Skip to main content

Molecular characteristics of antibiotic-resistant Escherichia coli and Klebsiella pneumoniae strains isolated from hospitalized patients in Tehran, Iran

Abstract

Background

We evaluated the distribution of carbapenem and colistin resistance mechanisms of clinical E. coli and K. pneumoniae isolates from Iran.

Methods

165 non-duplicate non-consecutive isolates of K. pneumoniae and E. coli were collected from hospitalized patients admitted to Iran's tertiary care hospitals from September 2016 to August 2018. The isolates were cultured from different clinical specimens, including wound, urine, blood, and tracheal aspirates. Antibiotic susceptibility testing was performed by disc diffusion and microdilution method according to the Clinical and Laboratory Standards Institute (CLSI) guideline. The presence of extended spectrum β-lactamases (ESBLs) genes, carbapenemase genes, as well as fosfomycin resistance genes, and colistin resistance genes was also examined by PCR-sequencing. The ability of biofilm formation was assessed with crystal violet staining method. The expression of colistin resistance genes were measured by quantitative reverse transcription-PCR (RT-qPCR) analysis to evaluate the association between gene upregulation and colistin resistance. Genotyping was performed using the multi-locus sequencing typing (MLST).

Results

Colistin and tigecycline were the most effective antimicrobial agents with 90.3% and 82.4% susceptibility. Notably, 16 (9.7%) isolates showed resistance to colistin. Overall, 33 (20%), 31 (18.8%), and 95 (57.6%) isolates were categorized as strong, moderate, and weak biofilm-producer, respectively. Additionally, blaTEM, blaSHV, blaCTX-M, blaNDM-1, blaOXA-48-like and blaNDM-6 resistance genes were detected in 98 (59.4%), 54 (32.7%), 77 (46.7%), 3 (1.8%), 17 (10.30%) and 3 (1.8%) isolates, respectively. Inactivation of mgrB gene due to nonsense mutations and insertion of IS elements was observed in 6 colistin resistant isolates. Colistin resistance was found to be linked to upregulation of pmrA-C, pmrK, phoP, and phoQ genes. Three of blaNDM-1 and 3 of blaNDM-6 variants were found to be carried by IncL/M and IncF plasmid, respectively. MLST revealed that blaNDM positive isolates were clonally related and belonged to three distinct clonal complexes, including ST147, ST15 and ST3299.

Conclusions

The large-scale surveillance and effective infection control measures are also urgently needed to prevent the outbreak of diverse carbapenem- and colistin-resistant isolates in the future.

Background

Enterobacteriaceae are opportunistic pathogens that cause severe nosocomial infections, including urinary tract infections (UTIs), bloodstream infections, abdominal infections, and ventilator-associated pneumonia [1, 2]. Escherichia coli and Klebsiella pneumoniae are two important members of Enterobacteriaceae that have the ability to develop resistance to various classes of antibiotics. Nowadays, carbapenem antibiotics are recommended as the last-line therapy for MDR strains of K. pneumoniae and E. coli infections [1, 3]. However, increasing rate of resistance to carbapenems has complicated the treatment process and led to untreatable hospital infections [1, 4]. Resistance to carbapenems in Enterobacteriaceae is mainly mediated by the production of carbapenem-hydrolyzing enzymes (carbapenemases), among which Klebsiella pneumoniae carbapenemase (KPC), metallo-β-lactamases (VIM, IMP, NDM), and OXA-48 type of enzymes are the most common. Mobile genetic elements, including plasmids, transposons, and integrons are involved in the dissemination of related encoding genes [5,6,7].

New Delhi metallo-β-lactamase-1 (NDM-1) is one of the most important type of carbapenemases in carbapenem-resistant Enterobacteriaceae (CRE) [8, 9]. The blaNDM-positive strains are usually resistant to most antimicrobial agents in addition to β-lactams due to the co-existence of other resistance mechanisms [10]. Such resistant strains have known as the leading cause of infections associated with high mortality worldwide, representing a significant challenge for clinical management and public health [11]. Under these conditions, clinicians rely on a few alternative antibiotics e.g., colistin, fosfomycin, and tigecycline to treat infections caused by CRE [1, 12].

The old polymyxin antibiotic colistin (i.e., polymyxin E) is now recommended as the last choice for treatment of MDR Gram-negative bacteria, especially CRE infections [13]. The recent increase in the use of colistin in clinical practice, accompanied by its unbridled use in agriculture, have contributed to the rapid dissemination of resistance [14]. Colistin resistance is caused by decreases in the net negative charge of the outer membrane, loss of lipid A, or efflux pumps and plasmid-encoded mcr genes [15]. The mcr-1 gene uses a target site modification mechanism to protect bacteria from the action of colistin. The mcr gene is observed on transferable plasmid and encodes an enzyme called phosphatidylethanolamine transferase which transfers the phosphatidylethanolamine residue to lipid A [16].

The main purpose of this study was to evaluate the antimicrobial resistance patterns and molecular mechanisms of carbapenem and colistin resistance among the clinical isolates of E. coli and K. pneumoniae from hospitalized patients admitted to tertiary care hospitals in Tehran, Ahwaz, Kashan, Tabriz, Sari, Gorgan, Birjand and Babol. In addition, the ability of biofilm production as well as clonal and genetic diversity of isolates were examined.

Methods

Ethical statement

This study was approved by the Ethics Committee of Shahid Beheshti University of Medical Sciences “IR.SBMU.MSP.REC.1397. 629”. In order to maintain patients confidentiality participants were anonymous and no personal information was collected or included in the study.

Bacterial isolates

K. pneumoniae and E. coli isolates were collected from hospitalized patients infected in Iran hospitals from September 2016 to August 2018. The isolates were cultured from different clinical specimens, including wound, urine, blood, and tracheal aspirates. Each isolate was identified at species level based on the biochemical reactions, including reaction on SH2/indole/motility (SIM) medium, triple sugar iron (TSI) agar, urease production on urea agar, growth on Simmons'citrate agar medium, methyl red/Vogues-Proskauer (MR/VP), and ornithine decarboxylase (OD) test [17]. All isolates were stored in tryptic soy broth (TSB) tube with 20% glycerol at − 70 °C.

Antimicrobial susceptibility testing

Antimicrobial susceptibility of all E. coli and K. pneumoniae isolates was determined by the Kirby-Bauer disk diffusion method on Cation-Adjusted Mueller Hinton agar (Merck, Germany) and interpreted as recommended by the Clinical and Laboratory Standards Institute (2018 CLSI breakpoints) or Food and Drug Administration (FDA) breakpoints guidelines (for tigecycline) [18, 19]. Antibiotic discs used were as follow: penicillins [piperacillin (PIP, 100 μg)], β-lactam/β-lactamase inhibitor combinations [piperacillin/tazobactam (PTZ, 100/10 μg)], cephems [ceftazidime (CAZ, 30 μg), cefotaxime (CTX, 30 μg), cefepime (FEP, 30 μg), cefpodoxime (CPD, 30 μg)], monobactams [aztreonam (ATM, 30 μg)], carbapenems [imipenem (IPM, 10 μg), meropenem (MEM, 10 μg), ertapenem (ETP, 10 μg), doripenem (DOR, 10 μg)], aminoglicosides [gentamicin(GEN,10 μg)], Amikacin (AK, 30 μg)], Fluoroquinolones [ciprofloxacin (CIP, 5 μg)], inhibitors [trimethoprim-sulfamethoxazole (TS, 2.5 μg)], fosfomycins [fosfomycin/trometamol (FOT, 200 μg)], tigecycline (TGC, 15 μg), and nalidixic acid (NA, 30 μg), (Mast Group, Merseyside, UK). The minimum inhibitory concentrations (MICs) of seven antibiotics, including imipenem, meropenem, ceftazidime, cefotaxime, cefepime, ciprofloxacin, and colistin were determined by broth microdilution method on Cation-Adjusted Mueller Hinton broth (Merck, Germany), and the results were analyzed according to the CLSI guidelines [18]. The 2016 EUCAST breakpoints were used (available at http://www.eucast.org/clinical_breakpoints/) for colistin. The antibiotic powders were purchased from Sigma-Aldrich (St. Louis, MO, USA). E. coli ATCC 25922 was used as a quality control strain for disk diffusion and MIC results.

The CDC and the European Centre for Disease Prevention and Control (ECDC) have jointly developed definitions for multidrug-resistant (MDR), extensively drug-resistant (XDR) and pandrug-resistant (PDR) bacteria. MDR was defined as acquired non-susceptibility to at least one agent in three or more antimicrobial categories, XDR was defined as non-susceptibility to at least one agent in all but two or fewer antimicrobial categories and PDR was defined as non-susceptibility to all agents in all antimicrobial categories.

Phenotypic detection of β-lactamases

Detection of ESBLs was tested for all the isolates by combination disc diffusion test (CDDT) containing ceftazidime (CAZ) and cefotaxime (CTX) with CAZ 30 μg + clavulanic acid 10 μg and CTX 30 μg + clavulanic acid 10 μg per disc (Mast Group, Merseyside, UK). K. pneumoniae ATCC 700,603 and E. coli ATCC 25,922 were used as positive and negative controls for ESBL production, respectively [22].

Phenotypic detection of metallo β-lactamases

Combined disk diffusion test (CDDT) was performed for identification of MBLs by imipenem and meropenem (Mast Group, Merseyside, UK) alone and in combination with EDTA [20]. Pseudomonas aeruginosa ATCC 27853 and P. aeruginosa PA40 (Accession number: KM359725) were used as negative and positive controls for MBL production, respectively.

Screening for carbapenemase production

The Carba NP test was performed for the detection of carbapenemase activity in isolates as described previously [21, 22].

Biofilm formation assay

Assessment of biofilm formation was performed by the colorimetric microtiter plate assay in triplicates [20, 21]. Overnight cultures of bacterial isolates were suspended in tryptic soy broth (TSB) (Merck-Germany) at 37 °C. Then, 200 μL bacterial suspension with turbidity of 0.5 McFarland standard were transferred into the sterile 96-well polystyrene microplates (JET Biofil, Guangzhou, China). TSB without bacteria was used as negative control. After 24 h of incubation at 37 °C, each well was rinsed three times with phosphate buffered saline (PBS, pH 7.3) to remove any non-adherent cells. Fixation and staining the adherent cells were performed by methanol and 1% crystal violet (Merck, Germany). Then, plates were gently rinsed off with PBS and destained by 33% glacial acetic acid and finally OD of each well were measured at 492 nm. The criteria for categorization of isolates were including: strong biofilm producer (4 × ODc < OD), moderate biofilm producer (2 × ODc < OD < 4 × ODc), weak biofilm producer (ODc < OD < 2 × ODc) and no biofilm producer (OD < ODc) [23, 24].

Detection of resistance genes

DNA was extracted using the DNA extraction kit (High Pure PCR Template Preparation Kit-Roche, Germany, Lot. No. 10362400) according to the manufacturer's instruction. Detection of resistance genes among all isolates, including ESBL-encoding genes (blaTEMblaSHV, blaCTX-M, blaGES, blaPER, and blaVEB), carbapenemases genes (blaOXA-48, blaNDM, blaKPCblaVIM, and blaIMP), and two fosfomycin resistance genes (fosA and fosC2), was performed by polymerase chain reaction (PCR) amplification using the specific primers [25,26,27,28,29] and confirmed by sequencing. P. aeruginosa containing blaGES, blaPER, blaVEB, blaVIMblaIMP genes and K. pneumoniae containing other genes received from Shahid Beheshti University of Medical Sciences, Tehran, Iran, were used as positive controls. PCR products were purified using a PCR purification Kit (Bioneer Co., Korea) and then, nucleotide sequencing of amplicons was performed by an ABI PRISM 3700 sequencer (Macrogen Co., Korea). Nucleotide sequences were analyzed using Chromas software version 1.45 (http://www.technelysium.com.au) and NCBI BLAST program (https://blast.ncbi.nlm.nih.gov/Blast.cgi).

Molecular analysis of colistin resistance

Analysis of plasmid-mediated colistin resistance was performed by PCR amplification of mcr-1, mcr-2, mcr-3, and mcr-4 among all colistin-resistant K. pneumoniae isolates. All colistin-resistant K. pneumoniae isolates were also examined for the presence of mutations in the chromosomally-encoded modifications of the LPS, including mgrB, pmrA, pmrB, phoP, and phoQ genes [30, 31]. Insertion sequences (ISs) were identified using the IS finder tool (https://www-is.biotoul.fr/index.php). Genomic DNA from two colistin-sensitive K. pneumoniae clinical isolates and K. pneumoniae ATCC 700603 were used as control.

Real-time quantitative reverse transcription PCR

Colistin-resistant isolates were assessed for expression of pmrC, pmrA, pmrB, pmrD, pmrE, and pmrK genes using specific primers [29, 31, 32]. rpsL gene encoding a ribosomal protein was used as housekeeping gene to normalize the levels of transcripts tested. Total RNA was extracted from the cultures grown in the mid-log phase of growth in Luria–Bertani broth (Merck, Darmstadt, Germany) by the RNX-Plus Kit (Cat. No., RN7713C, Sinaclon, Iran) according to the manufacturer’s instruction. The contaminating DNA was removed by RNase-free DNase I (Fermentas, Thermo Fisher Scientific Inc., USA). The total RNA concentration was determined by Nanodrop (WPA Biowave II Nanospectrophotometer, USA). DNase-treated RNA was reverse-transcribed into cDNA using the Takara Kit (Japan). RNA samples were checked for contaminating DNA by PCR. Real-time PCR assay was performed on synthesized cDNA using the Power SYBR Green PCR Master Mix (Bioneer, Korea) on a Corbett Rotor-Gene 6000 real-time rotary analyzer (Corbett Life Science, Australia). Each amplification protocol included a first denaturation step of 10 min at 94 °C, followed by 40 cycles of 20 s at 94 °C and 45 s at 59 °C. All samples were run in triplicate. Data were compared to those obtained with the rpsL gene. The expression level of transcripts was calculated based on 2−ΔΔCT method (relative) against that for the susceptible isolate, K. pneumoniae ATCC 700603. Experiments were repeated three times. The parameter Ct was defined as the threshold cycle number at which the first detectable fluorescence generated by the binding of SYBR Green I dye to the minor groove of double-stranded DNA began to increase exponentially.

Plasmid manipulation and analysis

NDM positive strains were selected for plasmid analysis. Plasmid DNA of isolates, transconjugants, and transformants was extracted by using the Roche kit (Cat. No. 11 754 777 001) according to the manufacturer’s instructions. Electroporation was used to transform plasmids encoding blaNDM into E. coli TOP10. The blaNDM transformants were selected on MH agar (Merck-Germany) supplemented with meropenem (0.5 mg/L) (Sigma–Aldrich). Conjugation experiments were carried out in LB broth with sodium-azide-resistant E. coli J53AzR as the recipient. Cultures of donor and recipient cells in logarithmic phase were added to 4 mL of fresh LB broth and were then incubated at 37 °C overnight without shaking. The transconjugants were selected on MH agar (Merck-Germany) supplemented with meropenem (0.5 mg/L) or ceftazidime (1, 2 and 4 mg/L) with sodium azide (100 mg/L) (Sigma–Aldrich).

PCR-based replicon typing

All transconjugants and transformants were typed by a PCR method based on replicons of the major plasmid incompatibility groups among Enterobacteriaceae [33].

Multi-locus sequence type (MLST) analysis

Genotyping by MLST analysis was conducted to characterize diversity and epidemiology of blaNDM- carrying K. pneumoniae isolates [34]. Briefly, PCR for seven housekeeping genes, including rpoB, gapA, mdh, phoE, pgi, infB, and tonB was carried out. Results were analyzed according to the Institute Pasteur Klebsiella MLST database (https://bigsdb.pasteur.fr/klebsiella/klebsiella.html). Unique sequence (allele) number for each gene was assigned on the basis of the information in the K. pneumoniae MLST database to determine specific sequence types (ST). A combination of the allelic sequences of the seven genes yielded the allelic profile for each isolate.

Repetitive extragenic palindromic (rep)-PCR typing

Rep-PCR analyses were conducted with the single primer BoxA1R (5′-CTA CGG CAA GGC GAC GCT GAC G-3′) [35]. To determine phylogenetic relationships, rep-PCR profiles were analyzed by GelCompar II software (Applied Maths, Belgium) using the Pearson’s correlation coefficient with unweighted paired group method using arithmetic averages (UPGMA) as well as at the 80% similarity level [35].

Statistical analysis

Chi-squared test was performed using SPSS software, 21.0 (SPSS Inc., Chicago, IL, USA) to check for any significant differences between datasets. A significant level of P ≤ 0.05 was considered statistically significant.

Results

Bacterial isolates

165 non-duplicate non-consecutive isolates of E. coli and K. pneumoniae were collected from 73 (45.5%) females and 92 (54.5%) males admitted at five Iranian hospitals during the September 2016 to August 2018. The age range of patients was between 1 and 87 years. The origins of isolates were 114 in urine, 39 in tracheal aspirates, 4 in wounds, and 8 in blood.

Antimicrobial susceptibility

Antibiotic resistance patterns of 165 isolates of K. pneumoniae and E. coli are shown in Table 1. The lowest rate of resistance was observed against tigecycline (n = 9, 5.5%), and fosfomycin (n = 26, 15.8%). The number of isolates with multidrug-resistant (MDR), extensively drug-resistant (XDR), and pandrug-resistant (PDR) phenotype was 32 (E. coli: 27, K. pneumoniae: 5), 120 (E. coli: 77, K. pneumoniae: 43), 1 (K. pneumoniae: 1), respectively. The MIC ranges, MIC50, MIC90, and the percentages of isolates resistant, intermediate, or susceptible isolates to the seven antimicrobial agents are shown in Table 2.

Table 1 Antibiotic resistance patterns of 165 isolates of K. pneumoniae and E. coli
Table 2 MIC of the K. pneumoniae and E. coli clinical isolates (n = 165)

K54 was found to be non-susceptible to all antibiotics tested, which includes cephalosporins, penicillins, carbapenems, aztreonam, aminoglycosides, ciprofloxacin, colistin, tetracyclines, tigecycline, trimethoprim-sulfamethoxazole and fosfomycin (Table 3). Thus, the isolate can truly be described as pandrug-resistant.

Table 3 MIC and molecular features related to NDM-producing and colistin-resistant K. pneumoniae isolates

β-lactamase phenotype

The prevalence of ESBL-producing E. coli and K. pneumoniae was 49.6% (n = 82) and 26.6% (n = 44), respectively. The proportion of ESBL-producing E.coli and K. pneumoniae showing resistance to cephalosporin were significantly higher than non-ESBL-producing strains (p < 0.05).

Metallo β-lactamase phenotype

The prevalence of MBL-producing E. coli and K. pneumoniae were 1.8% (n = 2) and 38.5% (n = 20), respectively. All MBL-producing isolates were resistant to carbapenems and cephalosporins (P ≤ 0.05).

Carbapenemase phenotype

According to the results of the Carba NP test, only 22 K. pneumoniae isolates produced carbapenemase enzymes. As with the MBL phenotypes, all carbapenemase-producing isolates were resistant to carbapenem and cephalosporin antibiotics (p ≤ 0.05).

Biofilm phenotype

Biofilm phenotype accounted for 159 out of 165 isolates (96.36%): 33 isolates (20%) produced strong biofilm, 31 isolates (18.8%) produced moderate biofilm, and 95 isolates (57.6%) produced weak biofilm; whereas 6 isolates (3.6%) did not form biofilm. Among 82 ESBL-producing E. coli, 12 (14.63%) isolates were strong biofilm-producers, 11(13.41%) were moderate biofilm-producers, 55 (67%) were weak biofilm-producers, and 4 (4.88%) isolates produced no biofilm. Moreover, among the 44 ESBL-producing K. pneumoniae, 16 (36.36%) isolates were strong biofilm-producers, 12 (27.27%) were moderate biofilm-producers, and 16 (36.36%) isolates were identified as weak biofilm-producers.

Antimicrobial resistance genes

The prevalence of isolates carrying ESBL-encoding determinants was 78.2% (n = 129). The blaTEM, blaSHV, and blaCTX-M genes were detected in 98 (59.4%), 54 (32.7%), and 77 (46.7%) isolates, respectively; while no isolates were positive for the blaGES, blaPER, and blaVEB genes (Table 4). In addition, the prevalence of MBL-producing E. coli and K. pneumoniae were 1.8% (n = 2) and 38.5% (n = 20), respectively, of which 6 (6.5%) K. pneumoniae isolates were positive for for blaNDM gene (blaNDM-6: 3, blaNDM-1: 3) (Table 4). No blaIMP, blaVIM, blaSIM, blaGIM, blaSPM, and blaKPC genes were detected. The blaOXA-48-like gene was identified among 17 (10.30%) of isolates. While no plasmid-mediated colistin resistance genes of mcr-1, mcr-2, mcr-3, mcr-4, and mcr-4 were detected in isolates, 16 (9.7%) K. pneumoniae were identified as colistin-resistant. Moreover, the primers targeting fosA and fosC2 genes did not provide any amplicon in fosfomycin-resistant isolates. The results from real-time PCR analysis were consistent with PCR and sequencing.

Table 4 Prevalence of beta-lactamase genes among isolates

Molecular analysis of colistin resistance

The mcr-1, mcr-2, mcr-3, and mcr-4 genes were not found in any of the colistin-resistant isolates, we focused on other mechanisms of resistance, specifically mgrB gene inactivation and the presence of the mutations in the pmrA, pmrB, phoP, and phoQ genes. Sequence analysis of the mgrB gene showed that one isolate (K37) generated amplicon that was larger than those produced by K. pneumoniae K85 control isolate and colistin-susceptible K. pneumoniae ATCC 700603 strain. Amplicon sequencing revealed that insertional inactivation had occurred in the coding region of the K. pneumoniae K37 mgrB gene. Also, occurred at nucleotide 75 and was raised by insertional sequence that shared 99% identity at the nucleotide level with IS5 family of insertion sequences (Fig. 1). Insertional inactivation was not detected in other isolates tested. However, K83, K101, K50, and K130 isolates had premature amber stop codon (TAG) due to a C-to-T change at position 88 and K136 had premature opal stop codon (TGA) due to a C-to-A change at position 117, resulting in a truncated MgrB protein containing 29 and 39 amino acids, respectively. Amino acid substitutions were detected in PmrB, PhoP and PhoQ proteins. Nucleotide A at the position 469 of the pmrB gene was converted to C in K101 isolate, leading to Thr157Pro substitution. At nucleotide position of 171, the phoP gene underwent A to C conversion, resulting in single substitution Glu57Asp in the isolate K37. The isolate K83 showed nucleotide conversion A to G at the position 449 of phoQ gene, leading to substitution Asp150Gly. No amino acid substitutions were detected in PmrA protein.

Fig. 1
figure 1

Schematic representation of the different insertion events identified in the mgrB gene. a The intact mgrB gene as found in wild type isolates and isolate (b) mgrB truncated by IS5-like in k37 isolate. c mgrB truncated by IS5-like as identified by Laurent Poirel et al.[65]

Overexpression of pmrCAB, pmrHFIJKLM, and phoPQ operons

Expression level of pmr and pho genes was measured to evaluate the effect of mutations on colistin-resistant isolates. Results revealed increased expression level of 1.2–8.6 fold for pmrA, 1.57–5.09 fold for pmrB, 0.93–8.8 fold for pmrC, 2.17–17 fold for pmrK, 2.35–15.02 fold for phoP, and 2.13–9.28 fold for phoQ genes; whereas no differences in expression levels were observed for pmrD and pmrE genes (Fig. 2a). Analysis of mRNA transcript in K37 isolate with an inactivated mgrB gene revealed a significant increase in expression level of genes pmrA (8.6-fold), pmrB (5.2-fold), pmrC (7.3-fold), pmrK (17.1-fold), phoP (14.5-fold), and phoQ (9.3-fold). No insertional inactivation of mgrB gene was found in K83 and K101 isolates. Also, features of the colistin-resistant isolates has been showed in Table 5. Relative expression levels of genes in PDR strain shown in Fig. 2b.

Fig. 2 a
figure 2

Relative expression levels of the pmrA, pmrB, pmrC, pmrE, pmrD, pmrK, phoP and phoQ genes in colistin-resistant K. pneumoniae isolates. No differences in expression levels were observed for pmrD and pmrE genes. ATCC: K. pneumoniae ATCC 700603. b Relative expression levels of the pmrA, pmrB, pmrC, pmrK, phoP and phoQ genes in PDR strain (K54). ATCC: K. pneumoniae ATCC 700603

Table 5 Features of the colistin-resistant isolates

Transformation and conjugation assays

Plasmids carrying blaNDM-1 and blaNDM-6 genes in all six strains were successfully transferred to E. coli TOPO10 and E. coli J53 recipient strains. The antimicrobial resistance profile of the transformants and transconjugants are shown in Table 6. PCR confirmed the presence of the blaNDM-1 and blaNDM-6 genes in the transformants and transconjugants; all these isolates harbored also blaCTX-M, blaTEM and blaSHV genes (Table 6).

Table 6 The features related to NDM-producing K. pneumoniae isolates in Iran

Plasmid replicon typing

Plasmid replicon typing revealed that 3 blaNDM-1-carrying- and 3 blaNDM-6-carrying K. pneumoniae isolates contained plasmid types belonging to IncF and IncL/M, respectively (Table 6).

MLST analysis results

STs were identified among the 6 blaNDM-carrying K. pneumoniae isolates, including ST147 (n = 4), ST15 (n = 1), and ST3299 (n = 1). Among the isolates that belonged to ST147, 3 isolates were originated from urine specimens (Table 6).

Rep-PCR analysis

To evaluate the genetic diversity, 6 blaNDM-positive and 16 colistin-resistant isolates were subjected to rep-PCR fingerprinting. Isolates were divided into 3 common types (CT) containing 2–4 isolates and 12 single types (ST). Among these, a dominant clone was from Tehran and originated from urine samples. The genotypic pattern of the dominant clone revealed that all isolates harbored ESBL genes.

Discussion

The excessive and inappropriate use of antibiotics against microbial infections in Iran has led to increased rate of drug resistance in recent decades [36]. Today, clinicians rely increasingly on carbapenems (i.e., imipenem, meropenem, doripenem, etc.) to treat infections due to multidrug-resistant bacteria. CRE strains have been reported in several hospital outbreaks and have the propensity to spread rapidly at local, regional and international levels. The continual emergence of CREs is a major threat to public health worldwide [1]. The worsening condition is that CRE strains show resistance progressively toward a wide range of antimicrobial classes [36, 37] [38]. In this study, about 73.1% of K. pneumoniae and 28.3% of E. coli isolates were resistant to at least one of the carbapenems tested. Among the included isolates, the highest rates of resistance belonged to piperacillin (n = 161, 97.6%), nalidixic acid (n = 154, 93.3%), and cefotaxime (n = 153, 92.7%). On the other hand, the lowest resistance rate was observed for tigecycline (n = 9, 5.5%) followed by colistin (n = 16, 9.7%), and fosfomycin (n = 26, 15.8%), indicating that these antibiotics have increasingly become primary options for treatment of multi-resistant strains of K. pneumoniae and E. coli. Our results indicated that the resistance rate of K. pneumoniae isolates against colistin was 30.77% with the range MIC 4–128 μg/mL. Colistin remains the last line of defense against many Gram-negative bacilli. However, colistin-resistant and even pan-drug-resistant Gram-negative bacilli have already been reported [39]. According to reports from other studies around the world, the rate of colistin resistance among carbapenem-resistant K. pneumoniae has progressively increased from < 2% to 9%. In the last decade in Europe, resistance to colistin has increased to one third of carbapenem-resistant isolates. In addition, multiple outbreaks of colistin-resistant K. pneumoniae have been reported in different regions of the world [40, 41].

In this study, the prevalence of ESBL-producing E. coli and K. pneumoniae were 49.6% and 26.6%, respectively. To date, the ESBL and MBL enzymes has been identified in almost all of the world, including many countries in Asia, Africa, Americas, the Europe, and Australia [42, 43]. The high rate of ESBL and MBL prevalence in the world and its widespread dissemination is a cause of concern. The blaNDM are plasmid-mediated genes responsible for resistance to carbapenems and are often co-harbored with different resistance determinants, such as those encoding ESBL. In this study, 98 (59.4%), 54 (32.7%), 77 (46.7%), 3(1.8%) and 3(1.8%) isolates harbored blaTEM, blaSHV, blaCTX-M, blaNDM-1 and blaNDM-6 β-lactamase genes, respectively. All three K. pneumoniae isolates carrying blaNDM-6 and one isolate harboring blaNDM-1 belonged to the ST147 clone. While each of the two remaining isolates that were positive for blaNDM-1 belonged separately to the ST15 and ST3299 clone. The blaNDM-6-producing E. coli and K. pneumoniae have been reported in New Zealand (ST101) [26] and India [44]. Distribution of blaNDM-1 is greater than that of blaNDM-6 and was reported from most regions of the world [45, 46].

Plasmids are elements that spread easily. This is one of the most difficult challenges to counteract the dissemination of antibiotic resistance genes and nosocomial infections. Analysis of transformants and transconjugants in the current study revealed that the blaNDM-6 gene along with blaCTX-M-15, blaSHV, and blaTEM were carried on transferable plasmids belonging to the IncL/M, while blaNDM-1 gene was carried on transferable plasmids belonging to the IncF along with blaCTX-M-15, blaSHV, and blaTEM. Previous studies have reported that the spread of blaNDM-1 is linked to different types of IncA/C, IncF, IncN, and untypeable plasmids [47]. Transferable IncL/M and IncF plasmids have greatly contributed to the dissemination of antibiotic resistance genes, such as blaNDM-6, blaNDM-1, blaTEM, blaSHV as well as blaCTX-M-15 among enterobacterial species [20, 48]. Other study reported that IncL/M and IncF plasmids have the ability to transfer to the susceptible strain, contributing to dissemination of antibiotic resistance genes, such as blaNDM-1 and blaCTX-M-15 among K. pneumoniae [48, 49]. The three K. pneumoniae isolates carrying blaNDM-6 belonged to ST147, suggesting the possibility of nosocomial infection. ST147 is among the major successful K. pneumoniae clone and, usually, is linked to IncF plasmids with blaKPC [50].

Colistin is a last-resort antibiotic that has been reintroduced today in clinical practices to treat infections caused by MDR CREs [13]. Acquired resistance to colistin is mostly caused by chromosomal mutations. However, a new plasmid-mediated colistin resistance gene, mcr-1, encoding a phosphoethanolamine transferase, has recently been described in China [51]. In our study, plasmid encoded mcr-1, mcr-2, mcr-3, and mcr-4 genes were not detected in any of the isolates. This results are in line with observations from other studies [29, 52]. Despite low prevalence, various variants of this gene have been reported from different regions of the world, including Iran [53,54,55,56,57]. In addition, many studies have shown the role of chromosomally-mediated mechanisms in colistin resistance [58]. MgrB, a small transmembrane protein with 47 amino acids that regulates the pmrHFIJKLM operon through a signaling cascade of PhoPQ, PmrD, and PmrAB and mediates potent negative feedback on the PhoQ/PhoP regulatory system [59]. The insertional inactivation of mgrB has been shown to be associated with overexpression of the phoPQ and pmrHFIJKLM operons, leading to modification of the LPS target, and eventually occurrence of colistin resistance [60]. The insertional inactivation of mgrB gene due to IS5-like mobile element was observed in one isolate. In particular, the insertion of IS5-like mobile element at nucleotide 75 of mgrB gene was in the same position to that found in other study [30, 52]. Similarly, a truncated MgrB protein by non-sense mutations C88T and C117A was identified in five isolates of the current study, causing premature termination [29, 52]. Remarkably, nine isolates had a wild type mgrB gene and also showed no mutations in the other genes associated with resistance to colistin, suggesting the presence of unknown mechanism(s) for colistin resistance. In addition, the mutated PmrB protein, encoded by the pmrB gene, is a part of the pmrCAB operon, leading to lipopolysaccharide modification and resistance to colistin [31]. In the present study, the A469C mutation in pmrB gene led to amino acid substitution Thr157Pro. Jayol et al., identified a Thr residue at position 157, therefore reinforcing the hypothesis that Thr157Pro might play a key role in acquired resistance to colistin [31].

In this study, single–base pair substitutions, including A449G leading to substitution Asp150Gly and A171C leading to substitution Glu57Asp were identified within the phoQ and phoP sequences, respectively. In other studies, amino acid substitutions in the PhoQ gene have been associated with the colistin resistance phenotype Leu26Pro [61], Leu384Gln [62], Asp150Gly [63], Leu96Pro, and Leu348Gln [60]. In K. pneumoniae, amino acid substitutions, including Ser85Arg, Thr140Pro, Thr157Pro, Ser205Pro [60] and Thr 157Pro [31] in pmrB [62], Leu26Gln and Arg114Ala in phoP [60, 63] have been previously reported. In our study, as in Mateur et al., no mutation in the pmrA gene was observed [63].

Colistin resistance has been found to be associated with upregulation of pmrCAB and pmrHFIJKLM operons and pmrE gene, resulting in lipidA modification in LPS structure. In this study, the relative expression of pmrA, pmrB, pmrC, pmrK, phoP, and phoQ genes in isolates with mgrB mutation (caused by IS element or nonsense mutation) was significantly higher than that of the mgrB in wild type isolate and non-mutant colR isolates. In particular, overexpression of studied genes was observed in the mgrB-inactivated isolate compared to other isolates. Based on the results of this study and others, increased expression of the genes in mgrB-degraded isolates was more noticeable [29,30,31, 64]. Mutations in pmrA/pmrB genes resulted in upregulation of the pmrABC and pmrFHIJKLM operons and pmrE gene [31]. The current study revealed an overexpression of the pmrA, pmrB, pmrC, pmrK, phoP, and phoQ genes in the pmrB-mutated isolate compared to that of the pmrB gene in wild-type colR K. pneumoniae, confirming that the pmrB substitution could be responsible for increased expression levels of relevant genes. In the study of Jayol et al., the expression of pmrA, pmrB, pmrC, and pmrK genes in isolates with pmrB-mutation were significantly increased in comparison with the that of pmrB in wild type isolate [31]. Cheng et al., also found Arg256Gly replacement in the pmrB in 8 of 26 col-R isolates. All of these eight isolates had overexpressed pmrHFIJKLM operon [61].

Conclusion

The prevalence of carbapenem and colistin resistance isolates among the patients with life-threatening infections hospitalized in critical wards is alarming. Unnecessary prescribing of antimicrobial drugs in patients is associated with the eradication of normal flora, leading to spread of MDR and XDR isolates. The emergence and spread of blaNDM and other antibiotic resistance genes in K. pneumoniae and E. coli will further limit the treatment options and threaten the public health of world.

This study demonstrated that carbapenem and colistin resistance K. pneumoniae strains are an emerging threat in different units and should be managed by implementation of timely identification and strict isolation methods that will help to reduce their severe outcomes and mortality rate in critically-ill patients. This study revealed the rapid emergence of extensively-drug resistant K. pneumoniae and E. coli isolates in patients. In addition, we report for the first time a pan-drug resistant strain from Iran that could be a serious warning for the emergence of highly dangerous strains of nosocomial infections in the future.

The molecular mechanisms investigated in this study found to play a major role in development of resistance to antimicrobials, including carbapenem and colistin. Additional factors, such as increased amount of capsular polysaccharide, efflux pumps, and porins are mechanisms that still needs to be investigated.

Availability of data and materials

The datasets generated and analyzed during this reasearch were included in the main document of this manuscript.

Abbreviations

CLSI:

Clinical and Laboratory Standards Institute

TSB:

Tryptic soy broth

OD:

Optical density

PCR:

Polymerase chain reaction

RT-qPCR:

Quantitative reverse transcription-PCR

MLST:

Multi-locus sequencing typing

UTIs:

Urinary tract infections

MDR:

Multidrug-resistant

CRE:

Carbapenem-resistant Enterobacteriaceae

MBL:

Metallo-β-Lactamase

rep PCR:

Repetitive extragenic palindromic

References

  1. David S, Reuter S, Harris SR, Glasner C, Feltwell T, Argimon S, et al. Epidemic of carbapenem-resistant Klebsiella pneumoniae in Europe is driven by nosocomial spread. Nat Microbiol. 2019;4(11):1919–29.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Monegro AF, Regunath H. Hospital acquired infections. StatPearls: StatPearls Publishing; 2018.

    Google Scholar 

  3. Sharahi JY, Maleki DT, Rad ZR, Rad ZR, Goudarzi M, Shariati A, et al. In vitro antibacterial activity of curcumin-meropenem combination against extensively drug-resistant (XDR) bacteria isolated from burn wound infections. 2019.

  4. Tavakolian S, Goudarzi H, Faghihloo E. LPS Induces microRNAs-9,-192, and-205 overexpression in colorectal cell lines SW480, HCT116. Middle East J Cancer. 2020;11(1):72–9.

    CAS  Google Scholar 

  5. Nordmann P, Poirel L, Walsh TR, Livermore DM. The emerging NDM carbapenemases. Trends Microbiol. 2011;19(12):588–95.

    Article  CAS  PubMed  Google Scholar 

  6. Abbasi E, Goudarzi H, Hashemi A, Chirani AS, Ardebili A, Goudarzi M, et al. Decreased carO gene expression and OXA-type carbapenemases among extensively drug-resistant Acinetobacter baumannii strains isolated from burn patients in Tehran, Iran. Acta Microbiol Immunol Hungarica. 2020.

  7. Nasiri MJ, Mirsaeidi M, Mousavi SMJ, Arshadi M, Fardsanei F, Deihim B, et al. Prevalence and mechanisms of carbapenem resistance in Klebsiella pneumoniae and Escherichia coli: a systematic review and meta-analysis of cross-sectional studies from Iran. Microb Drug Resist. 2020;26(12):1491–502.

    Article  CAS  PubMed  Google Scholar 

  8. Bengtsson-Palme J, Larsson DJ. Antibiotic resistance genes in the environment: prioritizing risks. Nat Rev Microbiol. 2015;13(6):396.

    Article  CAS  PubMed  Google Scholar 

  9. Rahmati Roodsari M, Fallah F, Taherpour A, Hakemi Vala M, Hashemi A. Carbapenem-resistant bacteria and laboratory detection methods. Arch Pediatric Infect Dis. 2014;2(1):188–91.

    Google Scholar 

  10. Dortet L, Poirel L, Nordmann P. Worldwide dissemination of the NDM-type carbapenemases in Gram-negative bacteria. BioMed Res Intern. 2014;2014.

  11. Yoon E-J, Kang DY, Yang JW, Kim D, Lee H, Lee KJ, et al. New Delhi metallo-beta-lactamase-producing Enterobacteriaceae in South Korea between 2010 and 2015. Front Microbiol. 2018;9:571.

    Article  PubMed  PubMed Central  Google Scholar 

  12. Huang L, Hu YY, Zhang R. Prevalence of fosfomycin resistance and plasmid-mediated fosfomycin-modifying enzymes among carbapenem-resistant Enterobacteriaceae in Zhejiang, China. J Med Microbiol. 2017;66(9):1332–4.

    Article  CAS  PubMed  Google Scholar 

  13. Higashino HR, Marchi AP, Martins RCR, Batista MV, Neto LVP, de Castro Lima VAC, et al. Colistin-resistant Klebsiella pneumoniae co-harboring KPC and MCR-1 in a hematopoietic stem cell transplantation unit. Bone Marrow Transplant. 2019;54(7):1118–20.

    Article  PubMed  Google Scholar 

  14. Nation RL, Li J. Colistin in the 21st century. Curr Opin Infect Dis. 2009;22(6):535.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Kim S, Woo JH, Kim N, Kim MH, Kim SY, Son JH, et al. Characterization of chromosome-mediated colistin resistance in Escherichia coli isolates from livestock in Korea. Infect Drug Resist. 2019;12:3291.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Du H, Chen L, Tang Y-W, Kreiswirth BN. Emergence of the mcr-1 colistin resistance gene in carbapenem-resistant Enterobacteriaceae. Lancet Infect Dis. 2016;16(3):287–8.

    Article  CAS  PubMed  Google Scholar 

  17. Pajand O, Darabi N, Arab M, Ghorbani R, Bameri Z, Ebrahimi A, et al. The emergence of the hypervirulent Klebsiella pneumoniae (hvKp) strains among circulating clonal complex 147 (CC147) harbouring bla NDM/OXA-48 carbapenemases in a tertiary care center of Iran. Ann Clin Microbiol Antimicrob. 2020;19:1–9.

    Article  CAS  Google Scholar 

  18. Weinstein MP. Performance standards for antimicrobial susceptibility testing. Clinical and Laboratory Standards Institute; 2019.

  19. Veeraraghavan B, Poojary A, Shankar C, Bari AK, Kukreja S, Thukkaram B, et al. In-vitro activity of tigecycline and comparator agents against common pathogens: Indian experience. J Infect Dev Countries. 2019;13(03):245–50.

    Article  CAS  Google Scholar 

  20. Ahmad N, Ali SM, Khan AU. Molecular characterization of novel sequence type of carbapenem-resistant New Delhi metallo-β-lactamase-1-producing Klebsiella pneumoniae in the neonatal intensive care unit of an Indian hospital. Int J Antimicrob Agents. 2019;53(4):525–9.

    Article  CAS  PubMed  Google Scholar 

  21. Gupta V, Soni R, Jain N, Chander J. In vitro cost-effective methods to detect carbapenemases in Enterobacteriaceae. J Lab Phys. 2018;10(1):101.

    CAS  Google Scholar 

  22. Nordmann P, Poirel L, Dortet L. Rapid detection of carbapenemase-producing Enterobacteriaceae. Emerg Infect Dis. 2012;18(9):1503.

    Article  PubMed  PubMed Central  Google Scholar 

  23. Pompilio A, Pomponio S, Crocetta V, Gherardi G, Verginelli F, Fiscarelli E, et al. Phenotypic and genotypic characterization of Stenotrophomonas maltophilia isolates from patients with cystic fibrosis: genome diversity, biofilm formation, and virulence. BMC Microbiol. 2011;11(1):159.

    Article  PubMed  PubMed Central  Google Scholar 

  24. Stepanović S, Vuković D, Hola V, Bonaventura GD, Djukić S, Ćirković I, et al. Quantification of biofilm in microtiter plates: overview of testing conditions and practical recommendations for assessment of biofilm production by staphylococci. APMIS. 2007;115(8):891–9.

    Article  PubMed  Google Scholar 

  25. Heffernan HM, Woodhouse RE, Pope CE, Blackmore TK. Prevalence and types of extended-spectrum β-lactamases among urinary Escherichia coli and Klebsiella spp. in New Zealand. Int J Antimicrob Agents. 2009;34(6):544–9.

    Article  CAS  PubMed  Google Scholar 

  26. Williamson DA, Sidjabat HE, Freeman JT, Roberts SA, Silvey A, Woodhouse R, et al. Identification and molecular characterisation of New Delhi metallo-β-lactamase-1 (NDM-1)-and NDM-6-producing Enterobacteriaceae from New Zealand hospitals. Int J Antimicrob Agents. 2012;39(6):529–33.

    Article  CAS  PubMed  Google Scholar 

  27. Ma Y, Xu X, Guo Q, Wang P, Wang W, Wang M. Characterization of fosA5, a new plasmid-mediated fosfomycin resistance gene in Escherichia coli. Lett Appl Microbiol. 2015;60(3):259–64.

    Article  PubMed  CAS  Google Scholar 

  28. Liu Y-Y, Wang Y, Walsh TR, Yi L-X, Zhang R, Spencer J, et al. Emergence of plasmid-mediated colistin resistance mechanism MCR-1 in animals and human beings in China: a microbiological and molecular biological study. Lancet Infect Dis. 2016;16(2):161–8.

    Article  PubMed  CAS  Google Scholar 

  29. Haeili M, Javani A, Moradi J, Jafari Z, Feizabadi MM, Babaei E. MgrB alterations mediate colistin resistance in Klebsiella pneumoniae isolates from Iran. Front Microbiol. 2017;8:2470.

    Article  PubMed  PubMed Central  Google Scholar 

  30. Cannatelli A, D’Andrea MM, Giani T, Di Pilato V, Arena F, Ambretti S, et al. In vivo emergence of colistin resistance in Klebsiella pneumoniae producing KPC-type carbapenemases mediated by insertional inactivation of the PhoQ/PhoP mgrB regulator. Antimicrob Agents Chemother. 2013;57(11):5521–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Jayol A, Poirel L, Brink A, Villegas M-V, Yilmaz M, Nordmann P. Resistance to colistin associated with a single amino acid change in protein PmrB among Klebsiella pneumoniae isolates of worldwide origin. Antimicrob Agents Chemother. 2014;58(8):4762–6.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  32. Cannatelli A, Di Pilato V, Giani T, Arena F, Ambretti S, Gaibani P, et al. In vivo evolution to colistin resistance by PmrB sensor kinase mutation in KPC-producing Klebsiella pneumoniae is associated with low-dosage colistin treatment. Antimicrob Agents Chemother. 2014;58(8):4399–403.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  33. Carattoli A, Bertini A, Villa L, Falbo V, Hopkins KL, Threlfall EJ. Identification of plasmids by PCR-based replicon typing. J Microbiol Methods. 2005;63(3):219–28.

    Article  CAS  PubMed  Google Scholar 

  34. Kitchel B, Rasheed JK, Patel JB, Srinivasan A, Navon-Venezia S, Carmeli Y, et al. Molecular epidemiology of KPC-producing Klebsiella pneumoniae isolates in the United States: clonal expansion of multilocus sequence type 258. Antimicrob Agents Chemother. 2009;53(8):3365–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Adamek M, Overhage J, Bathe S, Winter J, Fischer R, Schwartz T. Genotyping of environmental and clinical Stenotrophomonas maltophilia isolates and their pathogenic potential. PLoS ONE. 2011;6(11):e27615.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Vaez H, Sahebkar A, Khademi F. Carbapenem-resistant Klebsiella Pneumoniae in Iran: a systematic review and meta-analysis. J Chemother. 2019;31(1):1–8.

    Article  PubMed  Google Scholar 

  37. Codjoe FS, Donkor ES. Carbapenem resistance: a review. Med Sci. 2018;6(1):1.

    Google Scholar 

  38. Hindler JA, Humphries RM. Colistin MIC variability by method for contemporary clinical isolates of multidrug-resistant Gram-negative bacilli. J Clin Microbiol. 2013;51(6):1678–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Paterson DL, Harris P. Colistin resistance: a major breach in our last line of defence. Lancet Infect Dis. 2016;16(2):132.

    Article  PubMed  Google Scholar 

  40. Marchaim D, Chopra T, Pogue JM, Perez F, Hujer AM, Rudin S, et al. Outbreak of colistin-resistant, carbapenem-resistant Klebsiella pneumoniae in metropolitan Detroit, Michigan. Antimicrob Agents Chemother. 2011;55(2):593–9.

    Article  CAS  PubMed  Google Scholar 

  41. Petrosillo N, Taglietti F, Granata G. Treatment options for colistin resistant Klebsiella pneumoniae: present and future. J Clin Med. 2019;8(7):934.

    Article  CAS  PubMed Central  Google Scholar 

  42. Berrazeg M, Diene S, Medjahed L, Parola P, Drissi M, Raoult D, et al. New Delhi Metallo-beta-lactamase around the world: an eReview using Google Maps. Eurosurveillance. 2014;19(20):20809.

    Article  PubMed  Google Scholar 

  43. Briongos-Figuero L, Gomez-Traveso T, Bachiller-Luque P, Dominguez-Gil Gonzalez M, Gómez-Nieto A, Palacios-Martín T, et al. Epidemiology, risk factors and comorbidity for urinary tract infections caused by extended-spectrum beta-lactamase (ESBL)-producing enterobacteria. Int J Clin Pract. 2012;66(9):891–6.

    Article  CAS  PubMed  Google Scholar 

  44. Rahman M, Shukla SK, Prasad KN, Ovejero CM, Pati BK, Tripathi A, et al. Prevalence and molecular characterisation of New Delhi metallo-β-lactamases NDM-1, NDM-5, NDM-6 and NDM-7 in multidrug-resistant Enterobacteriaceae from India. Int J Antimicrob Agents. 2014;44(1):30–7.

    Article  CAS  PubMed  Google Scholar 

  45. Bogaerts P, Verroken A, Jans B, Denis O, Glupczynski Y. Global spread of New Delhi metallo-β-lactamase 1. Lancet Infect Dis. 2010;10(12):831–2.

    Article  PubMed  Google Scholar 

  46. Leverstein-Van Hall MA, Stuart JC, Voets GM, Versteeg D, Tersmette T, Fluit AC. Global spread of New Delhi metallo-β-lactamase 1. Lancet Infect Dis. 2010;10(12):830–1.

    Article  PubMed  Google Scholar 

  47. Voulgari E, Gartzonika C, Vrioni G, Politi L, Priavali E, Levidiotou-Stefanou S, et al. The Balkan region: NDM-1-producing Klebsiella pneumoniae ST11 clonal strain causing outbreaks in Greece. J Antimicrob Chemother. 2014;69(8):2091–7.

    Article  CAS  PubMed  Google Scholar 

  48. Solgi H, Badmasti F, Giske CG, Aghamohammad S, Shahcheraghi F. Molecular epidemiology of NDM-1-and OXA-48-producing Klebsiella pneumoniae in an Iranian hospital: clonal dissemination of ST11 and ST893. J Antimicrob Chemother. 2018;73(6):1517–24.

    Article  CAS  PubMed  Google Scholar 

  49. Markovska R, Schneider I, Ivanova D, Mitov I, Bauernfeind A. Predominance of IncL/M and IncF plasmid types among CTX-M-ESBL-producing Escherichia coliand K lebsiella pneumoniae in Bulgarian hospitals. APMIS. 2014;122(7):608–15.

    Article  CAS  PubMed  Google Scholar 

  50. Protonotariou E, Poulou A, Politi L, Sgouropoulos I, Metallidis S, Kachrimanidou M, et al. Hospital outbreak due to a Klebsiella pneumoniae ST147 clonal strain co-producing KPC-2 and VIM-1 carbapenemases in a tertiary teaching hospital in Northern Greece. Int J Antimicrob Agents. 2018;52(3):331–7.

    Article  CAS  PubMed  Google Scholar 

  51. Yu H, Qu F, Shan B, Huang B, Jia W, Chen C, et al. Detection of the mcr-1 colistin resistance gene in carbapenem-resistant Enterobacteriaceae from different hospitals in China. Antimicrob Agents Chemother. 2016;60(8):5033–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Esposito EP, Cervoni M, Bernardo M, Crivaro V, Cuccurullo S, Imperi F, et al. Molecular epidemiology and virulence profiles of colistin-resistant Klebsiella pneumoniae blood isolates from the Hospital Agency “Ospedale dei Colli”, Naples, Italy. Front Microbiol. 2018;9:1463.

    Article  PubMed  PubMed Central  Google Scholar 

  53. Stoesser N, Mathers AJ, Moore CE, Day NP, Crook DW. Colistin resistance gene mcr-1 and pHNSHP45 plasmid in human isolates of Escherichia coli and Klebsiella pneumoniae. Lancet Infect Dis. 2016;16(3):285–6.

    Article  CAS  PubMed  Google Scholar 

  54. Di Pilato V, Arena F, Tascini C, Cannatelli A, De Angelis LH, Fortunato S, et al. mcr-1.2, a new mcr variant carried on a transferable plasmid from a colistin-resistant KPC carbapenemase-producing Klebsiella pneumoniae strain of sequence type 512. Antimicrob Agents Chemother. 2016;60(9):5612–5.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  55. Rapoport M, Faccone D, Pasteran F, Ceriana P, Albornoz E, Petroni A, et al. First description of mcr-1-mediated colistin resistance in human infections caused by Escherichia coli in Latin America. Antimicrob Agents Chemother. 2016;60(7):4412–3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Pishnian Z, Haeili M, Feizi A. Prevalence and molecular determinants of colistin resistance among commensal Enterobacteriaceae isolated from poultry in northwest of Iran. Gut pathogens. 2019;11(1):1–8.

    Article  Google Scholar 

  57. Moosavian M, Emam N. The first report of emerging mobilized colistin-resistance (mcr) genes and ERIC-PCR typing in Escherichia coli and Klebsiella pneumoniae clinical isolates in southwest Iran. Infect Drug Resist. 2019;12:1001.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Cannatelli A, Giani T, D’Andrea MM, Di Pilato V, Arena F, Conte V, et al. MgrB inactivation is a common mechanism of colistin resistance in KPC-producing Klebsiella pneumoniae of clinical origin. Antimicrob Agents Chemother. 2014;58(10):5696–703.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  59. Poirel L, Jayol A, Bontron S, Villegas M-V, Ozdamar M, Türkoglu S, et al. The mgrB gene as a key target for acquired resistance to colistin in Klebsiella pneumoniae. J Antimicrob Chemother. 2014;70(1):75–80.

    Article  PubMed  CAS  Google Scholar 

  60. Olaitan AO, Morand S, Rolain J-M. Mechanisms of polymyxin resistance: acquired and intrinsic resistance in bacteria. Front Microbiol. 2014;5:643.

    Article  PubMed  PubMed Central  Google Scholar 

  61. Cheng Y-H, Lin T-L, Pan Y-J, Wang Y-P, Lin Y-T, Wang J-T. Colistin resistance mechanisms in Klebsiella pneumoniae strains from Taiwan. Antimicrob Agents Chemother. 2015;59(5):2909–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Choi M-J, Park YK, Peck KR, Ko KS. Mutant prevention concentrations of colistin used in combination with other antimicrobial agents against Acinetobacter baumannii, Klebsiella pneumoniae and Pseudomonas aeruginosa clinical isolates. Int J Antimicrob Agents. 2014;44(5):475–6.

    Article  CAS  PubMed  Google Scholar 

  63. Mathur P, Veeraraghavan B, Devanga Ragupathi NK, Inbanathan FY, Khurana S, Bhardwaj N, et al. Multiple mutations in lipid-A modification pathway & novel fosA variants in colistin-resistant Klebsiella pneumoniae. Fut Sci OA. 2018;4(07):FSO319.

    Article  CAS  Google Scholar 

  64. Wright MS, Suzuki Y, Jones MB, Marshall SH, Rudin SD, van Duin D, et al. Genomic and transcriptomic analyses of colistin-resistant clinical isolates of Klebsiella pneumoniae reveal multiple pathways of resistance. Antimicrob Agents Chemother. 2015;59(1):536–43.

    Article  PubMed  CAS  Google Scholar 

  65. Poirel L, Jayol A, Bontron S, Villegas M-V, Ozdamar M, Türkoglu S, et al. The mgrB gene as a key target for acquired resistance to colistin in Klebsiella pneumoniae. J Antimicrob Chemother. 2015;70(1):75–80.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The present study was financially supported by the research Department of the School of Medicine, Shahid Beheshti University of Medical Sciences (Grant No.: 12606).

Funding

This study was financially supported by the Research Department of the School of Medicine, Shahid Beheshti University of Medical Sciences (Grant No. 12606).

Author information

Authors and Affiliations

Authors

Contributions

JY, AR, SD and AH conceived, designed and performed the experiments and analyzed the data. JY, AR, SD and AH wrote the paper. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Ali Hashemi.

Ethics declarations

Ethics approval and consent to participate

The clinical samples collected were in line with the patients’ diagnostic stages and no additional samples were taken. This research was approved by the Ethics Committee of Shahid Beheshti University of Medical Sciences with the ethical code number IR.SBMU.MSP.REC.1397. 629.

Consent for publication

Not applicable.

Competing interests

There are no conflicts of interest.

Additional information

Publisher's Note

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

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sharahi, J.Y., Hashemi, A., Ardebili, A. et al. Molecular characteristics of antibiotic-resistant Escherichia coli and Klebsiella pneumoniae strains isolated from hospitalized patients in Tehran, Iran. Ann Clin Microbiol Antimicrob 20, 32 (2021). https://doi.org/10.1186/s12941-021-00437-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12941-021-00437-8

Keywords