Next Article in Journal
Macrolide versus Non-Macrolide in Combination with Steroids for the Treatment of Lobar or Segmental Mycoplasma pneumoniae Pneumonia Unresponsive to Initial Macrolide Monotherapy
Next Article in Special Issue
ST11 Carbapenem-Resistant Klebsiella pneumoniae Clone Harboring blaNDM Replaced a blaKPC Clone in a Tertiary Hospital in China
Previous Article in Journal
Acinetobacter baylyi Strain BD413 Can Acquire an Antibiotic Resistance Gene by Natural Transformation on Lettuce Phylloplane and Enter the Endosphere
Previous Article in Special Issue
Occurrence, Phenotypic and Molecular Characteristics of Extended-Spectrum Beta-Lactamase-Producing Escherichia coli in Healthy Turkeys in Northern Egypt
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genomic Analysis of a mcr-9.1-Harbouring IncHI2-ST1 Plasmid from Enterobacter ludwigii Isolated in Fish Farming

1
National Reference Laboratory of Antibiotic Resistances and Healthcare Associated Infections, Department of Infectious Diseases, National Institute of Health Dr. Ricardo Jorge, 1649-016 Lisbon, Portugal
2
Centre for the Studies of Animal Science, Institute of Agrarian and Agri-Food Sciences and Technologies, University of Porto, 4051-401 Porto, Portugal
3
AL4AnimalS, Associate Laboratory for Animal and Veterinary Sciences, 1300-477 Lisboa, Portugal
4
Laboratory of Biology and Ecotoxicology, Department of Environmental Health, National Institute of Health Dr. Ricardo Jorge, 1649-016 Lisbon, Portugal
5
Division of Aquaculture, Upgrading and Bioprospecting, Portuguese Institute for the Sea and Atmosphere, IPMA, 1749-077 Lisbon, Portugal
6
CIIMAR, Interdisciplinary Centre of Marine and Environmental Research, University of Porto, Terminal de Cruzeiros do Porto, Av. General Norton de Matos s/n, 4450-208 Matosinhos, Portugal
7
Molecular Diagnostics Research Group, School of Biological and Chemical Sciences, National University of Ireland, H91 DK59 Galway, Ireland
8
Centre for One Health, Ryan Institute, National University of Ireland, H91 TK33 Galway, Ireland
9
CIISA, Center for Interdisciplinary Research in Animal Health, Faculty of Veterinary Medicine, University of Lisbon, 1300-477 Lisbon, Portugal
*
Author to whom correspondence should be addressed.
Antibiotics 2022, 11(9), 1232; https://doi.org/10.3390/antibiotics11091232
Submission received: 16 August 2022 / Revised: 2 September 2022 / Accepted: 5 September 2022 / Published: 10 September 2022

Abstract

:
This study analyzed the resistome, virulome and mobilome of an MCR-9-producing Enterobacter sp. identified in a muscle sample of seabream (Sparus aurata), collected in a land tank from multitrophic fish farming production. Average Nucleotide Identity analysis identified INSAq77 at the species level as an Enterobacter ludwigii INSAq77 strain that was resistant to chloramphenicol, florfenicol and fosfomycin and was susceptible to all other antibiotics tested. In silico antimicrobial resistance analyses revealed genes conferring in silico resistance to β-lactams (blaACT-88), chloramphenicol (catA4-type), fosfomycin (fosA2-type) and colistin (mcr-9.1), as well as several efflux pumps (e.g., oqxAB-type and mar operon). Further bioinformatics analysis revealed five plasmid replicon types, including the IncHI2/HI2A, which are linked to the worldwide dissemination of the mcr-9 gene in different antibiotic resistance reservoirs. The conserved nickel/copper operon rcnR-rcnA-pcoE-ISSgsp1-pcoS-IS903-mcr-9-wbuC was present, which may play a key role in copper tolerance under anaerobic growth and nickel homeostasis. These results highlight that antibiotic resistance in aquaculture are spreading through food, the environment and humans, which places this research in a One Health context. In fact, colistin is used as a last resort for the treatment of serious infections in clinical settings, thus mcr genes may represent a serious threat to human health.

Graphical Abstract

1. Introduction

The emergence of colistin resistance in the last years is a serious threat to the treatment of infections caused by multidrug-resistant bacteria in human medicine. Consequently, colistin, a last resort antibiotic, is categorized by the World Health Organization (WHO) as one of the highest priority, critically important antibiotics for human medicine. The use of colistin in veterinary medicine has been prohibited in various countries. However, colistin is still an antibiotic extensively used in veterinary medicine for infections caused by Enterobacterales [1]. In aquatic animal species, colistin is also used to treat bacterial infections. Nevertheless, its application in aquaculture is much less common than that used in terrestrial animals [1].
Acquired resistance to colistin has been mostly due to chromosomal mutation in the PmrA/PmrB and PhoP/PhoQ, two-component regulatory systems, or through the increased production of capsular polysaccharide [2,3]. Since the first report of the plasmid-mediated colistin resistance (PMCR) gene mcr-1 in an Escherichia coli strain isolated from a pig in China, we have seen an increase in the number of cases worldwide, in different bacterial species [4,5]. To date, ten PMCR genes (mcr-1 to mcr-10), including variants, have been described, mainly in Enterobacterales from animals, humans and the environment [6]. Originally, MCR-producing bacteria were mostly isolated from animals, including pigs [7]. The exception are the mcr-9 and mcr-10 that were identified in human isolates: Salmonella enterica serotype Typhimurium and Enterobacter roggenkampii, respectively [8,9].
Unlike the situation in terrestrial animals, PMCR in aquaculture has largely been ignored [10]. However, colistin-resistant bacteria have also been found in aquaculture [11,12], being considered by the authors as a major source of colistin resistance genes. Recent studies in China showed the presence of MCR-1- and MCR-3-producing bacteria isolated in the food chain, such as in aquaculture fish and shrimp [13] and in E. coli recovered from grass carp fish farms [14].
In Vietnam, mcr-1 was detected in E. coli isolated from striped catfish grown in ponds [15,16] and a study from Spain reported mcr-1 in S. enterica serovar Rissen isolated from mussels [17]. Recently, mcr-1 was detected in E. coli isolated from fish guts of rainbow trout in Lebanon [18]. In Czech Republic, mcr-type genes were detected in colistin-resistant Enterobacterales and Acinetobacter strains isolated from aquaculture products (frog legs, crab meat and pangasius meat) originating from Vietnam [19]. Shen and co-workers also demonstrated the association between aquaculture and a high incidence of mcr-1-positive E. coli carried by humans [13]. Furthermore, a comparative analysis of the resistome of integrated and monoculture aquaculture ponds using metagenomics suggest that freshwater aquaculture is rich in opportunistic pathogen-associated taxonomic groups that may host antibiotic-resistant genes (including mcr) associated with critically important antibiotics used in human medicine [20]. Indeed, it has been proposed that some PMCR genes may have originated in aquatic environments, since MCR-3, MCR-4 and MCR-7 proteins showed an elevated level of homology to phosphoethanolamine transferases found in aquatic bacteria [10,11].
The mcr-9 gene is an emerging variant of the PMCR determinants, which was first identified in 2019, in a S. enterica isolated from a human patient in the USA [8]. Along with mcr-1, among the mcr-like genes, mcr-9 is the most widely disseminated [7]. The mcr-9 gene can be found worldwide in different reservoirs (human, animal, food and environment) and in various species of Enterobacteriaceae [21], which makes this resistance mechanism a problem under the perspective of One Health.
This study aimed to analyse the resistome, virulome and mobilome of an MCR-9-producing Enterobacter sp. isolated from farmed Sparus aurata and, to our knowledge, this is the first description of the colistin resistance mcr-9 gene in the aquaculture environment.

2. Results and Discussion

The mcr-1 and mcr-9 variants are the most widespread mcr-family genes. The mcr-9 gene was identified in 40 countries through six continents, with 61.5% of the mcr-9-positive strains isolated in the United States [7]. In that study, S. enterica was the most common host species, especially in turkeys and chickens. Furthermore, other systematic reviews showed that isolates carrying mcr-9 were detected in 21 countries through six continents, mainly from Europe. mcr-9-positive isolates were disseminated by various genera and species of Enterobacteriaceae isolates among which Enterobacter spp. were predominant (37.0%) [21]. More than 50% of the isolates were from human origin, being 29.0%, 3.6% and 2.9% from animal, environmental and food, respectively.
Here, an mcr-9-producing isolate (INSAq77) identified in a muscle sample of a commercial size S. aurata, collected during the winter season (March 2018), in a land tank from a fish multitrophic farming, in the south of Portugal is described. To our knowledge is the first description of mcr-9 gene in the aquaculture environment.
Average Nucleotide Identity (ANI) analysis performed by NCBI identified INSAq77 at the species level as E. ludwigii. The genome sequences of E. cloacae are 98.98% identical by ANI to the E. ludwigii, with 82.5% coverage of the genome. Indeed, INSAq77 isolate was identified as E. cloacae by the VITEK®2 automated identification system (BioMérieux, Marcy-l’Étoile, France) and sequencing of the 16S rRNA gene. However, it is well known that precise species identification for the taxonomy of Enterobacter is complex [22] and that hsp60 gene sequencing showed a higher species diversity than MALDI-TOF [23]. Seven species have been grouped within the E. cloacae complex: Enterobacter cloacae, Enterobacter asburiae, Enterobacter hormaechei, Enterobacter kobei, E. ludwigii, Enterobacter mori and Enterobacter nimipressuralis, which share at least 60% similarity in their genome with E. cloacae [24].
E. ludwigii was first described as a novel Enterobacter species in 2005 [25]. All strains are naturally resistant to ampicillin, amoxicillin-clavulanic acid, and cefoxitin due to the production of a chromosomal AmpC β-lactamase. Antibiotic-resistant E. ludwigii has been found mainly in clinical samples [26], although a CTX-M-producing E. ludwigii was recently described in an environmental isolate collected from a wastewater treatment plant in India [27]. Recently, antibiotic-resistant E. ludwigii was also identified in India, in moribund goldfish collected from ornamental fish farms [28]. On the other hand, E. ludwigii has been suggested as a potential probiotic microorganism in agriculture and aquaculture [29,30].
INSAq77 strain was resistant to chloramphenicol (MIC 32 mg/L), florfenicol (32 mg/L) and fosfomycin (64 mg/L) but was susceptible to all other antibiotics tested; cefoxitin and amoxicillin/clavulanic acid are intrinsic resistances. The colistin MIC for MCR-9-producing E. ludwigii as 1 mg/L, within susceptible EUCAST breakpoint. Indeed, other studies have demonstrated that the presence of an MCR-9 enzyme not always is associated with colistin resistance [8,31,32]. Nevertheless, recent studies showed that mcr genes might enhance the survival ability of bacteria under clinical colistin pressure, thereby potentially leading to treatment failure [33,34].
This study also analyzed the resistome, virulome and mobilome of this MCR-9-producing E. ludwigii isolated from farmed Sparus aurata. The analysis of WGS yielded 225 contigs, ranging from 237 to 244,787 bp. The draft genome contained a total assembly length of 5,276,953 bp, with estimated depth coverage of 30.7×; the GC content was 54.1%.
The MCR-9-producing INSAq77 E. ludwigii isolate belonged to the ST1342 lineage, first reported here. The whole-genome SNP-based phylogenetic tree using the 75 E. ludwigii genomes indicated that INSAq77 is not closely related to the other studied isolates (Figure 1).
Indeed, INSAq77 has 12% of nucleotide sequence divergence with the closest strain (NZ_VLMJ00000000), an E. ludwigii isolated from the lung of a clinical patient, in 2016, in the USA (PRJNA553678) [35]. The two other MCR-9-producing isolates (NZ_JAGDFR00000000 and NZ_JAGDFs00000000) were grouped into another cluster.
In silico antimicrobial resistance analyses using ResFinder 4.1, with a threshold of 90% identity and a minimum length of 60%, revealed acquired genes conferring resistance to β-lactams (blaACT-88, here firstly identified), fosfomycin (fosA2-type) and colistin (mcr-9.1). Furthermore, a total of 21 genes were detected in silico by CARD RGI perfect, strict and loose algorithms, involved in efflux, transport and permeability, which might justify the florfenicol and chloramphenicol resistance identified by phenotypic methods (Table 1 and Table S1).
CARD loose algorithm (match bitscore less than the curated one blastp bitscore) [36] identified that INSAq77 also harbors a catA4-type gene (64.8% of identity), which might infer resistance to chloramphenicol. However, the low percentage of identity it is not enough to assure the phenicol resistance causality of catA4-type gene. Other resistance mechanisms might be involved; indeed, the multiple antibiotic resistance (mar) locus, a resistance-nodulation-cell division (RND) antibiotic efflux pump detected in silico by CARD RGI strict algorithm, has been reported to contribute to chloramphenicol resistance [37]. Furthermore, the multiple antibiotic resistance oqxAB-type locus, another RND multidrug efflux pump operon was detected, which has been reported to contribute to multidrug resistance [38]. Diminished susceptibility to different antibiotic classes (e.g., aminoglycosides, fluoroquinolones and tetracycline) were bioinformatically predicted (e.g., rsmA, adeF, ramA, acrABR and soxAS), although the isolate was phenotypically susceptible; this can be explained by the fact that efflux pumps are frequently associated with a low decrease in antibiotic susceptibility, which may not translate to a change in phenotype [39].
The acquired disinfectant resistance gene formA-type, a plasmid-mediated formaldehyde resistance mechanism [40], was also identified. The ability to survive aldehyde disinfection processes is clinically significant, with possible cross-resistance to antibiotics [41]. Furthermore, the INSAq77 isolate carried the terC virulence gene, commonly associated with IncHI2 plasmids and conferring resistance to tellurium, where soluble salts, especially potassium tellurite, were used clinically in humans as antimicrobial agents [42].
PathogenFinder predicted the strain as being “human pathogenic” with a probability of 77.7% due to the presence of 74 genes belonging to known pathogenic protein families (Table S2). Indeed, in addition to the known E. cloacae complex genes encoding pathogenic proteins and the homologous sequences of pathogenic proteins from Citrobacter koseri, Enterobacter spp., E. coli, S. enterica and Shigella spp. were found in the study. Mobile genetic elements (MGE), such as plasmids, prophages and transposons among others, are main drivers for the spread of antibiotic resistance [43]. In this study, nine insertion sequences (IS) were found using MobileElementFinder tool: IS26, ISKpn28, ISSen4-type, IS30-type, ISEcl1-type, ISKpn43-type, ISKpn24-type, IS100-type and ISPpu12-type.
A total of eleven prophage regions were also identified using PHASTER tool (Figure S1), of which two regions were intact (PHAGE_Salmon_SEN4_NC_029015 and PHAGE_Entero_HK542_NC_019769), eight regions were incomplete and one region was questionable (PHAGE_Shigel_SfIV_NC_022749). Figure 2 shows the schematic representation of the phage-related proteins identified in the intact and questionable prophages. The size of the three prophages ranged from 16.7Kb to 32.6Kp with an average GC content of 52.7%. These prophages were firstly described in S. enterica subspecies salamae collected in the Czech Republic [44], E. coli isolated in Hong Kong and Shigella flexneri collected in Bangladesh [45], corroborating that MGE can be excised and integrated from the chromosome and MGE into each other. Indeed, several studies have already shown the presence of mcr-type genes in prophages [46,47], indicating the role of these MGE in the dissemination of antibiotic resistance. Furthermore, two cryptic prophages were detected by PathogenFinder (CP4-6 and CP4-57) which, although they do not form active phage particles or lyse their captors, can be considered the relatively permanent reservoirs of antibiotic resistance genes [48].
Further bioinformatics analysis revealed the presence of five plasmid replicon types: ColE10, Col(pHAD28)-type, IncFIA(HI1)-type, IncR-type, IncHI2 and IncHI2A, the last two linked to the worldwide dissemination of mcr-9 gene [7,21]. The mcr-9 gene can be found in different reservoirs (human, animal, food and environment), in various species of Enterobacteriaceae strains, mostly associated with IncHI2/IncHI2A plasmid replicons [21]. Indeed, as observed in our study, several works demonstrated the prevalence of mcr-9-harboring IncHI2/IncHI2A plasmids among Enterobacteriaceae isolates: e.g., Carroll et al., in 2019, detected 59/65 assemblies where IncHI2 and/or IncHI2A plasmid replicon were present on the same contig as mcr-9 [8].
The mcr-9 gene was found in a 30,314 bp length contig, manually assembled after visual inspection and alignment of contigs (Aq77p_57, Aq77p_155, Aq77p_191, Aq77p_196, Aq77p_213) against themselves using CLC Genomics Finishing Module v.20.0.1 (QIAGEN, Aarhus); the GC content was 47.6%. The analysis of mcr-9-harbouring contig using the Microbial Nucleotide MegaBLAST analysis against the complete plasmids database revealed nine mcr-9-carrying IncHI2 plasmid sequences (>99.9% identity, >97% query coverage and e-value 0.0) from multiple species, collected in different antibiotic resistance reservoirs worldwide, including human clinical/colonization samples (Table 2). Of notice, three plasmids of sequence type 1 (ST1), accordingly with the IncHI2 pDLST scheme, were collected from environmental samples during an extended blaIMP-4-associated carbapenemase outbreak in an Australian hospital [50].
In INSAq77 IncHI2 plasmid and in all others studied here, except p565_1 from a C. freundii strain (NZ_CP038657, Figure 3), the mcr-9 gene was surrounded upstream by an IS903 element and downstream by a wbuC family gene, encoding a cupin fold metalloprotein, followed by IS26.
The mcr-9 gene is frequently associated with the wbuC gene in different bacterial species, suggesting an essential role of this wbuC gene for the activity of the MCR-9 enzyme [51]. Indeed, this gene was proposed to have transferred together with mcr-9 as a whole fragment from Buttiauxella spp. [51]. On the other hand, qseB/qseC two-component regulators were absent, which could explain the susceptibility to colistin of INSAq77 isolate. Indeed, based on previous studies, in the presence of subinhibitory concentrations of colistin, the qseB/qseC regulatory system can induce the expression of the mcr-9 gene, which results in an increase in MIC values [32,51,52]. The mcr-9 gene is being described as part of a mcr-9 cassette containing the rcnR-rcnA-pcoE-pcoS-IS903-mcr-9-wbuC core structure (Figure 3) [21].
As shown in Figure 3 and Figure 4, the genetic background immediately upstream of mcr-9 was consistent among IncHI2 mcr-9-bearing plasmids. The exception is the presence of an ISSgsp1 element from the IS66 family in the INSAq77 mcr-9-harbouring contig, showing 100% of identity with an isolate of Klebsiella pneumoniae collected, in 2012, from a human patient in the USA (CP007734). Furthermore, in this study, the presence of the conserved nickel/copper operon (i.e., rcnA, rcnR, pcoE and pcoS genes), which plays a key role in copper tolerance under anaerobic growth and nickel homeostasis in bacteria, was also detected. Downstream of the mcr-9 gene, the nucleotide sequences, including that of INSAq77, were genetically diverse (Figure 3). Indeed, different regions were present among the IncHI2 plasmids also analyzed in this study, namely regions involved in resistance to copper (pcoABCDRSE), silver (silESRCBAP), arsenic (arsCBRH) and/or mercury (merEDACPTR).
The abundance of mcr variants and alleles in bacteria isolated from aquatic reservoirs suggests that these enzymes may play another role, namely a defense system against natural peptides and/or bacteriophages [53]. mcr-9 gene was firstly described in the USA, in a clinical S. Typhimurium isolate, which demonstrates the high transmission potential of this colistin resistance determinant and places this research in a One Health context [54].

3. Materials and Methods

3.1. Study Design and Bacterial Identification

MCR-9-producing Enterobacter sp. INSAq77 was isolated from a seabream (S. aurata) of commercial-size (500–1500 g), which was collected in March 2018 in a land tank from a fish multitrophic farming [55]. This station is in the Ria Formosa Natural Park (south of Portugal) with a semi-intensive production system. Animal welfare was safeguarded during production and transport accordingly with the European Commission SANTE/2016/G2/009 recommendations [56]. Species identification was performed by VITEK®2 Automated Identification System (BioMérieux, Marcy-l’Étoile, France), using GN ID card and by amplification of the 16S rRNA gene, as previously described [55].

3.2. Antimicrobial Susceptibility Testing

Antibiotic susceptibility testing was performed by disk diffusion (amoxicillin/clavulanic acid, aztreonam, cefepime, cefotaxime, cefoxitin, ceftazidime, ciprofloxacin, ertapenem, gentamicin, imipenem, meropenem, piperacillin/tazobactam and trimethoprim/sulfamethoxazole; Bio-Rad, Marnes-la-Coquette, France) and minimum inhibitory concentration (MIC) by in house broth microdilution (colistin, chloramphenicol, florfenicol, flumequine and oxytetracycline) and E-test® (fosfomycin; bioMérieux, Hazelwood, MO, USA), as previously described [55].

3.3. Whole-Genome Sequencing

DNA was extracted from freshly grown overnight culture (MagnaPure 96 Instrument, Roche, Manheim, Germany) and was quantified using Qubit fluorometer (Thermo Fisher Scientific, Waltham, MA, USA). Dual-indexed Nextera XT kit was used to library preparation followed by paired-end sequencing (2 × 250 bp) on a MiSeq Illumina platform (Illumina Inc., San Diego, CA, USA), according to the manufacturer’s instructions.

3.4. Genome Annotation and Analysis

Genomes were de novo assembled using the INNUca v4.2.2 pipeline (https://github.com/B-UMMI/INNUca; accessed on 4 January 2022): after quality control analysis performed by FastQC v0.11.5 and cleaning (Trimmomatic v0.38), genomes were assembled with SPAdes 3.14.0 and subsequently improved using Pilon v1.23. In silico multilocus sequence type (MLST) prediction was performed using the MLST v2.19.0. A Prokka v1.13.3 was utilized to annotate the assemblies. Average Nucleotide identity (ANI) was performed at NCBI to confirm the INSAq77 bacterial species [57]. All de novo contigs were BLAST searched against GenBank’s non-redundant nucleotide collection (nr/nt) [58]. QIAGEN CLC Genome Finishing Module v.20.0.1 (QIAGEN, Aarhus, Denmark) was used for visual inspection and manual editing by the alignment of contigs using BLAST against the contigs themselves, allowing contig joining and scaffolding.

3.5. Phylogenomic Analyses of E. ludwigii Genomes

All E. ludwigii genomes (n = 76, Table S3) available at NCBI library were used on the genomic comparison process. Single nucleotide polymorphisms (SNPs) phylogenetic analysis was performed by using CSI Phylogeny v1.4 (https://cge.cbs.dtu.dk/services/CSIPhylogeny/; accessed on 4 January 2022) with default options (reference strain NZ_CP017279). Phylogenetic tree image was visualized and edited by FigTree v1.4.4 (https://tree.bio.ed.ac.uk/software/figtree/; accessed on 4 January 2022).

3.6. Resistome, Virulome and Mobilome Analysis

Online bioinformatics tools and databases available at the Center for Genomic Epidemiology (CGE) (www.genomicepidemiology.org; accessed on 4 January 2022) were used to investigate the presence of antimicrobial resistance genes (ResFinder 4.1), virulence factors (VirulenceFinder 2.0), plasmids (PlasmidFinder 2.1 and pMLST 2.0 IncHI2 DLST configuration), mobile genetic elements (MobileElementFinder v1.0.3) and pathogenicity (PathogenFinder 1.1). The Comprehensive Antibiotic Resistance Database (CARD) with the “perfect”, “strict” and “loose” default settings were also used to characterize antibiotic resistance [36]. ISsaga was used for the identification and annotation of insertion sequences [59]. PHASTER search web tool (https://phaster.ca; accessed on 4 January 2022) was applied to detect, identify and annotate prophage sequences [49]. All analyses were performed using default parameters.

3.7. Plasmid Characterization

BRIG v.0.95 was used to perform a circular comparison between the complete sequence of INSAq77 mcr-9-harbouring contig and the highly similar plasmids detected by performing BLAST against the Microbial Nucleotide BLAST database for complete plasmids (https://blast.ncbi.nlm.nih.gov/Blast.cgi; accessed on 4 January 2022). The genetic environment of the mcr-9 gene was manually revisited using CLC Genomics Workbench v20.0.4 (QIAGEN, Aarhus, Denmark). EasyFig v2.2.5 was used for the visualization and comparison of mcr-9 genetic environment [60].

4. Conclusions

This work reinforces the knowledge that water environments play a crucial role in the spread of antibiotic resistance and that important antibiotic resistance mechanisms, such as mcr genes conferring low- or medium/high-level resistance to colistin, are also present in aquaculture. This fact, allows antibiotic-resistant bacteria to spread through food and through the environment, resulting in serious threats to human health [61].
The use of phenotypic methods to determine susceptibility to antibiotics may be a limitation, as they may not identify the low expression associated with the presence of a particular gene, as in this case. Thus, the implementation of high throughput methods in laboratories, such as the WGS, will make an important contribution to the detection of under-expressed genes, mostly when they are of clinical importance. Thus, the presence of antibiotic-susceptible isolates in different settings, such as the INSAq77 mcr-9-carrying strain isolated in aquaculture, highlights the risk of the silent dissemination of important resistance determinants, among which, in fact, the genes encoding such PMCR are an important example. Of concern is also the possible co-selection of antibiotic-resistant bacteria when exposed to heavy metals (copper and zinc), often used as growth promoters in aquaculture and terrestrial animal farms.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/antibiotics11091232/s1, Figure S1: Presence of prophages in INSAq77 genome identified by PHASTER. A total of eleven prophage regions were identified, of which two regions were intact, eight regions were incomplete and one region was questionable. (a) Table with prophage characteristics; (b) Location of predicted prophages within INSAq77 contigs. Region types were marked with colors: intact (green), questionable (blue) and incomplete (red).; Table S1: Loose best-hit results (≥65% of identity), by predicted gene, obtained using the Resistance Gene Identifier (RGI).; Table S2: Results obtained from prediction of a bacteria’s pathogenicity towards human hosts using PathogenFinder (https://cge.cbs.dtu.dk/services/PathogenFinder/, accessed on 4 January 2022). Results highlighted in green are those not matching protein pathogenic families.; Table S3: Demographic and genomic characteristics of the E. ludwigii isolates used for the phylogenomic analysis.

Author Contributions

Conceptualization, V.M. and M.C.; methodology, V.M., V.S., N.M.B. and E.F.; validation, V.M. and M.C.; formal analysis, V.M. and V.S.; investigation, V.M., V.S., E.F., T.R., E.D. and M.C.; writing—original draft preparation, V.M.; writing—review and editing, V.M., V.S., T.R., N.M.B., E.F., T.S., E.D. and M.C.; visualization, V.M. and V.S.; supervision, M.C.; funding acquisition, M.C. All authors have read and agreed to the published version of the manuscript.

Funding

V.S. had her Ph.D. fellowship granted by FCT (Fundação para a Ciência e a Tecnologia) [SFRH/BD/133100/2017], co-financed by European Social Fund and the Operational Program for Human Capital (POCH), Portugal. The work was supported with funding from FCT/MCTES [UIDB/00211/2020] through national funds. This work was partially supported by funding from the European Union’s Horizon 2020 Research and Innovation programme under grant agreement No 773830: One Health European Joint Programme (WORLDCOM project), and partially from FCT/MCTES [UIDB/00211/2020] through national funds.

Data Availability Statement

The authors confirm all supporting data, code and protocols have been provided within the article or through supplementary data files. The E. ludwigii whole genome shotgun (WGS) project has the project accession JABRPH000000000. The new blaACT-type nucleotide sequence was submitted to the GenBank Database as blaACT-88 with accession number MW887657, after request of the new allele number to NCBI (https://www.ncbi.nlm.nih.gov/pathogens/submit-beta-lactamase/; accessed on 4 January 2022).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kempf, I.; Jouy, E.; Chauvin, C. Colistin use and colistin resistance in bacteria from animals. Int. J. Antimicrob. Agents 2016, 48, 598–606. [Google Scholar] [CrossRef]
  2. El-Sayed Ahmed, M.A.E.G.; Zhong, L.L.; Shen, C.; Yang, Y.; Doi, Y.; Tian, G.B. Colistin and its role in the Era of antibiotic resistance: An extended review (2000–2019). Emerg. Microbes Infect. 2020, 9, 868–885. [Google Scholar] [CrossRef] [PubMed]
  3. Poirel, L.; Jayol, A.; Nordmann, P. Polymyxins: Antibacterial Activity, Susceptibility Testing, and Resistance Mechanisms Encoded by Plasmids or Chromosomes. Clin. Microbiol. Rev. 2017, 30, 557–596. [Google Scholar] [CrossRef] [PubMed]
  4. Andrade, F.F.; Silva, D.; Rodrigues, A.; Pina-Vaz, C. Colistin update on its mechanism of action and resistance, present and future challenges. Microorganisms 2020, 8, 1716. [Google Scholar] [CrossRef] [PubMed]
  5. Sun, J.; Zhang, H.; Liu, Y.H.; Feng, Y. Towards Understanding MCR-like Colistin Resistance. Trends Microbiol. 2018, 26, 794–808. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, S.; Abbas, M.; Rehman, M.U.; Wang, M.; Jia, R.; Chen, S.; Liu, M.; Zhu, D.; Zhao, X.; Gao, Q.; et al. Updates on the global dissemination of colistin-resistant Escherichia coli: An emerging threat to public health. Sci. Total Environ. 2021, 799, 149280. [Google Scholar] [CrossRef]
  7. Ling, Z.; Yin, W.; Shen, Z.; Wang, Y.; Shen, J.; Walsh, T.R. Epidemiology of mobile colistin resistance genes mcr-1 to mcr-9. J. Antimicrob. Chemother. 2020, 75, 3087–3095. [Google Scholar] [CrossRef] [PubMed]
  8. Carroll, L.M.; Gaballa, A.; Guldimann, C.; Sullivan, G.; Henderson, L.O.; Wiedmanna, M. Identification of Novel Mobilized Colistin Resistance Gene mcr-9 in a Multidrug-Resistant, Colistin-Susceptible Salmonella enterica Serotype Typhimurium Isolate. MBio 2019, 10, e00853-19. [Google Scholar] [CrossRef]
  9. Wang, C.; Feng, Y.; Liu, L.; Wei, L.; Kang, M.; Zong, Z. Identification of novel mobile colistin resistance gene mcr-10. Emerg. Microbes. Infect. 2020, 9, 508–516. [Google Scholar] [CrossRef]
  10. Shen, Y.; Zhang, R.; Schwarz, S.; Wu, C.; Shen, J.; Walsh, T.R.; Wang, Y. Farm animals and aquaculture: Significant reservoirs of mobile colistin resistance genes. Environ. Microbiol. 2020, 22, 2469–2484. [Google Scholar] [CrossRef] [Green Version]
  11. Cabello, F.C.; Godfrey, H.P. Aquaculture, exaptation, and the origin of mcr-positive colistin resistance. Antimicrob. Agents Chemother. 2018, 62, 13–14. [Google Scholar] [CrossRef] [PubMed]
  12. Cabello, F.C.; Tomova, A.; Ivanova, L.; Godfrey, H.P. Aquaculture and mcr colistin resistance determinants. MBio 2017, 8, 4–5. [Google Scholar] [CrossRef] [PubMed]
  13. Lei, T.; Zhang, J.; Jiang, F.; He, M.; Zeng, H.; Chen, M.; Wu, S.; Wang, J.; Ding, Y.; Wu, Q. First detection of the plasmid-mediated colistin resistance gene mcr-1 in virulent Vibrio parahaemolyticus. Int. J. Food Microbiol. 2019, 308, 108290. [Google Scholar] [CrossRef] [PubMed]
  14. Lv, L.; Cao, Y.; Yu, P.; Huang, R.; Wang, J.; Wen, Q.; Zhi, C.; Zhang, Q.; Liu, J.-H. Detection of mcr-1 Gene among Escherichia coli Isolates from Farmed Fish and Characterization of mcr-1-Bearing IncP Plasmids. Antimicrob. Agents Chemother. 2018, 62, e02378-17. [Google Scholar] [CrossRef] [PubMed]
  15. Hoa, T.T.T.; Nakayama, T.; Huyen, H.M.; Harada, K.; Hinenoya, A.; Phuong, N.T.; Yamamoto, Y. Extended-spectrum beta-lactamase-producing Escherichia coli harbouring sul and mcr-1 genes isolates from fish gut contents in the Mekong Delta, Vietnam. Lett. Appl. Microbiol. 2020, 71, 78–85. [Google Scholar] [CrossRef]
  16. Yamaguchi, T.; Kawahara, R.; Harada, K.; Teruya, S.; Nakayama, T.; Motooka, D.; Nakamura, S.; Do Nguyen, P.; Kumeda, Y.; Van Dang, C.; et al. The presence of colistin resistance gene mcr-1 and -3 in ESBL producing Escherichia coli isolated from food in Ho Chi Minh City, Vietnam. FEMS Microbiol. Lett. 2018, 365, fny100. [Google Scholar] [CrossRef]
  17. Lozano-Leon, A.; Garcia-Omil, C.; Dalama, J.; Rodriguez-Souto, R.; Martinez-Urtaza, J.; Gonzalez-Escalona, N. Detection of colistin resistance mcr-1 gene in Salmonella enterica serovar Rissen isolated from mussels, Spain, 2012- to 2016. Euro Surveill. 2019, 24, 1900200. [Google Scholar] [CrossRef]
  18. Hassan, J.; Eddine, R.Z.; Mann, D.; Li, S.; Deng, X.; Saoud, I.P.; Kassem, I.I. The mobile colistin resistance gene, mcr-1.1, is carried on Incx4 plasmids in multidrug resistant E. coli isolated from rainbow trout aquaculture. Microorganisms 2020, 8, 1636. [Google Scholar] [CrossRef]
  19. Kalová, A.; Gelbíčová, T.; Overballe-Petersen, S.; Litrup, E.; Karpíšková, R. Characterisation of colistin-resistant Enterobacterales and Acinetobacter strains carrying mcr genes from asian aquaculture products. Antibiotics 2021, 10, 838. [Google Scholar] [CrossRef]
  20. Xu, C.; Lv, Z.; Shen, Y.; Liu, D.; Fu, Y.; Zhou, L.; Liu, W.; Chen, K.; Ye, H.; Xia, X.; et al. Metagenomic insights into differences in environmental resistome profiles between integrated and monoculture aquaculture farms in China. Environ. Int. 2020, 144, 106005. [Google Scholar] [CrossRef]
  21. Li, Y.; Dai, X.; Zeng, J.; Gao, Y.; Zhang, Z.; Zhang, L. Characterization of the global distribution and diversified plasmid reservoirs of the colistin resistance gene mcr-9. Sci. Rep. 2020, 10, 8113. [Google Scholar] [CrossRef] [PubMed]
  22. Wu, W.; Feng, Y.; Zong, Z. Precise Species Identification for Enterobacter: A Genome Sequence-Based Study with Reporting of Two Novel Species, Enterobacter quasiroggenkampii sp. nov. and Enterobacter quasimori sp. nov. mSystems 2020, 5, e00527-20. [Google Scholar] [CrossRef] [PubMed]
  23. Mateos, M.; Hernández-García, M.; Del Campo, R.; Martínez-García, L.; Gijón, D.; Morosini, M.I.; Ruiz-Garbajosa, P.; Cantón, R. Emergence and Persistence over Time of Carbapenemase-Producing Enterobacter Isolates in a Spanish University Hospital in Madrid, Spain (2005-2018). Microb. Drug Resist. 2021, 27, 895–903. [Google Scholar] [CrossRef] [PubMed]
  24. Mezzatesta, M.L.; Gona, F.; Stefani, S. Enterobacter cloacae complex: Clinical impact and emerging antibiotic resistance. Future Microbiol. 2012, 7, 887–902. [Google Scholar] [CrossRef]
  25. Hoffmann, H.; Stindl, S.; Stumpf, A.; Mehlen, A.; Monget, D.; Heesemann, J.; Schleifer, K.H.; Roggenkamp, A. Description of Enterobacter ludwigii sp. nov., a novel Enterobacter species of clinical relevance. Syst. Appl. Microbiol. 2005, 28, 206–212. [Google Scholar] [CrossRef]
  26. Davin-Regli, A.; Lavigne, J.P.; Pagès, J.M. Enterobacter spp.: Update on taxonomy, clinical aspects, and emerging antimicrobial resistance. Clin. Microbiol. Rev. 2019, 32, e00002-19. [Google Scholar] [CrossRef]
  27. Ali, A.; Sultan, I.; Mondal, A.H.; Siddiqui, M.T.; Gogry, F.A.; Haq, Q.M.R. Lentic and effluent water of Delhi-NCR: A reservoir of multidrug-resistant bacteria harbouring blaCTX-M, blaTEM and blaSHV type ESBL genes. J. Water Health 2021, 19, 592–603. [Google Scholar] [CrossRef]
  28. Preena, P.G.; Dharmaratnam, A.; Raj, N.S.; Raja, S.A.; Nair, R.R.; Swaminathan, T.R. Antibiotic-resistant Enterobacteriaceae from diseased freshwater goldfish. Arch. Microbiol. 2021, 203, 219–231. [Google Scholar] [CrossRef]
  29. Lee, K.E.; Adhikari, A.; Kang, S.M.; You, Y.H.; Joo, G.J.; Kim, J.H.; Kim, S.J.; Lee, I.J. Isolation and characterization of the high silicate and phosphate solubilizing novel strain Enterobacter ludwigii GAK2 that promotes growth in rice plants. Agronomy 2019, 9, 144. [Google Scholar] [CrossRef]
  30. Porto de Souza Vandenberghe, L.; Marcela Blandon Garcia, L.; Rodrigues, C.; Cândido Camara, M.; Vinícius de Melo Pereira, G.; de Oliveira, J.; Ricardo Soccol, C. Potential applications of plant probiotic microorganisms in agriculture and forestry. AIMS Microbiol. 2017, 3, 629–648. [Google Scholar] [CrossRef]
  31. Börjesson, S.; Greko, C.; Myrenås, M.; Landén, A.; Nilsson, O.; Pedersen, K. A link between the newly described colistin resistance gene mcr-9 and clinical Enterobacteriaceae isolates carrying blaSHV-12 from horses in Sweden. J. Glob. Antimicrob. Resist. 2020, 20, 285–289. [Google Scholar] [CrossRef] [PubMed]
  32. Tyson, G.H.; Li, C.; Hsu, C.H.; Ayers, S.; Borenstein, S.; Mukherjee, S.; Tran, T.T.; McDermot, P.F.; Zhao, S. The mcr-9 gene of Salmonella and Escherichia coli is not associated with colistin resistance in the United States. Antimicrob. Agents Chemother. 2020, 64, e00573-20. [Google Scholar] [CrossRef] [PubMed]
  33. Zhang, H.; Zhao, D.; Quan, J.; Hua, X.; Yu, Y. mcr-1 facilitated selection of high-level colistin-resistant mutants in Escherichia coli. Clin. Microbiol. Infect. 2019, 25, 517.e1–517.e4. [Google Scholar] [CrossRef] [PubMed]
  34. Zhu, X.Q.; Liu, Y.Y.; Wu, R.; Xun, H.; Sun, J.; Li, J.; Feng, Y.; Liu, J.H. Impact of mcr-1 on the Development of High Level Colistin Resistance in Klebsiella pneumoniae and Escherichia coli. Front. Microbiol. 2021, 12, 9666782. [Google Scholar] [CrossRef] [PubMed]
  35. Ferreira, I.; Beisken, S.; Lueftinger, L.; Weinmaier, T.; Klein, M.; Bacher, J.; Patel, R.; von Haeseler, A.; Posch, A.E. Species identification and antibiotic resistance prediction by analysis of whole-genome sequence data by use of ARESdb: An analysis of isolates from the unyvero lower respiratory tract infection trial. J. Clin. Microbiol. 2020, 58, e00273-20. [Google Scholar] [CrossRef]
  36. 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]
  37. Schwarz, S.; Kehrenberg, C.; Doublet, B.; Cloeckaert, A. Molecular basis of bacterial resistance to chloramphenicol and florfenicol. FEMS Microbiol. Rev. 2004, 28, 519–542. [Google Scholar] [CrossRef]
  38. Li, J.; Zhang, H.; Ning, J.; Sajid, A.; Cheng, G.; Yuan, Z.; Hao, H. The nature and epidemiology of OqxAB, a multidrug efflux pump. Antimicrob. Resist. Infect. Control 2019, 8, 44. [Google Scholar] [CrossRef]
  39. Fernández, L.; Hancock, R.E.W. Adaptive and mutational resistance: Role of porins and efflux pumps in drug resistance. Clin. Microbiol. Rev. 2012, 25, 661–681. [Google Scholar] [CrossRef]
  40. Kümmerle, N.; Feucht, H.H.; Kaulfers, P.M. Plasmid-mediated formaldehyde resistance in Escherichia coli: Characterization of resistance gene. Antimicrob. Agents Chemother. 1996, 40, 2276–2279. [Google Scholar] [CrossRef] [Green Version]
  41. Cloete, T.E. Resistance mechanisms of bacteria to antimicrobial compounds. Int. Biodeterior. Biodegrad. 2003, 51, 277–282. [Google Scholar] [CrossRef]
  42. Taylor, D.E. Bacterial tellurite resistance. Trends Microbiol. 1999, 7, 111–115. [Google Scholar] [CrossRef]
  43. 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] [PubMed]
  44. Mikalová, L.; Bosák, J.; Hříbková, H.; Dědičová, D.; Benada, O.; Šmarda, J.; Šmajs, D. Novel temperate phages of Salmonella enterica subsp. salamae and subsp. diarizonae and their activity against pathogenic S. enterica subsp. enterica isolates. PLoS ONE 2017, 12, 1–14. [Google Scholar] [CrossRef] [PubMed]
  45. Jakhetia, R.; Talukder, K.A.; Verma, N.K. Isolation, characterization and comparative genomics of bacteriophage SfIV: A novel serotype converting phage from Shigella flexneri. BMC Genom. 2013, 14, 677. [Google Scholar] [CrossRef] [PubMed]
  46. Pan, Y.; Fang, Y.; Feng, Y.; Lyu, N.; Chen, L.; Li, J.; Xu, X.; Zhu, B.; Hu, Y. Discovery of mcr-3.1 gene carried by a prophage located in a conjugative IncA/C2 plasmid from a Salmonella Choleraesuis clinical isolate. J. Infect. 2021, 82, 414–451. [Google Scholar] [CrossRef]
  47. Zhang, C.; Feng, Y.; Liu, F.; Jiang, H.; Qu, Z.; Lei, M.; Wang, J.; Zhang, B.; Hu, Y.; Ding, J.; et al. A Phage-Like IncY Plasmid Carrying the mcr-1 Gene in Escherichia coli from a Pig Farm in China. Antimicrob. Agents Chemother. 2017, 61, e02035-16. [Google Scholar] [CrossRef]
  48. Wang, X.; Wood, T.K. Cryptic prophages as targets for drug development. Drug Resist. Updat. 2016, 27, 30–38. [Google Scholar] [CrossRef]
  49. Arndt, D.; Grant, J.R.; Marcu, A.; Sajed, T.; Pon, A.; Liang, Y.; Wishart, D.S. PHASTER: A better, faster version of the PHAST phage search tool. Nucleic Acids Res. 2016, 44, W16–W21. [Google Scholar] [CrossRef]
  50. Kizny Gordon, A.; Phan, H.T.T.; Lipworth, S.I.; Cheong, E.; Gottlieb, T.; George, S.; Peto, T.E.A.; Mathers, A.J.; Walker, A.S.; Crook, D.W.; et al. Genomic dynamics of species and mobile genetic elements in a prolonged blaIMP-4-associated carbapenemase outbreak in an Australian hospital. J. Antimicrob. Chemother. 2020, 75, 873–882. [Google Scholar] [CrossRef]
  51. Kieffer, N.; Royer, G.; Decousser, J.-W.; Bourrel, A.-S.; Palmieri, M.; Ortiz De La Rosa, J.-M.; Jacquier, H.; Denamur, E.; Nordmann, P.; Poirel, L. mcr-9, an Inducible Gene Encoding an Acquired Phosphoethanolamine Transferase in Escherichia coli, and Its Origin. Antimicrob. Agents Chemother. 2019, 63, e00965-19. [Google Scholar] [CrossRef]
  52. Bitar, I.; Papagiannitsis, C.C.; Kraftova, L.; Chudejova, K.; Mattioni Marchetti, V.; Hrabak, J. Detection of Five mcr-9 -Carrying Enterobacterales Isolates in Four Czech Hospitals. mSphere 2020, 5, e01008-20. [Google Scholar] [CrossRef] [PubMed]
  53. Khedher, M.B.; Baron, S.A.; Riziki, T.; Ruimy, R.; Raoult, D.; Diene, S.M.; Rolain, J.M. Massive analysis of 64,628 bacterial genomes to decipher water reservoir and origin of mobile colistin resistance genes: Is there another role for these enzymes? Sci. Rep. 2020, 10, 5970. [Google Scholar] [CrossRef] [PubMed]
  54. Xiaomin, S.; Yiming, L.; Yuying, Y.; Zhangqi, S.; Yongning, W.; Shaolin, W. Global impact of mcr-1-positive Enterobacteriaceae bacteria on “one health”. Crit. Rev. Microbiol. 2020, 46, 565–577. [Google Scholar] [CrossRef] [PubMed]
  55. Salgueiro, V.; Manageiro, V.; Bandarra, N.M.; Reis, L.; Caniça, M. Bacterial Diversity and Antibiotic Susceptibility of Sparus aurata from Aquaculture. Microorganisms 2020, 8, 1343. [Google Scholar] [CrossRef] [PubMed]
  56. European Commission Directorate-General for Health and Food Safety; Braak, K.; Schrijver, R.; Bergevoet, R. Welfare of Farmed Fish: Common Practices during Transport and at Slaughter: Executive Summary; Publications Office of European Union: Luxembourg, 2017. [Google Scholar] [CrossRef]
  57. Ciufo, S.; Kannan, S.; Sharma, S.; Badretdin, A.; Clark, K.; Turner, S.; Brover, S.; Schoch, C.L.; Kimchi, A.; DiCuccio, M. Using average nucleotide identity to improve taxonomic assignments in prokaryotic genomes at the NCBI. Int. J. Syst. Evol. Microbiol. 2018, 68, 2386–2392. [Google Scholar] [CrossRef] [PubMed]
  58. Camacho, C.; Coulouris, G.; Avagyan, V.; Ma, N.; Papadopoulos, J.; Bealer, K.; Madden, T.L. BLAST+: Architecture and applications. BMC Bioinform. 2009, 10, 421. [Google Scholar] [CrossRef]
  59. Varani, A.M.; Siguier, P.; Gourbeyre, E.; Charneau, V.; Chandler, M. ISsaga is an ensemble of web-based methods for high throughput identification and semi-automatic annotation of insertion sequences in prokaryotic genomes. Genome Biol. 2011, 12, R30. [Google Scholar] [CrossRef]
  60. Sullivan, M.J.; Petty, N.K.; Beatson, S.A. Easyfig: A genome comparison visualizer. Bioinformatics 2011, 27, 1009–1010. [Google Scholar] [CrossRef]
  61. Cherak, Z.; Loucif, L.; Moussi, A.; Rolain, J.M. Epidemiology of mobile colistin resistance (mcr) genes in aquatic environments. J. Glob. Antimicrob. Resist. 2021, 27, 51–62. [Google Scholar] [CrossRef]
Figure 1. Whole-genome SNP-based phylogenetic tree showing the relationship between 75 E. ludwigii genomes. The scale bar indicates 4% of nucleotide sequence divergence. The numbers at the nodes indicate percentage bootstrap replicates of 100. Sequences in the tree are indicated as GenBank accession number. Strain of the present study and the other MCR-9-producing isolates are highlighted in yellow. Blue colour indicates cluster containing INSAq77.
Figure 1. Whole-genome SNP-based phylogenetic tree showing the relationship between 75 E. ludwigii genomes. The scale bar indicates 4% of nucleotide sequence divergence. The numbers at the nodes indicate percentage bootstrap replicates of 100. Sequences in the tree are indicated as GenBank accession number. Strain of the present study and the other MCR-9-producing isolates are highlighted in yellow. Blue colour indicates cluster containing INSAq77.
Antibiotics 11 01232 g001
Figure 2. Schematic representation of phage-related proteins identified in the intact and questionable prophages by PHASTER prophage database (https://phaster.ca, accessed on 4 January 2022) [49]. The arrow indicates the sequence orientation (5′ to 3′ above the black line and 3′ to 5′ under it). The abbreviations are: Att (phage attachment site), Coa (Phage coat protein), Fib (Phage Tail Fibre), Int (Phage integrase), Hyp (Hypothetical protein), Pla (Phage plate protein), PLP (Phage-like protein), Por (Portal protein), Sha (Phage tail shaft protein) and Ter (Terminase).
Figure 2. Schematic representation of phage-related proteins identified in the intact and questionable prophages by PHASTER prophage database (https://phaster.ca, accessed on 4 January 2022) [49]. The arrow indicates the sequence orientation (5′ to 3′ above the black line and 3′ to 5′ under it). The abbreviations are: Att (phage attachment site), Coa (Phage coat protein), Fib (Phage Tail Fibre), Int (Phage integrase), Hyp (Hypothetical protein), Pla (Phage plate protein), PLP (Phage-like protein), Por (Portal protein), Sha (Phage tail shaft protein) and Ter (Terminase).
Antibiotics 11 01232 g002
Figure 3. Comparative analysis of INSAq77 mcr-9.1-containing contig with nine closely related IncHI2 mcr-9-harboring plasmids using the BLAST Ring Image Generator.
Figure 3. Comparative analysis of INSAq77 mcr-9.1-containing contig with nine closely related IncHI2 mcr-9-harboring plasmids using the BLAST Ring Image Generator.
Antibiotics 11 01232 g003
Figure 4. Schematic representation of the genetic environment of INSAq77 mcr-9.1-containing contig with others IncHI2 mcr-9-harboring plasmids. Boxed arrows indicate direction of transcription for all genes. Blue bars: normal tblastx matches; red: Inverted matches; depth of shading: percentage blast match. Color-coding for the genes inside mcr-9 cassette: dark red, mcr-9 gene; cyan, mobile DNA; purple, other genes; grey, other CDSs. The scale is represented in base pairs.
Figure 4. Schematic representation of the genetic environment of INSAq77 mcr-9.1-containing contig with others IncHI2 mcr-9-harboring plasmids. Boxed arrows indicate direction of transcription for all genes. Blue bars: normal tblastx matches; red: Inverted matches; depth of shading: percentage blast match. Color-coding for the genes inside mcr-9 cassette: dark red, mcr-9 gene; cyan, mobile DNA; purple, other genes; grey, other CDSs. The scale is represented in base pairs.
Antibiotics 11 01232 g004
Table 1. Perfect and strict best-hit results by predicted gene, obtained using the Resistance Gene Identifier (RGI).
Table 1. Perfect and strict best-hit results by predicted gene, obtained using the Resistance Gene Identifier (RGI).
ContigRGI CriteriaARO TermDetection Criteria ModelAMR Gene FamilyDrug ClassResistance Mechanism% Identity Matching Region% Length Reference Sequence
INSAq77p_155Perfectmcr-9.1protein homologMCR phosphoethanolamine transferasepeptide antibioticantibiotic target alteration100.0100.0
INSAq77p_4StrictCRPprotein homologresistance-nodulation-cell division (RND) antibiotic efflux pumpmacrolide antibiotic, fluoroquinolone antibiotic, penamantibiotic efflux99.1100.0
INSAq77p_10StrictACT-12protein homologACT beta-lactamasecarbapenem, cephalosporin, cephamycin, penamantibiotic inactivation98.7100.0
INSAq77p_1StrictFosA2protein homologfosfomycin thiol transferasefosfomycinantibiotic inactivation98.6100.0
INSAq77p_4StrictEscherichia coli EF-Tu mutants (R234F)protein variantelfamycin-resistant EF-Tuelfamycin antibioticantibiotic target alteration98.596.3
INSAq77p_82StrictbaeRprotein homologresistance-nodulation-cell division (RND) antibiotic efflux pumpaminoglycoside antibiotic, aminocoumarin antibioticantibiotic efflux95.8100.0
INSAq77p_37StrictH-NSprotein homologmajor facilitator superfamily (MFS) antibiotic efflux pump, resistance-nodulation-cell division (RND) antibiotic efflux pumpmacrolide antibiotic, fluoroquinolone antibiotic, cephalosporin, cephamycin, penam, tetracycline antibioticantibiotic efflux95.6100.0
INSAq77p_47StrictmsbAprotein homologATP-binding cassette (ABC) antibiotic efflux pumpnitroimidazole antibioticantibiotic efflux94.7100.0
INSAq77p_3StrictemrRprotein homologmajor facilitator superfamily (MFS) antibiotic efflux pumpfluoroquinolone antibioticantibiotic efflux94.3100.0
INSAq77p_25StrictEscherichia coli UhpT mutant (E350Q)protein variantantibiotic-resistant UhpTfosfomycinantibiotic target alteration93.7100.0
INSAq77p_21StrictmarAprotein homologresistance-nodulation-cell division (RND) antibiotic efflux pump, General Bacterial Porin with reduced permeability to beta-lactamsfluoroquinolone antibiotic, monobactam, carbapenem, cephalosporin, glycylcycline, cephamycin, penam, tetracycline antibiotic, rifamycin antibiotic, phenicol antibiotic, triclosan, penemantibiotic efflux, reduced permeability to antibiotic93.699.2
INSAq77p_3_7StrictKlebsiella pneumoniae KpnHprotein homologmajor facilitator superfamily (MFS) antibiotic efflux pumpmacrolide antibiotic, fluoroquinolone antibiotic, aminoglycoside antibiotic, carbapenem, cephalosporin, penam, peptide antibiotic, penemantibiotic efflux92.2100.6
INSAq77p_11StrictoqxAprotein homologresistance-nodulation-cell division (RND) antibiotic efflux pumpfluoroquinolone antibiotic, glycylcycline, tetracycline antibiotic, diaminopyrimidine antibiotic, nitrofuran antibioticantibiotic efflux91.1100.0
INSAq77p_21StrictEscherichia coli marR mutant conferring antibiotic resistanceprotein overexpressionresistance-nodulation-cell division (RND) antibiotic efflux pumpfluoroquinolone antibiotic, cephalosporin, glycylcycline, penam, tetracycline antibiotic, rifamycin antibiotic, phenicol antibiotic, triclosanantibiotic target alteration, antibiotic efflux91.0100.0
INSAq77p_34StrictKlebsiella pneumoniae KpnFprotein homologmajor facilitator superfamily (MFS) antibiotic efflux pumpmacrolide antibiotic, aminoglycoside antibiotic, cephalosporin, tetracycline antibiotic, peptide antibiotic, rifamycin antibioticantibiotic efflux89.0100.0
INSAq77p_3StrictrsmAprotein homologresistance-nodulation-cell division (RND) antibiotic efflux pumpfluoroquinolone antibiotic, diaminopyrimidine antibiotic, phenicol antibioticantibiotic efflux85.3100.0
INSAq77p_13StrictEscherichia coli ampH beta-lactamaseprotein homologampC-type beta-lactamasecephalosporin, penamantibiotic inactivation85.2100.8
INSAq77p_34StrictKlebsiella pneumoniae KpnEprotein homologmajor facilitator superfamily (MFS) antibiotic efflux pumpmacrolide antibiotic, aminoglycoside antibiotic, cephalosporin, tetracycline antibiotic, peptide antibiotic, rifamycin antibioticantibiotic efflux82.083.3
INSAq77p_11StrictadeFprotein homologresistance-nodulation-cell division (RND) antibiotic efflux pumpfluoroquinolone antibiotic, tetracycline antibioticantibiotic efflux60.999.2
INSAq77p_1StrictHaemophilus influenzae PBP3 mutant (D350N, S357N)protein variantPenicillin-binding protein mutations conferring resistance to beta-lactam antibioticscephalosporin, cephamycin, penamantibiotic target alteration53.196.4
INSAq77p_9StrictadeFprotein homologresistance-nodulation-cell division (RND) antibiotic efflux pumpfluoroquinolone antibiotic, tetracycline antibioticantibiotic efflux41.297.9
Table 2. Comparison of the INSAq77 mcr-9-containing contigs with the top nine IncHI2 mcr-9-harboring plasmids showing the highest identities (>99.0%, E-value 0.0, query coverage >94.0%).
Table 2. Comparison of the INSAq77 mcr-9-containing contigs with the top nine IncHI2 mcr-9-harboring plasmids showing the highest identities (>99.0%, E-value 0.0, query coverage >94.0%).
Plasmid
(bp)
StrainIsolation Source/Country/YearIdentity
(%)
Query Cover (%)pMLST bAcquired Antibiotic and Desinfectant Resistance Genes cGenBank Acc. No.
INSAq77 IncHI2
(30,314) a
E. ludwigii INSAq77Seabream (Sparus aurata)/Portugal/2018--DLST1mcr-9.1JABRPH000000000
pSPRC-Echo1
(339,920)
E. hormaechei C15117Burns unit/Australia/200799.99%99.0%DLST1aac(6’)-IIc, aph(3’’)-Ib-type, aph(6)-Id, blaSHV-12, blaTEM-1B, catA2-type, dfrA19, mcr-9, qacE, qnrA1-type, sul1, sul2, tet(D)NZ_CP032842
p525011-HI2
(354,045)
C. freundii 525011unknown/China/2017100.00%97.0%untyped,
Nearest STs: 7,1,4,15
aac(3)-IId-type, aac(6’)-aph(2’’), aadA5, armA, blaTEM-1B, catA2-type, dfrA1-type, mcr-9, mph(E), msr(E), qacE, qnrA1-type, sul1, sul2NZ_MF344582
pOSUKPC4
(351,806)
E. hormaechei OSUKPC4_L Animal/USA/2016100.00%98.0%DLST1aadA1, aph(3’’)-Ib-type, aph(6)-Id, blaKPC-4, blaOXA-129, dfrA21, mcr-9, qacE, sul1, tet(B)NZ_CP024910
pOSUEC_D
(354,256)
E. hormaechei OSUVMCKPC4-2 Animal/USA/2016100.00%98.0%DLST1aadA1, aph(3’’)-Ib-type, aph(6)-Id, blaKPC-4, blaOXA-129, dfrA21, mcr-9, qacE, sul1, tet(B)NZ_CP029248
pK29
(269,674)
K. pneumoniae NK29Human/Taiwan/2001100.00%98.0%DLST1aadA2, blaCMY-8, blaCTX-M-62-type, catB2, mcr-9, qacE, sul1NC_010870
pE1_001
(357,530)
L. adecarboxylata E1Burns Unit Shower/Australia/2012100.00%98.0%DLST1formA-type, mcr-9NZ_CP042506
pE11_001
(339,433)
C. freundii E11Burns Unit Shower/Australia/2012100.00%98.0%DLST1formA-type, mcr-9NZ_CP042525
pE61_001
(357,530)
L. adecarboxylata E61Burns Unit Shower/Australia/2014100.00%98.0%DLST1formA-type, mcr-9NZ_CP042494
p565_1
(263,189)
C. freundii 565Human Stool/Spain/201499,68%95,0%DLST1aac(6’)-Ib-cr, aadA1, aadA2b-type, blaCTX-M-9, blaSHV-12, blaVIM-1, catA1-type, dfrA16, mcr-9-type, qacE, qnrA1-type, sul1NZ_CP038657
a Length of the mcr-9-containing contig. b Plasmid pMLST-2.0 Server c ResFinder-4.1 (Selected %ID threshold: 90%; selected minimum length: 60%).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Manageiro, V.; Salgueiro, V.; Rosado, T.; Bandarra, N.M.; Ferreira, E.; Smith, T.; Dias, E.; Caniça, M. Genomic Analysis of a mcr-9.1-Harbouring IncHI2-ST1 Plasmid from Enterobacter ludwigii Isolated in Fish Farming. Antibiotics 2022, 11, 1232. https://doi.org/10.3390/antibiotics11091232

AMA Style

Manageiro V, Salgueiro V, Rosado T, Bandarra NM, Ferreira E, Smith T, Dias E, Caniça M. Genomic Analysis of a mcr-9.1-Harbouring IncHI2-ST1 Plasmid from Enterobacter ludwigii Isolated in Fish Farming. Antibiotics. 2022; 11(9):1232. https://doi.org/10.3390/antibiotics11091232

Chicago/Turabian Style

Manageiro, Vera, Vanessa Salgueiro, Tânia Rosado, Narcisa M. Bandarra, Eugénia Ferreira, Terry Smith, Elsa Dias, and Manuela Caniça. 2022. "Genomic Analysis of a mcr-9.1-Harbouring IncHI2-ST1 Plasmid from Enterobacter ludwigii Isolated in Fish Farming" Antibiotics 11, no. 9: 1232. https://doi.org/10.3390/antibiotics11091232

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