Next Article in Journal
Epidemiological and Clinical Aspects of Cutaneous and Mucosal Leishmaniases in Portugal: Retrospective Analysis of Cases Diagnosed in Public Hospitals and Reported in the Literature between 2010 and 2020
Previous Article in Journal
Genomic Characterization of Listeria monocytogenes and Other Listeria Species Isolated from Sea Turtles
Previous Article in Special Issue
Contribution of the Type III Secretion System (T3SS2) of Vibrio parahaemolyticus in Mitochondrial Stress in Human Intestinal Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Non-O1/Non-O139 Vibrio cholerae—An Underestimated Foodborne Pathogen? An Overview of Its Virulence Genes and Regulatory Systems Involved in Pathogenesis

by
Quantao Zhang
,
Thomas Alter
and
Susanne Fleischmann
*
Institute of Food Safety and Food Hygiene, School of Veterinary Medicine, Freie Universität Berlin, Königsweg 69, 14163 Berlin, Germany
*
Author to whom correspondence should be addressed.
Microorganisms 2024, 12(4), 818; https://doi.org/10.3390/microorganisms12040818
Submission received: 29 February 2024 / Revised: 5 April 2024 / Accepted: 15 April 2024 / Published: 18 April 2024
(This article belongs to the Special Issue Vibrio Virulence)

Abstract

:
In recent years, the number of foodborne infections with non-O1 and non-O139 Vibrio cholerae (NOVC) has increased worldwide. These have ranged from sporadic infection cases to localized outbreaks. The majority of case reports describe self-limiting gastroenteritis. However, severe gastroenteritis and even cholera-like symptoms have also been described. All reported diarrheal cases can be traced back to the consumption of contaminated seafood. As climate change alters the habitats and distribution patterns of aquatic bacteria, there is a possibility that the number of infections and outbreaks caused by Vibrio spp. will further increase, especially in countries where raw or undercooked seafood is consumed or clean drinking water is lacking. Against this background, this review article focuses on a possible infection pathway and how NOVC can survive in the human host after oral ingestion, colonize intestinal epithelial cells, express virulence factors causing diarrhea, and is excreted by the human host to return to the environment.

1. Introduction

Vibrio (V.) cholerae is mainly known as the causative agent of the endemic and epidemic diarrheal disease cholera. However, V. cholerae is a globally distributed aquatic commensal that has been classified into more than 200 serogroups. Only two serogroups, O1 and O139, have the ability to cause pandemic cholera outbreaks. Since 1961, the V. cholerae serotype O1 biotype El Tor has been the predominant strain in the seventh pandemic, and since 1992, the V. cholerae serotype O139 has been the predominant strain in the eighth pandemic. Both pandemics are still ongoing today [1,2]. The World Health Organization (WHO) reported outbreaks of cholera in 30 countries in Asia, Africa, and America between 1 January and 15 December 2023, with over 667,000 cases and 4000 deaths [3].
Non-O1 and non-O139 V. cholerae (NOVC) serogroups are less in the focus of public health interest compared to O1 and O139 V. cholerae as they cause single-case disease or even localized outbreaks with milder and often self-limiting symptoms. In fact, one or both of the main virulence factors, cholera toxin (CT) and toxin-coregulated pilus (TCP), are missing in their genomes. Nevertheless, NOVCs are playing an increasingly important role in public health worldwide. Several studies have shown that the number of infections and outbreaks caused by NOVC has increased over time, being positively correlated with the progressive rise in seawater temperatures [4,5,6,7]. This is promoted by the anthropization of coastal regions, the increasing global trade of seafood, the trend towards the consumption of raw seafood (e.g., oysters and sushi), and the increasing number of immunocompromised people, especially older people with pre-existing diseases [7,8]. Particularly people with a compromised immune system can suffer from severe diarrhea with cholera-like symptoms. Bacteremia can also be caused by an orally acquired infection via the infiltration of NOVC in the bloodstream through the portal vein and intestinal lymphatic system [8,9,10]. Individual infection cases that could be clearly attributed to the consumption of contaminated seafood have been described in Spain [11], Italy [12], Portugal [13], India [14], Australia [15], the USA [16], and Iran [17]. Localized NOVC outbreaks have been reported in India and Thailand in the past [18,19,20,21,22,23]. Meanwhile, NOVC outbreaks have also been described in the USA [24,25], China [26,27], and Chile [28] which have also been linked to seafood consumption.
Octavia et al., 2013, pointed out that a combination of virulence factors in the genome of clinical NOVC is a prerequisite for a successful infection process [29]. The combined virulence factors identified in the genomes of NOVC isolated from the above-mentioned infection cases and local outbreaks are the Vibrio pathogenicity islands VSP-2 and VPI-2, genomic islands (GI) encoding type III (T3SS) and type VI secretion systems (T6SS), enterotoxins (RtxA and Stn), and the hemolysin HlyA. We were able to show that genes encoding these virulence factors are also present in NOVC isolated from seafood and the environment in previous studies [30,31]. Further investigations showed that other virulence genes are also present in the NOVC genomes which could also play a role in the infection process, such as hapA for hemagglutinin protease; mshA for mannose sensitive hemagglutinin; and frhA, gbpA, and mam7 for non-specific adhesins [30].
In addition to the presence of virulence factors, genes involved in host adaption and colonization are also required in the pathogenicity process. Before a successful infection, pathogens need to survive the host defense system such as acidic pH values in the stomach, anti-microbial peptides, reactive oxygen species (ROS), and an already predominant gut microbiota [32]. Mucosal penetration and epithelial cell attachment in the small intestine are also necessary for the final infection and proliferation of the bacteria [33]. V. cholerae has evolved a complex regulation system to ensure proper arrangement of various effective factors throughout the infection inside a human host, such as the quorum sensing system, two-component system, histone-like nucleoid structuring protein (Hns), small molecule signals (c-di-GMP), biofilm promotor and motility repressor modulation, and wide spectrum regulator (cAMP-CRP) [34].
Thus far, there are several comprehensive overviews about the virulence-associated genes in both O1/O139 V. cholerae and NOVCs [35,36]. Nevertheless, it should be noted that the relationship between virulence factors and a resulting infection is complex, and an interaction network rather than individual virulence factors must be considered at this point. The previous findings on virulence-associated genes and their interaction with other genetic features involved in the infection process will be discussed in this review article. Furthermore, a genetic model of a theoretical infection caused by NOVC inspired by the Kyoto Encyclopedia of Genes and Genomes (KEGG) mapping tool [32,37,38,39] was developed (see Figure 1). The whole infection workflow was divided into five stages as follows. Stage 1: survival in host gastrointestinal tract; stage 2: localization and penetration of the mucus layer in the small intestine; stage 3: intestinal epithelial cell colonization; stage 4: virulence gene expression; stage 5: detachment from the epithelial cells to return in the environment.

2. Stage 1: Survival in the Gastrointestinal Tract

After oral ingestion, pathogenic bacteria will encounter a set of host-derived defense systems, including chemical and biological barriers, when entering the stomach and arriving at the small intestine. Therefore, various genes involved in adaptation processes as a response to these conditions can be found in NOVCs to ensure that they reach the small intestine to interact with epithelial cells [40,41]. The stage 1 section therefore describes the adaptation to low pH values in the stomach [41,42]; the adaptation to reactive nitrogen and oxygen species in the stomach [43,44]; changes in porin channel size to prevent the diffusion of harmful molecules into the bacterial cell such as bile salts from the gallbladder in the duodenum [45]; efflux pumps to displace harmful molecules such as bile in the duodenum and antimicrobial peptides in the small intestine [46]; the formation of protective biofilms to protect bacteria against antimicrobial substances from the stomach, duodenum, and the small intestine; and the T6SS to compete with the predominant gut microbiome [47]. All mechanisms and genes involved in stage 1 are shown in orange in Figure 1.

2.1. Acid Tolerance Response

A common feature of diarrheal pathogens is the acid tolerance response (ATR), necessary to survive the acidic pH environment in the stomach, which is a prerequisite for subsequent successful intestinal colonization [42]. In V. cholerae, the cadABC operon first described in Escherichia (E.) coli is important for protecting the bacteria from acid hydrolysis [40]. The genetic presence of the cadABC operon was identified in 90% of NOVC isolated from seafood and the environment in our previous studies, with genetic identities over 90% compared to the V. cholerae O1 El Tor biotype, suggesting a fully functional cadABC operon [30,31]. In particular, the cadA gene encodes a lysine decarboxylase that binds protons through the production of cadaverine and carbon dioxide. Finally, this antiporter system transfers protons out of the bacterial cell and neutralizes the pH value [40]. Kovacikova et al., 2010, mentioned that cadC, the regulator of the cadABC operon, can be directly activated by aphB encoding a cytoplasmic DNA-binding protein which will be upregulated during acid stress [48]. Additionally, the expression of clcA, a hydrochloric acid transporter, regulated by aphB, plays a role in neutralizing the pH value in the bacterial cell [49].

2.2. Adaptation to Reactive Nitrogen and Oxygen Species

In the stomach, nitrite from food and saliva that is exposed to the acidic milieu results in acidified nitrite, which can be reduced by reactive nitrogen species (RNS) to antimicrobially active nitric oxide. By the detoxification of RNS, the expression of the genes nnrS and hmpA plays an important role in V. cholerae, and these genes were also identified in NOVCs from seafood and the environment [30,31]. Both genes encode enzymes that are capable of destroying nitric oxide. It is assumed that the regulator for both genes is norR, although this regulator is not stimulated by nitric oxide [43].
In diarrheal diseases, the level of reactive oxygen species (ROS) in the host gastrointestinal tract increases, resulting in damage to the bacterial cell structure as an immune defense. In V. cholerae, genes with ROS resistance activity have been identified as part of ROS removal. Superoxide dismutases such as manganese-binding SodA, for example, convert superoxide into hydrogen peroxide and oxygen. Catalases such as KatB and KatG later detoxify peroxides into water and oxygen [50,51]. The organic hydroperoxidase OhrA and preoxiredoxins such as PrxA and AphC cleave organic (alkyl) hydroperoxides [52,53]. Two homologs of ohrR, the gene for hydrogen peroxide resistance in E. coli, were found as well in V. cholerae, namely oxyR1 and oxyR2, which have a modulating function on prxA and aphC, respectively [52].

2.3. Resistance Nodulation Division (RND) Efflux Pump

After passing through the stomach, pathogenic bacteria in the small intestine must resist against host-derived bile salts, organic acids, and antimicrobial peptides. The main systems which help the bacteria to pump numerous poisonous compounds out of the cell are efflux pumps [54]. The RND efflux pump is a multi-functional unit in both O1/O139 V. cholerae and NOVC [55], encoded by a vex gene cluster (vexAB, vexCD, vexEF, vexGH, vexIJK, and vexLM) and a shared outer membrane porin encoded by tolC. Using an infant mouse model, small intestine colonization deficiency was found in different RND mutants [46,56].

2.4. Outer Membrane Protein (OMP)

Another response to toxic components such as bile salts is the alteration of porins in the bacterial cell membrane. In V. cholerae, expression and upregulation of ompU take place when bile salts are present. OmpU is widely present in NOVC [9,57,58] and should have the same response to the presence of bile. Due to its smaller channel size which prevents the influx of bile salts into the bacterial cell, the porin OmpU will be replaced instead of a larger channel porin such as OmpT [45,59].

2.5. Biofilm Formation

Biofilms provide a continuous protective cover around bacterial cells against multiple harmful components and play an important role in environmental adaptation and survival in the host [39]. In V. cholerae, Vibrio polysaccharides encoded by vpsA to vpsK and other biofilm-forming proteins encoded by the genes rbmA, rbmC, and bap1 are the main structure components that build a stable biofilm [60,61], while vpsR and vpsT serve as transcriptional activators [62,63]. Another small molecule signal, c-di-GMP, also has a positive effect on biofilm formation through the upregulation of vpsR and vpsT [64]. The intracellular concentration of c-di-GMP can be increased in the presence of bile [65]. The presence of vpsR in NOVC was confirmed by Dua et al. (2018) [66], and variations in VpsR between O1/O139 V. cholerae and NOVC were identified in our previous studies, including point mutation and gene fragment deletions [30]. However, 99% of NOVCs could form stable biofilms in our previous studies, and all the biofilm-relevant genes were present in NOVCs [30,31].

2.6. Type IV Secretion System (T6SS)

Entering the small intestine, the T6SS plays an important role in competition between microorganisms, so that the distribution of commensals in the intestine is altered [47,67]. Therefore, it might act in stage 1 as well as in stage 4. In our previous studies, all NOVCs contained the T6SS [30,31]. In the V. cholerae strain O1 C6706, the T6SS is repressed at low cell density by quorum sensing (QS) molecules [68]. In contrast, N-acetyl glucosamine (GlcNac) can be sensed by the O1 V. cholerae serotype, which leads to tfoX (a major regulator of T6SS) expression followed by T6SS activation [69]. The regulation network of NOVCs is complex and not fully explored, but these regulators might have similar effects in NOVCs [70]. Three regulatory genes, hapR, tfoX, and cytR, achieve their T6SS regulation through the QS- and TfoX-dependent regulator (QstR) [34].

3. Stage 2: Localization and Penetration of the Mucus Layer in the Small Intestine

To cause diarrhea, V. cholerae need to reach the small intestinal epithelial cells to penetrate them. However, the intestinal epithelium is covered by a mucus layer (approximately 150 μm thick), making the ability to penetrate mucus important [32]. Motility is therefore necessary and responsible for a targeted direction [71], while the contribution of chemotaxis remains controversial [72,73]. In contrast to O1/O139 V. cholerae serotypes, whose fitness is supported by genetic features on Vibrio seventh pathogenicity islands 1 and 2 (VSP-1 and VSP-2), the movement of NOVC through the mucosa could be supported by the hemagglutinin protease and neuraminidase, which act as mucinases and are encoded on Vibrio pathogenicity island 2 (VPI-2) [1,74,75]. In addition, environmental NOVCs isolated from food and water sources carry not only the pathogenicity island VPI-2 but also the pathogenicity island VSP-2 in their genome [30,31]. All mechanisms and genes involved in stage 2 are shown in purple in Figure 1.

3.1. Motility via Flagella

The motility-related genes in NOVC were detected through Gene Oncology analysis and the KEGG pathway, which indicate the similar function of motility between NOVC and O1/O139 V. cholerae as these databases are mainly built based on research on O1/O139 V. cholerae [76]. All the motility-associated genes were identified to 100% in NOVC according to our previous studies [30,31]. As a highly motile bacterium, the driving force of V. cholerae is provided by a single polar flagellum. Motility is also functional in host environment adaptation, including nutrient acquisition and toxic component avoidance [77]. The flagellar motility of V. cholerae is important to move the bacteria through the mucus layer [78]. The structure of the flagellum and its four-hierarchy regulatory system was already described by Syed et al. (2009) [79]. The whole flagella system is regulated by sigma factor 54 FlrA, the downstream activator FlrC, and the alternative sigma 28 factor FliA [79]. The motility of V. cholerae also declined due to c-di-GMP. Furthermore, several genes with motility regulation activity were reported. The multifunctional regulation gene csrA could upregulate flrC [80]. Besides, under a high-speed microscope, arcA/cytR and the O-antigen synthesis gene cmd were found to promote motility with an unclear mechanism [81].

3.2. Chemotaxis

The chemotaxis system can recognize chemical signals and regulate the motility and swimming behavior of V. cholerae. At first, a common chemotaxis model of E. coli was identified and subsequently applied in a chemotaxis study of V. cholerae [82] including methyl-accepting chemotaxis proteins (MCPs) encoded by cheW, cheA, cheY, cheR, and cheB. As the chemotaxis system in NOVC has not been explored and the chemotaxis-related genes in O1/O139 V. cholerae were identified in several NOVCs, we suspected that the chemotaxis system in NOVCs might play a similar role. In our previous study, we confirmed the presence of the genes cheA, cheY, and cheR in all analyzed NOVCs, while cheW and cheR were present in 32% of the strains [30]. Transmitted signals can be caught by the cytoplasmic linker protein cheW and transmitted to the two-component system cheA/cheY. Phosphate-activated cheY binds to the flagella motor and causes a reverse rotation direction, from left to right [72]. The genes cheR and cheB play a role in the transfer of methyl groups, which contributes to adaptation to a stable background level of attractants [82]. On the other hand, V. cholerae (both O1/O139 and NOVC) have a far more complex chemotaxis system than E. coli, with 68 related ORFs categorized into three clusters. Among those, cluster II seems to play a similar role in E. coli [83]. Later research reported that cluster I components are assembled into the supramolecular signaling complex in response to reduced cellular energy states, raising the possibility that the cluster I complex plays a role in sensing and signaling under microaerobic environments, such as in the host intestine [84]. The general stress regulator RpoS and autoinducer 1 in quorum sensing could regulate the expression of cluster III [85].

3.3. Vibrio Pathogenicity Island 2 (VPI-2)

VPI-2 (located between vc1758 and vc1809) was identified in O1/O139 V. cholerae and NOVCs. However, Jermyn and Boyd (2005), Haley et al. (2014), and Takahashi et al. (2021) studied the genetic variation of VPI-2 in NOVCs and showed that NOVCs could harbor an incomplete VPI-2 compared to O1/O139 [74,86,87]. This variation might result from the horizontal gene transfer of VPI-2 from the ancestors V. mimicus and O1/O139 V. cholerae [74]. VPI-2 contains the neuraminidase-encoding gene nanH [88] which plays a role in altering mucus structure by cleaving sialic acid groups (GM1 gangliosides) on the epithelial cell surface. Further studies on VPI-2 revealed that nanA, nanE, nanK, and nagA1, which are also localized on VPI-2, can catalyze the metabolism of N-acetylneuraminic acid, which is a component of mucin [6]. All of these functional genes in VPI-2 were identified in 33% of NOVCs isolated from seafood and the environment [30,31].
Vertebrate hosts could limit the zinc level for bacteria as a defense strategy. Zinc deficiency activates Vibrio energy taxis system A (VerA), which is also encoded on the pathogenicity island VSP-2. In addition, VerA could trigger the expression of aerB transcribing a methyl-accepting chemotaxis protein, which could bind cheW and affect the flagellum rotation and motility [89,90].

3.4. Hemagglutinin Protease HapA

The hemagglutinin protease HapA, encoded by hapA, is suggested to be responsible for altering the mucus layer and playing a role in mucus layer penetration during initial infection for both O1/O139 V. cholerae and NOVC [91,92].

4. Stage 3: Intestinal Epithelial Cell Colonization

After the localization of intestinal epithelial cells, V. cholerae must attach to their surface, whereby the type IV pili and the T3SS play a crucial role [93,94]. Subsequently, non-specific adhesins can be secreted via these systems [32]. In contrast to O1 and O139 V. cholerae, the T3SS plays an important role for NOVC in attachment and colonization when TCP is not present in the genome [94]. All mechanisms and genes involved in stage 3 are shown in green in Figure 1.

4.1. Type IV Pili

Type IV pili, encoded by mshA, play a role in the braking and anchoring function of V. cholerae during the landing process on the epithelial cell surface [95]. In addition, MshA pili cause an irreversible attachment and microcolony formation [93]. The presence of mshA in 27% of NOVCs was confirmed in our previous studies [30,31]. At the beginning stage after landing on the epithelial cell surface, several transient non-specific adhesins were secreted to bind the component of small intestine epithelial cells, including multivalent adhesion molecule 7 (mam7, binding with fibronectin and phosphatidic acid), GlcNAc binding protein A (encoded by gbpA), and flagellum-regulated hemagglutinin A (encoded by frhA, binding calcium) [79,96,97]. The presence rates of mam7, gbpA, and frhA in NOVCs were detected as 100%, 94%, and 22%, respectively, in our previous studies [30,31]. Sperandio et al., 1995, stated that a potential adherence factor to epithelial cells could be the outer membrane protein U (OmpU) [98]. This finding is supported by Potapova et al., 2024, who also mentioned that OmpU could also regulate the biofilm matrix assembly [99].

4.2. Type III Secretion System (T3SS)

The T3SS is suggested to have an important role in the colonization of intestinal epithelial cells by NOVCs when TCP is absent. Dziejman et al., 2005, showed using a rabbit and mouse model that the TCP-negative NOVC strain AM-19226 could colonize the intestinal epithelial cell surface through the T3SS [94]. The whole island contains 47 ORFs from A33_1660 to A33_1706. However, an exact mechanism of the T3SS in colonization has not yet been fully identified, although possible functions of several effectors have been addressed: VopF (A33_1696) and VopM (A33_1684) are two effectors in the core region with actin alteration activities, which could disrupt the cell structure and contribute to colonization [100]. VopM can bind F-actin and also plays an important role in colonization by remodeling the intestinal brush border, which facilitates bacterial adhesion [101]. The colonization activity of VopX (A33_1663) in AM-19226 was also stated by Alam et al. [102], and a contradictive result was reported by Chaand et al. [103]. Meanwhile, the T3SS is important for toxicity and toxin transfer; therefore, this part is explained further in stage 4.

5. Stage 4: Virulence Factor Expression

In contrast to the V. cholerae serovars O1 and O139, which express cholera toxin (CTX) and its accessory toxins within the CTX phage, various toxins can be produced by NOVCs after colonization of the small intestine. Currently, four secreted proteins with direct toxic effects shown on cell lines and in animal models have been identified: the hemolysin HlyA, repeats-in-toxin (RTX), heat-stable enterotoxin (ST), and cholix toxin (ChxA) [104,105,106,107]. The expression of these toxins leads to an alteration in the morphology of epithelial cells, cell damage, and subsequently to the death of the cells [35]. Similar to V. cholerae serotypes O1 and O139, whose virulence is supported by genetic features on VPI-1 and VPI-2 [1], VPI-2 was also identified in NOVC environmental isolates from food and water [30,31]. In addition to the toxin genes, the T3SS also plays an important role for NOVC toxicity by secreting virulence factors from the bacteria to the host cells when TCP is missing in the genome [94]. All mechanisms and genes involved in stage 4 are shown in red in Figure 1.

5.1. Toxin Expression

The hemolysin HlyA (also called V. cholerae cytolysin, VCC) could both lyse erythrocytes and form beta barrel pores on epithelial cells [104], followed by cytoskeleton damage, cell lysis, and diarrhea. The iron extracted from the cells in this way serves as a nutrient supplier for NOVCs [108,109]. The transcription of hlyA in V. cholerae is regulated by QS molecules, which regulate hlyU, resulting in the highest transcription of hlyA in the early mid-logarithmic growth phase [110].
Repeats-in-toxin (RTX) is a large protein (around 3500 to 5300 amino acids) widely present in many bacteria which could cause tight junction loss in lung and intestinal epithelial cells [111]. The in vivo toxicity of RTX in hlyA-harboring V. cholerae tends to present as innate immune evasion rather than diarrhea [105,112]. Three major functional units of RTX were found in V. cholerae O1 El Tor N16961. The actin cross-linking domain (ACD) is responsible for cytoskeleton disruption, the Rho GTPase-inactivation domain (RID) for cell rounding, and the alpha/beta hydrolase domain (ABH) for autophagic/endosomal trafficking inhibition. An additional cysteine protease domain is responsible for effectors’ autoprocessing and distribution. The combination of RID and ABH could reduce the inflammatory response caused by ACD [113]. Compared to the El Tor O1 serogroup, NOVCs have more variations in their RTX domain [114]. A nucleotide cluster with five ORFs is responsible for the coding of RTX: the encoding toxin gene rtxA, the activator gene rtxC, and the associated ABC transportation gene cluster rtxBDE [113]. The whole RTX complex was identified in 61% of NOVCs in our previous studies [30,31].
Heat-stable enterotoxin (ST, encoded by stn) is a known toxin in E. coli and was also identified in the genome of NOVC [106,115]. The in vivo toxicity was attributed to fluid accumulation in mouse intestine [106]. The toxin consists of two domains, STa and STb. STa leads to anion secretion and calcium absorption, while STb could decrease the expression of the tight junction proteins ZO-1 and occludin [116].
The cholix toxin ChxA interacts with prohibitin and could therefore cause mitochondrial dysfunction and cytoskeletal remodeling. It is able to bind the lipoprotein receptors of the intestinal epithelial cells and inhibit protein synthesis by ADP-ribosylation. The in vivo toxicity presented as liver damage and final death through mouse assay [107,117]. Tangestani et al., 2020, also confirmed the presence of cholix toxin in NOVC [17].

5.2. Type III Secretion System (T3SS)

The T3SS plays an important role in NOVC after the colonization of intestinal epithelial cells in toxicity. Dziejman et al., 2005, suggested using a mouse model that the T3SS-positive NOVC strain AM-19226 causes mouse death in contrast to a T3SS-negative mutant strain [94]. The protein VopF contains three WASP homology 2 (WH2) actin-binding domains, which could remodel the actin cytoskeleton in eukaryotic host cells. The actin polymerization disorder triggered by VopF is essential for T3SS-mediated intestinal cell damage in AM-19226 [118]. The mechanism might be that VopF could induce cortical actin depolymerization and aberrant localization of the tight junction protein ZO-1, resulting in loosening of the tight junction between intestinal epithelial cells and causing diarrhea [119]. However, Miller et al., 2016, observed that cell death and disruption of the tight junction are independent of VopF [120]. It has been suggested that VopE (A33_1662) impairs mitochondrial dynamics and stimulates the innate immune pathway [121]. Furthermore, the in vivo toxicity of VopE was verified in an infant rabbit and mouse model [122]. The regulators VttRA and VttRB, which show homology with ToxR, can control T3SS activity both during colonization and pathogenesis [123].

Bacteremia Caused by NOVCs

When NOVCs enter and colonize the small intestine (as described in stages 1 to 3), they could enter the bloodstream through the portal vein and the intestinal lymphatic system [8,9]. The immune system, macrophages, and specific antibodies are involved in the blood defense system, indicating that genes for immune modulation are important for NOVCs to cause blood infections [124,125]. Hemolytic properties, such as the presence of HlyA, suggest their ability to enter the bloodstream and lyse erythrocytes [104]. RTX could protect NOVCs from neutrophil-dependent clearance [105]. Biofilms could protect NOVCs from leukocytes of the human immune system. Additionally, NOVCs can form biofilms on the eukaryotic cell surface, causing a concentration of MshA and HapA, which could increase the local hemolysin level lysing immune cells [126].

6. Stage 5: Detachment from the Epithelial Cells

At the end of the infection cycle, NOVCs return to the environment through watery to bloody diarrhea. The symptoms of infection can be closely similar to those of the cholera caused by the serotypes O1 and O139. The starvation/stationary phase alternative sigma factor RpoS positively controls the expression of HapR, a gene involved in flagella assembly and chemotaxis. This enables the detachment and migration of NOVCs from the epithelial cells into the lumen of the intestine [127]. After the activation by RpoS, the hemagglutinin protease HapA encoded by hapA is responsible for detachment from intestinal epithelial cells [75]. Apart from the mucinase activity, HapA could degrade GbpA, the non-specific adhesin in colonization [128]. Furthermore, a set of potential biofilm degradation genes were also identified by Bridges et al., 2020 [129]. These genes include ribosome-associated GTPase encoded by bipA; c-di-GMP phosphodiesterases encoded by cdgG, cdgI, rocS, and mbaA; a polyamine transporter encoded by potD1; a peptidase encoded by lapG; a polysaccharide lyase encoded by rbmB; and a chemotaxis regulator encoded by cheY3. All genes were controlled by the two-component system dbfS/dbfR [129]. All the mentioned mechanisms and genes involved in stage 5 are shown in blue in Figure 1.

7. Multifunctional Regulation System

NOVCs have evolved several regulators to ensure the expression of genes that lead to successful colonization of the intestinal tract. In addition, NOVCs have evolved a number of adaptive mechanisms to adapt to both the environment and the human host as well as to the transition between host and environment [34]. One such multifunctional regulation system is quorum sensing. By cell-to-cell communication, NOVC is able to adjust the cell density. Three QS pathways through different chemical signals have evolved: cholera autoinducer 1 (CAI-1), autoinducer 2 (AI-2), and 3,5-dimethylpyrazin-2-ol (DPO) [130,131,132]. The downstream genes of CAI-1 and AI-2 are cqsS and luxPQ, respectively, followed by luxO, aphA, and hapR (Figure 1) [133]. DPO is the third QS signal mechanism, which can be sensed by vqmA [134], followed by the release of the small molecule vqmR to downregulate rtxA and vpsR [135,136].
Two-component systems are another set of regulators with a wide range of functions. Within the two-component system varS/varA, a receptor for QS and environmental signals represents, together with its downstream gene csrA, a multifunctional regulator in biofilm regulation, iron metabolism, virulence gene expression, and motility [80,137,138]. Another two-component system is vprA/vprB, which is involved in polymyxin and bile resistance [139], which also demonstrated a mutant strain showing colonization failure in host intestine in a mouse model. The gene set vxrA/vxrB could upregulate T6SS expression and biofilm formation [140,141]. The expression of phoR/phoB is activated by phosphate limitation, followed by repression of biofilm-related genes and upregulation of motility [142]. The qseB/qseC gene set is a receptor of the hormones epinephrine and norepinephrine and could affect bacteria motility through triggering pomB expression [143,144]. To adapt to oxygen-poor conditions, acrB/acrA could upregulate toxT and enhance biofilm formation and ROS resistance [145]. chiS is both the monitor and regulator of (ClcNac)2, which is important for intestinal epithelial cell adherence and gut fluid accumulation [146,147].
Two global regulators are histone-like nucleoid structuring protein (HNS) and cyclic AMP-activated global transcriptional regulator (cAMP-CRP). HNS acts as a mediator at the late stage of infection with a repressive effect on a large number of virulence-associated genes such as hemolysin hlyA [148], repeat-in-toxin rtxA [149], Vibrio polysaccharide vps [51], and the T6SS [150], while it could also promote motility and the detachment-dependent protein HapA [151]. The global regulator CRP is the receptor of cAMP, the secondary messenger, and acts as a key regulator of many genes in response to lifestyle changes, including the genes rtxBDE and hlyA [152]. Moreover, CRP represses biofilm formation by repressing the genes vpsR, vpsT, and vpsL and, at the same time, activating the high cell density regulator HapR [153].

8. Schematic Infection Pathway of NOVC

Based on the research mentioned above, a schematic map of virulence-associated genes in NOVC was established and is summarized in Figure 1.

9. Conclusions

As this review article shows, the oral infection of human hosts by pathogenic NOVC is a complex process that depends on the infectivity of the bacterial cells and their ability to survive the harsh conditions in the host until they return to the environment.
It is known that the virulence profile of NOVCs varies, but among them, there are strains expressing all or most of the virulence genes and regulatory systems described in this review article, possibly leading to a pathogenesis ranging from self-limiting diarrheal diseases to cholera-like symptoms and/or bacteremia. Thus, this review article provides an overview of a variety of virulence-associated genes and regulatory systems supporting the understanding of how and why foodborne NOVCs can cause infections.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/microorganisms12040818/s1, Table S1: Additional information of virulence-associated genes in NOVC.

Author Contributions

Conceptualization, T.A. and S.F.; investigation, Q.Z.; visualization, Q.Z.; writing—original draft, Q.Z. and S.F.; writing—review and editing, T.A. and S.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

We thank the China Scholarship Council (CSC) for the scholarship of Q.Z.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Banerjee, R.; Das, B.; Nair, G.B.; Basak, S. Dynamics in genome evolution of Vibrio cholerae. Infect. Genet. Evol. 2014, 23, 32–41. [Google Scholar] [CrossRef] [PubMed]
  2. Faruque, S.M.; Albert, M.J.; Mekalanos, J.J. Epidemiology, genetics, and ecology of toxigenic Vibrio cholerae. Microbiol. Mol. Biol. Rev. 1998, 62, 1301–1314. [Google Scholar] [CrossRef]
  3. WHO. Multi-Country Outbreak of Cholera, External Situation Report n.10; WHO: Geneva, Switzerland, 2024.
  4. Ottaviani, D.; Medici, L.; Talevi, G.; Napoleoni, M.; Serratore, P.; Zavatta, E.; Bignami, G.; Masini, L.; Chierichetti, S.; Fisichella, S. Molecular characterization and drug susceptibility of non-O1/O139 V. cholerae strains of seafood, environmental and clinical origin, Italy. Food Microbiol. 2018, 72, 82–88. [Google Scholar] [CrossRef]
  5. Amato, E.; Riess, M.; Thomas-Lopez, D.; Linkevicius, M.; Pitkänen, T.; Wołkowicz, T.; Rjabinina, J.; Jernberg, C.; Hjertqvist, M.; MacDonald, E. Epidemiological and microbiological investigation of a large increase in vibriosis, northern Europe, 2018. Eurosurveillance 2022, 27, 2101088. [Google Scholar] [CrossRef] [PubMed]
  6. Roux, F.L.; Wegner, K.M.; Baker-Austin, C.; Vezzulli, L.; Osorio, C.R.; Amaro, C.; Ritchie, J.M.; Defoirdt, T.; Destoumieux-Garzón, D.; Blokesch, M. The emergence of Vibrio pathogens in Europe: Ecology, evolution, and pathogenesis (Paris, 11–12th March 2015). Front. Microbiol. 2015, 6, 830. [Google Scholar] [CrossRef] [PubMed]
  7. BfR. Bakterielle Lebensmittelinfektionen durch Vibrionen: Gesundheitliche Bewertung zum Vorkommen von Vibrio spp. (Nicht-Cholera-Vibrionen) in Lebensmitteln: Stellungnahme Nr. 011/2022 des BfR vom 13. April 2022. BfR-Stellungnahmen 2022, 2022, 1–39. [Google Scholar] [CrossRef]
  8. 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, 575. [Google Scholar] [CrossRef]
  9. Schirmeister, F.; Dieckmann, R.; Bechlars, S.; Bier, N.; Faruque, S.; Strauch, E. Genetic and phenotypic analysis of Vibrio cholerae non-O1, non-O139 isolated from German and Austrian patients. Eur. J. Clin. Microbiol. Infect. Dis. 2014, 33, 767–778. [Google Scholar] [CrossRef]
  10. Restrepo, D.; Huprikar, S.S.; VanHorn, K.; Bottone, E.J. O1 and non-O1 Vibrio cholerae bacteremia produced by hemolytic strains. Diagn. Microbiol. Infect. Dis. 2006, 54, 145–148. [Google Scholar] [CrossRef]
  11. Barceló, C.; Sánchez, N.; Martínez, B. Vibrio cholerae non 01 sepsis in a healthy patient: Review of reported cases in Spain. Rev. Clin. Esp. 1998, 198, 850–851. [Google Scholar]
  12. Ottaviani, D.; Leoni, F.; Rocchegiani, E.; Santarelli, S.; Masini, L.; Di Trani, V.; Canonico, C.; Pianetti, A.; Tega, L.; Carraturo, A. Prevalence and virulence properties of non-O1 non-O139 Vibrio cholerae strains from seafood and clinical samples collected in Italy. Int. J. Food Microbiol. 2009, 132, 47–53. [Google Scholar] [CrossRef]
  13. Albuquerque, A.; Cardoso, H.; Pinheiro, D.; Macedo, G. Vibrio cholerae non-O1 and non-O139 bacteremia in a non-traveler Portuguese cirrhotic patient: First case report. Gastroenterol. Y Hepatol. 2013, 36, 309–310. [Google Scholar] [CrossRef]
  14. Dutta, D.; Chowdhury, G.; Pazhani, G.P.; Guin, S.; Dutta, S.; Ghosh, S.; Rajendran, K.; Nandy, R.K.; Mukhopadhyay, A.K.; Bhattacharya, M.K. Vibrio cholerae non-O1, non-O139 serogroups and cholera-like diarrhea, Kolkata, India. Emerg. Infect. Dis. 2013, 19, 464. [Google Scholar] [CrossRef]
  15. Hsu, C.-Y.; Pollett, S.; Ferguson, P.; McMullan, B.J.; Sheppeard, V.; Mahady, S.E. Locally acquired severe non-O1 and non-O139 Vibrio cholerae infection associated with ingestion of imported seafood. Med. J. Aust. 2013, 199, 26–27. [Google Scholar] [CrossRef]
  16. Hasan, N.A.; Rezayat, T.; Blatz, P.J.; Choi, S.Y.; Griffitt, K.J.; Rashed, S.M.; Huq, A.; Conger, N.G.; Colwell, R.R.; Grimes, D.J. Nontoxigenic Vibrio cholerae non-O1/O139 isolate from a case of human gastroenteritis in the US Gulf Coast. J. Clin. Microbiol. 2015, 53, 9–14. [Google Scholar] [CrossRef]
  17. Tangestani, M.G.; Alinezhad, J.; Khajeian, A.; Gharibi, S.; Haghighi, M.A. Identification of cholix toxin gene in Vibrio cholerae non-O1/non-O139 isolated from diarrhea patients in Bushehr, Iran. Iran. J. Microbiol. 2020, 12, 273. [Google Scholar]
  18. Dalsgaard, A.; Forslund, A.; Bodhidatta, L.; Serichantalergs, O.; Pitarangsi, C.; Pang, L.; Shimada, T.; Echeverria, P. A high proportion of Vibrio cholerae strains isolated from children with diarrhoea in Bangkok, Thailand are multiple antibiotic resistant and belong to heterogenous non-O1, non-O139 O-serotypes. Epidemiol. Infect. 1999, 122, 217–226. [Google Scholar] [CrossRef]
  19. Bagchi, K.; Echeverria, P.; Arthur, J.; Sethabutr, O.; Serichantalergs, O.; Hoge, C. Epidemic of diarrhea caused by Vibrio cholerae non-O1 that produced heat-stable toxin among Khmers in a camp in Thailand. J. Clin. Microbiol. 1993, 31, 1315–1317. [Google Scholar] [CrossRef]
  20. Ramamurthy, T.; Bag, P.K.; Pal, A.; Bhattacharya, S.; Bhattacharya, M.; Shimada, T.; Takeda, T.; Karasawa, T.; Kurazono, H.; Takeda, Y. Virulence patterns of Vibrio cholerae non-01 strains isolated from hospitalised patients with acute diarrhoea in Calcutta, India. J. Med. Microbiol. 1993, 39, 310–317. [Google Scholar] [CrossRef]
  21. Rudra, S.; Mahajan, R.; Mathur, M.; Kathuria, K.; Talwar, V. Cluster of cases of clinical cholera due to Vibrio cholerae 010 in east Delhi. Indian J. Med. Res. 1996, 103, 71–73. [Google Scholar]
  22. Sharma, C.; Thungapathra, M.; Ghosh, A.; Mukhopadhyay, A.K.; Basu, A.; Mitra, R.; Basu, I.; Bhattacharya, S.; Shimada, T.; Ramamurthy, T. Molecular analysis of non-O1, non-O139 Vibrio cholerae associated with an unusual upsurge in the incidence of cholera-like disease in Calcutta, India. J. Clin. Microbiol. 1998, 36, 756–763. [Google Scholar] [CrossRef] [PubMed]
  23. Bhattacharya, M.; Dutta, D.; Bhattacharya, S.; Deb, A.; Mukhopadhyay, A.; Nair, G.; Shimada, T.; Takeda, Y.; Chowdhury, A.; Mahalanabis, D. Association of a disease approximating cholera caused by Vibrio cholerae of serogroups other than O1 and O139. Epidemiol. Infect. 1998, 120, 1–5. [Google Scholar] [CrossRef] [PubMed]
  24. Onifade, T.M.; Hutchinson, R.; Van Zile, K.; Bodager, D.; Baker, R.; Blackmore, C. Toxin producing Vibrio cholerae O75 outbreak, United States, march to April 2011. Eurosurveillance 2011, 16, 19870. [Google Scholar] [CrossRef] [PubMed]
  25. Tobin-D’Angelo, M.; Smith, A.R.; Bulens, S.N.; Thomas, S.; Hodel, M.; Izumiya, H.; Arakawa, E.; Morita, M.; Watanabe, H.; Marin, C. Severe diarrhea caused by cholera toxin–producing Vibrio cholerae serogroup O75 infections acquired in the southeastern United States. Clin. Infect. Dis. 2008, 47, 1035–1040. [Google Scholar] [CrossRef] [PubMed]
  26. Luo, Y.; Ye, J.; Jin, D.; Ding, G.; Zhang, Z.; Mei, L.; Octavia, S.; Lan, R. Molecular analysis of non-O1/non-O139 Vibrio cholerae isolated from hospitalised patients in China. BMC Microbiol. 2013, 13, 52. [Google Scholar] [CrossRef] [PubMed]
  27. Ke, B.; Pang, B.; He, D.; Xu, J.; Chen, Q.; Liang, J.; Chen, J.; Li, Z.; Zhou, H.; Deng, X. Phylogenetic analysis of serogroup O5 Vibrio cholerae that caused successive cholera outbreaks—Guangdong Province, China, 2020–2021. China CDC Wkly. 2022, 4, 238. [Google Scholar] [CrossRef] [PubMed]
  28. Arteaga, M.; Velasco, J.; Rodriguez, S.; Vidal, M.; Arellano, C.; Silva, F.; Carreño, L.J.; Vidal, R.; Montero, D.A. Genomic characterization of the non-O1/non-O139 Vibrio cholerae strain that caused a gastroenteritis outbreak in Santiago, Chile, 2018. Microb. Genom. 2020, 6, e000340. [Google Scholar] [CrossRef] [PubMed]
  29. Octavia, S.; Salim, A.; Kurniawan, J.; Lam, C.; Leung, Q.; Ahsan, S.; Reeves, P.R.; Nair, G.B.; Lan, R. Population structure and evolution of non-O1/non-O139 Vibrio cholerae by multilocus sequence typing. PLoS ONE 2013, 8, e65342. [Google Scholar] [CrossRef]
  30. Zhang, Q.; Alter, T.; Strauch, E.; Hammerl, J.A.; Schwartz, K.; Borowiak, M.; Deneke, C.; Fleischmann, S. Genetic and Phenotypic Virulence Potential of Non-O1/Non-O139 Vibrio cholerae Isolated from German Retail Seafood. Microorganisms 2023, 11, 2751. [Google Scholar] [CrossRef]
  31. Zhang, Q.; Alter, T.; Strauch, E.; Eichhorn, I.; Borowiak, M.; Deneke, C.; Fleischmann, S. German coasts harbor non-O1/non-O139 Vibrio cholerae with clinical virulence gene profiles. Infect. Genet. Evol. 2024, 120, 105587. [Google Scholar] [CrossRef]
  32. Almagro-Moreno, S.; Pruss, K.; Taylor, R.K. Intestinal colonization dynamics of Vibrio cholerae. PLoS Pathog. 2015, 11, e1004787. [Google Scholar] [CrossRef] [PubMed]
  33. Jones, G.W.; Isaacson, R.E. Proteinaceous bacterial adhesins and their receptors. CRC Crit. Rev. Microbiol. 1982, 10, 229–260. [Google Scholar] [CrossRef] [PubMed]
  34. Lo Scrudato, M.; Blokesch, M. The regulatory network of natural competence and transformation of Vibrio cholerae. PLoS Genet. 2012, 8, e1002778. [Google Scholar] [CrossRef] [PubMed]
  35. Ramamurthy, T.; Nandy, R.K.; Mukhopadhyay, A.K.; Dutta, S.; Mutreja, A.; Okamoto, K.; Miyoshi, S.-I.; Nair, G.B.; Ghosh, A. Virulence regulation and innate host response in the pathogenicity of Vibrio cholerae. Front. Cell. Infect. Microbiol. 2020, 10, 572096. [Google Scholar] [CrossRef] [PubMed]
  36. Montero, D.A.; Vidal, R.M.; Velasco, J.; George, S.; Lucero, Y.; Gómez, L.A.; Carreño, L.J.; García-Betancourt, R.; O’Ryan, M. Vibrio cholerae, classification, pathogenesis, immune response, and trends in vaccine development. Front. Med. 2023, 10, 1155751. [Google Scholar] [CrossRef] [PubMed]
  37. Kanehisa, M.; Sato, Y.; Kawashima, M. KEGG mapping tools for uncovering hidden features in biological data. Protein Sci. 2022, 31, 47–53. [Google Scholar] [CrossRef] [PubMed]
  38. Silva, A.J.; Benitez, J.A. Vibrio cholerae biofilms and cholera pathogenesis. PLoS Neglected Trop. Dis. 2016, 10, e0004330. [Google Scholar] [CrossRef]
  39. Teschler, J.K.; Zamorano-Sánchez, D.; Utada, A.S.; Warner, C.J.; Wong, G.C.; Linington, R.G.; Yildiz, F.H. Living in the matrix: Assembly and control of Vibrio cholerae biofilms. Nat. Rev. Microbiol. 2015, 13, 255–268. [Google Scholar] [CrossRef]
  40. Merrell, D.S.; Hava, D.L.; Camilli, A. Identification of novel factors involved in colonization and acid tolerance of Vibrio cholerae. Mol. Microbiol. 2002, 43, 1471–1491. [Google Scholar] [CrossRef]
  41. Zhou, Y.; Lee, Z.L.; Zhu, J. On or Off: Life-Changing Decisions Made by Vibrio cholerae under Stress. Infect. Microbes Dis. 2020, 2, 127. [Google Scholar] [CrossRef]
  42. Merrell, D.S.; Camilli, A. The cadA gene of Vibrio cholerae is induced during infection and plays a role in acid tolerance. Mol. Microbiol. 1999, 34, 836–849. [Google Scholar] [CrossRef] [PubMed]
  43. Stern, A.M.; Hay, A.J.; Liu, Z.; Desland, F.A.; Zhang, J.; Zhong, Z.; Zhu, J. The NorR regulon is critical for Vibrio cholerae resistance to nitric oxide and sustained colonization of the intestines. MBio 2012, 3, e00013-12. [Google Scholar] [CrossRef] [PubMed]
  44. Frey, R.S.; Ushio–Fukai, M.; Malik, A.B. NADPH oxidase-dependent signaling in endothelial cells: Role in physiology and pathophysiology. Antioxid. Redox Signal. 2009, 11, 791–810. [Google Scholar] [CrossRef] [PubMed]
  45. Provenzano, D.; Klose, K.E. Altered expression of the ToxR-regulated porins OmpU and OmpT diminishes Vibrio cholerae bile resistance, virulence factor expression, and intestinal colonization. Proc. Natl. Acad. Sci. USA 2000, 97, 10220–10224. [Google Scholar] [CrossRef] [PubMed]
  46. Bina, X.R.; Provenzano, D.; Nguyen, N.; Bina, J.E. Vibrio cholerae RND family efflux systems are required for antimicrobial resistance, optimal virulence factor production, and colonization of the infant mouse small intestine. Infect. Immun. 2008, 76, 3595–3605. [Google Scholar] [CrossRef] [PubMed]
  47. Zhao, W.; Caro, F.; Robins, W.; Mekalanos, J.J. Antagonism toward the intestinal microbiota and its effect on Vibrio cholerae virulence. Science 2018, 359, 210–213. [Google Scholar] [CrossRef]
  48. Kovacikova, G.; Lin, W.; Skorupski, K. The LysR-type virulence activator AphB regulates the expression of genes in Vibrio cholerae in response to low pH and anaerobiosis. J. Bacteriol. 2010, 192, 4181–4191. [Google Scholar] [CrossRef] [PubMed]
  49. Cakar, F.; Zingl, F.G.; Moisi, M.; Reidl, J.; Schild, S. In vivo repressed genes of Vibrio cholerae reveal inverse requirements of an H+/Cl− transporter along the gastrointestinal passage. Proc. Natl. Acad. Sci. USA 2018, 115, E2376–E2385. [Google Scholar] [CrossRef]
  50. Fernandez, N.L.; Waters, C.M. Cyclic di-GMP increases catalase production and hydrogen peroxide tolerance in Vibrio cholerae. Appl. Environ. Microbiol. 2019, 85, e01043-19. [Google Scholar] [CrossRef]
  51. Wang, H.; Ayala, J.C.; Silva, A.J.; Benitez, J.A. The histone-like nucleoid structuring protein (H-NS) is a repressor of Vibrio cholerae exopolysaccharide biosynthesis (vps) genes. Appl. Environ. Microbiol. 2012, 78, 2482–2488. [Google Scholar] [CrossRef]
  52. Liu, Z.; Wang, H.; Zhou, Z.; Sheng, Y.; Naseer, N.; Kan, B.; Zhu, J. Thiol-based switch mechanism of virulence regulator AphB modulates oxidative stress response in Vibrio cholerae. Mol. Microbiol. 2016, 102, 939–949. [Google Scholar] [CrossRef] [PubMed]
  53. Cha, M.-K.; Hong, S.-K.; Lee, D.-S.; Kim, I.-H. Vibrio cholerae thiol peroxidase-glutaredoxin fusion is a 2-Cys TSA/AhpC subfamily acting as a lipid hydroperoxide reductase. J. Biol. Chem. 2004, 279, 11035–11041. [Google Scholar] [CrossRef]
  54. Bina, X.R.; Howard, M.F.; Taylor-Mulneix, D.L.; Ante, V.M.; Kunkle, D.E.; Bina, J.E. The Vibrio cholerae RND efflux systems impact virulence factor production and adaptive responses via periplasmic sensor proteins. PLoS Pathog. 2018, 14, e1006804. [Google Scholar] [CrossRef] [PubMed]
  55. Mukherjee, M.; Kakarla, P.; Kumar, S.; Gonzalez, E.; Floyd, J.T.; Inupakutika, M.; Devireddy, A.R.; Tirrell, S.R.; Bruns, M.; He, G. Comparative genome analysis of non-toxigenic non-O1 versus toxigenic O1 Vibrio cholerae. Genom. Discov. 2014, 2, 1. [Google Scholar] [CrossRef] [PubMed]
  56. Kunkle, D.E.; Bina, X.R.; Bina, J.E. The Vibrio cholerae VexGH RND efflux system maintains cellular homeostasis by effluxing vibriobactin. MBio 2017, 8, e00126-17. [Google Scholar] [CrossRef] [PubMed]
  57. Singh, D.; Matte, M.H.; Matté, G.R.; Jiang, S.; Sabeena, F.; Shukla, B.; Sanyal, S.; Huq, A.; Colwell, R. Molecular analysis of Vibrio cholerae O1, O139, non-O1, and non-O139 strains: Clonal relationships between clinical and environmental isolates. Appl. Environ. Microbiol. 2001, 67, 910–921. [Google Scholar] [CrossRef] [PubMed]
  58. Jiang, F.; Bi, R.; Deng, L.; Kang, H.; Gu, B.; Ma, P. Virulence-associated genes and molecular characteristics of non-O1/non-O139 Vibrio cholerae isolated from hepatitis B cirrhosis patients in China. Int. J. Infect. Dis. 2018, 74, 117–122. [Google Scholar] [CrossRef]
  59. Pennetzdorfer, N.; Höfler, T.; Wölflingseder, M.; Tutz, S.; Schild, S.; Reidl, J. σE controlled regulation of porin OmpU in Vibrio cholerae. Mol. Microbiol. 2021, 115, 1244–1261. [Google Scholar] [CrossRef] [PubMed]
  60. Giglio, K.M.; Fong, J.C.; Yildiz, F.H.; Sondermann, H. Structural basis for biofilm formation via the Vibrio cholerae matrix protein RbmA. J. Bacteriol. 2013, 195, 3277–3286. [Google Scholar] [CrossRef]
  61. Fong, J.C.; Yildiz, F.H. The rbmBCDEF gene cluster modulates development of rugose colony morphology and biofilm formation in Vibrio cholerae. J. Bacteriol. 2007, 189, 2319–2330. [Google Scholar] [CrossRef]
  62. Zamorano-Sánchez, D.; Fong, J.C.; Kilic, S.; Erill, I.; Yildiz, F.H. Identification and characterization of VpsR and VpsT binding sites in Vibrio cholerae. J. Bacteriol. 2015, 197, 1221–1235. [Google Scholar] [CrossRef] [PubMed]
  63. Beyhan, S.; Bilecen, K.; Salama, S.R.; Casper-Lindley, C.; Yildiz, F.H. Regulation of rugosity and biofilm formation in Vibrio cholerae: Comparison of VpsT and VpsR regulons and epistasis analysis of vpsT, vpsR, and hapR. J. Bacteriol. 2007, 189, 388–402. [Google Scholar] [CrossRef] [PubMed]
  64. Krasteva, P.V.; Fong, J.C.; Shikuma, N.J.; Beyhan, S.; Navarro, M.V.; Yildiz, F.H.; Sondermann, H. Vibrio cholerae VpsT regulates matrix production and motility by directly sensing cyclic di-GMP. Science 2010, 327, 866–868. [Google Scholar] [CrossRef] [PubMed]
  65. Koestler, B.J.; Waters, C.M. Bile acids and bicarbonate inversely regulate intracellular cyclic di-GMP in Vibrio cholerae. Infect. Immun. 2014, 82, 3002–3014. [Google Scholar] [CrossRef] [PubMed]
  66. Dua, P.; Karmakar, A.; Ghosh, C. Virulence gene profiles, biofilm formation, and antimicrobial resistance of Vibrio cholerae non-O1/non-O139 bacteria isolated from West Bengal, India. Heliyon 2018, 4, e01040. [Google Scholar] [CrossRef] [PubMed]
  67. Logan, S.L.; Thomas, J.; Yan, J.; Baker, R.P.; Shields, D.S.; Xavier, J.B.; Hammer, B.K.; Parthasarathy, R. The Vibrio cholerae type VI secretion system can modulate host intestinal mechanics to displace gut bacterial symbionts. Proc. Natl. Acad. Sci. USA 2018, 115, E3779–E3787. [Google Scholar] [CrossRef] [PubMed]
  68. Jaskolska, M.; Stutzmann, S.; Stoudmann, C.; Blokesch, M. QstR-dependent regulation of natural competence and type VI secretion in Vibrio cholerae. Nucleic Acids Res. 2018, 46, 10619–10634. [Google Scholar] [CrossRef] [PubMed]
  69. Borgeaud, S.; Metzger, L.C.; Scrignari, T.; Blokesch, M. The type VI secretion system of Vibrio cholerae fosters horizontal gene transfer. Science 2015, 347, 63–67. [Google Scholar] [CrossRef] [PubMed]
  70. Crisan, C.V.; Hammer, B.K. The Vibrio cholerae type VI secretion system: Toxins, regulators and consequences. Environ. Microbiol. 2020, 22, 4112–4122. [Google Scholar] [CrossRef]
  71. Krukonis, E.S.; DiRita, V.J. From motility to virulence: Sensing and responding to environmental signals in Vibrio cholerae. Curr. Opin. Microbiol. 2003, 6, 186–190. [Google Scholar] [CrossRef]
  72. Butler, S.M.; Camilli, A. Both chemotaxis and net motility greatly influence the infectivity of Vibrio cholerae. Proc. Natl. Acad. Sci. USA 2004, 101, 5018–5023. [Google Scholar] [CrossRef]
  73. Millet, Y.A.; Alvarez, D.; Ringgaard, S.; von Andrian, U.H.; Davis, B.M.; Waldor, M.K. Insights into Vibrio cholerae intestinal colonization from monitoring fluorescently labeled bacteria. PLoS Pathog. 2014, 10, e1004405. [Google Scholar] [CrossRef]
  74. Jermyn, W.S.; Boyd, E.F. Molecular evolution of Vibrio pathogenicity island-2 (VPI-2): Mosaic structure among Vibrio cholerae and Vibrio mimicus natural isolates. Microbiology 2005, 151, 311–322. [Google Scholar] [CrossRef]
  75. Finkelstein, R.A.; Boesman-Finkelstein, M.; Chang, Y.; Häse, C. Vibrio cholerae hemagglutinin/protease, colonial variation, virulence, and detachment. Infect. Immun. 1992, 60, 472–478. [Google Scholar] [CrossRef]
  76. Zhou, Y.; Gu, S.; Li, J.; Ji, P.; Zhang, Y.; Wu, C.; Jiang, Q.; Gao, X.; Zhang, X. Complete genome analysis of highly pathogenic non-O1/O139 Vibrio cholerae isolated from Macrobrachium rosenbergii reveals pathogenicity and antibiotic resistance-related genes. Front. Vet. Sci. 2022, 9, 882885. [Google Scholar] [CrossRef]
  77. Ottemann, K.M.; Miller, J.F. Roles for motility in bacterial–host interactions. Mol. Microbiol. 1997, 24, 1109–1117. [Google Scholar] [CrossRef]
  78. Nhu, N.T.; Lee, J.S.; Wang, H.J.; Dufour, Y.S. Alkaline pH increases swimming speed and facilitates mucus penetration for Vibrio cholerae. J. Bacteriol. 2021, 203, e00607-20. [Google Scholar] [CrossRef]
  79. Syed, K.A.; Beyhan, S.; Correa, N.; Queen, J.; Liu, J.; Peng, F.; Satchell, K.J.; Yildiz, F.; Klose, K.E. The Vibrio cholerae flagellar regulatory hierarchy controls expression of virulence factors. J. Bacteriol. 2009, 191, 6555–6570. [Google Scholar] [CrossRef]
  80. Butz, H.A.; Mey, A.R.; Ciosek, A.L.; Crofts, A.A.; Davies, B.W.; Payne, S.M. Regulatory effects of CsrA in Vibrio cholerae. MBio 2021, 12, e03380-20. [Google Scholar] [CrossRef] [PubMed]
  81. Li, Y.; Yan, J.; Guo, X.; Wang, X.; Liu, F.; Cao, B. The global regulators ArcA and CytR collaboratively modulate Vibrio cholerae motility. BMC Microbiol. 2022, 22, 22. [Google Scholar] [CrossRef] [PubMed]
  82. Boin, M.A.; Austin, M.J.; Häse, C.C. Chemotaxis in Vibrio cholerae. FEMS Microbiol. Lett. 2004, 239, 1–8. [Google Scholar] [CrossRef]
  83. Yang, W.; Alvarado, A.; Glatter, T.; Ringgaard, S.; Briegel, A. Baseplate variability of Vibrio cholerae chemoreceptor arrays. Proc. Natl. Acad. Sci. USA 2018, 115, 13365–13370. [Google Scholar] [CrossRef]
  84. Hiremath, G.; Hyakutake, A.; Yamamoto, K.; Ebisawa, T.; Nakamura, T.; Nishiyama, S.i.; Homma, M.; Kawagishi, I. Hypoxia-induced localization of chemotaxis-related signaling proteins in V ibrio cholerae. Mol. Microbiol. 2015, 95, 780–790. [Google Scholar] [CrossRef]
  85. Ringgaard, S.; Hubbard, T.; Mandlik, A.; Davis, B.M.; Waldor, M.K. RpoS and quorum sensing control expression and polar localization of V ibrio cholerae chemotaxis cluster III proteins in vitro and in vivo. Mol. Microbiol. 2015, 97, 660–675. [Google Scholar] [CrossRef]
  86. Haley, B.J.; Choi, S.Y.; Grim, C.J.; Onifade, T.J.; Cinar, H.N.; Tall, B.D.; Taviani, E.; Hasan, N.A.; Abdullah, A.H.; Carter, L. Genomic and phenotypic characterization of Vibrio cholerae non-O1 isolates from a US Gulf Coast cholera outbreak. PLoS ONE 2014, 9, e86264. [Google Scholar] [CrossRef]
  87. Takahashi, E.; Ochi, S.; Morita, D.; Morita, M.; Ohnishi, M.; Koley, H.; Dutta, M.; Chowdhury, G.; Mukhopadhyay, A.K.; Dutta, S. Virulence of cholera toxin gene-positive Vibrio cholerae non-O1/non-O139 strains isolated from environmental water in Kolkata, India. Front. Microbiol. 2021, 12, 726273. [Google Scholar] [CrossRef]
  88. Galen, J.E.; Ketley, J.; Fasano, A.; Richardson, S.; Wasserman, S.; Kaper, J. Role of Vibrio cholerae neuraminidase in the function of cholera toxin. Infect. Immun. 1992, 60, 406–415. [Google Scholar] [CrossRef]
  89. Murphy, S.G.; Alvarez, L.; Adams, M.C.; Liu, S.; Chappie, J.S.; Cava, F.; Dörr, T. Endopeptidase regulation as a novel function of the Zur-dependent zinc starvation response. MBio 2019, 10, e02620-18. [Google Scholar] [CrossRef]
  90. Murphy, S.G.; Murtha, A.N.; Zhao, Z.; Alvarez, L.; Diebold, P.; Shin, J.-H.; VanNieuwenhze, M.S.; Cava, F.; Dörr, T. Class A penicillin-binding protein-mediated cell wall synthesis promotes structural integrity during peptidoglycan endopeptidase insufficiency in Vibrio cholerae. MBio 2021, 12, e03596-20. [Google Scholar] [CrossRef]
  91. Silva, A.J.; Pham, K.; Benitez, J.A. Haemagglutinin/protease expression and mucin gel penetration in El Tor biotype Vibrio cholerae. Microbiology 2003, 149, 1883–1891. [Google Scholar] [CrossRef]
  92. Ceccarelli, D.; Chen, A.; Hasan, N.A.; Rashed, S.M.; Huq, A.; Colwell, R.R. Non-O1/non-O139 Vibrio cholerae carrying multiple virulence factors and V. cholerae O1 in the Chesapeake Bay, Maryland. Appl. Environ. Microbiol. 2015, 81, 1909–1918. [Google Scholar] [CrossRef]
  93. Utada, A.S.; Bennett, R.R.; Fong, J.C.; Gibiansky, M.L.; Yildiz, F.H.; Golestanian, R.; Wong, G.C. Vibrio cholerae use pili and flagella synergistically to effect motility switching and conditional surface attachment. Nat. Commun. 2014, 5, 4913. [Google Scholar] [CrossRef]
  94. Dziejman, M.; Serruto, D.; Tam, V.C.; Sturtevant, D.; Diraphat, P.; Faruque, S.M.; Rahman, M.H.; Heidelberg, J.F.; Decker, J.; Li, L. Genomic characterization of non-O1, non-O139 Vibrio cholerae reveals genes for a type III secretion system. Proc. Natl. Acad. Sci. USA 2005, 102, 3465–3470. [Google Scholar] [CrossRef]
  95. Zhang, W.; Luo, M.; Feng, C.; Liu, H.; Zhang, H.; Bennett, R.R.; Utada, A.S.; Liu, Z.; Zhao, K. Crash landing of Vibrio cholerae by MSHA pili-assisted braking and anchoring in a viscoelastic environment. Elife 2021, 10, e60655. [Google Scholar] [CrossRef]
  96. Krachler, A.M.; Ham, H.; Orth, K. Outer membrane adhesion factor multivalent adhesion molecule 7 initiates host cell binding during infection by gram-negative pathogens. Proc. Natl. Acad. Sci. USA 2011, 108, 11614–11619. [Google Scholar] [CrossRef]
  97. Wong, E.; Vaaje-Kolstad, G.; Ghosh, A.; Hurtado-Guerrero, R.; Konarev, P.V.; Ibrahim, A.F.; Svergun, D.I.; Eijsink, V.G.; Chatterjee, N.S.; van Aalten, D.M. The Vibrio cholerae colonization factor GbpA possesses a modular structure that governs binding to different host surfaces. PLoS Pathog. 2012, 8, e1002373. [Google Scholar] [CrossRef]
  98. Sperandio, V.; Giron, J.A.; Silveira, W.D.; Kaper, J.B. The OmpU outer membrane protein, a potential adherence factor of Vibrio cholerae. Infect. Immun. 1995, 63, 4433–4438. [Google Scholar] [CrossRef]
  99. Potapova, A.; Garvey, W.; Dahl, P.; Guo, S.; Chang, Y.; Schwechheimer, C.; Trebino, M.A.; Floyd, K.A.; Phinney, B.S.; Liu, J. Outer membrane vesicles and the outer membrane protein OmpU govern Vibrio cholerae biofilm matrix assembly. Mbio 2024, 15, e03304-23. [Google Scholar] [CrossRef]
  100. Miller, K.A.; Tomberlin, K.F.; Dziejman, M. Vibrio variations on a type three theme. Curr. Opin. Microbiol. 2019, 47, 66–73. [Google Scholar] [CrossRef]
  101. Zhou, X.; Massol, R.H.; Nakamura, F.; Chen, X.; Gewurz, B.E.; Davis, B.M.; Lencer, W.I.; Waldor, M.K. Remodeling of the intestinal brush border underlies adhesion and virulence of an enteric pathogen. MBio 2014, 5, e01639-14. [Google Scholar] [CrossRef]
  102. Alam, A.; Miller, K.A.; Chaand, M.; Butler, J.S.; Dziejman, M. Identification of Vibrio cholerae type III secretion system effector proteins. Infect. Immun. 2011, 79, 1728–1740. [Google Scholar] [CrossRef]
  103. Chaand, M.; Miller, K.A.; Sofia, M.K.; Schlesener, C.; Weaver, J.W.; Sood, V.; Dziejman, M. Type three secretion system island-encoded proteins required for colonization by non-O1/non-O139 serogroup Vibrio cholerae. Infect. Immun. 2015, 83, 2862–2869. [Google Scholar] [CrossRef]
  104. Krasilnikov, O.V.; Muratkhodjaev, J.N.; Zitzer, A.O. The mode of action of Vibrio cholerae cytolysin. The influences on both erythrocytes and planar lipid bilayers. Biochim. Et Biophys. Acta (BBA)-Biomembr. 1992, 1111, 7–16. [Google Scholar] [CrossRef]
  105. Queen, J.; Satchell, K.J.F. Neutrophils are essential for containment of Vibrio cholerae to the intestine during the proinflammatory phase of infection. Infect. Immun. 2012, 80, 2905–2913. [Google Scholar] [CrossRef]
  106. Arita, M.; Takeda, T.; Honda, T.; Miwatani, T. Purification and characterization of Vibrio cholerae non-O1 heat-stable enterotoxin. Infect. Immun. 1986, 52, 45–49. [Google Scholar] [CrossRef]
  107. Jørgensen, R.; Purdy, A.E.; Fieldhouse, R.J.; Kimber, M.S.; Bartlett, D.H.; Merrill, A.R. Cholix toxin, a novel ADP-ribosylating factor from Vibrio cholerae. J. Biol. Chem. 2008, 283, 10671–10678. [Google Scholar] [CrossRef]
  108. Saka, H.A.; Bidinost, C.; Sola, C.; Carranza, P.; Collino, C.; Ortiz, S.; Echenique, J.R.; Bocco, J.L. Vibrio cholerae cytolysin is essential for high enterotoxicity and apoptosis induction produced by a cholera toxin gene-negative V. cholerae non-O1, non-O139 strain. Microb. Pathog. 2008, 44, 118–128. [Google Scholar] [CrossRef]
  109. Olson, R.; Gouaux, E. Crystal structure of the Vibrio cholerae cytolysin (VCC) pro-toxin and its assembly into a heptameric transmembrane pore. J. Mol. Biol. 2005, 350, 997–1016. [Google Scholar] [CrossRef]
  110. Gao, H.; Xu, J.; Lu, X.; Li, J.; Lou, J.; Zhao, H.; Diao, B.; Shi, Q.; Zhang, Y.; Kan, B. Expression of hemolysin is regulated under the collective actions of HapR, Fur, and HlyU in Vibrio cholerae El Tor serogroup O1. Front. Microbiol. 2018, 9, 1310. [Google Scholar] [CrossRef] [PubMed]
  111. Satchell, K.J.F. MARTX, multifunctional autoprocessing repeats-in-toxin toxins. Infect. Immun. 2007, 75, 5079–5084. [Google Scholar] [CrossRef] [PubMed]
  112. Olivier, V.; Queen, J.; Satchell, K.J. Successful small intestine colonization of adult mice by Vibrio cholerae requires ketamine anesthesia and accessory toxins. PLoS ONE 2009, 4, e7352. [Google Scholar] [CrossRef]
  113. Satchell, K.J. Multifunctional-autoprocessing repeats-in-toxin (MARTX) Toxins of Vibrios. Microbiol. Spectr. 2015, 3. [Google Scholar] [CrossRef]
  114. Dolores, J.; Satchell, K.J. Analysis of Vibrio cholerae genome sequences reveals unique rtxA variants in environmental strains and an rtxA-null mutation in recent altered El Tor isolates. MBio 2013, 4, e00624. [Google Scholar] [CrossRef]
  115. Spangler, B.D. Structure and function of cholera toxin and the related Escherichia coli heat-labile enterotoxin. Microbiol. Rev. 1992, 56, 622–647. [Google Scholar] [CrossRef]
  116. Peterson, J.W.; Whipp, S.C. Comparison of the mechanisms of action of cholera toxin and the heat-stable enterotoxins of Escherichia coli. Infect. Immun. 1995, 63, 1452–1461. [Google Scholar] [CrossRef]
  117. Awasthi, S.P.; Asakura, M.; Chowdhury, N.; Neogi, S.B.; Hinenoya, A.; Golbar, H.M.; Yamate, J.; Arakawa, E.; Tada, T.; Ramamurthy, T. Novel cholix toxin variants, ADP-ribosylating toxins in Vibrio cholerae non-O1/non-O139 strains, and their pathogenicity. Infect. Immun. 2013, 81, 531–541. [Google Scholar] [CrossRef]
  118. Tam, V.C.; Serruto, D.; Dziejman, M.; Brieher, W.; Mekalanos, J.J. A type III secretion system in Vibrio cholerae translocates a formin/spire hybrid-like actin nucleator to promote intestinal colonization. Cell Host Microbe 2007, 1, 95–107. [Google Scholar] [CrossRef]
  119. Tam, V.C.; Suzuki, M.; Coughlin, M.; Saslowsky, D.; Biswas, K.; Lencer, W.I.; Faruque, S.M.; Mekalanos, J.J. Functional analysis of VopF activity required for colonization in Vibrio cholerae. MBio 2010, 1, e00289-10. [Google Scholar] [CrossRef]
  120. Miller, K.A.; Chaand, M.; Gregoire, S.; Yoshida, T.; Beck, L.A.; Ivanov, A.I.; Dziejman, M. Characterization of V. cholerae T3SS-dependent cytotoxicity in cultured intestinal epithelial cells. Cell Microbiol. 2016, 18, 1857–1870. [Google Scholar] [CrossRef] [PubMed]
  121. Suzuki, M.; Danilchanka, O.; Mekalanos, J.J. Vibrio cholerae T3SS effector VopE modulates mitochondrial dynamics and innate immune signaling by targeting Miro GTPases. Cell Host Microbe 2014, 16, 581–591. [Google Scholar] [CrossRef] [PubMed]
  122. Shin, O.S.; Tam, V.C.; Suzuki, M.; Ritchie, J.M.; Bronson, R.T.; Waldor, M.K.; Mekalanos, J.J. Type III secretion is essential for the rapidly fatal diarrheal disease caused by non-O1, non-O139 Vibrio cholerae. MBio 2011, 2, e00106-11. [Google Scholar] [CrossRef]
  123. Chaand, M.; Dziejman, M. Vibrio cholerae VttRA and VttRB regulatory influences extend beyond the type 3 secretion system genomic island. J. Bacteriol. 2013, 195, 2424–2436. [Google Scholar] [CrossRef]
  124. Kashimura, M. The human spleen as the center of the blood defense system. Int. J. Hematol. 2020, 112, 147–158. [Google Scholar] [CrossRef]
  125. Kwiecinski, J.M.; Horswill, A.R. Staphylococcus aureus bloodstream infections: Pathogenesis and regulatory mechanisms. Curr. Opin. Microbiol. 2020, 53, 51–60. [Google Scholar] [CrossRef]
  126. Vidakovic, L.; Mikhaleva, S.; Jeckel, H.; Nisnevich, V.; Strenger, K.; Neuhaus, K.; Raveendran, K.; Ben-Moshe, N.B.; Aznaourova, M.; Nosho, K. Biofilm formation on human immune cells is a multicellular predation strategy of Vibrio cholerae. Cell 2023, 186, 2690–2704.e1–e10. [Google Scholar] [CrossRef]
  127. Nielsen, A.T.; Dolganov, N.A.; Otto, G.; Miller, M.C.; Wu, C.Y.; Schoolnik, G.K. RpoS controls the Vibrio cholerae mucosal escape response. PLoS Pathog. 2006, 2, e109. [Google Scholar] [CrossRef]
  128. Jude, B.A.; Martinez, R.M.; Skorupski, K.; Taylor, R.K. Levels of the secreted Vibrio cholerae attachment factor GbpA are modulated by quorum-sensing-induced proteolysis. J. Bacteriol. 2009, 191, 6911–6917. [Google Scholar] [CrossRef]
  129. Bridges, A.A.; Fei, C.; Bassler, B.L. Identification of signaling pathways, matrix-digestion enzymes, and motility components controlling Vibrio cholerae biofilm dispersal. Proc. Natl. Acad. Sci. USA 2020, 117, 32639–32647. [Google Scholar] [CrossRef]
  130. Perez, L.J.; Ng, W.-L.; Marano, P.; Brook, K.; Bassler, B.L.; Semmelhack, M.F. Role of the CAI-1 fatty acid tail in the Vibrio cholerae quorum sensing response. J. Med. Chem. 2012, 55, 9669–9681. [Google Scholar] [CrossRef]
  131. Papenfort, K.; Silpe, J.E.; Schramma, K.R.; Cong, J.-P.; Seyedsayamdost, M.R.; Bassler, B.L. A Vibrio cholerae autoinducer–receptor pair that controls biofilm formation. Nat. Chem. Biol. 2017, 13, 551–557. [Google Scholar] [CrossRef]
  132. Winzer, K.; Hardie, K.; Williams, P. LuxS and autoinducer-2: Their contribution to quorum. Adv. Appl. Microbiol. 2003, 53, 291. [Google Scholar]
  133. Higgins, D.A.; Pomianek, M.E.; Kraml, C.M.; Taylor, R.K.; Semmelhack, M.F.; Bassler, B.L. The major Vibrio cholerae autoinducer and its role in virulence factor production. Nature 2007, 450, 883–886. [Google Scholar] [CrossRef]
  134. Huang, X.; Duddy, O.P.; Silpe, J.E.; Paczkowski, J.E.; Cong, J.; Henke, B.R.; Bassler, B.L. Mechanism underlying autoinducer recognition in the Vibrio cholerae DPO-VqmA quorum-sensing pathway. J. Biol. Chem. 2020, 295, 2916–2931. [Google Scholar] [CrossRef]
  135. Papenfort, K.; Förstner, K.U.; Cong, J.-P.; Sharma, C.M.; Bassler, B.L. Differential RNA-seq of Vibrio cholerae identifies the VqmR small RNA as a regulator of biofilm formation. Proc. Natl. Acad. Sci. USA 2015, 112, E766–E775. [Google Scholar] [CrossRef]
  136. Defoirdt, T. Amino acid–derived quorum sensing molecules controlling the virulence of vibrios (and beyond). PLoS Pathog. 2019, 15, e1007815. [Google Scholar] [CrossRef]
  137. Butz, H.A.; Mey, A.R.; Ciosek, A.L.; Payne, S.M. Vibrio cholerae CsrA directly regulates varA to increase expression of the three nonredundant Csr small RNAs. MBio 2019, 10, e01042-19. [Google Scholar] [CrossRef]
  138. Jang, J.; Jung, K.-T.; Yoo, C.-K.; Rhie, G.-e. Regulation of hemagglutinin/protease expression by the VarS/VarA–CsrA/B/C/D system in Vibrio cholerae. Microb. Pathog. 2010, 48, 245–250. [Google Scholar] [CrossRef]
  139. Herrera, C.M.; Crofts, A.A.; Henderson, J.C.; Pingali, S.C.; Davies, B.W.; Trent, M.S. The Vibrio cholerae VprA-VprB two-component system controls virulence through endotoxin modification. MBio 2014, 5, e02283-14. [Google Scholar] [CrossRef]
  140. Teschler, J.K.; Cheng, A.T.; Yildiz, F.H. The two-component signal transduction system VxrAB positively regulates Vibrio cholerae biofilm formation. J. Bacteriol. 2017, 199, e00139-17. [Google Scholar] [CrossRef]
  141. Cheng, A.T.; Ottemann, K.M.; Yildiz, F.H. Vibrio cholerae response regulator VxrB controls colonization and regulates the type VI secretion system. PLoS Pathog. 2015, 11, e1004933. [Google Scholar] [CrossRef]
  142. Pratt, J.T.; McDonough, E.; Camilli, A. PhoB regulates motility, biofilms, and cyclic di-GMP in Vibrio cholerae. J. Bacteriol. 2009, 191, 6632–6642. [Google Scholar] [CrossRef]
  143. Barrasso, K.; Watve, S.; Simpson, C.A.; Geyman, L.J.; van Kessel, J.C.; Ng, W.-L. Dual-function quorum-sensing systems in bacterial pathogens and symbionts. PLoS Pathog. 2020, 16, e1008934. [Google Scholar] [CrossRef]
  144. Halang, P.; Toulouse, C.; Geißel, B.; Michel, B.; Flauger, B.; Müller, M.; Voegele, R.T.; Stefanski, V.; Steuber, J. Response of Vibrio cholerae to the catecholamine hormones epinephrine and norepinephrine. J. Bacteriol. 2015, 197, 3769–3778. [Google Scholar] [CrossRef]
  145. Sengupta, N.; Paul, K.; Chowdhury, R. The global regulator ArcA modulates expression of virulence factors in Vibrio cholerae. Infect. Immun. 2003, 71, 5583–5589. [Google Scholar] [CrossRef]
  146. Chourashi, R.; Das, S.; Dhar, D.; Okamoto, K.; Mukhopadhyay, A.K.; Chatterjee, N.S. Chitin-induced T6SS in Vibrio cholerae is dependent on ChiS activation. Microbiology 2018, 164, 751–763. [Google Scholar] [CrossRef]
  147. Klancher, C.A.; Yamamoto, S.; Dalia, T.N.; Dalia, A.B. ChiS is a noncanonical DNA-binding hybrid sensor kinase that directly regulates the chitin utilization program in Vibrio cholerae. Proc. Natl. Acad. Sci. USA 2020, 117, 20180–20189. [Google Scholar] [CrossRef]
  148. García, J.; Madrid, C.; Cendra, M.; Juárez, A.; Pons, M. N9L and L9N mutations toggle Hha binding and hemolysin regulation by Escherichia coli and Vibrio cholerae H-NS. FEBS Lett. 2009, 583, 2911–2916. [Google Scholar] [CrossRef]
  149. Wang, H.; Chen, S.; Zhang, J.; Rothenbacher, F.P.; Jiang, T.; Kan, B.; Zhong, Z.; Zhu, J. Catalases promote resistance of oxidative stress in Vibrio cholerae. PLoS ONE 2012, 7, e53383. [Google Scholar] [CrossRef]
  150. Kitaoka, M.; Miyata, S.T.; Brooks, T.M.; Unterweger, D.; Pukatzki, S. VasH is a transcriptional regulator of the type VI secretion system functional in endemic and pandemic Vibrio cholerae. J. Bacteriol. 2011, 193, 6471–6482. [Google Scholar] [CrossRef]
  151. Silva, A.J.; Sultan, S.Z.; Liang, W.; Benitez, J.A. Role of the histone-like nucleoid structuring protein in the regulation of rpoS and RpoS-dependent genes in Vibrio cholerae. J. Bacteriol. 2008, 190, 7335–7345. [Google Scholar] [CrossRef]
  152. Manneh-Roussel, J.; Haycocks, J.R.; Magán, A.; Perez-Soto, N.; Voelz, K.; Camilli, A.; Krachler, A.-M.; Grainger, D.C. cAMP receptor protein controls Vibrio cholerae gene expression in response to host colonization. MBio 2018, 9, e00966-18. [Google Scholar] [CrossRef] [PubMed]
  153. Liu, C.; Sun, D.; Zhu, J.; Liu, J.; Liu, W. The regulation of bacterial biofilm formation by cAMP-CRP: A mini-review. Front. Microbiol. 2020, 11, 802. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The map of virulence-associated genes and regulatory systems in NOVC: positive relationships are labeled with red arrows and negative relationships are labeled with blue arrows. The whole infection procedure is separated into five stages. Stage 1: survival in host gastrointestinal tract (in orange); stage 2: localization and penetration of the mucus layer in the small intestine (in purple); stage 3: intestinal epithelial cell colonization (in green); stage 4: virulence gene expression (in red); stage 5: detachment from the epithelial cells to return in the environment (in blue). The detailed information is shown in Table S1 in the Supplementary Files.
Figure 1. The map of virulence-associated genes and regulatory systems in NOVC: positive relationships are labeled with red arrows and negative relationships are labeled with blue arrows. The whole infection procedure is separated into five stages. Stage 1: survival in host gastrointestinal tract (in orange); stage 2: localization and penetration of the mucus layer in the small intestine (in purple); stage 3: intestinal epithelial cell colonization (in green); stage 4: virulence gene expression (in red); stage 5: detachment from the epithelial cells to return in the environment (in blue). The detailed information is shown in Table S1 in the Supplementary Files.
Microorganisms 12 00818 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, Q.; Alter, T.; Fleischmann, S. Non-O1/Non-O139 Vibrio cholerae—An Underestimated Foodborne Pathogen? An Overview of Its Virulence Genes and Regulatory Systems Involved in Pathogenesis. Microorganisms 2024, 12, 818. https://doi.org/10.3390/microorganisms12040818

AMA Style

Zhang Q, Alter T, Fleischmann S. Non-O1/Non-O139 Vibrio cholerae—An Underestimated Foodborne Pathogen? An Overview of Its Virulence Genes and Regulatory Systems Involved in Pathogenesis. Microorganisms. 2024; 12(4):818. https://doi.org/10.3390/microorganisms12040818

Chicago/Turabian Style

Zhang, Quantao, Thomas Alter, and Susanne Fleischmann. 2024. "Non-O1/Non-O139 Vibrio cholerae—An Underestimated Foodborne Pathogen? An Overview of Its Virulence Genes and Regulatory Systems Involved in Pathogenesis" Microorganisms 12, no. 4: 818. https://doi.org/10.3390/microorganisms12040818

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop