Next Article in Journal
Gut Microbiome and Lipidome Signatures in Irritable Bowel Syndrome Patients from a Low-Income, Food-Desert Area: A Pilot Study
Previous Article in Journal
Review and Perspectives on Bifidobacterium lactis for Infants’ and Children’s Health
Previous Article in Special Issue
Persistent Dysbiosis, Parasite Rise and Growth Impairment in Aquacultured European Seabass after Oxytetracycline Treatment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Tools to Enumerate and Predict Distribution Patterns of Environmental Vibrio vulnificus and Vibrio parahaemolyticus

by
Lisa A. Waidner
1,* and
Trupti V. Potdukhe
2
1
Hal Marcus College of Science and Engineering, University of West Florida, 11000 University Pkwy, Building 58, Room 108, Pensacola, FL 32514, USA
2
GEMS Program, College of Medicine, University of Illinois Chicago, 1853 W. Polk St., Chicago, IL 60612, USA
*
Author to whom correspondence should be addressed.
Microorganisms 2023, 11(10), 2502; https://doi.org/10.3390/microorganisms11102502
Submission received: 7 September 2023 / Revised: 28 September 2023 / Accepted: 3 October 2023 / Published: 5 October 2023
(This article belongs to the Special Issue Host–Bacteria Interactions in Aquaculture Systems)

Abstract

:
Vibrio vulnificus (Vv) and Vibrio parahaemolyticus (Vp) are water- and foodborne bacteria that can cause several distinct human diseases, collectively called vibriosis. The success of oyster aquaculture is negatively impacted by high Vibrio abundances. Myriad environmental factors affect the distribution of pathogenic Vibrio, including temperature, salinity, eutrophication, extreme weather events, and plankton loads, including harmful algal blooms. In this paper, we synthesize the current understanding of ecological drivers of Vv and Vp and provide a summary of various tools used to enumerate Vv and Vp in a variety of environments and environmental samples. We also highlight the limitations and benefits of each of the measurement tools and propose example alternative tools for more specific enumeration of pathogenic Vv and Vp. Improvement of molecular methods can tighten better predictive models that are potentially important for mitigation in more controlled environments such as aquaculture.

1. Introduction

Vibrio are Gram-negative, flagellar bacteria that typically live in marine to estuarine waters [1]. Three major human pathogens in the genus include Vibrio vulnificus, V. parahaemolyticus, and V. cholerae, the first two of which are considered in this review. As facultative anaerobes, Vibrio metabolic demands can be met even in suboxic waters [2]. Vibrio vulnificus (henceforth abbreviated Vv) can enter open cuts or wounds via exposure to contaminated seawater and may cause necrotizing fasciitis (infamously known as “flesh-eating disease”). Systemic cases often necessitate surgery or peripheral amputations, and in rare cases, result in death [3,4]. Vv can also cause gastrointestinal disease through the ingestion of raw or undercooked shellfish and is the leading cause of reported acute gastroenteritis and septicemia cases due to foodborne illness [5]. Over 95% of all Vv cells associated with oysters are in the meat [6], and roughly 50% of Vv human infections are transmitted through food consumption [7]. Similarly, V. parahaemolyticus (Vp) is transmitted through the consumption of raw or undercooked shellfish, especially oysters [8]. Both Vv and Vp can exist in the water column in a planktonic state, but in response to certain molecular signaling events, can become sessile on surfaces such as shellfish, finfish, plankton, or suspended particulate material [9,10]. Although antibiotic-resistant V. cholerae resulting from the use of aquaculture probiotics may confer resistance to non-cholera Vibrio species, as reviewed elsewhere [11], numbers of isolates from environmental samples that are identified as V. cholerae can be quite low [12]. This review focuses on environmental and climate-related parameters known to be associated with higher incidences of Vv and Vp in several types of environmental samples.
Unfortunately, the global occurrence of outbreaks and rates of reported Vv and Vp cases may increase [13,14,15]. Vibriosis cases are predicted to rise, in part, due to anthropogenic climate change and associated extreme meteorological events and downstream changes to the waterbodies where Vibrio naturally reside [16,17,18,19]. Antibiotics usage in aquaculture areas may also lead to Vibrio populations with antibiotic tolerance (“persisters”, please see below). Additional considerations are the specific genetic and phenotypic properties of strains, serovars, and “serovariants”, which are extensive in both Vv and Vp species, as evident from just a single study of oyster-associated strains [20]. An increase in variant numbers among strains, i.e., increased genetic diversity, should be considered, especially when enumerating with functional genes, such as those that confer fitness in aquaculture facilities, fitness for sessile living state, and/or higher pathogenicity to humans [21]. Hence, it is particularly important to recognize the limitations and strengths of the various molecular tools for detection of Vibrio in environmental samples.

2. Objectives

The primary aims of this review are to (i) provide an overview of virulence genes used to call for and enumerate Vv and Vp in environmental samples, (ii) summarize current knowledge of environmental factors driving Vv and Vp success, and (iii) present examples of alternative molecular tools for ecological surveys and better predictions of Vv and Vp distribution changes with various environmental and climate-related perturbations.

3. Virulence Factors and Associated Marker Genes

3.1. Culture-Dependent and -Independent Method Considerations

Microbiologists often use culture-dependent methods. However, the limitation of first relying on growth in liquid or solid media before enumerating the bacteria may lead us to underestimate the numbers in the substrates of interest, such as fish tissue, sediment, and water column samples. The “viable but not culturable” (VBNC) state is a survival mechanism in which bacteria are alive, but do not grow in culture media. Entering the VBNC state often occurs in response to unfavorable conditions, such as, in the case of Vv, extreme salinity changes or large temperature drops [22]. “Persister” cells are those that are possibly induced by the usage of antibiotics in aquaculture areas, as previously reviewed elsewhere [23] Briefly, persisters are similar to VBNC cells, in that both types are more stress-tolerant than vegetative cells, but persisters are in a less dormant state than VBNC cells; additionally, persister cells are generally antibiotic-tolerant. Some surveys of the isolates of aquaculture samples resulted in significant number of isolates with multiple Antibiotic Resistance Genes (ARGs), especially from aquaculture waters where antibiotics usage was low [24]. A few ARG marker genes are provided in Table 1, and a more comprehensive list is available in [24]. For further review on Vv VBNC and persisters, please see [23], and for a proteomics study of resuscitation mechanisms of Vp from the VBNC state, see [25]. An extensive review of Vibrio VBNC and persister states in survival and epidemiological implications was recently published elsewhere [26]. While lower temperatures may lead to underestimation of environmental or mammalian–host-associated Vv [27,28], the opposite may be true for Vp in controlled culture conditions where salinity, not temperature, is the important factor affecting the VBNC state [29]. In slightly different culture conditions, increased temperatures can somewhat reverse the VBNC issue [30]. Additionally, physiological responses of oysters to hypoxia may influence the accumulation of Vibrio in the meat [31]. Other environmental stressors on the host organism, such as salinity, temperature, and metal concentrations may also influence Vibrio loads in the oyster [6,32].
A variety of methods have been reported for enumeration and screening for Vibrio in water, sediment and aquaculture samples. These include those based on solely culture-independent (“molecular”) tools, where DNA or RNA is extracted directly from the substrate and evaluated using deep sequencing of a marker gene (“amplicon” libraries), qPCR, RT-qPCR, metagenomics, or metatranscriptomics. Since the culture-dependent step is skipped, the bacterial VBNC state does not affect detection. In waterbodies annually experiencing wide temperature ranges, hence possibly inducing the VBNC state of Vibrio, considerations should be made when enumerating during cold months. For example, in the Chesapeake Bay, Vv could comprise ~10% to 50% of culturable bacterial populations in oysters [33], easily enumerated via direct-plating/colony hybridization (DP/CH; i.e., a method that is inherently culture-based). However, in the same estuary, Vibrio numbers declined with temperature, thereby evading detection with these standard culture methods [34].
Currently, most studies rely on a combination of enrichment followed by molecular verification. An example is MPN (most probable number)-qPCR (MPN-qPCR) or MPN-PCR, where the substrate is diluted in liquid culture, counts are based on the MPN method, and then turbid cultures are subjected to either non-quantitative PCR or quantitative PCR (qPCR). The non-quantitative (PCR) or quantitative (qPCR) step is designed to verify presence, or to obtain gene abundances from the species or strains of interest. MPN-PCR and MPN-qPCR have long been used for pathogen screening of environmental or food samples [35,36,37]. Other methods based on culture-dependent, followed by culture-independent tools include DP/CH and DP-colony PCR, both relying on the success of the Vibrio to first grow on a solid culture medium. All such methods, liquid- and solid-culture-based (DP/CH, DP/CH-followed by PCR confirmation, DP followed by colony PCR, MPN-PCR, and MPN-qPCR) rely on the “cultivability” of the species of interest. Thus, rather than using only conventional microbiological methods for detection and quantification of bacterial communities, it can be more valuable to use molecular methods that evaluate DNA or RNA for confirming the presence/absence or providing quantities of the strains of interest. Care must be taken to interpret the findings of any type of study (culture-dependent, or -independent) when modeling predictions of increased pathogenic Vv or Vp with any environmental or ecological factor [38]. A newer type of quantitative PCR (digital droplet PCR, ddPCR) has recently been used to quantify pathogenic Vibrio species in the environment [39]. Benefits of pathogenic screening bypassing the culture step, and performing qPCR or RT-qPCR directly on DNA or RNA extracted from the substrate, has been extensively reviewed elsewhere [40,41,42].

3.2. Marker Genes

Sequencing of conserved taxonomic marker genes, such as the 16S rRNA gene, 23S rRNA gene, and rpoA, can help in the taxonomic identification of various Vibrio isolates, regardless of species, strain, or pathogenicity status [43,44]. However, due to the highly conserved sequences of these phylogenetic markers, PCR of these genes are usually not suitable for screening for species or specific strains, and thus cannot be used to reliably differentiate between pathogenic and non-pathogenic strains. Genus-level molecular tools were used in earlier studies of environmental Vibrio [44,45,46], some of which are included in this review for describing the early findings of environmental parameters driving Vibrio success. “Functional” genes are now more often used as molecular markers to enumerate and/or score samples as positive or negative for Vibrio. Both Vv and Vp possess specialized genes that contribute to pathogen virulence by promoting phenotypes such as increased epithelial (intestinal) cell adhesion or erythrocyte (red blood cell) lysis (i.e., hemolysis). Some enteropathic bacteria are capable of host invasion through mucus-binding proteins, including those encoded by gbpA of V. cholerae [47] and V. vulnificus [48]. An early phenotypic screening tool for pathogenic Vibrio was the modified Elek test method [49]. A positive result, called the Kanagawa phenomenon, is characterized by the halo formation around colonies grown on Wagatsuma agar, signifying the colony contains a Vp strain actively expressing the ß-hemolysin. The Kanagawa phenomenon test must be validated with culture-independent methods, since pathogenic strains from human patients, fish and water that score tdh-positive do not always present the Kanagawa phenotype [50].
The Vv-specific virulence gene, vvhA, encodes an extracellular cytolysin (VVH) which, if present, often confers higher likelihood of pathogenesis, possibly by helping the Vv invade host intestines, thereby allowing for its entry into the bloodstream [51]. Published vvhA PCR primer pairs (Table 1) are routinely used to detect Vv in DNA from seawater, sediments, fish or shellfish tissues. Recently, vvhA has been used to enumerate total pathogenic Vv, while the virulence-correlated gene (vcg) is now considered as a stronger indicator of virulence in Vv. Both environmental (vcgE) and clinical (vcgC) genes are strongly associated with distinct types of the hemolysin gene, vvhA, among clinical and environmental isolates [52,53].
Table 1. Noteworthy Vibrio marker genes. Common marker genes that are used to identify and enumerate Vibrio vulnificus (Vv) and V. parahaemolyticus (Vp) in environmental samples are listed here. Although most studies rely on PCR to detect species-specific genes, selected genes from recent publications relying on transcriptomics, genomics or metagenomics are also presented.
Table 1. Noteworthy Vibrio marker genes. Common marker genes that are used to identify and enumerate Vibrio vulnificus (Vv) and V. parahaemolyticus (Vp) in environmental samples are listed here. Although most studies rely on PCR to detect species-specific genes, selected genes from recent publications relying on transcriptomics, genomics or metagenomics are also presented.
Gene of InterestSpeciesEncodes 1Mode of PathogenesisPCR Primers 2Reference 3
VvVp
vvhA VVH cytolysin (hemolysin)Extracellular cytotoxin, cytolytic pore-forming protein, pore-forming cholesterol-dependent cytolysinvvhA_F/R[34]
vvhA-F/R[54]
vvhA 1973 rev/vvhA 1795[55]
F-vvh785/R-vvh990[56,57]
tdh Thermostable direct hemolysinIncrease permeability of host cell membranes, act as toxic porins, result in cell lysistdh86F/tdh331R[58]
trh tdh-related hemolysintrh90F/500R
tlh ThermoLabile hemolysintlh-F/R[59]
tdh, trh with IAC4 see abovesee abovetdh, trh with IAC[60]
tdh, trh, tlh with IAC4 tdh, trh, tlh with IAC[61]
ToxRRegulatory transcriptional membrane proteinStrongly associated with the upregulation of the gene encoding tdh in Vp [62] or vvhA in Vv [63]UToxF/vvtoxR (Vv); UToxF/vptoxR (Vp)[64]
vcgC, vcgE “Virulence-correlated gene” (C, clinical; E, environmental)Strongly associated with the expression of vvhAvcgCP1/P3vcgEP2/P3[52,53]
vpm Vp metalloproteaseShows proteolytic activity towards type I collagen; degrades host tissue and may promote pathogen invasionvpm1/2[65]
pilA Type IV pilinInvolved in biofilm formation, adherence to human epithelial cells, and oyster colonization [66,67,68] VvpAF3/VvpBR6[69]
pilF PilF-F/R[70,71]
brpL, brpG BrpL, BrpGInvolved in extracellular polysaccharide production, biofilm formation; affects rugose colony phenotypeN/A[72]
hupA TonB-dependent heme and hemoglobin receptorImportant for iron acquisition during infections [73]hupA_F1/R1[74]
sodB Iron superoxide dismutasePromotes survival and virulence under acid stress and phagocyte engulfment [75]FeSOD_F1/R1
vvpE ferric uptake regulator (Fur)Regulation of vvhA expressionN/A[76,77]
rtxAMultifunctional-autoprocessing repeats-in-toxin (MARTX)Improves antiphagocytosis, colonization, and dissemination to the bloodstream and other organsN/A[78,79]
plpA Toxic lipoprotein phospholipase A2Causes necrotic cell death of epithelial cells, lyses human erythrocytes of membranes containing sufficient phosphatidylcholineN/A[80,81]
wza Polysaccharide export lipoproteinFormation of capsule to allow bacterium to evade host immune system; increases Vv survival in the presence of serumwza_F/R[82]
OmpUConserved outer membrane protein, fibronectin binding proteinToxic porin that forms nonspecific β-barrel channels allowing free diffusion of hydrophilic molecules across outer membrane; shown to exhibit resistance to bile and antibacterial peptidesOmpU R-F/R[83]
gbpA GbpA, GlcNAc-binding protein AMucin-binding protein involved in colonization of host intestineGBPA_qRT-F/R[48]
Example regulatory genes used as PCR markers
HlyUDNA binding proteinPromotes activation of ExsA, a master regulator of type III secretion systems that control >40 genes, many of which are involved in virulenceNT398/399[84,85]
RseB Periplasmic negative regulator of the alternative sigma factor EControls an extensive regulon involved in responding to cell envelope stresses; associated with the colony morphotype of extracellular polysaccharidesrseB5′USER/rseB3′USER[86]
rtxA RTX toxin proteinExported via Type I Secretion System, many roles in infection, including growth in host, and host cell necrosis and apoptosis N/A[87,88]
vvpR, smcR VvpR, SmcRHomolog of LuxR gene in V. fischeri; quorum sensing master regulator, in Vv, regulates transcription of the maltose regulonN/A[87,89]
Example ARGs or ARG regulator genes used as PCR markers 5
blaCARB-17 like element ß-lactamase Intrinsic Vp ampR, disrupts beta-lactam ring present in beta-lactam-class antibioticsCARB-VP-F/R[90]
tetB tetracycline efflux MFS transporterFor active pumping of the tetracycline compound out of the celltetBDEFHJ-F, tetBD-R[91]
qnrSQnrS family quinolone resistance pentapeptide repeat proteinTetR/AcrR family transcriptional regulator associated with resistance to fluoroquinolonesqnrS-F/R[24]
1 Name of protein encoded by gene of interest. 2 Published PCR primer pair name, name of primer used to call for gene of interest using molecular methods (N/A: PCR was not used, gene sequence groups identified via deep-sequencing or genomics/metagenomics tools). 3 Reference: study in which PCR primer pair was first developed and tested or employed to enumerate Vv or Vp in environmental samples. 4 IAC = Internal Amplification Control. 5 ARG = Antibiotic Resistance Gene. For ARGs, a more comprehensive list of PCR targets for screening of aquaculture samples is provided by [24].
The pilus has a major role in Vibrio pathogenesis. In pathogenic Vv strains, pilF and pilA encode major pilus structural subunits. Those, and potential regulator genes like gacA that may control pil expression, are involved in biofilm formation, oyster colonization, and adherence to host gut epithelial cells [66,67,68,92]. Common PCR primer pairs for enumerating the pilA and pilF genes show promise for seafood pathogen screening [69,70,71]. The expression of the toxic lipoprotein, plp, causes necrotic cell death of epithelial cells and lyses human erythrocytes of membranes. For example, Vp persisting in Pacific oyster tissues rely on types I and IV pili, as well as polar and lateral flagellar systems [93]. Not only has this study helped us to better understand the pilus mechanism at the Vp-oyster interface, but also provided a gateway for more in-depth studies, for evaluating potential effects of relevant environmental parameters, including joint salinity-pH condition. Expression of the aforementioned genes is directly involved in pathogenicity. However, important regulatory genes responsible for the expression of virulence genes are also used to assess potential pathogen loads (Table 1, regulatory gene examples). These include transcriptional activator HlyU [85], periplasmic regulator RseB [86], heme and hemoglobin receptor hupA [74], and quorum-sensing master regulator SmcR [76]. Another example is BrpR, a regulatory protein upstream of several genes known to be involved in Vv biofilm formation and rugose colony formation [72], indicative of the potential for pathogenicity (see Table 1, brpL, brpG).
Similarly, there are several virulence-associated genes in Vp. Generally, the hemolysin genes of major interest for screening environmental samples are tdh, tlh, and trh. Like VVH, Vp hemolysins increase the permeability of host cell membranes, act as toxic porins and could result in cell lysis [94]. The genes encoding the thermostable-direct hemolysin (tdh) and the thermostable-direct hemolysin-related hemolysin (trh) are the most commonly used Vp virulence gene markers (Table 1). The gene encoding the thermolabile hemolysin (tlh) is typically a species-specific marker that is used to call for and enumerate all Vp, regardless of pathogenicity. Not all strains of Vp are pathogenic, and samples that score tdh- or trh-positive (or both) are generally only those that are scored as “hemolysin-producing” [59]. Interestingly, synteny analyses of Vp hemolysin and nearby genes indicate gene order may confer differences in pathogenicity of strains isolated from human wounds or feces, water samples, and shellfish [95].
The ToxR protein, first described in V. cholerae, is a regulatory protein strongly associated with the upregulation of the gene encoding tdh in Vp [62] or vvhA in Vv [63]. The toxR gene can be used to identify pathogenic Vp and Vv in environmental studies [96,97], and a homolog is present in V. alginolyticus, another important related pathogen [64]. The Vp metalloprotease, encoded by vpm, promotes pathogen invasion via degradation of fish host tissue, especially type I collagen [65]; the vpm gene is another example of an alternative PCR target used to screen environmental samples [97]. Table 1 presents the selected PCR primer pairs for these and other example marker genes, along with additional genes described in other molecular-based studies using tools other than PCR. Described below are many of the environmental parameters thought to drive Vibrio success, inferred from studies using a combination of culture-dependent and -independent techniques.

4. Predictors, Published Patterns, and Correlated Ecological Factors

Several water quality parameters may have strong influence on pathogenic Vibrio dynamics, partly due to their known optimal ranges. Temperature and salinity are often reported as drivers of Vibrio populations, regardless of geographical location of the waterbody. A variety of parameters are correlated with abundance and distribution in various regions and waterbody types and can be generally grouped into three categories: i) hydrological and/or meteorological factors, ii) abiotic factors that may be indicative of biological activity, and iii) purely biological factors. Additional considerations should be made when evaluating aquaculture studies, such as the now-known strains of Vp that are resistant to food pressure treatments [98,99]. An overview of environmental parameters potentially driving Vibrio success are provided here and in Table 2.

4.1. Oceanographic, Hydrographic, and Meteorological

4.1.1. Temperature

Strong, positive correlations between water temperature and Vv distribution and/or abundance have been demonstrated in numerous studies, such as [46,55,110,114,120,123,140] among others. Using a combination of DP-CH and MPN-qPCR with primer pairs previously described [61], Vibrio concentrations were enumerated in surface waters, sediments, and oysters in coastal waters of two states in the subtropical northern Gulf of Mexico region [100]. Several environmental parameters, especially sea surface temperature (SST), contribute to Vibrio success in one or more of these substrates, along with chlorophyll a (chl a) and turbidity in the water column. Although many of the earlier studies finding increased Vibrio with higher sea surface temperatures were focused on subtropical or tropical locales, this strong positive relationship is also seen in additional, more temperate regions. For example, in North Carolina Atlantic east coast estuaries, DP/CH also confirmed strong relationships between temperature and Vv and Vp, but additional abiotic and biotic factors were also important considerations [12]. In the North Sea, where water temperatures were low and salinities high, infection occurrence was lower [141]. Warm waters were associated with high concentrations of Vv and Vp in surface waters of subtropical northern Gulf of Mexico but also in more temperate waters in the Pacific Northwest in the Chesapeake Bay [101].
There is little “correlation” of Vibrio with SST when short-term, within-season, evaluations are conducted in both culture-dependent [128] and culture-independent [134,142] types of studies. This is especially evident from studies where the entire sampling period occurs in the warm summer months [134] or from single-year evaluations where temperature is a strong factor [143]. Long-term, there are positive relationships between increased temperature with Vibrio in the zooplankton and phytoplankton seawater fraction [122], in the water column [112], and with Vp associated with shellfish [106]. However, in some cases, there may be a non-significant relationship of long-term effects of SST with Vv and Vp abundances in the water column [18]. Evaluation of epidemiological data and concomitant increases in sea surface temperatures [144,145,146] and other studies, reviewed by [147], are especially important for forming more confidence in ecological and environmental predictor models.
Increases in water temperature could also decrease water quality and induce favorable conditions for pathogen growth, such as salinity changes during droughts. For example, insubtropical surface waters in the northern Gulf of Mexico near Galveston, TX, USA contained the highest abundances of culturable Vv in waters with salinities of 7–16% and water temperatures between 30℃ and 37℃ [103,148]. Temperature was a driving factor in waters near Honolulu, HI, where Vv was low during the summer dry season, when waters were warm, and importantly, more saline [104]. In certain cases, droughts can promote distribution and abundance of pathogenic bacteria. However, some droughts do not seem to have much of an effect. For example, abundance of Vv in North Carolina estuaries was in fact reduced during a drought season compared to an above-average freshwater inflow event [149]. Similarly, a longer-term study (10 years) in the Neuse River Estuary showed that culturable Vibrio numbers were not necessarily associated with the three most commonly reported factors for predicting estuarine Vibrio abundance: salinity, temperature, and dissolved oxygen [150]. In a subtropical estuary, Charlotte Harbor of southwest Florida, there were no significant relationships between the presence of Vv and pH, salinity, turbidity, dissolved nutrients, or estuarine bacteria, an outlier from other studies [151]. These conflicting findings suggest that increased temperature coupled with a specific salinity range may provide more favorable Vibrio growth conditions.

4.1.2. Salinity

Increases in salinity due to drought may actually decrease the levels of Vv associated with oysters, even when both culture-dependent and culture-independent methods of Vibrio measurement are employed [152]. Hindsight analyses indicate that Vv are more abundant in “wet” vs. “dry” years in the temperate Chesapeake Bay system [153]. In certain areas, such as subtropical estuaries, salinity can be a stronger determinant for Vibrio success than temperature. Vibrio are considered to be mesohalophiles (3–16 PSU). Thus, ecological surveys commonly report a negative relationship with salinity, especially at sites proximal to freshwater input [100,118,127,128]. Additional considerations should also include prevalence of the bacteria in surface waters as compared to bottom waters, where surface and bottom Vv may be differentially affected by temperature in the long term [154], a phenomenon likely present in any type of stratified estuarine environment. Some studies suggest that numbers increase proportionally with salinity. For example, a study near the Mississippi Gulf Coast reported a significant relationship between salinity (and turbidity) with Vp in surface waters and oysters [129]. Vibrio exposure can increase from exposure to low-salinity standing water caused by hurricane-induced flooding [155], but prior wind (see Section 4.1.4 and storm surge may also contribute to this problem. Systems affected by abundant freshwater influence, such as Indian tropical monsoonal estuaries, may also experience high Vv and Vp abundances, but in addition to salinity, other correlative factors such as eutrophication (see Section 4.2), reduced flushing leading to hypoxia, and wind direction reversal, are important [156].
Neither salinity nor SST alone is a strong predictor of pathogenic Vv and Vp, suggesting the need to consider other environmental factors influencing their distribution. Many studies of environmental samples consider the specific relationship between temperature and salinity to be the most primary consideration in Vibrio ecology. Warm waters with moderate-to-high salinity are typical for high abundances of Vibrio measured in both temperate and warmer locales [55,110]. The first study to detect and quantify co-occurring Vibrio species in West African coastal waters also confirmed that SST and salinity were driving environmental factors of Vibrio densities, and Vv was highest at the end of the wet season [124]. However, wind events (highest in this region’s dry season) were not considered. In myriad examinations among a variety of geographic regions and types of waterbodies, the one environmental parameter that is least predictive is salinity, where Vibrio abundances can be positively or negatively correlated with it, or where bimodal or non-linear models best fit the data (see “Relationship” column in Table 2).

4.1.3. Alkalinity

Fluctuations in pH levels may also affect Vibrio ecology, particularly when preceded by heavy rainfall [157,158]. In Dauphin Island Bay, AL, USA, the optimal water pH for Vv and Vp were found to be ~7.0–9.0 [127]. In more temperate South Carolina, USA, Vp abundances were negatively correlated with pH, where HABs were prevalent in coastal stormwater detention ponds [115]. In a group of subtropical estuaries, pH was found to be correlated positively with Vv, but negatively with Vp [128]. Evidently, variation in driving factors can be highly region-specific and potentially seasonally driven, but tidal influence or freshwater inputs on fluctuating pH may also be the driving factor in Vibrio abundances. Tidal influence is an especially important predictive factor in estuaries, particularly shallow ones or those subjected to stratification [110,128,132].

4.1.4. Wind

Importantly, besides SST, other considerations such as wind speed and air temperature should be included in ecological models. In a limited geographical area (tropical Florida, USA) there were relationships between coastal clinical cases and multiple meteorological factors [159]. One study elucidated the ecology and culturable Vv and Vp abundances in three substrates (surface waters, bottom sediments, and invertebrate biofilms) within the Pensacola Bay System, FL. High concentrations of Vv and Vp in surface waters may have been due to high wind speeds during the days leading up to sampling, which caused resuspension bacteria that would have otherwise been present within the sediments [128]. Additionally, wind-induced surface agitation may promote oxygenation, favoring Vibrio species whose oxygen preferences align with surface water conditions. For example, in the Chesapeake Bay, differential abundances of Vv and Vp were seen before and after an extreme storm event (Hurricane Irene), possibly because of increased turbidity from vertical mixing. Abundances in the water column, sediment, or both, were significantly correlated with several environmental parameters, notably secchi depth, total suspended solids, and tidal height [132]. Wind direction is also an important consideration, where the shape of the estuary dictates mixing depending on the vector [131], or where wind direction reversal, such as before or after significant storm events [156], can influence the water column Vv and/or Vp abundances.

4.1.5. Turbidity

In coastal environments, an increased rate of fluvial input often increases turbidity, possibly driving Vibrio success [12,100,129]. Mixing and resuspension of bottom sediments transports sediment-dwelling Vibrio upward into surface waters, increasing their overall distribution and consequently reintroducing them to the water column. Not only does particle association allow Vibrio to evade grazing, particles are sources of nutrients and organic carbon [160,161]. In a recent study of coastal Alabama, USA, 30–50% of total Vibrio was associated with particles sized 5 μm or larger in a freshwater-driven system [131]. In a nearby northern Gulf of Mexico coastal region, light attenuation was, not unexpectedly, positively correlated with total suspended solids, but neither surface nor sedimentary Vibrio abundances were correlated with turbidity or light attenuation [128].

4.2. Abiotic Factors

Chlorophyll and Nutrients

Chlorophyll a (chl a) and nutrient concentrations are considered as abiotic factors, but they are indicative of biological activity within the ecosystem. Both biotic and abiotic factors generate selective pressures in ecosystems, which facilitate the emergence of pathogenic traits in the community [162]. In a northern temperate estuary, Vp associated with oysters were positively correlated with water column chl a [163], and the relationships of sedimentary and water column Vibrio with chl a were significant in several other regions [46,110,147]. Two regions near the subtropical Gulf of Mexico coast showed positive correlations of Vv and Vp with chl a in surface waters and sediments [100,128]. Eutrophication often promotes algal blooms. As blooms die and decompose, the dissolved and particulate organic matter released become fodder for heterotrophic bacteria such as Vv and Vp, which are efficient scavengers that often outcompete other non-pathogenic heterotrophs.
There is a clear connection between elevated nutrients and concentrations of Vv and Vp. In nutrient-rich detention ponds of South Carolina, USA [115], concentrations of a variety of nutrients—dissolved organic nitrogen, dissolved inorganic nitrogen, total dissolved phosphorus, and silicate—were important, where almost all correlated positively with both Vv and Vp. Similarly, in the Pensacola Bay System, surface water Vp positively correlated with total dissolved inorganic nitrogen and total Kjeldahl nitrogen [128]. Another study showed that Vp was moderately correlated with total dissolved phosphorus and total dissolved nitrogen in Rehoboth Bay, Delaware, USA [133]. Conversely, a strong inverse relationship was observed with ammonia for both Vv and Vp in North Carolina estuaries [12]. However, a predictive modelling survey in this region indicated that high dissolved organic carbon alone may not have a direct impact but can likely promote more rapid Vibrio growth when temperature and salinity ranges are viable [164].

4.3. Biological Factors

4.3.1. Other Bacterial Pathogens

A positive correlation of Vibrio with total bacteria and/or coliforms is not unexpected, especially when Vibrio abundances are enumerated with culture-based methods [165]. Relationships between surface water Vv and fecal enterococci [124] was consistent with other earlier, culture-based studies [12,166]. Co-occurrences of Vp and other pathogenic bacteria were seen in a northern estuary with fecal contamination, concomitant with positive correlations with particulate matter and zooplankton abundances [167], again suggesting multivariate considerations must be made in predictive modeling. Indeed, in North Carolina estuaries, correlations of Vv, Vp, or both, were seen with several chemical, physical, and biological parameters, including salinity, turbidity, dissolved oxygen, concentrations of inorganic nutrients, and loads of other human pathogens such as E. coli and total coliform bacteria [12].

4.3.2. Cyanobacterial and Eukaryotic Algal Blooms

The two different potentially pathogenic Vibrio species may respond differently to blooms following extreme storm events. In the Neuse River Estuary, North Carolina, USA, after an extreme storm event (Hurricane Matthew), several types of HABs were observed in surface waters, including cyanobacterial types and eukaryotic dinoflagellates, both measured using deep sequencing of the hsp60 gene [136]. Although Vibrio abundances were estimated via relative abundances in deep sequencing, significant correlations with the typical “environmental factors” were seen, but differing relationships with HABs, depending on the type of HAB and the Vibrio species, were also observed. Correlations with a cyanobacterial bloom with Vp and Vv were positive and negative, respectively, but both Vibrio species were positively correlated with dinoflagellate blooms. Therefore, after extreme storm events, water quality, including various photosynthetic pigments such as chl a (as a proxy for potential HABs) should therefore be monitored closely when assessing aquaculture risks. These risks were also assessed in coastal Mediterranean, in a fish aquaculture area, where several water quality markers were correlated with multiple Vibrio species, based on 16S rRNA profiles and culturable bacteria counts [168]. Interestingly, but not surprisingly, bacterial species often associated with fish lesions and/or mortality were associated with relatively high levels of the human pathogens in the Vv and Vp species markers found in the deep sequencing amplicon data.
Species within the copiotrophic Vibrio genus are often associated with surfaces, which could include association with harmful algal bloom (HAB)-forming cyanobacteria, eukaryotic phytoplankton and zooplankton. Indeed, a combination of higher SST and chitinous plankton abundance are associated with higher Vv [143], and distinct pathogenic types are associated with distinct groups of chitinous eukaryotic plankton and follow particular patterns with respect to environmental conditions such as salinity and temperature [39]. Capabilities of extracellular polysaccharide matrix formation for Vibrio and their resulting increased fitness were previously reviewed elsewhere [160,169]. Archived plankton samples may be screened with molecular methods to further understand long-term SST increases and prediction of Vibrio dynamics. Using genus-specific PCR primers, Vibrio associated with plankton were positively correlated with SST, in addition to other climate indices, namely, the Northern Hemisphere Temperature and the Atlantic Multidecadal Oscillation [122].
Co-occurrences of high relative abundances of Vibrio and HAB species are also of particular interest. In temperate coastal and inland waters of Delaware, USA, higher abundances of Vibrio are concomitant with various eukaryotic species, often associated with blooms, as assessed with both size fractionation and in whole water analyses [44,133]. The early Delaware study relied solely on phylogenetic markers to identify relative abundances of Vibrio species of interest, but the 2022 Rosales study in the same geographic region used more specific genetic markers that are most likely to reflect the more pathogenic strains of Vv and Vp. Additionally, in subtropical estuaries of the Gulf of Mexico coast (Alabama USA), the findings are similar to those of temperate Delaware [131]. Size fractionation prior to DNA extraction and analysis using qPCR with the most reliable marker genes indicate that Vv and Vp are associated with blooms of Akashiwo sanguinea and Heterocapsa spp. Even more notably, and supporting other findings [136], the effect of wind prior to the HAB bloom influences the strength of the responding Vibrio increased abundances.
In both northern and southern temperate regions, as well as subtropical waters, abundances are positively correlated with higher relative amounts of markers of algal or cyanobacterial biomass, including but not limited to photosynthetic pigments, microscopy counts, HAB marker genes, or relative HAB abundances based on phylogenetic marker gene datasets [131,137]. Geographic considerations, including the maximum or persistence of SST must also be included in predictive modeling. In a Prorocentrum bloom of temperate coastal surface seawater (East China Sea), based on Vibrio-specific genes in a metagenomics dataset, as well as 16S rRNA relative abundances in amplicon deep sequencing data, there was a strong (but not statistically significant) association between Vibrio genera gene abundances and HAB algal biomass [135]. In contrast to findings of coastal South Carolina, USA, tidal creeks and nearby retention waterbodies, where increases in either or both Vv and Vp were associated with harmful cyanobacterial blooms and harmful eukaryotic algal blooms [115], there were no such associations in a partner study in Puget Sound of Washington state, USA, a more temperate environment [170]. Notably, correlations of Vp with HABs are more pronounced with higher SST. In the Puget Sound, there were clear blooms of three harmful species during the study period (Pseudo-nitzschia, Alexandrium and Dinophysis), but these events did not coincide with increases in Vp [170]. Also in this study, the only water column nutrient significantly associated with Vp was silicate, unlike the findings of many other studies suggesting strong correlations between Vv and Vp abundances with nitrogen- and phosphorous-containing nutrients (see Section 4.2).
The known association of Vv and Vp in the “particle-associated” fraction (often meaning associated with eukaryotic plankton) also emphasizes the need for size fractionation of water column samples prior to screening for the bacteria. This type of size fractionation is rare, but is becoming more common in environmental studies such as [131], resulting in insightful conclusions about the Vibrio ecology considerations when planning and modeling to improve aquaculture safety. Indeed, the composition of the plankton community may dictate the relative success of Vv and Vp [131] or bacteria in the Vibrio genus [143]. Based on a study using only culture-independent methods, relative abundances of planktonic Vibrio are found to be higher in the FL fraction as compared to PA (>3 micron), but the qPCR primers used were only able to enumerate Vibrio at the genus level [142].

4.3.3. Seagrass Density

Coastal areas with high water movement can facilitate Vibrio transport and distribution. As recently examined with culture-dependent tools [128], it has long been known that Vibrio are enriched in sediments, where their abundances can exceed water counts by a log or more, and temporal fluctuations of Vv and Vp abundances in sediments are pronounced and tied closely to SST [171]. Wind intensity, changes in direction, and changes in wind patterns may mix sedimentary Vibrio into the water column, increasing chances of human encounters or bringing more into contact with oyster and other shellfish. Fortunately, it is possible that these effects may be mitigated by submerged aquatic vegetation that stabilize sediments. In temperate eelgrass (Zostera marina) meadows, all Vibrio and Vv abundances were reduced, based on culture-dependent measurements [138].
Similarly, a negative correlation of Vibrio plus other pathogens was seen with other types of seagrasses in tropical coasts of Indonesia based on deep sequencing of the 16S rRNA gene [139]. Unfortunately, inferring the relationship of grass density with strain- and species-specific Vibrio abundances is not possible with this method, and to date, the number of seagrass density studies is limited. Additionally, it has recently been suggested that surfaces of macroalgae or seagrasses could harbor higher Vibrio densities than in environs absent of macrophytes [172]. The promise of decreased pathogen loads in waterbodies with high-density seagrass beds, however, is likely due to the sediment stabilization by healthy seagrasses.

4.3.4. Other Anthropogenic Factors

In addition to all the environmental factors described above, introduction of microplastics and heavy metals into waterbodies near aquaculture facilities may also increase abundances. For example, Vp enumerated with trh qPCR was positively correlated with heavy metals in water samples, but the impact of metals was also seen for general bacteria, E. coli, Pseudomonas aeruginosa, among other groups [134]. Additionally, growth and biofilm formation by both pathogenic and non-pathogenic types of Vv are positively correlated with increasing iron concentrations [173]. Introduction of heavy metals, antibiotics, or both, may affect the rate or mechanism ofARG transfer on transposable elements [174]. There are several recent general reviews on antibiotic resistance genes in Vp and Vv, including [81,175,176].
Extracellular polysaccharide matrix formation by Vv and Vp leads to fitness on surfaces, both sessile and planktonic [160,169]. Therefore, it is not surprising that microplastics loads in the water column are a risk factor for enhanced Vibrio abundances. Vibrio bacteria are also enriched in the plastisphere. A survey from the North and Baltic Seas reported that Vibrio-attached microplastics tend to be more brittle, perhaps due to degradation by high winds and storm events [177]. These degraded microplastics may increase surface area and create greater attachment sites for these particle-associated microbes. More general molecular tools for enumerating bacterial families, including Vibrioaceae, were previously used in the context of fluorescence in situ hybridization (FISH) to follow succession of colonizers on plastic surfaces; Vibrio are important in early biofilm colonization [178].
In northern estuarine waters, Vibrio have different responses to various artificially introduced substrates, where wood or plastics result in the expression of different groups of genes, resulting in different proteomes of the different substrates’ bacteria [179]. Colonizing sargassum and microplastics in subtropical waters of the Caribbean and Sargasso Seas all apparently rely on the expression of the pilus-related MSHA operon, including the mshA gene. The MSHA is necessary for adhesion and biofilm formation, regardless of substrate [180]. As extensively reviewed [181], sessile Vibrio abundances on these artificial surfaces are similar to those in naturally occurring substrates, both in fixed environments as well as on planktonic surfaces such as floating particles or larger plankton.

4.3.5. All Factors

Historically, several basic water quality parameters such as temperature and salinity were thought to have strong influence on pathogenic Vibrio dynamics, partly due to their known optimal ranges. As described in detail in Section 4.1 through Section 4.3, myriad examples exist demonstrating a combination of factors, both biotic and abiotic, that correlate very well with increased Vv and Vp abundances. Wind-mixing of sediments, seagrass bed densities, and other climate-related factors can influence both Vibrio distribution, as well as the distribution of other planktonic groups such as zooplankton and phytoplankton, now known to be the biotic predictors of Vibrio blooms. Importantly, it is now clear that multivariate analyses are needed to predict Vv and Vp abundances’ correlations with meterological, biological and abiotic environmental factors [39,182], since oceanographic and hydrographic perturbations affect all trophic levels, including those to which Vibrio may be attached. In addition, interpretation of correlative relationships should also consider the enumerating tools used, which are overviewed in Section 5 (below).

5. Future Outlooks, Advancements, and New Molecular Tools

Ecological or environmental parameter predictions, to date, are based on various enumeration methods, each of which has its own set of limitations. Limitations of enumeration methods could include sensitivity, PCR primer bias, or underestimations due to “cultivability” of the environmental strains of interest. Common to many types of naturally occurring bacteria, the “Viable but Not-Culturable” (VBNC) state can confound models. Bacteria in the VBNC state may cause us to miss those “overwintering” Vibrio from sediments, or Vibrio associated with oysters, when using culture-based methods [32,183]. The in situ water temperature also dictates culture growth rates, in particular environmental Vv from cooler samples [184]. Although the VBNC state, metabolic survival and pathogenesis responses of these bacteria have been extensively reviewed elsewhere [26,185,186,187], most current environmental monitoring types of studies still do not consider the impact of inclusion of culture-based methods on Vibrio abundance underestimates. Modeling based on data from most studies using partially culture-based methods described in Table 2, such as liquid culture-based (MPN-) or solid medium (DP-) culturing are at risk of underestimating Vv and Vp that are possibly pathogenic but resist growth on our current culture media.
Abundances gleaned from metagenomics studies of samples associated with enteric pathogens can be as reliable as, or at least consistent with, culture-based methods. For screening environmental samples, tying strain taxonomy with important functional genes such as individual ARGs or multi-drug resistance transporters is possible with metagenomics [188]. Other more specialized types of molecular-only types of screening tools are available, such as LAMP-PCR-based assays for Vp, can be used, but these methods are limiting due to cost, extensive pretreatment needs, or other problems [189]. Care should be taken with interpretations of studies using tools enumerating Vibrio at the genus level, or even at the species level without considering relative species abundances of pathogenic and non-pathogenic species [18].
Sensitivity of current assays, both those that include a culture-dependent step, as well as those that bypass enrichments, is generally not a problem. For example, methods in DP studies or others that use an enrichment step prior to PCR/qPCR, all have very low detection limits, ranging from ~<1 [21] to 5–15 colony forming unit (CFU) per mL of seawater [54,128]. From oysters, DP methods can detect as low as 5–10 CFU per gram of oyster homogenate sample [121,190]. Multi-step assays that rely on liquid culture-based enrichment steps first, prior to PCR or qPCR, are even more sensitive, since the enrichment allows for growth prior to detection. Ranges of sensitivity for these types of studies are quite variable. For example, in a variety of substrates from both subtropical and temperate locales, the lower detection limit of either Vv or Vp in all substrates tested was 1 to >80,000 CFU per unit of substrate tested, specifically 1–250 CFU/mL (water), 10–25,000 CFU/g (oyster), and 100 to 83,333 CFU/g (sediment) [101].
PCR primer bias could also lead to underestimation of true Vibrio abundances. For example, most of the recent studies employing PCR-based verification use PCR primers designed and published over a decade ago (“Gene” column in Table 2). One exception is a recent study that developed newer functional gene assays, including for one for sodB and related genes, which may assist Vv in transitioning from the environment into a virulent state associated with its host [74]. PCR primer biases may result from direct PCR (or methods that use a culture-based step first followed by PCR or qPCR), since primer pairs (Table 1) may not match well to all representative strains’ sequences. PCR primer bias does not create an issue with sensitivity, rather the community composition of pathogenic strains of Vibrio may not be accurately represented with use of outdated primer sequences. To remedy this, continual development of qPCR primer pairs should be carried out, refining them to be inclusive of more recently submitted sequences from known pathogenic strains.
However, there is hope; examples of more efficient direct analyses of DNA are becoming more prevalent, such as direct qPCR [191], and metagenomics [192], where both of these types of methods by-pass the culture-dependent step completely. Sensitivity of qPCR-only studies is, like with methods containing a culture step, very good. For example, Vp detection is as low as 1 to 10 copies per mL of seawater with trh primers [134]; with these and other primers for Vp and for other Vibrio species, a similar range is detectable with the very sensitive qPCR method [137]. Detection of Vv with direct qPCR is also very sensitive; cell equivalents in DNA enumerated with vvhA qPCR is as low as 15 CFU per mg of homogenized fish tissue [54]. PCR primer bias can be overcome with the use of PCR-independent detection of functional genes in genome segments of the entire microbial consortia. For instance, in metagenomics, sensitivity may be a problem, but the benefits gleaned (new sequences of functional genes obtained without a PCR step) may lead to improved PCR and qPCR primer pairs [192].
Increases in vibriosis associated with finfish aquaculture has been reviewed extensively elsewhere [193]. A recent systematic review of foodborne-Vibrio evaluated temperature and other non-environmental parameters provides insights into predictions of Vibrio distributions in finfish aquaculture [194]. Although specific serovars can affect the health of aquaculture species such as eels and other teleosts [195], the scope of this review is limited to human pathogens among Vv and Vp, and thus, particular attention is paid to increasing Vibrio abundances in shellfish [190]. Oysters, often eaten raw, are well-known to harbor high abundances of both Vv and Vp, in their liquor, on their shells, and/or in the meat. Seafood-borne infections are dominated by these two species as well as V. cholerae [64]. Oyster-derived Vp outbreaks in northern US environments on both Atlantic and Pacific coasts and in the Gulf of Mexico region [196], as well as Vv outbreaks in a variety of locales [197] indicate the pervasive nature of these naturally occurring species.
Numerous earlier, and all subsequent, studies (too many to list, but many of which are listed with “Oyster” in the Substrate column of Table 2) emphasize the need for good screening methods to improve aquaculture efforts. Importantly, there may be differential influences in the changes in salinity, TSS, or even water depth on the loads of pathogenic Vv in oysters and clams, as well as an increased chance of finding pathogenic Vp in cooler, rather than warmer, waters [38]. These studies have demonstrated how using genomics and molecular methods to call for and/or enumerate these genes can distinguish virulent strains of Vibrio in environmental samples.
Genes of interest for many studies focused on those in pathogenic Vv and Vp, including tdh, trh, and vvhA. A 1999 study successfully defined and developed a 5′ nuclease probe-based qPCR method for quantifying Vp in oyster tissue homogenate [59]. When compared to positive detection in pure culture strains, trh, but not tdh, was detected in one strain, suggesting that assaying for both genes may be necessary to detect all hemolysin-producing strains. Some published primer pairs also exist for detection and enumeration of Vv with qPCR forgoing the use of probes [56]. Such a design relies solely on SYBR Green I fluorescence in qPCR for the detection of amplicon production during the course of the PCR, a more cost-effective approach than probe-dependent assays. Additionally, careful PCR primer design may improve detection since TaqMan probes add a layer of specificity, reducing the ability to include degeneracy codes to better capture the full diversity of pathogenic strains’ gene sequences [198].
Pathogenicity-associated target genes may appear to be present in species other than Vp. For example, trh, a target gene that was only reported to be present in Vp [199] was found in what was later identified as V. alginolyticus, a near-neighbor species, suggesting a positive PCR result could be associated with presence of pathogenic Vp [200]. The lineage of a newly evolved pathogenic Vp strain expanded its distribution along the North American Atlantic coast. It was dubbed as a new “sequence type”, number 631 (ST631) exhibiting a similar virulence gene profile to members of a related ST36 clade, a group of Vp known to contain both tdh and trh [201]. Methods of detecting such highly adaptable pathogens need to be attuned often to accommodate the effects of a rapidly changing environment.
Increasing global temperatures due to climate change portends an increase in frequency and/or intensity of heavy rainstorms, which could promote Vibrio abundance. However, there are now better predictive tools such as machine learning modeling of hydrodynamic factors influencing particle movement [202], along with multivariate analyses to predict Vv and Vp abundances’ correlations with meteorological, biological and abiotic environmental factors [39,182]. There is promise for improved predictive models without sampling oyster tissues; this is especially true in aquaculture environments with low tidal range. Recently described in a study of water column and oyster abundances, the predictive patterns with fluctuations in environmental parameters may be determined by water column abundances alone [110], potentially allowing for time and cost-savings. Finally, the knowledge that healthy densities of submerged aquatic vegetative material can possibly mitigate some of these problems is a good start; further studies on seagrass beds and associated water column, sediment and shellfish-associated Vibrio are needed. These improved models and additional information on the potential for seagrass mitigation may assist local, regional, and other policy-makers in planning improvements for aquaculture facilities and human recreation. Methodological shifts in measurement approaches, away from those that depend on culturing as a first step, and towards evaluation of DNA or RNA extracted directly from the substrate will also allow for more rapid detection, accurate quantification, and a more comprehensive characterization of the types of bacteria present.

Author Contributions

L.A.W. and T.V.P. contributed equally to the literature review, interpretation and writing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to thank the two reviewers who provided insightful and detailed feedback of an earlier version of the manuscript, resulting in significant improvements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Coia, J.; Cubie, H. Vibrio Species. In The Immunoassay Kit Directory: Volume 1: Part 3 December 1995; Coia, J., Cubie, H., Eds.; The Immunoassay Kit Directory; Springer Netherlands: Dordrecht, The Netherlands, 1995; pp. 975–976. ISBN 978-94-009-0359-3. [Google Scholar]
  2. Drake, S.L.; DePaola, A.; Jaykus, L.-A. An Overview of Vibrio Vulnificus and Vibrio Parahaemolyticus. Compr. Rev. Food Sci. Food Saf. 2007, 6, 120–144. [Google Scholar] [CrossRef]
  3. Heng, S.-P.; Letchumanan, V.; Deng, C.-Y.; Ab Mutalib, N.-S.; Khan, T.M.; Chuah, L.-H.; Chan, K.-G.; Goh, B.-H.; Pusparajah, P.; Lee, L.-H. Vibrio Vulnificus: An Environmental and Clinical Burden. Front. Microbiol. 2017, 8, 997. [Google Scholar] [CrossRef] [PubMed]
  4. Baker-Austin, C.; Oliver, J.D. Vibrio Vulnificus. Trends Microbiol. 2020, 28, 81–82. [Google Scholar] [CrossRef]
  5. Bross, M.H.; Soch, K.; Morales, R.; Mitchell, R.B. Vibrio Vulnificus Infection: Diagnosis and Treatment. Am. Fam. Physician 2007, 76, 539–544. [Google Scholar] [PubMed]
  6. Froelich, B.; Oliver, J.D. The Interactions of Vibrio Vulnificus and the Oyster Crassostrea Virginica. Microb. Ecol. 2013, 65, 807–816. [Google Scholar] [CrossRef] [PubMed]
  7. Mead, P.S.; Slutsker, L.; Dietz, V.; McCaig, L.F.; Bresee, J.S.; Shapiro, C.; Griffin, P.M.; Tauxe, R.V. Food-Related Illness and Death in the United States. Emerg. Infect. Dis. 1999, 5, 607–625. [Google Scholar] [CrossRef] [PubMed]
  8. Letchumanan, V.; Chan, K.-G.; Lee, L.-H. Vibrio Parahaemolyticus: A Review on the Pathogenesis, Prevalence, and Advance Molecular Identification Techniques. Front. Microbiol. 2014, 5, 705. [Google Scholar] [CrossRef]
  9. Gode-Potratz, C.J.; Kustusch, R.J.; Breheny, P.J.; Weiss, D.S.; McCarter, L.L. Surface Sensing in Vibrio Parahaemolyticus Triggers a Programme of Gene Expression That Promotes Colonization and Virulence. Mol. Microbiol. 2011, 79, 240–263. [Google Scholar] [CrossRef] [PubMed]
  10. Phippen, B.L.; Oliver, J.D. Clinical and Environmental Genotypes of Vibrio Vulnificus Display Distinct, Quorum-Sensing-Mediated, Chitin Detachment Dynamics. Pathog. Dis. 2015, 73, ftv072. [Google Scholar] [CrossRef]
  11. Balcázar, J.L.; de Blas, I.; Ruiz-Zarzuela, I.; Cunningham, D.; Vendrell, D.; Múzquiz, J.L. The Role of Probiotics in Aquaculture. Vet. Microbiol. 2006, 114, 173–186. [Google Scholar] [CrossRef]
  12. Blackwell, K.D.; Oliver, J.D. The Ecology of Vibrio Vulnificus, Vibrio Cholerae, and Vibrio Parahaemolyticus in North Carolina Estuaries. J. Microbiol. 2008, 46, 146–153. [Google Scholar] [CrossRef] [PubMed]
  13. Newton, A.; Kendall, M.; Vugia, D.J.; Henao, O.L.; Mahon, B.E. Increasing Rates of Vibriosis in the United States, 1996–2010: Review of Surveillance Data From 2 Systems. Clin. Infect. Dis. 2012, 54, S391–S395. [Google Scholar] [CrossRef] [PubMed]
  14. Pazhani, G.P.; Bhowmik, S.K.; Ghosh, S.; Guin, S.; Dutta, S.; Rajendran, K.; Saha, D.R.; Nandy, R.K.; Bhattacharya, M.K.; Mukhopadhyay, A.K.; et al. Trends in the Epidemiology of Pandemic and Non-Pandemic Strains of Vibrio Parahaemolyticus Isolated from Diarrheal Patients in Kolkata, India. PLoS Negl. Trop. Dis. 2014, 8, e2815. [Google Scholar] [CrossRef]
  15. Park, M.S.; Park, K.H.; Bahk, G.J. Interrelationships between Multiple Climatic Factors and Incidence of Foodborne Diseases. Int. J. Environ. Res. Public. Health 2018, 15, 2482. [Google Scholar] [CrossRef] [PubMed]
  16. Malham, S.K.; Rajko-Nenow, P.; Howlett, E.; Tuson, E.K.; Perkins, T.L.; Pallett, D.W.; Wang, H.; Jago, C.F.; Jones, D.L.; McDonald, J.E. The Interaction of Human Microbial Pathogens, Particulate Material and Nutrients in Estuarine Environments and Their Impacts on Recreational and Shellfish Waters. Environ. Sci. Process. Impacts 2014, 16, 2145–2155. [Google Scholar] [CrossRef]
  17. Esteves, K.; Hervio-Heath, D.; Mosser, T.; Rodier, C.; Tournoud, M.-G.; Jumas-Bilak, E.; Colwell, R.R.; Monfort, P. Rapid Proliferation of Vibrio Parahaemolyticus, Vibrio Vulnificus, and Vibrio Cholerae during Freshwater Flash Floods in French Mediterranean Coastal Lagoons. Appl. Environ. Microbiol. 2015, 81, 7600–7609. [Google Scholar] [CrossRef]
  18. Froelich, B.; Gonzalez, R.; Blackwood, D.; Lauer, K.; Noble, R. Decadal Monitoring Reveals an Increase in Vibrio Spp. Concentrations in the Neuse River Estuary, North Carolina, USA. PLoS ONE 2019, 14, e0215254. [Google Scholar] [CrossRef]
  19. Froelich, B.A.; Daines, D.A. In Hot Water: Effects of Climate Change on Vibrio–Human Interactions. Environ. Microbiol. 2020, 22, 4101–4111. [Google Scholar] [CrossRef]
  20. Elmahdi, S.; Parveen, S.; Ossai, S.; DaSilva, L.V.; Jahncke, M.; Bowers, J.; Jacobs, J. Vibrio Parahaemolyticus and Vibrio Vulnificus Recovered from Oysters during an Oyster Relay Study. Appl. Environ. Microbiol. 2018, 84, e01790-17. [Google Scholar] [CrossRef]
  21. Noorian, P.; Hoque, M.M.; Espinoza-Vergara, G.; McDougald, D. Environmental Reservoirs of Pathogenic Vibrio Spp. and Their Role in Disease: The List Keeps Expanding. Adv. Exp. Med. Biol. 2023, 1404, 99–126. [Google Scholar] [CrossRef]
  22. Baker-Austin, C.; Oliver, J.D. Vibrio Vulnificus: New Insights into a Deadly Opportunistic Pathogen. Environ. Microbiol. 2018, 20, 423–430. [Google Scholar] [CrossRef]
  23. Ayrapetyan, M.; Williams, T.; Oliver, J.D. Relationship between the Viable but Nonculturable State and Antibiotic Persister Cells. J. Bacteriol. 2018, 200, e00249-18. [Google Scholar] [CrossRef]
  24. Sony, M.; Sumithra, T.G.; Anusree, V.N.; Amala, P.V.; Reshma, K.J.; Alex, S.; Sanil, N.K. Antimicrobial Resistance and Virulence Characteristics of Vibrio Vulnificus, Vibrio Parahaemolyticus and Vibrio Harveyi from Natural Disease Outbreaks of Marine/Estuarine Fishes. Aquaculture 2021, 539, 736608. [Google Scholar] [CrossRef]
  25. Wagley, S.; Morcrette, H.; Kovacs-Simon, A.; Yang, Z.R.; Power, A.; Tennant, R.K.; Love, J.; Murray, N.; Titball, R.W.; Butler, C.S. Bacterial Dormancy: A Subpopulation of Viable but Non-Culturable Cells Demonstrates Better Fitness for Revival. PLoS Pathog. 2021, 17, e1009194. [Google Scholar] [CrossRef] [PubMed]
  26. Gao, R.; Liao, X.; Zhao, X.; Liu, D.; Ding, T. The Diagnostic Tools for Viable but Nonculturable Pathogens in the Food Industry: Current Status and Future Prospects. Compr. Rev. Food Sci. Food Saf. 2021, 20, 2146–2175. [Google Scholar] [CrossRef] [PubMed]
  27. Oliver, J.D.; Bockian, R. In Vivo Resuscitation, and Virulence towards Mice, of Viable but Nonculturable Cells of Vibrio Vulnificus. Appl. Environ. Microbiol. 1995, 61, 2620–2623. [Google Scholar] [CrossRef]
  28. Franco, S.L.M.; Swenson, G.J.; Long, R.A. Year Round Patchiness of Vibrio Vulnificus within a Temperate Texas Bay. J. Appl. Microbiol. 2012, 112, 593–604. [Google Scholar] [CrossRef]
  29. Wong, H.C.; Wang, P. Induction of Viable but Nonculturable State in Vibrio Parahaemolyticus and Its Susceptibility to Environmental Stresses. J. Appl. Microbiol. 2004, 96, 359–366. [Google Scholar] [CrossRef]
  30. Coutard, F.; Crassous, P.; Droguet, M.; Gobin, E.; Colwell, R.R.; Pommepuy, M.; Hervio-Heath, D. Recovery in Culture of Viable but Nonculturable Vibrio Parahaemolyticus: Regrowth or Resuscitation? ISME J. 2007, 1, 111–120. [Google Scholar] [CrossRef]
  31. Furr, D.; Ketchum, R.N.; Phippen, B.L.; Reitzel, A.M.; Ivanina, A.V. Physiological Variation in Response to Vibrio and Hypoxia by Aquacultured Eastern Oysters in the Southeastern United States. Integr. Comp. Biol. 2021, 61, 1715–1729. [Google Scholar] [CrossRef]
  32. Motes, M.L.; DePaola, A.; Cook, D.W.; Veazey, J.E.; Hunsucker, J.C.; Garthright, W.E.; Blodgett, R.J.; Chirtel, S.J. Influence of Water Temperature and Salinity onVibrio Vulnificus in Northern Gulf and Atlantic Coast Oysters (Crassostrea Virginica). Appl. Environ. Microbiol. 1998, 64, 1459–1465. [Google Scholar] [CrossRef] [PubMed]
  33. Wright, A.C.; Hill, R.T.; Johnson, J.A.; Roghman, M.C.; Colwell, R.R.; Morris, J.G. Distribution of Vibrio Vulnificus in the Chesapeake Bay. Appl. Environ. Microbiol. 1996, 62, 717–724. [Google Scholar] [CrossRef]
  34. Campbell, M.S.; Wright, A.C. Real-Time PCR Analysis of Vibrio Vulnificus from Oysters. Appl. Environ. Microbiol. 2003, 69, 7137–7144. [Google Scholar] [CrossRef]
  35. Picard, C.; Ponsonnet, C.; Paget, E.; Nesme, X.; Simonet, P. Detection and Enumeration of Bacteria in Soil by Direct DNA Extraction and Polymerase Chain Reaction. Appl. Environ. Microbiol. 1992, 58, 2717–2722. [Google Scholar] [CrossRef]
  36. Mäntynen, V.; Niemelä, S.; Kaijalainen, S.; Pirhonen, T.; Lindström, K. MPN-PCR—Quantification Method for Staphylococcal Enterotoxin c 1 Gene from Fresh Cheese. Int. J. Food Microbiol. 1997, 36, 135–143. [Google Scholar] [CrossRef] [PubMed]
  37. Sharma, S.; Radl, V.; Hai, B.; Kloos, K.; Mrkonjic Fuka, M.; Engel, M.; Schauss, K.; Schloter, M. Quantification of Functional Genes from Procaryotes in Soil by PCR. J. Microbiol. Methods 2007, 68, 445–452. [Google Scholar] [CrossRef] [PubMed]
  38. Williams, T.C.; Froelich, B.A.; Phippen, B.; Fowler, P.; Noble, R.T.; Oliver, J.D. Different Abundance and Correlational Patterns Exist between Total and Presumed Pathogenic Vibrio Vulnificus and V. Parahaemolyticus in Shellfish and Waters along the North Carolina Coast. FEMS Microbiol. Ecol. 2017, 93, fix071. [Google Scholar] [CrossRef]
  39. Diner, R.E.; Kaul, D.; Rabines, A.; Zheng, H.; Steele, J.A.; Griffith, J.F.; Allen, A.E. Pathogenic Vibrio Species Are Associated with Distinct Environmental Niches and Planktonic Taxa in Southern California (USA) Aquatic Microbiomes. mSystems 2021, 6, e00571-21. [Google Scholar] [CrossRef]
  40. Hirsch, P.R.; Mauchline, T.H.; Clark, I.M. Culture-Independent Molecular Techniques for Soil Microbial Ecology. Soil Biol. Biochem. 2010, 42, 878–887. [Google Scholar] [CrossRef]
  41. Botes, M.; de Kwaadsteniet, M.; Cloete, T.E. Application of Quantitative PCR for the Detection of Microorganisms in Water. Anal. Bioanal. Chem. 2013, 405, 91–108. [Google Scholar] [CrossRef]
  42. Li, J.; Zhu, Y.; Wu, X.; Hoffmann, M.R. Rapid Detection Methods for Bacterial Pathogens in Ambient Waters at the Point of Sample Collection: A Brief Review. Clin. Infect. Dis. 2020, 71, S84–S90. [Google Scholar] [CrossRef]
  43. Dalmasso, A.; La Neve, F.; Suffredini, E.; Croci, L.; Serracca, L.; Bottero, M.T.; Civera, T. Development of a PCR Assay Targeting the rpoA Gene for the Screening of Vibrio Genus. Food Anal. Methods 2009, 2, 317–324. [Google Scholar] [CrossRef]
  44. Main, C.R.; Salvitti, L.R.; Whereat, E.B.; Coyne, K.J. Community-Level and Species-Specific Associations between Phytoplankton and Particle-Associated Vibrio Species in Delaware’s Inland Bays. Appl. Environ. Microbiol. 2015, 81, 5703–5713. [Google Scholar] [CrossRef]
  45. Thompson, J.R.; Randa, M.A.; Marcelino, L.A.; Tomita-Mitchell, A.; Lim, E.; Polz, M.F. Diversity and Dynamics of a North Atlantic Coastal Vibrio Community. Appl. Environ. Microbiol. 2004, 70, 4103–4110. [Google Scholar] [CrossRef]
  46. Randa, M.A.; Polz, M.F.; Lim, E. Effects of Temperature and Salinity on Vibrio Vulnificus Population Dynamics as Assessed by Quantitative PCR. Appl. Environ. Microbiol. 2004, 70, 5469–5476. [Google Scholar] [CrossRef] [PubMed]
  47. Stauder, M.; Huq, A.; Pezzati, E.; Grim, C.J.; Ramoino, P.; Pane, L.; Colwell, R.R.; Pruzzo, C.; Vezzulli, L. Role of GbpA Protein, an Important Virulence-Related Colonization Factor, for Vibrio Cholerae’s Survival in the Aquatic Environment. Environ. Microbiol. Rep. 2012, 4, 439–445. [Google Scholar] [CrossRef]
  48. Jang, K.K.; Gil, S.Y.; Lim, J.G.; Choi, S.H. Regulatory Characteristics of Vibrio Vulnificus gbpA Gene Encoding a Mucin-Binding Protein Essential for Pathogenesis. J. Biol. Chem. 2016, 291, 5774. [Google Scholar] [CrossRef]
  49. Honda, T.; Chearskul, S.; Takeda, Y.; Miwatani, T. Immunological Methods for Detection of Kanagawa Phenomenon of Vibrio Parahaemolyticus. J. Clin. Microbiol. 1980, 11, 600. [Google Scholar] [CrossRef] [PubMed]
  50. Hongping, W.; Jilun, Z.; Ting, J.; Yixi, B.; Xiaoming, Z. Insufficiency of the Kanagawa Hemolytic Test for Detecting Pathogenic Vibrio Parahaemolyticus in Shanghai, China. Diagn. Microbiol. Infect. Dis. 2011, 69, 7–11. [Google Scholar] [CrossRef] [PubMed]
  51. Yuan, Y.; Feng, Z.; Wang, J. Vibrio Vulnificus Hemolysin: Biological Activity, Regulation of vvhA Expression, and Role in Pathogenesis. Front. Immunol. 2020, 11, 599439. [Google Scholar] [CrossRef]
  52. Rosche, T.M.; Yano, Y.; Oliver, J.D. A Rapid and Simple PCR Analysis Indicates There Are Two Subgroups of Vibrio Vulnificus Which Correlate with Clinical or Environmental Isolation. Microbiol. Immunol. 2005, 49, 381–389. [Google Scholar] [CrossRef]
  53. Warner, E.B.; Oliver, J.D. Multiplex PCR Assay for Detection and Simultaneous Differentiation of Genotypes of Vibrio Vulnificus Biotype 1. Foodborne Pathog. Dis. 2008, 5, 691–693. [Google Scholar] [CrossRef]
  54. Sanjuán, E.; Amaro, C. Multiplex PCR Assay for Detection of Vibrio Vulnificus Biotype 2 and Simultaneous Discrimination of Serovar E Strains. Appl. Environ. Microbiol. 2007, 73, 2029–2032. [Google Scholar] [CrossRef]
  55. Wetz, J.J.; Blackwood, A.D.; Fries, J.S.; Williams, Z.F.; Noble, R.T. Trends in Total Vibrio Spp. and Vibrio Vulnificus Concentrations in the Eutrophic Neuse River Estuary, North Carolina, during Storm Events. Aquat. Microb. Ecol. 2008, 53, 141–149. [Google Scholar] [CrossRef]
  56. Panicker, G.; Myers, M.L.; Bej, A.K. Rapid Detection of Vibrio Vulnificus in Shellfish and Gulf of Mexico Water by Real-Time PCR. Appl. Environ. Microbiol. 2004, 70, 498–507. [Google Scholar] [CrossRef] [PubMed]
  57. Panicker, G.; Vickery, M.C.L.; Bej, A.K. Multiplex PCR Detection of Clinical and Environmental Strains of Vibrio Vulnificus in Shellfish. Can. J. Microbiol. 2004, 50, 911–922. [Google Scholar] [CrossRef] [PubMed]
  58. Gutierrez West, C.K.; Klein, S.L.; Lovell, C.R. High Frequency of Virulence Factor Genes Tdh, Trh, and Tlh in Vibrio Parahaemolyticus Strains Isolated from a Pristine Estuary. Appl. Environ. Microbiol. 2013, 79, 2247–2252. [Google Scholar] [CrossRef] [PubMed]
  59. Bej, A.K.; Patterson, D.P.; Brasher, C.W.; Vickery, M.C.; Jones, D.D.; Kaysner, C.A. Detection of Total and Hemolysin-Producing Vibrio Parahaemolyticus in Shellfish Using Multiplex PCR Amplification of Tl, Tdh and Trh. J. Microbiol. Methods 1999, 36, 215–225. [Google Scholar] [CrossRef]
  60. Kinsey, T.P.; Lydon, K.A.; Bowers, J.C.; Jones, J.L. Effects of Dry Storage and Resubmersion of Oysters on Total Vibrio Vulnificus and Total and Pathogenic (Tdh+/Trh+) Vibrio Parahaemolyticus Levels. J. Food Prot. 2015, 78, 1574–1580. [Google Scholar] [CrossRef]
  61. Nordstrom, J.L.; Vickery, M.C.L.; Blackstone, G.M.; Murray, S.L.; DePaola, A. Development of a Multiplex Real-Time PCR Assay with an Internal Amplification Control for the Detection of Total and Pathogenic Vibrio Parahaemolyticus Bacteria in Oysters. Appl. Environ. Microbiol. 2007, 73, 5840–5847. [Google Scholar] [CrossRef]
  62. Lin, Z.; Kumagai, K.; Baba, K.; Mekalanos, J.J.; Nishibuchi, M. Vibrio Parahaemolyticus Has a Homolog of the Vibrio Cholerae toxRS Operon That Mediates Environmentally Induced Regulation of the Thermostable Direct Hemolysin Gene. J. Bacteriol. 1993, 175, 3844–3855. [Google Scholar] [CrossRef]
  63. Lee, S.E.; Shin, S.H.; Kim, S.Y.; Kim, Y.R.; Shin, D.H.; Chung, S.S.; Lee, Z.H.; Lee, J.Y.; Jeong, K.C.; Choi, S.H.; et al. Vibrio Vulnificus Has the Transmembrane Transcription Activator ToxRS Stimulating the Expression of the Hemolysin Gene vvhA. J. Bacteriol. 2000, 182, 3405–3415. [Google Scholar] [CrossRef] [PubMed]
  64. Bauer, A.; Rørvik, L.M. A Novel Multiplex PCR for the Identification of Vibrio Parahaemolyticus, Vibrio Cholerae and Vibrio Vulnificus. Lett. Appl. Microbiol. 2007, 45, 371–375. [Google Scholar] [CrossRef]
  65. Luan, X.; Chen, J.; Zhang, X.-H.; Li, Y.; Hu, G. Expression and Characterization of a Metalloprotease from a Vibrio Parahaemolyticus Isolate. Can. J. Microbiol. 2007, 53, 1168–1173. [Google Scholar] [CrossRef]
  66. Paranjpye, R.N.; Strom, M.S. A Vibrio Vulnificus Type IV Pilin Contributes to Biofilm Formation, Adherence to Epithelial Cells, and Virulence. Infect. Immun. 2005, 73, 1411–1422. [Google Scholar] [CrossRef] [PubMed]
  67. Gauthier, J.D.; Jones, M.K.; Thiaville, P.; Joseph, J.L.; Swain, R.A.; Krediet, C.J.; Gulig, P.A.; Teplitski, M.; Wright, A.C. Role of GacA in Virulence of Vibrio Vulnificus. Microbiol. Read. Engl. 2010, 156, 3722–3733. [Google Scholar] [CrossRef]
  68. Paranjpye, R.; Hamel, O.S.; Stojanovski, A.; Liermann, M. Genetic Diversity of Clinical and Environmental Vibrio Parahaemolyticus Strains from the Pacific Northwest. Appl. Environ. Microbiol. 2012, 78, 8631–8638. [Google Scholar] [CrossRef]
  69. Chattopadhyay, S.; Paranjpye, R.N.; Dykhuizen, D.E.; Sokurenko, E.V.; Strom, M.S. Comparative Evolutionary Analysis of the Major Structural Subunit of Vibrio Vulnificus Type IV Pili. Mol. Biol. Evol. 2009, 26, 2185–2196. [Google Scholar] [CrossRef] [PubMed]
  70. Roig, F.J.; Sanjuán, E.; Llorens, A.; Amaro, C. pilF Polymorphism-Based PCR To Distinguish Vibrio Vulnificus Strains Potentially Dangerous to Public Health. Appl. Environ. Microbiol. 2010, 76, 1328–1333. [Google Scholar] [CrossRef]
  71. Roig, A.P.; Carmona-Salido, H.; Sanjuán, E.; Fouz, B.; Amaro, C. A Multiplex PCR for the Detection of Vibrio Vulnificus Hazardous to Human and/or Animal Health from Seafood. Int. J. Food Microbiol. 2022, 377, 109778. [Google Scholar] [CrossRef]
  72. Hwang, S.-H.; Im, H.; Choi, S.H. A Master Regulator BrpR Coordinates the Expression of Multiple Loci for Robust Biofilm and Rugose Colony Development in Vibrio Vulnificus. Front. Microbiol. 2021, 12, 679854. [Google Scholar] [CrossRef]
  73. Pajuelo, D.; Lee, C.-T.; Roig, F.J.; Lemos, M.L.; Hor, L.-I.; Amaro, C. Host-Nonspecific Iron Acquisition Systems and Virulence in the Zoonotic Serovar of Vibrio Vulnificus. Infect. Immun. 2014, 82, 731–744. [Google Scholar] [CrossRef] [PubMed]
  74. Conrad, J.W.; Harwood, V.J. Sewage Promotes Vibrio Vulnificus Growth and Alters Gene Transcription in Vibrio Vulnificus CMCP6. Microbiol. Spectr. 2022, 10, e0191321. [Google Scholar] [CrossRef]
  75. Kang, I.-H.; Kim, J.-S.; Lee, J.K. The Virulence of Vibrio Vulnificus Is Affected by the Cellular Level of Superoxide Dismutase Activity. J. Microbiol. Biotechnol. 2007, 17, 1399–1402. [Google Scholar] [PubMed]
  76. Kim, C.-M.; Shin, S.-H. Vibrio Vulnificus Quorum-Sensing System Operates in Cirrhotic Ascites, a Human Ex Vivo Experimental System. Ann. Microbiol. 2013, 63, 403–408. [Google Scholar] [CrossRef]
  77. Lee, H.-J.; Kim, J.-A.; Lee, M.-A.; Park, S.-J.; Lee, K.-H. Regulation of Haemolysin (VvhA) Production by Ferric Uptake Regulator (Fur) in Vibrio Vulnificus: Repression of vvhA Transcription by Fur and Proteolysis of VvhA by Fur-Repressive Exoproteases. Mol. Microbiol. 2013, 88, 813–826. [Google Scholar] [CrossRef]
  78. Kim, B.S.; Gavin, H.E.; Satchell, K.J.F. Distinct Roles of the Repeat-Containing Regions and Effector Domains of the Vibrio Vulnificus Multifunctional-Autoprocessing Repeats-in-Toxin (MARTX) Toxin. mBio 2015, 6, e00324-15. [Google Scholar] [CrossRef]
  79. Kim, B.S. The Modes of Action of MARTX Toxin Effector Domains. Toxins 2018, 10, 507. [Google Scholar] [CrossRef]
  80. Jang, K.K.; Lee, Z.-W.; Kim, B.; Jung, Y.H.; Han, H.J.; Kim, M.H.; Kim, B.S.; Choi, S.H. Identification and Characterization of Vibrio Vulnificus plpA Encoding a Phospholipase A2 Essential for Pathogenesis. J. Biol. Chem. 2017, 292, 17129–17143. [Google Scholar] [CrossRef] [PubMed]
  81. Choi, G.; Choi, S.H. Complex Regulatory Networks of Virulence Factors in Vibrio Vulnificus. Trends Microbiol. 2022, 30, 1205–1216. [Google Scholar] [CrossRef]
  82. Williams, T.C.; Ayrapetyan, M.; Ryan, H.; Oliver, J.D. Serum Survival of Vibrio Vulnificus: Role of Genotype, Capsule, Complement, Clinical Origin, and in Situ Incubation. Pathog. Basel Switz. 2014, 3, 822–832. [Google Scholar] [CrossRef] [PubMed]
  83. Goo, S.Y.; Lee, H.-J.; Kim, W.H.; Han, K.-L.; Park, D.-K.; Lee, H.-J.; Kim, S.M.; Kim, K.-S.; Lee, K.-H.; Park, S.-J. Identification of OmpU of Vibrio Vulnificus as a Fibronectin-Binding Protein and Its Role in Bacterial Pathogenesis. Infect. Immun. 2006, 74, 5586. [Google Scholar] [CrossRef] [PubMed]
  84. Liu, M.; Alice, A.F.; Naka, H.; Crosa, J.H. The HlyU Protein Is a Positive Regulator of rtxA1, a Gene Responsible for Cytotoxicity and Virulence in the Human Pathogen Vibrio Vulnificus. Infect. Immun. 2007, 75, 3282–3289. [Google Scholar] [CrossRef]
  85. Getz, L.J.; Thomas, N.A. The Transcriptional Regulator HlyU Positively Regulates Expression of exsA, Leading to Type III Secretion System 1 Activation in Vibrio Parahaemolyticus. J. Bacteriol. 2018, 200, e00653-17. [Google Scholar] [CrossRef]
  86. Brown, R.N.; Gulig, P.A. Roles of RseB, sigmaE, and DegP in Virulence and Phase Variation of Colony Morphotype of Vibrio Vulnificus. Infect. Immun. 2009, 77, 3768–3781. [Google Scholar] [CrossRef] [PubMed]
  87. Chen, C.-Y.; Wu, K.-M.; Chang, Y.-C.; Chang, C.-H.; Tsai, H.-C.; Liao, T.-L.; Liu, Y.-M.; Chen, H.-J.; Shen, A.B.-T.; Li, J.-C.; et al. Comparative Genome Analysis of Vibrio Vulnificus, a Marine Pathogen. Genome Res. 2003, 13, 2577–2587. [Google Scholar] [CrossRef]
  88. Lee, A.; Kim, M.S.; Cho, D.; Jang, K.K.; Choi, S.H.; Kim, T.S. Vibrio Vulnificus RtxA Is a Major Factor Driving Inflammatory T Helper Type 17 Cell Responses in Vitro and in Vivo. Front. Immunol. 2018, 9, 2095. [Google Scholar] [CrossRef] [PubMed]
  89. Girard, L. Quorum Sensing in Vibrio Spp.: The Complexity of Multiple Signalling Molecules in Marine and Aquatic Environments. Crit. Rev. Microbiol. 2019, 45, 451–471. [Google Scholar] [CrossRef]
  90. Li, R.; Chiou, J.; Chan, E.W.-C.; Chen, S. A Novel PCR-Based Approach for Accurate Identification of Vibrio Parahaemolyticus. Front. Microbiol. 2016, 7, 44. [Google Scholar] [CrossRef]
  91. Furushita, M.; Shiba, T.; Maeda, T.; Yahata, M.; Kaneoka, A.; Takahashi, Y.; Torii, K.; Hasegawa, T.; Ohta, M. Similarity of Tetracycline Resistance Genes Isolated from Fish Farm Bacteria to Those from Clinical Isolates. Appl. Environ. Microbiol. 2003, 69, 5336–5342. [Google Scholar] [CrossRef]
  92. Srivastava, M.; Tucker, M.S.; Gulig, P.A.; Wright, A.C. Phase Variation, Capsular Polysaccharide, Pilus and Flagella Contribute to Uptake of Vibrio Vulnificus by the Eastern Oyster (Crassostrea Virginica). Environ. Microbiol. 2009, 11, 1934–1944. [Google Scholar] [CrossRef]
  93. Aagesen, A.M.; Phuvasate, S.; Su, Y.-C.; Häse, C.C. Persistence of Vibrio Parahaemolyticus in the Pacific Oyster, Crassostrea Gigas, Is a Multifactorial Process Involving Pili and Flagella but Not Type III Secretion Systems or Phase Variation. Appl. Environ. Microbiol. 2013, 79, 3303–3305. [Google Scholar] [CrossRef] [PubMed]
  94. Narayanan, S.V.; Joseph, T.C.; Peeralil, S.; Mothadaka, M.P.; Lalitha, K.V. Prevalence, Virulence Characterization, AMR Pattern and Genetic Relatedness of Vibrio Parahaemolyticus Isolates From Retail Seafood of Kerala, India. Front. Microbiol. 2020, 11, 592. [Google Scholar] [CrossRef] [PubMed]
  95. Ellingsen, A.B.; Olsen, J.S.; Granum, P.E.; Rørvik, L.M.; González-Escalona, N. Genetic Characterization of Trh Positive Vibrio Spp. Isolated from Norway. Front. Cell. Infect. Microbiol. 2013, 3, 107. [Google Scholar] [CrossRef]
  96. Williams, T.C.; Blackman, E.R.; Morrison, S.S.; Gibas, C.J.; Oliver, J.D. Transcriptome Sequencing Reveals the Virulence and Environmental Genetic Programs of Vibrio Vulnificus Exposed to Host and Estuarine Conditions. PLoS ONE 2014, 9, e114376. [Google Scholar] [CrossRef]
  97. Almuhaideb, E.; Chintapenta, L.K.; Abbott, A.; Parveen, S.; Ozbay, G. Assessment of Vibrio Parahaemolyticus Levels in Oysters (Crassostrea Virginica) and Seawater in Delaware Bay in Relation to Environmental Conditions and the Prevalence of Molecular Markers to Identify Pathogenic Vibrio Parahaemolyticus Strains. PLoS ONE 2020, 15, e0242229. [Google Scholar] [CrossRef]
  98. Feng, L.; Xu, M.; Zhu, J.; Lu, H. Genetic Basis of High-Pressure Tolerance of a Vibrio Parahaemolyticus Mutant and Its Pathogenicity. Front. Microbiol. 2022, 13, 827856. [Google Scholar] [CrossRef] [PubMed]
  99. Liang, S.; Zhang, T.; Liu, Z.; Wang, J.; Zhu, C.; Kong, Q.; Fu, X.; Mou, H. Response Mechanism of Vibrio Parahaemolyticus at High Pressure Revealed by Transcriptomic Analysis. Appl. Microbiol. Biotechnol. 2022, 106, 5615–5628. [Google Scholar] [CrossRef] [PubMed]
  100. Johnson, C.N.; Flowers, A.R.; Noriea, N.F.; Zimmerman, A.M.; Bowers, J.C.; DePaola, A.; Grimes, D.J. Relationships between Environmental Factors and Pathogenic Vibrios in the Northern Gulf of Mexico. Appl. Environ. Microbiol. 2010, 76, 7076–7084. [Google Scholar] [CrossRef] [PubMed]
  101. Johnson, C.N.; Bowers, J.C.; Griffitt, K.J.; Molina, V.; Clostio, R.W.; Pei, S.; Laws, E.; Paranjpye, R.N.; Strom, M.S.; Chen, A.; et al. Ecology of Vibrio Parahaemolyticus and Vibrio Vulnificus in the Coastal and Estuarine Waters of Louisiana, Maryland, Mississippi, and Washington (United States). Appl. Environ. Microbiol. 2012, 78, 7249–7257. [Google Scholar] [CrossRef]
  102. McCarthy, S.A.; DePaola, A.; Cook, D.W.; Kaysner, C.A.; Hill, W.E. Evaluation of Alkaline Phosphatase- and Digoxigenin-Labelled Probes for Detection of the Thermolabile Hemolysin (Tlh) Gene of Vibrio Parahaemolyticus. Lett. Appl. Microbiol. 1999, 28, 66–70. [Google Scholar] [CrossRef]
  103. DePaola, A.; Nordstrom, J.L.; Bowers, J.C.; Wells, J.G.; Cook, D.W. Seasonal Abundance of Total and Pathogenic Vibrio Parahaemolyticus in Alabama Oysters. Appl. Environ. Microbiol. 2003, 69, 1521–1526. [Google Scholar] [CrossRef] [PubMed]
  104. Nigro, O.D.; James-Davis, L.I.; De Carlo, E.H.; Li, Y.-H.; Steward, G.F. Variable Freshwater Influences on the Abundance of Vibrio Vulnificus in a Tropical Urban Estuary. Appl. Environ. Microbiol. 2022, 88, e0188421. [Google Scholar] [CrossRef] [PubMed]
  105. Baker-Austin, C.; Gore, A.; Oliver, J.D.; Rangdale, R.; McArthur, J.V.; Lees, D.N. Rapid in Situ Detection of Virulent Vibrio Vulnificus Strains in Raw Oyster Matrices Using Real-Time PCR. Environ. Microbiol. Rep. 2010, 2, 76–80. [Google Scholar] [CrossRef]
  106. Cruz, C.D.; Hedderley, D.; Fletcher, G.C. Long-Term Study of Vibrio Parahaemolyticus Prevalence and Distribution in New Zealand Shellfish. Appl. Environ. Microbiol. 2015, 81, 2320–2327. [Google Scholar] [CrossRef]
  107. Shirai, H.; Ito, H.; Hirayama, T.; Nakamoto, Y.; Nakabayashi, N.; Kumagai, K.; Takeda, Y.; Nishibuchi, M. Molecular Epidemiologic Evidence for Association of Thermostable Direct Hemolysin (TDH) and TDH-Related Hemolysin of Vibrio Parahaemolyticus with Gastroenteritis. Infect. Immun. 1990, 58, 3568–3573. [Google Scholar] [CrossRef]
  108. Tada, J.; Ohashi, T.; Nishimura, N.; Shirasaki, Y.; Ozaki, H.; Fukushima, S.; Takano, J.; Nishibuchi, M.; Takeda, Y. Detection of the Thermostable Direct Hemolysin Gene (Tdh) and the Thermostable Direct Hemolysin-Related Hemolysin Gene (Trh) of Vibrio Parahaemolyticus by Polymerase Chain Reaction. Mol. Cell. Probes 1992, 6, 477–487. [Google Scholar] [CrossRef]
  109. Chen, A.J.; Hasan, N.A.; Haley, B.J.; Taviani, E.; Tarnowski, M.; Brohawn, K.; Johnson, C.N.; Colwell, R.R.; Huq, A. Characterization of Pathogenic Vibrio Parahaemolyticus from the Chesapeake Bay, Maryland. Front. Microbiol. 2017, 8, 2460. [Google Scholar] [CrossRef]
  110. Audemard, C.; Ben-Horin, T.; Kator, H.I.; Reece, K.S. Vibrio Vulnificus and Vibrio Parahaemolyticus in Oysters under Low Tidal Range Conditions: Is Seawater Analysis Useful for Risk Assessment? Foods 2022, 11, 4065. [Google Scholar] [CrossRef]
  111. Baker-Austin, C.; Lemm, E.; Hartnell, R.; Lowther, J.; Onley, R.; Amaro, C.; Oliver, J.D.; Lees, D. pilF Polymorphism-Based Real-Time PCR to Distinguish Vibrio Vulnificus Strains of Human Health Relevance. Food Microbiol. 2012, 30, 17–23. [Google Scholar] [CrossRef] [PubMed]
  112. Brumfield, K.D.; Chen, A.J.; Gangwar, M.; Usmani, M.; Hasan, N.A.; Jutla, A.S.; Huq, A.; Colwell, R.R. Environmental Factors Influencing Occurrence of Vibrio Parahaemolyticus and Vibrio Vulnificus. Appl. Environ. Microbiol. 2023, 89, e00307-23. [Google Scholar] [CrossRef]
  113. Wright, A.C.; Miceli, G.A.; Landry, W.L.; Christy, J.B.; Watkins, W.D.; Morris, J.G. Rapid Identification of Vibrio Vulnificus on Nonselective Media with an Alkaline Phosphatase-Labeled Oligonucleotide Probe. Appl. Environ. Microbiol. 1993, 59, 541–546. [Google Scholar] [CrossRef]
  114. King, N.J.; Pirikahu, S.; Fletcher, G.C.; Pattis, I.; Roughan, B.; Merien, A.-M.P. Correlations between Environmental Conditions and Vibrio Parahaemolyticus or Vibrio Vulnificus in Pacific Oysters from New Zealand Coastal Waters. N. Z. J. Mar. Freshw. Res. 2020, 55, 393–410. [Google Scholar] [CrossRef]
  115. Greenfield, D.I.; Gooch Moore, J.; Stewart, J.R.; Hilborn, E.D.; George, B.J.; Li, Q.; Dickerson, J.; Keppler, C.K.; Sandifer, P.A. Temporal and Environmental Factors Driving Vibrio Vulnificus and V. Parahaemolyticus Populations and Their Associations with Harmful Algal Blooms in South Carolina Detention Ponds and Receiving Tidal Creeks. GeoHealth 2017, 1, 306–317. [Google Scholar] [CrossRef] [PubMed]
  116. Panicker, G.; Bej, A.K. Real-Time PCR Detection of Vibrio Vulnificus in Oysters: Comparison of Oligonucleotide Primers and Probes Targeting vvhA. Appl. Environ. Microbiol. 2005, 71, 5702–5709. [Google Scholar] [CrossRef] [PubMed]
  117. Anthony, R.M.; Brown, T.J.; French, G.L. Rapid Diagnosis of Bacteremia by Universal Amplification of 23S Ribosomal DNA Followed by Hybridization to an Oligonucleotide Array. J. Clin. Microbiol. 2000, 38, 781–788. [Google Scholar] [CrossRef] [PubMed]
  118. Canellas, A.L.B.; Lopes, I.R.; Mello, M.P.; Paranhos, R.; de Oliveira, B.F.R.; Laport, M.S. Vibrio Species in an Urban Tropical Estuary: Antimicrobial Susceptibility, Interaction with Environmental Parameters, and Possible Public Health Outcomes. Microorganisms 2021, 9, 1007. [Google Scholar] [CrossRef]
  119. Kim, Y.B.; Okuda, J.; Matsumoto, C.; Takahashi, N.; Hashimoto, S.; Nishibuchi, M. Identification of Vibrio Parahaemolyticus Strains at the Species Level by PCR Targeted to the toxR Gene. J. Clin. Microbiol. 1999, 37, 1173–1177. [Google Scholar] [CrossRef]
  120. Chase, E.; Young, S.; Harwood, V.J. Sediment and Vegetation as Reservoirs of Vibrio Vulnificus in the Tampa Bay Estuary and Gulf of Mexico. Appl. Environ. Microbiol. 2015, 81, 2489–2494. [Google Scholar] [CrossRef]
  121. Parveen, S.; Hettiarachchi, K.A.; Bowers, J.C.; Jones, J.L.; Tamplin, M.L.; McKay, R.; Beatty, W.; Brohawn, K.; Dasilva, L.V.; Depaola, A. Seasonal Distribution of Total and Pathogenic Vibrio Parahaemolyticus in Chesapeake Bay Oysters and Waters. Int. J. Food Microbiol. 2008, 128, 354–361. [Google Scholar] [CrossRef]
  122. Vezzulli, L.; Grande, C.; Reid, P.C.; Hélaouët, P.; Edwards, M.; Höfle, M.G.; Brettar, I.; Colwell, R.R.; Pruzzo, C. Climate Influence on Vibrio and Associated Human Diseases during the Past Half-Century in the Coastal North Atlantic. Proc. Natl. Acad. Sci. USA 2016, 113, E5062–E5071. [Google Scholar] [CrossRef]
  123. Deeb, R.; Tufford, D.; Scott, G.I.; Moore, J.G.; Dow, K. Impact of Climate Change on Vibrio Vulnificus Abundance and Exposure Risk. Estuaries Coasts 2018, 41, 2289–2303. [Google Scholar] [CrossRef] [PubMed]
  124. Machado, A.; Bordalo, A.A. Detection and Quantification of Vibrio Cholerae, Vibrio Parahaemolyticus, and Vibrio Vulnificus in Coastal Waters of Guinea-Bissau (West Africa). EcoHealth 2016, 13, 339–349. [Google Scholar] [CrossRef]
  125. Brasher, C.W.; DePaola, A.; Jones, D.D.; Bej, A.K. Detection of Microbial Pathogens in Shellfish with Multiplex PCR. Curr. Microbiol. 1998, 37, 101–107. [Google Scholar] [CrossRef]
  126. Roque, A.; Lopez-Joven, C.; Lacuesta, B.; Elandaloussi, L.; Wagley, S.; Furones, M.D.; Ruiz-Zarzuela, I.; de Blas, I.; Rangdale, R.; Gomez-Gil, B. Detection and Identification of Tdh- and Trh-Positive Vibrio Parahaemolyticus Strains from Four Species of Cultured Bivalve Molluscs on the Spanish Mediterranean Coast. Appl. Environ. Microbiol. 2009, 75, 7574–7577. [Google Scholar] [CrossRef]
  127. Givens, C.E.; Bowers, J.C.; DePaola, A.; Hollibaugh, J.T.; Jones, J.L. Occurrence and Distribution of Vibrio Vulnificus and Vibrio Parahaemolyticus—Potential Roles for Fish, Oyster, Sediment and Water. Lett. Appl. Microbiol. 2014, 58, 503–510. [Google Scholar] [CrossRef]
  128. Potdukhe, T.V.; Caffrey, J.M.; Rothfus, M.J.; Daniel, C.E.; Swords, M.E.; Albrecht, B.B.; Jeffrey, W.H.; Waidner, L.A. Viable Putative Vibrio Vulnificus and Parahaemolyticus in the Pensacola and Perdido Bays: Water Column, Sediments, and Invertebrate Biofilms. Front. Mar. Sci. 2021, 8, 645755. [Google Scholar] [CrossRef]
  129. Zimmerman, A.M.; DePaola, A.; Bowers, J.C.; Krantz, J.A.; Nordstrom, J.L.; Johnson, C.N.; Grimes, D.J. Variability of Total and Pathogenic Vibrio Parahaemolyticus Densities in Northern Gulf of Mexico Water and Oysters. Appl. Environ. Microbiol. 2007, 73, 7589–7596. [Google Scholar] [CrossRef]
  130. Hughes, S.N.; Greig, D.J.; Miller, W.A.; Byrne, B.A.; Gulland, F.M.D.; Harvey, J.T. Dynamics of Vibrio with Virulence Genes Detected in Pacific Harbor Seals (Phoca Vitulina Richardii) Off California: Implications for Marine Mammal Health. Microb. Ecol. 2013, 65, 982–994. [Google Scholar] [CrossRef] [PubMed]
  131. Morrison, B.H. Tracking Vibrio: Population Dynamics and Community Ecology in Alabama Estuaries. Ph.D. Thesis, University of South Alabama, Mobile, AL, USA, 2021. [Google Scholar]
  132. Shaw, K.S.; Jacobs, J.M.; Crump, B.C. Impact of Hurricane Irene on Vibrio Vulnificus and Vibrio Parahaemolyticus Concentrations in Surface Water, Sediment, and Cultured Oysters in the Chesapeake Bay, MD, USA. Front. Microbiol. 2014, 5, 204. [Google Scholar] [CrossRef]
  133. Rosales, D.; Ellett, A.; Jacobs, J.; Ozbay, G.; Parveen, S.; Pitula, J. Investigating the Relationship between Nitrate, Total Dissolved Nitrogen, and Phosphate with Abundance of Pathogenic Vibrios and Harmful Algal Blooms in Rehoboth Bay, Delaware. Appl. Environ. Microbiol. 2022, 88, e0035622. [Google Scholar] [CrossRef]
  134. Huang, H.; Zan, S.; Ming, H.; Li, M.; Xu, J.; Xie, Y.; Wang, W.; Fan, J. 16S Next-Generation Sequencing and Quantitative PCR Reveal the Distribution of Potential Pathogens in the Liaohe Estuary. Front. Mar. Sci. 2022, 9, 1044260. [Google Scholar] [CrossRef]
  135. Kim, D.D.; Wan, L.; Cao, X.; Klisarova, D.; Gerdzhikov, D.; Zhou, Y.; Song, C.; Yoon, S. Metagenomic Insights into Co-Proliferation of Vibrio Spp. and Dinoflagellates Prorocentrum during a Spring Algal Bloom in the Coastal East China Sea. Water Res. 2021, 204, 117625. [Google Scholar] [CrossRef]
  136. Jesser, K.J.; Noble, R.T. Vibrio Ecology in the Neuse River Estuary, North Carolina, Characterized by Next-Generation Amplicon Sequencing of the Gene Encoding Heat Shock Protein 60 (Hsp60). Appl. Environ. Microbiol. 2018, 84, e00333-18. [Google Scholar] [CrossRef]
  137. Asplund, M.E.; Rehnstam-Holm, A.-S.; Atnur, V.; Raghunath, P.; Saravanan, V.; Härnström, K.; Collin, B.; Karunasagar, I.; Godhe, A. Water Column Dynamics of Vibrio in Relation to Phytoplankton Community Composition and Environmental Conditions in a Tropical Coastal Area. Environ. Microbiol. 2011, 13, 2738–2751. [Google Scholar] [CrossRef]
  138. Reusch, T.B.H.; Schubert, P.R.; Marten, S.-M.; Gill, D.; Karez, R.; Busch, K.; Hentschel, U. Lower Vibrio Spp. Abundances in Zostera Marina Leaf Canopies Suggest a Novel Ecosystem Function for Temperate Seagrass Beds. Mar. Biol. 2021, 168, 149. [Google Scholar] [CrossRef]
  139. Lamb, J.B.; van de Water, J.A.J.M.; Bourne, D.G.; Altier, C.; Hein, M.Y.; Fiorenza, E.A.; Abu, N.; Jompa, J.; Harvell, C.D. Seagrass Ecosystems Reduce Exposure to Bacterial Pathogens of Humans, Fishes, and Invertebrates. Science 2017, 355, 731–733. [Google Scholar] [CrossRef] [PubMed]
  140. Staley, C.; Chase, E.; Harwood, V.J. Detection and Differentiation of Vibrio Vulnificus and V. Sinaloensis in Water and Oysters of a Gulf of Mexico Estuary. Environ. Microbiol. 2013, 15, 623–633. [Google Scholar] [CrossRef] [PubMed]
  141. Fleischmann, S.; Herrig, I.; Wesp, J.; Stiedl, J.; Reifferscheid, G.; Strauch, E.; Alter, T.; Brennholt, N. Prevalence and Distribution of Potentially Human Pathogenic Vibrio Spp. on German North and Baltic Sea Coasts. Front. Cell. Infect. Microbiol. 2022, 12, 846819. [Google Scholar] [CrossRef] [PubMed]
  142. Liang, J.; Liu, J.; Wang, X.; Lin, H.; Liu, J.; Zhou, S.; Sun, H.; Zhang, X.-H. Spatiotemporal Dynamics of Free-Living and Particle-Associated Vibrio Communities in the Northern Chinese Marginal Seas. Appl. Environ. Microbiol. 2019, 85, e00217-19. [Google Scholar] [CrossRef]
  143. Turner, J.W.; Good, B.; Cole, D.; Lipp, E.K. Plankton Composition and Environmental Factors Contribute to Vibrio Seasonality. ISME J. 2009, 3, 1082–1092. [Google Scholar] [CrossRef] [PubMed]
  144. Martinez-Urtaza, J.; Bowers, J.C.; Trinanes, J.; DePaola, A. Climate Anomalies and the Increasing Risk of Vibrio Parahaemolyticus and Vibrio Vulnificus Illnesses. Food Res. Int. 2010, 43, 1780–1790. [Google Scholar] [CrossRef]
  145. Baker-Austin, C.; Trinanes, J.A.; Taylor, N.G.H.; Hartnell, R.; Siitonen, A.; Martinez-Urtaza, J. Emerging Vibrio Risk at High Latitudes in Response to Ocean Warming. Nat. Clim. Chang. 2013, 3, 73–77. [Google Scholar] [CrossRef]
  146. Hounmanou, Y.M.G.; Engberg, J.; Bjerre, K.D.; Holt, H.M.; Olesen, B.; Voldstedlund, M.; Dalsgaard, A.; Ethelberg, S. Correlation of High Seawater Temperature with Vibrio and Shewanella Infections, Denmark, 2010–2018. Emerg. Infect. Dis. J. 2023, 29, 3. [Google Scholar] [CrossRef]
  147. Brumfield, K.D.; Usmani, M.; Chen, K.M.; Gangwar, M.; Jutla, A.S.; Huq, A.; Colwell, R.R. Environmental Parameters Associated with Incidence and Transmission of Pathogenic Vibrio Spp. Environ. Microbiol. 2021, 23, 7314–7340. [Google Scholar] [CrossRef] [PubMed]
  148. Kelly, M.T. Effect of Temperature and Salinity on Vibrio (Beneckea) Vulnificus Occurrence in a Gulf Coast Environment. Appl. Environ. Microbiol. 1982, 44, 820–824. [Google Scholar] [CrossRef]
  149. Wetz, M.S.; Yoskowitz, D.W. An “extreme” Future for Estuaries? Effects of Extreme Climatic Events on Estuarine Water Quality and Ecology. Mar. Pollut. Bull. 2013, 69, 7–18. [Google Scholar] [CrossRef]
  150. Froelich, B.A.; Noble, R.T. Factors Affecting the Uptake and Retention of Vibrio Vulnificus in Oysters. Appl. Environ. Microbiol. 2014, 80, 7454–7459. [Google Scholar] [CrossRef]
  151. Lipp, E.K.; Rodriguez-Palacios, C.; Rose, J.B. Occurrence and Distribution of the Human Pathogen Vibrio Vulnificus in a Subtropical Gulf of Mexico Estuary. In The Ecology and Etiology of Newly Emerging Marine Diseases; Porter, J.W., Ed.; Developments in Hydrobiology; Springer Netherlands: Dordrecht, The Netherlands, 2001; pp. 165–173. ISBN 978-94-017-3284-0. [Google Scholar]
  152. Froelich, B.A.; Williams, T.C.; Noble, R.T.; Oliver, J.D. Apparent Loss of Vibrio Vulnificus from North Carolina Oysters Coincides with a Drought-Induced Increase in Salinity. Appl. Environ. Microbiol. 2012, 78, 3885–3889. [Google Scholar] [CrossRef] [PubMed]
  153. Banakar, V.; Constantin de Magny, G.; Jacobs, J.; Murtugudde, R.; Huq, A.; Wood, R.J.; Colwell, R.R. Temporal and Spatial Variability in the Distribution of Vibrio Vulnificus in the Chesapeake Bay: A Hindcast Study. EcoHealth 2011, 8, 456–467. [Google Scholar] [CrossRef] [PubMed]
  154. Wetz, J.J.; Blackwood, A.D.; Fries, J.S.; Williams, Z.F.; Noble, R.T. Quantification of Vibrio Vulnificus in an Estuarine Environment: A Multi-Year Analysis Using QPCR. Estuaries Coasts 2014, 37, 421–435. [Google Scholar] [CrossRef]
  155. Frank, A.M.; Cains, M.G.; Henshel, D.S. A Predictive Human Health Risk Assessment of Non-Choleraic Vibrio Spp. during Hurricane-Driven Flooding Events in Coastal South Carolina, USA. Atmosphere 2021, 12, 269. [Google Scholar] [CrossRef]
  156. Krishna, K.; Veettil, V.P.; Anas, A.; Nair, S. Hydrological Regulation of Vibrio Dynamics in a Tropical Monsoonal Estuary: A Classification and Regression Tree Approach. Environ. Sci. Pollut. Res. 2020, 28, 724–737. [Google Scholar] [CrossRef]
  157. López-Pérez, M.; Jayakumar, J.M.; Grant, T.-A.; Zaragoza-Solas, A.; Cabello-Yeves, P.J.; Almagro-Moreno, S. Ecological Diversification Reveals Routes of Pathogen Emergence in Endemic Vibrio Vulnificus Populations. Proc. Natl. Acad. Sci. USA 2021, 118, e2103470118. [Google Scholar] [CrossRef]
  158. López-Pérez, M.; Jayakumar, J.M.; Haro-Moreno, J.M.; Zaragoza-Solas, A.; Reddi, G.; Rodriguez-Valera, F.; Shapiro, O.H.; Alam, M.; Almagro-Moreno, S. Evolutionary Model of Cluster Divergence of the Emergent Marine Pathogen Vibrio Vulnificus: From Genotype to Ecotype. mBio 2019, 10. [Google Scholar] [CrossRef]
  159. Ayala, A.J.; Kabengele, K.; Almagro-Moreno, S.; Ogbunugafor, C.B. Meteorological Associations of Vibrio Vulnificus Clinical Infections in Tropical Settings: Correlations with Air Pressure, Wind Speed, and Temperature. PLoS Negl. Trop. Dis. 2023, 17, e0011461. [Google Scholar] [CrossRef] [PubMed]
  160. Takemura, A.F.; Chien, D.M.; Polz, M.F. Associations and Dynamics of Vibrionaceae in the Environment, from the Genus to the Population Level. Front. Microbiol. 2014, 5, 38. [Google Scholar] [CrossRef]
  161. Kirchman, D.L. Processes in Microbial Ecology, 2nd ed.; Oxford University Press: New York, NY, USA, 2018; ISBN 978-0-19-878941-3. [Google Scholar]
  162. Balasubramanian, D.; López-Pérez, M.; Grant, T.-A.; Ogbunugafor, C.B.; Almagro-Moreno, S. Molecular Mechanisms and Drivers of Pathogen Emergence. Trends Microbiol. 2022, 30, 898–911. [Google Scholar] [CrossRef]
  163. Urquhart, E.A.; Jones, S.H.; Yu, J.W.; Schuster, B.M.; Marcinkiewicz, A.L.; Whistler, C.A.; Cooper, V.S. Environmental Conditions Associated with Elevated Vibrio Parahaemolyticus Concentrations in Great Bay Estuary, New Hampshire. PLoS ONE 2016, 11, e0155018. [Google Scholar] [CrossRef] [PubMed]
  164. Hsieh, J.L.; Fries, J.S.; Noble, R.T. Dynamics and Predictive Modelling of Vibrio Spp. in the Neuse River Estuary, North Carolina, USA. Environ. Microbiol. 2008, 10, 57–64. [Google Scholar] [CrossRef] [PubMed]
  165. Pfeffer, C.S.; Hite, M.F.; Oliver, J.D. Ecology of Vibrio Vulnificus in Estuarine Waters of Eastern North Carolina. Appl. Environ. Microbiol. 2003, 69, 3526–3531. [Google Scholar] [CrossRef] [PubMed]
  166. Rivera, S.; Lugo, T.; Hazen, T.C. Autecology of Vibrio Vulnificus and Vibrio Parahaemolyticus in Tropical Waters. Water Res. 1989, 23, 923–931. [Google Scholar] [CrossRef]
  167. Watkins, W.D.; Cabelli, V.J. Effect of Fecal Pollution on Vibrio Parahaemolyticus Densities in an Estuarine Environment. Appl. Environ. Microbiol. 1985, 49, 1307–1313. [Google Scholar] [CrossRef]
  168. Stabili, L.; Di Salvo, M.; Alifano, P.; Talà, A. An Integrative, Multiparametric Approach for the Comprehensive Assessment of Microbial Quality and Pollution in Aquaculture Systems. Microb. Ecol. 2022, 83, 271–283. [Google Scholar] [CrossRef]
  169. Johnson, C.N. Fitness Factors in Vibrios: A Mini-Review. Microb. Ecol. 2013, 65, 826–851. [Google Scholar] [CrossRef] [PubMed]
  170. Paranjpye, R.N.; Nilsson, W.B.; Liermann, M.; Hilborn, E.D.; George, B.J.; Li, Q.; Bill, B.D.; Trainer, V.L.; Strom, M.S.; Sandifer, P.A. Environmental Influences on the Seasonal Distribution of Vibrio Parahaemolyticus in the Pacific Northwest of the USA. FEMS Microbiol. Ecol. 2015, 91, fiv121. [Google Scholar] [CrossRef]
  171. Vezzulli, L.; Pezzati, E.; Moreno, M.; Fabiano, M.; Pane, L.; Pruzzo, C. VibrioSea Consortium Benthic Ecology of Vibrio Spp. and Pathogenic Vibrio Species in a Coastal Mediterranean Environment (La Spezia Gulf, Italy). Microb. Ecol. 2009, 58, 808–818. [Google Scholar] [CrossRef]
  172. Kalvaitienė, G.; Vaičiūtė, D.; Bučas, M.; Gyraitė, G.; Kataržytė, M. Macrophytes and Their Wrack as a Habitat for Faecal Indicator Bacteria and Vibrio in Coastal Marine Environments. Mar. Pollut. Bull. 2023, 194, 115325. [Google Scholar] [CrossRef]
  173. Çam, S.; Brinkmeyer, R. The Effects of Temperature, pH, and Iron on Biofilm Formation by Clinical versus Environmental Strains of Vibrio Vulnificus. Folia Microbiol. 2020, 65, 557–566. [Google Scholar] [CrossRef]
  174. Pavón, A.; Riquelme, D.; Jaña, V.; Iribarren, C.; Manzano, C.; Lopez-Joven, C.; Reyes-Cerpa, S.; Navarrete, P.; Pavez, L.; García, K. The High Risk of Bivalve Farming in Coastal Areas with Heavy Metal Pollution and Antibiotic-Resistant Bacteria: A Chilean Perspective. Front. Cell. Infect. Microbiol. 2022, 12, 377. [Google Scholar] [CrossRef]
  175. Dutta, D.; Kaushik, A.; Kumar, D.; Bag, S. Foodborne Pathogenic Vibrios: Antimicrobial Resistance. Front. Microbiol. 2021, 12, 638331. [Google Scholar]
  176. Stephen, J.; Lekshmi, M.; Ammini, P.; Kumar, S.H.; Varela, M.F. Membrane Efflux Pumps of Pathogenic Vibrio Species: Role in Antimicrobial Resistance and Virulence. Microorganisms 2022, 10, 382. [Google Scholar] [CrossRef]
  177. Kirstein, I.V.; Kirmizi, S.; Wichels, A.; Garin-Fernandez, A.; Erler, R.; Löder, M.; Gerdts, G. Dangerous Hitchhikers? Evidence for Potentially Pathogenic Vibrio Spp. on Microplastic Particles. Mar. Environ. Res. 2016, 120, 1–8. [Google Scholar] [CrossRef] [PubMed]
  178. Schlundt, C.; Mark Welch, J.L.; Knochel, A.M.; Zettler, E.R.; Amaral-Zettler, L.A. Spatial Structure in the “Plastisphere”: Molecular Resources for Imaging Microscopic Communities on Plastic Marine Debris. Mol. Ecol. Resour. 2020, 20, 620–634. [Google Scholar] [CrossRef] [PubMed]
  179. Oberbeckmann, S.; Bartosik, D.; Huang, S.; Werner, J.; Hirschfeld, C.; Wibberg, D.; Heiden, S.E.; Bunk, B.; Overmann, J.; Becher, D.; et al. Genomic and Proteomic Profiles of Biofilms on Microplastics Are Decoupled from Artificial Surface Properties. Environ. Microbiol. 2021, 23, 3099–3115. [Google Scholar] [CrossRef] [PubMed]
  180. Mincer, T.J.; Bos, R.P.; Zettler, E.R.; Zhao, S.; Asbun, A.A.; Orsi, W.D.; Guzzetta, V.S.; Amaral-Zettler, L.A. Sargasso Sea Vibrio Bacteria: Underexplored Potential Pathovars in a Perturbed Habitat. Water Res. 2023, 242, 120033. [Google Scholar] [CrossRef]
  181. Bowley, J.; Baker-Austin, C.; Porter, A.; Hartnell, R.; Lewis, C. Oceanic Hitchhikers—Assessing Pathogen Risks from Marine Microplastic. Trends Microbiol. 2021, 29, 107–116. [Google Scholar] [CrossRef]
  182. Campbell, A.M.; Hauton, C.; Baker-Austin, C.; van Aerle, R.; Martinez-Urtaza, J. An Integrated Eco-Evolutionary Framework to Predict Population-Level Responses of Climate-Sensitive Pathogens. Curr. Opin. Biotechnol. 2023, 80, 102898. [Google Scholar] [CrossRef]
  183. Oliver, J.D.; Hite, F.; McDougald, D.; Andon, N.L.; Simpson, L.M. Entry into, and Resuscitation from, the Viable but Nonculturable State by Vibrio Vulnificus in an Estuarine Environment. Appl. Environ. Microbiol. 1995, 61, 2624–2630. [Google Scholar] [CrossRef]
  184. Oliver, J.D.; Nilsson, L.; Kjelleberg, S. Formation of Nonculturable Vibrio Vulnificus Cells and Its Relationship to the Starvation State. Appl. Environ. Microbiol. 1991, 57, 2640–2644. [Google Scholar] [CrossRef]
  185. Yoon, J.-H.; Lee, S.-Y. Characteristics of Viable-but-Nonculturable Vibrio Parahaemolyticus Induced by Nutrient-Deficiency at Cold Temperature. Crit. Rev. Food Sci. Nutr. 2020, 60, 1302–1320. [Google Scholar] [CrossRef]
  186. Wagley, S. The Viable but Non-Culturable (VBNC) State in Vibrio Species: Why Studying the VBNC State Now Is More Exciting than Ever. In Vibrio spp. Infections; Almagro-Moreno, S., Pukatzki, S., Eds.; Advances in Experimental Medicine and Biology; Springer International Publishing: Cham, Switzerland, 2023; pp. 253–268. ISBN 978-3-031-22997-8. [Google Scholar]
  187. Pan, H.; Ren, Q. Wake Up! Resuscitation of Viable but Nonculturable Bacteria: Mechanism and Potential Application. Foods 2023, 12, 82. [Google Scholar] [CrossRef]
  188. Gabashvili, E.; Kobakhidze, S.; Chkhikvishvili, T.; Tabatadze, L.; Tsiklauri, R.; Dadiani, K.; Koulouris, S.; Kotetishvili, M. Metagenomic and Recombination Analyses of Antimicrobial Resistance Genes from Recreational Waters of Black Sea Coastal Areas and Other Marine Environments Unveil Extensive Evidence for Their Both Intrageneric and Intergeneric Transmission across Genetically Very Diverse Microbial Communities. Mar. Genom. 2022, 61, 100916. [Google Scholar] [CrossRef]
  189. Zhao, X.; Li, M.; Liu, Y. Microfluidic-Based Approaches for Foodborne Pathogen Detection. Microorganisms 2019, 7, 381. [Google Scholar] [CrossRef] [PubMed]
  190. Harrison, J.; Nelson, K.; Morcrette, H.; Morcrette, C.; Preston, J.; Helmer, L.; Titball, R.W.; Butler, C.S.; Wagley, S. The Increased Prevalence of Vibrio Species and the First Reporting of Vibrio Jasicida and Vibrio Rotiferianus at UK Shellfish Sites. Water Res. 2022, 211, 117942. [Google Scholar] [CrossRef]
  191. Di Salvo, E.; Panebianco, F.; Panebianco, A.; Ziino, G. Quantitative Detection of Viable but Nonculturable Vibrio Parahaemolyticus in Frozen Bivalve Molluscs. Foods 2023, 12, 2373. [Google Scholar] [CrossRef] [PubMed]
  192. Huang, A.D.; Luo, C.; Pena-Gonzalez, A.; Weigand, M.R.; Tarr, C.L.; Konstantinidis, K.T. Metagenomics of Two Severe Foodborne Outbreaks Provides Diagnostic Signatures and Signs of Coinfection Not Attainable by Traditional Methods. Appl. Environ. Microbiol. 2017, 83, e02577-16. [Google Scholar] [CrossRef]
  193. Sanches-Fernandes, G.M.M.; Sá-Correia, I.; Costa, R. Vibriosis Outbreaks in Aquaculture: Addressing Environmental and Public Health Concerns and Preventive Therapies Using Gilthead Seabream Farming as a Model System. Front. Microbiol. 2022, 13. [Google Scholar]
  194. Ma, J.-Y.; Zhu, X.-K.; Hu, R.-G.; Qi, Z.-Z.; Sun, W.-C.; Hao, Z.-P.; Cong, W.; Kang, Y.-H. A Systematic Review, Meta-Analysis and Meta-Regression of the Global Prevalence of Foodborne Vibrio Spp. Infection in Fishes: A Persistent Public Health Concern. Mar. Pollut. Bull. 2023, 187, 114521. [Google Scholar] [CrossRef]
  195. Amaro, C.; Sanjuán, E.; Fouz, B.; Pajuelo, D.; Lee, C.-T.; Hor, L.-I.; Barrera, R. The Fish Pathogen Vibrio Vulnificus Biotype 2: Epidemiology, Phylogeny, and Virulence Factors Involved in Warm-Water Vibriosis. Microbiol. Spectr. 2015, 3, 3. [Google Scholar] [CrossRef]
  196. DePaola, A.; Kaysner, C.A.; Bowers, J.; Cook, D.W. Environmental Investigations of Vibrio Parahaemolyticus in Oysters after Outbreaks in Washington, Texas, and New York (1997 and 1998). Appl. Environ. Microbiol. 2000, 66, 4649–4654. [Google Scholar] [CrossRef]
  197. Cook, D.W.; O’leary, P.; Hunsucker, J.C.; Sloan, E.M.; Bowers, J.C.; Blodgett, R.J.; Depaola, A. Vibrio Vulnificus and Vibrio Parahaemolyticus in U.S. Retail Shell Oysters: A National Survey from June 1998 to July 1999. J. Food Prot. 2002, 65, 79–87. [Google Scholar] [CrossRef] [PubMed]
  198. Malinen, E.; Kassinen, A.; Rinttilä, T.; Palva, A. Comparison of Real-Time PCR with SYBR Green I or 5′-Nuclease Assays and Dot-Blot Hybridization with rDNA-Targeted Oligonucleotide Probes in Quantification of Selected Faecal Bacteria. Microbiol. Read. Engl. 2003, 149, 269–277. [Google Scholar] [CrossRef] [PubMed]
  199. Lida, T.; Park, K.-S.; Suthienkul, O.; Kozawa, J.; Yamaichi, Y.; Yamamoto, K.; Honda, T. Close Proximity of the Tdh, Trh and Ure Genes on the Chromosome of Vibrio Parahaemolyticus. Microbiol. Read. Engl. 1998, 144 Pt 9, 2517–2523. [Google Scholar] [CrossRef]
  200. González-Escalona, N.; Blackstone, G.M.; DePaola, A. Characterization of a Vibrio Alginolyticus Strain, Isolated from Alaskan Oysters, Carrying a Hemolysin Gene Similar to the Thermostable Direct Hemolysin-Related Hemolysin Gene (Trh) of Vibrio Parahaemolyticus. Appl. Environ. Microbiol. 2006, 72, 7925–7929. [Google Scholar] [CrossRef] [PubMed]
  201. Xu, F.; Gonzalez-Escalona, N.; Haendiges, J.; Myers, R.A.; Ferguson, J.; Stiles, T.; Hickey, E.; Moore, M.; Hickey, J.M.; Schillaci, C.; et al. Sequence Type 631 Vibrio Parahaemolyticus, an Emerging Foodborne Pathogen in North America. J. Clin. Microbiol. 2017, 55, 645–648. [Google Scholar] [CrossRef] [PubMed]
  202. Stachurska, B.; Mahdavi-Meymand, A.; Sulisz, W. Machine Learning Methodology for Determination of Sediment Particle Velocities over Sandy and Rippled Bed. Measurement 2022, 197, 111332. [Google Scholar] [CrossRef]
Table 2. Known environmental factors associated with changes in Vv and Vp abundances. Studies listed show strong positive or negative correlations of Vibrio species with factors listed (Env. Column). Method: tools used to enumerate Vv or Vp are indicated (if PCR and/or qPCR primers were used, the original reference for the primer sequences is provided). Quantification methods for studies examining correlating environmental factors included: MPN (most probable number), DP (direct plating), CH (colony hybridization), PCR (polymerase chain reaction) or qPCR (quantitative PCR), BOX-A1R-based repetitive extragenic palindromic-PCR (BOX-PCR), or simply “culture”, indicating liquid culture methods. CHROMAgar™ is a specialized direct plating method and is listed separately from DP. Some studies used only metagenomics and/or deep sequencing of amplicon libraries of a single gene, where the genes were either 16S rRNA or other phylogenetic markers. Genes specific to Vv and/or Vp are listed in Table 1. Positive, negative, or other non-linear relationships of environmental factors with either or both Vv and Vp species are indicated, where a check mark indicates the effected Vibrio species, and a check in parentheses indicates that species-level confirmation is not possible due to the limitations of the enumerating tool used.
Table 2. Known environmental factors associated with changes in Vv and Vp abundances. Studies listed show strong positive or negative correlations of Vibrio species with factors listed (Env. Column). Method: tools used to enumerate Vv or Vp are indicated (if PCR and/or qPCR primers were used, the original reference for the primer sequences is provided). Quantification methods for studies examining correlating environmental factors included: MPN (most probable number), DP (direct plating), CH (colony hybridization), PCR (polymerase chain reaction) or qPCR (quantitative PCR), BOX-A1R-based repetitive extragenic palindromic-PCR (BOX-PCR), or simply “culture”, indicating liquid culture methods. CHROMAgar™ is a specialized direct plating method and is listed separately from DP. Some studies used only metagenomics and/or deep sequencing of amplicon libraries of a single gene, where the genes were either 16S rRNA or other phylogenetic markers. Genes specific to Vv and/or Vp are listed in Table 1. Positive, negative, or other non-linear relationships of environmental factors with either or both Vv and Vp species are indicated, where a check mark indicates the effected Vibrio species, and a check in parentheses indicates that species-level confirmation is not possible due to the limitations of the enumerating tool used.
StudyAsso. 1Env. 2VvVpSubstrateMethod 3Gene(s) 4
Oceanographic, hydrographic, meteorological
[12]PositiveTemperatureSurface watersDP/CH, conf. PCRtlh [59]
PositiveSediments
[100]PositiveSurface watersDP/CH, MPN-qPCRtdh, trh, tlh [61]
[101]PositiveSurface watersDP/CH, MPN-qPCRtdh, trh, tlh [61] vvhA [102]
PositiveSediments
Positive Oysters
[103]Negative OystersDP/CH, conf. PCRtdh [59]
[104]Positive Surface watersDP, qPCRvcgC [105]; vvhA [34]
[55]Positive Surface watersMPN-qPCRvvhA [55]
[106]Positive OystersMPN-PCRtdh [107]; trh [108]
[109]Negative Surface watersMPN-PCRtlh, trh, tdh [59]
[97]Positive Surface watersMPN-PCRtlh, tdh, trh, toxR, vpm [61,65]
Positive Oysters
[110]PositiveSurface watersMPN-qPCRvvhA [34]; pilF [111]; trh, tdh, tlh [61]
[112]PositiveSurface watersDP/CH, conf. PCRtdh [102]; trh [61]; vvhA [113]
[114]Positive OystersN/A; meta-analysis of oyster surveys and previous publications
[115]Positive Surface watersDP/CH, conf. PCRtdh/trh [61]; vvhA [116]
[46]Positive Surface watersMPN-qPCR23S rRNA gene nested with Vv-specific 23S [117]
[118]Positive Surface watersBOX-PCRToxR [119]
[120]Positive Surface watersqPCR, culturevvhA [34]
Positive Oysters
Positive Vegetation
[121]Positive OystersDP/CH, MPN-qPCRtdh, trh, tlh [61]
[122]Positive(✓)(✓)Plankton fraction (phytoplankton and zooplankton)qPCRVibrio genus 16S rRNA gene [45]
[123]Positive Surface watersDP/PCR, qPCRvvhA [116]
[124]NegativeModerate to low salinitiesSurface watersMPN-qPCRtdh, trh [59]; vvhA [125]
[104]Nonlinear Surface watersMPN-qPCRvcgC [105]
[110]Negative Surface watersMPN-qPCRvvhA [34]
Negative Oysters
[55]Negative Surface watersMPN-qPCRvvhA [55]
[110]ConvexOystersMPN, qPCRtdh, trh, tlh [61]; vvhA [34]
[100]bimodal Surface watersDP/CH, MPN-qPCRvvhA [102]; tdh, trh, tlh [61]
Oysterstdh, trh, tlh [61]
Sedimentstdh, trh, tlh [61]
[126]Positive ShellfishDP/CHtdh, trh [59]
[112]NonlinearSurface watersDP/CH, conf. PCRtdh [102]; trh [61]; vvhA [113]
[115]NegativeSurface watersDP/CH, conf. PCRtdh, trh, tlh [61]; vvhA [116]
[46]Positive Surface watersMPN-qPCR23S rRNA gene nested with Vv-specific 23S [117]
[127]NegativeFish intestinesqPCR, DP-CH Vibrio-specific 16S rRNA [45]
[103]Bimodal OystersDP/CH, conf. PCRtdh [59]
[118]Negative Surface watersBOX-PCRToxR [119]
[120]Positive WaterqPCR, culturevvhA [34]
Positive Sediments
[128]Negative Surface watersDP, conf. PCRtdh, trh, tlh [61]
[123]Positive Surface watersDP/PCR, qPCRvvhA [116]
[100]Positive Surface watersDP/CH, MPN-qPCRtdh, trh, tlh [61]
Positive Oysters
Positive
Positive Sediments
Positive
Positive
[110]Negative Surface watersMPN-qPCRpilF [111]; vvhA [34]
[129]Positive Surface watersDP/CH, MPN-qPCRtdh, trh, tlh [61]
Positive Oysters
[130]Negative SealsDP/CH, conf. PCRtdh, trh [59]
[121]Positive OystersDP/CH, MPN-qPCRtdh [61]
[123]Positive Surface watersDP/PCR, qPCRvvhA [116]
[128]Positive7.0 > pH < 9.0 5 Bottom watersDP, conf. PCRvvhA [116]; tdh, trh, tlh [61]
Negative Surface, bottom waters
[115]Negative Surface watersDP/CH, conf. PCRtdh, trh, tlh [61]
[131]NegativeSurface watersMPN-qPCRtdh, trh, tlh [61]
[128]PositiveTidal influence 6 Surface watersDP/CH, conf. PCRvvhA [116]; tdh, trh, tlh [61]
Positive Sediments
[132]NegativeTurbulence 7 SedimentsMPN-PCRvcgC [105]; tdh, trh, tlh [61]
Negative
Negative Surface waters
[110]Negative OystersMPN-qPCRpilF [111]; vvhA [34]
[128]PositiveSurface watersDP, conf. PCRvvhA [116]
[131]NegativeSurface watersMPN-qPCRvvhA [34]; vvhA [34]; tdh, trh, tlh [61]
[129]PositiveTurbidity Surface watersDP/CH, MPN-qPCRtdh, trh, tlh [61]
Positive Oysters
[131]PositiveSurface watersMPN-qPCRvvhA [34]; tdh, trh, tlh [61]
[129]Positive Surface watersDP/CH, MPN-qPCRtdh, trh, tlh [61]
Positive Oysters
Abiotic factors that may be indicative of biological factors
[100]PositiveChlorophyllSurface watersMPN-qPCRtdh, trh, tlh [61]
PositiveSediments
[110]Positive Surface watersMPN-qPCRtdh, trh, tlh [61]
[112]PositiveSurface watersDP/CH, conf. PCRtdh [102]; trh [61]; vvhA [113]
Nutrients 8
[133]PositiveTDN Surface watersMPN-PCRfor Vv, vvhA [116], and vcg [105]; for Vp, tdh, trh, tlh [61]
[115]PositiveTDP Surface watersDP/CH, conf. PCRtdh, trh [61]; vvhA [116]
PositiveTDP
PositivePO4
PositiveDIN
PositivePO4
PositiveSi
PositiveTDN
PositiveDON
PositiveTP
PositiveTDP
[118]NegativeTN Surface watersBOX-PCRToxR [119]
NegativeTP
[128]PositiveDIN Surface watersDP, conf. PCRtdh, trh, tlh [61]
PositiveTKN
[134]NegativePO43− Surface watersqPCRtdh, trh [61]
NegativeNO3
NegativeNO2
NegativeNH4+
Biological factors
[124]PositiveVibrio choleraeSurface watersMPN-PCRtdh, trh [59]; vvhA [125]
[12]Positive Surface watersDP/CH, conf. PCRtdh, trh [59]
[124]PositiveFIB 9 Surface watersMPN-PCRtdh, trh [59]
Positive Surface watersvvhA [125]
[12]PositiveSurface watersDP/CH, conf. PCRtdh, trh [59]
[74]Positive Surface watersqPCRvvhA [34]
Positive MesocosmRT-qPCRsodB [74]
Other organisms 10
[115]PositiveMulti-species harmful cyanobacteria bloom 11Surface watersDP/CH, conf. PCRtdh, trh, tlh [61]; vvhA [116]
PositiveMulti-species HAB
PositiveH. rotundata, non-HAB euglenophyte Eutreptiella spp. 12
[44]PositiveHeterosigma spp., total dinoflagellates(✓)(✓)Surface watersqPCRrpoA [43]
[133]PositiveH. akashiwo, G. instriatum 13 Surface watersMPN-PCRtdh, trh, tlh [61]
vvhA [116]; vcgC [105]
[135]PositiveProrocentrum(✓)(✓)Surface watersMetagenomics, 16S deep seq.Phylogenetic markers, full genome
[136]PositiveDinoflagellates 14(✓)(✓)Seawaterhsp60, 16S deep seq.Phylogenetic markers
NegativeCyanobacterial bloom 14(✓)
Positive (✓)
[131]PositiveA. sanguinea, Heterocapsa spp. 15 Surface watersMPN-qPCRvvhA [34]; tdh, trh, tlh [61]
[137]PositiveDiatom-dominated phytoplanktonbloom Surface watersqPCRtdh, trh, tlh [61]
[138]NegativeSeagrass Δ 16(✓)(✓)Surface watersCHROMagarN/A
[139]Negative(✓)(✓)16S deep seq.Phylogenetic markers
1 Association, or relationship, with indicated environmental factor. 2 Correlated environmental factor. 3 Quantification method used; Conf. PCR: confirmed with PCR; deep seq.: deep sequencing. 4 Gene(s) detected with probes or amplified via PCR or qPCR. Primer pair or probe name and reference provided. 5 Reported pH conditions between 7.0 and 9.0. 6 Factors directly associated with tidal influence, including tidal range, tidal coefficient, or salinity stratification. 7 External physical factors related to resuspension of the water column, including wind direction, wind speeds, and storm events. 8 Nutrient concentrations. TDN: total dissolved nitrogen. TDP: total dissolved phosphorus. DIN: dissolved inorganic nitrogen. Si: silicate. DON: dissolved organic nitrogen. TP: total phosphorus. TN: total nitrogen. TKN: total Kjeldahl nitrogen. 9 Fecal indicator bacteria (fecal enterococci, total coliforms). 10 Other microorganisms including cyanobacterial harmful algal blooms (HABs), eukaryotic phytoplankton HABs, zooplankton. 11 Species include Cylindrospermopsis raciborskii, Anabaena spp. 12 Heterocapsa rotundata. 13 Heterosigma akashiwo, Gyrodinium instriatum. 14 following extreme storm event. 15 Akashiwo sanguinea. 16 Change in seagrass density.
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

Waidner, L.A.; Potdukhe, T.V. Tools to Enumerate and Predict Distribution Patterns of Environmental Vibrio vulnificus and Vibrio parahaemolyticus. Microorganisms 2023, 11, 2502. https://doi.org/10.3390/microorganisms11102502

AMA Style

Waidner LA, Potdukhe TV. Tools to Enumerate and Predict Distribution Patterns of Environmental Vibrio vulnificus and Vibrio parahaemolyticus. Microorganisms. 2023; 11(10):2502. https://doi.org/10.3390/microorganisms11102502

Chicago/Turabian Style

Waidner, Lisa A., and Trupti V. Potdukhe. 2023. "Tools to Enumerate and Predict Distribution Patterns of Environmental Vibrio vulnificus and Vibrio parahaemolyticus" Microorganisms 11, no. 10: 2502. https://doi.org/10.3390/microorganisms11102502

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