Skip to main content

Occurrence of virulence determinants in vibrio cholerae, vibrio mimicus, vibrio alginolyticus, and vibrio parahaemolyticus isolates from important water resources of Eastern Cape, South Africa

Abstract

Background

Virulence determinants are crucial to the risk assessment of pathogens in an environment. This study investigated the presence of eleven key virulence-associated genes in Vibrio cholerae (n = 111) and Vibrio mimicus (n = 22) and eight virulence determinants in Vibrio alginolyticus (n = 65) and Vibrio parahaemolyticus (n = 17) isolated from six important water resources in Eastern Cape, South Africa, using PCR techniques. The multiple virulence gene indexes (MVGI) for sampling sites and isolates as well as hotspots for potential vibriosis outbreaks among sampling sites were determined statistically based on the comparison of MVGI.

Result

The PCR assay showed that all the V. cholerae isolates belong to non-O1/non-O139 serogroups. Of the isolates, Vibrio Cholera (84%), V. mimicus (73%), V. alginolyticus (91%) and V. parahaemolyticus (100%) isolates harboured at least one of the virulence-associated genes investigated. The virulence gene combinations detected in isolates varied at sampling site and across sites. Typical virulence-associated determinants of V. cholerae were detected in V. mimicus while that of V. parahaemolyticus were detected in V. alginolyticus. The isolates with the highest MVGI were recovered from three estuaries (Sunday river, Swartkopps river, buffalo river) and a freshwater resource (Lashinton river). The cumulative MVGI for V. cholerae, V. mimicus, V. alginolyticus and V. parahaemolyticus isolates were 0.34, 0.20, 0.45, and 0.40 respectively. The targeted Vibrio spp. in increasing order of the public health risk posed in our study areas based on the MVGI is V. alginolyticus > V. parahaemolyticus > V. cholerae > V. mimicus. Five (sites SR, PA5, PA6, EL4 and EL6) out of the seventeen sampling sites were detected as the hotspots for potential cholera-like infection and vibriosis outbreaks.

Conclusions

Our findings suggest that humans having contact with water resources in our study areas are exposed to potential public health risks owing to the detection of virulent determinants in human pathogenic Vibrio spp. recovered from the water resources. The study affirms the relevancy of environmental Vibrio species to the epidemiology of vibriosis, cholera and cholera-like infections. Hence we suggest a monitoring program for human pathogenic Vibrio spp. in the environment most especially surface water that humans have contact with regularly.

Peer Review reports

Introduction

Water has been identified as one of the earth’s most precious and threatened resources which must be well protected and enhanced for good human health [1]. The causative agents of health challenges that usually emanate from water are diverse nonetheless, the role of bacteria especially the pathogenic ones cannot be overemphasized. V. cholerae and its close relative V. mimicus are bacteria of public health importance, especially as an etiological agent of waterborne infections. V. cholerae serotypes O1 and O139 are famous for the several cholera pandemics recorded over the years while non-O1/non-O139 serotypes have been responsible for several cholera-like outbreaks [2,3,4,5,6,7,8,9]. Also, V. mimicus which is a close relative of V. cholerae has been implicated in outbreaks of diseases such as gastroenteritis, ear infections and severe cholera-like diarrhoea in the time past [10,11,12]. The commonly reported virulence determinants of V. cholerae irrespective of the serotype are hemolysin (hlyA), the actin cross-linking repeats toxin (rtxA), hemagglutinin protease (hap), type III and VI secretion systems, vibrio pathogenic highland (vpi), neuraminidase-encoding (nanH) gene, toxin (NAG-ST and ctx genes), accessory cholera enterotoxin (ace), zot (zonula occludens toxin), transmembrane regulatory protein gene (toxR) and toxin-coregulated pili (tcp) [2, 9, 13,14,15,16,17,18]. Other important vibrios pathogenic island (VPI) associated virulence determinants include ald, tag, toxT, acfB, acfC, orfZ, orfW, int, LJ and RJ genetic elements. Also, cep and orfU are the other two CTX-associated virulence genetic elements of importance that are found in V. cholerae [19]. Although it is scarce in the literature, some of the V. cholerae and V. parahaemolyticus typical genes most especially toxin and toxin regulatory genes have been detected in V. mimicus [20].

Vibrio parahaemolyticus is an established human pathogen that causes gastroenteritis, wound infections and some other human diseases [21]. The pathogen is the most implicated Vibrio spp. in seafood gastroenteritis. The key virulent signatures found in V. parahaemolyticus are the thermostable direct hemolysin (TDH), thermor-stable-related hemolysin (TRH) and the thermo labile hemolysin (TLH) genes. Of the three, TDH and TRH genes have been proposed as the most important virulence factors for V. parahaemolyticus human infections while the role of TLH gene in Vibrio parahaemolyticusrelated human infections is regarded unclear [21,22,23,24,25]. However, recent studies suggested that thermolabile hemolysin (TLH) gene could be as important as TDH and TRH in human infection episodes [24, 25]. The gene was reported to up-regulate in the human gastrointestinal model and also lyse human erythrocytes [24,25,26,27]. Vibrio alginolyticus (formerly V. parahaemolyticus biotype 2) was considered a non-human pathogen however, it has recently become a bacteria of public health concern because of its involvement in human disease conditions [28]. Although there is a paucity of information on the virulence capability of the organism, the available documentation showed that the organism is a potential reservoir of many virulence genes known in other members of the Vibrio genus [29,30,31,32,33]. For example, the tdh, trh and tlh genes and other virulence genes commonly found in V. parahaemolyticus, vopD gene [30,31,32]and vopB genes belonging to the T3SS, and vgrG, hcp and vasH genes of the T6SS have been detected in V. alginolyticus [34]. Aside the waterborne and foodborne infections cause by V. cholerae and its close relative V. mimicus; V. parahaemolyticus and V. alginolyticus, they also cause economic losses in mariculture and aquaculture farms round the globe. Mariculture, aquaculture and recreational fishing are common activities that exposes human to different types of surface water resources in the Eastern Cape Province.

The presence/detection of bacteria in water resources is not enough to ascertain the magnitude of public health risk posed by the bacteria since the degree of pathogenicity, severity and treatability of infections is directly proportional to the number of virulence and resistance genes present/acquired by pathogens [35].

The detection of virulence determinants in the pathogenic bacteria isolated from water resources is essential for microbial risk assessment and this kind of assessment will enhance appropriate decision-making in the epidemiology of pathogenic Vibrio species. Hence, we elucidated the presence of virulence determinants in four medically important Vibrio species earlier isolated from important water resources in the Eastern Cape as reported in one of our previous studies [36] and determine hotspots for potential cholera and vibriosis outbreaks among our sampling sites.

Result

The non-detection of rfb-O1 and rfb-O139 genes confirms all V. cholerae isolates in this study as members of non-O1/non-O139 serogroup. The prevalence of targeted virulence determinants in isolates from freshwater and brackish water is given in Table 1 while the variability in the gene combinations detected in the Vibrio species isolates is given in Table 2. Eighty-four per cent of V. cholera (n = 93), 73% of V. mimicus (n = 16), 91% of V. alginolyticus (n = 60) isolates and 100% of V. parahaemolytic (n = 17) harboured at least one of the virulence-associated genes investigated (Tables S1-S8). The gel pictures showing the expected DNA band sizes of the regions of interest on the targeted genes are given in Plates S1-S10.

Table 1 Prevalence (%) of targeted virulence genes in isolates from freshwater and brackish water samples

Variability in virulence genes combination, multiple virulence gene index (MVGI), and determination of hotspots for potential cholera and vibriosis outbreak based on targeted virulence determinants

In this study, different combinations of targeted virulence genes were detected in the Vibrio spp. isolates and these are given in Table 2. Nine different gene combinations types were found among V. cholerae from brackish water samples while fifteen were found among V. cholerae from freshwater samples. Two different combination types were detected among V. mimicus from brackish water samples while seven were found among V. mimicus recovered from freshwater samples. One combination type was found among V. alginolyticus from freshwater but eleven among V. alginolyticus from brackish water samples. Seven different virulence gene combination types were found among V. parahaemolyticus isolates from brackish water samples. The site with the most diverse Vibrio species base on virulence gene combination types detected was SKR. The list of isolates with their corresponding sampling site and characteristic MVGI is given in Tables S1-S8. The CMVGI for the isolates and sampling sites is given in Table 3 with V. cholerae having the highest CMVGI of 0.36 among isolates from fresh water. On the other hand, V. alginolyticus had the highest CMVGI of 0.48 among isolates from brackish water. Site EL4 has the highest CMVGI of 0.45 among freshwater sampling sites while site SR has the highest CMVGI of 0.54 among the sites from brackish water sampling sites. Leven test revealed that only MVGI data for freshwater isolates is parametric variable while MVGI data for brackish water isolates, freshwater and brackish water sampling sites are non-parametric variables. The ANOVA showed that mean CMVGI for freshwater isolates are significantly different while Welch ANOVA showed that mean CMVGI for brackish water isolates, freshwater and brackish water sampling sites are significantly different. The benferroni Post Hoc test for fresh water isolates showed that CMVGI for V. cholerae and V. mimicus are not significantly different while Game-Howell Post Hoc test showed that CMVGI of V. alginolytiucs and V. parahaemolyticus of the isolates from brackish water were not significantly differenttoo. Hence, of the four Vibrio spp. used for this study, V. cholerae and V. mimicus are the most probable organisms to cause vibrio-related infections at the freshwater sampling environ while V. alginolytiucs and V. parahaemolyticus are most likely bacteria to cause vibrio-related infections at the brackish water sampling sites based on criterion set in section 2.3. Of the brackish water sampling sites, the Game-Howell Post Hoc test revealed that site SR’s CMVGI is significantly higher than that of other four brackish water sampling sites (PA7, SKR, EL5 and EL6). However, the CMVGI of those four other brackish water sampling sites despite been numerically different in the order EL6 > SKR > EL5 > PA7, are not statistically different. On the other hand, sites PA1, PA2, PA3, PA5 and PA6 have CMVGI that are not significantly different from CMVGI of site EL4 which has the highest CMVGI among the nine freshwater sampling sites. Therefore, the statistical result suggested sites PA1, PA2, PA3, PA5, PA6 and EL4 as freshwater hotspots and site SR as the brackish water hotspot for possible relatively high vibrio-related infections outbreaks among the sixteen sampling sites from which isolates used for this study were recovered. The ANOVA and Welch statistical test results are given in Table 4 while Post Hoc test results are in Table S10.

Table 2 Different virulence gene combinations detected in isolates
Table 3 Descriptive statistics of Multiple virulence gene indexes (MVGI) of isolates and sampling sites
Table 4 Statistical comparison of the CMVGI of four Vibrio spp. Isolates and the sampling sites

Discussion

This study reveals the prevalence of eleven virulence genes in the non-O1/non-O139 V. cholerae and V. mimicus and eight virulence genes in V. alginolyticus and V. parahaemolyticus isolates. The low prevalence of key cholera-associated virulence factors (zot, tcp, ctx and ace) and the relatively high prevalence of other virulence-associated factors (hyla, rtx, toxR, ompU, vpi) in the non-O1/non-O139 V. cholerae is in concordance with the previous studies [9, 18, 37,38,39,40,41,42,43,44,45]. The two most essential virulence genes for cholera epidemics and pandemics are tcp and ctx genes [46,47,48]. Interestingly, some earlier reports detected the two genes in non-O1/non-O139 V. cholerae strains [49, 50] such as O141, O75, O27, O37, O53, and O65 strains [51,52,53,54,55] while some others like the current study did not [56,57,58,59]. The presence of the two genes in non-O1/non-O139 strains suggests that toxigenic non-O1/non-O139 strains of V. cholerae exist. Although CTX+ and TCP+V. cholerae non-O1/non-O139 were not encountered in the present study, ctx but tcp+V. cholerae non-O1/non-O139 was detected and the ability of ctxand tcp to acquire the two genes has been reported [60, 61]. Although Vibrio cholerae non-O1/non-O139 strains even those carrying tcp and ctx do not cause cholera epidemic/pandemic, they have been implicated in vibrioses such as gastroenteritis, ear infections, septicemia, and cholera-like infections which are sometimes severe and fatal most especially in immunocompromised patients [8, 59, 62, 63]. The work of [64] showed that several other virulence determinants like those found in V. cholerae in this study work in synergy and are responsible for the pathogenicity of non-O1/non-O139 serogroups of V. cholerae that are negative for tcp and ctx genes. The non-detection of the ctx gene and detection of a relatively low tcp+V cholerae in this study does not guarantee that our sampling areas are free of possible future cholera or fatal vibrio related diarrheal disease outbreaks for three reasons. Firstly, molecular studies have shown that the non-toxigenic strain of V. cholerae can be transformed into toxigenic strain by CTXφ (V. cholerae phage) infestation and antigenic shift that results from the homologous recombination-mediated exchange of O-antigen biosynthesis (wb*) clusters between toxigenic and non-toxigenic strains of V. cholerae [55, 65]. Secondly, the isolation of tcp+ but ctxV. cholerae strains from Sunday and Kowie River water samples suggests that the two rivers are potential hotspots for cholera causing Vibrio cholerae in the future since tcp gene plays a pivotal role in V. cholerae pathogenicity in term of the cholera outbreak. The toxin coregulated pili (TCP) gene detected in V. cholerae from the two rivers acts as a receptor for CTXφ and when CTXφ infests non-toxigenic V. cholerae, it can lead to the emergence of a new toxigenic strain [61, 66]. Thirdly, previous works have shown that Type III Secretion system (T3SS) is relatively common in non-O1/non-O139 V. cholerae, and the T3SS+ non-O1/non-O139 V. cholerae causes severe and fatal diarrhea even more rapidly than V. cholerae O1 in animals model. The T3SS is not commonly found in V. cholerae O1 and O139 serogroups [18, 67,68,69,70]. Also, toxR, vpi, ompU, hyla, and rtx virulence genes are important drivers of Vibrio spp. related infections. The relatively high prevalence of the five aforementioned genes in Vibrio spp. isolates in this study suggests that a sizable number of the V. cholerae and V. mimicus isolates in our study area have virulence gene architecture that only need to acquire a few more virulence determinants to become epidemics causing Vibrio species. The high prevalence of ToxR gene in V. cholerae and V. mimicus isolates further affirms the clinical importance of the isolates. Vibrio spp. with this gene will possibly not find it difficult to express their virulence determinants in a suitable host such as humans since the virulence functions of V. cholerae and V. mimicus are regulated by the transmembrane protein encoded by ToxR gene. [34, 71]. The VPI region of Vibrio genome usually has many features that are typical of pathogenicity islands and these include presence of low G + C content (35%) compared to the rest of the genome (48%), phage-like attachment (att), a transposase-like gene, and a phage-like integrase gene (int) among others [72]. The ompU gene is essential for the adhesion of Vibrio spp. to its host and reduces the permeation of antibiotics across the membrane barriers [73]. The hylA gene is an important virulence determinant of non-O1 and non-O139 Vibrio cholerae. It encodes the hemolysin gene, which plays a very important role in cytotoxicity and apoptosis [74].

Vibrio mimicus had been implicated in both waterborne and foodborne disease outbreaks [10, 75, 76] however, there is paucity of information on the virulence determinants of V. mimicus in the literature. Vibrio mimicus causes infections such as gastroenteritis, ear infections, and severe cholera-like diarrhea. Some virulence determinants peculiar to V. cholerae have been detected in V. mimicus isolates [11, 77]. In this study, all the virulence-associated genes detected in V. cholerae were also detected in at least one V. mimicus isolate except for zot and tcp genes. This finding supports some of the few reports that have shown that V. mimicus could carry similar virulence determinants cassette commonly found in V. cholerae [10, 20, 78] and this could have something to do with both species sharing common ancestors [77]. The presence of typical V. cholerae virulence determinants in Vibrio mimicus suggests that the bacterium should be one of the Vibrio pathogens to be investigated during cholera outbreaks especially when Vibrio cholerae cannot be isolated from samples collected for investigation.

The relatively high prevalence of typical Vibrio parahaemolyticus virulence genes in Vibrio alginolyticus in this study supports earlier reports that showed that most Vibrio spp. possesses atypical virulence genes in addition to their typical virulence determinants [79]. For example, the work of [80] detected V. cholerae ctxAB gene in V. diabolicu, V. alginolyticus and V. parahaemolyticus while studies carried out by [30] and [81] reported V. parahaemolyticus tdh gene in V. harveyi and V. mimicus respectively. Also, V. cholerae virulence determinants specifically zot, ace and vpi genes, have been detected in V. alginolyticus [32, 33]. V. parahaemolyticus is one of the important indicators recommended for accessing the microbial quality of seafood but V. alginolyticus is not [82]. The findings of this study and that of earlier reports show that it is important to investigate both V. alginolyticus and V. parahaemolyticus simultaneously whenever vibriosis peculiar to V. parahaemolyticus is suspected or in the microbial assessment of seafood that requires testing for V. parahaemolyticus.

In this study, none of the targeted virulence genes was detected in some of our isolates Tables S1-S8. However, these isolates are potential agents of vibrio-related infections since they can acquire virulence factors via horizontal gene transfer. Horizontal gene transfer and recombination are common phenomena that play significant roles in the evolution and emergence of new pathogenic strains among Gram-negative organisms [83, 84]. It is common knowledge that horizontal gene transfer plays a vital role in the exchange of genetic materials among Vibrio spp. [84,85,86,87,88,89].

It was observed in this study that the four targeted Vibrio species with varying combinations of virulence determinants co-existed in the same ecological niche and this has been reported in some earlier work [9, 29, 37, 79, 80, 90, 91]. The most diverse Vibrio spp. in this study based on the different combinations of the targeted virulence determinants were found at sampling sites PA4, PA5, PA6, PA7, EL6, SR, and SKR where the level of pollution and anthropogenic activities were relatively high. Pollution promotes horizontal gene transfer and enhances bacterial virulence and antibiotic resistance [79, 92]. This could explain the high variability in terms of virulence determinants combinations found among population of each of the four targeted Vibrio spp. at the sampling sites PA4, PA5, PA6, PA7, EL6, SR, and SKR. The presence of strains with relatively high variability in virulence-associated gene combinations found at these seven sampling sites is of great public health concern. It has been shown that each virulence determinant that makes up the virulence determinant combinations found in the strains interact synergistically to achieve disease conditions [64]. The presence of different strains of V. cholerae, V. mimicus, V. alginolyticus and V. parahaemolyticus harboring various virulence determinants combinations in the water resources sampled interfere with the resourcefulness of the water resources. Water resources contaminated with these pathogens are not good for drinking, irrigation, recreational and agricultural purposes because it has been established that such water resources could lead to disease outbreaks [8, 93, 94]. As such, the presence of virulence determinants carrying Vibrio spp. in the water resources especially the freshwater resources in a province that suffers yearly water scarcity [95,96,97] calls for concern since surface water are key and integral part of water supply in South Africa. People most especially in the rural areas in the absence of sustainable access to potable water, seek for alternative sources to meet their fundamental needs, and surface water is the first point of call as it is easy to access and use [98,99,100,101].

There was no known cholera outbreak ongoing when the isolates that were used for this study were recovered from the aquatic milieu and this we believe, accounted for our inability to detect the cholera pandemic and epidemic-causing V. cholerae O1 and O139 serotypes among the V. cholerae isolates in our study. Nevertheless, the possibility of future outbreaks of vibrio-related illnesses including cholera has been articulated above. Also, the current virulence signature of the V. cholerae and V. mimicus suggests that some of the sampling areas where the isolates were recovered are at the risk of vibriosis and immunocompromised individuals such as people with HIV and tuberculosis are at higher risk [102]. Compared to other provinces in South Africa, the burden of HIV and tuberculosis is relatively high in Eastern Cape Province most especially in the Sarah Baartman district municipality [94, 103, 104]. Unfortunately, 80%, 40%, 76%, and 63% of the virulence containing V. cholerae non-O1/non-O139, V. mimicus, V. alginolyticus and V. parahaemolyticus respectively were recovered from water resources in this district municipality.

In summary, this study which is the first of its kind in the study area shows that the four Vibrio species isolates investigated especially those from water resources in Sarah Baartman district municipality, have the potential to cause cholera-like infection and other vibrioses. The virulence gene combinations detected in isolates varied at each sampling site and across sites. All the water resources that harbored the various genotypes (based on the virulence genes combinations) of the four Vibrio spp. are important water resources that are normally used for agricultural (fishing and irrigation), recreational and spiritual ablution purposes.

Furthermore, typical virulence-associated determinants of V. cholerae (vpi, toxR, ompU, tcp, hyla, rtxA and rtxC) were detected in V. mimicus while that of V. parahaemolyticus (vpi, tlh, vppC, vopB2, vgrG and hcp) were detected in V. alginoyiticus. The isolates with the highest MVGI among each of the four targeted Vibrio spp. were recovered from three estuaries (Sunday river, Swartkopps river, buffalo river) and a freshwater resource (Lashinton river). The cumulative MVGI for V. cholerae, V. mimicus, V. alginolyticus and V. parahaemolyticus isolates were 0.34, 0.20, 0.45, and 0.40 respectively. The presence of pathogenic Vibrio spp. in the water resources especially the freshwater resources in a country that is freshwater stressed, calls for concern. The targeted pathogens in increasing order of public health risk posed based on the MVGI can be represented as V. alginolyticus > V. parahaemolyticus > V. cholerae > V. mimicus. Eight (sites SR, PA1, PA2, PA3, PA5, PA6, EL4, and EL6) out of the sixteen sampling sites were detected as the hotspots for potential cholera and vibriosis based on MVGI analysis. Although the MVGI analysis for brackish water sampling sites detected only site SR as the hotspot for vibrio-related infections among the brackish water sampling sites, sites EL6 and SKR can also be considered as important hotspot for vibrio-related infections since they have CMVGI of approximately 0.4. The virulence genes detected in the four medically important Vibrio spp. isolates in this study are key to the pathophysiology of cholera and vibriosis. Therefore, to prevent outbreaks of vibrio-related infections in the study area, epidemiological proactive measures such as creation of awareness and Vibrio monitoring programes for our sampling sites most especially rivers where the hotspots are located (Kowie, Bloukrans, Lashinton, Kubusi, Buffallo, Swartkopps and Sunday rivers) is hereby advocated. The relatively high prevalence of virulence-associated genes in isolates from water resources where the level of pollution was relatively high suggests the need to prioritize and enforce hygiene as a major component of the monitoring program.

Contribution to knowledge

The information on the virulence capability of vibrio isolates from the aquatic environment in Eastern Cape Province is limited in the literature. We believe that the data generated from our study are relevant to microbial risk assessment of Vibrio spp. and support the development of environmental regulations that will help in preventing cholera and vibriosis outbreaks at our sampling sites most especially the hotspots. The study also provides information on the virulence of V. mimicus and V. alginolyticus, two Vibrio spp. that are not commonly reported from water sources.

Materials and methods

Isolates

The V. cholerae (n = 111), V. mimicus (n = 22), V. alginolyticus (n = 65) and V. parahaemolyticus (n = 17) used for this study were environmental isolates recovered from water samples collected from six important water resources in Eastern Cape, Province, South Africa as earlier reported [36]. The water resources from which the isolates were recovered include Kowie River and two of its tributaries, Sunday River, Swartkopps River, Buffalo River, Kubusi River and one of its tributaries, and two of the University of Fort Hare’s Dams.

DNA extraction and PCR assay

An 18 h old culture of all isolates was prepared from 20% glycerol stock culture stored at -80 oC using freshly prepared sterile brain heart infusion broth. The isolates were afterwards streaked out on a freshly prepared nutrient agar plate containing 1% NaCl and incubated at 37 oC for 18 h. After the incubation period, using the boiling method, the genetic material of a distinct colony of isolates was extracted as described by [36] and used as the template for the PCR assay. Firstly, V. cholerae isolates were delineated into O1, O139 or non-O1/non-O139 strains using rfb-O1 primer specific for V. cholerae O1 strain and rfb-O139 primer specific for O139 strain. This was followed by the molecular detection of eleven virulence determinants (rtxA, rtxC, toxR, ompU, ctx, ctxB, vpi, hylA, tcpA, Zot, ace) in V. cholerae and V. mimicus isolates and eighth virulence determinants (vpi, tdh, trh, tlh, vppC, vopB2, vgrG, hcp) in V. alginolyticus and V. parahaemolyticus. The primers sequences, expected amplicon sizes and thermal cycler conditions employed in this study and the PCR type employed for the amplification of the specific regions of the targeted genes are given in the Table S9. The duplex and singleplex PCR assays were carried out in a 25 µl reaction mixture. The reaction mixture and volume of each component of the mixture are as enunciated in an earlier study [105]. Amplicons sizes were resolved on 1.5% agarose gel and the resulting gel was stained with ethidium bromide (1 µg/mL) and viewed under a transilluminator. A reaction mixture without the DNA template and E. coli ATCC 35,150 was used as negative and internal controls for the PCR assays respectively.

Multiple virulence gene index (MVGI) determination and statistical analysis

The multiple virulence gene indexes (MVGI) were determined using the equation I below as earlier reported [79]. The cumulative multiple virulence gene indexes (CMVGI) for the freshwater, brackish water sampling sites, and each of the four Vibrio spp. that were targeted in this study were determined using equation II below.

$$MVGI{\text{ }} = {\text{ }}VGD/VGT$$
(I)
$$CMVGI{\text{ }} = {\text{ }}AVGD/\left( {VGT{\text{ }}x{\text{ }}NI} \right)$$
(II)

Where MVGI = multiple virulence gene index, CMVGI = cumulative multiple virulence gene index, VGD = total number of virulence genes detected, VGT = total number of virulence genes targeted, AVGD = aggregate virulence gene detected, NI = number of isolates.

The presence or absence of pollution at the water sampling sites was determined using criteria (color, turbidity, temperature, suspended solids, foam) for determining the presence of physical pollutants [106]. To understand the association between pollution and the prevalence of virulence genes in isolates, a correlation analysis was carried out on the MVGI of isolates versus the detection of pollution at the sampling sites. Detection of pollution was treated as a dichotomous variable such that detection of pollution was coded as one (1) while non-detection was coded as zero (0).

The hotspots for potential cholera-like and vibriosis infections were determined by comparing CMVGI across sites statistically. Since salinity affects the prevalence of Vibrio spp. in aquatic milieu, brackish water, and freshwater sampling sites were analyzed independently. To decide on parametric or non-parametric tools for analysis, variables were subjected to Levene test. One-way Analysis of Variance (ANOVA) and Bonferroni Post Hoc Test were employed to compare CMVGI across sites for parametric variables while Welch ANOVA test and Game-Howell Post Hoc test were employed for non-parametric variables.

The specie with the highest CMVGI and any other specie with CMVGI that is not significantly different from that of the species with the highest CMVGI out of the four Vibrio spp. were regarded as the Vibrio spp. of utmost public health importance in our sampling areas. The different combinations of virulence determinants detected in each of the four Vibrio spp. were also noted.

The site having the highest CMVGI served as the reference site for the determination of hotspots among our sampling sites. The site(s) with CMVGI that is not significantly different from the CMVGI of the reference site (the first hotspot) were regarded as additional hotspots for possible cholera and vibriosis outbreaks. The p-value was set at 0.05 for all analyses.

Data Availability

All data generated or analysed during this study are included in this published article [and its supplementary information files].

References

  1. Heijnen HA, editor. Water for development: Safe water for all Bangladesh perspective. Dhaka: DPHE-WHO Programme; Sustainable Development and Healthy Environment; 2002. pp. 1–140. Available: https://www.ircwash.org/sites/default/files/822-BD02-17375.pdf.

  2. Hasan NA, Rezayat T, Blatz PJ, Choi SY, Griffitt KJ, Rashed SM, et al. Nontoxigenic Vibrio cholerae non-O1/O139 isolate from a case of human gastroenteritis in the U.S. Gulf Coast. J Clin Microbiol. 2015;53:9–14. https://doi.org/10.1128/JCM.02187-14.

    Article  CAS  PubMed  Google Scholar 

  3. Ali M, Nelson AR, Lopez AL, Sack DA. Updated Global Burden of Cholera in Endemic Countries. Remais J V., editor. PLoS Negl Trop Dis. 2015;9: e0003832. https://doi.org/10.1371/journal.pntd.0003832.

  4. Azman AS, Rudolph KE, Cummings DAT, Lessler J. The incubation period of Cholera: a systematic review. J Infect. 2013;66:432–8. https://doi.org/10.1016/j.jinf.2012.11.013.

    Article  PubMed  Google Scholar 

  5. WHO. Weekly epidemiological record Relevé épidémiologique hebdomadaire. 2018 [cited 4 Feb 2019]. Available: http://www.who.int/wer.

  6. WHO. Weekly epidemiological record Relevé épidémiologique hebdomadaire. 2017 [cited 4 Feb 2019]. Available: http://www.who.

  7. Dutta D, Chowdhury G, Pazhani GP, Guin S, Dutta S, Ghosh S, et al. Vibrio cholerae non-O1, non-O139 serogroups and cholera-like diarrhea, Kolkata, India. Emerg Infect Dis. 2013;19:464–7. https://doi.org/10.3201/eid1903.121156.

    Article  PubMed  PubMed Central  Google Scholar 

  8. Morris JG. Infections due to non-O1/O139 Vibrio cholerae - UpToDate. [cited 4 Feb 2019]. Available: https://www.uptodate.com/contents/infections-due-to-non-o1-o139-vibrio-cholerae.

  9. Ceccarelli D, Chen A, Hasan NA, Rashed SM, Huq A, Colwell RR. Non-O1/non-O139 Vibrio cholerae carrying multiple virulence factors and V. Cholerae O1 in the Chesapeake Bay, Maryland. Appl Environ Microbiol. 2015;81:1909–18. https://doi.org/10.1128/AEM.03540-14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Guardiola-Avila I, Acedo-Felix E, Sifuentes-Romero I, Yepiz-Plascencia G, Gomez-Gil B, Noriega-Orozco L. Molecular and genomic characterization of vibrio mimicus isolated from a frozen shrimp processing facility in Mexico. PLoS ONE. 2016;11:1–15. https://doi.org/10.1371/journal.pone.0144885.

    Article  CAS  Google Scholar 

  11. Singh DV, Isac SR, Colwell RR. Development of a Hexaplex PCR Assay for Rapid Detection of Virulence and Regulatory genes in Vibrio cholerae and Vibrio mimicus. 2002;40: 4321–4. https://doi.org/10.1128/JCM.40.11.4321.

  12. Amin RA, Salem AM. Specific detection of pathogenic Vibrio species in Shellfish by using Multiplex polymerase chain reaction. 2012;8: 525–31.

  13. Lin W, Fullner KJ, Clayton R, Sexton JA, Rogers MB, Calia KE, et al. Identification of a vibrio cholerae RTX toxin gene cluster that is tightly linked to the Cholera toxin prophage. Proc Natl Acad Sci U S A. 1999;96:1071–6. https://doi.org/10.1073/pnas.96.3.1071.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Jiang F, Bi RR, Deng LH, Kang HQ, 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;74:117–22. https://doi.org/10.1016/j.ijid.2018.06.021.

    Article  CAS  PubMed  Google Scholar 

  15. Ghosh C, Nandy RK, Dasgupta SK, Nair GB, Hall RH, Ghose AC. A search for cholera toxin (CT), toxin coregulated pilus (TCP), the regulatory element ToxR and other virulence factors in non-01/non-0139 Vibrio cholerae. Microb Pathog. 1997;22: 199–208. Available: http://www.ncbi.nlm.nih.gov/pubmed/9140915.

  16. O’Shea YA, Reen FJ, Quirke AM, Boyd EF. Evolutionary genetic analysis of the emergence of epidemic Vibrio cholerae isolates on the basis of comparative nucleotide sequence analysis and multilocus virulence gene profiles. J Clin Microbiol. 2004;42:4657–71. https://doi.org/10.1128/JCM.42.10.4657-4671.2004.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Reidl J, Klose KE. Vibrio cholerae and Cholera: out of the water and into the host. FEMS Microbiol Rev. 2002;26:125–39. https://doi.org/10.1111/j.1574-6976.2002.tb00605.x.

    Article  CAS  PubMed  Google Scholar 

  18. Schirmeister F, Dieckmann R, Bechlars S, Bier N, Faruque SM, 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;33:767–78. https://doi.org/10.1007/s10096-013-2011-9.

    Article  CAS  PubMed  Google Scholar 

  19. Sarkar A, Nandy RK, Nair GB, Ghose AC. Vibrio pathogenicity island and Cholera toxin genetic element-associated virulence genes and their expression in non-O1 non-O139 strains of Vibrio cholerae. Infect Immun. 2002;70:4735–42. https://doi.org/10.1128/IAI.70.8.4735-4742.2002.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Shinoda S, Nakagawa T, Shi L, Bi K, Kanoh Y, Tomochika K, et al. Distribution of Virulence-Associated genes in Vibrio mimicus isolates from Clinical and Environmental origins. Microbiol Immunol. 2004;48:547–51. https://doi.org/10.1111/j.1348-0421.2004.tb03551.x.

    Article  CAS  PubMed  Google Scholar 

  21. Gutierrez West CK, Klein SL, Lovell CR. High frequency of virulence factor genes tdh, trh, and tlh in Vibrio parahaemolyticus strains isolated from a pristine estuary. Appl Environ Microbiol. 2013;79:2247–52. https://doi.org/10.1128/AEM.03792-12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Raimondi F, Kao JPY, Fiorentini C, Fabbri A, Donelli G, Gasparini N, et al. Enterotoxicity and cytotoxicity of Vibrio parahaemolyticus Thermostable Direct Hemolysin in in Vitro systems. Infect Immun. 2000;68:3180–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Zhang XH, Austin B. Haemolysins in Vibrio species. J Appl Microbiol. 2005;98:1011–9. https://doi.org/10.1111/j.1365-2672.2005.02583.x.

    Article  CAS  PubMed  Google Scholar 

  24. Letchumanan V, Chan KG, Khan TM, Bukhari SI, Mutalib NSA, Goh BH, et al. Bile sensing: the activation of Vibrio parahaemolyticus virulence. Front Microbiol. 2017;8:1–6. https://doi.org/10.3389/fmicb.2017.00728.

    Article  Google Scholar 

  25. Wang R, Zhong Y, Gu X, Yuan J, Saeed AF, Wang S. The pathogenesis, detection, and prevention of Vibrio parahaemolyticus. Front Microbiol. 2015;6:1–13. https://doi.org/10.3389/fmicb.2015.00144.

    Article  Google Scholar 

  26. Gotoh K, Kodama T, Hiyoshi H, Izutsu K, Park KS, Dryselius R, et al. Bile acid-induced virulence gene expression of Vibrio parahaemolyticus reveals a novel therapeutic potential for bile acid sequestrants. PLoS ONE. 2010;5. https://doi.org/10.1371/journal.pone.0013365.

  27. Broberg CA, Calder TJ, Orth K. Vibrio parahaemolyticus cell biology and pathogenicity determinants. Microbes Infect. 2011;13:992–1001. https://doi.org/10.1016/j.micinf.2011.06.013.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Citil BE, Derin S, Sankur F, Sahan M, Citil MU, Report C. Case Report Vibrio alginolyticus Associated Chronic Myringitis Acquired in Mediterranean Waters of Turkey. 2015;2015. https://doi.org/10.1155/2015/187212.

  29. Sabir M, Ennaji Moulay M, Cohen N. Vibrio Alginolyticus: An Emerging Pathogen of Foodborne Diseases. Int J Sci Technol. 2013;2: 302–309. Available: http://www.journalofsciences-technology.org/archive/2013/april_vol_2_no_4/66925136139495.pdf.

  30. Gennari M, Ghidini V, Caburlotto G, Lleo MM. Virulence genes and pathogenicity islands in environmental Vibrio strains nonpathogenic to humans. FEMS Microbiol Ecol. 2012;82:563–73. https://doi.org/10.1111/j.1574-6941.2012.01427.x.

    Article  CAS  PubMed  Google Scholar 

  31. Xie Z-Y, Hu C-Q, Chen C, Zhang L-P, Ren C-H. Investigation of seven Vibrio virulence genes among Vibrio alginolyticus and Vibrio parahaemolyticus strains from the coastal mariculture systems in Guangdong, China. Lett Appl Microbiol. 2005;41:202–7. https://doi.org/10.1111/j.1472-765X.2005.01688.x.

    Article  CAS  PubMed  Google Scholar 

  32. Sechi LA, Duprè I, Deriu A, Fadda G, Zanetti S. Distribution of Vibrio cholerae virulence genes among different Vibrio species isolated in Sardinia, Italy. J Appl Microbiol. 2000;88:475–81. https://doi.org/10.1046/j.1365-2672.2000.00982.x.

    Article  CAS  PubMed  Google Scholar 

  33. Snoussi M, Noumi E, Usai D, Sechi LA, Zanetti S, Bakhrouf A. Distribution of some virulence related-properties of Vibrio alginolyticus strains isolated from Mediterranean seawater (Bay of Khenis, Tunisia): investigation of eight Vibrio cholerae virulence genes. World J Microbiol Biotechnol. 2008;24:2133–41. https://doi.org/10.1007/s11274-008-9719-1.

    Article  CAS  Google Scholar 

  34. Xie Z-Y, Hu C-Q, Chen C, Zhang L-P, Ren C-H (2005) Investigation of seven Vibrio virulence genes among Vibrio alginolyticus and Vibrio parahaemolyticus strains from the coastal mariculture systems in Guangdong, China. Lett Appl Microbiol 41:202–207. https://doi.org/10.1111/j.1472-765X.2005.01688.x

    Article  CAS  PubMed  Google Scholar 

  35. Hernández-Robles MF, Álvarez-Contreras AK, Juárez-García P, Natividad-Bonifacio I, Curiel-Quesada E, Vázquez-Salinas C, et al. Virulence factors and antimicrobial resistance in environmental strains of Vibrio alginolyticus. Int Microbiol. 2016;19:191–8. https://doi.org/10.2436/20.1501.01.277.

    Article  CAS  PubMed  Google Scholar 

  36. El-Baky RMA, Ibrahim RA, Mohamed DS, Ahmed EF, Hashem ZS. Prevalence of virulence genes and their association with antimicrobial resistance among pathogenic E. Coli isolated from Egyptian patients with different clinical Infections. Infect Drug Resist. 2020;13:1221–36. https://doi.org/10.2147/IDR.S241073.

    Article  Google Scholar 

  37. Abioye OE, Osunla C, Okoh A. Molecular detection and distribution of six medically important Vibrio spp. in selected freshwater and Brackish Water resources in Eastern Cape Province, South Africa. 2021;12. https://doi.org/10.3389/fmicb.2021.617703.

  38. Sack RB, Choi J, Grim CJ, Sadique A, Taviani E, Colwell RR, et al. Distribution of virulence genes in clinical and environmental Vibrio cholerae strains in Bangladesh. Appl Environ Microbiol. 2013;79:5782–5. https://doi.org/10.1128/aem.01113-13.

    Article  PubMed  PubMed Central  Google Scholar 

  39. De Menezes FGR, Rodriguez MTT, de Carvalho FCT, Rebouças RH, Costa RA, de Sousa OV, et al. Pathogenic Vibrio species isolated from estuarine environments (Ceará, Brazil) - antimicrobial resistance and virulence potential profiles. An Acad Bras Cienc. 2017;89:1175–88. https://doi.org/10.1590/0001-3765201720160191.

    Article  CAS  PubMed  Google Scholar 

  40. Bidinost C, Saka HA, Aliendro O, Sola C, Panzetta-Duttari G, Carranza P, et al. Virulence factors of non-O1 non-O139 Vibrio cholerae isolated in Córdoba, Argentina. Rev Argent Microbiol. 2004;36:158–63.

    CAS  PubMed  Google Scholar 

  41. Sharma C, Thungapathra M, Ghosh A, Mukhopadhyay AK, Basu A, Mitra R, et al. 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;36:756–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Kumar A, Thomas S, Harichandran D, Karim S, Khan S, Meparambu D. Fatal non-O1, non-O139 Vibrio cholerae septicaemia in a patient with chronic Liver Disease. J Med Microbiol. 2013;62:917–21. https://doi.org/10.1099/jmm.0.049296-0.

    Article  PubMed  Google Scholar 

  43. Lu Z, Wang Y, Jin Y, Chen B, Bai Y, Shao C, et al. A case of non-O1/non-O139 Vibrio cholerae Septicemia and Meningitis in a neonate. Int J Infect Dis. 2015;35:117–9. https://doi.org/10.1016/j.ijid.2015.05.004.

    Article  PubMed  Google Scholar 

  44. Lu B, Zhu F, Huang L, Cui Y, Li D, Li F, et al. The first case of bacteraemia due to non-O1/non-O139 Vibrio cholerae in a type 2 Diabetes Mellitus patient in mainland China. Int J Infect Dis. 2014;25:116–8. https://doi.org/10.1016/j.ijid.2014.04.015.

    Article  PubMed  Google Scholar 

  45. Theophilo GND, Rodrigues DDP, Leal NC, Hofer E. Distribution of virulence markers in clinical and environmental Vibrio cholerae non-O1/non-O139 strains isolated in Brazil from 1991 to 2000. Rev Inst Med Trop Sao Paulo. 2006;48:65–70. https://doi.org/10.1590/S0036-46652006000200002.

    Article  PubMed  Google Scholar 

  46. Chatterjee S, Patra T, Ghosh K, Raychoudhuri A, Pazhani GP, Das M, et al. Vibrio cholerae O1 clinical strains isolated in 1992 in Kolkata with progenitor traits of the 2004 Mozambique variant. J Med Microbiol. 2009;58:239–47. https://doi.org/10.1099/jmm.0.003780-0.

    Article  CAS  PubMed  Google Scholar 

  47. Faruque SM, Kamruzzaman M, Meraj IM, Chowdhury N, Nair GB, Sack RB, et al. Pathogenic potential of environmental Vibrio cholerae strains carrying genetic variants of the toxin-coregulated pilus pathogenicity island. Infect Immun. 2003;71:1020–5. https://doi.org/10.1128/IAI.71.2.1020-1025.2003.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Sánchez J, Holmgren J. Virulence factors, pathogenesis and vaccine protection in Cholera and ETEC diarrhea. Curr Opin Immunol. 2005;17:388–98. https://doi.org/10.1016/J.COI.2005.06.007.

    Article  PubMed  Google Scholar 

  49. Childers BM, Klose KE. Regulation of virulence in Vibrio cholerae: the ToxR regulon. Future Microbiol. 2007;2:335–44. https://doi.org/10.2217/17460913.2.3.335.

    Article  CAS  PubMed  Google Scholar 

  50. Takahashi E, Ochi S, Mizuno T, Morita D, Morita M, Ohnishi M, et al. Virulence of Cholera Toxin Gene-positive Vibrio cholerae Non-O1/non-O139 strains isolated from Environmental Water in Kolkata, India. Front Microbiol. 2021;12:1–13. https://doi.org/10.3389/fmicb.2021.726273.

    Article  Google Scholar 

  51. Chakraborty S, Mukhopadhyay AK, Kumar Bhadra R, Nath Ghosh A, Mitra R, Shimada T et al. Virulence Genes in Environmental Strains of Vibrio cholerae. Appl Environ Microbiol. 2000. Available: http://aem.asm.org/.

  52. Dalsgaard A, Serichantalergs O, Forslund A, Lin W, Mekalanos J, Mintz E, et al. Clinical and environmental isolates of Vibrio cholerae serogroup O141 carry the CTX phage and the genes encoding the toxin-coregulated pili. J Clin Microbiol. 2001;39:4086–92. https://doi.org/10.1128/JCM.39.11.4086-4092.2001.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Tobin-D’Angelo M, Smith AR, Bulens SN, Thomas S, Hodel M, Izumiya H, et al. Severe Diarrhea caused by Cholera Toxin–Producing Vibrio cholerae Serogroup O75 Infections acquired in the Southeastern United States. Clin Infect Dis. 2008;47:1035–40. https://doi.org/10.1086/591973.

    Article  PubMed  Google Scholar 

  54. Haley BJ, Choi SY, Grim CJ, Onifade TJ, Cinar HN, Tall BD, et al. Genomic and phenotypic characterization of Vibrio cholerae Non-O1 isolates from a US Gulf Coast Cholera Outbreak. Chang Y-F. Editor PLoS One. 2014;9:e86264. https://doi.org/10.1371/journal.pone.0086264.

    Article  CAS  Google Scholar 

  55. Crowe SJ, Newton AE, Gould LH, Parsons MB, Stroika S, Bopp CA, et al. Vibriosis, not Cholera: toxigenic Vibrio cholerae non-O1, non-O139 Infections in the United States, 1984–2014. Epidemiol Infect. 2016;144:3335–41. https://doi.org/10.1017/S0950268816001783.

    Article  CAS  PubMed  Google Scholar 

  56. Li M, Shimada T, Morris JG, Sulakvelidze A, Sozhamannan S. Evidence for the emergence of Non-O1 and Non-O139 Vibrio cholerae strains with pathogenic potential by Exchange of O-Antigen biosynthesis regions. Infect Immun. 2003;71:588–8. https://doi.org/10.1128/IAI.71.1.588.2003.

    Article  CAS  PubMed Central  Google Scholar 

  57. Iyer L, Vadivelu J, Puthucheary SD. Detection of Virulence Associated Genes, Haemolysin and Protease amongst Vibrio cholerae Isolated in Malaysia. 2000. Available: https://about.jstor.org/terms.

  58. Jagadeeshan S, Kumar P, Abraham WP, Thomas S. Multiresistant Vibrio cholerae non-O1/non-O139 from waters in South India: resistance patterns and virulence-associated gene profiles. J Basic Microbiol. 2009;49:538–44. https://doi.org/10.1002/jobm.200900085.

    Article  CAS  PubMed  Google Scholar 

  59. Shinoda S, Nakagawa T, Hirakawab N, Miyoshi S, Arakawac E, Ramamurthyd T et al. Molecular Epidemiological Studies of Vibrio cholerae in BengalRegion. Biocontrol Sci. 2008. Available: https://www.jstage.jst.go.jp/article/bio1996/13/1/13_1_1/_pdf.

  60. Chowdhury G, Joshi S, Bhattacharya S, Sekar U, Birajdar B, Bhattacharyya A, et al. Extraintestinal Infections caused by non-toxigenic Vibrio cholerae non-O1/non-O139. Front Microbiol. 2016;7:144. https://doi.org/10.3389/fmicb.2016.00144.

    Article  PubMed  PubMed Central  Google Scholar 

  61. Dziejman M, Serruto D, Tam VC, Sturtevant D, Diraphat P, Faruque SM et al. Type III Secretion System Source: Proceedings of the National Academy of Sciences of the United States of Linked references are available on JSTOR for this article : Genomic characterization of non-01, non-0139 Vibrio cholerae reveals genes for a type 1. 2005.

  62. Waldor MK, Mekalanos JJ. Lysogenic conversion by a filamentous phage encoding Cholera toxin. Sci (80-). 1996;272:1910–3. https://doi.org/10.1126/SCIENCE.272.5270.1910.

    Article  CAS  Google Scholar 

  63. Petsaris O, Nousbaum JB, Quilici ML, Coadou GL, Payan, Abalain ML. Non-O1, non-O139 Vibrio cholerae bacteraemia in a cirrhotic patient. J Med Microbiol. 2010;59: 1260–1262. doi:I https://doi.org/10.1099/jmm.0.021014-0.

  64. Zmeter C, Tabaja H, Sharara AI, Kanj SS. Non-O1, non-O139 Vibrio cholerae Septicemia at a tertiary care center in Beirut, Lebanon; a case report and review. J Infect Public Health. 2018;11:601–4. https://doi.org/10.1016/J.JIPH.2018.01.001.

    Article  PubMed  Google Scholar 

  65. Rajpara N, Vinothkumar K, Mohanty P, Singh AK, Singh R, Sinha R et al. Synergistic Effect of Various Virulence Factors Leading to High Toxicity of Environmental V. cholerae Non-O1/ Non-O139 Isolates Lacking ctx Gene: Comparative Study with Clinical Strains. Hensel M, editor. PLoS One. 2013;8: e76200. https://doi.org/10.1371/journal.pone.0076200.

  66. Davis BM, Kimsey HH, Chang W, Waldor MK. The Vibrio cholerae O139 Calcutta bacteriophage CTX? Is infectious and encodes a novel repressor. Med J Aust. 2007;187:345–7. https://doi.org/10.5694/j.1326-5377.2007.tb01278.x.

    Article  Google Scholar 

  67. Mukhopadhyay AK, Mitra R, Colwell RR, Shimada T, Takeda Y, Faruque SM, et al. Virulence genes in environmental strains of Vibrio cholerae. Appl Environ Microbiol. 2002;66:4022–8. https://doi.org/10.1128/aem.66.9.4022-4028.2000.

    Article  Google Scholar 

  68. Tulatorn S, Preeprem S, Vuddhakul V, Mittraparp-arthorn P. Comparison of virulence gene profiles and genomic fingerprints of Vibrio cholerae O1 and non-O1/non-O139 isolates from diarrheal patients in southern Thailand. Trop Med Health. 2018;46:1–8. https://doi.org/10.1186/s41182-018-0113-x.

    Article  Google Scholar 

  69. Dziejman M, Serruto D, Tam VC, Sturtevant D, Diraphat P, Faruque SM, 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;102:3465–70. https://doi.org/10.1073/pnas.0409918102.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Chaand M, Miller KA, Sofia MK, Schlesener C, Weaver JWA, Sood V, et al. Type three secretion system island-encoded proteins required for colonization by non-O1/non-O139 serogroup vibrio cholerae. Infect Immun. 2015;83:2862–9. https://doi.org/10.1128/IAI.03020-14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Shin OS, Tam VC, Suzuki M, Ritchie JM, Bronson RT, Waldor MK, et al. Type III secretion is essential for the rapidly fatal Diarrheal Disease caused by non-o1, non-o139 vibrio cholerae. MBio. 2011;2. https://doi.org/10.1128/mBio.00106-11.

  72. Skorupski K, Taylor RK. Control of the ToxR virulence regulon in Vibrio cholerae by environmental stimuli. Mol Microbiol. 1997;25:1003–9. https://doi.org/10.1046/j.1365-2958.1997.5481909.x.

    Article  CAS  PubMed  Google Scholar 

  73. Rajanna C, Wang J, Zhang D, Xu Z, Ali A, Hou Y-M, et al. The vibrio pathogenicity island of epidemic Vibrio cholerae forms precise extrachromosomal circular excision products. J Bacteriol. 2003;185:6893–901. https://doi.org/10.1128/JB.185.23.6893-6901.2003.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Wibbenmeyer JA, Provenzano D, Landry CF, Klose KE, Delcour AH. Vibrio cholerae OmpU and OmpT porins are differentially affected by bile. Infect Immun. 2002;70:121. https://doi.org/10.1128/IAI.70.1.121-126.2002.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Kanoktippornchai B, Chomvarin C, Hahnvajanawong C, Nutrawong T. Role of HLYA-positive vibrio cholerae non-o1/non-o139 on apoptosis and cytotoxicity in a Chinese hamster ovary cell line. Southeast Asian J Trop Med Public Health. 2015;45:1365–75.

    Google Scholar 

  76. Chitov T, Wongdao S, Thatum W, Puprae T, Sisuwan P. Occurrence of potentially pathogenic Vibrio species in raw, processed, and ready-to-eat seafood and seafood products. Maejo Int J Sci Technol. 2009;3:88–98.

    Google Scholar 

  77. Chitov T, Kirikaew P, Yungyune P, Ruengprapan N, Sontikun K. An incidence of large foodborne outbreak associated with Vibrio mimicus. Eur J Clin Microbiol Infect Dis. 2009;28:421–4. https://doi.org/10.1007/s10096-008-0639-7.

    Article  CAS  PubMed  Google Scholar 

  78. Hasan NA, Grim CJ, Haley BJ, Chun J, Alam M, Taviani E et al. Comparative genomics of clinical and environmental Vibrio mimicus. 2010. https://doi.org/10.1073/pnas.1013825107 /-/DCSupplemental.www.pnas.org/cgi/doi/10.1073/pnas.1013825107.

  79. Yu Z, Wang E, Geng Y, Wang K, Chen D, Huang X, et al. Complete genome analysis of Vibrio mimicus strain SCCF01, a highly virulent isolate from the freshwater catfish. Virulence. 2020;11:23–31. https://doi.org/10.1080/21505594.2019.1702797.

    Article  CAS  PubMed  Google Scholar 

  80. Deng Y, Xu L, Chen H, Liu S, Guo Z, Cheng C, et al. Prevalence, virulence genes, and antimicrobial resistance of Vibrio species isolated from diseased marine fish in South China. Sci Rep. 2020;1–8. https://doi.org/10.1038/s41598-020-71288-0.

  81. Hossain S, Wickramanayake MVKS, Dahanayake PS, Heo G-J. Occurrence of virulence and extended-spectrum β-Lactamase determinants in Vibrio Spp. Isolated from marketed hard-shelled mussel (Mytilus coruscus). Microb Drug Resist. 2020;26:391–401. https://doi.org/10.1089/mdr.2019.0131.

    Article  CAS  PubMed  Google Scholar 

  82. Shi L, Miyoshi SI, Bi K, Masami N, Hiura M, Tomochika K, et al. Presence of hemolysin genes (vmh, tdh and hlx) in isolates of Vibrio mimicus determined by polymerase chain reaction. J Heal Sci. 2009;55:857–9. https://doi.org/10.1248/jhs.55.857.

    Article  Google Scholar 

  83. Health Protection Agency. Guidelines for assessing the Microbiological Safety of Ready-to-eat Foods placed on the market. London: Heal Prot Agency; 2009. p. 33.

    Google Scholar 

  84. Almagro-Moreno S, Napolitano MG, Boyd EF. Excision dynamics of Vibrio pathogenicity island-2 from Vibrio cholerae: role of a recombination directionality factor VefA. BMC Microbiol. 2010;10:306. https://doi.org/10.1186/1471-2180-10-306.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Gyles C, Boerlin P. Horizontally transferred genetic elements and their role in Pathogenesis of Bacterial Disease. 2014 [cited 8 Mar 2019]. https://doi.org/10.1177/0300985813511131.

  86. Mel SF, Mekalanos JJ. Modulation of Horizontal Gene Minireview Transfer in Pathogenic Bacteria by In Vivo Signals Here we briefly review two examples in diverse patho-gen-host systems, one well-studied and one newly-dis-*Department of Microbiology and Molecular Genetics †. Cell. 1996. Available: https://www.cell.com/action/showPdf?pii=S0092-8674%2800%2981986-8.

  87. Thomas J, Watve SS, Ratcliff WC, Hammer BK. Horizontal gene transfer of functional type VI killing genes by Natural Transformation. MBio. 2017;8:e00654–17. https://doi.org/10.1128/mBio.00654-17.

    Article  PubMed  PubMed Central  Google Scholar 

  88. Le Roux F, Blokesch M. Eco-evolutionary dynamics linked to horizontal gene transfer in Vibrios. Annu Rev Microbiol. 2018;72:89–110. https://doi.org/10.1146/annurev-micro-090817-062148.

    Article  CAS  PubMed  Google Scholar 

  89. Theethakaew C, Feil EJ, Castillo-Ramírez S, Aanensen DM, Suthienkul O, Neil DM, et al. Genetic relationships of Vibrio parahaemolyticus isolates from clinical, human carrier, and environmental sources in Thailand, determined by multilocus sequence analysis. Appl Environ Microbiol. 2013;79:2358–70. https://doi.org/10.1128/AEM.03067-12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Beceiro A, Tomás M, Bou G. Antimicrobial resistance and virulence: a successful or deleterious association in the bacterial world? Clin Microbiol Rev. 2013;26:185–230. https://doi.org/10.1128/CMR.00059-12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Khaleel SH, Al-Azawia IH, Khlebos AH. Genotyping of vibrio Cholera for virulence factors in diwaniyah city - Iraq. J Pure Appl Microbiol. 2018;12:777–82. https://doi.org/10.22207/JPAM.12.2.38.

    Article  CAS  Google Scholar 

  92. Mohamad N, Amal MNA, Saad MZ, Yasin ISM, Zulkiply NA, Mustafa M, et al. Virulence-associated genes and antibiotic resistance patterns of Vibrio spp. isolated from cultured marine fishes in Malaysia. BMC Vet Res. 2019;15:176. https://doi.org/10.1186/s12917-019-1907-8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Schafer A, Kalinowski J, Puhler A. Increased fertility of Corynebacterium glutamicum recipients in intergeneric matings with Escherichia coli after stress exposure. Appl Environ Microbiol. 1994;60:756–9. https://doi.org/10.1128/aem.60.2.756-759.1994.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Baker-Austin C 1, Oliver JD, Munirul Alam M, Ali A, Waldor MK, Qadri F et al. Vibrio spp. infections. Nat Rev. 2018;4:8: 20. Available: file:///C:/Users/hp/Desktop/Primer_1531468992_1_Vibriosis.pdf.

  95. Louisiana Office of Public Health. CHOLERA & OTHER VIBRIOS. 2006. Available: www.infectiousdisease.dhh.louisiana.gov

  96. CDP Global Water Report. Lead Partner National Business Initiative Report written by Irbaris and Incite Sustainability Recognising the strategic value of water. 2013. Available: http://www.nbi.org.za/wp-content/uploads/2016/06/CDP-2012-water-report.pdf.

  97. Western Cape Government. Western Cape Sustainable Water Management Plan-2012. 2012. Available: https://www.greencape.co.za/assets/Water-Sector-Desk-Content/WC-DEAP-DWS-Western-Cape-sustainable-water-management-plan-2012.pdf.

  98. Lewis RSA. Statement by the Wildlife and Environment Society of South Africa, on outspoken address at canoe marathon on South Africa’s potential water crisis by Richard Lewis (08/11/2012). 2012 [cited 26 Feb 2019]. Available: http://www.polity.org.za/article/sa-statement-by-the-wildlife-and-environment-society-of-south-africa-on-outspoken-address-at-canoe-marathon-on-south-africas-potential-water-crisis-by-richard-lewis08112012-2012-11-08.

  99. Madilonga RT, Edokpayi JN, Volenzo ET, Durowoju OS, Odiyo JO. Water quality assessment and evaluation of human health risk in mutangwi river, Limpopo province, South Africa. Int J Environ Res Public Health. 2021;18. https://doi.org/10.3390/ijerph18136765.

  100. Bosman C. Guideline for the Management of Waterborne Epidemics, with the emphasis on Cholera. Management. 2002; 1–52. Available: http://www.dwa.gov.za/Dir_WQM/docs/CHOLERA.pdf.

  101. Wilma OSS. A CSIR perspective 2010. 2010.

  102. Edokpayi JN, Odiyo JO, Popoola EO, Msagati TAM. Evaluation of Microbiological and Physicochemical Parameters of Alternative Source of Drinking Water: a case study of Nzhelele River, South Africa. Open Microbiol J. 2018;12:18–27. https://doi.org/10.2174/1874285801812010018.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Turner JW. Environmental factors and reservoir shifts contribute to the seasonality of pathogenic. University of Georgia; 2010.

  104. Spotlight. The numbers: HIV and TB in South Africa • Spotlight. 2018 [cited 21 Feb 2019]. Available: https://www.spotlightnsp.co.za/2018/07/04/the-numbers-hiv-and-tb-in-south-africa/.

  105. MartinVA. Mid-year population estimates 2017. 2017. Available: www.statssa.gov.zainfo@statssa.gov.za.

  106. Kanabus A. TB Statistics for South Africa | National & provincial: “Information about Tuberculosis" 2018 [cited 21 Feb 2019]. Available: https://www.tbfacts.org/tb-statistics-south-africa/

  107. Abioye OE, Okoh AI. Limpet (Scutellastra cochlear) recovered from some estuaries in the Eastern Cape Province, South Africa Act as reservoirs of pathogenic Vibrio species. Front Public Heal. 2018;6. https://doi.org/10.3389/fpubh.2018.00237.

  108. Englande AJ, Krenkel P, Shamas J. Wastewater Treatment &Water Reclamation. Reference Module in Earth systems and Environmental sciences. Elsevier Inc.; 2015. doi:https://doi.org/10.1016/b978-0-12-409548-9.09508-7.

Download references

Acknowledgements

Not applicable.

Funding

This work was supported by Water Research Commission (Grant number: K5 Number: 2432) and South Africa and Medical Research Council (SAMRC), South Africa.

Author information

Authors and Affiliations

Authors

Contributions

OEA and AIO initiated the research topic, AIO provided materials for the study, OEA and NN structured the methods, OEA and CAO carried out the experiment, OEA carried out the statistical analysis and wrote the manuscript including preparation of Tables and additional document, CAO, NN and AIO proofread and correct the manuscript. All authors approved the manuscript.

Corresponding author

Correspondence to Oluwatayo E. Abioye.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary Material 1

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Abioye, O.E., Osunla, C.A., Nontongana, N. et al. Occurrence of virulence determinants in vibrio cholerae, vibrio mimicus, vibrio alginolyticus, and vibrio parahaemolyticus isolates from important water resources of Eastern Cape, South Africa. BMC Microbiol 23, 316 (2023). https://doi.org/10.1186/s12866-023-03060-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12866-023-03060-z

Keywords