Next Article in Journal
Development of a Genoserotyping Method for Salmonella Infantis Detection on the Basis of Pangenome Analysis
Next Article in Special Issue
Mobilome Analysis of Achromobacter spp. Isolates from Chronic and Occasional Lung Infection in Cystic Fibrosis Patients
Previous Article in Journal
Rock Surface Fungi in Deep Continental Biosphere—Exploration of Microbial Community Formation with Subsurface In Situ Biofilm Trap
Previous Article in Special Issue
Plasmid Identification and Plasmid-Mediated Antimicrobial Gene Detection in Norwegian Isolates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Whole Genome Sequence-Based Prediction of Resistance Determinants in High-Level Multidrug-Resistant Campylobacter jejuni Isolates in Lithuania

by
Jurgita Aksomaitiene
1,*,
Aleksandr Novoslavskij
1,
Egle Kudirkiene
2,
Ausra Gabinaitiene
1 and
Mindaugas Malakauskas
1
1
Department of Food Safety and Quality, Faculty of Veterinary Medicine, Veterinary Academy, Lithuanian University of Health Sciences, Tilzes str. 18, LT 47181 Kaunas, Lithuania
2
Department of Veterinary and Animal Sciences, Faculty of Health and Medical Sciences, University of Copenhagen, Stigbøjlen 4, 1870 Frederiksberg C, Denmark
*
Author to whom correspondence should be addressed.
Microorganisms 2021, 9(1), 66; https://doi.org/10.3390/microorganisms9010066
Submission received: 28 November 2020 / Revised: 19 December 2020 / Accepted: 24 December 2020 / Published: 29 December 2020
(This article belongs to the Special Issue Bacterial Genomes and Evolution by Horizontal Gene Transfer (HGT))

Abstract

:
Spread of antibiotic resistance via mobile genetic elements associates with transfer of genes providing resistance against multiple antibiotics. Use of various comparative genomics analysis techniques enables to find intrinsic and acquired genes associated with phenotypic antimicrobial resistance (AMR) in Campylobacter jejuni genome sequences with exceptionally high-level multidrug resistance. In this study, we used whole genome sequences of seven C. jejuni to identify isolate-specific genomic features associated with resistance and virulence determinants and their role in multidrug resistance (MDR). All isolates were phenotypically highly resistant to tetracycline, ciprofloxacin, and ceftriaxone (MIC range from 64 to ≥256 µg/mL). Besides, two C. jejuni isolates were resistant to gentamicin, and one was resistant to erythromycin. The extensive drug-resistance profiles were confirmed for the two C. jejuni isolates assigned to ST-4447 (CC179). The most occurring genetic antimicrobial-resistance determinants were tetO, beta-lactamase, and multidrug efflux pumps. In this study, mobile genetic elements (MGEs) were detected in genomic islands carrying genes that confer resistance to MDR, underline their importance for disseminating antibiotic resistance in C. jejuni. The genomic approach showed a diverse distribution of virulence markers, including both plasmids and phage sequences that serve as horizontal gene transfer tools. The study findings describe in silico prediction of AMR and virulence genetics determinants combined with phenotypic AMR detection in multidrug-resistant C. jejuni isolates from Lithuania.

1. Introduction

Campylobacteriosis is one of the most common foodborne bacterial diseases worldwide [1]. The reported confirmed cases of human campylobacteriosis reached 246,571 in 2018 and remained the most frequently reported foodborne illness in the EU with notification rate of 64.1 per 100,000 population. [2]. Campylobacter jejuni is an important zoonotic pathogen causing significant infections, especially in young children, geriatric and immunocompromised patients [3,4]. Bacterial antibiotic resistance, especially multidrug resistance (MDR) and extensive drug resistance (XDR), has become a global emerging threat to public health systems [5,6]. Multidrug-resistance bacteria are frequently detected in humans and animals from developed and developing countries and pose a serious threat to human health [7].
Bacteria can acquire antimicrobial resistance through two leading pathways: chromosomal mutation and the acquisition of mobile genetic elements (MGEs) by horizontal gene transfer (HGT). Horizontal gene transfer of mobile genetic elements allows bacteria to exchange the genetic materials among pathogenic and non-pathogenic bacteria from different environments [8,9]. The HGT partly causes an increase in the adaptability of bacteria to environmental changes [10]. The transposition of MGEs can radically alter genome structure and genome sequence of bacteria, as antibiotic resistance is often spread via mobile genetic elements, and carry resistance against multiple antibiotics. Such bacteria with acquired AMR can spread and be transmitted to another environment, as resistant bacteria transfer from wild birds and cattle host to humans [11,12]. A better understanding of antibiotic resistance genes’ (ARGs’) circulation within bacterial species in different host niches and the mobility of these genes between different hosts could be important for identifying and analyzing multidrug resistance [13].
The study aimed to search the possible links among the phenotypic multidrug antimicrobial resistance and whole genome sequencing data (WGS) of C. jejuni isolates from different sources (cattle and wild bird feces).

2. Materials and Methods

2.1. Study Isolates

In total, seven C. jejuni isolates from bacterial culture collection of the Department of Food Safety and Quality of Lithuanian University of Health Sciences were tested in this study. These isolates were previously characterized by Multi Locus Sequence Typing (MLST) and assigned to CC179 and CC21 clonal complexes, with wide spread in Lithuania [14,15]. The isolates were stored at −80 °C in brain heart infusion broth (BHI) (Oxoid, Basingstoke, UK) with 30% glycerol (Stanlab, Lublin, Poland). The isolates’ recovery were performed by plating the stocks on Blood agar base No. 2 (Oxoid, Basingstoke, UK) supplemented with 5% defibrinated horse blood (E&O Laboratories Limited, Scotland, UK) and further incubation under microaerophilic conditions (5% oxygen, 10% carbon dioxide, and 85% nitrogen) at 42 °C for 48 h.

2.2. Antimicrobial Susceptibility Testing

All isolates were tested for antimicrobial susceptibility to erythromycin, tetracycline, gentamicin, ciprofloxacin, and ceftriaxone (all Sigma-Aldrich, Saint Louis, MO, USA). The minimum inhibitory concentration (MIC) was determined by the agar dilution method performed according to recommendations by the Clinical and Laboratory Standards Institute (CLSI) guidelines [16]. Isolates were cultured on Mueller-Hinton agar (Liofilchem, Italy) plates with dilutions ranging from 0.25 to 256 mg/mL for all antimicrobials. For each individual C. jejuni isolate, 5 µL of approximately 1 × 107 CFU/mL (OD600 = 0.1) bacterial suspension dissolved in phosphate-buffered saline (PBS) (E&O Laboratories Limited, Scotland, UK) were spotted onto Mueller-Hinton agar plates containing the corresponding antimicrobial agent concentration. The inoculated plates were incubated under microaerophilic conditions at 42 °C temperature for 24 h. The determination of MIC for each isolate was performed in triplicate. The MIC values were defined as the lowest concentration that produces complete inhibition of C. jejuni growth. For quality control, the reference isolate C. jejuni ATCC 33,560 was included. These breakpoints defined by the National Antimicrobial Resistance Monitoring System for Enteric Bacteria (NARMS) for antimicrobial susceptibility were used: for erythromycin ≥8 mg/mL, for tetracycline ≥2 mg/mL, for ciprofloxacin ≥1 mg/mL, for gentamicin ≥4 mg/mL, and for ceftriaxone ≥8 mg/mL. Isolates showing resistance to three or more antimicrobials were considered as multidrug-resistant.

2.3. Whole Genome Sequencing

DNA extraction was performed using the PureLink Genomic DNA Kit (Invitrogen, Carlsbad, CA, USA) according to the manufacturer’s instructions and finally eluted in 50 µL of sterile Mili-Q water. The concentration and integrity of DNA were quantified by Qubit 3.0 Fluorometer (Thermo Fisher Scientific, Waltham, MA, USA) with the double-stranded DNA (dsDNA) assay HS kit (lifetechnologies, Eugene, OR, USA) and 1% agarose gel, respectively.
According to the manufacturer’s instructions, DNA libraries were prepared using the Nextera XT library preparation kit (Illumina, San Diego, CA, USA). The sequencing was carried out at the NGS-MiSeq core facility of the University of Copenhagen using an Illumina MiSeq platform (Illumina, San Diego, CA, USA) with 250 bp paired-end read format and aiming to obtain an average genome depth of 50X. CLC Genomics Workbench version 6.5.1 was used for the adapter and quality trimming of the raw reads. Sequence reads were de novo assembled into contigs using SPAdes v.3.10 assembler [17]. The quality of the assembly was evaluated with QUAST v.2.3 (3). The subsystems’ annotation was obtained using the SEED-based automated annotation system after the data were uploaded to RAST (Rapid Annotation using Subsystem Technology) [18,19] genome server. Ribosomal multilocus sequence typing (rMLST) employing 53 genes encoding the bacterial ribosome protein subunits (rps genes) was performed using the Genome Comparator module of the BIGSdb platform on the PubMLST website [20]. BlastKOALA (https://www.kegg.jp/blastkoala/) [21] was used to perform KO (KEGG Orthology) assignments to characterize individual gene functions and reconstruct KEGG pathways, BRITE hierarchies, and KEGG modules. The presence of potential genes encoding antibiotic resistance was checked using the NCBI AMRFinder v.3.1.1 tool and the ResFinder v.3.0 and PointFinder v.3.1.0 (https://cge.cbs.dtu.dk/services/ResFinder/) [22] databases using thresholds of 90% identity and 60% gene coverage. Besides, a resistome prediction which uses BLAST algorithms to search of the AMR genes and SNPs was performed with Resistance Gene Identifier (RGI v.4.2.2) [23] with a previous coding sequence of the genome submission to the Comprehensive Antimicrobial Resistance Database (https://card.mcmaster.ca/analyze/rgi). Detection of antibiotic-resistance genes, mobile genetic elements, and mutation were performed through alignment to perform multiple alignments of the query with reference sequence downloaded from the NCBI reference (RefSeq) database using the CLUSTAW alignment tool https://www.genome.jp/tools-bin/clustalw [24]. CRISPR finder (https://crispr.i2bc.paris-saclay.fr/Server/) [25] and PathogenFinder 1.1 (https://cge.cbs.dtu.dk/services/PathogenFinder/) [26] databases of the Center for Genomic Epidemiology were used for the potential prediction of pathogenicity. A genome BLAST atlas was generated using isolates of C. jejuni comparison against the reference genome NCTC11168 (AL111168.1) using Gview (https://server.gview.ca/) [27]. For the dataset, alignment was used with BLASTn analysis with an e-value of 1 × 10−10, coverage 100%, and identity 80%. The IslandViewer version 4 server (https://www.pathogenomics.sfu.ca/islandviewer/) [28] was used to predict the putative genomic islands (GIs).

2.4. Data Availability

All genomes were submitted to NCBI under the following Accession Numbers: SAMN08794492; SAMN08794493; SAMN08803042; SAMN08803043; SAMN08803044; SAMN08803060; SAMN08803062 (BioProject PRJNA445645).

3. Results and Discussion

3.1. Phenotypic Antimicrobial Resistance Determination

All seven C. jejuni isolates were resistant to three or more antibiotics, and were identified as multidrug-resistant (MDR). MIC data are provided in Table 1. All isolates were phenotypically highly resistant to tetracycline, ciprofloxacin, and ceftriaxone (MIC range 64 ≥ 256 µg/mL). Besides, two C. jejuni isolates were resistant to gentamicin, and one was resistant to erythromycin. The extensive drug-resistance profiles were confirmed for the two C. jejuni isolates assigned to ST-4447 (CC179) (Table 2).

3.2. Genomics

All isolates with high-level multidrug resistance were further characterized by whole genome sequencing to identify transferable genes encoding antimicrobial resistance.
The general genome stats suggest that genome sizes range from 1.68 to 1.70 Mbp, with the average 30% G + C content, well within the range of available genome sequences. The summary of the genomes’ sequencing assignment is listed in Table 3.

3.3. Whole Genome Sequence-Based Genotypic Predictions of Antibiotic-Resistance Genes

The genomes of C. jejuni isolates were clustered into orthologue groups and annotated in RAST with the aim to identify traits involved in antimicrobial resistance and survival. Based on RAST analysis, 77 genes in three cattle, in each genome assigned to ST-21 (CC21), were annotated in association with virulence, disease, and defense. The analysis revealed the presence of virulence marker genes associated with adhesion (cadF and pEB1), invasion (yidC and yidD), and cytotoxin production (cdtA, cdtB, and cdtC). Nine protein-coding genes in genomes of CCm32 and CCm33 isolates were identified in phages, prophages, and transposable elements’ category including phages’ proteins involved in phage replication process, phage tail, and phage capsid proteins. Figure 1 shows the diagram of the genes associated with the functional categories of examined isolates.
The WGS data of examined isolates were mapped based on intrinsic and acquired genes known to be associated with phenotypic AMR. The genomes were also manually searched for genes known to being involved in AMR and virulence. The isolates CCm31 and CCm32 assigned to CC179 harbored cobalt-zinc-cadmium resistance determinants composed of czc, chr, ncc, and mer genes responsible for resistance to Zn, Cr, Ni and Hg, respectively. In three isolates, CCm35, CCm36, and CCm37, all assigned to CC21 (ST-21), platinum drug resistance ctpA, and a cationic antimicrobial peptide (CAMP) system genes cluster were identified (Table 4). The ctpA gene encodes the C-terminal processing protease for the photosystem’s D1 protein II reaction center complex related to virulence and cytotoxicity against host cells [29,30].
Type IV secretion system (T4S) genes virB2, virB4, virB8 and virB9 were identified in genomes of C. jejuni CCm33, CCm36, CCm37 isolates. The operon of the cmeABC multidrug efflux pump, consisting of cmeA, cmeB, and cmeC genes were identified in the genomes of three isolates (Table 4). The cmeABC, a resistance-nodulation-division (RND) type of efflux pump, contributes significantly to both intrinsic and acquired resistance to various antimicrobials in C. jejuni [31].
Additionally, the pmrA efflux pump, which belongs to the resistance-nodulation-division family of transporters and contributes to multidrug resistance of antimicrobials, was found in two C. jejuni isolates. The tetO gene, which codes the resistance to tetracycline, was detected in all C. jejuni isolates from cattle and in one isolate from a wild bird. Another tetracycline resistance gene tetM was found in the same two isolates, which had a tetO gene. The β-lactamase resistance gene blaOXA-448 was identified in all C. jejuni isolates assigned to clonal complex CC179; however, blaOXA-61 was identified in all isolates assigned to clonal complex CC21. Among the resistant isolates, several genes coding the virulence factors were found. The chemotaxis and flagellar motility genes, trg, flgE, and biofilm dispersion bdlA gene with increased adherent properties required for biofilm formation, were identified. These genes were identified in all C. jejuni isolates, which shows their widespread dissemination throughout C. jejuni genomes.

3.4. Point Mutation

Nucleotide and amino acid changes of C. jejuni genomic sequences are shown in Table 5. All CIP-resistant isolates harbored gyrA point mutation T86I. The G→A transversion in the rpsL gene was detected in two ST-4447 isolates. The six different point mutations, including deletion Lys→del in the L22 ribosomal protein gene rpIV, were observed in ST-6424. The amino acid changes in the cmeR gene, including D121N and E159K, were identified in two ST-21 isolates. Fourteen non-synonymous mutations were detected in the 23S rRNA gene.

3.5. Mobile Genetic Elements: Genomic Islands, Prophages and Plasmids

The different genomic approaches analysis revealed the prevalence of MGEs, including plasmids, pathogenicity islands, and bacteriophages in C. jejuni isolates. Most of the AMR and virulence factors were distributed within genomic island (GI) regions. Analysis of ARGs’ composition based on GI revealed 18 GIs with the size ranging in length from 5.61 kb to 58.83 kb. The largest GI was detected in the CCm36 (58.83 kb), containing gene tetO encoding tetracycline resistance, trg gene encoding chemotaxis, virB2 gene, putative DNA-invertase pinR, and ansZ coding L-asparaginase 2. In total, six GI (size 5.73 to 45.9 kb) were uniquely found in one of the ST-4447 (CCm33) isolate, which contains prophage CPS-53 integrase intS, flgE encoding flagellar motility, virB2, virB4, virB8, and virB9 genes. GIs play crucial roles in microbial genome evolution and adaptation of microbes to environments as part of a flexible gene pool [32]. Whole genome sequencing also revealed a 131.1 kb phage harbored by one isolate (C. jejuni CCm33) with high homology (identity 98%; e value 1.37 × 10−118) to Campylobacter phage PC5 (KX229736.1). Prophages are genomes of temperate phages that have infected a susceptible host bacterium, where they either integrate into the chromosome or exist as circular or linear plasmids [33].
The plasmid of 20,765 bp was identified in two C. jejuni isolates assigned to ST-21 (CC21) with high homology of repUS59 (pSSG1) plasmid sequence (FR824044) in the NCBI repository database. Furthermore, the nucleotide sequence comparison of genomic island in CCm37 showed presence of pTet plasmids with the tetO gene, L-asparginase 2, flagellin A and various homologous hypothetical genes. These acquired elements expand the genetic flexibility of pathogens. Plasmids and bacteriophages contribute to the C. jejuni evolution via adaptation and survivability with the novel functions integrated into the chromosome [34,35].

3.6. BLAST Identification and Diverse Genomic Locations of the C. jejuni

BLAST genomic atlas provides detailed valuable genomic insights that support genome-wide gene characterization of C. jejuni isolates and reveals how similar any genome is to the reference genome (NCTC 11168). Analysis of the G + C content distribution showed localized peak deviations from the genome’s average GC content. At positions from 71,887 to 50,973 bp, the chemotactic transducer gene pctC was detected in CCm33, CCm36, and CCm37 isolates. This gene consists of 2102 bp and encodes a transmembrane signaling receptor activity, which is to enhance the ability of signal transduction [36]. Genomic comparison with the reference C. jejuni NCTC11168 isolate genome revealed that the ssa1 gene encoding serotype 1-specific antigen (identity 100%) was found in all ST-21 C. jejuni isolates. This serotype-specific antigen has been shown to be involved in the endopeptidase activity regulation complex of Neisseria meningitides M0579 and Pasteurella haemolytica but is not present in C. jejuni NCTC 11168. The multidrug resistance gene MdtG (identity 97.88%) was observed in all ST-4447 isolates.
Moreover, the BLAST search of C. jejuni ST-21 isolates revealed a novel acquired complex that harbors the β-lactamase resistance gene blaOXA-133 (100% identity) belonging to the class-D β-lactamase family, yafP gene, and ykkC multidrug resistance gene in the region from 273,300 to 282,500 bases (Figure 2). The changes, most often associated with MGEs that have been acquired by HGT, can initiate a rapid mechanism for acquiring genes and confer novel function [37].
In conclusion, this knowledge provides insights on the distribution and genetic content of MGEs in multidrug Campylobacter jejuni isolates. Mobile genetic elements are principally important to facilitate horizontal genetic exchange and, therefore, induce the acquisition and spread of resistance genes. This may allow an assessment of how genes carried by mobile genetic elements can contribute to traits that are responsible for antimicrobial resistance and virulence. These findings highlighted the role of resistance determinants in the epidemiology of MGEs in C. jejuni genome sequences and indicate the potential for bacteria’s genomic plasticity. In this context, a better understanding of the acquisition of resistance determinants is critical to understand the capacity of resistance genes to spread in C. jejuni population.

Author Contributions

M.M. supervised the work; J.A. and E.K. performed the bioinformatics analysis; J.A. wrote the manuscript with input from all authors; M.M., E.K., A.G. and A.N. contributed to interpreting the results. All authors contributed to manuscript revision. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Oyarzabal, O.A.; Carrillo, C.D. Chapter 4—Isolation, identification, and typing of Campylobacter strains from food samples. In Campylobacter; Klein, G., Ed.; Academic Press: Cambridge, MA, USA, 2017; pp. 61–83. [Google Scholar]
  2. European Food Safety Authority and European Centre for Disease Prevention and Control (EFSA and ECDC). The European Union One Health 2018 Zoonoses Report. EFSA J. 2019, 17, e05926. [Google Scholar] [CrossRef] [Green Version]
  3. Sakran, W.; Hexner-Erlichman, Z.; Spiegel, R.; Batheesh, H.; Halevy, R.; Koren, A. Campylobacter gastroenteritis in children in north-eastern Israel comparison with other common pathogens. Sci. Rep. 2020, 10, 5823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Abebe, E.; Gugsa, G.; Ahmed, M. Review on Major Food-Borne Zoonotic Bacterial Pathogens. J. Trop. Med. 2020, 2020, e4674235. [Google Scholar] [CrossRef] [PubMed]
  5. Magiorakos, A.-P.; Srinivasan, A.; Carey, R.B.; Carmeli, Y.; Falagas, M.E.; Giske, C.G.; Harbarth, S.; Hindler, J.F.; Kahlmeter, G.; Olsson-Liljequist, B.; et al. Multidrug-resistant, extensively drug-resistant and pandrug-resistant bacteria: An international expert proposal for interim standard definitions for acquired resistance. Clin. Microbiol. Infect. 2012, 18, 268–281. [Google Scholar] [CrossRef] [Green Version]
  6. Van Duin, D.; Paterson, D. Multidrug Resistant Bacteria in the Community: Trends and Lessons Learned. Infect. Dis. Clin. N. Am. 2016, 30, 377–390. [Google Scholar] [CrossRef] [Green Version]
  7. Doyle, M.E. Multidrug-resistant pathogens in the food supply. Foodborne Pathog. Dis. 2015, 12, 261–279. [Google Scholar] [CrossRef]
  8. Bello-López, J.M.; Cabrero-Martínez, O.A.; Ibáñez-Cervantes, G.; Hernández-Cortez, C.; Pelcastre-Rodríguez, L.I.; Gonzalez-Avila, L.U.; Castro-Escarpulli, G. Horizontal Gene Transfer and Its Association with Antibiotic Resistance in the Genus Aeromonas spp. Microorganisms 2019, 7, 363. [Google Scholar] [CrossRef] [Green Version]
  9. Lee, K.; Kim, D.-W.; Lee, D.-H.; Kim, Y.-S.; Bu, J.-H.; Cha, J.-H.; Thawng, C.N.; Hwang, E.-M.; Seong, H.J.; Sul, W.J.; et al. Mobile resistome of human gut and pathogen drives anthropogenic bloom of antibiotic resistance. Microbiome 2020, 8, 2. [Google Scholar] [CrossRef] [Green Version]
  10. Emamalipour, M.; Seidi, K.; Zununi Vahed, S.; Jahanban-Esfahlan, A.; Jaymand, M.; Majdi, H.; Amoozgar, Z.; Chitkushev, L.T.; Javaheri, T.; Jahanban-Esfahlan, R.; et al. Horizontal Gene Transfer: From Evolutionary Flexibility to Disease Progression. Front. Cell Dev. Biol. 2020, 8, 229. [Google Scholar] [CrossRef]
  11. Partridge, S.R.; Kwong, S.M.; Firth, N.; Jensen, S.O. Mobile Genetic Elements Associated with Antimicrobial Resistance. Clin. Microbiol. Rev. 2018, 31, e00088-17. [Google Scholar] [CrossRef] [Green Version]
  12. Waddell, S.; Barnstedt, O.; Treiber, C. Chapter Four—Neural Transposition in the Drosophila Brain: Is It All Bad News? In Advances in Genetics; Epigenetic Shaping of Sociosexual Interactions; Yamamoto, D., Ed.; Academic Press: Cambridge, MA, USA, 2014; Volume 86, pp. 65–92. [Google Scholar]
  13. Chen, Y.-M.; Holmes, E.C.; Chen, X.; Tian, J.-H.; Lin, X.-D.; Qin, X.-C.; Gao, W.-H.; Liu, J.; Wu, Z.; Zhang, Y. Diverse and abundant resistome in terrestrial and aquatic vertebrates revealed by transcriptional analysis. Sci. Rep. 2020, 10, 18870. [Google Scholar] [CrossRef] [PubMed]
  14. Ramonaite, S.; Kudirkiene, E.; Tamuleviciene, E.; Leviniene, G.; Malakauskas, A.; Gölz, G.; Alter, T.; Malakauskas, M. Prevalence and genotypes of Campylobacter jejuni from urban environmental sources in comparison with clinical isolates from children. J. Med. Microbiol. 2014, 63 Pt 9, 1205–1213. [Google Scholar] [CrossRef]
  15. Ramonaite, S.; Tamuleviciene, E.; Alter, T.; Kasnauskyte, N.; Malakauskas, M. MLST genotypes of Campylobacter jejuni isolated from broiler products, dairy cattle and human campylobacteriosis cases in Lithuania. BMC Infect. Dis. 2017, 17, 430. [Google Scholar] [CrossRef] [PubMed]
  16. Patel and Clinical and Laboratory Standards Institute—2017—Performance Standards for Antimicrobial Susceptibility Testing. Available online: https://clsi.org/media/1469/m100s27_sample.pdf (accessed on 1 October 2020).
  17. Bankevich, A.; Nurk, S.; Antipov, D.; Gurevich, A.A.; Dvorkin, M.; Kulikov, A.S.; Lesin, V.M.; Nikolenko, S.I.; Pham, S.; Prjibelski, A.D.; et al. SPAdes: A New Genome Assembly Algorithm and Its Applications to Single-Cell Sequencing. J. Comput. Biol. 2012, 19, 455–477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Aziz, R.K.; Bartels, D.; Best, A.A.; DeJongh, M.; Disz, T.; Edwards, R.A.; Formsma, K.; Gerdes, S.; Glass, E.M.; Kubal, M.; et al. The RAST Server: Rapid Annotations using Subsystems Technology. BMC Genom. 2008, 9, 75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Overbeek, R.; Olson, R.; Pusch, G.D.; Olsen, G.J.; Davis, J.J.; Disz, T.; Edwards, R.A.; Gerdes, S.; Parrello, B.; Shukla, M.; et al. The SEED and the Rapid Annotation of microbial genomes using Subsystems Technology (RAST). Nucleic Acids Res. 2014, 42, D206–D214. [Google Scholar] [CrossRef]
  20. Jolley, K.A.; Bliss, C.M.; Bennett, J.S.; Bratcher, H.B.; Brehony, C.; Colles, F.M.; Wimalarathna, H.; Harrison, O.B.; Sheppard, S.K.; Cody, A.J.; et al. Ribosomal multilocus sequence typing: Universal characterization of bacteria from domain to strain. Microbiology 2012, 158 Pt 4, 1005–1015. [Google Scholar] [CrossRef]
  21. Kanehisa, M.; Sato, Y.; Morishima, K. BlastKOALA and GhostKOALA: KEGG Tools for Functional Characterization of Genome and Metagenome Sequences. J. Mol. Biol. 2016, 428, 726–731. [Google Scholar] [CrossRef] [Green Version]
  22. Zankari, E.; Hasman, H.; Cosentino, S.; Vestergaard, M.; Rasmussen, S.; Lund, O.; Aarestrup, F.M.; Larsen, M.V. Identification of acquired antimicrobial resistance genes. J. Antimicrob. Chemother. 2012, 67, 2640–2644. [Google Scholar] [CrossRef]
  23. 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]
  24. Thompson, J.D.; Gibson, T.J.; Higgins, D.G. Multiple Sequence Alignment Using ClustalW and ClustalX. Curr. Protoc. Bioinform. 2003, 1, 2.3.1–2.3.22. [Google Scholar] [CrossRef] [PubMed]
  25. Grissa, I.; Vergnaud, G.; Pourcel, C. CRISPRcompar: A website to compare clustered regularly interspaced short palindromic repeats. Nucleic Acids Res. 2008, 36, W145–W148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Cosentino, S.; Voldby Larsen, M.; Møller Aarestrup, F.; Lund, O. PathogenFinder—Distinguishing friend from foe using bacterial whole genome sequence data. PLoS ONE 2013, 8, e77302. [Google Scholar] [CrossRef] [Green Version]
  27. Petkau, A.; Stuart-Edwards, M.; Stothard, P.; Van Domselaar, G. Interactive microbial genome visualization with GView. Bioinformatics 2010, 26, 3125–3126. [Google Scholar] [CrossRef]
  28. Bertelli, C.; Laird, M.R.; Williams, K.P.; Simon Fraser University Research Computing Group; Lau, B.Y.; Hoad, G.; Winsor, G.L.; Brinkman, F.S.L. IslandViewer 4: Expanded prediction of genomic islands for larger-scale datasets. Nucleic Acids Res. 2017, 45, W30–W35. [Google Scholar] [CrossRef]
  29. Anbudurai, P.R.; Mor, T.S.; Ohad, I.; Shestakov, S.V.; Pakrasi, H.B. The ctpA gene encodes the C-terminal processing protease for the D1 protein of the photosystem II reaction center complex. Proc. Natl. Acad. Sci. USA 1994, 91, 8082–8086. [Google Scholar] [CrossRef] [Green Version]
  30. Inagaki, N.; Yamamoto, Y.; Mori, H.; Satoh, K. Carboxyl-terminal processing protease for the D1 precursor protein: Cloning and sequencing of the spinach cDNA. Plant Mol. Biol. 1996, 30, 39–50. [Google Scholar] [CrossRef]
  31. Su, C.-C.; Yin, L.; Kumar, N.; Dai, L.; Radhakrishnan, A.; Bolla, J.R.; Lei, H.-T.; Chou, T.-H.; Delmar, J.A.; Rajashankar, K.R.; et al. Structures and transport dynamics of a Campylobacter jejuni multidrug efflux pump. Nat. Commun. 2017, 8, 171. [Google Scholar] [CrossRef]
  32. Hacker, J.; Carniel, E. Ecological fitness, genomic islands and bacterial pathogenicity. EMBO Rep. 2001, 2, 5–81. [Google Scholar] [CrossRef]
  33. Uelze, L.; Grützke, J.; Borowiak, M.; Hammerl, J.A.; Juraschek, K.; Deneke, C.; Tausch, S.H.; Malorny, B. Typing methods based on whole genome sequencing data. One Health Outlook 2020, 2, 3. [Google Scholar] [CrossRef] [Green Version]
  34. Dobrindt, U.; Hacker, J. Mobile genetic elements and pathogenicity islands encoding bacterial toxins. In The Comprehensive Sourcebook of Bacterial Protein Toxins; Elsevier Ltd.: Amsterdam, The Netherlands, 2006; pp. 44–63. [Google Scholar]
  35. Oliveira, P.H.; Touchon, M.; Cury, J.; Rocha, E.P.C. The chromosomal organization of horizontal gene transfer in bacteria. Nat. Commun. 2017, 8, 841. [Google Scholar] [CrossRef] [PubMed]
  36. Parales, R.E.; Ditty, J.L. Chemotaxis. In Handbook of Hydrocarbon and Lipid Microbiology; Timmis, K.N., Ed.; Springer: Berlin/Heidelberg, Germany, 2010; pp. 1529–1543. [Google Scholar] [CrossRef]
  37. Sheppard, S.K.; Guttman, D.S.; Fitzgerald, J.R. Population genomics of bacterial host adaptation. Nat. Rev. Genet. 2018, 19, 549–565. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Subsystem category distribution of seven C. jejuni isolates.
Figure 1. Subsystem category distribution of seven C. jejuni isolates.
Microorganisms 09 00066 g001
Figure 2. Circular genomic atlas of 7 C. jejuni isolates in comparison to the reference C. jejuni NCTC11168 genome. The circle is divided into arcs representing the genomes as labeled. The black histogram represents the G + C content, and the purple-green histogram represents the G + C deviation.
Figure 2. Circular genomic atlas of 7 C. jejuni isolates in comparison to the reference C. jejuni NCTC11168 genome. The circle is divided into arcs representing the genomes as labeled. The black histogram represents the G + C content, and the purple-green histogram represents the G + C deviation.
Microorganisms 09 00066 g002
Table 1. Minimum inhibitory concentration among Campylobacter jejuni isolates from cattle and wild birds.
Table 1. Minimum inhibitory concentration among Campylobacter jejuni isolates from cattle and wild birds.
Antimicrobial Agent
IsolateTETCIPGENAXOERY
MIC, µg/mL
CCm2612812821284
CCm31>256640.51280.5
CCm321282568648
CCm33>25612841280.25
CCm351281280.51280.5
CCm362561280.252560.5
CCm3764640.25640.5
Breakpoint≥2≥1≥4≥8≥8
TET, tetracycline; ERY, erythromycin; CIP, ciprofloxacin; GEN, gentamicin; AXO, ceftriaxone; MIC, minimum inhibitory concentration.
Table 2. Phenotypic antimicrobial resistance profiles of C. jejuni isolates from cattle and wild birds.
Table 2. Phenotypic antimicrobial resistance profiles of C. jejuni isolates from cattle and wild birds.
IsolateSourceClonal ComplexSequence TypeAntimicrobial Resistance Profile
CCm26Wild birdCC179ST-6424TET + CIP + AXO
CCm31Wild birdCC179ST-4447TET + CIP + AXO
CCm32Wild birdCC179ST-4447TET + CIP + AXO + GEN + ERY *
CCm33Wild birdCC179ST-4447TET + CIP + AXO + GEN *
CCm35CattleCC21ST-21TET + CIP + AXO
CCm36CattleCC21ST-21TET + CIP + AXO
CCm37CattleCC21ST-21TET + CIP + AXO
TET, tetracycline; ERY, erythromycin; CIP, ciprofloxacin; GEN, gentamicin; AXO, ceftriaxone; CC, clonal complex; ST, sequence type; * extensive drug resistant (XRD) isolates.
Table 3. List of genomic data of C. jejuni isolates.
Table 3. List of genomic data of C. jejuni isolates.
CCm26CCm31CCm32CCm33CCm35CCm36CCm37
Metrics of sequence data
Coverage (x)160120180400220180160
Genomic data report
Size (bp)1682833168519116850331721979170139117051641705114
No. of contigs31592529232421
GC (%)30.330.430.430.330.330.330.3
N5012285462137158860184236116706122859135281
L504944665
L75917871098
Genes1768177217701814178117811779
CDs1725172917271771173717381736
Subsystems
rRNAs2222222
tRNAs40404040414040
Table 4. Genetic determinants associated of virulence and resistance markers found in C. jejuni. czc, cobalt zinc cadmium resistance system; ctpA, platinum resistance; AMPs, antimicrobial peptides; CAMP, cationic antimicrobial peptide; T4S, Type IV secretion system; flgE, flagellar motility; trg, chemotaxis; bdlA, biofilm formation; +, positive; −, negative; +/−, uncomplete system.
Table 4. Genetic determinants associated of virulence and resistance markers found in C. jejuni. czc, cobalt zinc cadmium resistance system; ctpA, platinum resistance; AMPs, antimicrobial peptides; CAMP, cationic antimicrobial peptide; T4S, Type IV secretion system; flgE, flagellar motility; trg, chemotaxis; bdlA, biofilm formation; +, positive; −, negative; +/−, uncomplete system.
C. jejuniVirulence MarkersResistance Markers
Heavy Metal ResistanceAMPs Sensing SystemInvasionMultidrug Efflux PupmsTetracyclineβ-Lactams
czcctpACAMPT4SflgE, trg, bdlAcmeABCpmrAtetOtetMblaOXA-448blaOXA-61blaOXA-451blaOXA-133
CCm26+++++++
CCm31+++++
CCm32+++++
CCm33+/−++++
CCm35++++++++
CCm36+++/−++++++++
CCm37+++/−++++++++++
Table 5. Nucleotide and amino acid changes of C. jejuni genomic sequences.
Table 5. Nucleotide and amino acid changes of C. jejuni genomic sequences.
L22cmeRgyrArpsL23S rRNA
MutationNucleotide ChangeAmino Acid ChangeMutationNucleotide ChangeAmino Acid ChangeMutationNucleotide ChangeAmino Acid ChangeMutationNucleotide ChangeAmino Acid ChangeMutationNucleotide Change
I165V 1ATT→GTTIle→ValG144D 1,2,3,4,5GGT→GATGly→AspR285K 1,2,3,4AGG→AAGArg→LysA119T 3,4GCT→ACTAla→Thr287G > A 5,6,7G→A
S109A 1TCT→GCTSer→AlaS207G 1,2,3,4AGC→GGCSer→GlyA312T 1,2,3,4GCT→ACTAla→Thr 296C > G 5,6,7C→G
T119A 1ACT→GCTThr→AlaD121N 6,7GAC→AACAsp→AsnA664V 1,2,3,4GCC→GTCAla→Val 298G > A 5,6,7G→A
T120P 1ACA→CCAThr→ProE159K 6,7GAA→AAAGlu→LysT665S 1,2,3,4ACT→AGTThr→Ser 327G > A 5,6,7G→A
V137A 1GTG→GCGVal→Ala T804A 1,2,3,4ACA→GCAThr→Ala 364G > C 5,6,7G→C
K123→del 1AAA→delLys→del T86I 1,2,3,4,5,6,7ACA→ATATrr→Ile 554A > C 5,6,7A→C
571T > G 5,6,7T→G
1027A > G 5,6,7A→G
1485C > T 4C→T
1735T > C 4T→C
1739T > C 4T→C
1752T > C 4T→C
1759A > G 4A→G
1761G > A 4G→A
Superscript numbers indicate the C. jejuni isolates harboring specific nucleotide and amino acid changes: 1 CCm26; 2 CCm31; 3 CCm32; 4 CCm33; 5 CCm35; 6 CCm36; 7 CCm37.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Aksomaitiene, J.; Novoslavskij, A.; Kudirkiene, E.; Gabinaitiene, A.; Malakauskas, M. Whole Genome Sequence-Based Prediction of Resistance Determinants in High-Level Multidrug-Resistant Campylobacter jejuni Isolates in Lithuania. Microorganisms 2021, 9, 66. https://doi.org/10.3390/microorganisms9010066

AMA Style

Aksomaitiene J, Novoslavskij A, Kudirkiene E, Gabinaitiene A, Malakauskas M. Whole Genome Sequence-Based Prediction of Resistance Determinants in High-Level Multidrug-Resistant Campylobacter jejuni Isolates in Lithuania. Microorganisms. 2021; 9(1):66. https://doi.org/10.3390/microorganisms9010066

Chicago/Turabian Style

Aksomaitiene, Jurgita, Aleksandr Novoslavskij, Egle Kudirkiene, Ausra Gabinaitiene, and Mindaugas Malakauskas. 2021. "Whole Genome Sequence-Based Prediction of Resistance Determinants in High-Level Multidrug-Resistant Campylobacter jejuni Isolates in Lithuania" Microorganisms 9, no. 1: 66. https://doi.org/10.3390/microorganisms9010066

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