Next Article in Journal
Bacterial Communities in Lanna Fermented Soybeans from Three Different Ethnolinguistic Groups in Northern Thailand
Previous Article in Journal
Quiescence of Escherichia coli Aerosols to Survive Mechanical Stress during High-Velocity Collection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genomic Characterization and Wetland Occurrence of a Novel Campylobacter Isolate from Canada Geese

1
Oak Ridge Institute for Science and Education, Oakridge, TN 37830, USA
2
Office of Research and Development, United States Environmental Protection Agency, Cincinnati, OH 45268, USA
3
EPA National Student Services Contact (NSSC), Cincinnati, OH 45268, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Microorganisms 2023, 11(3), 648; https://doi.org/10.3390/microorganisms11030648
Submission received: 24 January 2023 / Revised: 23 February 2023 / Accepted: 1 March 2023 / Published: 3 March 2023
(This article belongs to the Section Systems Microbiology)

Abstract

:
Populations of resident, non-migratory Canada geese are rapidly increasing. Canada geese are known to transmit viral and bacterial diseases, posing a possible threat to human health. The most prevalent pathogens vectored by geese are Campylobacter species, yet the current understanding of the identity and virulence of these pathogens is limited. In our previous study, we observed a high prevalence of Campylobacter spp. in the Banklick Creek wetland—a constructed treatment wetland (CTW) located in northern KY (USA) used to understand sources of fecal contamination originating from humans and waterfowl frequenting the area. To identify the types of Campylobacter spp. found contaminating the CTW, we performed genetic analyses of Campylobacter 16s ribosomal RNA amplified from CTW water samples and collected fecal material from birds frequenting those areas. Our results showed a high occurrence of a Campylobacter canadensis-like clade from the sampling sites. Whole-genome sequence analyses of an isolate from Canada goose fecal material, called MG1, were used to confirm the identity of the CTW isolates. Further, we examined the phylogenomic position, virulence gene content, and antimicrobial resistance gene profile of MG1. Lastly, we developed an MG1-specific real-time PCR assay and confirmed the presence of MG1 in Canada goose fecal samples surrounding the CTW. Our findings reveal that the Canada goose-vectored Campylobacter sp. MG1 is a novel isolate compared to C. canadensis that possesses possible zoonotic potential, which may be of human health concern.

1. Introduction

From 1970 to 2012, resident non-migratory Canada geese (Branta canadensis) populations increased from 1.26 million to 5.69 million in North America [1]. While this increase may support recreational activities such as bird watching, high goose occupancy can lead to undesirable consequences such as agricultural damage, accumulation of fecal material on surfaces, and degraded water quality [2]. Waterfowl such as Canada geese are known to transmit viral and bacterial diseases of human and agricultural concern [3,4]. Their feces are a source of pathogens such as avian influenza, Arcobacter, Escherichia coli, Salmonella, Giardia, Cryptosporidium, and Campylobacter, which can be released into freshwater reservoirs [3,5,6,7,8]. Among those potential pathogens, Campylobacter spp. can be highly prevalent (11.2%) in Canada geese, suggesting that these birds may play a role in dispersing Campylobacter within their frequented environments [9].
Any environment frequented by wild birds, and waterfowl specifically, could contain a diverse array of Campylobacter spp. (for review see [10]). Sanitation District No. 1 (SD1) in Fort Wright, Kentucky (USA) installed a constructed treatment wetland (CTW) in 2011 to divert and treat a portion of the Banklick Creek, a watershed regularly impacted by combined sewer overflows (CSOs) and sanitary sewer overflows (SSOs) in the area. CTWs can be an effective means to passively treat water by removing harmful chemical and microbial contaminates, but since they closely mimic natural wetlands, they can attract native wildlife such as waterfowl. Previous research using qPCR analyses of the treated water throughout the system showed that Campylobacter and avian marker signals (targeting Helobacter sp.) were detected at high frequencies where bird occupancy was commonly observed [11]. Moreover, various waterfowl species, such as Canada goose, were frequently found throughout the CTW. However, more in-depth characterization of the Campylobacter spp. diversity detected in this wetland, as well as establishing if Canada geese could be vectors of human pathogenic Campylobacter strains in these environments were not conducted. To understand the molecular basis of pathogenicity, the identification of virulence factors is needed to elucidate the potential of a Campylobacter isolate to cause disease [12]. For those well-known Campylobacter species (C. jejuni and C. coli) which cause human (and animal) infections, several putative virulence factors (genes) that contribute to motility, intestinal adhesion, colonization, toxin production, and invasion have been identified as critical to their human and animal pathogenicity. For example, the genes cdtA, cdtB, and cdtC [13,14] encode Campylobacter cytolethal distending toxin, causing host cell cycle arrest, cell distention, and eventually cell death [15]. Some genes such as flaA encoding flagellin [16] and cadF encoding a protein that interacts with a host extracellular matrix protein fibronectin [17] are required for Campylobacter adherence to, and colonization of the host cell surface. Although Campylobacter spp. have been isolated from Canada geese as mentioned above, documentation addressing their human pathogenicity is rare, and data on the virulence of Campylobacter strains isolated from the large resident populations of Canada geese in urban and suburban settings are limited.
In this study, we began by using genetic analysis to parse the possible identity of targeted Campylobacter 16s rRNA genes amplified from the CTW. We found that a large portion of the sequenced clones originated from a Campylobacter canadensis-like species of unknown identity. Next, we tried to isolate the potential Campylobacter spp. from CTW Canada goose fecal samples, which failed despite multiple attempts. Assuming the age of the fecal samples was hampering our efforts, we collected fresh fecal samples from a Mason city park (40 miles away from the CTW), located near our laboratories, where we could monitor goose activity and collect droppings immediately. From the Mason samples, we identified a unique isolate, which we called MG1. Initial 16s rRNA sequence analysis showed high similarity between MG1 and isolates from the CTW. To further understand MG1 at a genetic level we performed genome sequencing and assembled an annotated draft genome. We used phylogenomics to explore the relationship of the MG1 isolate to other Campylobacter species and confirmed the identity of the unknown clone sequences from the CTW as most closely related to MG1. We also surveyed the MG1 genome for loci associated with virulence and disease in humans to begin to understand the zoonotic potential of the isolate. Finally, we designed a unique real-time PCR assay for detecting and tracking MG1 in goose fecal contamination in CTW water samples. Our findings help to understand host-specific strains of Campylobacter spp. and aid in microbiological water source tracking. These data contribute to an overall effort to safeguard public water ecosystems and will assist in preventing future human and animal exposure.

2. Materials and Methods

2.1. Fecal Sample Collection, Isolation and DNA Extraction

Initially, two collections of Canada goose fecal material with an unknown time of defecation were collected from the CTW (Ft. Wright, KY, USA) in fall 2017. After failing to isolate Campylobacter spp. from these samples, fresh Canada goose fecal samples were collected from local, easily accessible city park lawns where geese are frequently found and monitored (Mason, OH, USA) during the following year. Fecal samples were collected just after defecation, immediately transferred to a sterile 50 mL conical tube (ThermoFisher Scientific, Waltham, MA, USA), and brought to the US EPA laboratory in Cincinnati, OH for further processing. Each fecal pellet was submerged in sterile 1× PBS solution and vortexed to homogenize the sample. Duplicate aliquots (100 µL) were diluted 10-fold in sterile PBS and cultured as previously described [18]. Briefly, each sample was spread on two 47-mm diameter polycarbonate filters of 0.6-µm pore size (GE Water & Process Technologies, Addison, IL, USA) placed on tryptic soy agar TSA w/5% Sheep’s Blood (BD biosciences, San Jose, CA, USA). The plates were incubated at 37 °C, which was found to be the optimal temperature for the recovery and growth of most Campylobacter species [19], for 30 min. Following incubation, the filters were removed, and the plates were grown in a microaerophilic chamber at 37 °C for 48 h. Presumptive Campylobacter-positive colonies were picked and streaked on TSA w/5% Sheep’s Blood plates and grown at 37 °C in a microaerophilic chamber for 48 h for isolation. From the plates, a single colony was selected for DNA extraction using the Tissue and Cell Lysis kit (Epicenter Technologies Corp. Madison, WI, USA) following the manufacturer’s instructions. DNA was screened for Campylobacter using the 16s PCR assay described by Linton et al. 1996 [20]. The isolates that were positive by PCR were stored at −80 °C in tryptic soy yeast broth (TSY) containing 15% glycerol.

2.2. Illumina MiSeq Genomic Sequencing, Read Processing, and Assembly/Annotation

From positive Campylobacter isolates (see above), a sequencing library was prepared using Nextera XT (Illumina) following the manufacturer’s protocol. The libraries were denatured and diluted following the MiSeq Denature and Dilute library guide to a final concentration of 10 pM and mixed with 10% PhiX sequencing control. The pooled library was then loaded into a MiSeq Reagent Kit v3 and run using 300 base-pair (bp) paired-end chemistry (Illumina, San Diego, CA, USA). During this effort, 3,722,434 paired reads were produced for further analysis. Raw reads with primers and adapters removed were then processed. Reads were quality checked using Fastqc v.0.11.9 [21] and cleaned using Trimmomatic v.0.39 [22] to perform quality trimming. Trimmomatic was run with the following parameters, LEADING:2 TRAILING:20 SLIDINGWINDOW:5:20 MINLEN:150. After trimming, Fastqc was run again. Approximately 9% of reads were removed by trimming. The genome was assembled with the SPAdes assembler v3.11.1 [23] using default parameters, and whole-genome annotation was performed with PROKKA v1.14.6 [24]. Assembly quality was evaluated with QUAST v5.0.2 [25]. Genome assembly information is presented in Table 1.

2.3. Antibiotic Resistance Gene, Virulence Gene, Lipooligosaccharide Biosynthesis Loci, and Capsular Polysaccharide Analyses

To check for antimicrobial resistance genes, we used the Comprehensive Antibiotic Resistance Database (CARD, version 3.2.3) RGI tool (version 5.2.1) that predicts Open Reading Frames (ORFs) using Prodigal, homolog detection using DIAMOND, and Strict significance based on CARD curated bitscore cutoffs [26]. For analysis of virulence genes, lipooligosaccharide (LOS), and capsular polysaccharide (CPS) regions, we used BLAST v2.9.0+ [27]. Custom BLAST databases were generated from the PROKKA amino acid annotations from the MG1 isolate genome. Query virulence gene sequences were taken from Campylobacter jejuni NTCT11168 or other species where appropriate. For all BLAST analyses, alignments were inspected by hand, and e-value cutoffs were made at 1 × 10−5. Reciprocal BLAST searches were performed to confirm orthology when positive hits were detected.

2.4. Water Sampling, PCR, Cloning, and Sanger Sequencing for Banklick Creek Wetland Samples

The wetland sampling has been described in our previous study [11]. Briefly, samples were collected from June through September 2017 from five sites in the Banklick Creek Treatment Wetland located in Fort Wright, Kentucky (39°01′14.2″ N 84°31′42.1″ W) (Figure S1). Water samples (~10 L) were collected in a sterilized carboy from each of the sampling locations over a 1-h long sampling event, and immediately transported to the U.S. EPA laboratory located in Cincinnati, Ohio for processing. Bird occupancy was also recorded at each of the five sampling locations during each of the 15 independent sampling events. Recorded bird occupancy was then totaled for the entirety of this study at each sample site. For the collected water samples, duplicate 200 mL aliquots from each 10 L carboy were passed through a 47 mm polyvinylidene difluoride (PVDF) filter with a nominal pore size of 0.45 μm (Millipore, Burlington, MA, USA) (Total: 15 × 5 × 2). Each membrane was transferred to a Lysing Matrix A bead tube (MP Biomedicals, Santa Ana, CA, USA), and the resulting filter tubes were stored at −80 °C until extraction. To each filter tube, 600 μL RLT Plus buffer was added and the tube underwent bead beating (5000 reciprocations/min) for 30 s, cooling on ice for 5 min, exposure to another round of bead beating, and centrifugation for 5 min at 12,000 RPM. Nucleic acids were then extracted from the supernatant using the AllPrep DNA/RNA Mini Kit (Qiagen, Germantown, MD, USA), following the manufacturer’s instructions. A Campylobacter genus-specific PCR assay was used to amplify 28 of the samples which were positive via the previous qPCR assay for Campylobacter genera [11,20,28] (Table S1). The amplicons were cloned into pCR4.1 TOPO (Invitrogen, Carlsbad, CA, USA) to aid in further identification. Raw sequences were processed with Sequencher 5.2.4 software (Gene Codes, Ann Arbor, MI, USA) for editing, initial comparisons, alignment, homology searches, and a chimera check as previously described [29].

2.5. 16s rRNA Phylogenetics and Phylogenomic Analyses

To resolve the phylogenomic position of Campylobacter isolate MG1, we utilized GToTree v1.6.11 [30] with a variety of Campylobacter genomes from GenBank. GToTree automates querying and downloading all genomes from GenBank, translating the MG1 isolate genome assembly in all open reading frames, filtering all genomes for 119 possible single-copy genes against the Proteobacteria HMM-gene set, retaining only single-copy genes present in at least 90% of all genomes which, for the genomes we chose for analysis, retained ~108 genes, on average. GToTree then performs multiple sequence alignments followed by phylogenetic reconstruction of the concatenated multiple sequence alignment using heuristic neighbor-joining with the consensus tree calculated from 1000 bootstrap replicates. Average nucleotide identity (ANI) was used to contrast closely related genomes and assess species similarity [31,32].
To analyze the phylogenetic distribution of Campylobacter isolates detected at sites 3 and 5 in the Banklick Creek Wetland area (see Figure S1), we gathered the 16s rRNA nucleotide sequences (see above), together with nine additional 16s rRNA sequences from known Campylobacter spp., as well as the 16s sequence from the MG1 isolate, for a total of 38 nucleotide sequences. Sequences were compiled and aligned using MUSCLE in MEGA X [33] with default settings. Alignments were manually inspected and trimmed to remove gaps. The evolutionary history was inferred using the neighbor-joining method in MEGA X. The bootstrap consensus tree was inferred from 500 replicates. The evolutionary distances were computed using the Maximum Composite Likelihood method to estimate the base substitution rate. All positions containing gaps and missing data following alignment were eliminated (complete deletion option). In a total, 661 positions were included in the final dataset.

2.6. MG1 16s rRNA Primer Design

To design primers specific to MG1 16s rRNA, we obtained 16s rRNA nucleotide sequences from five Campylobacter species (C. lari—accession: CP046243.1, C. coli—accession: CP083814.1, C. rectus—accession: CP012543.1, C. jejuni—accession: CP048760.1, and C. canadensis—accession: CP035946.1) and performed multiple sequence alignment against the MG1 isolate 16s rRNA nucleotide sequence (Figure S2). We inspected the alignment and designed primers and a probe placed in regions with the greatest number of nucleotide differences. Using BLAST, we inspected the sequence region of each species where primers were designed to check for a lack of inter-species variation. Three sets of real-time PCR assays were designed, and in silica analysis was conducted for the oligo sequence of each set to ensure that at least one primer and the probe of each assay was specific to MG1. It is worth noting that, because of the sequence conservation, we were unable to locate a target region for MG1 where the primers were highly unique compared to C. canadensis; however, the forward primer sequence contained at least 1 unique residue, while the probe sequence contained 3 unique residues. Lastly, we evaluated the three assays in terms of their sensitivity and amplification efficiency and a single primer/probe set was selected and used for this study. The MG1 assay was also cross-tested against various environmental and laboratory Campylobacter strains and three clinical strains to assure no false-positive detection occurred [18]. For this analysis, real-time PCR was performed as described below (also see Table S2).

2.7. Real-Time PCR Analysis

Real-time PCR assays were used to detect the MG1 Campylobacter isolate for samples from Canada geese excreta. Real-time PCR was performed on a QuantStudio 6 Flex thermal cycler using Applied Biosystems® TaqMan Environmental Master Mix (EMM) 2.0 (Life Technologies, Grand Island, NY, USA). The real-time PCR assay was carried out in a 20 μL volume and contained 2 μL DNA, 10 μL 2× EMM, 3 μL of forward and reverse primers and probe mixture (100 nM), and 5 μL DNase-free water. The cycling conditions consisted of a 2 min hold at 50 °C, then a 10 min hold at 95 °C, followed by 40 cycles of 15 s at 95 °C and 1 min at 60 °C. The standards of Campylobacter spp. (MG1) were diluted to yield a series of 10-fold concentrations ranging from 101 to 107 copies, prepared in triplicates, and then used for standard curves.

3. Results

3.1. Bird Activities and Campylobacter Genetic Analysis from Banklick Creek Wetlands

At the Banklick Creek CTW (Figure S1), we observed that bird occupancies varied among the five sites with different vegetative covers. At sites 1 and 2 (no vegetative cover), negligible numbers of birds (i.e., 1) were observed only during one and two sampling events for these sites, respectively. At site 3 (most vegetative cover), birds were observed during 60% of the sampling events (ranging from 1 to 10 birds). Decreased bird occupancy was noted at site 4 (less vegetative cover) with ranges of 2–10 birds present in only 20% of sampling events. At site 5 (outlet), no bird occupancy was observed during the entirety of this study. The previously detected Campylobacter quantity showed a positive correlation with bird occupancies and avian marker qPCR signals [11].
We next analyzed the composition of the Campylobacter isolates cloned from sites 3 and 5 at the CTW. To assess the identity of the isolates present at the two sites and compare the taxonomic composition of isolates derived from each location, we cloned and sequenced the 16s rRNA region of 28 isolates from both sites and phylogenetically compared them to 16s sequences of various Campylobacter species (Figure 1). A majority of the 28 isolates clustered within a clade containing C. canadensis; however, the 16s sequences were, on average, only 94.4% identical between C. canadensis and the unknown clones. These data suggested that the clones may be a unique species compared to C. canadensis [34]. Accordingly, significant efforts were made to isolate Campylobacter spp. from the Canada goose fecal samples collected from the Banklick Creek CTW. The isolation effort initially targeted Canada goose fecal samples collected from different dates at the CTW, but this effort failed likely due to the age of feces collected. Instead, we acquired fresh Canada goose feces from local locations (e.g., Mason, Ohio) for our analysis.

3.2. Campylobacter MG1 Genome and MG1 Phylogenomic and Phylogenetic Analysis

After successfully isolating a novel Campylobacter isolate, termed MG1, from goose fecal samples (see materials and methods) we performed genome sequencing. The assembled MG1 genome was of good quality with high coverage, containing 1.9 Mb of genomic sequence composed of 92 contigs with an N50 length of 1.7 × 105 bp. Our annotation efforts produced 1908 genes (Table 1). We used phylogenomics to better understand the phylogenetic relationship of MG1 to other Campylobacter species. The novel MG1 isolate clustered within a distinct clade that includes C. canadensis and a recently published Campylobacter isolate taken from a cloacal swab of a Canada goose from California (strain RM12654, accession: GCA_020137585.1) (Figure 2). To further understand the relationship between the clade containing C. canadensis, strain RM12654, and MG1 we directly aligned 16s rRNA sequences and performed an average nucleotide identity (ANI) analysis among the genomes. Compared to C. canadensis and RM12654 (GenBank: CP035946.1 and MW131451.1, respectively), the MG1 16s rRNA sequences were 96.8% and 100% similar, and the ANI values were 74.6% and 88.8%, respectively. The level of variation (3.2%) in the 16S rRNA sequences between the two clades (MG1 and C. canadensis) was significant according to [34]. Next, we repeated our phylogenetic analysis of the CTW clones and integrated the 16s rRNA sequence from MG1. The 20 clones of interest clustered within a clade containing the MG1 isolate and were on average 98.5% identical (97.7–98.9%) (Figure 1). The variations (1.2%) among the 20 clones suggested a reasonable intraspecies diversity according to previous analysis for Campylobacter species [34]. The nearest outgroup to this clade was C. canadensis—matching our initial analysis and the above phylogenomics (Figure 1 and Figure 2).

3.3. Virulence Gene Content and Antibiotic Resistome of the MG1 Strain

To determine the virulence gene content of the MG1 isolate, we assembled a list of potential virulence genes reported to contribute to the virulence of Campylobacter in the literature [35,36,37,38]. These included categories of virulence such as genes and gene clusters involved in adhesion, invasion, CPS synthesis, LOS synthesis, and type IV secretion (T4SS). We used amino acid sequences of these genes to query the amino acid sequences of annotated genes from the MG1 isolate using BLASTp.
Most categories of virulence factors were identified in the MG1 genome, including factors associated with adhesion and colonization (cadF, jlpA, racR, racS, and peb genes), invasion (ciaB, iamA, and some but not all flagella-associated genes), T4SS genes (albeit with weak orthology), and other factors. We did not detect toxin-producing genes in MG1 (cdt genes) (Table 2). Next, we explored genes associated with LOS and CPS. Specifically, we searched for the genes involved in sialic acid synthesis and translocation (cst-III, neuB1, neuC1, and neuA1) that define the group 1 category of the LOS locus, including types A, B, C, R, M, and V—the types often associated with neural disease development (including Guillain–Barré syndrome and Miller Fisher syndrome) [36]. MG1 possessed orthologs of all 4 LOS genes associated with group 1 LOS types (Table 3). We also explored the CPS loci in MG1 and found an intact CPS region containing ~19 CPS orthologs, 12 of which were syntenous across a singular contig. However, we cannot rule out the possibility that the fractionation of our genome assembly hindered our ability to detect synteny of the entire CPS locus (Table 4). Lastly, we checked for the presence of antimicrobial resistance (AMR) genes in the MG1 isolate but did not detect any AMR-associated genes.

3.4. MG1-Specific Detection in Canada Goose Fecal Samples

As a final step in our analysis, we designed primers and a probe specific to the MG1 isolate 16s rRNA (see materials and methods) and tested our assay against an array of Campylobacter species to confirm the primer/probe specificity (Table S2). Once confirmed, we assessed the prevalence of the MG1 isolate directly from Canada goose fecal samples we collected from the areas just outside of sites 3 and 5 at Banklick Creek CTW. MG1 was detected in each of the samples we collected (Table S3).

4. Discussion

4.1. Bird Activities and Campylobacter Genetic Analysis from Banklick Creek Wetlands

We found all but eight of the sequenced isolates clustered within a clade containing MG1. We then confirmed that these isolates were likely the same species based on 16s rRNA homology. This finding implies that MG1 is pervasive in the CTW where Canada geese frequent. The nearest outgroup to the MG1 clade was C. canadensis, a novel Campylobacter isolate obtained from cloacal swabs from captive adult whooping cranes (Grus americana) [39]. A small subset of three clones from site 3 formed a clade with C. lanienae, a natural inhabitant of healthy farm and feral animals [40,41,42,43]. The remaining five clones clustered with the additional Campylobacter spp. included in the phylogeny. The prevalence of Canada geese within the CTW suggested that geese were the predominant source of MG1 detected in the wetland area. In light of these data, future studies will focus on the prevalence of MG1 in other habitats that have high Canada goose occupancy.

4.2. Campylobacter MG1 Genome and MG1 Phylogenomic and Phylogenetic Analysis

Our initial phylogenomics suggested that MG1 is a novel or unique isolate separate from C. canadensis. These data also suggest that the MG1 isolate is likely the same species as the RM12654 strain [31,34]. This close relationship to the RM12654 strain from a Canada goose in California indicates that MG1 and MG1-related species may be present in Canada geese from across the United States. Nevertheless, a DNA sequence analysis of 16s rRNA genes for eight isolates from chicken carcasses collected from Saudi Arabia showed the highest identity to Campylobacter strain RM12654 16s ribosomal RNA gene [44], suggesting that this Campylobacter sp., originally isolated from Canada geese, is also able to reside in chickens, although it is currently unclear which animal could be the original host. The host lability of Campylobacter spp. and their ability to overlap between humans, domestic animals, and wild birds further highlights the public health concern generated by the ability of wild birds to mobilize these strains across large geographic distances [10].
Altogether, our data support our previous results describing Campylobacter spp. present at sites 3 and 5 [11]. Our data also show that the majority of the clones we sequenced from the sites were phylogenetically identified as most closely related to MG1. Unfortunately, our sampling scheme was not structured to target times of high goose occupancy. In addition to testing pervasiveness across environments, future studies should monitor how the abundance of MG1 changes in single sites within the CTW over time with the fluctuation in Canada geese occupancy.

4.3. Virulence Gene Content and Antibiotic Resistome of the MG1 Strain

Our analysis of virulence gene content in MG1 showed that the isolate contains many genes associated with virulent tendencies among Campylobacter spp. The presence of LOS group 1 and CPS loci indicate that this isolate can trigger neural diseases often associated with other Campylobacter spp.—although the direct association with these loci and disease is an active area of research [36]. Despite these findings, the percent amino acid similarity we detected was generally low, although this is likely a product of the evolutionary distance between our query sequences (predominantly C. jejuni) and the MG1 isolate. Further, we failed to identify cdt genes within MG1 which, when absent in C. jejuni, was shown previously to lessen the severity of infections [45]. We did not detect isolates similar to MG1 derived from human patients infected with Campylobacter. Combined, this may suggest that MG1 poses minimal infectious risk to humans, and the virulence genes we detected generally function to facilitate infection in avian hosts. Despite this, the genetic potential for future human-related virulence cannot be ruled out. The absence of AMR and toxin-producing genes present in MG1 lessens the danger of human and animal transmission or health effects should MG1 be found in non-avian hosts in the future.
The lack of AMR genes is a promising finding given the increasing prevalence of resistance to common antibiotics found among Campylobacter spp. frequently found in wild birds [10,46]. These findings may imply a lack of significant crossover between MG1, MG1 avian vectors, and human/human-related waste, as AMR prevalence is associated with the propensity for microbial hosts to reside within and feed upon human waste [47], although we cannot be certain of the causitive nature of AMR absence in MG1. It is also important to consider that these findings are all based on in silico genome exploration and gene orthology detection. Accordingly, these data are a first-pass examination of the MG1 isolate and ultimately require the addition of future experiments using in vivo infectious model systems to confirm bona fide virulent tendencies of the isolate.

4.4. MG1-Specific Detection in Canada Goose Fecal Samples

Positive detection of MG1 in Canada geese fecal samples via our designed real-time PCR assay confirmed the pervasive nature of the MG1 isolate within Canada geese. Combined with the original isolation of MG1 in CTW and Mason, Ohio along with the detection of RM12654 from a Canada goose in California (GenBank: GCA_020137585.1) (see Figure 2), these findings highlight the likelihood of widespread presence of MG1 in this avian species and environments frequented by them. Unfortunately, because of sample degradation, we were unable to use our assay to directly quantify MG1 content within the Banklick Creek sites from which the isolate was initially detected. The real-time PCR assay we developed will be a critical tool in addressing this and other MG1-related questions in future work.

5. Summary and Conclusions

We have detected a novel Campylobacter isolate, MG1, that was isolated from Canada goose fecal samples. We provided a draft genome assembly and genome overview with a particular emphasis on genetic signatures most often associated with virulence and disease, vis à vis a risk to human and animal health. We have also generated specific primer and probe combinations that can be used for tracking Canada goose fecal contamination. Using these assays, we confirmed the prevalence of MG1 in Canada goose fecal material. Our clone sequencing also showed a high occurrence of MG1-related isolates across a wetland ecosystem frequented by geese. Our findings emphasize the role that waterfowl, specifically Canada geese, may play in acting as a distribution vector for Campylobacter species including MG1. In this study, we adopted a selective enrichment approach to identify thermotolerant Campylobacter isolates derived from birds that are capable of thriving at human body temperature (37 °C). This allowed us to assess their zoonotic potential and the likelihood of cross-species infection in humans. Moreover, our analysis revealed the presence of multiple MG1 virulence factors, which have been linked to virulence among Campylobacter spp. These findings suggest that MG1 may pose a significant threat to human health in the future. Through our MG1-specific assay, future work should focus on examining the prevalence and distribution of this isolate in the environment and, if cases should arise, in human clinical isolates. Combined, these findings can be used to better understand the spread, occurrence, and microbial sources of Campylobacter species present in public water sources, and the risk these organisms pose to public health. The identification of this new Campylobacter isolate, MG1, in waterbodies used for source drinking water further highlights the importance of using a One Health approach to consider environmental sources and wild animals as another potential reservoir for human-infectious Campylobacter spp., and not just anthropogenic or animal/livestock used for human consumption as sources.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/microorganisms11030648/s1.

Author Contributions

D.M.L. performed genome assembly, all in silico analyses, and prepared the manuscript. K.D.M. performed DNA extraction and the real-time PCR experiments. I.S. performed sample preparation, genome sequencing, MG1 isolation, and management of sample processing. S.K. performed sampling, PCR, and clone-sequencing from Banklick Creek. B.R.M. established collaboration with Sanitation District 1 (SD1) of Northern Kentucky, gained access to Banklick Creek treatment wetland, collected water samples, and developed the research and sampling plan. R.A.H. and E.N.V. provided scientific and technical directions and participated in discussions of sampling plan and processing. J.L. provided scientific and technical directions and maintained responsibility for the whole study. All authors participated in editing and review of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Office of Research and Development, U.S. EPA under its research program: Safe and Sustainable Water Resources (SSWR) 403.1.1.

Data Availability Statement

All raw genome sequencing data have been deposited in the NCBI sequence read archive under accession number PRJNA875047. The assembled genome sequence has been deposited at DDBJ/ENA/GenBank under accession number JANYME000000000.

Acknowledgments

This research was supported by the Office of Research and Development, U.S. EPA under its research program: Safe and Sustainable Water Resources (SSWR) 403.1.1. We also thank Craig Frye, Sanitation District No.1, for assistance with sampling collection. This project was also supported in part by an appointment to the Research Participation Program at the ORD, U.S. Environmental Protection Agency for D.M.L. and K.M., administered by the Oak Ridge Institute for Science and Education through an interagency agreement between the U.S. Department of Energy and U.S. EPA. The views expressed in this manuscript are those of the authors and do not necessarily represent the views or policies of the U.S. EPA. It has been subjected to Agency review and approved for publication. Any use of trade, firm, or product names is for descriptive purposes only and does not imply endorsement by the U.S. Government.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dolbeer, R.A.; Seubert, J.L.; Begier, M.J. Population trends of resident and migratory Canada geese in relation to strikes with civil aircraft. Hum. Wildl. Interact. 2014, 8, 88–99. [Google Scholar]
  2. Titchenell, M.A.; Lynch, W.E., Jr. Coping with Canada geese: Conflict management and damage prevention strategies. In The Ohio State University Agriculture and Natural Resources Fact Sheet; The Ohio State University: Columbus, OH, USA, 2010. [Google Scholar]
  3. Clark, L. A review of pathogens of agricultural and human health interest found in Canada Geese. USDA Natl. Wildl. Res. Cent. Staff. Publ. 2003, 205, 326–334. [Google Scholar]
  4. Elmberg, J.; Berg, C.; Lerner, H.; Waldenström, J.; Hessel, R. Potential disease transmission from wild geese and swans to livestock, poultry and humans: A review of the scientific literature from a One Health perspective. Infect. Ecol. Epidemiol. 2017, 7, 1300450. [Google Scholar] [CrossRef] [PubMed]
  5. Devane, M.; Nicol, C.; Ball, A.; Klena, J.; Scholes, P.; Hudson, J.; Baker, M.; Gilpin, B.; Garrett, N.; Savill, M. The occurrence of Campylobacter subtypes in environmental reservoirs and potential transmission routes. J. Appl. Microbiol. 2005, 98, 980–990. [Google Scholar] [CrossRef] [PubMed]
  6. Edge, T.A.; Hill, S. Multiple lines of evidence to identify the sources of fecal pollution at a freshwater beach in Hamilton Harbour, Lake Ontario. Water Res. 2007, 41, 3585–3594. [Google Scholar] [CrossRef] [PubMed]
  7. Fallacara, D.; Monahan, C.; Morishita, T.; Wack, R. Fecal shedding and antimicrobial susceptibility of selected bacterial pathogens and a survey of intestinal parasites in free-living waterfowl. Avian Dis. 2001, 45, 128–135. [Google Scholar] [CrossRef]
  8. Gorham, T.J.; Lee, J. Pathogen Loading from Canada Geese Faeces in Freshwater: Potential Risks to Human Health through Recreational Water Exposure. Zoonoses Public Health 2016, 63, 177–190. [Google Scholar] [CrossRef]
  9. Vogt, N.A.; Pearl, D.L.; Taboada, E.N.; Mutschall, S.K.; Janecko, N.; Reid-Smith, R.; Jardine, C.M. Epidemiology of Campylobacter, Salmonella and antimicrobial resistant Escherichia coli in free-living Canada geese (Branta canadensis) from three sources in southern Ontario. Zoonoses Public Health 2018, 65, 873–886. [Google Scholar] [CrossRef]
  10. Ahmed, N.A.; Gulhan, T. Campylobacter in Wild Birds: Is It an Animal and Public Health Concern? Front. Microbiol. 2022, 12, 812591. [Google Scholar] [CrossRef]
  11. McMINN, B.R.; Klemm, S.; Korajkic, A.; Wyatt, K.M.; Herrmann, M.P.; Haugland, R.A.; Lu, J.; Villegas, E.N.; Frye, C. A constructed wetland for treatment of an impacted waterway and the influence of native waterfowl on its perceived effectiveness. Ecol. Eng. 2019, 128, 48–56. [Google Scholar] [CrossRef]
  12. Thakur, S.; Zhao, S.; McDermott, P.F.; Harbottle, H.; Abbott, J.; English, L.; Gebreyes, W.A.; White, D.G. Antimicrobial Resistance, Virulence, and Genotypic Profile Comparison of Campylobacter jejuni and Campylobacter coli Isolated from Humans and Retail Meats. Foodborne Pathog. Dis. 2010, 7, 835–844. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Lara-Tejero, M.; Galán, J.E. CdtA, CdtB, and CdtC Form a Tripartite Complex That Is Required for Cytolethal Distending Toxin Activity. Infect. Immun. 2001, 69, 4358–4365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Purdy, D.; Buswell, C.M.; Hodgson, A.E.; McAlpine, K.; Henderson, I.; Leach, S.A. Characterisation of cytolethal distending toxin (CDT) mutants of Campylobacter jejuni. J. Med. Microbiol. 2000, 49, 473–479. [Google Scholar] [CrossRef] [PubMed]
  15. Whitehouse, C.A.; Balbo, P.B.; Pesci, E.C.; Cottle, D.L.; Mirabito, P.M.; Pickett, C.L. Campylobacter jejuni Cytolethal Distending Toxin Causes a G2-Phase Cell Cycle Block. Infect. Immun. 1998, 66, 1934–1940. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Wassenaar, T.M.; Bleumink-Pluym, N.M.; Van Der Zeijst, B.A. Inactivation of Campylobacter jejuni flagellin genes by homologous recombination demonstrates that flaA but not flaB is required for invasion. EMBO J. 1991, 10, 2055–2061. [Google Scholar] [CrossRef]
  17. Monteville, M.R.; Konkel, M.E. Fibronectin-Facilitated Invasion of T84 Eukaryotic Cells by Campylobacter jejuni Occurs Preferentially at the Basolateral Cell Surface. Infect. Immun. 2002, 70, 6665–6671. [Google Scholar] [CrossRef] [Green Version]
  18. Lye, D.; Struewing, I.; Gruber, T.M.; Oshima, K.; Villegas, E.N.; Lu, J. A Gallus gallus model for determining infectivity of zoonotic Campylobacter. Front. Microbiol. 2019, 10, 2292. [Google Scholar] [CrossRef]
  19. Hsieh, Y.-H.; Simpson, S.; Kerdahi, K.; Sulaiman, I.M. A Comparative Evaluation Study of Growth Conditions for Culturing the Isolates of Campylobacter spp. Curr. Microbiol. 2018, 75, 71–78. [Google Scholar] [CrossRef]
  20. Linton, D.; Owen, R.; Stanley, J. Rapid identification by PCR of the genus Campylobacter and of five Campylobacter species enteropathogenic for man and animals. Res. Microbiol. 1996, 147, 707–718. [Google Scholar] [CrossRef]
  21. Andrews, S. FastQC: A Quality Control Tool for High throughput Sequence Data. 2010. Available online: https://www.bioinformatics.babraham.ac.uk/projects/fastqc/ (accessed on 4 September 2022).
  22. Bolger, A.M.; Lohse, M.; Usadel, B. Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics 2014, 30, 2114–2120. [Google Scholar] [CrossRef] [Green Version]
  23. Prjibelski, A.; Antipov, D.; Meleshko, D.; Lapidus, A.; Korobeynikov, A. Using SPAdes de novo assembler. Curr. Protoc. Bioinform. 2020, 70, e102. [Google Scholar] [CrossRef] [PubMed]
  24. Seemann, T. Prokka: Rapid Prokaryotic Genome Annotation. Bioinformatics 2014, 30, 2068–2069. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Gurevich, A.; Saveliev, V.; Vyahhi, N.; Tesler, G. QUAST: Quality assessment tool for genome assemblies. Bioinformatics 2013, 29, 1072–1075. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Alcock, B.P.; Raphenya, A.R.; Lau, T.T.Y.; Tsang, K.K.; Bouchard, M.; Edalatmand, A.; Huynh, W.; Nguyen, A.-L.V.; Cheng, A.A.; Liu, S.; et al. CARD 2020: Antibiotic resistome surveillance with the comprehensive antibiotic resistance database. Nucleic Acids Res. 2020, 48, D517–D525. [Google Scholar] [CrossRef] [PubMed]
  27. Camacho, C.; Coulouris, G.; Avagyan, V.; Ma, N.; Papadopoulos, J.; Bealer, K.; Madden, T.L. BLAST+: Architecture and applications. BMC Bioinform. 2009, 10, 421. [Google Scholar] [CrossRef] [Green Version]
  28. Lund, M.; Nordentoft, S.; Pedersen, K.; Madsen, M. Detection of Campylobacter spp. in Chicken Fecal Samples by Real-TimePCR. J. Clin. Microbiol. 2004, 42, 5125–5132. [Google Scholar] [CrossRef] [Green Version]
  29. Lu, J.; Ryu, H.; Domingo, J.W.S.; Griffith, J.F.; Ashbolt, N. Molecular detection of Campylobacter spp. in California gull (Larus californicus) excreta. Appl. Environ. Microbiol. 2011, 77, 5034–5039. [Google Scholar] [CrossRef] [Green Version]
  30. Lee, M.D. GToTree: A user-friendly workflow for phylogenomics. Bioinformatics 2019, 35, 4162–4164. [Google Scholar] [CrossRef] [Green Version]
  31. Konstantinidis, K.T.; Tiedje, J.M. Prokaryotic taxonomy and phylogeny in the genomic era: Advancements and challenges ahead. Curr. Opin. Microbiol. 2007, 10, 504–509. [Google Scholar] [CrossRef]
  32. Yoon, S.-H.; Ha, S.-M.; Lim, J.; Kwon, S.; Chun, J. A large-scale evaluation of algorithms to calculate average nucleotide identity. Antonie Van Leeuwenhoek 2017, 110, 1281–1286. [Google Scholar] [CrossRef]
  33. Kumar, S.; Stecher, G.; Li, M.; Knyaz, C.; Tamura, K. MEGA X: Molecular Evolutionary Genetics Analysis across Computing Platforms. Mol. Biol. Evol. 2018, 35, 1547–1549. [Google Scholar] [CrossRef] [PubMed]
  34. Gorkiewicz, G.; Feierl, G.; Schober, C.; Dieber, F.; Köfer, J.; Zechner, R.; Zechner, E.L. Species-Specific Identification of Campylobacters by Partial 16S rRNA Gene Sequencing. J. Clin. Microbiol. 2003, 41, 2537–2546. [Google Scholar] [CrossRef] [Green Version]
  35. Dasti, J.I.; Tareen, A.M.; Lugert, R.; Zautner, A.E.; Groß, U. Campylobacter jejuni: A brief overview on pathogenicity-associated factors and disease-mediating mechanisms. Int. J. Med. Microbiol. 2010, 300, 205–211. [Google Scholar] [CrossRef] [PubMed]
  36. Hameed, A.; Woodacre, A.; Machado, L.R.; Marsden, G.L. An Updated Classification System and Review of the Lipooligosaccharide Biosynthesis Gene Locus in Campylobacter jejuni. Front. Microbiol. 2020, 11, 677. [Google Scholar] [CrossRef] [PubMed]
  37. Müller, J.; Schulze, F.; Müller, W.; Hänel, I. PCR detection of virulence-associated genes in Campylobacter jejuni strains with differential ability to invade Caco-2 cells and to colonize the chick gut. Vet. Microbiol. 2006, 113, 123–129. [Google Scholar] [CrossRef] [PubMed]
  38. Rojas, J.D.; Reynolds, N.D.; Pike, B.L.; Espinoza, N.M.; Kuroiwa, J.; Jani, V.; Ríos, P.A.; Nunez, R.G.; Yori, P.P.; Bernal, M.; et al. Distribution of Capsular Types of Campylobacter jejuni Isolates from Symptomatic and Asymptomatic Children in Peru. Am. J. Trop. Med. Hyg. 2019, 101, 541–548. [Google Scholar] [CrossRef]
  39. Inglis, G.D.; Hoar, B.M.; Whiteside, D.P.; Morck, D.W. Campylobacter canadensis sp. nov., from captive whooping cranes in Canada. Int. J. Syst. Evol. Microbiol. 2007, 57, 2636–2644. [Google Scholar] [CrossRef] [Green Version]
  40. Acik, M.; Karahan, M.; Ongor, H.; Karagulle, B.; Cetinkaja, B. The first isolation of Campylobacter lanienae from chickens. Revue. Méd. Vét. 2013, 164, 368–373. [Google Scholar]
  41. Inglis, G.; Kalischuk, L.; Busz, H. A survey of Campylobacter species shed in faeces of beef cattle using polymerase chain reaction. Can. J. Microbiol. 2003, 49, 655–661. [Google Scholar] [CrossRef]
  42. Sasaki, Y.; Fujisawa, T.; Ogikubo, K.; Ohzono, T.; Ishihara, K.; Takahashi, T. Characterization of Campylobacter lanienae from Pig Feces. J. Vet. Med. Sci. 2003, 65, 129–131. [Google Scholar] [CrossRef] [Green Version]
  43. Schweitzer, N.; Damjanova, I.; Kaszanyitzky, E.; Ursu, K.; Samu, P.; Tóth, Á.G.; Dan, A. Molecular characterization of Campylobacter lanienae strains isolated from food-producing animals. Foodborne Pathog. Dis. 2011, 8, 615–621. [Google Scholar] [CrossRef] [PubMed]
  44. Alarjani, K.M.; Elkhadragy, M.F.; Al-Masoud, A.H.; Yehia, H.M. Detection of Campylobacter jejuni and Salmonella typhimurium in chicken using PCR for virulence factor hipO and invA genes (Saudi Arabia). Biosci. Rep. 2021, 41, BSR20211790. [Google Scholar] [CrossRef] [PubMed]
  45. Sen, K.; Lu, J.; Mukherjee, P.; Berglund, T.; Varughese, E.; Mukhopadhyay, A.K. Campylobacter jejuni Colonization in the Crow Gut Involves Many Deletions within the Cytolethal Distending Toxin Gene Cluster. Appl. Environ. Microbiol. 2018, 84, e01893-01817. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Du, J.; Luo, J.; Huang, J.; Wang, C.; Li, M.; Wang, B.; Wang, B.; Chang, H.; Ji, J.; Sen, K.; et al. Emergence of Genetic Diversity and Multi-Drug Resistant Campylobacter jejuni from Wild Birds in Beijing, China. Front. Microbiol. 2019, 10, 2433. [Google Scholar] [CrossRef] [PubMed]
  47. Gil, J.C.; Hird, S.M. Multi-omics characterization of the Canada goose fecal microbiome reveals selective efficacy of simulated metagenomes. Microbiol. Spectr. 2022, 10, e02384-22. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Phylogenetic analysis of Campylobacter 16s sequences isolated from Banklick Creek Wetlands site 3 (blue dots) and 5 (purple dots). The neighbor-joining phylogenetic tree is based on nucleotide alignments of 38 sequences composed of 661 nucleotide positions. The optimal tree is shown and drawn to scale in units of the number of base substitutions per site (scale bar = 0.02 substitutions). The percentage of replicate trees in which the associated taxa clustered together in the bootstrap test (500 replicates) are shown next to the branches. E. coli (strain KBKVG4) was used to root the tree. The majority of sequenced isolates clustered within a clade containing MG1 (red vertical line). The closest outgroup for the MG1-containing clade was occupied by C. canadensis (green vertical line).
Figure 1. Phylogenetic analysis of Campylobacter 16s sequences isolated from Banklick Creek Wetlands site 3 (blue dots) and 5 (purple dots). The neighbor-joining phylogenetic tree is based on nucleotide alignments of 38 sequences composed of 661 nucleotide positions. The optimal tree is shown and drawn to scale in units of the number of base substitutions per site (scale bar = 0.02 substitutions). The percentage of replicate trees in which the associated taxa clustered together in the bootstrap test (500 replicates) are shown next to the branches. E. coli (strain KBKVG4) was used to root the tree. The majority of sequenced isolates clustered within a clade containing MG1 (red vertical line). The closest outgroup for the MG1-containing clade was occupied by C. canadensis (green vertical line).
Microorganisms 11 00648 g001
Figure 2. Phylogenomic analysis of the MG1 isolate. The phylogenomic tree was generated using 118 possible single copy genes specific to Proteobacteria and generated via GToTree. Escherichia coli (E. coli—GCF_000005845) was used to root the tree. Nodes are annotated with bootstrap support (dot = 100% agreement). The scale bar indicates substitutions per site.
Figure 2. Phylogenomic analysis of the MG1 isolate. The phylogenomic tree was generated using 118 possible single copy genes specific to Proteobacteria and generated via GToTree. Escherichia coli (E. coli—GCF_000005845) was used to root the tree. Nodes are annotated with bootstrap support (dot = 100% agreement). The scale bar indicates substitutions per site.
Microorganisms 11 00648 g002
Table 1. MG1 genome assembly statistics.
Table 1. MG1 genome assembly statistics.
Coverage (x)No. of ContigsAssembly Size (bp)Contig N50 (bp)G+C Content (%)Gene Annotation Data (no.)
GenesCDSrRNAstRNAs
914921,900,729168,48826.75190818663 (2 *)39
* two rRNAs after manual curation.
Table 2. Presumptive MG1 virulence genes identified by BLAST analysis of protein sequences. Query sequences are predominantly from C. jejuni (see Materials and Methods).
Table 2. Presumptive MG1 virulence genes identified by BLAST analysis of protein sequences. Query sequences are predominantly from C. jejuni (see Materials and Methods).
Query IDGene DescriptionCategoryMG1 Best Hit BLAST Results
MG1 prokka_aa * IDIdentity (%)e-Value
YP_002344858.1outer membrane fibronectin-binding (cadF)AdhesionPROKKA_0105441.42.48 × 10−74
YP_002344378.1surface-exposed lipoprotein (jlpA)AdhesionPROKKA_0172026.71.79 × 10−18
YP_002344652.1two-component regulator (racR)AdhesionPROKKA_0156665.52.56 × 10−109
YP_002344653.1sensor histidine kinase (racS)AdhesionPROKKA_0156734.91.30 × 10−56
YP_002344319.1bifunctional adhesin/ABC transporter (peb1A)AdhesionPROKKA_0122428.66.41 × 10−21
YP_002344320.1amino acid ABC transporter ATP-binding protein (pebC)AdhesionPROKKA_0122352.53.95 × 10−90
YP_002344743.1enterochelin uptake substrate-binding protein (ceuE)InvasionNo significant hits
YP_002344312.1invasion antigen B (ciaB)InvasionPROKKA_0152840.02.92 × 10−126
YP_002345016.1ABC transporter ATP-binding protein (iamA)InvasionPROKKA_0011537.02.65 × 10−45
YP_002343524.1flagellar motor switch protein (fliM)InvasionPROKKA_0160127.52.43 × 10−23
YP_002343523.1flagellar motor switch protein (fliY)InvasionPROKKA_0160228.65.29 × 10−23
YP_002344200.1sensor histidine kinase (flgS)InvasionPROKKA_0169848.04.48 × 10−26
YP_002345095.1flagellar hook protein (flgE)InvasionNo significant hits
YP_002344726.1flagellin B (flaB)InvasionNo significant hits
YP_002343773.1flagellar biosynthesis protein (flhB)InvasionNo significant hits
YP_002343959.1flagellar basal body rod protein (flgB)InvasionNo significant hits
YP_002344727.1flagellin A (flaA)InvasionNo significant hits
YP_002344282.1flagellar biosynthesis protein (flhA)InvasionNo significant hits
YP_002344851.1type II protein secretion system protein E (ctsE)otherPROKKA_0119341.43.65 × 10−144
YP_002344617.1sensor histidine kinase (Cj1226c)otherPROKKA_0119751.24.16 × 10−137
YP_002344072.1response regulator domain protein (cbrR)otherPROKKA_0016841.97.56 × 10−107
YP_002344977.1two-component regulator (Cj1608)otherPROKKA_0149053.28.10 × 10−104
YP_002344737.1fibronectin/fibrinogen-binding (Cj1349c)otherPROKKA_0112241.73.04 × 10−98
YP_002344670.1fibronectin domain-containing lipoprotein (Cj1279c)otherPROKKA_0024639.24.35 × 10−94
YP_002344872.1two-component sensor (Cj1492c)otherPROKKA_0172238.09.64 × 10−83
YP_002344532.1beta-1,3 galactosyltransferase (wlaN)otherPROKKA_0046653.43.78 × 10−69
YP_002343492.1cytochrome C551 peroxidase (docA)otherPROKKA_0079840.61.68 × 10−63
YP_002344871.1two-component regulator (Cj1491c)otherPROKKA_0172145.71.85 × 10−54
YP_002343491.1MCP-domain signal transduction protein (Cj0019c)otherPROKKA_0176337.41.34 × 10−33
YP_002344318.1amino acid ABC transporter permease (Cj0920c)otherPROKKA_0122129.78.74 × 10−33
YP_002344289.1putative sensory transduction transcriptional regulator (Cj0890c)otherPROKKA_0124127.74.18 × 10−27
YP_002344614.1putative two-component regulator (dccR)otherPROKKA_0119630.02.33 × 10−24
WP_002826431.1virB11T4SSPROKKA_0106928.81.5 × 10−37
WP_012662267.1virB10T4SSNo significant hits
WP_002834097.1virB9T4SSNo significant hits
WP_012662258.1virB4T4SSPROKKA_0071123.52.12 × 10−5
WP_012662270.1virD4T4SSPROKKA_0021426.53.64 × 10−11
YP_002343541.1cytolethal distending toxin B (cdtB)ToxinNo significant hits
YP_002343539.1cytolethal distending toxin C (cdtC)ToxinNo significant hits
YP_002343540.1cytolethal distending toxin A (cdtA)ToxinNo significant hits
* prokka_aa are the prokka annotated amino acid sequences from MG1.
Table 3. Presumptive MG1 group 1 lipooligosaccharide loci identified by BLAST analysis of protein sequences.
Table 3. Presumptive MG1 group 1 lipooligosaccharide loci identified by BLAST analysis of protein sequences.
C. jejuni NCTC11168MG1 Best Hit BLAST Result
Query IDNameMG1 prokka_aa * IDIdentity (%)e-Value
YP_002344533.1cstIIIPROKKA_0046136.77.89 × 10−40
YP_002344534.1neuB1PROKKA_0046264.39.39 × 10−154
YP_002344535.1neuC1PROKKA_0046349.64.04 × 10−117
YP_002344536.1neuA1PROKKA_0046439.66.58 × 10−33
* prokka_aa are the prokka annotated amino acid sequences from MG1.
Table 4. Presumptive MG1 capsular polysaccharide loci identified by BLAST analysis of protein sequences.
Table 4. Presumptive MG1 capsular polysaccharide loci identified by BLAST analysis of protein sequences.
C. jejuni NCTC11168MG1 Best Hit BLAST Results
Query IDGene Name *MG1 prokka_aa ** IDSyntenous ***Identity (%)e-Value
YP_002344796.1kpsSPROKKA_0018044.51.11 × 10−109
YP_002344797.1kpsCPROKKA_0017254.40
YP_002344798.1cysCNo significant hits
YP_002344799.1PROKKA_00458 33.83.12 × 10−8
YP_002344800.1PROKKA_00080 23.04.74 × 10−7
YP_002344801.1No significant hits
YP_002344802.1PROKKA_00794 32.66.31 × 10−9
YP_002344803.1No significant hits
YP_002344804.1No significant hits
YP_002344805.1No significant hits
YP_002344806.1hddCPROKKA_01216 29.02.93 × 10−10
YP_002344807.1gmhA2PROKKA_00892 49.12.72 × 10−47
YP_002344808.1hddANo significant hits
YP_002344809.1No significant hits
YP_002344810.1PROKKA_01532 26.74.69 × 10−17
YP_002344811.1fclNo significant hits
YP_002344812.1No significant hits
YP_002344813.1rfbCNo significant hits
YP_002344814.1hddCNo significant hits
YP_002344815.1No significant hits
YP_002344816.1No significant hits
YP_002344817.1PROKKA_0017636.05.90 × 10−23
YP_002344818.1No significant hits
YP_002344819.1PROKKA_0018925.93.38 × 10−11
YP_002344820.1PROKKA_0018930.65.00 × 10−20
YP_002344821.1PROKKA_0017634.28.19 × 10−22
YP_002344822.1glfPROKKA_0017536.82.25 × 10−78
YP_002344823.1PROKKA_0017333.57.61 × 10−52
YP_002344824.1kfiDNo significant hits
YP_002344825.1No significant hits
YP_002344826.1kpsFPROKKA_00337 42.03.12 × 10−69
YP_002344827.1kpsDPROKKA_0018455.10
YP_002344828.1kpsEPROKKA_0018343.35.24 × 10−87
YP_002344829.1kpsTPROKKA_0018263.62.84 × 10−110
YP_002344830.1kpsMPROKKA_0018143.93.50 × 10−67
* Dash symbol indicates genes are unnamed, ** prokka_aa are the prokka annotated amino acid sequences from MG1, *** hits are located along the same contig. Dot (•) confirms synteny.
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

Linz, D.M.; McIntosh, K.D.; Struewing, I.; Klemm, S.; McMinn, B.R.; Haugland, R.A.; Villegas, E.N.; Lu, J. Genomic Characterization and Wetland Occurrence of a Novel Campylobacter Isolate from Canada Geese. Microorganisms 2023, 11, 648. https://doi.org/10.3390/microorganisms11030648

AMA Style

Linz DM, McIntosh KD, Struewing I, Klemm S, McMinn BR, Haugland RA, Villegas EN, Lu J. Genomic Characterization and Wetland Occurrence of a Novel Campylobacter Isolate from Canada Geese. Microorganisms. 2023; 11(3):648. https://doi.org/10.3390/microorganisms11030648

Chicago/Turabian Style

Linz, David M., Kyle D. McIntosh, Ian Struewing, Sara Klemm, Brian R. McMinn, Richard A. Haugland, Eric N. Villegas, and Jingrang Lu. 2023. "Genomic Characterization and Wetland Occurrence of a Novel Campylobacter Isolate from Canada Geese" Microorganisms 11, no. 3: 648. https://doi.org/10.3390/microorganisms11030648

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