Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Characterization of the emerging zoonotic pathogen Arcobacter thereius by whole genome sequencing and comparative genomics

  • Francesca Rovetto,

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Validation, Visualization, Writing – original draft, Writing – review & editing

    Affiliations Department of Veterinary Public Health and Food Safety, Faculty of Veterinary Medicine, Ghent University, Salisburylaan 133, Merelbeke, Belgium, Department of Forestry, Agriculture and Food Sciences, University of Torino, Largo Braccini 2, Grugliasco, Italy

  • Aurélien Carlier,

    Roles Conceptualization, Data curation, Formal analysis, Supervision, Validation, Writing – original draft, Writing – review & editing

    Affiliation Laboratory of Microbiology, Faculty of Sciences, Ghent University, K. L. Ledeganckstraat 35, Ghent, Belgium

  • Anne-Marie Van den Abeele,

    Roles Formal analysis

    Affiliation Microbiology laboratory, Saint-Lucas Hospital, Groenebriel 1, Ghent, Belgium

  • Koen Illeghems,

    Roles Conceptualization, Data curation, Formal analysis, Writing – original draft

    Affiliation Department of Veterinary Public Health and Food Safety, Faculty of Veterinary Medicine, Ghent University, Salisburylaan 133, Merelbeke, Belgium

  • Filip Van Nieuwerburgh,

    Roles Formal analysis

    Affiliation Laboratory of Pharmaceutical Biotechnology, Faculty of Pharmaceutical Sciences, Ghent University, Harelbekestraat 72, Ghent, Belgium

  • Luca Cocolin,

    Roles Funding acquisition, Supervision, Writing – original draft, Writing – review & editing

    Affiliation Department of Forestry, Agriculture and Food Sciences, University of Torino, Largo Braccini 2, Grugliasco, Italy

  • Kurt Houf

    Roles Conceptualization, Funding acquisition, Investigation, Methodology, Project administration, Supervision, Writing – original draft, Writing – review & editing

    Kurt.houf@UGent.be

    Affiliation Department of Veterinary Public Health and Food Safety, Faculty of Veterinary Medicine, Ghent University, Salisburylaan 133, Merelbeke, Belgium

Abstract

Four Arcobacter species have been associated with human disease, and based on current knowledge, these Gram negative bacteria are considered as potential food and waterborne zoonotic pathogens. At present, only the genome of the species Arcobacter butzleri has been analysed, and still little is known about their physiology and genetics. The species Arcobacter thereius has first been isolated from tissue of aborted piglets, duck and pig faeces, and recently from stool of human patients with enteritis. In the present study, the complete genome and analysis of the A. thereius type strain LMG24486T, as well as the comparative genome analysis with 8 other A. thereius strains are presented. Genome analysis revealed metabolic pathways for the utilization of amino acids, which represent the main source of energy, together with the presence of genes encoding for respiration-associated and chemotaxis proteins. Comparative genome analysis with the A. butzleri type strain RM4018 revealed a large correlation, though also unique features. Furthermore, in silico DDH and ANI based analysis of the nine A. thereius strains disclosed clustering into two closely related genotypes. No discriminatory differences in genome content nor phenotypic behaviour were detected, though recently the species Arcobacter porcinus was proposed to encompass part of the formerly identified Arcobacter thereius strains. The report of the presence of virulence associated genes in A. thereius, the presence of antibiotic resistance genes, verified by in vitro susceptibility testing, as well as other pathogenic related relevant features, support the classification of A. thereius as an emerging pathogen.

Introduction

The genus Arcobacter was created in 1991 as a second genus within the family Campylobacteraceae to include bacteria which differ from the closely related Campylobacter species by their aerotolerance and ability to grow at temperatures below 30° [1]. At the time of writing, 21 species have been characterized including the new species Arcobacter ebronensis, Arcobacter aquimarinus, Arcobacter faecis, Arcobacter lanthierii and Arcobacter pacificus [25]. These species are isolated from environmental matrices or shellfish. Six species are commonly isolated from food of animal sources across the world [6]. In particular the species Arcobacter butzleri, Arcobacter cryaerophilus and Arcobacter skirrowii are incriminated as food and waterborne pathogens for humans [7]. In contrast to animals where infection is asymptomatic, Arcobacter infection in humans seems to cause enteritis and sometimes bacteraemia, with clinical signs similar to those of campylobacteriosis, but with a higher frequency of persistent watery diarrhoea [810]. Contaminated drinking water and the manipulation or consumption of raw or undercooked food are likely to be the infection sources [7].

Recently, the species Arcobacter thereius was isolated from the stool of two hospitalized patients with symptoms of enteritis [9]. This species had first been isolated during a Danish study on the prevalence of campylobacteria in ducks and in internal organs of aborted piglets [11,12]. Some of the isolated Gram negative, rod-shaped, slightly curved, not able to degrade urea, non-spore-forming, oxidase, catalase and nitrate reduction positive, bacteria clustered in a distinct phenon within the genus Arcobacter, and were later characterized by Houf et.al. [13] as representing a novel species, A. thereius. Though the species has been isolated relatively frequently from the faeces of healthy pigs, still almost nothing is known about its behaviour [14]. Houf et al. already reported its lack of growth at 37°C under conditions which cultivate other host-associated arcobacters [13], and the generally slow and difficult in vitro culture has troubled many researchers since. Furthermore, in contrast to the other mammal-associated species, A. thereius isolates are hard to type at strain level, in fact pulsed-field gel electrophoresis (PFGE), amplified fragment-length polymorphism (AFLP) and enterobacterial repetitive intergenic consensus PCR (ERIC-PCR) showed to be not discriminatory for A. thereius [15], and the already identified virulence associated genes in the other human associated Arcobacter species seem to be lacking in A. thereius [16]. Since A. thereius is in vitro non-fermenting, nor oxidising carbohydrates, as for the other members of the family Campylobacteraceae, genome analysis represents an ultimate approach to elucidate this species full potential.

The present study presents the full genome analysis of the A. thereius type strain LMG24486T, as well as the sequences of eight other strains originating from different matrices. The genome of the type strain is analysed with a focus on metabolic pathways, virulence factors, antibiotic resistance genes and adaptation to the environment. The genome is then compared with the only other yet available full genome of a human associated species: A. butzleri type strain RM4018 (= LMG10828T) [17]. Subsequently, strain variation in A. thereius is assessed in order to comprehend the previously reported strain high heterogeneity [13].

Materials and methods

1. Bacterial strains and growth conditions

For genome analysis, nine A. thereius strains were included (Table 1). Five strains were isolated during a Danish surveillance study: strains DU19 and DU22 from duck faeces, and strains LMG24487, 11743–4 and LMG24486T, including the type strain, from the internal organs of spontaneous porcine abortions [11,12]. Four unrelated strains, isolated from porcine faeces, were randomly chosen from the Arcobacter strain bank of the Department of Veterinary Public Health and Food Safety, Faculty of Veterinary, Ghent University, Belgium [14]. Both A. thereius strains previously isolated from humans, have been preserved on pearls, and showed to be no longer culturable, and therefore were not included in the present study.

thumbnail
Table 1. General features of the Arcobacter thereius genome sequences studied.

https://doi.org/10.1371/journal.pone.0180493.t001

Strains were stored at -80°C in full-horse blood. Re-cultivation was performed by inoculation of 10 μL from the stock onto blood agar plates, and incubation at 28°C for 72 h in microaerobic conditions by evacuating 80% of the normal atmosphere and introducing a gas mixture of 8% CO2, 8% H2 and 84% N2 into a jar.

2. Genome analysis

2.1. DNA extraction and genome sequencing.

DNA extraction was performed using the DNeasy Tissue Kit (50) (Qiagen, Venlo, The Netherlands) according to the manufacturer’s instructions, followed by evaluation of the DNA integrity by size separation of 10 μL of each DNA sample by agarose gel electrophoresis (1.5%). An Illumina paired-end library was generated from 1 μg of genomic DNA for each of the nine strains. The DNA was fragmented to 800 bp using Covaris S2 sonication (Covaris, MA, USA) and a sequence library was made for each sample using the NEBNext Ultra DNA Library Prep Kit (New England Biolabs, Ipswich, MA, USA). During the library preparation, a size selection was performed before and after the enrichment PCR using an Invitrogen 1% E-gel (ThermoFisher Scientific, Waltham, MA, USA) to select for fragments between 850 and 1200 bp. The libraries were equimolarly pooled and were sequenced on an Illumina MiSeq sequencer (Illumina, San Diego, CA, USA), generating 2x300 bp paired-end reads for each sequenced fragment. Base calling and primary quality assessments were performed using Illumina’s Basespace genomics cloud computing environment. Using CLC bio Genomics Workbench version 8.0, the reads were trimmed based on quality scores, library prep adapters were removed, and resulting sequences shorter than 100 nt were discarded. The quality trimmed reads were used to perform a de novo assembly with CLC bio Genomics Workbench version 8.0 using standard settings. The resulting scaffolds were further extended and joined into a final set by SSPACE version 2.0 [18] using all quality filtered reads. The remaining gaps inside the final scaffolds were partially or completely removed using GapFiller version 1.10 [19] and the quality filtered reads.

Additionally, the genome of the type strain, A. thereius LMG24486T was sequenced using PacBio technology [20], following the chemistry Polymerase P6/ Chemistry C4.0 and the type of the library “20 kb Template Preparation Using BluePippin Size-Selection system” protocol from PacBio (ref 100-286-000-07) with the actual fragment distribution on BioAnalyser >12kb. The long reads, together with the scaffolds generated by CLC bio Genomics Workbench starting from Illumina paired-end data, were submitted to a hybrid assembly using SSpace-Longread version 1.1 [21]. The gaps in the final scaffolds were partially or completely closed using GapFiller.

2.2. Genome annotation and analysis.

The assembled genome sequences of the nine A. thereius strains were annotated using a local installation of the Prokka platform v1.11 [22], using the feature prediction tools Prodigal v2.60 [23], ARAGORN v1.2 [24], Barrnap v.0.5 (http://www.vicbioinformatics.com/software.barrnap.shtml), Infernal v1.1.1, and SignalP v4.1 [25,26]. Further, protein-encoding sequences (CDS) were annotated by BLAST analysis (e-value cut-off of 10−6 and 50% of identity) using all available bacterial proteins in the UniProt database 2015_04 [27] and HMMER analysis using the Pfam [28] and TIGRFAM [29] databases. The genome sequences of the A. thereius strains were also annotated by the Rapid Annotation using Subsystem Technology (RAST) [30,31] platform. Annotations from Prokka and RAST were merged in one final annotation file in order to reduce the number of hypothetical proteins. CDS of interest were additionally analysed by PSI-BLAST (default setting) and metabolic pathways were constructed using the Kyoto Encyclopedia of Genes and Genomes (KEGG) database [32]. Plasmid sequences were identified using PlasmidFinder version 1.2 [33] and the PATRIC [34] plasmid database (plasmid_seq, version 30/6/2014) and prophage regions were detected using PHAST [35]. Antibiotic resistance genes were searched for by BLASTn analysis using the antibiotic resistance gene-annotation database (ARG-ANNOT; [36]) using default parameters and the comprehensive antibiotic resistance database (CARD, [37]) by BLASTp analysis with an default e-value cut-off of 10−30; virulence factors were identified by BLASTn analysis using the human pathogenic bacteria virulence factor database (VFDB;[3840]) with default parameters.

For A. thereius LMG 24486T, clustered regularly insterspaced short palindromic repeat (CRISPR) regions were identified through CRISPRFinder [41] and a genome plot was generated with DNAPlotter [42]. The origin of chromosomal replication was predicted with the Ori-Finder tool [43]. Genomic islands were searched for with IslandViewer [44] and bacteriocins with BAGEL3 [45].

2.3 Comparative genome analysis.

A maximum likelihood phylogenetic tree of a concatenated alignment of 239 single-copy orthologs was reconstructed. One-to-one ortholog genes were obtained by comparing a genome database comprising the A. thereius strains that have been sequenced in this study, all other Arcobacter genomes available from NCBI and Campylobacter jejuni NCTC11168 as outgroup. Orthologues were predicted using the OrthoMCL software v1.4 [46] using the following settings: BLASTp e-value cut-off: 10−6; identity cut-off 50%; reciprocal hit length cut-off: 50%. Further, in order to include only orthologs without evidence of recombination or gene conversion, a stringent filtering step was added. Briefly, the probability of recombination for each orthologous group was calculated using the PhiPack software [47], and discarding alignments without sufficient phylogenetic signal or with a p-value < 0.05 (with 1000 permutations). Nucleotide corresponding to 239 single copy ortholog sequence sets were aligned with MUSCLE [48], trimmed with TrimAL [49] to remove columns with >80% gaps and concatenated. The maximum likelihood tree was built with RaxML [50] using the GTRGAMMA substitution model and 100 bootstrap replicates.

Comparative analysis between A. thereius LMG24486T and A. butzleri RM4018 and within the A. thereius strains were performed using the OrthoMCL software v1.4 [46] using the following settings: BLASTp e-value cut-off: 10−6; identity cut-off 50%; reciprocal hit length cut-off: 50%. EDGAR [51] comparative genome analysis was also used to obtain the core and the pan genome plot of A. thereius species and to construct a synteny plot between A. thereius LMG24486T and A. butzleri RM4018 following default parameters.

A phylogenetic analysis using MEGA 6 [52] was performed based on the amino acid sequence of the virulence genes that were in common in all nine A. thereius strains. Further, amino acid sequences of Campylobacter and/or Helicobacter species were included in the analysis in the cases where an appropriate amino acid sequence could be retrieved from the NCBI data repository. Each dendrogram was made using the Neighbor-Joining method. Bootstrap values of >50% generated from 100 replicates are shown next to the branches. The evolutionary distance were computed using the p-distance method and are in the units of the number of amino acid differences per site.

The Average Nucleotide Identity based on BLAST+ calculation (ANIb) was performed for the nine A. thereius strains using JSpecies Web Server (JSpace WS) with default parameters [53]. The genome of A. butzleri RM4018 and A. cibarius LMG21996 were used as an outgroup. In addition, in silico DNA-DNA hybridisation was calculated for all A. thereius strains using the genome-to-genome distance calculator following as alignment method the recommended GGDC 2 BLAST+ [54]. Genomes of A. butzleri RM4018 and A. cibarius LMG21996 were included as comparison.

MLST profiles were obtained for the nine A. thereius strains using the Arcobacter pubMLST database (http://pubmlst.org/Arcobacter/) and according to the protocol of Miller et al. [55]. The nucleotides sequences were aligned and concatenated in the order aspA, atpA, glnA, gltA, pmg and tkt. The gene glyA was however not included as it was absent in A. thereius strains 213, 216, 440, 452 and 11743–4. A dendrogram was constructed with MEGA 6 using the Neighbor-joining method. Bootstrap values of >50% generated from 100 replicates are shown next to the branches. The evolutionary distance were computed using the p-distance method and are in units of the number of bases differences per site.

Alignment of the nine genomes of A. thereius has been generate with ProgressiveMauve [56] using default parameters. Clusters of orthologous group families (COG) were assigned, for all the nine A. thereius strains, using the rps-BLAST program against the NCBI Conserved Domain Database (CDD) and with an e-value cut-off of 1.0 × 10−3. Only the top hits were retained and the genes belonging to each COG categories were counted, and the distribution were compared using a Pearson chi-squared test using the IBM SPSS statistic program v23.

2.4 Data availability.

The annotated genome sequences were deposited in the DDBJ/EMBL/GenBank database. The sequencing projects and accession numbers are listed in Table 2.

thumbnail
Table 2. Genbank accession numbers of A. thereius genomes.

https://doi.org/10.1371/journal.pone.0180493.t002

3. Growth, motility and antimicrobial susceptibility of Arcobacter thereius.

The nine A. thereius strains included in the present study were tested for antimicrobial susceptibility, growth and motility capacity. The latter two parameters were tested under different temperature and atmospheric conditions as shown in Table 3. Briefly, cell suspensions of 1 McFarland for each strains were prepared in 2 mL Peptone Water (Oxoid, Basingstoke, UK). For the growth test, a drop of 30 μL of each suspension was placed on blood agar plates. For the motility test, a 100 μl drop of each suspension was placed in the centre of a Petri dish containing a semi-solid media (24 g/L Arcobacter broth, 3 g/L Agar Technical n° 3 (Oxoid)).

thumbnail
Table 3. Characteristics of the phenotypic behaviour and distribution of antimicrobial susceptibility of the nine A. thereius strains sequenced in this study.

Susceptibility breakpoint (mg/L): erythromycin: ≤ 8; ciprofloxacin: ≤ 0,5; ampicillin: ≤ 2; tetracycline: ≤ 2; gentamicin: ≤ 2; streptomycin: ≤ 16; chloramphenicol: ≤ 8; spectinomycin: ≤ 32.

https://doi.org/10.1371/journal.pone.0180493.t003

Based on the genomic findings, antimicrobial susceptibility to eight antibiotics: erythromycin, ciprofloxacin, ampicillin, tetracycline, gentamicin, streptomycin, chloramphenicol and spectinomycin was determined by the gradient strip diffusion method as previously described by Van den Abeele et al. [57]. In brief, an inoculum of 1 McFarland in 0.9% NaCl was prepared, plated on Mueller Hinton agar supplemented with horse blood (MH II-F agar, bioMérieux, Marcy L’étoile, France), and incubated at 35°C in microaerobic atmosphere for 48 hours. Quality control strains Campylobacter jejuni ATCC33560 and A. butzleri RM4018 were included. Due to the fastidious growth, plates were read after 48 hours. EUCAST breakpoints were used for erythromycin, ciprofloxacin and tetracycline (Campylobacter coli), ampicillin, (non-species-related PK/PD), gentamicin and chloramphenicol (Enterobacteriaceae) and spectinomycin (Neisseria gonorrhoeae) [58]. For streptomycin, NARMS resistance breakpoints were applied [59].

Results and discussion

1. General genomic features of the type strain, Arcobacter thereius LMG 24486T

1.1. Genome assembly and annotation.

Hybrid sequence assembly of the Illumina paired-end reads and PacBio long reads resulted in a single contig representing the circular chromosome of A. thereius LMG24486T, for a total sequence length of 1,909,306 bp with a GC content of 26.93% (Tables 1 and 4, Fig 1). The absence of plasmids in A. thereius has been already previously reported, and is confirmed in the present study [60]. Genome annotation revealed the presence of 1925 protein coding genes (CDS), three rRNA operons (all 16S rRNA genes were identical), and 46 tRNA genes. Three Cas/CRISPR regions were found in the genome of A. thereius LMG24486T, consisting of 28 (position 1,211,914–1,213,792), 4 (position 1,242,675–1,243,017), and 27 (position 1,243,120–1,245,145) spacer sequences, respectively (Fig 1; S1 Table). The genome sequence of A. thereius LMG24486T contained no predicted bacteriocin-associated genes, prophages, and no genomic islands.

thumbnail
Fig 1. General features of the Arcobacter thereius LMG24486T genome sequence.

The circles represent (from the outside to the inside): circle 1, position (bp); circle 2, CDS transcribed clockwise; circle 3, CDS transcribed anti-clockwise; circle 4, GC content plotted using default settings; circle 5, GC skew plotted using a window size of 25,000 and default step size. The predicted oriC, difH site, and CRISPR sequences are indicated by black arrows. The difH site had a consecutive 31-bp sequence match with that of A. butzleri RM4018 [61].

https://doi.org/10.1371/journal.pone.0180493.g001

2. Central metabolism

2.1. Sugar metabolism.

Arcobacter thereius LMG24486T is unable to metabolize carbohydrates through the Embden Meyerhof-Parnas pathway due to the absence of genes encoding for the key enzymes 6-phosphofructokinase and glucose-1-phosphate phosphodismutase. However, as all genes encoding for the other enzymes necessary in glycolysis and a gene encoding fructose-1,6-bisphosphatase are present (AA347_01347; AA347_01345; AA347_00800; AA347_01516; AA347_00648; AA347_00647; AA347_00887; AA347_00564; AA347_00705), this suggests that the Embdem Meyerhof pathway operates towards gluconeogenesis. This functionality has also been described in the genome of the asaccharolytic Campylobacter jejuni [62]. All genes encoding enzymes of the non-oxidative branch of the pentose phosphate pathway (PPP) were present (AA347_01712; AA347_00012; AA347_01343; AA347_00280; AA347_00243; AA347_01540), indicating that sugars can be metabolized using this pathway, while the oxidative branch of the PPP is not active. Other pathways that allow bacteria to metabolize sugars, such as the Entner-Doudoroff pathway are not present. In A. thereius, a complete pathway for the degradation of pyruvate, which can be formed through amino acid degradation, is present (see below). A gene cluster encoding the pyruvate dehydrogenase complex allowing formation of acetyl-CoA is present (AA347_01751, AA347_01752, and AA347_01753). The latter can be channelled into the tricarboxylic acid (TCA) cycle or it can be dehydrogenated to acetate by the aldehyde dehydrogenase (AA347_00799). Further, oxaloacetate can be formed from pyruvate by carboxylation by pyruvate carboxylase (AA347_01516). Oxaloacetate can be channelled in the TCA cycle if the quantity of energy is low or it can be converted to glucose if the energy charge is high [63].

All genes encoding the enzymes of the TCA were detected, except for the one encoding the succinyl-CoA synthetase. This conformation has also been reported previously in A. butzleri RM4018 [17] and other species of the family Epsilonproteobacteria, such as Helicobacter pylori [64], Campylobacter coli RM2228, C. upsaliensis RM3195 and C. lari RM2100 [65]. In contrast, the genome sequence of C. jejuni RM1221 contains all genes encoding the TCA [64,65].

2.2. Amino acid metabolism.

Since A. thereius LMG24486T is unable to use carbohydrates as an energy source, alternative metabolic pathways are needed. Amino acids and their degradation products seems to be good candidates since they are present in the habitat of arcobacters, including the animal and human gut. The genome sequence of A. thereius LMG24486T harbours genes encoding different amino acid transporters, such as symporters specific for glutamate and aspartate/sodium (AA347_00247), an arginine/ornithine antiporter (AA347_00156; AA347_00225; AA347_00537; AA347_01038), and a methionine ABC transporter (AA347_00336). This methionine transporter is absent in A. butzleri RM4018 genome, and it could be specific for A. thereius. Furthermore, nine genes were identified encoding several ABC transporter ATP-binding proteins (AA347_00335; AA347_00462; AA347_00463; AA347_01165; AA347_01313; AA347_01549; AA347_01654; AA347_01671) that could not be further specified based on genome annotation. Nineteen transporters that are present in the genome of A. butzleri RM4018 were absent in all nine A. thereius strains sequenced. These comprised 17 ABC transporter ATP-binding proteins, one sodium:alanine symporter and a sodium:hydrogen antiporter. These additional transporters might allow A. butzleri to use more extracellular substrates compared to A. thereius. For A. thereius LMG24486T, four genes encoding aminopeptidases were retrieved, encompassing a leucine aminopeptidase (pepA; AA347_01678), a methionine aminopeptidase (map, AA347_01831), one generic aminopeptidase (AA347_00091) and an oligopeptidase A (prlC, AA347_01499) with a broad specificity.

Arcobacter thereius LMG24486T has a limited capacity to catabolize amino acids, as only genes involved in aspartate, L-glutamate, and serine catabolism were found in its genome (Fig 2).

thumbnail
Fig 2. The proteolytic metabolism of Arcobacter thereius LMG24486T.

(1) D-3-phosphoglycerate dehydrogenase (AA347_00201); (2) Phosphoserine aminotransferase (AA347_00366); (3) Phosphoserine phosphatase (AA347_00244); (4) L-serine deaminase (AA347_00179); (5) Malate dehydrogenase (AA347_01362); (6) Pyruvate kinase (AA347_00705); (7) Phosphoenolpyruvate carboxykinase (AA347_01523); (8) Pyruvate dehydrogenase E1 component (AA347_01757); (9) Dihydrolipoyl dehydrogenase (AA347_01759); (10) Dihydrolipoyllysine-residue acetyltransferase (AA347_01758); (11) Acetyl-coenzyme A synthetase (AA347_00006); (12) Aldehyde dehydrogenase (AA347_00801); (13) NADP-specific glutamate dehydrogenase (AA347_00755); (14) Glutamine synthetase 1 (AA347_01959); (15) glucosamine 6-phosphate synthase (AA347_00043); (16) Carbamoyl-phosphate synthase small chain, Carbamoyl-phosphate synthase arginine-specific large chain (AA347_00362; AA347_01184); (17) Amino-acid acetyltransferase (AA347_01854); (18) Acetylglutamate kinase (AA347_00182); (19) N-acetyl-gamma-glutamyl-phosphate reductase (AA347_01878); (20) Acetylornithine aminotransferase (AA347_00770); (21) Arginine biosynthesis (AA347_01036); (22) Glutamate 5-kinase (AA347_00930); (23) Gamma-glutamyl phosphate reductase (AA347_01121); (24) Ornithine aminotransferase (AA347_01752); (25) Pyrroline-5-carboxylate reductase (AA347_01948); (26) Aspartate aminotransferase (AA347_01647); (27) Aspartate ammonia-lyase (AA347_00952); (28) L-asparaginase 2 (AA347_00953); (29) argininosuccinate synthase (AA347_01336); (30) Argininosuccinate lyase (AA347_01731); (31) Biosynthetic arginine decarboxylase (AA347_01873); (32) agmatinase (AA347_00964); (33) spermidine synthetase (AA347_00329); (34) aspartate kinase (AA347_01503); (35) aspartate-semialdehyde dehydrogenase (AA347_01039); (36) homoserine dehydrogenase (AA347_00407); (37) homoserine kinase (AA347_00208); (38) threonine synthase (AA347_00183); (39) Homoserine O-acetyltransferase (AA347_01182); (40) Cystathionine gamma-synthase (AA347_01684).

https://doi.org/10.1371/journal.pone.0180493.g002

A gene encoding aspartate aminotransferase was found in the genome of A. thereius LMG24486T (Fig 2, reaction 26, AA347_01647), which enables this strain to produce oxaloacetate from aspartate. Another important aspartate catabolic pathway involves aspartase (Fig 2, reaction 27. aspA, AA347_00952), that allows the production of fumarate and ammonia. Aspartase is a central enzyme in the amino acid catabolism in C. jejuni, and seems to have an important effect on growth in complex media, as a C. jejuni mutant lacking a functional aspartase fails to utilize aspartate, glutamate, glutamine and proline [66].

Moreover, aspartase may play a role in the formation of fumarate, an alternative electron acceptor during growth, especially in low oxygen conditions. One of the mechanism of oxygen sensing that has been suggested for aspartase involves the regulator of the CmeABC multidrug efflux transporter, CmeR [67]. In A. thereius LMG24486T, the gene encoding for CmeR is not present, although an anaerobic regulatory protein (AA347_00524) and a transcriptional regulator (AA347_00794) are present, which function typically in response to environmental change. Therefore, they may be involved in modulation of gene expression by the presence of oxygen. Other reactions involving aspartate allow the production of homoserine (Fig 2, reactions 34; 35; 36. AA347_01503; AA347_01039; AA347_00407), threonine (Fig 2, reaction 37; 38; 04. AA347_00208; AA347_00183; AA347_00179) and L-cystathionine (Fig 2, reactions 39; 40. AA347_01182; AA347_01684). Through aspartate, A. thereius LMG24486T is able to produce the biogenic amines agmatine, putrescine and spermidine (Fig 2, reactions 29; 30; 31; 32; 33. AA347_01336; AA347_01731; AA347_01873; AA347_00964; AA347_00329). Their presence in food is related to the activity of decarboxylase by the microorganism and their occurrence can be detect, most of the time, in meat and meat products [68]. Biogenic amines have been detected in other Gram-negative bacteria, but not in Arcobacter yet, and this can represent and essential aspect in prevention in food industry.

Glutamate can be metabolized by A. thereius LMG24486T by a NADP-dependent glutamate dehydrogenase (Fig 2, reaction 13; AA347_00755) and a glutamate synthase (Fig 2, reaction 14; AA347_01959). Further, ornithine can be formed out of glutamate by an amino acid acetyltransferase (Fig 2, reaction 17; AA347_01854), and L-proline can be formed by a glutamate-5-kinase (Fig 2, reaction 22; AA347_00930). Ornithine can be transported through an ornithine/arginine antiporter.

Serine can be deaminated through a serine deaminase (Fig 2, reaction 4 AA347_00179), a pyridoxal 5-phosphate (PLP)-dependent enzyme that converts serine to pyruvate, thereby generating ammonia. Further, serine can be converted in tryptophan or glycine by a tryptophan synthase (AA347_01111, AA347_01680) and a serine hydroxymethyltransferase (AA347_01745).

2.3. Respiration.

Arcobacter thereius has previously been described as microaerophilic bacteria with an optimal temperature range for growth of 21 to 30°C. Growth at 37°C in air or in anaerobic conditions was not observed [13]. This seems however to contrast with its natural presence in the pig and duck gut. All 9 A. thereius strains tested were able to grow and were motile at 28°C in aerobic and microaerobic conditions, but growth as well as motility were completely absent at 37°C and 42°C in all atmospheres tested (Table 3). Remarkably, the genome of A. thereius LMG24486T harbours full clusters of genes encoding for aerobic and microaerobic respiration. Concerning oxygen-dependent electron transport, a NADH:quinone oxidoreductase (Complex I; AA347_00727—AA347_00740) is present together with the aldehyde dehydrogenase complex (AA347_00801), allowing oxidation of NADH by complex I and reduction of NADP+ by an aldehyde dehydrogenase.

A cytochrome bc1 complex (AA347_00184; AA347_00185; AA347_00186), a cytochrome c oxidase (AA347_00228; AA347_00230; AA347_00231; AA347_00379), and a complete cluster of genes encoding a F0/F1 ATPase (AA347_00180; AA347_00936—AA347_00942; AA374_01050) are present, enabling production of ATP from ADP or to generate a transmembrane ion gradient. Three genes encoding a cytochrome c peroxidase were found (AA347_00640; AA347_01278; AA347_00869) which may be important in the detoxification of hydrogen peroxide in the periplasm [64].

In A. thereius LMG 24486T, electrons can be fed to the quinone oxidoreductase by the membrane associated FeNi hydrogenase encoded by hydABCD (AA347_00601—AA347_00604), a malate oxidoreductase (AA347_00064) and a variety of other primary substrate dehydrogenases, for example glutamate dehydrogenase, isocitrate dehydrogenase or pyruvate dehydrogenase. The groups of genes hyaABCD and hupLS encoding for another membrane associate FeNi hydrogenase and for an uptake hydrogenase respectively, result as unique gene for A. butzleri RM4018; although the gene cluster hypABCDEF (AA347_00606 –AA347_00615) is present in both of the species.

In contrast to the results of the growth tests in anaerobic condition, both in the present study (Table 3), as well as in previous ones [1114], the presence of genes encoding fumarate reductase FrdABC (AA347_00741—AA347_00742) together with the additional presence of the nap operon (napDLFHG; AA347_00782-AA347_00787), encoding for proteins involved in nitrate reduction, provides the possibility for anaerobic growth using fumarate and nitrate as electron acceptors instead of oxygen [64]. The ability of A. thereius to reduce nitrate has already been reported by Houf et al. [13] Analysis revealed also the presence of the arsenic resistance operon ars (arsCBR; AA347_01174; AA347_00273; AA347_00274). The system is regulated by the repressor protein ArsR, while ArsB is the determinant of the membrane efflux protein that confers resistance by pumping arsenic from the cell, and ArsC is an arsenate reductase. The ars operon has already been described as part of the E. coli chromosome [69] and on a plasmid in Staphylococcus aureus [70]. The oxyanions of arsenic can be used in anaerobic respiration as terminal electron acceptors, and their oxidation can be coupled with oxidation of other organic substrates (pyruvate, acetate) or hydrogen [71], providing energy for active growth and metabolic activity.

3. Clinical relevant features within Arcobacter thereius LMG24486T

3.1. Lipooligosaccharides.

Lipopolysaccharide (LPS) and lipooligosaccharide (LOS) are important constituents of the outer membrane of bacteria and they can have a role as virulence factor or in antibiotic resistance mechanism [72]. Campylobacter jejuni has been described as only able to express LOS structure and no LPS [73]. The same capacity was retrieved in the genome sequence of A. thereius LMG24486T where a lipooligosaccharide (LOS) biosynthesis gene cluster (AA347_01016 –AA347_01027) was found. The structure of this gene cluster resembles the organization of A. butzleri RM4018 and is present and conserved in all A. thereius strains sequenced in the present study. The genes waaC and waaF (AA347_01016; AA347_01027) encode a heptosyltransferase I, which is responsible for the linkage between the L-glycero-D-manno-heptose (HEP) and 3-deoxy-D-manno-octulsonic (KDO) and, for a heptosyltransferase II which catalyses the transfer of a second HEP to HEP I [74,75]. It has been shown in C. jejuni that a mutation in waaF increases the susceptibility to different hydrophobic antibiotics, such as novobiocin, and a mutation in waaC results in an incomplete LOS [72,76,77]; also their position in the different genomes of A. thereius stains sequenced in this study, resemble the organization of C. jejuni [65]. The kps genes involved in the capsular production are not present in A. thereius genomes, as well as, genes encoding for the O-antigen.

3.2. Flagellum.

Arcobacter thereius possesses a polar flagellum [13] and all genes involved in flagella biosynthesis (flhA, flhB, flhF, flgL, flgC, flgE, flgB, flgG, flgK, flgH, flgI, fliI, fliR, fliK, fliE, fliN, fliG, fliF, fliS, fliD, fliQ, fliM, flaB) were found in A. thereius LMG24486T, which were organized into three gene clusters (AA347_00097—AA347_00126; AA347_00256—AA347_00267; AA347_00588—AA347_00589). The same gene organization is also found in the other A. thereius sequenced, suggesting that this gene region is conserved within this species.

Arcobacter thereius LMG24486T contains a complete pseudaminic acid biosynthesis pathway (pseBCFGHI; AA347_00580 –AA347_00585), which has been reported to constitute a major virulence factor in C. jejuni [78]. Interestingly, this gene region contained also an acetylase (AA347_00585), as well as a gene with high similarity to C. jejuni pseH (AA347_00584), which may be involved in flagellar glycosylation. This pathway was only partially present in A. butzleri RM4018, as both the pseG (encoding for an hydrolase with important function during the fourth step of the pseudoaminic acid biosynthesis pathway [78]) and pseH genes were missing, suggesting that this post-translationally modification system through O-linked glycosylation is not functional in A. butzleri [79].

3.3. Chemotaxis.

Interaction between bacteria and the environment is a fundamental aspect for the survival and adaptation of the microorganism. In fact, bacteria have a lot of different mechanisms which enable them to respond to environmental changes. A. thereius LMG22486T harbours 18 methyl-accepting chemotaxis proteins, nine two-component response regulators, and eight sensor histidine kinases. During the comparison, in the genome of A. butzleri RM4018 36 sensor histidine kinase, 41 response regulators have been found as unique features of A. butzleri RM4018 [17]. This indicates that A. thereius might be less reactive to the surrounding changes, although a full set of genes encoding for chemotaxis proteins were present, including CheARDB (AA347_01490 –AA347_01494), CheV (AA347_01730), CheY (AA347_01490). These proteins play a role in signal transmission from the receptor to the flagellum to regulate or change the bacterial movements.

3.4. Antibiotic resistance.

An in silico search for genes putatively involved in antibiotic resistance revealed the presence of several genes within the A. thereius LMG24486T genome that could contribute to antibiotic resistance. For instance, a gene cluster encoding a CmeABC efflux pump was found (AA347_00510 –AA347_00512), which is associated with efflux of various antibiotics in C. jejuni [80,81]. This gene cluster was found in all A. thereius strains sequenced (S2 Table). The β-lactam resistance mechanisms of A. thereius LMG24486T are unclear as no genes encoding for β-lactamases were found, which have been reported before in A. butzleri [17]. However, an lrgAB operon was found (AA347_00278 –AA347_00279), described to be involved in penicillin tolerance in Staphyloccocus aureus [82]. Evidence was found for the presence of resistance mechanisms towards quinolones, such as a Thr85Ser mutation in the DNA gyrase subunit A. This mutation has been reported to be responsible for resistance of Arcobacter cibarius towards ciprofloxacin [83]. The Thr85Ser mutation in GyrA was also present in strains DU22 (AAX29_01876), 440 (AAX25_00506), and 452 (AAX26_01016), whereas a Thr85Ile mutation was found in strain 11743–4 (AAX28_1873). The latter mutation was considered responsible of conferring a high ciprofloxacin resistance in C. coli, A. butzleri and A. cryaerophilus [83,84]. Arcobacter thereius LMG24486T possesses no specific mechanisms towards macrolide resistance, such as the presence of a major outer membrane porin (MOMP) or macrolide resistance-associated mutations. The latter include specific point mutations at positions 2074 and 2075 in the 23S rRNA gene, which are involved in erythromycin resistance in Campylobacter species [81,85]. However, MOMP was present in strains A. thereius 213, A. thereius 216, A. thereius 440, A. thereius 11743–4, and A. thereius DU22 (S4 Table), indicating a potential resistance towards macrolides, though not phenotypical expressed.

Further, evidence for chloramphenicol resistance was found, as a chloramphenicol O-acetyltransferase was present (cat; AA347_01813), which was also observed in the A. thereius strains 440 and DU22.

Several other antibiotic resistant related genes were found in the other A. thereius strains sequenced, that were not present in the type strain. For example, a gene encoding adenylyltransferase (aadA25; AAD) was found in the genomes of A. thereius strains 213 (AAW29_01543), 216 (AAW30_00055), DU19 (AAX30_01180), and 11743–4 (AAX28_00760), which is involved in resistance towards streptomycin/spectinomycin in Pasteurella multocida [86]. Furthermore, a gene encoding an energy-depended membrane-associated protein TetA (AAX28_02020) was found in the genome of A. thereius 11743–4. This efflux pump, usually present in Gram-negative bacteria, confers resistance to tetracycline by exporting it out of the cell thereby reducing the intracellular concentration [87]. This is the first report of this protein in the family Campylobacteriaceae.

The susceptibility results obtained by gradient strip diffusion for the nine A. thereius strains are shown in Table 3. Of all strains, 80–100% were susceptible to erythromycin, ciprofloxacin, tetracycline and gentamicin. These results resemble those obtained by the genome analysis. In fact none of the strains harbour the point mutation in the 23 rRNA gene causing a specific resistance to erythromycin and, for tetracycline resistance, only the strain A. thereius 11743–4 carried the TetA protein. Interesting, the ciprofloxacin susceptibility results showed that only A. thereius 11743–4, carrying the DNA GyraseA point mutation Thr85Ile, is resistant against ciprofloxacin while, A. thereius DU22, 440 and 452, that carry the Thr85Ser point mutation remain sensitive towards ciprofloxacin. Streptomycin and chloramphenicol show MICs around breakpoint value, while 80–100% of the strains are resistant to spectinomycin. A. thereius 11743–4 presented with a divergent resistance pattern in comparison to the other strains, pointing towards acquired multiresistance, which has to be further investigated.

3.5. Virulence associated genes.

Several virulence factors were detected in the genome of A. thereius LMG24486T, among which the fibronectin binding protein Cj1349 (AA347_00304), the invasion protein CiaB (AA347_00973), the virulence factor MviN (AA347_01129), the phospholipase PldA (AA347_01541), the hemolysin TlyA (AA347_01277), and the enterobactin receptor IrgA (AA347_01908 and AA347_01909) [17,88]. The presence of two copies of the latter gene is in contrast with the genome of A. butzleri RM4018, which encodes a putative siderophore esterase IroE adjacent to a single copy of the IrgA gene [17]. Next to the absence of IroE, other virulence factors reported in A. butzleri, A. cryaerophilus and A. skirrowii such as HecAB [88] and CadF [16,88], were not found in A. thereius LMG24486T. The virulence factors present in the other eight A. thereius strains sequenced in this study were identical to the type strain (S3 Table), except for A. thereius DU22, which possessed the hecAB virulence genes (AAX29_00055 –AAX29_00056). The putative virulence genes cj1349, ciaB, mviN, pldA, tlyA and irgA were conserved within the species A. thereius (Fig 3). Although the presence of these virulence factor were reported before in Arcobacter species [17,88], this is the first report of their occurrence in A. thereius. Indeed, traditional PCR detection methods fail when applied on A. thereius strains [16], which might be related to differences in DNA composition due to evolutionary modifications of the virulence genes.

thumbnail
Fig 3. Phylogenetic analysis of the six putative virulence proteins of Arcobacter thereius.

A Neighbor-Joining phylogenetic tree representing the six putative virulence proteins found in A. thereius. Arcobacter, Campylobacter and Helicobacter species are included as outgroup. The scale bar represent substitution per site.

https://doi.org/10.1371/journal.pone.0180493.g003

4. Analysis of the genome variability within A. thereius

Comparison of the genomes from all A. thereius strains sequenced in the present study, displays a similar genomic architecture, although a few structural rearrangements were found (S4 Table, S1 Fig). We examined whether the genome content of A. thereius strains isolated from the cloaca of ducks were distinct from those isolated from pig faeces or the tissue of aborted piglets. A phylogenetic tree based on the alignment of 239 single-copy orthologs was constructed (Fig 4). The interspersed positions of the pig faeces, piglet tissue and duck cloaca isolates on the species tree, together with the short length of the branches, indicate that a single, homogenous population of A. thereius is maintained in the different animal populations. This hypothesis is further supported by the fact that there were only minor differences in the distribution of genes into clusters of orthologous groups of proteins (COG) functional categories across genomes (S2 Fig). This result indicate a limited functional variability among different A. thereius strains. For example, only two genes without a clear link to disease or habitat were present in both strains isolated from cloaca from ducks but absent in all other strains, encoding a serine hydroxymethyltransferase (DU19: AAX30_00966 and DU22: AAX29_01527) and a UDP-N-acetyl-D-glucosamine 6-dehydrogenase (DU22: AAX29_01729 and DU19: AAX30_01828). The seven strains isolated from pigs contained two genes which were absent in all strains originating from duck cloaca, both encoding hypothetical proteins which are adjacent to each other. The genome of A.thereius LMG24486T harboured 394 accessory genes (S5 Table), i.e., genes without a homolog in the other A. thereius genomes. Most of these genes were organised among 18 islands. These strain-specific genes included 26 genes coding for phage integrases, tyrosine recombinases, prophage integrases and CRISPR- associated endonucleases. Further, two complete genes clusters coding for the FeNi hydrogenase and for the HypABCDEF complex (AA347_00599 –AA347_00602; AA347_00604 –AA347_00613; discussed above) were detected in the type strain. The occurrence of two unique additional hydrogenase might enable this strain to obtain energy in microaerobic environments. Because they are not found in related, pathogenic strains, these genes, unique to A. thereius LMG24486T, might confer an advantage in some aspects of A. thereius’ behaviour, but it is unlikely that they play a role in pathogenicity.

thumbnail
Fig 4. Phylogenetic tree of Arcobacter genus.

A maximum likelihood phylogenetic tree constructed using 239 single-copy orthologs and representing the A. thereius strains sequenced in this study and all other Arcobacter genomes available. C. jejuni was not included in the phylogenetic tree.“White dots” = A. thereius isolated from piglet aborted foetus; “Black dots” = A. thereius isolated from pig faeces; “Yellow dots” = A. thereius isolated from cloaca content duck.

https://doi.org/10.1371/journal.pone.0180493.g004

Though in the initial characterization of the species, DNA-DNA hybridization of strains LMG24486T and A. thereius LMG24487 exhibited a mean DNA-DNA relatedness value of 79% [13], in silico DDH as well as Average Nucleotide Identity (ANI) analysis suggest a separation into two species, each inclosing one of the strains. The in silico DDH values and the ANI data for the nine A. thereius strains are shown in Table 5 and Table 6. Combining these findings with the results obtained by the phylogenetic analysis, there is indeed a consistent split into two closely related subclusters within the set of A. thereius strains examined. However, strains of both clusters have been included in the previous characterization of the species [13], with over 60 phenotypic characteristics determined. No differential diagnostic test (a strict requirement for the description of a new species) could be identified. Furthermore, there is currently also no geographic, biological niche, genome content, nor clinical relevance indicating the need for the relocation into a new species. Because the taxonomic criteria for the creation of a new species are not fulfilled, the strains have to be considered as closely related genotypes of the same species. However, recently, a study based on ANI and isDDH data analysis on the strains included in the present study proposed the classification into a new species, Arcobacter porcinus, including the majority of strains previously identified as A. thereius [89]. Further studies seem necessary in order to provide a clear overview in the taxonomy of the Arcobacter genus.

thumbnail
Table 5. Estimation of in silico DDH for the nine A. thereius strains.

The confidential interval is shown in square brackets. A. butzleri RM4018 and A. cibarius LMG21996 were included as references. The symbol “*” is added when the same genomes are compared.

https://doi.org/10.1371/journal.pone.0180493.t005

thumbnail
Table 6. Estimation of the Average Nucleotide Identity (ANI) for the nine A. thereius strains with A. butzleri RM4018 and A. cibarius LMG21996 were included as outgroup.

ANI results are based on BLAST+ analysis and expressed as [aligned nucleotides] [%].The symbol “*” is added when the same genomes are compared.

https://doi.org/10.1371/journal.pone.0180493.t006

MLST analysis on the nine A. thereius genomes revealed the consistent presence of the loci aspA, atpA, glnA, gltA, pmg and tkt. In contrast glyA locus was lacking in A. thereius 213, 216, 440, 452 and 11743–4, while two copies were present in the genomes of A. thereius DU19, DU22, LMG24487 and LMG24486T (S3 Fig). The MLST analysis of the concatenated genes confirmed again the presence of two closely related genotypes, as also found by DDH and ANI analysis.

Fig 5 shows a plot of the core and the pan genome of the nine A. thereius strains included in the present study. Based on EDGAR analysis using one complete genome (A. thereius LMG24486T) and eight draft genome sequences, the A. thereius core genome showed to have a stable size with the largest proportion of the genes part of the core-genome, while the pan-genome is predicted to be open.

thumbnail
Fig 5. Core and pan genome development plot.

Prediction of the development of the core and pan genome for the species Arcobacter thereius, based on the nine genome sequences determined in this study. Starting with the first genome, a sequence of core and pan genome sizes is calculated by iteratively adding one genome at time to the comparison and observed value are depicted in full circle [51].

https://doi.org/10.1371/journal.pone.0180493.g005

Of the eight A. thereius genomes sequenced, A. thereius 440 recovered from pig faeces and LMG24487 isolated from piglet aborted foetus, carried the gene encoding for the zonula occludens toxin (ZOT) (AAX25_01078; AAX27_00411). The ZOT is known to work on the intracellular tight junction and allows pathogenic bacteria, like Vibrio cholera or Neisseria meningitidis, to increase tissue permeability. Recently, this gene has been detected also in the genome of Campylobacter concisus 13826 [90], a non-jejuni Campylobacter found in samples of patients with gastrointestinal disorders and now suggested as a potential pathogen [91]. According to Kaakoush et al. [90], the presence of the ZOT gene can have an important role in the pathogenesis of C. concisus, allowing the bacteria to attach and invade host cells through a paracellular mechanism. It was not detect in other Arcobacter species, and further research is needed, taking into account current genome knowledge, to elucidate which genes participate in cell adhesion and invasion.

A gene cluster, coding for type IV secretion system (virB4, virB6, virB9, virB10, virB11; AAW29_01609; AAW29_01614; AAW29_01615; AAW29_01616; AAW29_001618) was present in the genome of A. thereius 213 as a singleton. The type IV secretion contributes in a lot of transport like, exchange of genetic material between bacteria, movement of plasmid and injection of virulence factor in the host cell. The subgroup VirB is specific for the T-DNA transfer and is built with 11 different proteins (VirB1-VirB11), but a conserved core of five proteins is always present (VirB4; VirB7; VirB9; VirB10; VirB11) [92]. This secretion system has been described as part of the large plasmid (AC1119) found in A. butzleri and it could play an important role in gene transfer within Arcobacter [60].

5. Genome comparison of A. thereius LMG24486T and A. butzleri RM4018

Arcobacter butzleri is the most representative human related Arcobacter and, after the announcement of its genome, its role as potential human pathogen has been confirmed in several reported cases. The genomes of A. thereius LMG24486T and A. butzleri RM4018 show little synteny (Fig 6), which supports the view that A. thereius appeared to have unique characteristics, different from those previously reported for A. butzleri [17]. Indeed, the genome sequences of A. thereius LMG24486T and A. butzleri RM4018 shared 1474 (ortholog) genes, A. thereius LMG24486T contained 393 singletons (among which 219 encode hypothetical proteins), and A. butzleri RM4018 contained 688 singletons (among which 344 encode hypothetical proteins). Besides the differences and similarities already mentioned above, among the singletons of A. butzleri RM4018 a complete cluster of genes encoding for sulphur uptake and assimilation was present [17]. However, as the sox gene cluster, necessary for the sulphur oxidation, were also present in A. thereius LMG24486T (soxA,Z,Y; AA347_00874—AA347_00876), this strain could be able to oxidise sulphite to sulphate.

thumbnail
Fig 6. Comparative genome analysis of Arcobacter species.

Synteny plot depicting the finished genome sequences of A. thereius LMG24486T and A. butzleri RM4018. Each dot represents a reciprocal best BLAST hit. For the x-axis, the genomic position is depicted, for the y-axis, 100% corresponds with 2,341,251 bp.

https://doi.org/10.1371/journal.pone.0180493.g006

Further, a thiamine biosynthetic gene cluster was found (thiDEHGFS, AA347_00322 –AA347_00327), enabling thiamine autotrophy in A. thereius LMG24486T. Production of thiamine is an important mechanism because it is an essential cofactor of different metabolic enzymes and it was not present in other Arcobacter [17,93] suggesting that this feature is unique for A. thereius. However, it remains to be determined if this is going to be maintained when genomes of other Arcobacter species will be available.

Another interesting difference present as a singleton in A. thereius LMG24486Tgenome are the type I, II and III restriction endonuclease. These enzymes protect bacteria against invasion of foreign DNA, and differ in their way of recognition and cleavage [94]. Type I restriction enzymes consist in three subunit, HsdS (AA347_01253), HsdM (AA347_0152) and HsdR (AA347_01255; AA347_00856; AA347_00561) and, they are responsible for modification, restriction and sequence recognition [94]. For the type III restriction enzyme only the subunit Res (AA347_01092) has been found in the genome of A. thereius LMG24486T meaning that this enzyme is not working, although two type II enzymes were harboured (AA347_1323; AA347_01567).

In contrast to A. butzleri RM4018, the genome sequence of LMG24486T contains the genes of the ectoine biosynthesis pathway (ectABC, AA347_00353 –AA347_00355). Ectoine is a compatible solute, important for its function as osmoprotector; it helps microorganism to survive extreme osmotic stress and temperature stress. As A. thereius LMG 24486T contains also an aspartate kinase (AA347_01497), and an aspartate semialdehyde dehydrogenase (AA347_01037), ectoine may be produced out of aspartate. Further, all six genes involved in urea degradation (ureABCDEFD) in A. butzleri RM4018 [17] were not present in either of the A. thereius strains.

Conclusion

Whole genome analysis of nine A. thereius strains, including the type strain, confirmed that they do not ferment nor oxidize carbohydrates, and energy provision depends rather on a limited group of amino acids. The species is regarded as difficult to grow under laboratory conditions, and growth temperature and atmosphere requirements are not exactly similar to those previous described for A. butzleri. However, we could not identify the possible genetic determinants for these differences in the genomes of A. thereius.

Arcobacter thereius is predominantly present in the pig intestinal tract, but, due to the apparent lack of virulence factors previously based on specific PCR detection, has not been considered an important species from the food safety perspective so far. However, the species was recently isolated from the stool of human enteritis patients, and the present study reveals the presence of six out of eight virulence associated genes previously reported in A. butzleri. Further research will elucidate the role of these genes in the potential pathogenicity of A. thereius.

Comparative genome and phylogenetic analysis of the nine A. thereius strains revealed the delineation of two closely related subgroups, for which, besides the group including the A. thereius reference strain, a new species, A. porcinus has recently be proposed [89]. However, at present, no clear phenotypic or genomic differences are yet identified to support this creation of a new species. Comparative genome analysis with the other human associated species, A. butzleri, reveals a large correlation, though also unique features are present. The occurrence of virulence associated genes, genes for antibiotic resistance as well as other pathogenic related relevant features, justifies a further exploration of the reservoir, transmission routes, consolidated in a risk assessment in both human and veterinary medicine.

Supporting information

S1 Fig. Alignment with ProgressiveMauve (default parameters) of the nine A. thereius strains sequenced.

https://doi.org/10.1371/journal.pone.0180493.s001

(TIF)

S2 Fig. Distribution of genes in COG functional categories.

The number of genes in each categories was compared among the nine A. thereius strains sequenced. No significant differences in the distribution of genes belonging to a functional category among the different strains has been found (see Materials and Method). (C) = Energy production and conversion; (D) = Cell cycle control, cell division, chromosome partitioning; (E) = Amino acid transport and metabolism; (F) = Nucleotide transport and metabolism; (G) = Carbohydrate transport and metabolism; (H) = Coenzyme transport and metabolism; (I) = Lipid transport and metabolism; (J) = Translation, ribosomal structure and biogenesis; (K) = Transcription; (L) = Replication, recombination and repair; (M) = Cell/wall/membrane/envelope biogenesis; (N) = Cell motility; (O) = Posttranslational modification, protein turnover, chaperones; (P) = Inorganic ion transport and metabolism; (Q) = Secondary metabolites biosynthesis, transport and catabolism; (R) = General function protection only; (S) = Function unknown; (T) = Signal transduction mechanisms; (U) = Intracellular trafficking, secretion, and vascular transport; (V) = Defence mechanisms.

https://doi.org/10.1371/journal.pone.0180493.s002

(PDF)

S3 Fig. Phylogenetic analysis of the six genes used in the MLST typing scheme for Arcobacter spp.

A neighbour-joining phylogenetic tree representing the six genes involved in the MLST method. Arcobacter and Campylobacter species are included as outgroup. For Campylobacter jejuni ATCC43439 the gene uncA (atpA) has been used.

https://doi.org/10.1371/journal.pone.0180493.s003

(TIF)

S1 Table. CRISPR sequence of A. thereius LMG24486T genome.

https://doi.org/10.1371/journal.pone.0180493.s004

(XLSX)

S2 Table. Position of the genes involved in the antibiotic resistance in the different A. thereius strains.

https://doi.org/10.1371/journal.pone.0180493.s005

(XLSX)

S3 Table. Position of the virulence genes in the different A. thereius strains.

https://doi.org/10.1371/journal.pone.0180493.s006

(XLSX)

S4 Table. General features of the nine A. thereius strains sequenced in this study.

https://doi.org/10.1371/journal.pone.0180493.s007

(XLSX)

S5 Table. Accessory genes of A. thereius LMG24486T organise within 18 clusters.

https://doi.org/10.1371/journal.pone.0180493.s008

(XLSX)

Acknowledgments

The authors thank Dr. Jochen Blom and Prof. Dr. Alexander Goesmann (Institute of Bioinformatics and Systems Biology, Justus-Liebig-University, Giessen, Germany) for performing the EDGAR computational analyses and Dr. Yannick Gansemans for the kind assistance in the genome assembly.

References

  1. 1. Vandamme P, Falsen E, Rossau R, Hoste B, Segers P,Tytgat R, et al. (1991) Emendation of Generic Descriptions and Proposal of Arcobacter gen. nov. Int J Syst Evol Microbiol 41: 88–103.
  2. 2. Levican A, Rubio-Arcos S, Martinez-Murcia A, Collado L, Figueras MJ (2015) Arcobacter ebronensis sp. nov. and Arcobacter aquimarinus sp. nov., two new species isolated from marine environment. Syst Appl Microbiol 38: 30–35. pmid:25497285
  3. 3. Whiteduck-Léveillée K, Whiteduck-Léveillée J, Cloutier M, Tambong JT, Xu R, Toop E, et al. (2016) Identification, characterization and description of Arcobacter faecis sp. nov., isolated from a human waste septic tank. Syst Appl Microbiol 39: 93–99. pmid:26723853
  4. 4. Whiteduck-Léveillée K, Whiteduck-Léveillée J, Cloutier M, Tambong JT, Xu R, Toop E et al. (2015) Arcobacter lanthieri sp. Nov., isolated from pig and dairy cattle manure. Int J Syst Evol Microbiol 65: 2709–2716. pmid:25977280
  5. 5. Zhang Z, Yu C, Wang X, Yu S, Zhang XH (2016) Arcobacter pacificus sp. nov., isolated from seawater of the south pacific Gyre. Int J Syst Evol Microbiol 66: 542–547. pmid:26556763
  6. 6. Ho HTK, Lipman LJ a, Gaastra W (2006) Arcobacter, what is known and unknown about a potential foodborne zoonotic agent! Vet Microbiol 115: 1–13. pmid:16621345
  7. 7. Collado L, Figueras MJ (2011) Taxonomy, epidemiology, and clinical relevance of the genus Arcobacter. Clin Microbiol Rev 24: 174–192. pmid:21233511
  8. 8. Vandenberg O, Dediste A, Houf K, Ibekwem S, Souayah H, Cadranel S, et al. (2004) Arcobacter species in humans. Emerg Infect Dis 10: 1863–1867. pmid:15504280
  9. 9. Van den Abeele A, Vogelaers D, Van Hende J, Houf K (2014) Prevalence of Arcobacter Species among Humans, Belgium,2008–2013. Emerg Infect Dis 20: 1731–1734. pmid:25271569
  10. 10. Wybo I, Breynaert J, Lauwers S, Lindenburg F, Houf K (2004) Isolation of Arcobacter skirrowii from a Patient with Chronic Diarrhea. J Clin Microbiol 42: 1851–1852. pmid:15071069
  11. 11. On SLW, Jensen TK, Bille-Hansen V, Jorsal SE, Vandamme P (2002) Prevalence and diversity of Arcobacter spp. isolated from the internal organs of spontaneous porcine abortions in Denmark. Vet Microbiol 85: 159–167. pmid:11844622
  12. 12. On SLW, Harrington CS, Atabay HI (2003) Differentiation of Arcobacter species by numerical analysis of AFLP profiles and description of a novel Arcobacter from pig abortions and turkey faeces. J Appl Microbiol 95: 1096–1105. pmid:14633039
  13. 13. Houf K, On SLW, Coenye T, Debruyne L, De Smet S, Vandamme P. (2009) Arcobacter thereius sp. nov., isolated from pigs and ducks. Int J Syst Evol Microbiol 59: 2599–2604. pmid:19622651
  14. 14. De Smet S, De Zutter L, Debruyne L, Vangroenweghe F, Vandamme P, Houf K. (2011) Arcobacter population dynamics in pigs on farrow-to-finish farms. Appl Environ Microbiol 77: 1732–1738. pmid:21216902
  15. 15. Douidah L, De Zutter L, Baré J, Houf K (2014) Towards a typing strategy for Arcobacter species isolated from humans and animals and assessment of the in vitro genomic stability. Foodborne Pathog Dis 11: 272–280. pmid:24400986
  16. 16. Levican A, Alkeskas A, Günter C, Forsythe SJ, Figueras MJ (2013) Adherence to and invasion of human intestinal cells by Arcobacter species and their virulence genotypes. Appl Environ Microbiol 79: 4951–4957. pmid:23770897
  17. 17. Miller WG, Parker CT, Rubenfield M, Mendz GL, Wosten MMSM, Ussery DW, et al. (2007) The Complete Genome Sequence and Analysis of the Epsilonproteobacterium Arcobacter butzleri. PLoS One 2: e1358. pmid:18159241
  18. 18. Boetzer M, Henkel C V, Jansen HJ, Butler D, Pirovano W (2011) Scaffolding pre-assembled contigs using SSPACE. Bioinformatics 27: 578–579. pmid:21149342
  19. 19. Boetzer M, Pirovano W (2012) Toward almost closed genomes with GapFiller. Genome Biol 13: R56. pmid:22731987
  20. 20. Eid J, Fehr A, Gray J, Luong K, Lyle J, Otto G, et al. (2009) Real-Time DNA sequencing from single polymerase molecules. Science (80-) 323: 133–138.
  21. 21. Boetzer M, Pirovano W (2014) SSPACE-LongRead: scaffolding bacterial draft genomes using long read sequence information. BMC Bioinformatics 15: 211. pmid:24950923
  22. 22. Seemann T (2014) Prokka: Rapid prokaryotic genome annotation. Bioinformatics 30: 2068–2069. pmid:24642063
  23. 23. Hyatt D, Chen G-L, Locascio PF, Land ML, Larimer FW, Hauser LJ. (2010) Prodigal: prokaryotic gene recognition and translation initiation site identification. BMC Bioinformatics 11: 119. pmid:20211023
  24. 24. Laslett D, Canback B (2004) ARAGORN, a program to detect tRNA genes and tmRNA genes in nucleotide sequences. Nucleic Acids Res 32: 11–16. pmid:14704338
  25. 25. Nawrocki EP, Eddy SR (2013) Infernal 1.1: 100-fold faster RNA homology searches. Bioinformatics 29: 2933–2935. pmid:24008419
  26. 26. Petersen TN, Brunak S, Von Heijne G, Nielsen H (2011) SignalP 4.0: discriminating signal peptides from transmembrane regions. Nat Meth 8: 785–786.
  27. 27. Magrane M, Consortium UP (2011) UniProt Knowledgebase: A hub of integrated protein data. Database 2011: 1–13. pmid:21447597
  28. 28. Finn RD, Bateman A, Clements J, Coggill P, Eberhardt RY, Eddy SR, et al. (2014) Pfam: the protein families database. Nucleic Acids Res 42: D222–D230. pmid:24288371
  29. 29. Haft DH, Jeremy D, White S, White O (2003) The TIGRFAMs database of protein families. Nucleic Acids Res 31: 371–373. pmid:12520025
  30. 30. Aziz RK, Bartels D, Best AA, DeJongh M, Disz T, Edwards RA, et al. (2008) The RAST Server: rapid annotations using subsystems technology. BMC Genomics 9: 75. pmid:18261238
  31. 31. Overbeek R, Olson R, Pusch GD, Olsen GJ, Davis JJ, Disz T, et al. (2014) The SEED and the Rapid Annotation of microbial genomes using Subsystems Technology (RAST). Nucleic Acids Res 42: D206–14. pmid:24293654
  32. 32. Kanehisa M, Goto S (1999) KEGG: Kyoto encyclopedia of genes and genomes. Nucleic Acids Res 27: 29–34. pmid:9847135
  33. 33. Carattoli A, Zankari E, Garciá-Fernández A, Larsen MV, Lund O, Villa L, et al. (2014) PlasmidFinder and pMLST: in silico detection and typing of plasmid. Antimicrob Agents Chemother 58: 3895–3903. pmid:24777092
  34. 34. Wattam AR, Abraham D, Dalay O, Disz TL, Driscoll T, Gabbard JL, et al. (2014) PATRIC, the bacterial bioinformatics database and analysis resource. Nucleic Acids Res 42: D581–D591. pmid:24225323
  35. 35. Zhou Y, Liang Y, Lynch KH, Dennis JJ, Wishart DS (2011) PHAST: A Fast Phage Search Tool. Nucleic Acids Res 39: 1–6.
  36. 36. Gupta SK, Padmanabhan BR, Diene SM, Lopez-Rojas R, Kempf M, Landraud L, et al. (2014) ARG-annot, a new bioinformatic tool to discover antibiotic resistance genes in bacterial genomes. Antimicrob Agents Chemother 58: 212–220. pmid:24145532
  37. 37. McArthur AG, Waglechner N, Nizam F, Yan A, Azad MA, Baylay AJ, et al. (2013) The comprehensive antibiotic resistance database. Antimicrob Agents Chemother 57: 3348–3357. pmid:23650175
  38. 38. Chen L, Xiong Z, Sun L, Yang J, Jin Q (2012) VFDB 2012 update: toward the genetic diversity and molecular evolution of bacterial virulence factors. Nucleic Acids Res 40: D641–D645. pmid:22067448
  39. 39. Chen L, Yang J, Yu J, Yao Z, Sun L, Shen Y, et al. (2004) VFDB: a reference database for bacterial virulence factors. Nucleic Acids Res 33: D325–D328. pmid:15608208
  40. 40. Yang J, Chen L, Sun L, Yu J, Jin Q (2007) VFDB 2008 release: an enhanced web-based resource for comparative pathogenomics. Nucleic Acids Res 36: D539–D542. pmid:17984080
  41. 41. Grissa I, Vergnaud G, Pourcel C (2008) CRISPRFinder: a web tool to identify clustered regularly interspaced short palindromic repeats. Nucleic Acids Res 36: 52–57. pmid:18442988
  42. 42. Carver T, Thomson N, Bleasby A, Berriman M, Parkhill J (2009) DNAPlotter: Circular and linear interactive genome visualization. Bioinformatics 25: 119–120. pmid:18990721
  43. 43. Gao F, Zhang C- T (2008) Ori-Finder: a web-based system for finding oriCs in unannotated bacterial genomes. BMC Bioinformatics 9: 79. pmid:18237442
  44. 44. Dhillon BK, Chiu TA, Laird MR, Langille MGI, Brinkman FSL (2013) IslandViewer update: Improved genomic island discovery and visualization. Nucleic Acids Res 41: 129–132. pmid:23677610
  45. 45. Van Heel AJ, de Jong A, Montalbán-López M, Kok J, Kuipers OP (2013) BAGEL3: Automated identification of genes encoding bacteriocins and (non-)bactericidal posttranslationally modified peptides. Nucleic Acids Res 41: 448–453. pmid:23677608
  46. 46. Li L, Christian J, Stoeckert J, Roos DS (2003) OrthoMCL: Identification of Ortholog Groups for Eukaryotic Genomes. Genome Res 13: 2178–2189. pmid:12952885
  47. 47. Bruen T (2005) PhiPack: PHI test and other tests of recombination.
  48. 48. Edgar RC (2004) MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acid Res 32: 1792–1797. pmid:15034147
  49. 49. Capella-Gutierrez S, Silla-Martinez JM, Gabaldon T (2009) trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25: 1972–1973. pmid:19505945
  50. 50. Stamatakis A (2014) RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 30: 1312–1313. pmid:24451623
  51. 51. Blom J, Albaum SP, Doppmeier D, Pühler A, Vorhölter F-J, Vorhölter FJ, et al. (2009) EDGAR: a software framework for the comparative analysis of prokaryotic genomes. BMC Bioinformatics 10: 154. pmid:19457249
  52. 52. Tamura K, Stecher G, Peterson D, Filipski A, Kumar S (2013) MEGA6: Molecular evolutionary genetics analysis version 6.0. Mol Biol Evol 30: 2725–2729. pmid:24132122
  53. 53. Richter M, Rosselló-Móra R, Glöckner FO, Peplies J (2015) JSpeciesWS: a web server for prokaryotic species circumscription based on pairwise genome comparison. Bioinformatics 32: btv681-. pmid:26576653
  54. 54. Meier-Kolthoff Jan P, Alexander F Auch H-PK and G M (2013) Genome sequence-based species delimitation with confidence intervals and improved distance functions. BMC Bioinformatics 14.
  55. 55. Miller WG, Wesley IV, On SLW, Houf K, Mégraud F, Wang G, et al. (2009) First multi-locus sequence typing scheme for Arcobacter spp. BMC Microbiol 9: 196. pmid:19751525
  56. 56. Darling AE, Mau B, Perna NT (2010) ProgressiveMauve: Multiple genome alignment with gene gain, loss and rearrangement. PLoS One 5: 1–17. pmid:20593022
  57. 57. Van den Abeele AM, Vogelaers D, Vanlaere E, Houf K (2016) Antimicrobial susceptibility testing of Arcobacter butzleri and Arcobacter cryaerophilus strains isolated from Belgian patients. J Antimicrob Chemother 71: 1241–1244. pmid:26851610
  58. 58. EUCAST (2016) The European Committee on Antimicrobial Susceptibility Testing. Breakpoint tables for interpretation of MICs and zone diameters. Version 6.0, 2016. http://www.eucast.org: 0–77.
  59. 59. Centres for Disease Control and Prevention (CDC) (2012) National Antimicrobial Resistance Monitoring System for Enteric Bacteria (NARMS): Human Isolates Final Report, 2010. 104: 75. 10.1093/infdis/jis901.
  60. 60. Douidah L, De Zutter L, Van Nieuwerburgh F, Deforce D, Ingmer H, Vandenberg O, et al. (2014) Presence and analysis of plasmids in human and animal associated arcobacter species. PLoS One 9: e85487. pmid:24465575
  61. 61. Carnoy C, Roten CA (2009) The dif/Xer recombination systems in proteobacteria. PLoS One 4. pmid:19727445
  62. 62. Parkhill J, Wren BW, Mungall K, Ketley JM, Churcher C, Basham D, et al. (2000) The genome sequence of the food-borne pathogen Campylobacter jejuni reveals hypervariable sequences. Nature 403: 665–668. pmid:10688204
  63. 63. Berg JM, Tymoczko JL, Stryer L (2002) The Citric Acid Cycle Is a Source of Biosynthetic Precursors. Biochemestry. W H Freeman.
  64. 64. Kelly DJ (2001) The physiology and metabolism of Campylobacter jejuni and Helicobacter pylori. Symp Ser Soc Appl Microbiol 90: 16S–24S.
  65. 65. Fouts DE, Mongodin EF, Mandrell RE, Miller WG, Rasko DA, Ravel J, et al. (2005) Major structural differences and novel potential virulence mechanisms from the genomes of multiple Campylobacter species. PLoS Biol 3: e15. pmid:15660156
  66. 66. Guccione E, Del Rocio LKM, Pearson BM, Hitchin E, Mulholland F, Van Diemen PM, et al. (2008) Amino acid-dependent growth of Campylobacter jejuni: Key roles for aspartase (AspA) under microaerobic and oxygen-limited conditions and identification of AspB (Cj0762), essential for growth on glutamate. Mol Microbiol 69: 77–93. pmid:18433445
  67. 67. Guo B, Wang Y, Shi F, Barton YW, Plummer P, Reynolds DL, et al. (2008) CmeR functions as a pleiotropic regulator and is required for optimal colonization of Campylobacter jejuni in vivo. J Bacteriol 190: 1879–1890. pmid:18178742
  68. 68. Buňková L, Buňka F, Klčovská P, Mrkvička V, Doležalová M and Kràčmar S. (2010) Formation of biogenic amines by Gram-negative bacteria isolated from poultry skin. Food Chem 121: 203–206.
  69. 69. Diorio C, Cai J, Marmor J, Shinder R, DuBow MS (1995) An Escherichia coli chromosomal ars operon homolog is functional in arsenic detoxification and is conserved in gram-negative bacteria. J Bacteriol 177: 2050–2056. pmid:7721697
  70. 70. Ji G, Silver S (1992) Reduction of Arsenate to Arsenite by the ArsC Protein of the Arsenic Resistance Operon of Staphylococcus aureus Plasmid pI258. Proc Natl Acad Sci 89: 9474–9478. pmid:1409657
  71. 71. Stolz J, Oremland R (1999) Bacterial respiration of arsenic and selenium. FEMS Microbiol Rev 23: 615–627. pmid:10525169
  72. 72. Oldfield NJ, Moran AP, Millar LA., Prendergast MM, Ketley JM (2002) Characterization of the Campylobacter jejuni heptosyltransferase II gene, waaF, provides genetic evidence that extracellular polysaccharide is lipid a core independent. J Bacteriol 184: 2100–2107. pmid:11914340
  73. 73. Karlyshev AV., Linton D, Gregson NA, Lastovica AJ, Wren BW (2000) Genetic and biochemical evidence of a Campylobacter jejuni capsular polysaccharide that accounts for Penner serotype specificity. Mol Microbiol 35: 529–541. pmid:10672176
  74. 74. Sirisena DM, Brozek K a, MacLachlan PR, Sanderson KE, Raetz CR (1992) The rfaC gene of Salmonella typhimurium. Cloning, sequencing, and enzymatic function in heptose transfer to lipopolysaccharide. J Biol Chem 267: 18874–18884. pmid:1527014
  75. 75. Schwan ET, Robertson BD, Brade H, van Putten JPM (1995) Gonococcal rfaF mutants express Rd2 chemotype LPS and do not enter epithelial host cells. Mol Microbiol 15: 267–275. pmid:7746148
  76. 76. Kanipes MI, Papp-szabo E, Guerry P, Monteiro MA (2006) Mutation of waaC, Encoding Heptosyltransferase I in Campylobacter jejuni 81–176, Affects the Structure of both Lipooligosaccharide and Capsular Carbohydrate. 188: 3273–3279. pmid:16621820
  77. 77. Klena JD, Gray S a., Konkel ME (1998) Cloning, sequencing, and characterization of the lipopolysaccharide biosynthetic enzyme heptosyltransferase I gene (waaC) from Campylobacter jejuni and Campylobacter coli. Gene 222: 177–185. pmid:9831648
  78. 78. Liu F, Tanner ME (2006) PseG of Pseudaminic Acid Biosynthesis: a UDP-sugar hydrolase as masked glycosyltransferase. J Biol Chem 281: 20902–20909. pmid:16728396
  79. 79. Miller WG, Parker CT (2010) Campylobacter and Arcobacter. In: Fratamico P, Liu Y, Kathariou S, editors. Genomes of Foodborne and Waterborne Pathogens. American Society for Microbiology Press. pp. 49–65.
  80. 80. Lin J, Michel LO, Zhang Q (2002) CmeABC functions as a multidrug efflux system in Campylobacter jejuni. Antimicrob Agents Chemother 46: 2124–2131. pmid:12069964
  81. 81. Iovine NM (2013) Resistance mechanisms in Campylobacter jejuni. Virulence 4: 230–240. pmid:23406779
  82. 82. Groicher KH, Firek B a, Fujimoto DF, Bayles KW (2000) The Staphylococcus aureus lrgAB operon modulates murein hydrolase activity and penicillin tolerance. J Bacteriol 182: 1794–1801. pmid:10714982
  83. 83. Abdelbaqi K, Ménard A, Prouzet-Mauleon V, Bringaud F, Lehours P and Mégraud F. (2007) Nucleotide sequence of the gyrA gene of Arcobacter species and characterization of human ciprofloxacin-resistant clinical isolates. FEMS Immunol Med Microbiol 49: 337–345. pmid:17378897
  84. 84. Carattoli A, Dionisi AM, Luzzi I (2002) Use of a LightCycler gyrA mutation assay for identification of ciprofloxacin-resistant Campylobacter coli. 214: 87–93. pmid:12204377
  85. 85. Ren GWN, Wang Y, Shen Z, Chen X, Shen J, Wu C. (2011) Rapid detection of point mutations in domain V of the 23S rRNA gene in erythromycin-resistant Campylobacter isolates by pyrosequencing. Foodborne Pathog Dis 8: 375–379. pmid:21166562
  86. 86. Kehrenberg C, Catry B, Haesebrouck F, De Kruif A, Schwarz S (2005) Novel spectinomycin/streptomycin resistance gene, aadA14, from Pasteurella multocida. Antimicrob Agents Chemother 49: 3046–3049. pmid:15980396
  87. 87. Roberts MC (1996) Tetracycline resistance determinants: mechanisms of action, regulation of expression, genetic mobility, and distribution. FEMS Microbiol Rev 19: 1–24. pmid:8916553
  88. 88. Douidah L, De Zutter L, Baré J, De Vos P, Vandamme P, Vandenberd O, et al. (2012) Occurrence of putative virulence genes in Arcobacter species isolated from humans and animals. J Clin Microbiol 50: 735–741. pmid:22170914
  89. 89. Figueras MJ, Levican A, Collado L (2017) “Arcobacter porcinus” sp. nov., a novel Arcobacter species uncovered by Arcobacter thereius. New Microbes New Infect 15: 104–106. pmid:28070334
  90. 90. Kaakoush NO, Man SM, Lamb S, Raftery MJ, Wilkins MR, Kovach Z, et al. (2010) The secretome of Campylobacter concisus. FEBS J 277: 1606–1617. pmid:20148967
  91. 91. Vandamme P., Falsen E, Pot B., Hoste B. Kersters K., De Ley J (1987) Identification of EF group 22 Campylobacters from gastroenteritis cases as Campylobacter concius. J Clin Microbiol 25: 2398–2399.
  92. 92. Backert S, Meyer TF (2006) Type IV secretion systems and their effectors in bacterial pathogenesis. Curr Opin Microbiol 9: 207–217. pmid:16529981
  93. 93. Pati A, Gronow S, Lapidus A, Copeland A, Glavina Del Rio T, Nolan D, et al. (2010) Complete genome sequence of Arcobacter nitrofigilis type strain (CI). Stand Genomic Sci 2: 300–308. pmid:21304714
  94. 94. Pingoud A, Fuxreiter M, Pingoud V, Wende W (2005) Type II restriction endonucleases: Structure and mechanism. Cell Mol Life Sci 62: 685–707. pmid:15770420