growth Factor Inhibits Campylobacter jejuni -Induced Claudin-4 Disruption, Loss of Epithelial Barrier Function, and Escherichia coli Translocation. Infect Immun. 2008;76(8):3390–3398. doi:10.1128/IAI.01698-07.
  • Ilyas B, Tsai CN, Coombes BK. Evolution of Salmonella-Host Cell Interactions through a Dynamic Bacterial Genome. Front Cell Infect Microbiol. 2017;7:428. doi:10.3389/fcimb.2017.00428.
  • Ruby T, McLaughlin L, Gopinath S, Monack D. Salmonella ‘s long-term relationship with its host. FEMS Microbiol Rev. 2012;36(3):600–615. doi:10.1111/j.1574-6976.2012.00332.x.
  • Guerry P, Poly F, Riddle M, Maue AC, Chen YH, Monteiro MA. Campylobacter polysaccharide capsules: virulence and vaccines. Front Cell Infect Microbiol. 2012;2:7. doi:10.3389/fcimb.2012.00007.
  • Kreutzberger MAB, Ewing C, Poly F, Wang F, Egelman EH. Atomic structure of the Campylobacter jejuni flagellar filament reveals how ε Proteobacteria escaped Toll-like receptor 5 surveillance. Proc Natl Acad Sci U S A. 2020;117(29):16985–16991. doi:10.1073/pnas.2010996117.
  • Frirdich E, Biboy J, Pryjma M, Lee J, Huynh S, Parker CT, Girardin SE, Vollmer W, Gaynor EC. The Campylobacter jejuni helical to coccoid transition involves changes to peptidoglycan and the ability to elicit an immune response. Mol Microbiol. 2019;112(1):280–301. doi:10.1111/mmi.14269.
  • Neal-McKinney JM, Konkel ME. The Campylobacter jejuni CiaC virulence protein is secreted from the flagellum and delivered to the cytosol of host cells. Front Cell Infect Microbiol. 2012;2:31. doi:10.3389/fcimb.2012.00031.
  • Konkel ME, Klena JD, Rivera-Amill V, Monteville MR, Biswas D, Raphael B, Mickelson J. Secretion of Virulence Proteins from Campylobacter jejuni Is Dependent on a Functional Flagellar Export Apparatus. J Bacteriol. 2004;186(11):3296–3303. doi:10.1128/JB.186.11.3296-3303.2004.
  • Lara-Tejero M, Galan JE. A bacterial toxin that controls cell cycle progression as a deoxyribonuclease I-like protein. Science. 2000;290(5490):354–357. doi:10.1126/science.290.5490.354.
  • Tang Y, Fang L, Xu C, Zhang Q. Antibiotic resistance trends and mechanisms in the foodborne pathogen, Campylobacter. Anim Health Res Rev. 2017;18(2):87–98. doi:10.1017/S1466252317000135.
  • Lozupone CA, Stombaugh JI, Gordon JI, Jansson JK, Knight R. Diversity, stability and resilience of the human gut microbiota. Nature. 2012;489(7415):220–230. doi:10.1038/nature11550.
  • Bandoy DDR, Weimer BC. Biological Machine Learning Combined with Campylobacter Population Genomics Reveals Virulence Gene Allelic Variants Cause Disease. Microorganisms. 2020;8(4):549. doi:10.3390/microorganisms8040549.
  • Artymovich K, Kim J-S, Linz JE, Hall DF, Kelley LE, Kalbach HL, Kathariou S, Gaymer J, Paschke B. A “successful allele” at Campylobacter jejuni contingency locus Cj0170 regulates motility; “successful alleles” at locus Cj0045 are strongly associated with mouse colonization. Food Microbiol. 2013;34(2):425–430. doi:10.1016/j.fm.2013.01.007.
  • Gupta V, Gulati P, Bhagat N, Dhar MS, Virdi JS. Detection of Yersinia enterocolitica in food: an overview. Eur J Clinic Microbiol Infect Dis. 2015;34(4):641–650. doi:10.1007/s10096-014-2276-7.
  • Schubert KA, Xu Y, Shao F, Auerbuch V. The Yersinia Type III Secretion System as a Tool for Studying Cytosolic Innate Immune Surveillance. Annu Rev Microbiol. 2020;74(1):221–245. doi:10.1146/annurev-micro-020518-120221.
  • Zhang L, Mei M, Yu C, Shen W, Ma L, He J, Yi LI. The Functions of Effector Proteins in Yersinia Virulence. Pol J Microbiol. 2016;65(1):5–12. doi:10.5604/17331331.1197324.
  • Grabowski B, Schmidt MA, Rüter C. Immunomodulatory Yersinia outer proteins (Yops)–useful tools for bacteria and humans alike. Virulence. 2017;8(7):1124–1147. doi:10.1080/21505594.2017.1303588.
  • Carter PB. Animal model of human disease. Yersinia enteritis. Animal model: oral Yersinia enterocolitica infection of mice. Am J Pathol. 1975;81:703–706.
  • Hancock GE, Schaedler RW, MacDonald TT. Yersinia enterocolitica infection in resistant and susceptible strains of mice. Infect Immun. 1986;53(1):26–31. doi:10.1128/iai.53.1.26-31.1986.
  • Hooker-Romero D, Schwiesow L, Wei Y, Auerbuch V. Mouse Models of Yersiniosis. Methods Mol Biol. 2019;2010:41–53.
  • Fahlgren A, Avican K, Westermark L, Nordfelth R, Fallman M, Bäumler AJ. Colonization of cecum is important for development of persistent infection by Yersinia pseudotuberculosis. Infect Immun. 2014;82(8):3471–3482. doi:10.1128/IAI.01793-14.
  • Ducarmon QR, Zwittink RD, Hornung BVH, Van Schaik W, Young VB, Kuijper EJ. Gut Microbiota and Colonization Resistance against Bacterial Enteric Infection. Microbiol Mol Biol Rev. 2019;83.
  • Isberg RR, Falkow S. A single genetic locus encoded by Yersinia pseudotuberculosis permits invasion of cultured animal cells by Escherichia coli K-12. Nature. 1985;317(6034):262–264. doi:10.1038/317262a0.
  • Eitel J, Dersch P. The YadA Protein of Yersinia pseudotuberculosis Mediates High-Efficiency Uptake into Human Cells under Environmental Conditions in Which Invasin Is Repressed. Infect Immun. 2002;70(9):4880–4891. doi:10.1128/IAI.70.9.4880-4891.2002.
  • Isberg RR, Voorhis DL, Falkow S. Identification of invasin: a protein that allows enteric bacteria to penetrate cultured mammalian cells. Cell. 1987;50(5):769–778. doi:10.1016/0092-8674(87)90335-7.
  • Drechsler-Hake D, Alamir H, Hahn J, Gunter M, Wagner S, Schutz M, Bohn E, Schenke-Layland K, Pisano F, Dersch P, et al. Mononuclear phagocytes contribute to intestinal invasion and dissemination of Yersinia enterocolitica. Int J Med Microbiol. 2016;306(6):357–366. doi:10.1016/j.ijmm.2016.04.002.
  • Xu H, Yang J, Gao W, Li L, Li P, Zhang L, Gong Y-N, Peng X, Xi JJ, Chen S, et al. Innate immune sensing of bacterial modifications of Rho GTPases by the Pyrin inflammasome. Nature. 2014;513(7517):237–241. doi:10.1038/nature13449.
  • Straley SC, Cibull ML. Differential clearance and host-pathogen interactions of YopE- and YopK- YopL- Yersinia pestis in BALB/c mice. Infect Immun. 1989;57(4):1200–1210. doi:10.1128/iai.57.4.1200-1210.1989.
  • Mills SD, Boland A, Sory MP, van der Smissen P, Kerbourch C, Finlay BB, Cornelis GR. Yersinia enterocolitica induces apoptosis in macrophages by a process requiring functional type III secretion and translocation mechanisms and involving YopP,presumably acting as an effector protein. Proc Natl Acad Sci U S A. 1997;94(23):12638–12643. doi:10.1073/pnas.94.23.12638.
  • Monack DM, Mecsas J, Ghori N, Falkow S. Yersinia signals macrophages to undergo apoptosis and YopJ is necessary for this cell death. Proc Natl Acad Sci U S A. 1997;94(19):10385–10390. doi:10.1073/pnas.94.19.10385.
  • Malireddi RKS, Kesavardhana S, Kanneganti TD. ZBP1 and TAK1: master Regulators of NLRP3 Inflammasome/Pyroptosis, Apoptosis, and Necroptosis (PAN-optosis). Front Cell Infect Microbiol. 2019;9:406. doi:10.3389/fcimb.2019.00406.
  • Malireddi RKS, Kesavardhana S, Karki R, Kancharana B, Burton AR, Kanneganti TD. RIPK1 Distinctly Regulates Yersinia -Induced Inflammatory Cell Death, PANoptosis. Immunohorizons. 2020;4(12):789–796. doi:10.4049/immunohorizons.2000097.
  • Mares CA, Lugo FP, Albataineh M, Goins BA, Newton IG, Isberg RR, Bergman MA. Heightened Virulence of Yersinia Is Associated with Decreased Function of the YopJ Protein. Infect Immun. 2021;89(12):e0043021. doi:10.1128/IAI.00430-21.
  • Le Baut G, O’Brien C, Pavli P, Roy M, Seksik P, Treton X, Nancey S, Barnich N, Bezault M, Auzolle C, et al. Prevalence of Yersinia Species in the Ileum of Crohn’s Disease Patients and Controls. Front Cell Infect Microbiol. 2018;8:336. doi:10.3389/fcimb.2018.00336.
  • Honda K, Iwanaga N, Izumi Y, Tsuji Y, Kawahara C, Michitsuji T, Higashi S, Kawakami A, Migita K. Reactive Arthritis Caused by Yersinia enterocolitica Enteritis. Intern Med. 2017;56(10):1239–1242. doi:10.2169/internalmedicine.56.7888.
  • Avican K, Fahlgren A, Huss M, Heroven AK, Beckstette M, Dersch P, Fällman M. Reprogramming of Yersinia from virulent to persistent mode revealed by complex in vivo RNA-seq analysis. PLoS Pathog. 2015;11(1):e1004600. doi:10.1371/journal.ppat.1004600.
  • Wang H, Avican K, Fahlgren A, Erttmann SF, Nuss AM, Dersch P, Fallman M, Edgren T, Wolf-Watz H. Increased plasmid copy number is essential for YersiniaT3SS function and virulence. Science. 2016;353(6298):492–495. doi:10.1126/science.aaf7501.
  • Schneiders S, Hechard T, Edgren T, Avican K, Fallman M, Fahlgren A, et al. Spatiotemporal Variations in Growth Rate and Virulence Plasmid Copy Number during Yersinia pseudotuberculosis Infection. Infect Immun. 2021;89(4).
  • Fonseca D, Morais D, Hand W, Timothy H, Gerner S-J, Michael Y, Zaretsky G, Arielle B, Allyson L, Trinchieri G, et al. Microbiota-Dependent Sequelae of Acute Infection Compromise Tissue-Specific Immunity. Cell. 2015;163(2):354–366. doi:10.1016/j.cell.2015.08.030.
  • Randolph GJ, Bala S, Rahier JF, Johnson MW, Wang PL, Nalbantoglu I, Dubuquoy L, Chau A, Pariente B, Kartheuser A, et al. Lymphoid Aggregates Remodel Lymphatic Collecting Vessels that Serve Mesenteric Lymph Nodes in Crohn Disease. Am J Pathol. 2016;186(12):3066–3073. doi:10.1016/j.ajpath.2016.07.026.
  • Weil AA, Becker RL, Harris JB. Vibrio cholerae at the Intersection of Immunity and the Microbiome. mSphere. 2019;4(6).
  • Baker-Austin C, Oliver JD, Alam M, Ali A, Waldor MK, Qadri F, Martinez-Urtaza J. Vibrio spp. infections. Nature Rev Dis Primers. 2018;4(1):1–19. doi:10.1038/s41572-018-0005-8.
  • Baker-Austin C, Trinanes J, Gonzalez-Escalona N, Martinez-Urtaza J. Non-Cholera Vibrios: the Microbial Barometer of Climate Change. Trends Microbiol. 2017;25(1):76–84. doi:10.1016/j.tim.2016.09.008.
  • Ali M, Qadri F, Kim DR, Islam MT, Im J, Ahmmed F, Khan AI, Zaman K, Marks F, Kim JH, et al. Effectiveness of a killed whole-cell oral cholera vaccine in Bangladesh: further follow-up of a cluster-randomised trial. Lancet Infect Dis. 2021;21(10):1407–1414. doi:10.1016/S1473-3099(20)30781-7.
  • Deshayes S, Daurel C, Cattoir V, Parienti -J-J, Quilici M-L, De La Blanchardière A. Non-O1, non-O139 Vibrio cholerae bacteraemia: case report and literature review. SpringerPlus. 2015;4(1):4. doi:10.1186/2193-1801-4-4.
  • Ramamurthy T, Nandy RK, Mukhopadhyay AK, Dutta S, Mutreja A, Okamoto K, Miyoshi S-I, Nair GB, Ghosh A. Virulence Regulation and Innate Host Response in the Pathogenicity of Vibrio cholerae. Front Cell Infect Microbiol. 2020;10:572096. doi:10.3389/fcimb.2020.572096.
  • Miller KA, Tomberlin KF, Dziejman M. Vibrio variations on a type three theme. Curr Opin Microbiol. 2019;47:66–73. doi:10.1016/j.mib.2018.12.001.
  • Clemens JD, Nair GB, Ahmed T, Qadri F, Holmgren JC. Cholera. Lancet. 2017;390(10101):1539–1549. doi:10.1016/S0140-6736(17)30559-7.
  • Sit B, Fakoya B, Waldor MK. Animal models for dissecting Vibrio cholerae intestinal pathogenesis and immunity. Curr Opin Microbiol. 2022;65:1–7. doi:10.1016/j.mib.2021.09.007.
  • Blow NS, Salomon RN, Garrity K, Reveillaud I, Kopin A, Jackson FR, Watnick PI. Vibrio cholerae Infection of Drosophila melanogaster Mimics the Human Disease Cholera. PLoS Pathog. 2005;1(1):e8. doi:10.1371/journal.ppat.0010008.
  • Runft DL, Mitchell KC, Abuaita BH, Allen JP, Bajer S, Ginsburg K, Neely MN, Withey JH. Zebrafish as a Natural Host Model for Vibrio cholerae Colonization and Transmission. Appl Environ Microbiol. 2014;80(5):1710–1717. doi:10.1128/AEM.03580-13.
  • Vaitkevicius K, Lindmark B, Ou G, Song T, Toma C, Iwanaga M, Zhu J, Andersson A, Hammarström M-L, Tuck S, et al. A Vibrio cholerae protease needed for killing of Caenorhabditis elegans has a role in protection from natural predator grazing. Proc Natl Acad Sci U S A. 2006;103(24):9280–9285. doi:10.1073/pnas.0601754103.
  • Zhao W, Caro F, Robins W, Mekalanos JJ. Antagonism toward the intestinal microbiota and its effect on Vibrio cholerae virulence. Science. 2018;359(6372):210–213. doi:10.1126/science.aap8775.
  • Bourque DL, Bhuiyan TR, Genereux DP, Rashu R, Ellis CN, Chowdhury F, et al. Analysis of the Human Mucosal Response to Cholera Reveals Sustained Activation of Innate Immune Signaling Pathways. Infect Immun. 2018;86(2).
  • Silva AJ, Pham K, Benitez JA. Haemagglutinin/protease expression and mucin gel penetration in El Tor biotype Vibrio cholerae. Microbiology. 2003;149(7):1883–1891. doi:10.1099/mic.0.26086-0.
  • Martens EC, Neumann M, Desai MS. Interactions of commensal and pathogenic microorganisms with the intestinal mucosal barrier. Nat Rev Microbiol. 2018;16(8):457–470. doi:10.1038/s41579-018-0036-x.
  • Bartlett TM, Bratton BP, Duvshani A, Miguel A, Sheng Y, Martin NR, Nguyen JP, Persat A, Desmarais SM, VanNieuwenhze MS, et al. A Periplasmic Polymer Curves Vibrio cholerae and Promotes Pathogenesis. Cell. 2017;168(1–2):172–85.e15. doi:10.1016/j.cell.2016.12.019.
  • Herrington DA, Hall RH, Losonsky G, Mekalanos JJ, Taylor RK, Levine MM. Toxin, toxin-coregulated pili, and the toxR regulon are essential for Vibrio cholerae pathogenesis in humans. J Exp Med. 1988;168(4):1487–1492. doi:10.1084/jem.168.4.1487.
  • Midani FS, Weil AA, Chowdhury F, Begum YA, Khan AI, Debela MD, Durand HK, Reese AT, Nimmagadda SN, Silverman JD, et al. Human Gut Microbiota Predicts Susceptibility to Vibrio cholerae Infection. J Infect Dis. 2018;218(4):645–653. doi:10.1093/infdis/jiy192.
  • Levade I, Saber MM, Midani FS, Chowdhury F, Khan AI, Begum YA, Ryan ET, David LA, Calderwood SB, Harris JB, et al. Predicting Vibrio cholerae Infection and Disease Severity Using Metagenomics in a Prospective Cohort Study. J Infect Dis. 2021;223(2):342–351. doi:10.1093/infdis/jiaa358.
  • Alavi S, Mitchell JD, Cho JY, Liu R, Macbeth JC, Hsiao A. Interpersonal Gut Microbiome Variation Drives Susceptibility and Resistance to Cholera Infection. Cell. 2020;181(7):1533–46.e13. doi:10.1016/j.cell.2020.05.036.
  • Hsiao A, Ahmed AMS, Subramanian S, Griffin NW, Drewry LL, Petri WA, Haque R, Ahmed T, Gordon JI. Members of the human gut microbiota involved in recovery from Vibrio cholerae infection. Nature. 2014;515(7527):423–426. doi:10.1038/nature13738.
  • Barrasso K, Chac D, Debela MD, Geigel C, Steenhaut A, Rivera Seda A, et al. Impact of a human gut microbe on Vibrio cholerae host colonization through biofilm enhancement. eLife. 2022;11.
  • Cho JY, Liu R, Macbeth JC, Hsiao A. The Interface of Vibrio cholerae and the Gut Microbiome. Gut Microbes. 2021;13(1):1937015. doi:10.1080/19490976.2021.1937015.
  • Wu Z, Nybom P, Magnusson K-E. Distinct effects of Vibrio cholerae haemagglutinin/protease on the structure and localization of the tight junction-associated proteins occludin and ZO-1. Cell Microbiol. 2000;2(1):11–17. doi:10.1046/j.1462-5822.2000.00025.x.
  • Miller KA, Chaand M, Gregoire S, Yoshida T, Beck LA, Ivanov AI, et al. Characterization of V . cholerae T3SS-dependent cytotoxicity in cultured intestinal epithelial cells. Cell Microbiol. 2016;18(12):1857–1870. doi:10.1111/cmi.12629.
  • Dziejman M, Serruto D, Tam VC, Sturtevant D, Diraphat P, Faruque SM, Rahman MH, Heidelberg JF, Decker J, Li L, et al. Genomic characterization of non-O1, non-O139 Vibrio cholerae reveals genes for a type III secretion system. Proc Natl Acad Sci U S A. 2005;102(9):3465–3470. doi:10.1073/pnas.0409918102.
  • Hubbard TP, Chao MC, Abel S, Blondel CJ, Abel Zur Wiesch P, Zhou X, Davis BM, Waldor MK. Genetic analysis of Vibrio parahaemolyticus intestinal colonization. Proc Natl Acad Sci U S A. 2016;113(22):6283–6288. doi:10.1073/pnas.1601718113.
  • Miller KA, Sofia MK, Weaver JWA, Seward CH, Dziejman M, DiRita VJ. Regulation by ToxR-Like Proteins Converges on vttR B Expression To Control Type 3 Secretion System-Dependent Caco2-BBE Cytotoxicity in Vibrio cholerae. J Bacteriol. 2016;198(11):1675–1682. doi:10.1128/JB.00130-16.
  • Peschek N, Herzog R, Singh PK, Sprenger M, Meyer F, Fröhlich KS, et al. RNA-mediated control of cell shape modulates antibiotic resistance in Vibrio cholerae. Nat Commun. 2020;11.
  • Das B, Verma J, Kumar P, Ghosh A, Ramamurthy T. Antibiotic resistance in Vibrio cholerae: understanding the ecology of resistance genes and mechanisms. Vaccine. 2020;38(Suppl 1):A83–A92. doi:10.1016/j.vaccine.2019.06.031.
  • Bloom BR, Shevach E. Requirement for T cells in the production of migration inhibitory factor. J Exp Med. 1975;142(5):1306–1311. doi:10.1084/jem.142.5.1306.
  • Becker D, Selbach M, Rollenhagen C, Ballmaier M, Meyer TF, Mann M, Bumann D. Robust Salmonella metabolism limits possibilities for new antimicrobials. Nature. 2006;440(7082):303–307. doi:10.1038/nature04616.