Skip to main content

Molecular epidemiology of blaCTX-M gene-producing uropathogenic Escherichia coli among Iranian kidney transplant patients: clonal dissemination of CC131 and CC10

Abstract

Background

This study aimed to investigate the phylogenetic characterization and virulence traits of uropathogenic Escherichia coli (UPEC) isolated from kidney transplant patients (KTPs) as well as non-KTPs and analyze the clonal distribution of Extended spectrum β-lactamases (ESBLs)-producing UPEC containing blaCTX-M gene.

Methods

To this end, we determined virulence marker and the phylogenetic characterization of UPEC in non-KTPs (n = 65) and KTPs (n = 46). The non-KTPs were considered the control group of the study. Also, according to the Achtman scheme, we performed multilocus sequence typing to assess the relationship between twenty-nine of ESBL-producing isolates containing blaCTX-M gene.

Results

According to the results of PCR assay, the prevalence of virulence factor genes ranged from 0% (cnf and papG III) to 93.7% (fimH). Also, KTP isolates significantly differed from non-KTP isolates only in terms of the prevalence of pap GI elements. Moreover, the most frequent UPEC isolates were in phylogenetic group B2, followed by group D (18.9%), and group A (13.5%). Furthermore, except for phylogenetic group C, there was no significant correlation between phylogenetic distribution in KTPs and non-KTPs. Additionally, MLST analysis of blaCTX-M carrying isolates identified 18 unique sequence types (ST) the most common of which was ST131 (24.1%), followed by ST1193 (10.3%), while fourteen STs were detected only once.

Conclusions

The results further revealed significant differences between the UPEC isolates from KTPs and non-KTPs regarding the phylogroups C and PAI gene. Based on MLST analysis, we also observed a relatively high diversity in UPEC isolates obtained from KTPs and non-KTPs. Moreover, clonal complex (CC) 131 and ST131 were found to be the most prevalent clones and ST types, respectively. Besides, for the first time, ST8503 were reported in KTPs. These results suggested regular studies on characterization of UPEC isolates among KTPs.

Background

Recently, in kidney transplant patients (KTPs), urinary tract infections (UTI) have been complicated by the widespread disseminated bacterial agents, particularly Escherichia coli, which are resistant to first-line antimicrobial options [1, 2]. Uropathogenic Escherichia coli (UPEC) are common human pathogens associated with both nosocomial and community-acquired UTI [3].

The worldwide increasing extended-spectrum β-lactamases (ESBL) and multi-drug resistance (MDR) of E. coli strains are responsible for a high proportion of UTI frequently affecting KTP [4]. Production of ESBL is one of the primary interfering mechanisms in resistance of E. coli to β-lactam antibiotics, including first- and third-generation cephalosporins. According to several reports, the most prevalent ESBL clone is related to the wide spreading of CTX-M, which has been replaced with blaSHV and blaTEM, as the most prevalent ESBL genes [5,6,7].

The dissemination of CTX-M family among UPEC is mediated by the rapid spread of plasmids and mobile genetic elements. UTI with blaCTX-M-15-producing-UPEC in humans is a major concern especially in KTP [8].

Moreover, E. coli can be typed based on their O serotypes, phylogroups (A, B1, B2, C, D, E, and F), and sequence types (ST), which express different pattern of virulence factors (VFs) [9, 10]. Most UPEC strains infecting immunocompetent humans belong to phylogroups group B2 and, to a lesser extent, phylogroups D, which can overcome host defenses and invade host tissues in presence of VFs [10, 11].

Previous studies have also revealed that many UPEC isolates belong to a small number of clonal groups presenting a range of VFs and variable antibiotic resistance pattern. According to the literature, ST 131, ST69, ST73, and ST95 are pandemic UPEC clones. However, recently, other clones, such as ST10 and ST127 have emerged among UPEC strains [12,13,14]. ST131 containing CTX-15 enzyme is associated with high potential virulence and high level of antibiotic resistance (two major public health problems) [15].

Furthermore, since UTI caused by UPEC in KTPs is considered one of the most important post-transplantation complications increasing the risks of graft rejection, it is critical to determine the characterization and population structure of ESBL-producing UPEC [11].

In Iran, previous studies focused on the investigation of UPEC isolated from hospitalized patients in clinical samples [16,17,18]; thus, no sufficient data are available about the UPEC isolated from KTPs [19, 20]. Also, up until today, there is no reports about the STs of UPEC isolates that have distributed in Iran were available.

Accordingly, we planned to investigate the phylogenetic characterization and virulence traits of UPEC isolated from KTPs and non-KTPs with symptoms of UTI in Iran and analyze the clonal distribution of ESBL-producing UPEC containing CTX-M isolated from KTPs as well as non-KTPs isolates.

Materials and methods

Study design and setting

In this case–control study, UPEC isolates were collected from June 2019 to October 2019 from a total number of 111 patients, who were assigned to two groups of non-KTPs (n = 65) and KTPs (n = 46). All the patients had UTI symptoms. This study was performed in three laboratory centers as well as two nephrology private clinics affiliated to Isfahan University of Medical Sciences (IUMS) [21]. This study was also confirmed and permitted by the Ethics Committee of IUMS (IR.MUI.MED.REC.1398.202). It is necessary to mention that all the UPEC isolates were confirmed in previous our study [21].

Antimicrobial susceptibility testing and phenotypic assays

Antimicrobial susceptibility testing was carried out using the disc diffusion method based on CLSI guidelines [21]. Based on CLSI recommendations, the double-disk synergy test (DDST) was used as a phenotypic method to detect ESBL-producing strains [21, 22].

Molecular detection of resistance genes

The presence of ESBLs genes including blaCTX-M, blaTEM, blaSHV were determined using PCR assays as previous described [23, 24]. The PCR amplicons of ESBL-producing isolates contain blaCTX-M gene were sequenced, and the DNA sequence of each gene was assigned in the GenBank nucleotide database at http://www.ncbi.nlm.nih.gov/blast/.

Virulence gene profiling

The presence of the most relevant UPEC genes were detected by PCR using a set of specific primers as previously described. They included adhesins (papG I, II, III, sfaDE, afaBC, fimH), siderophores (iutA and chuA), miscellaneous (PAI), and toxins (hlyA, and cnf1) [25, 26]. Moreover, the virulence score (VF score; number of VF genes) was considered as the sum of all VFs for each isolates. Positive controls for (afa, fimH, hly A and iutA) and (pap G1, pap G III and chuA) genes were obtained from previous study performed by Malekzadegan et al. and Hojabri et al., respectively [27, 28].

Phylogenetic analysis and multilocus sequence typing (MLST)

Each isolate was assigned to one of the major E. coli phylogroups (A, B1, B2, C, D, E, and F) by quadruplex PCR as described previously [29]. MLST was performed to assess the twenty-nine of ESBL-producing isolates contain blaCTX-M gene according to the Achtman scheme using amplification of the seven housekeeping genes including adk, fumC, gyrB, icd, mdh, purA, and recA available at MLST website (http://mlst.warwick.ac.uk/mlst/dbs/Ecoli). STs were assigned to clonal complex (CCs), single-locus variants (SLV), double-locus variants (DLV), and singletons using the goeBURST algorithm using PHYLOViZ 2.0 software [30].

GenBank accession numbers

The nucleotide sequences of the blaCTX-M gene have been assigned to GenBank under the following accession numbers:

MT274743, MN997127, MT007111, MT274741, MT274742, MT293606, MT293607, MT293608, MT293609, MT293610, MT293611, MT293612, MT310667, MT310668, MT310669, MT310670, MT310671, MT310672, MT310673, MT310674, MT310675, MT310676, MT310677, MT310678, MT310679, MT310680, MT501676, MT501677.

Statistical analysis

Continuous and categorical data were reported as mean ± standard deviation (SD) and frequency (percentage). Categorical data were compared between groups by using chi-squared test and pairwise comparisons were conducted by adopting Bonferroni correction for multiple testing. Continuous variables (VF and antibiotic resistance scores) were compared between groups by using one-way Analysis Of Variance (ANOVA) and Gabriel post hoc test has been conducted for pirwise comparisons. All statistical analyses were performed using SPSS 16 (IBM Corp., USA). Statistical significance level was considered as P value of < 0.05.

Results

According to the results of PCR assay, 90.9% of the UPEC isolates was positive for one or more of the VF genes (Table 1). Among all of the isolates, the prevalence of individual VF genes ranged from 0% (cnf and papG III) to 93.7% (fimH) (Additional file 1: Figure S1A–G).

Table 1 Distribution of genes encoding virulence factors among KTP and non-KTP isolates

Also, KTP isolates significantly differed from non-KTP isolates only in terms of pap GI elements prevalence [52.2% (24/46) versus 29.2% (19/65), P = 0.01], not regarding any other studied virulence genes. In addition, the overall median aggregate VF scores were higher for KTP isolates (median VF score 4.97, range 2–8) than those for non-KTP isolates (median VF score 4.72, range 2–9). However, there was no significant correlation between VF score and type of patient. Further, the way how genes encoding VFs were distributed among KTP and non-KTP isolates is shown in Table 1

Phylogenetic groups

According to the quadruplex PCR assay results, most (38.7%) of the UPES isolates belonged to phylogenetic group B2, followed by group D (18.9%), group A (13.5%), B1 (9%), C (5.4%), E (4.5%), and F (3.6%) (Additional file 1: Figure S2). The remaining 6.3% of the isolates were found to be un-typeable. Phylogenetic group B2, D, A, and F were also more prevalent in the KTPs than the non-KTPs isolates, while the phylogenetic group B1, C, and E were more frequent in control group. However, except for phylogenetic group C that was more significantly prevalent in control group, there was no significant difference between KTPs and non-KTPs regarding phylogenetic distribution. Distribution of the phylogenetic groups is summarized in Table 2.

Table 2 Distribution of VF genes according to phylogroups

Correlations between phylogroups and VAGs

According to distribution of VFs in phylogroups, fimH, PAI, and iutA were more frequent among B2 members followed by fimH, iutA, and sfa, which were more frequent among phylogroups D and fimH, papGI and iutA more common among phylogroups A and B1 members. Additionally, sfa were more prevalent in phylogroup C than other groups. Frequency of VF genes based on phylogroups is shown in more details in Table 2.

Furthermore, phylogenetic groups considerably varied in VF content so that the highest median VF scores (and ranges) were 6 (3–7) for isolates belonged to phylogroup D, followed by phylogroup B2, E with 5 (3–8), and 5 (4–6), respectively, while the lowest median VF score (and range) were 3.3 (2–4) for isolates belonged to phylogroup C. Statistical analysis also showed no significant differences between VF score and various phylogroups, except for phylogroup B2 and C. Moreover, among different phylogroups, pai gene was statistically significant in phylogroup B2 in comparison to other phylogroups.

CTX-M group (mainly CTX-M-15) was also found to be more prevalent in B2, D, and A phylogroups, while blaTEM was more frequent in D, F, and A phylogroups.

Phylogroups further varied in antibiotic susceptibility pattern; the most susceptible phylogroup was phylogroups C (resistance score median 2 and range 0–8), and the most resistant phylogroup was phylogroup B1 (median 6 and range 0–11), followed by phylogroups D and B2 (median 6 and range 0–11). Overall, we noticed that 35% of the UPEC strains were ESBL-producer [21]. Isolates belonged to phylogroups F and B2 exhibited a high prevalence (83.3%) of ESBL. Distribution of the antimicrobial resistance profiles according to phylogroups is also shown in Table 3.

Table 3 Distribution of the antimicrobial resistance profiles according to phylogroups

Among 31/5% (35/111) of ESBL-producing isolates, 82.9% (29/35) contained blaCTX-M genes. Of these blaCTX-M positive isolates, 27 isolates carried blaCTX-M-15, and 2 isolates carried blaCTX-M-55.

MLST analysis of blaCTX-M carrying isolates identified 18 unique STs, the most common were ST131 (7; 24.1%) followed by, ST1193 (3; 10.3%), ST38 (2; 6.9%), ST8503 (2; 6.9%) and fourteen STs were detected only once ST69, ST90, ST167, ST127, ST405, ST617, ST636, ST648, ST744, ST838, ST2115, ST4516, ST4988 and ST5114. Overall, using the goeBURST algorithm, STs were belonged to eight CCs including CC131 (ST131, 8503 and 838), CC10 (ST167, 744, 4988 and 617), C14 (ST1193), CC23 (ST90), CC38 (ST38), CC69 (ST69), CC405 (ST405) and CC648 (ST648) and remaining STs were classified as singleton. The distribution of STs and characteristics of the 29 blaCTX-M-producing UPEC is shown in Table 4. Moreover, phylogenic analysis of our STs in each CCs has been shown in Fig. 1.

Table 4 Characteristics of the 29 blaCTX-M -producing UPEC
Fig. 1
figure 1

Phylogenetic tree of MLST for the blaCTX-M positive strains; clonal complexes are shown with haloes. The colors of the circles correspond to the type of patients (Red: transplanted patients; Blue: Control group)

Discussion

In this case–control study, we characterized the virulence potential, phylogenetic group distributions, and clonal diversity in KTPs and non-KTPs with UTI caused by UPEC. To date, only some studies have been carried out on the above-mentioned items in KTPs. Therefore, this is the first study to simultaneously investigate these aspects among KTPs in Iran [19, 20] to assist physicians to conform with the latest pathogenic strains causing UTI by having a better perception of the virulence markers of UPEC strains, especially in transplant patients.

UPEC isolates produce different types of virulence factor, such as adhesion, siderophores and toxins, which are vital for attaching to, colonizing, and invading their host. Of VFs, genes encoding adhesions such as type‐1 fimbriae, which are essential for establishment and progression of UTI, are the most common occurring ones in UPEC isolates. As we expected, the results revealed that fimH was the most frequent virulence genes followed by sfa and papG I. These findings are in consistent with results of other studies indicated that fimH is the most predominant virulence factor between UPEC isolates [16, 27, 31,32,33,34].

In agreement with a study conducted in another part of Iran (79.4%) as well as in South Korea (100%), in the present study, the sfa gene was the second most common adhesion gene (48.6%) among the studied isolates [27, 35].

In several other studies performed on KTPs, the sfa was reported to include 26.1% [20], 19% [36], and 0% [34] of UPEC isolates. All of these percentages are lower than what we detected in the current work.

In this study, the second most common VF gene was found to be iutA. iutA is the siderophore-mediated uptake systems in UPEC strain that contribute to resistance to serum killing. As shown by findings of the current work, 64.9% of UPEC was PCR-positive for iutA. These results are broadly consistent with those reported considering the KTP [34], community [14], and hospitalized population [27].

In the presents study, although KTP isolates had higher VFs score than non-KTP isolates, we did not find significant difference between the VF scores of KTP and non-KTP isolates. However, among adhesion factors, pap G1 was significantly more frequent in KTP isolates than non-KTP isolates (p = 0.01). Distribution of VFs is usually varied in UPEC. This discrepancy could be due to the difference in disease type, geographic area, and the studied population. Based on phylogenetic analysis, E. coli pathotypes have been classified into four main phylogenetic groups of A, B1, B2, and D. According to previous studies, B2 and D are the most frequent phylogroups among UPEC isolates, whereas phylogroups A and B1 have been mostly recognized as commensal E. coli isolates [35, 37].

This study also revealed that the major phylogenetic group in KTP isolates was B2 (39.1%), followed by D (21.7%), and A (19.6%), while the predominance of phylogenetic groups in non-KTP isolates were B2 (38.7%), followed by D (18.9%), and B1 (9%). However, there was no significant difference between phylogenetic dissemination in KTPs and non-KTPs, except for phylogenetic group C that was significantly predominant in the control group. These findings are in agreement with several other studies, which indicated that the majority of UPEC strains, with an increased potential to cause UTIs, belonged to group B2 and D [31], while diarrheagenic or commensals strains often belonged to phylogroups A and B1 [38].

Furthermore, it should be pointed out that this study’s findings suggested that UTIs can have extraintestinal origin even in immunosuppressed patients such as KTPs.

In 2019, Basha et al. described the phylogroup frequency of UPEC strains in KTP whereby the most common groups were B2, A, and B1, while in control group, B2, B1, and D were the most common groups. These findings strongly support the present work’s reports [39]. Similarly, Tashk et al. revealed that the majority of UPEC isolates from KTPs belonged to phylogroups A, B1, and F, which is uncommon for UPEC strain [40].

Unlike the current study, Khairy et al. reported that the predominant phylogenetic groups among UPEC isolates were groups A, followed by B2 and D [41]. Also, in another study conducted by Farajzadah Sheikh et al., phylogroup D (58%) was detected as the most common phylogroups of UPEC strains [16]. These discrepancies suggested that distribution of phylogenetic groups might vary across different regions, features of host, environmental, social, and geographic conditions.

In another part of our study, finding revealed that phylogroup B1, followed by phylogroups D and B2 had high level of resistance score. On the other hand, UPEC isolates belonged to phylogroups F and B2 exhibited a high frequency of ESBL strain. According to previous report, isolates belonged to phylogenetic group B2 and D had high level of antibiotic resistance compared to other groups, which our results in partially in agreement with their results [33, 42].

Besides, resistant rate of nitrofurantoin among different phylogenetic group was significant and member of phylogroups E and D had high level of resistance than other phylogroups. However, due to the appropriate level of sensitivity to nitrofurantoin compared to other oral antibiotics, nitrofurantoin can be recommended as a suitable oral drug to be prescribed by physicians to these patients.

Moreover, member of phylogroups B2 and D as most important phylogroups, had high level of sensitivity to carbapenems, piperacillin–tazobactam and amikacin.

Therefore, based on previous findings and our results, the type of phylogenetic group and the pattern of antibiotic resistance associated with it may be appropriate and useful in the correct administration of antibiotics, especially in patients with kidney transplant.

In the present study, according to the observed population structure of E. coli, members of CC131 predominantly belonged to phylogenetic group B2. By the widely used quadruplex PCR method, we further confirmed that 90% of members of CC131 found in this study could be classified into B2 phylogenetic group.

Among the ESBL-producing UPEC isolates, blaCTX-M-15 is mostly associated with the ST131 clonal group, which has emerged in many parts of the world since 2000 [42]. Our results confirmed studies reported blaCTX-M-15 as the dominant CTX-M type among E. coli ST131 isolates in western Asia (68%) [43], Iran [44], and Spain (95%) [45].

On the other hand, a recent meta-analysis study described that ST131 clone occurrence among ESBL-producing E. coli was 42.7%, which is partially comparable with this work’s results with 34.5%. The observed discrepancies among the studies’ results were mostly related to different sample sizes, studies’ population, or detected O-serogroups.

We further perceived that ST8503 as a member of CC131 belonging to phylogroup B2 co-harbored blaCTX-M, and blaTEM genes. ST8503 as an MDR ESBL-producing ST having high pathogenicity had not previously been reported from KTPs. So, our results confirmed the emergence of blaCTX-M-15-producing ST8503 E. coli in the studied isolates.

Within the last several years, multiple studies across the world have reported rise of a new pandemic fluoroquinolones-resistant clonal group, E. coli ST1193. In this study, B2-ST1193 was the second most prevalent ST (10.3%). We also discerned that all of the B2-ST1193 isolates were resistant to ciprofloxacin and harboured blaCTX-M-15. Moreover, ST1193 has recently emerged from ST14CC as an important cause of extraintestinal disease in humans [46]. This is similar to some previous observations for different populations, such as community, hospital, and KTPs from which UPEC isolates had been derived in Switzerland [14], Iran [47], and Germany [39].

ST 38, on the other hand, has been identified as a hybrid of UPEC and enteroaggregative E. coli (EAEC) that is commonly associated with UTIs. In this study, ST410 was found in both groups of patients, one isolate, which co-produced blaCTX-M-15, blaSHV and blaTEM genes, belonged to phylogenetic group B2, while another ST38 E. coli, which had been previously reported in Germany, was associated with KTPs and belonged to phylogenetic group D [39]. Moreover, a report from Saudi Arabia stated that ST38 E. coli strains producing ESBL were one of the predominant ST in clinical settings [48].

We also noted that CC10 was the second predominant lineage disseminated among KTPs and non-KTPs. All members of CC10 were MDR and ESBL producer. They could also carry the blaCTX-M gene: of them, ST744 and ST617 carry blaCTX-M-55 gene. Previous studies have also identified the association between members of CC10 and KTPs and hospital-acquired UTI [34, 39].

ST648 is one of the most successful and persistent clones and is also the most predominant genotype in CC648. In line with previous reports in Iran [49], we found ST648 to be associated with KTPs, phylogenetic group B2, and blaCTX-M-15 production in E. coli isolates.

In contrast to previous works, which announced ST69 as the predominant ST among KTPs and other population, we rarely found this ST among the studied strains so that only one ST69 was reported [39]. However, as previously reported, the ST69 have been predominantly detected among KTPs [39]. Moreover, in a study conducted in Iran, ST69 was reported with different antimicrobial resistance patterns [47].

In the current study, B2-ST405 was isolated from a KTP and considered as high-risk epidemic clone owing to its ability to co-produce blaCTX-M-15, blaTEM, fimH, PAI, sfa, iut, and papG1 genes and disseminate β-lactamases. ST405 had not been already reported from KTPs. So, we reported for the first time the emergence of ST405 in KTPs [34, 39].

The main finding of this study was that the ST8503 and ST405 clones had not been previously reported among KTPs. this could be indicative of the fact that the spread of these ESBL-producing ST is an alarm about infection control policies. Thus, it is essential to recognize the clonal relatedness in the KTPs to amplify the surveillance and infection control system in clinical setting.

Another issue investigated in the current work was detection of isolates belonging to CC131, CC648, CC69, and CC 14 (international epidemic UPEC clones rapidly transmittable) in Iran.

The present research had limitations. The first, UPEC isolates was obtained only from there hospitals; thus the results should be interpreted with caution. Second, the impossibility of using typing methods such as pulsed-field gel electrophoresis (PFGE).

In conclusion, in this work, which was the first report on phylogenetic characterization and MLST analysis of UPEC isolates obtained from KTPs in Iran, we revealed significant differences between KTPs and non-KTPs regarding the phylogroups C and PAI gene of the UPEC isolates. We also observed differences between UPEC isolates obtained from the two studied group based on factors, such as the patients’ phylogenetic background (phylogroups and MLST analysis), their antibiotic resistance, and VF scores.

Moreover, based on MLST analysis, a relatively high diversity was observed in UPEC isolates obtained from KTPs and non-KTPs. Eighteen ST and nine clonal complexes were also identified. Pandemic UPEC clones including CC131 were also found to be the most prevalent ST types and clones. Further, ST 8503 was reported for the first time in KTPs. These results suggested regular studies on characterization of UPEC isolates among KTPs to help clinicians prescribe the most appropriate antibiotic and prevent further development of antimicrobial drug resistance.

Availability of data and materials

Not applicable.

Abbreviations

ESBL:

Extended-spectrum β-lactamases

CLSI:

Clinical and Laboratory Standards Institute

CCs:

Clonal complex

DLV:

Double-locus variants

DDST:

Double-disk synergy test

KTP:

Kidney transplant patients

ST:

Sequence types

SLV:

Single-locus variants

SD:

Standard deviation

UPEC:

Uropathogenic Escherichia coli

UTI:

Urinary tract infections

MLST:

Multilocus sequence typing

VFs:

Virulence factors

References

  1. Espinar MJ, Miranda IM, Costa-de-Oliveira S, et al. Urinary tract infections in kidney transplant patients due to Escherichia coli and Klebsiella pneumoniae-producing extended-spectrum beta-lactamases: risk factors and molecular epidemiology. PLoS ONE. 2015;10:e0134737.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  2. Vidal E, Cervera C, Cordero E, et al. Management of urinary tract infection in solid organ transplant recipients: consensus statement of the group for the study of infection in transplant recipients (GESITRA) of the Spanish society of infectious diseases and clinical microbiology (SEIMC) and the Spanish network for research in infectious diseases (REIPI). Enferm Infecc Microbiol Clin. 2015;33:679.e1-679.e21.

    Article  Google Scholar 

  3. Pilmis B, Scemla A, Join-Lambert O, et al. ESBL-producing enterobacteriaceae-related urinary tract infections in kidney transplant recipients: incidence and risk factors for recurrence. Infect Dis. 2015;47:714–8.

    Article  CAS  Google Scholar 

  4. Cloeckaert A, Rodríguez-Baño J, Mingorance J, et al. Virulence profiles of bacteremic extended-spectrum β-lactamase-producing Escherichia coli: association with epidemiological and clinical features. PLoS ONE. 2012;7:e44238.

    Article  CAS  Google Scholar 

  5. Yu Y, Ji S, Chen Y, et al. Resistance of strains producing extended-spectrum β-lactamases and genotype distribution in China. J Infect. 2007;54:53–7.

    Article  PubMed  Google Scholar 

  6. D’Andrea MM, Arena F, Pallecchi L, et al. CTX-M-type β-lactamases: a successful story of antibiotic resistance. Int J Med Microbiol. 2013;303:305–17.

    Article  PubMed  CAS  Google Scholar 

  7. Nabti LZ, Sahli F, Radji N, et al. High prevalence of multidrug-resistant Escherichia coli in urine samples from inpatients and outpatients at a tertiary care hospital in Sétif, Algeria. Microb Drug Resist. 2019;25:386–93.

    Article  CAS  PubMed  Google Scholar 

  8. Irrgang A, Falgenhauer L, Fischer J, et al. CTX-M-15-producing E. coli isolates from food products in Germany are mainly associated with an IncF-type plasmid and belong to two predominant clonal E. coli lineages. Front Microbiol. 2017;8:2318.

    Article  PubMed  PubMed Central  Google Scholar 

  9. Manges AR, Johnson JR. Food-borne origins of Escherichia coli causing extraintestinal infections. Clin Infect Dis. 2012;55:712–9.

    Article  PubMed  Google Scholar 

  10. Raimondi S, Righini L, Candeliere F, et al. Antibiotic resistance, virulence factors, phenotyping, and genotyping of E. coli isolated from the feces of healthy subjects. Microorganisms. 2019;7:251.

    Article  CAS  PubMed Central  Google Scholar 

  11. Bert F, Johnson JR, Ouattara B, et al. Genetic diversity and virulence profiles of Escherichia coli isolates causing spontaneous bacterial peritonitis and bacteremia in patients with cirrhosis. J Clin Microbiol. 2010;48:2709–14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Tartof SY, Solberg OD, Manges AR, et al. Analysis of a uropathogenic Escherichia coli clonal group by multilocus sequence typing. J Clin Microbiol. 2005;43:5860–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Chakravortty D, Alghoribi MF, Gibreel TM, et al. Galleria mellonella infection model demonstrates high lethality of ST69 and ST127 uropathogenic E. coli. PLoS ONE. 2014;9:e101547.

    Article  CAS  Google Scholar 

  14. Nüesch-Inderbinen MT, Baschera M, Zurfluh K, et al. Clonal diversity, virulence potential and antimicrobial resistance of Escherichia coli causing community acquired urinary tract infection in Switzerland. Front Microbiol. 2017;8:2334.

    Article  PubMed  PubMed Central  Google Scholar 

  15. Nicolas-Chanoine MH, Bertrand X, Madec JY. Escherichia coli ST131, an intriguing clonal group. Clin Microbiol Rev. 2014;27:543–74.

    Article  PubMed  PubMed Central  Google Scholar 

  16. Farajzadah Sheikh A, Goodarzi H, Yadyad MJ, et al. Virulence-associated genes and drug susceptibility patterns of uropathogenic Escherichia coli isolated from patients with urinary tract infection. Infect Drug Resist. 2019;12:2039–47.

    Article  PubMed  PubMed Central  Google Scholar 

  17. Shahbazi S, Asadi Karam MR, Habibi M, et al. Distribution of extended-spectrum beta-lactam, quinolone and carbapenem resistance genes, and genetic diversity among uropathogenic Escherichia coli isolates in Tehran, Iran. J Glob Antimicrob Resist. 2018;14:118–25.

    Article  PubMed  Google Scholar 

  18. Shahraki-Zahedani S, Moghadampour M, Bokaeian M, et al. Prevalence of CTX-M-8 and CTX-M-15 type extended-spectrum beta-lactamases between Klebsiella pneumoniae spp. isolated from Zahedan, Southeast Iran. J Chemother. 2016;28:343–5.

    Article  CAS  PubMed  Google Scholar 

  19. Kashef Nejad M, Jazani NH, Sharifi Y. Urinary tract infections among kidney transplant patients due to extended-spectrum beta-lactamase-producing bacteria. Microb Pathog. 2017;107:276–9.

    Article  PubMed  Google Scholar 

  20. Mohammadzadeh M, Tavakoli M, Yaslianifard S, et al. Genetic diversity and antibiotic susceptibility of uropathogenic Escherichia coli isolates from kidney transplant recipients. Infect Drug Resist. 2019;12:1795–803.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Halaji M, Shahidi S, Atapour A, et al. Characterization of extended-spectrum β-lactamase-producing uropathogenic Escherichia coli among Iranian kidney transplant patients. Infect Drug Resist. 2020;13:1429–37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Wayne P. Clinical and Laboratory Standards Institute: performance standards for antimicrobial susceptibility testing: twenty-fourth informational supplement, M100-S24. Clinical and Laboratory Standards Institute (CLSI); 2019. p. 39.

  23. Edelstein M, Pimkin M, Palagin I, et al. Prevalence and molecular epidemiology of CTX-M extended-spectrum-lactamase-producing Escherichia coli and Klebsiella pneumoniae in Russian hospitals. Antimicrob Agents Chemother. 2003;47:3724–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Wiegand I, Geiss HK, Mack D, et al. Detection of extended-spectrum beta-lactamases among enterobacteriaceae by use of semiautomated microbiology systems and manual detection procedures. J Clin Microbiol. 2007;45:1167–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Watt S, Lanotte P, Mereghetti L, et al. Escherichia coli strains from pregnant women and neonates: intraspecies genetic distribution and prevalence of virulence factors. J Clin Microbiol. 2003;41:1929–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. White AP, Sibley KA, Sibley CD, et al. Intergenic sequence comparison of Escherichia coli isolates reveals lifestyle adaptations but not host specificity. Appl Environ Microbiol. 2011;77:7620–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Malekzadegan Y, Khashei R, Sedigh Ebrahim-Saraie H, et al. Distribution of virulence genes and their association with antimicrobial resistance among uropathogenic Escherichia coli isolates from Iranian patients. BMC Infect Dis. 2018;18:1–9.

    Article  CAS  Google Scholar 

  28. Hojabri Z, Mirmohammadkhani M, Darabi N, et al. Characterization of antibiotic-susceptibility patterns and virulence genes of five major sequence types of Escherichia coli isolates cultured from extraintestinal specimens: a 1-year surveillance study from Iran. Infect Drug Resist. 2019;12:893–903.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Clermont O, Christenson JK, Denamur E, et al. The Clermont Escherichia coli phylo-typing method revisited: improvement of specificity and detection of new phylo-groups. Environ Microbiol Rep. 2013;5:58–65.

    Article  CAS  PubMed  Google Scholar 

  30. Nascimento M, Sousa A, Ramirez M, et al. PHYLOViZ 2.0: providing scalable data integration and visualization for multiple phylogenetic inference methods. Bioinformatics. 2017;33:128–9.

    Article  CAS  PubMed  Google Scholar 

  31. Bahadori M, Motamedifar M, Derakhshandeh A, et al. Genetic relatedness of the Escherichia coli fecal population and strains causing urinary tract infection in the same host. MicrobiologyOpen. 2019;8:e00759.

    Article  PubMed  Google Scholar 

  32. Hojabri Z, Darabi N, Arab M, et al. Clonal diversity, virulence genes content and subclone status of Escherichia coli sequence type 131: comparative analysis of E. coli ST131 and non-ST131 isolates from Iran. BMC Microbiol. 2019;19:1–10.

    Article  CAS  Google Scholar 

  33. Raeispour M, Ranjbar R. Antibiotic resistance, virulence factors and genotyping of uropathogenic Escherichia coli strains. Antimicrob Resist Infect Control. 2018;7:1–9.

    Article  Google Scholar 

  34. Coussement J, Argudín MA, Heinrichs A, et al. Host and microbial factors in kidney transplant recipients with Escherichia coli acute pyelonephritis or asymptomatic bacteriuria: a prospective study using whole-genome sequencing. Nephrol Dial Transplant. 2019;34:878–85.

    Article  CAS  PubMed  Google Scholar 

  35. Lee JH, Subhadra B, Son YJ, et al. Phylogenetic group distributions, virulence factors and antimicrobial resistance properties of uropathogenic Escherichia colistrains isolated from patients with urinary tract infections in South Korea. Lett Appl Microbiol. 2016;62:84–90.

    Article  CAS  PubMed  Google Scholar 

  36. Merçon M, Regua-Mangia AH, Teixeira LM, et al. Urinary tract infections in renal transplant recipients: virulence traits of uropathogenic Escherichia coli. Transplant Proc. 2010;42:483–5.

    Article  PubMed  Google Scholar 

  37. Munkhdelger Y, Gunregjav N, Dorjpurev A, et al. Detection of virulence genes, phylogenetic group and antibiotic resistance of uropathogenic Escherichia coli in Mongolia. J Infect Dev Ctries. 2017;11:51–7.

    Article  CAS  PubMed  Google Scholar 

  38. Tourret J, Denamur E. Population phylogenomics of extraintestinal pathogenic Escherichia coli. Microbiol Spectr. 2016. https://doi.org/10.1128/microbiolspec.

    Article  PubMed  Google Scholar 

  39. Abo Basha J, Kiel M, Görlich D, et al. Phenotypic and genotypic characterization of Escherichia coli causing urinary tract infections in kidney-transplanted patients. J Clin Med. 2019;8:988.

    Article  CAS  PubMed Central  Google Scholar 

  40. Tashk P, Lecronier M, Clermont O, et al. Épidémiologie moléculaire et cinétique des infections urinaires précoces à Escherichia coli chez les sujets transplantés d’un rein. Nephrol Ther. 2017;13:236–44.

    Article  PubMed  Google Scholar 

  41. Thumbikat P, Khairy RM, Mohamed ES, et al. Phylogenic classification and virulence genes profiles of uropathogenic E. coli and diarrhegenic E. coli strains isolated from community acquired infections. PLoS ONE. 2019;14:e0222441.

    Article  CAS  Google Scholar 

  42. Matsumura Y, Yamamoto M, Nagao M, et al. Emergence and spread of B2-ST131-O25b, B2-ST131-O16 and D-ST405 clonal groups among extended-spectrum-β-lactamase-producing Escherichia coli in Japan. J Antimicrob Chemother. 2012;67:2612–20.

    Article  CAS  PubMed  Google Scholar 

  43. Jafari A, Falahatkar S, Delpasand K, et al. Emergence of Escherichia coli ST131 causing urinary tract infection in Western Asia: a systematic review and meta-analysis. Microb Drug Resist. 2020;26(11):1357–64.

    Article  CAS  PubMed  Google Scholar 

  44. Moghanni M, Ghazvini K, Farsiani H, et al. High prevalence of sequence type 131 isolates producing CTX-M-15 among extended-spectrum β-lactamase-producing Escherichia coli strains in northeast Iran. J Glob Antimicrob Resist. 2018;15:74–8.

    Article  PubMed  Google Scholar 

  45. Merino I, Shaw E, Horcajada JP, et al. CTX-M-15-H30Rx-ST131 subclone is one of the main causes of healthcare-associated ESBL-producing Escherichia coli bacteraemia of urinary origin in Spain. J Antimicrob Chemother. 2016;71:2125–30.

    Article  PubMed  Google Scholar 

  46. Johnson TJ, Elnekave E, Miller EA, et al. Phylogenomic analysis of extraintestinal pathogenic Escherichia coli sequence type 1193, an emerging multidrug-resistant clonal group. Antimicrob Agents Chemother. 2019;63:e01913-18.

    Article  PubMed  Google Scholar 

  47. Neamati F, Khorshidi A, Moniri R, et al. Molecular epidemiology of antimicrobial resistance of uropathogenic Escherichia coli isolates from patients with urinary tract infections in a tertiary teaching hospital in Iran. Microb Drug Resist. 2020;26:60–70.

    Article  CAS  PubMed  Google Scholar 

  48. Alghoribi MF, Gibreel TM, Farnham G, et al. Antibiotic-resistant ST38, ST131 and ST405 strains are the leading uropathogenic Escherichia coli clones in Riyadh, Saudi Arabia. J Antimicrob Chemother. 2015;70:2757–62.

    Article  CAS  PubMed  Google Scholar 

  49. Solgi H, Giske CG, Badmasti F, et al. Emergence of carbapenem resistant Escherichia coli isolates producing bla NDM and bla OXA-48-like carried on IncA/C and IncL/M plasmids at two Iranian university hospitals. Infect Genet Evol. 2017;55:318–23.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors would like to appreciate the Department of Microbiology of IUMS.

Funding

The authors are grateful to office of vice-chancellor for Research of Isfahan University of Medical Sciences for the support of the current study.

Author information

Authors and Affiliations

Authors

Contributions

SAH designed the study and reviewed the manuscript, and edited the final version. MH contributed to design the study, collected the data, and drafted the manuscript. SSH, BA and AA contributed to design the study, collected the data and participated in drafting the manuscript. AF analyzed the data, reviewed the manuscript, and edited the final version. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Seyed Asghar Havaei.

Ethics declarations

Ethics approval and consent to participate

This study was also confirmed and permitted by the Ethics Committee of Isfahan University of Medical Sciences (IR.MUI.MED.REC.1398.202). These patients routinely referred to the microbiology laboratory and a urine sample was taken and then we only used bacteria isolates obtained from urine culture.

Consent for publication

Not applicable.

Competing interests

The authors report no conflicts of interest in this work.

Additional information

Publisher's Note

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

Supplementary Information

Additional file 1: Figure S1.

(A) Agarose gel electrophoresis of PCR products for sfa (1177 bp) gene. M: 100 bp DNA size marker. (B) Agarose gel electrophoresis of PCR products for hlyA (1177 bp) gene. M: 100 bp DNA size marker. (C) Agarose gel electrophoresis of PCR products for papG III (258 bp) gene. M: 100 bp DNA size marker. (D) Agarose gel electrophoresis of PCR products for iutA (300 bp) gene. M: 100 bp DNA size marker. (E) Agarose gel electrophoresis of PCR products for papG1 (461 bp) gene. M: 100 bp DNA size marker. (F) Agarose gel electrophoresis of PCR products for pai (930 bp) gene. M: 100 bp DNA size marker. (G) Agarose gel electrophoresis of PCR products for fimH (508 bp) and (afa 750 bp) genes. M: 100 bp DNA size marker. Figure S2. Agarose gel electrophoresis of Quadruplex PCR profiles of Clermont phylo-typing method. D: phylogenetic group D; A: phylogenetic group A; B2: phylogenetic group B2; Un: Unknown phylogenetic group M: 100 bp DNA size marker.

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

Halaji, M., Shahidi, S., Ataei, B. et al. Molecular epidemiology of blaCTX-M gene-producing uropathogenic Escherichia coli among Iranian kidney transplant patients: clonal dissemination of CC131 and CC10. Ann Clin Microbiol Antimicrob 20, 65 (2021). https://doi.org/10.1186/s12941-021-00470-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12941-021-00470-7

Keywords