Skip to main content

Identification and whole-genome sequencing analysis of Vibrio vulnificus strains causing pearl gentian grouper disease in China

Abstract

Vibrio vulnificus is a pathogenic bacterium that causes disease in marine fish, affecting fish farming and human health worldwide. In May 2021, in the Bohai Bay region, a disease broke out in commercially farmed pearl gentian grouper (♀Epinephelus fuscoguttatus × ♂Epinephelus lanceolatus), causing huge economic losses. The diseased fish had skin lesions, water accumulation in their abdomens, and showed tissue and organ damage. V. vulnificus biotype 2 has been reported in eels and other marine fish, but it is less reported in pearl gentian grouper. In this study, the pathogenic strain isolated from diseased fish was identified as V. vulnificus EPL 0201 biotype 2 on the basis of physiological and biochemical characteristics and the results of 16S rRNA gene and gyrB sequencing, virulence gene detection, and recursive infection experiments. To gain a comprehensive understanding of the pathogenicity and drug resistance of this strain, whole-genome sequencing was performed. Whole-genome analysis showed that the gene map of this strain was complete. The Virulence Factor Database annotation results showed that this strain had the key virulence factor genes vvhA and rtxA, which cause host disease. In addition, this strain had genes conferring resistance against cephalosporins, aminoglycosides, tetracyclines, and sulfonamides. Antimicrobial susceptibility testing confirmed the presence of these resistance genes identified in the genome. The results of this study show that V. vulnificus EPL 0201 biotype 2 is a multi-drug resistant strain with high pathogenicity.

Peer Review reports

Introduction

Vibrio disease, also known as pathogenic hemorrhagic sepsis, is a bacterial disease caused by pathogenic Vibrio species, which affects a variety of aquatic animals, including fish, shellfish, shrimp, and crabs [1]. V. vulnificus belongs to the genus Vibrio and is a gram-negative, rod-shaped, halophilic opportunistic pathogen. It was first identified as halophilic Vibrio in 1976 [2] and was named V. vulnificus in 1979 [3]. V. vulnificus is an important pathogen of human beings and aquatic animals, and can be divided into three biotypes. V. vulnificus biotype 1 represents the first isolated subtype, which mainly infects humans [4]. The route of infection can be through direct contact with seawater, wounds, and the consumption of aquatic products, such as raw shellfish, carrying V. vulnificus [5, 6]. V. vulnificus infection can cause sepsis and seriously endanger human health [7]. This infection has also been reported in fish [8]. V. vulnificus biotype 2 was first identified in European eels (Anguilla anguilla) and was isolated from infected eels [9]. Diseases caused by V. vulnificus biotype 2 affect eels and other mariculture fish, especially in fish farms. In recent years, there have been many reports of diseases caused by V. vulnificus in marine fish, including grouper (Epinephelus coioides), half-smooth tongue sole (Cynoglossus semilaevis), large yellow croaker (Pseudosciaena crocea), mullet goby (Mugilogobius chulae), and gilthead seabream (Sparus aurata) [10,11,12,13,14]. Symptoms include skin lesions, and in severe cases, the disease may cause ulcers, abdominal effusion, enteritis, and tissue and organ lesions. The disease has a rapid onset and causes high mortality. Some researchers have suggested that all strains of V. vulnificus biotype 2 be considered pathogens with a special ability to cause Vibrio infections in fish, suggesting the name “Piscis” for this biotype [15]. V. vulnificus biotype 3 has only been reported in Israel and causes wound infection and bacteremia in humans [16].

In May 2021, a disease broke out in a pearl gentian grouper breeding pond in a marine farm in the Bohai Bay area and caused serious economic losses. The laboratory obtained dominant bacteria from the diseased pearl gentian grouper and named it EPL 0201. In this study, we isolated the pathogen; performed bioinformatics analysis, virulence gene testing, API 20 Bacterial Identification System analysis, recursive infection experiments, and whole-genome sequencing; and tested antimicrobial drug susceptibility to investigate the pathogenic and resistance mechanisms of the bacterium. Our findings provide the basis for further research on the pathogenicity of this bacterium and the development of strategies to prevent the spread of infection.

Material and methods

Diseased fish

Diseased and moribund pearl gentian groupers (body length 15–25 cm; body weight 50–150 g) were collected from a seawater factory farm in the Binhai New district, Tianjin City, China, where the motility of the cultured pearl gentian grouper was > 75%. The diseased fish were transferred to the Tianjin Agricultural University Key Lab of Aqua-Ecology and Aquaculture, and after anesthetization using MS-222 (Sigma-Aldrich, St. Louis, MO, USA), a camera (Nikon, Tokyo, Japan) was used to record their symptoms.

Isolation and identification of bacteria

Under sterile conditions, the gills, surface mucus, and diseased tissue of the diseased fish were observed under a microscope by wet mount, and no fungi and parasites were found. Subsequently, tissues from the spleen, kidney, and intestine of the diseased fish were taken under sterile conditions, pulverized with sterile PBS buffer, and then streaked on BHI fixed medium plates (Landbridge, Beijing, China), followed by culturing at 28 °C for 24 h. Dominant uniform bacterial colonies were observed on the culture plates after incubation. A single dominant bacterial colony was selected and inoculated on the same medium to obtain a pure isolate. The pure isolate thus obtained was tentatively named EPL 0201 was stored at − 80 °C in BHI broth containing sterile glycerol at a final concentration of 15% (v/v). This bacterial isolate was gram stained, and its morphology was observed using a light microscope (Leica, Wetzlar, Germany). The isolates were fixed in 2.5% electron microscope fixative solution for 2 h in the dark, following which the morphology was observed using a transmission electron microscope (Hitachi, Tokyo, Japan). Molecular identification of the isolate was performed via 16S rRNA gene and gyrase B (gyrB) sequencing [17, 18]. The 16S rRNA gene and gyrB amplification products were sent to Sangon Biotech (Shanghai, China) for sequencing. Nucleotide sequences were submitted to GenBank (https://www.ncbi.nlm.nih.gov/genbank/) [19]. Similarity analysis of the 16S rRNA gene and gyrB sequences was performed using the EzBioCloud database (https://www.ezbiocloud.net/) [20]. MEGA 7.0 software (https://www.megasoftware/) was used for phylogenetic analysis based on individual and concatenated sequences [21]. Distances and clustering were estimated using the neighbor-joining method (bootstrap = 1000).

Testing for virulence genes

The presence of some virulence genes, including vcgC, vvhA, rtxA, Wza, and ompU, was detected by performing PCR with their respective specific primers listed in Supplementary Table F1 [22,23,24].

vcgC is a unique virulence gene of V. vulnificus biotype 1 and has weak pathogenicity but is important for distinguishing between V. vulnificus biotypes 1 and 2 [22]. vvhA (hemolysin A) codes for a 51-kDa water-soluble protein, which is the only cytotoxin extracellularly secreted by V. vulnificus. The V. vulnificus strains carrying this virulence factor gene are highly pathogenic [25]. vvhA is associated with tissue necrosis in the small intestine, which is often followed by the spread of infection to the bloodstream and other tissues [26]. rtxA (cytolysin) is the most interesting virulence gene carried by V. vulnificus biotype 2 and is highly pathogenic [27]. After the bacteria come into contact with host cell, the gene produces a cytotoxin that causes the host cell to lyse. At present, it is believed that the pathogenesis of V. vulnificus infection mainly depends on the synergistic effect of multiple virulence factor genes. In addition to the well-defined main virulence genes vvhA and rtxA, other virulence genes, such as Wza (capsular polysaccharide outer membrane protein) and OmpU (conserved outer membrane porin), are thought to play a role in the pathogenic process of V. vulnificus [28]. Wza plays an important role in protecting bacteria from the host’s innate immune response and is widespread among pathogenic Vibrio spp.[29]. OmpU can mediate bacterial adhesion to host cells and enhance the pathogenicity of V. vulnificus [28].

Physiological and biochemical characteristics

The physiological and biochemical characteristics of the bacterium were analyzed using the API 20E bacterial identification system [30]. The results of the physiological and biochemical characterization of the isolated strain EPL 0201 were evaluated according to Bergey’s Manual of Determinative Bacteriology [31] and Bacteria and Fungi from Fish and Other Aquatic Animals: A Practical Identification Manual [32].

Recursive infection

For recursive infection experiments, healthy pearl gentian groupers (body length, 15–20 cm; body weight, 50–100 g) were purchased from Qianhaiyuan Farm in Tianjin, China, and kept in a 1500-L circulating water system similar to the farm environment for 2 weeks. During this period, the fish were given commercial feed (Haitong, Weifang, China) twice daily, and the water temperature was maintained at 25 °C ± 0.5 °C with 25 ppt salinity.

Before the recursive infection experiments, the fish were randomly divided into 6 groups of 12 each. Each group was placed in an 80-L water tank. Bacterial suspension (1.0 × 107–1.0 × 103 CFU mL−1) was prepared with BSMFL01 (Baso, Guangzhou, China). The fishes in each experimental group were intraperitoneally injected with 0.1 mL of the bacterial suspension, whereas those in the control group were injected with 0.1 mL of 0.85% NaCl. The experimental fish were continuously observed for 7 days after injection, during which normal feeding, water changes, infection incidence, and mortality were recorded, and the LD50 was calculated using the Bliss method [33].

Whole-genome sequencing and analysis

The EPL 0201 strain was grown and harvested in the late exponential growth phase. Total genomic DNA (gDNA) was extracted using bacterial gDNA extraction kits (TaKaRa, Tokyo, Japan). The extracted DNA was detected using agarose gel electrophoresis, and the DNA quality was analyzed using IMPLEN P300 (IMPLEN, Munich, Germany). High-quality bacterial DNA was sent to BioMarker (Beijing, China) and sequenced using the PacBio RSII high-throughput sequencing platform (Pacific Biosciences, Menlo Park, CA, USA).

Genome functional annotation

Public database annotations

The whole-genome sequence of the EPL 0201 strain was obtained using the GenBank non-redundant protein (NR) database [34]. Evolutionary genealogy of genes: Non-supervised Orthologous Groups (eggNOG) [35]; Gene Ontology (GO) [36]; and the Kyoto Encyclopedia of Genes and Genomes (KEGG) [37] were used to predict gene functions.

Proprietary database annotation

We used the Virulence Factor Database (VFDB) [38]and the Comprehensive Antibiotic Research Database (CARD) [39] to identify virulence and resistance factors.

Antimicrobial susceptibility

Antimicrobial susceptibility testing was performed using the Kirby–Bauer paper diffusion method [40]. The bacterial suspension was evenly distributed on Mueller–Hinton agar plates using sterile cotton swabs, and antibiotic disks (Binhe Microorganism Reagent, Hangzhou, China) were placed on the surface of the culture plate using sterile tweezers. The plates were subsequently incubated at 28 °C for 24 h, following which the zone of inhibition was measured and the colonies were counted using a colony counter (Czone 8, Shineso Science & Technology Co., Ltd., Hangzhou, China). The sensitivity was determined according to the manufacturer’s instructions (Binhe Microorganism Reagent).

Results

Identification of isolates

Gram staining indicated that the isolated strain was a rod-shaped gram-negative bacteria (Fig. 1A). Electron microscopy revealed curved rod-shaped bacteria with micro-arc-shaped terminal single flagella (Fig. 1B). The 16S rRNA gene and gyrB sequences were 1434 and 1154 bp in length, respectively, and showed 100% identity with V. vulnificus KU245729.1 and 99.86% identity with V. vulnificus EU118215.1, respectively. A phylogenetic tree was constructed using MEGA 7.0 (Fig. 2A, B), and the EPL 0201 strain was found to cluster with V. vulnificus [19]. The virulence factor gene vcgC is unique to V. vulnificus biotype 1 and can be used to differentiate between V. vulnificus biotypes 1 and 2 [41]. vvhA is a key pathogenic gene for determining the pathogenicity of V. vulnificus, and 90% of pathogenic V. vulnificus carry this factor [27]. The agarose gel electrophoresis did not show the presence of vcgC and OmpU but showed the presence of vvhA, rtxA, and Wza in the PCR products (Fig. 2C). Therefore, it was preliminarily concluded that the EPL 0201 strain was V. vulnificus biotype 2 and had certain pathogenicity.

Fig. 1
figure 1

Varying morphology and flagella of strain EPL 0201 observed under light and electron microscopes. A Micro-curved rod-shaped bacteria (black arrow). B Micro-curved rod-shaped bacteria and flagella observed under an electron microscope

Fig. 2
figure 2

Molecular identification and virulence gene detection of strain EPL 0201. A Phylogenetic tree of Vibrio spp. based on the 16S rRNA gene sequence constructed using the neighbor-joining method. Bootstrap values were based on 1000 replicates, and percentage bootstrap values are shown at each node. B Phylogenetic tree of Vibrio spp. based on gyrB sequence constructed using the neighbor-joining method. Bootstrap values were based on 1000 replicates, and percentage bootstrap values are shown at each node. C Agarose gel electrophoresis did not show the presence of vcgC and OmpU but showed the presence of vvhA, rtxA, and Wza in the PCR products. M, Marker; Neg, negative control. Original agarose gel electropherogram (Supplementary Fig. S6)

Analysis of physiological and biochemical characteristics

The API 20E is a useful tool for delineating V. vulnificus, can identified V. vulnificus biotypes 1 and 2 with 85 ~ 99% accuracy [42, 43]. The main phenotypic characteristics that differentiate V. vulnificus biotype 1 from V. vulnificus biotype 2 are indole production, ornithine decarboxylation, growth at 42 °C, and acid production from mannitol [43]. The code of the EPL 0201 strain in the API 20E coding manual was 500,600,557 (Table 1). This code was consistent with the V. vulnificus ATCC 33,149 biotype 2 strain. Furthermore, this strain did not grow in 0, 8, or 10% NaCl peptone water but did grow in 3, 6, and 7% NaCl peptone water, which is typical of V. vulnificus [43]. These results were consistent with the physiological and biochemical characteristics of V. vulnificus biotype 2, and the strain was determined to be V. vulnificus biotype 2.

Table 1 Physiological and biochemical characteristics of pathogenic strain EPL 0201 and other V. vulnificus strains

Recursive infection

The results of the recursive infection test using the V. vulnificus EPL 0201 strain isolated from the diseased pearl gentian groupers showed that all fishes died 7 days after the injection of the bacterial suspension at a concentration of 1 × 107 CFU mL−1. In addition, mortality of varying degrees was observed after 7 days in the groups injected the bacterial suspension at a concentration of 1 × 106–1 × 104 CFU mL−1. The symptoms of the diseased fish in the recursive infection experiment were same as those of the naturally occurring fish [44]. The diseased fish developed skin lesions, water accumulated in their abdomens, and their tissues and organs were diseased and necrotic (Fig. 3). The fishes in the control group (0.85% NaCl) and the other experimental group (1 × 103 CFU mL−1) did not show any symptoms and none died (Table 2). Bacteria were isolated from the spleen, kidney, and intestinal tract of the dying fish, and a large number of bacteria with the same colony morphology, size, and physiological and biochemical characteristics as the original infecting V. vulnificus EPL 0201 could be isolated. Four strains of V. vulnificus identified based on the physiological and biochemical characteristics were randomly selected for virulence gene detection, and the results showed that they all carried vvhA and rtxA. The LD50 calculated using the Bliss method was 1.097 × 105 CFU g−1 (Table 2). The results of the recursive infection experiment showed that V. vulnificus EPL 0201 had high pathogenicity.

Fig. 3
figure 3

Symptoms of fish tested for recursive infection. A Body chart of experimental fish in control group. B Abdominal cavity of fish in control group (white arrow). C Phosphorus loss in a large area of the skin (white arrow). D Hemorrhagic necrosis of the liver (aa white arrow); a large amount of fluid in the abdominal cavity (ab white arrow); and intestinal inflammation (ac white arrow). E Agarose gel electrophoresis of PCR products of vvhA and rtxA from the suspected V. vulnificus isolates. M, Marker (unit, bp); 1–4, PCR products of vvhA from V. vulnificus in samples from four diseased fishes in the 1 × 107–1 × 104 CFU mL−1 groups; 5–8, PCR products of rtxA from V. vulnificus in samples from 4 diseased fishes in the 1 × 107–1 × 104 CFU mL.−1 groups; 9, negative control; 10–11, positive control (PCR products of vvhA and rtxA from EPL 0201 isolate). Original agarose gel electropherogram (Fig. S6)

Table 2 Results of recursive infection test following intraperitoneal injection

Whole-genome sequencing analysis of EPL 0201

Sequencing data quality control information

The gDNA of V. vulnificus EPL 0201 was detected using 1% agarose gel electrophoresis. The four groups of bacteria-like DNA bands were complete and clear, with no tailing and no residual RNA (Supplementary Fig. S1). The DNA quality test results showed that the OD260 nm/OD280 nm values were between 1.7 and 2.2, which were within the appropriate limit. Sequencing was performed using the PacBio RSII high-throughput sequencing platform. The raw data is shown in Supplementary Table F2, the filtered data in Supplementary Table F3, and the length distribution of the filtered data in Supplementary Fig. S2.

Genome assembly

The filtered subreads data were assembled using HGAP software (Supplementary Table F4). For assembled and predicted genomic information, such as tRNA, rRNA, repeats, GC content, and gene function information, we used Circos v0.66 software to draw circular genome maps (Fig. 4) [45]. The strain was found to have three complete chromosomes (Chr1, Chr2, and Plas1) and two closed circular plasmids (Plas2 and Plas3). Circular genome map assembly metrics is presented in Supplementary Table F5.

Fig. 4
figure 4

GC content and average depth correlation analysis chart. Note: The outermost circle represents the genome size, with each scale being 5 kb. The second circle and the third circle represent genes on the positive and negative strands of the genome. The different colors represent different COG functional classifications. The fourth circle represents a repetitive sequence, the fifth circle represents tRNA (blue) and rRNA (purple), and the sixth circle represents the GC content. The light yellow part indicates that the GC content of the region is higher than the average GC content of the genome, whereas the blue part indicates that the GC content of the region is lower than the average GC content of the genome. The higher the peak value, the greater the difference in GC content. The innermost circle represents GC skew, wherein the dark gray parts represent the regions where the G content is greater than the C content and the red parts represent the regions where the C content is greater than the G content

Functional annotation of the genome

NR database annotations

NR protein database is a non-redundant protein database created and maintained by the NCBI. The database includes comprehensive protein sequence and annotation information along with corresponding species information. The NR database generally enables a more comprehensive annotation of genes compared with other NCBI databases. However, one disadvantage of this database is that many protein sequences and annotations have not been verified, and the reliability needs to be improved.

Annotation using the NR database revealed that Vibrio accounted for 96.48% of the organisms showing sequence homology with the EPL 0201 strain (Fig. 5). Further, V. vulnificus accounted for 75.34% and others accounted for 3.52% of the organisms showing sequence homology with the EPL 0201 strain.

Fig. 5
figure 5

Species distribution map of the results of sequence alignment using the NR database. Note: Different colors represent different species

eggNOG database annotation

The eggNOG database is a database of biological orthologous gene clusters [35]. It is continuously updated based on the COG database [46], which contains data about clusters of orthologous groups. The proteins are assumed to be derived from an ancestral protein and have the same function. The database is generated by comparing the protein sequences of complete genomes.

Using the data from eggNOG (Supplementary Fig. S3), the genome was divided into 25 categories with a total of 4000 genes. The number of S (function unknown) annotations was significantly higher than that of other groups. A (RNA processing and modification), B (chromatin structure and dynamics), W (extracellular structures), Y (nuclear structure), and Z (cytoskeleton) were the five categories with the lowest proportions.

GO database annotation

GO is a database established by the Gene Ontology Consortium to build a framework in which gene and protein functions are defined and described for multiple species [36]. The highest levels of categorization in the GO database are cellular component, which is used to describe subcellular structure, location, and macromolecular complexes; molecular function, which is used to describe the functions of individual gene products; and biological process, which is used to describe the biological process in which the gene product participates.

The GO annotations indicated that the genes were divided into 36 functional groups and summarized into three categories (Supplementary Fig. S4). The EPL 0201 strain had 4,803 annotations related to cellular component, 4,149 annotations related to molecular function, and 6,103 annotations related to biological process. Among the 36 functional components, the 3 categories with the highest number of genes were catalytic activity, metabolic process, and binding, whereas the 3 categories with the lowest number of genes were cell killing, developmental process, and biological adhesion.

KEGG database annotation

The KEGG database is a genomic information database for the systematic analysis of gene function [37]. This database integrates genomic, biochemical, and omics information and produces biological pathway maps of different types of biological processes, thereby helping researchers to study gene and expression information as a whole network.

The KEGG annotations revealed 49 biological pathway genes. The biological pathways were divided into four categories, with each category further subdivided into two levels (Supplementary Fig. S5). Ribosomes in genetic information processing, two-component systems in environmental information processing, bacterial chemotaxis in cellular processes, and biosynthesis of amino acids in metabolism were most highly represented in the pathways.

VFDB annotations

VFDB is a database that contains information on most of the virulence factors from bacterial pathogens that have been completely described [38]. These pathogenic factors help pathogenic bacteria establish infection in the host, respond to the host’s immune system, survive in the host, and cause disease. The database includes two virulence factor sequence datasets. Set A is the experimentally confirmed core dataset, whereas set B contains both the experimentally confirmed virulence factors in set A and predicted virulence factors. This database can help predict possible virulence factors in bacterial genomes that have not been studied in detail.

VFDB annotation results showed that the EPL 0201 strain contained the vvhA, rtxA, and Wza virulence factors (Table 3). This finding was consistent with the previously reported major virulence factors of V. vulnificus [47,48,49].

Table 3 VFDB annotation statistics

CARD database annotations

CARD is a continuously updated database of information related to antibiotic resistance genes. It contains information describing antibiotics and their targets, including antibiotic resistance genes, related proteins, and mechanisms of antibiotic resistance. At the core of CARD is a highly developed Antibiotic Resistance Ontology for the classification of genetic data related to antibiotic resistance [39].

A search using the CARD database showed that the EPL 0201 strain contained H-NS, TEM, sul1, tet, parE, APH, and other resistance genes (Table 4). These resistance genes produce resistance to macrolides, penam, cephalosporin, carbapenems, aminoglycosides, tetracyclines, fluoroquinolones, sulfonamides, and rifampicin antibiotics.

Table 4 CARD database annotation statistics

Antimicrobial susceptibility test

In total, 48 kinds of antibiotics were selected from 11 types of antibacterial drugs to conduct antibacterial drug susceptibility tests on V. vulnificus EPL 0201. The results showed that V. vulnificus EPL 0201 was sensitive to β-lactams such as imipenem, meropenem, piperacillin, and mezlocillin; macrolides such as azithromycin, erythromycin, roxithromycin, and midememycin; nitrofurans such as furazolidone; and aminoglycosides such as tobramycin (Table 5). It also showed varying degrees of resistance to different kinds of antibiotics, including β-lactams (ceftazolin, cefotaxime, ceftazidime, and cefoxitin), tetracyclines (tetracycline and doxycycline), sulfonamides (trimethoprim), quinolones (norfloxacin, ofloxacin, ciprofloxacin, and levofloxacin), aminoglycosides (kanamycin, spectinomycin, and neomycin), glycopeptides (teicoplanin and vancomycin), and lincosamides (lincomycin and clindamycin). These results indicate that V. vulnificus is a multi-drug resistant strain and confirmed the existence of its resistance genes.

Table 5 Antimicrobial susceptibility testing of the pathogenic strain EPL 0201

Discussion

Isolation and identification of the EPL 0201 strain

In this study, the EPL 0201 strain was isolated from the kidney, spleen, and intestine of diseased gentian groupers. The 16S rRNA gene sequence is the genetic marker most commonly used for bacterial taxonomy [50]; however, owing to the extremely high similarity of the 16S rRNA gene sequences in Vibrio species, the accuracy of identification at the species level is limited [51]. V. vulnificus biotypes are complex, and routine identification of 16S rRNA can lead to erroneous results [43]. Therefore, identification was performed using the gyrB gene, the API 20E identification system, and the virulence gene. The gyrB gene is variable and conservative, owing to its genetic codon usage that allows the DNA sequence to undergo relatively many substitutions without changing the amino acid sequence [52]. Therefore, this gene is valuable for the discrimination of Vibrio spp.

The 16S rRNA gene and gyrB sequences of the EPL 0201 strain were analyzed using nucleotide BLAST against the NCBI database. More than 99% homology with the sequence of the standard strain of V. vulnificus was observed. A phylogenetic tree was also constructed, and the results showed that the EPL 0201 strain clustered with V. vulnificus. Several reports on V. vulnificus have indicated that the API 20E system is effective for identifying V. vulnificus, with 85–99% accuracy [42, 43]. The code of the EPL 0201 strain, according to the API 20E coding manual is 500600557, which is consistent with the physiological and biochemical characteristics of V. vulnificus strain ATCC 33,149. In order to improve the accuracy and reliability of the identification, and to avoid false negative results caused by API 20E in the identification of trauma biotypes 1 and 2, resulting in identification errors, additional experiments with NaCl peptone water supplementation were performed, based on the study by Biosca et al. [43]. The results were as expected and the EPL 0201 strain was identified as V. vulnificus biotype 2. The presence of several virulence genes in the EPL 0201 strain indicated that the synergistic effects conferred by combinations of these genes contribute to its pathogenicity. This result is in accord with the observation that combinations of virulence genes are commonly present in particularly virulent strains of V. vulnificus [27]. According to research by Shao et al. [30], a bacterial strain can be considered highly virulent if its LD50 values are within the range of 105–106 CFU g−1 fish body weight. In our study, the LD50 of the EPL 0201 strain was calculated as 1.097 × 105 CFU g−1, indicating that it is highly virulent to the pearl gentian grouper. Fish infected with V. vulnificus tended to exhibit various clinical symptoms, including skin lesions, water accumulation in their abdomens, and diseased and necrotic tissues and organs. Different V. vulnificus strains sometimes, but not always, cause similar clinical symptoms. Even the same V. vulnificus could cause divergent clinical symptoms in different fish. In this study, pearl gentian grouper infected with the EPL 0201 strain developed abdominal effusions. However, in related studies, this symptom was not present after infection with V. vulnificus[44, 53, 54]. The results showed that the pearl gentian grouper infected with V. vulnificus EPL 0201 strain developed skin lesions. This is the same as the findings of Thawonsuwan et al. [44]. However, in the study by Fouz and Li et al. [53, 54], this symptom did not manifest itself after infection with a V. vulnificus strain.

Whole-genome sequencing and annotation of the EPL 0201 strain

V. vulnificus causes a large number of deaths in marine aquaculture fish and seriously threatens the development of aquaculture. It can also infect humans and other mammals, thereby impacting public health. Advances in bioinformatics and genome sequencing have facilitated the understanding of microbial diversity, evolution, and interspecies interactions [55]. Whole-genome sequencing technology has been widely used in the identification and analysis of pathogenic bacteria affecting animals and plants and can aid the systematic study of the evolution, pathogenic mechanisms, and interaction mechanisms of pathogenic bacteria at the molecular level. It provides valuable reference data for the identification and classification of pathogenic bacteria, the detection of drug resistance, the formulation of disease prevention strategies, and the development of vaccines [56, 57].

In recent years, there have been increasing numbers of reports on V. vulnificus, but there are few reports describing the isolation of this bacterium from the pearl gentian grouper, and research on the pathogenesis of this bacterium is still incomplete [11, 44]. In this study, V. vulnificus was isolated from diseased pearl gentian grouper and systematically analyzed. The analysis of the genome sequence of EPL 0201 can provide a theoretical basis for the in-depth study of its mechanisms of pathogenesis. In this study, the whole-genome sequence of V. vulnificus EPL0201 was determined, and the genome (5,769,851 bp) was assembled into three chromosomes and two plasmids. EPL 0201 had a larger genome than that of the reference strain for this bacterium, from the JGI IMG/MER database. It has been reported that the larger the genome, the more likely it is to contain metabolism-related and resistance genes, providing a unique mechanism for adaptation to different living environments [58]. VFDB annotations confirmed that the strain carried three virulence factor genes: vvhA, rtxA, and Wza. Related studies have shown that these genes are closely related to bacterial pathogenicity [29, 59]. The virulence genes were identified via different strains is not completely consistent between studies; thus, virulence genes should be studied in greater depth [58]. KEGG database annotation found that this strain has abundant, complete metabolic pathways. The presence of multiple metabolic pathways enable organisms to continuously exchange substances and energy, maintain metabolism, and respond to the external environment in a timely manner, thereby enhancing the environmental adaptability of strains [60].

Drug resistance analysis of V. vulnificus EPL 0201

Aquatic bacteria, which are able to adapt to their environment, usually contain several antibiotic resistance factors, which enable them to resist antibiotics that may be present in aquaculture ecosystems [61]. Analysis using the CARD database revealed that the EPL 0201 strain had resistance genes such as TEM, sul1, tet, parE, and APH, which can confer resistance to β-lactams (carbapenems), aminoglycosides, tetracyclines, fluoroquinolones, sulfonamides, and rifampicin. This result suggests that the isolate is a multi-drug resistant strain. According to the results of the drug susceptibility test, the EPL 0201 strain was resistant to antibiotics such as β-lactams (cephalosporins), aminoglycosides, tetracyclines, quinolones, sulfonamides, glycopeptides, and lincosamides. These results showed that this strain was multi-drug resistant V. vulnificus and confirmed the resistant genes identified in its genome. However, we found that the strain was highly sensitive to carbapenem and macrolides, a finding which may be related to the type and frequency of carbapenem antibiotics used in this area or to the weakening and loss of function caused by the mutation of resistance genes [23, 62]. Studies have also reported that environmental factors may mediate the mutation or methylation of bacterial resistance genes, thereby changing their resistance [63]. These results also indicate that some drug resistance phenotypes are not completely consistent with resistance gene carriers. Interestingly, the EPL 0201 strain showed varying degrees of resistance to 13 of the 22 antimicrobial agents recommended by the Centers for Disease Control and Prevention for the treatment of Vibrio spp. infection [64], indicating that this strain has gradually become insensitive to the quick-acting drugs used for the treatment of V. vulnificus infection, which should be a cause for alarm.

Availability of data and materials

The 16S rRNA gene sequence of EPL 0201 strain has been submitted to GenBank (https://www.ncbi.nlm.nih.gov/genbank/; accession number OL687343).

References

  1. Vinh DC, Mubareka S, Fatoye B, Plourde P, Orr P. Vibrio vulnificus septicemia after handling Tilapia species fish: a Canadian case report and review. Can J Infect Dis Med Microbiol. 2016;17:129–32. https://doi.org/10.1155/2006/164681.

    Article  Google Scholar 

  2. Hollis DG, Weaver RE, Baker CN, Thornsberry C. Halophilic Vibrio species isolated from blood cultures. J Clin Microbiol. 1976;3:425–31. https://doi.org/10.1128/jcm.3.4.425-431.1976.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Farmer J. Vibrio (“Beneckea”) vulnificus, the bacterium associated with sepsis, septicaemia, and the sea. The Lancet. 1979;2:903. https://doi.org/10.1016/s0140-6736(79)92715-6.

    Article  Google Scholar 

  4. Watanabe H, Miyoshi SI, Kawase T, Tomochika KI, Shinoda S. High growing ability of Vibrio vulnificus biotype 1 is essential for production of a toxic metalloprotease causing systemic diseases in humans. Microb Pathog. 2004;36(3):117–23. https://doi.org/10.1016/j.micpath.2003.10.001.

    Article  CAS  PubMed  Google Scholar 

  5. Chuang YC, Yuan CY, Liu CY, Lan CK, Huang AH. Vibrio vulnificus infection in Taiwan: report of 28 cases and review of clinical manifestations and treatment. Clin Infect Dis. 1992;15:271–6. https://doi.org/10.1093/clinids/15.2.271.

    Article  CAS  PubMed  Google Scholar 

  6. Oliver JD. Wound infections caused by Vibrio vulnificus and other marine bacteria. Epidemiol Infect. 2005;133:383–91. https://doi.org/10.1017/S0950268805003894.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Strom MS, Paranjpye RN. Epidemiology and pathogenesis of Vibrio vulnificus. Microbes Infect. 2000;2:177–88. https://doi.org/10.1016/S1286-4579(00)00270-7.

    Article  CAS  PubMed  Google Scholar 

  8. Jones MK, Oliver JD. Vibrio vulnificus: disease and pathogenesis. Infect Immun. 2009;77:1723–33. https://doi.org/10.1128/iai.01046-08.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Amarco C, Biosca EG, Esteve C, Fouz B, Toranzo AE. Comparative study of phenotypic and virulence properties in Vibrio vulnificus biotypes 1 and 2 obtained from a European eel farm experiencing mortalities. Dis Aquat Org. 1992;13:29–35. https://doi.org/10.3354/dao013029.

    Article  Google Scholar 

  10. Pan CY, Huang TC, Wang YD, Yeh YC, Hui CF, Chen JY. Oral administration of recombinant epinecidin-1 protected grouper (Epinephelus coioides) and zebrafish (Danio rerio) from Vibrio vulnificus infection and enhanced immune-related gene expressions. Fish Shellfish Immunol. 2012;32:947–57. https://doi.org/10.1016/j.fsi.2012.01.023.

    Article  CAS  PubMed  Google Scholar 

  11. Tan J, Hu X, Lü A, Liu X, Sun J, Niu Y. Skin proteome profiling of tongue sole (Cynoglossus semilaevis) challenged with Vibrio vulnificus. Fish Shellfish Immunol. 2020;106:1052–66. https://doi.org/10.1016/j.fsi.2020.09.020.

    Article  CAS  PubMed  Google Scholar 

  12. Liu L, Ge M, Zheng X, Tao Z, Zhou S, Wang G. Investigation of Vibrio alginolyticus, V. harveyi, and V. parahaemolyticus in large yellow croaker, Pseudosciaena crocea (Richardson) reared in Xiangshan Bay China. Aquac Rep. 2016;3:220–4. https://doi.org/10.1016/j.aqrep.2016.04.004.

    Article  Google Scholar 

  13. Guo Z, Gao N, Wu Y, Zhang L. The simultaneous uptake of dietary and waterborne Cd in gastrointestinal tracts of marine yellowstripe goby Mugilogobius chulae. Environ Pollut. 2017;223:31–41. https://doi.org/10.1016/j.envpol.2016.12.007.

    Article  CAS  PubMed  Google Scholar 

  14. Khemiss F, Ahmadi S, Massoudi R, Ghoul-Mazgar S, Safta S, Moshtaghie AA, Saïdane D. Effect of in vitro exposure to Vibrio vulnificus on hydroelectrolytic transport and structural changes of sea bream (Sparus aurata L.) intestine. Fish Physiol Biochem. 2009;35:541–9. https://doi.org/10.1007/s10695-008-9265-7.

    Article  CAS  PubMed  Google Scholar 

  15. Roig FJ, González-Candelas F, Sanjuán E, Fouz B, Feil EJ, Llorens C, Baker-Austin C, Oliver JD, Danin-Poleg Y, Gibas CJ, Kashi Y, Gulig PA, Morrison SS, Amaro C. Phylogeny of Vibrio vulnificus from the analysis of the core-genome: implications for intra-species taxonomy. Front Microbiol. 2017;8:2613. https://doi.org/10.3389/fmicb.2017.02613.

    Article  PubMed  Google Scholar 

  16. Bisharat N, Cohen DI, Harding RM, Falush D, Crook DW, Peto T, Maiden MC. HybridVibrio vulnificus. Emerg Infect Dis. 2005;11:30–5. https://doi.org/10.3201/eid1101.040440.

    Article  PubMed  PubMed Central  Google Scholar 

  17. Weisburg WG, Barns SM, Pelletier DA, Lane DJ. 16S ribosomal DNA amplification for phylogenetic study. J Bacteriol. 1991;173:697–703. https://doi.org/10.1128/jb.173.2.697-703.1991.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Sawabe T, Kita-Tsukamoto K, Thompson FL. Inferring the evolutionary history of vibrios by means of multilocus sequence analysis. J Bacteriol. 2007;189:7932–6. https://doi.org/10.1128/JB.00693-07.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Sayers EW, Barrett T, Benson DA, Bolton E, Bryant SH, Canese K, Ye J. Database resources of the National Center for Biotechnology Information. Nucleic Acids Res. 2010;39((Database)):D38–51. https://doi.org/10.1093/nar/gkq1172.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Yoon SH, Ha SM, Kwon S, Lim J, Kim Y, Seo H, Chun J. Introducing EzBioCloud: a taxonomically united database of 16S rRNA gene sequences and whole-genome assemblies. Int J Syst Evol Microbiol. 2017;67:1613–7. https://doi.org/10.1099/ijsem.0.001755.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Kumar S, Stecher G, Tamura K. MEGA7: Molecular Evolutionary Genetics Analysis Version 7.0 for Bigger Datasets. Mol Biol Evol. 2016;33:1870–4. https://doi.org/10.1093/molbev/msw054.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Warner E, Oliver JD. Population structures of two genotypes of Vibrio vulnificus in oysters (Crassostrea virginica) and seawater. Appl Environ Microbiol. 2008;74:80–5. https://doi.org/10.1128/aem.01434-07.

    Article  CAS  PubMed  Google Scholar 

  23. Baker-Austin C, Mcarthur JV, Lindell AH, Wright MS, Tuckfield RC, Gooch J, Warner L, Oliver J, Stepanauskas R. Multi-site analysis reveals widespread antibiotic resistance in the marine pathogen Vibrio vulnificus. Microb Ecol. 2009;57:151–9. https://doi.org/10.1007/s00248-008-9413-8.

    Article  CAS  PubMed  Google Scholar 

  24. Rivera ING, Chun J, Huq A, Sack RB, Colwell RR. Genotypes associated with virulence in environmental isolates of Vibrio cholerae. Appl Environ Microb. 2001;67:2421–9. https://doi.org/10.1128/aem.67.6.2421-2429.2001.

    Article  CAS  Google Scholar 

  25. Wuermeling HB. Use-risk consideration in antibacterial chemotherapy. Infection. 1991;19(Supplement 1):S13–5. https://doi.org/10.1007/bf01644728.

    Article  PubMed  Google Scholar 

  26. Jeong HG, Satchell KJF. Correction: additive function of Vibrio vulnificus MARTX Vv and VvhA cytolysins promotes rapid growth and epithelial tissue necrosis during intestinal infection. PLOS Pathogens. 2019;15:e1007895. https://doi.org/10.1371/journal.ppat.1007895.

    Article  PubMed  PubMed Central  Google Scholar 

  27. Amaro C, Sanjuán E, Fouz B, Pajuelo D, Lee CT, Hor LI, Barrera R. The fish pathogen Vibrio vulnificus biotype 2: epidemiology, phylogeny, and virulence factors involved in warm-water vibriosis. Microb Spectr. 2015;3:3. https://doi.org/10.1128/microbiolspec.ve-0005-2014.

    Article  Google Scholar 

  28. Guo S, Hu L, Feng J, Lin P, He L, Yan Q. Immunogenicity of a bivalent protein as a vaccine against Edwardsiella anguillarum and Vibrio vulnificus in Japanese eel (Anguilla japonica). Microbi Open. 2019;8:e00766. https://doi.org/10.1002/mbo3.766.

    Article  CAS  Google Scholar 

  29. Wright AC, Powell JL, Kaper JB, Morris JG. Identification of a group 1-like capsular polysaccharide operon for Vibrio vulnificus. Infect Immun. 2001;69:6893–901. https://doi.org/10.1128/iai.69.11.6893-6901.2001.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Shao P, Yong P, Zhou W, Sun J, Wang Y, Tang Q, Ren S, Wu Z, Zhao C, Xu Y, Wang X. First isolation of Photobacterium damselae subsp. damselae from half-smooth tongue sole suffering from skin-ulceration disease. Aquaculture. 2019;511:734208. https://doi.org/10.1016/j.aquaculture.2019.734208.

    Article  Google Scholar 

  31. Garrity G, Brenner D, Krieg NR, Staley JT, Krieg N. The Proteobacteria, PartC: the alpha-, beta-, delta-, and Epsilonproteobacteria. New York: Bergey’s manual of systematic bacteriology. Springer; 2005. p. 647–57.

    Google Scholar 

  32. Buller NB. Bacteria and fungi from fish and other aquatic animals: a practical identification manual. 2014. https://doi.org/10.1079/9781845938055.0000.

    Article  Google Scholar 

  33. Finney DJ. The median lethal dose and its estimation. Arch Toxicol. 1985;56:215–8. https://doi.org/10.1007/bf00295156.

    Article  CAS  PubMed  Google Scholar 

  34. Fu L, Niu B, Zhu Z, Wu S, Li W. CD-HIT: accelerated for clustering the next-generation sequencing data. Bioinformatics. 2012;28:3150–2. https://doi.org/10.1093/bioinformatics/bts565.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Huerta-Cepas J, Szklarczyk D, Heller D, Hernández-Plaza A, Forslund SK, Cook H, Mende DR, Letunic I, Rattei T, Jensen LJ, von Mering C, Bork P. eggNOG 5.0: a hierarchical, functionally and phylogenetically annotated orthology resource based on 5090 organisms and 2502 viruses. Nucleic Acids Res. 2019;47:D309–14. https://doi.org/10.1093/nar/gky1085.

    Article  CAS  PubMed  Google Scholar 

  36. Ashburner M, Ball CA, Blake JA, Botstein D, Butler H, Cherry JM, Davis AP, Dolinski K, Dwight SS, Eppig JT, Harris MA, Hill DP, Issel-Tarver L, Kasarskis A, Lewis S, Matese JC, Richardson JE, Ringwald M, Rubin GM, Sherlock G. Gene ontology: tool for the unification of biology The Gene Ontology Consortium. Nat Gen. 2000;25:25–9. https://doi.org/10.1038/75556.

    Article  CAS  Google Scholar 

  37. Minoru K, Susumu G, Shuichi K, Yasushi O, Masahiro H. The KEGG resource for deciphering the genome. Nucleic Acids Res. 2004;32:277–80. https://doi.org/10.1093/nar/gkh063.

    Article  CAS  Google Scholar 

  38. Chen L, Jian Y, Yu J, Yao Z, Sun L, Shen Y, Qi J. VFDB: a reference database for bacterial virulence factors. Nucleic Acids Res. 2005;33((suppl_1)):325–8. https://doi.org/10.1093/nar/gki008.

    Article  CAS  Google Scholar 

  39. Jia B, Raphenya AR, Alcock B, Waglechner N, Guo P, Tsang KK, Lago BA, Dave BM, Pereira S, Sharma AN, Doshi S, Courtot M, Lo R, Williams LE, Frye JG, Elsayegh T, Sardar D, Westman EL, Pawlowski AC, Johnson TA, Brinkman FS, Wright GD, McArthur AG. CARD 2017: expansion and model-centric curation of the comprehensive antibiotic resistance database. Nucleic Acids Res. 2017;45:566–73. https://doi.org/10.1093/nar/gkw1004.

    Article  CAS  Google Scholar 

  40. Shao P, Yong P, Wang X, Xie S, Fan Y, Zang L, Cui L, Sun J. Isolation, identification, and histopathological analysis of Vibrio tubiashii from lined seahorse hippocampus erectus. Dis Aquat Org. 2019;133:195–205. https://doi.org/10.3354/dao03350.

    Article  CAS  Google Scholar 

  41. Baker-Austin C, Gore A, Oliver JD, Rangdale R, McArthur JV, Lees DN. Rapid in situ detection of virulent Vibrio vulnificus strains in raw oyster matrices using real-time PCR. Environ Microb Rep. 2010;2:76–80. https://doi.org/10.1111/j.1758-2229.2009.00092.x.

    Article  CAS  Google Scholar 

  42. Macdonell MT, Singleton FL, Hood MA. Diluent composition for use of API 20E in characterizing marine and estuarine bacteria. Appl Environ Microb. 1982;44:423–7. https://doi.org/10.1007/BF00448446.

    Article  CAS  Google Scholar 

  43. Biosca EG, Esteve C, Garay E, Amaro C. Evaluation of the API 20E system for identification and discrimination of Vibrio vulnificus biotypes 1 and 2. J Fish Dis. 1993;16:79–82. https://doi.org/10.1111/j.1365-2761.1993.tb00850.x.

    Article  Google Scholar 

  44. Thawonsuwan J, Kasornchandra J, Soonsan P, Keawtapee C. Isolation of vibrio vulnificus biotype 1 from disease outbreaks on cultured tiger grouper epinephelus fuscoguttatus forsskal, 1775. Fish Pathol. 2016;51((Special-issue)):S39–45. https://doi.org/10.3147/jsfp.51.S39.

    Article  Google Scholar 

  45. Krzywinski M, Schein J, Birol I, Connors J, Gascoyne R, Horsman D, Jones SJ, Marra MA. Circos: an information aesthetic for comparative genomics. Genome Res. 2009;19:1639–45. https://doi.org/10.1101/gr.092759.109.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Galperin MY, Wolf YI, Makarova KS, Vera Alvarez R, Landsman D, Koonin EV. COG database update: focus on microbial diversity, model organisms, and widespread pathogens. Nucleic Acids Res. 2021;49:D274–81. https://doi.org/10.1093/nar/gkaa1018.

    Article  CAS  PubMed  Google Scholar 

  47. Lee KJ, Lee NY, Han YS, Kim J, Lee KH, Park SJ. Functional characterization of the IlpA protein of Vibrio vulnificus as an adhesin and its role in bacterial pathogenesis. Infect Immun. 2010;78:2408–17. https://doi.org/10.1128/iai.01194-09.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Lee SJ, Jung YH, Oh SY, Song EJ, Choi SH, Han HJ. Vibrio vulnificus VvhA induces NF-κB-dependent mitochondrial cell death via lipid raft-mediated ROS production in intestinal epithelial cells. Cell Death Dis. 2015;6:1655. https://doi.org/10.1038/cddis.2015.19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Goo SY, Lee HJ, Kim WH, Han KL, Park DK, Lee HJ, Kim SM, Kim KS, Lee KH, Park SJ. Identification of OmpU of Vibrio vulnificus as a fibronectin-binding protein and its role in bacterial pathogenesis. Infect Immun. 2006;74:5586–94. https://doi.org/10.1128/iai.00171-06.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Busse H-J, Denner EBM, Lubitz W. Classification and identification of bacteria: current approaches to an old problem. Overview of methods used in bacterial systematics. J Biotechnol. 1996;47:3–38. https://doi.org/10.1016/0168-1656(96)01379-x.

    Article  CAS  PubMed  Google Scholar 

  51. Wang X, Yuan Y, Wang H, Zhou Q, Zhu J, Chang Y, 2009. Phylogenetic analysis of four vibrio strains of pathogenic bacteria based on hemolysin genes and 16S rRNA genes. 3rd International Conference on Bioinformatics and Biomedical Engineering. https://doi.org/10.1109/icbbe.2009.5163190.

  52. Kasai H, Watanabe K, Gasteiger E, Bairoch A, Isono K, Yamamoto S. Construction of the gyrB database for the identification and classification of bacteria. Genome informatica seria workshop. Genome Inform. 1998;9:13–21. https://doi.org/10.11234/gi1990.9.13.

    Article  CAS  Google Scholar 

  53. Fouz B, Amaro C. Isolation of a new serovar of Vibrio vulnificus pathogenic for eels cultured in freshwater farms. Aquaculture. 2003;217:677–82. https://doi.org/10.1016/s0044-8486(02)00200-4.

    Article  Google Scholar 

  54. Li G, Zhao D, Huang L, Sun J, Gao D, Wang H, Liang L. Identification and phylogenetic analysis of Vibrio vulnificus isolated from diseased Trachinotus ovatus in cage mariculture. Aquaculture. 2006;261(1):17–25. https://doi.org/10.1016/j.aquaculture.2006.07.01.

    Article  CAS  Google Scholar 

  55. Dolled-Filhart MP, Lee M, Ou-yang CW, Haraksingh RR, Lin JC-H. Computational and bioinformatics frameworks for next-generation whole exome and genome sequencing. ScientificWorldJournal. 2013;2013:730210. https://doi.org/10.1155/2013/730210.

    Article  PubMed  PubMed Central  Google Scholar 

  56. Biswas S, Raoult D, Rolain JM. A bioinformatic approach to understanding antibiotic resistance in intracellular bacteria through whole genome analysis. Int J Antimicrob Agents. 2008;32:207–20. https://doi.org/10.1016/j.ijantimicag.2008.03.0.

    Article  CAS  PubMed  Google Scholar 

  57. Pan Y, Yang X, Duan J, Lu N, Leung AS, Tran V, Hu Y, Wu N, Liu D, Wang Z, Yu X, Chen C, Zhang Y, Wan K, Liu J, Zhu B. Whole-genome sequences of four Mycobacterium bovis BCG vaccine strains. J Bacteriol. 2011;193:3152–3. https://doi.org/10.1128/jb.00405-11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Liu J, Zhao Z, Deng Y, Shi Y, Liu Y, Wu C, Luo P, Hu C. Complete genome sequence of Vibrio campbellii LMB 29 isolated from red drum with four native megaplasmids. Front Microbiol. 2017;8:2035. https://doi.org/10.3389/fmicb.2017.02035.

    Article  PubMed  PubMed Central  Google Scholar 

  59. Song EJ, Lee SJ, Lim HS, Kim JS, Jang KK, Choi SH, Han HJ. Vibrio vulnificus VvhA induces autophagy-related cell death through the lipid raft-dependent c-Src/NOX signaling pathway. Sci Rep. 2016;6:27080. https://doi.org/10.1038/srep27080.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Maughan H, Nicholson WL. Increased fitness and alteration of metabolic pathways during Bacillus subtilis evolution in the laboratory. Appl Environ Microbiol. 2011;77:4105–18. https://doi.org/10.1128/aem.00374-11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Marti E, Variatza E, Balcazar JL. The role of aquatic ecosystems as reservoirs of antibiotic resistance. Trends Microbiol. 2014;22:36–41. https://doi.org/10.1016/j.tim.2013.11.001.

    Article  CAS  PubMed  Google Scholar 

  62. Nascimento SM, dos Fernandes Vieira RH, Theophilo GN, Dos Prazeres Rodrigues D, Vieira GH. Vibrio vulnificus as a health hazard for shrimp consumers. Rev Inst de Med Trop São Paulo. 2001;43:263–6. https://doi.org/10.1590/s0036-46652001000500005.

    Article  Google Scholar 

  63. Yan L, Liu D, Wang XH, Wang Y, Zhang B, Wang M, Xu H. Bacterial plasmid-mediated quinolone resistance genes in aquatic environments in china. Sci Rep. 2017;7:40610. https://doi.org/10.1038/srep40610.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Shaw KS, Rosenberg Goldstein RE, He X, Jacobs JM, Crump BC, Sapkota AR. Antimicrobial susceptibility of Vibrio vulnificus and Vibrio parahaemolyticus recovered from recreational and commercial areas of Chesapeake Bay and Maryland Coastal Bays. PLOS ONE. 2014;9:e89616. https://doi.org/10.1371/journal.pone.0089616.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We would like to thank NES (lunwen@nesediting.com) for English language editing.

Funding

This research was jointly funded by Tianjin Municipal Science and Technology Commission (18YFZCNC01190) and Tianjin Natural Science Foundation Project (18JCQNJC14800).

Author information

Authors and Affiliations

Authors

Contributions

ZW and YW performed the experiments and wrote the manuscript. HX, QY, ZY, JZ, JG, XJ, and YD are responsible for experimental collection and manuscript writing. XW and PS performed the experimental design and revised the manuscript. The author(s) read and approved the final manuscript.

Corresponding authors

Correspondence to Xiaoyu Wang or Peng Shao.

Ethics declarations

Ethics approval and consent to participate

All animal experiments were approved by the Animal Experiment Ethics Committee at Tianjin Agricultural University, China.

All methods were performed in accordance with relevant guidelines and regulations.

All methods are reported in accordance with ARRIVE guidelines (https://arriveguidelines.org) for the reporting of animal experiments.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

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

Supplementary Information

Additional file 1: Table F1.

Primers used in this study. Table F2. Raw sequencing data. Table F3. Filtered sequencing data. Table F4. Summarized data of gene prediction. Table F5. Assembly index. Fig. S1. Electrophoresis of the DNA isolated from the EPL 0201 strain. Fig. S2. Sequencing data length distribution of the EPL 0201 train. Fig. S3. Function classification statistics of eggNOG functional genes. Note: The abscissa represents the content of each eggNOG classification, and the ordinate represent the relative content of the number of corresponding functional genes. Fig. S4. GO function annotation clasification statistics chart. Note: The abscissa represents the content of each GO category, the left of the ordinate represents the percentage of genes, and the right of the ordinate represents the numbers of genes. Fig. S5. KEGG annotation classification statistics. Note: The ordinate represents the KEGG secondary classification, and the abscissa represents the percentage. Fig. S6. Original agarose gel electropherogram. Fig. S7. pictures of naturally occurring fish. Fig. S8. Electrophoresis of the DNA isolated from teh EPL 0201 strain (original image).

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wu, Z., Wu, Y., Gao, H. et al. Identification and whole-genome sequencing analysis of Vibrio vulnificus strains causing pearl gentian grouper disease in China. BMC Microbiol 22, 200 (2022). https://doi.org/10.1186/s12866-022-02610-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12866-022-02610-1

Keywords