Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2019 Mar 1.
Published in final edited form as: Int J Antimicrob Agents. 2018 Jan 31;51(3):479–483. doi: 10.1016/j.ijantimicag.2018.01.012

Chromosomal Mutations that Accompany qnr in Clinical Isolates of Escherichia coli

Laura Vinué a,*, David C Hooper a, George A Jacoby b
PMCID: PMC5849519  NIHMSID: NIHMS936036  PMID: 29360506

Abstract

We examined 13 qnr-positive and 14 qnr-negative clinical isolates of Escherichia coli for mutations previously seen in a qnr-containing laboratory strain exposed to supra MIC concentrations of ciprofloxacin. Among the qnr-positive strains, those with ciprofloxacin MICs of ≥ 2 µg/ml had at least one mutation in gyrA. Mutations in parC were present in strains with a ciprofloxacin MIC of ≥128 µg/ml. The 6 most ciprofloxacin-resistant strains contained additional plasmid-mediated quinolone resistance determinants. aac(6’)-Ib-cr was found in 5 of the 6 strains. Eleven of the 13 strains had alterations in MarR, 9 in SoxR, and 5 had mutations in AcrR, All had elevated expression of at least one efflux pump gene, predominantly acrA (92% of the strains), followed by mdtE (54%) and ydhE (46%). Nine had functionally silent alterations in rfa, 2 had mutations in gmhB, and one of these also a mutation in surA. An E. coli with ciprofloxacin MIC of 1024 µg/ml contained 4 different plasmid-mediated quinolone resistance determinants as well as gyrA, parC, parE and pump overexpression mutations. Nine of the 14 qnr-negative strains had mutations in topoisomerase genes with the ciprofloxacin MIC starting at 0.25 µg/ml and reaching 256 µg/ml. The three most resistant strains also had mutations in parE. Twelve had alterations in MarR, 10 in SoxR and 5 in AcrR. Ten of the 14 strains had elevated expression of efflux pumps with acrA (71.4%), followed by ydhE (50%) and mdtE (14.3%). A diversity of resistance mechanisms occurs in clinical isolates with and without qnr genes.

Keywords: E. coli, quinolones, qnr, mutations, clinical isolates

1. Introduction

Plasmid-mediated qnr genes are often found in clinical isolates with mutations in the quinolone resistance determining region of type II topoisomerase genes [[1], [2], [3], [4]], providing an additive effect on resistance [[5]]. Hence, it was unexpected that gyrA mutations were absent or very rare when mutants with higher levels of ciprofloxacin resistance were selected from a qnr-containing strain of E. coli [[6], [7], [8]]. Selected instead are mutations in marR and soxS regulator genes that increase expression of the AcrAB efflux pump, increased expression of MdtEF and YdhE efflux pumps, and mutations in genes for inner core lipopolysacharide (LPS) synthesis. The LPS defects also cause reduced stability of the outer cell membrane and hypersensitivity to hydrophobic antibiotics such as novobiocin. How commonly such mutations accompany qnr in clinical isolates is not yet known. We examined a set of clinical E. coli strains containing qnrA, qnrB, and qnrS alleles for these associated mutations in comparison to their occurrence in control strains lacking qnr genes.

2. Materials and methods

2.1 Bacterial strains and susceptibility testing

Twenty-seven clinical E. coli isolates previously characterized for plasmid-mediated quinolone resistance genes were evaluated. Thirteen were qnr-positive and 14 were qnr-negative [[9], [9], [11], [12]]. The qnr-negative strains were collected at Seoul National University Hospital (South Korea) in the same time periods, in the intervals from 1998 to 2006, as qnr-positive strains (6–42, 6–75, 3–41, 3–48, 6–74). Other qnr-positive strains (4, 7, 10, 12, 29 and 76) came from a hospital in Shanghai (China) between March 2000 and March 2001. Each strain was from a different patient and had a unique plasmid profile [[9]]. E. coli J53 AziR [[13]] was used as a recipient in outcrosses with 100 µg/ml of sodium azide for counterselection plus either ciprofloxacin (0.5 µg/ml), ampicillin (100 µg/ml), kanamycin (25 µg/ml) or chloramphenicol (32 µg/ml). Plasmid DNAs from the resulting outcrosses were isolated with the Qiagen Plasmid MIDI kit (Qiagen, Valencia, CA).

MICs of ciprofloxacin and novobiocin (Sigma-Aldrich) were determined by agar dilution on Mueller-Hinton agar at 37°C with an inoculum of ~104 CFU following CLSI guidelines [[14]]. Susceptibility testing to 13 antimicrobial agents (amikacin, ampicillin, cefepime, cefotaxime, cefotetan, ceftazidime, chloramphenicol, gentamicin, kanamycin, streptomycin, tetracycline, tobramycin and trimethoprim-sulfamethoxazole (BD diagnostics) was performed for all strains by disk diffusion [[14]].

2.2 PCR and DNA sequencing

We amplified by PCR and sequenced the quinolone-resistance determining regions (QRDRs) of gyrA, gyrB, parC and parE genes, other plasmid-mediated quinolone resistance genes not previously studied in these strains, such as aac(6’)Ib-cr, oxqAB, qepA1, regulator genes such as acrR, marR, soxR, lipopolysaccharide biosynthesis genes rfaD, rfaE, rfaF and gmhB, and other genes surA and rpoB. DNA templates were prepared by boiling, and the primers used were described previously [[8], [12], [15], [16], [17]]. We used the Maxima Hot Start PCR Master Mix (Thermo Scientific, Waltham, MA) in a final volume of 50 µl. PCR products were purified using a PCR purification kit (Qiagen, Valencia, CA) and sent for sequencing by the Tufts University Core Facility, Boston, MA.

2.3 Relative expression of genes encoding efflux pumps and regulators

Reverse transcription followed by real-time quantitative PCR (RT-qPCR) was used to determine the expression levels of selected efflux pump genes, including acrA, mdtE, ydhE, and regulators such as marA and soxS. Comparison was made to expression of the housekeeping gene mdh. Primers used and specifications of RNA extraction and generation of cDNA were as previously described [[8]]. At least three different assays with three different RNA extractions were performed for each gene tested. E. coli J53 AziR was used as a reference to calculate relative expression.

2.4 Site-directed point mutations in RfaF in the E. coli chromosome

To evaluate the significance of alterations found in RfaF, we introduced single point mutations in the chromosomal rfaF gene of E. coli HS996 by site-directed mutagenesis using the Red®/ET® recombination system (Gene Bridges, Heidelberg). For verification of the correct rfaF mutations, PCR amplification and sequencing was employed as described previously [[8]].

3. Results

Thirteen qnr-positive and 14 qnr-negative clinical E. coli isolates were studied to determine if they contained gene mutations found in an earlier study evaluating events after selection of J53 pMG252 (qnrA1) mutants with increased ciprofloxacin resistance [[8]].

3.1 Susceptibility

Overall qnr-positive strains were more resistant to antimicrobial agents (besides ciprofloxacin and novobiocin) than qnr-negative strains (Supplementary table). Twelve of 13 qnr-positive strains were multiresistant (resistant to three or more classes of antimicrobials), while only 7 of 14 of the qnr-negative strains were multiresistant, and two qnr-negative strains were pansusceptible.

3.2 Quinolone resistance determinants

By themselves qnr genes provide a modest loss of ciprofloxacin susceptibility with MICs in E. coli of 0.25 – 0.5 µg/ml compared to an MIC of ~0.010 µg/ml for a fully susceptible strain. The ciprofloxacin MIC in the clinical strains containing qnr genes ranged from 0.125 to 1024 µg/ml (Table 1). Strains with ciprofloxacin MICs of 2 µg/ml and above had at least one mutation in gyrA as did one strain with an MIC of 0.5 µg/ml. Mutations in parC were present in strains with a ciprofloxacin MIC of at least 128 µg/ml. The six most ciprofloxacin-resistant strains contained additional plasmid-mediated quinolone resistance (PMQR) determinants. In particular, aac(6’)-Ib-cr was found in 5 of the 6 strains. Strain 76 was remarkable for a ciprofloxacin MIC of 1024 µg/ml. It contained qnrA1, aac(6’)-Ib-cr, oqxAB, and qepA1 in addition to mutations in gyrA, parC and parE genes. On outcross from this strain to E. coli J53 single transconjugants contained only oqxAB or qepA1, while transconjugants with all four PMQR genes located in a single plasmid had a ciprofloxacin MIC of 3–4 µg/ml. Nine of the 14 qnr-negative strains had mutations in topoisomerase genes, with the ciprofloxacin MIC reaching 256 µg/ml in a strain with qepA1. Another qnr-negative strain contained oqxAB. Interestingly, the 3 most resistant strains showed different mutations in ParE (D420N, S458A, and S458T). Mutations in the QRDR of GyrB were not detected in any strain.

Table 1.

Ciprofloxacin and novobiocin MICs of qnr-positive and qnr-negative strains with the quinolone resistance determinants found.

Strains MIC (µg/ml) PMQR Topoisomerase mutationsa Other
PMQR
Reference


Ciprofloxacin Novobiocin GyrA ParC ParE
6–42 0.125 40 qnrB1 WTb WT WT - [12]
EC35 0.25 80 qnrB6 WT WT WT - [11]
UAB4 0.5 40 qnrA1 WT WT WT - [9]
6–75 0.5 80 qnrB19 WT WT WT - [12]
3–41 0.5 80 qnrB4 S83L WT WT - [12]
3–48 1 80 qnrS1 WT WT WT - [12]
6–74 2 80 qnrS1 S83L WT WT - [12]
4 128 160 qnrA1 S83L, D87N S80I WT aac(6’)-Ib-cr [10]
29 128 160 qnrA1 S83L, D87N S80I WT aac(6’)-Ib-cr [10]
7 256 80 qnrA1 S83L, D87N S80I, E84G WT - [10]
10 256 320 qnrA1 S83L, D87N S80I WT aac(6’)-Ib-cr [10]
12 128 320 qnrA1 S83L, D87N S80I WT aac(6’)-Ib-cr [10]
76 1024 40 qnrA1 S83L, D87N S80I S458A aac(6’)-Ib-cr, qepA1, oqxAB [10]
3–62 0.016 320 - WT WT WT - [12]
6–66 0.016 40 - WT WT WT - [12]
3–25 0.03 80 - WT WT WT - [12]
3–7 0.06 80 - WT WT WT - [12]
3–33 0.06 80 - WT WT WT - [12]
6–52 0.25 80 - S83L WT WT - [12]
4–76 0.5 80 - S83L WT WT - [12]
5–58 0.5 80 - S83L E84G WT - [12]
5–66 1 80 - S83L S80I WT - [12]
5–65 4 40 - S83L S80I WT - [12]
5–81 32 80 - S83L, D87N S80I WT - [12]
6–13 64 160 - S83L, D87N S80I S458A oqxAB [12]
5–33 64 160 - S83L, D87N S80I D420N - [12]
5–59 256 160 - S83L, D87N S80I S458T qepA1 [12]
a

Mutations in gyrB were not detected.

b

Wild type.

3.3 Expression of efflux pumps and mutations in regulators

Eleven of the 13 qnr-positive strains had alterations in MarR, 9 had alterations in SoxR, and 5 had mutations in AcrR, using the E. coli K-12 strain MG1655 sequence for comparison (GenBank accession number U00096.3).

Among the MarR alterations, the S3N, G103S, Y137H amino acid changes are known not to affect repressor activity [[18]]. The remaining K62R mutation has been identified before in clinical isolates but showed no evidence of contributing to organic solvent tolerance, suggesting that it also is a silent mutation with regard to resistance [[19]]. Four qnr-positive strains had 2.4- to 8.5-fold increased marA expression without evident MarR alterations, suggesting a role for additional regulators. The T38S and G74R changes in SoxR have been previously identified in E. coli clinical isolates from Spain that had increased basal soxS expression [[20]]. However, no consistent increase in soxS expression was seen in our strains with these mutations. The SoxR A111T alteration observed in strain 76 also did not elevate soxS transcript levels, confirming earlier observations [[20]]. Five of the 10 mutations detected in AcrR caused frameshifts and were associated with 2.1- to 11.9 -fold increased expression of acrA. All the qnr-positive strains had elevated expression of at least one efflux pump gene, predominantly acrA (92% of the strains) followed by mdtE (54%) and ydhE (46%).

Similar mutations were found in the qnr-negative clinical E. coli isolates. Twelve had alterations in MarR, 10 in SoxR, and 5 in AcrR. One of the amino acid changes in MarR (D76G), one in SoxR (I140V), and 4 in AcrR (I113V, T213I, N214T and T32P) were not seen in the qnr-positive strains. Ten of the 14 strains had elevated expression of efflux pump genes with acrA (71.4%) predominating, followed by ydhE (50%) and mdtE (14.3%).

Recently, Pietsch et al., [[15]] identified mutations in rpoB, the gene coding for the β-subunit of RNA polymerase, as novel contributors to ciprofloxacin resistance via increased expression of the ydhE (also known as mdtK) efflux pump gene. We sequenced the entire rpoB gene for all strains with and without increase ydhE expression, but detected no rpoB mutations.

3.4. LPS defects and novobiocin susceptibility

E. coli mutants lacking heptose in the LPS core display a variety of phenotypes due to the reduced stability of the outer membrane, including hypersensitivity to hydrophobic antibiotics such as novobiocin. In our strains the novobiocin MICs ranged from 40 to 320 µg/ml with somewhat lower values than previously seen in laboratory strains [[8]]. In an attempt to find out if the strains had defects on the LPS pathway we sequenced rfaD, rfaE, rfaF and gmhB genes and using the E. coli K-12 strain MG1655 sequence for comparison. Two qnr-negative strains had a RfaE (A245T) mutation, 9 of 13 qnr-positive strains had alterations in Rfa as did all the qnr-negative strains, and 2 qnr-positive strains had 2 mutations in GmhB. (Table 3).

Table 3.

LPS core biosynthesis mutations detected.

Protein Mutation detected Strains
RfaE A245T 3–25, 5–58
RfaF I136V 6–42, 4, 29, 7, 10, 12, 6–66, 3–7, 3–33, 6–52, 5–66, 5–65, 5–81, 5–59
I136V, I144V, A348V 6–75, 6–74
I136V, A141S 76, 3–62, 5–33
I136V, V335A 3–25, 4–76, 5–58
I136V, Y287F 6–13
GmhB V145A, I161V, Q191Stop UAB4
V145A, Q191Stop 3–48
SurA P346A 3–48

Defects in these genes are known to cause novobiocin susceptibility, but the specific mutations we observed have not been published. We created point mutations (I136V and A141S) in the rfaF gene of E. coli HS996 by site-directed mutagenesis. The MICs for ciprofloxacin (0.008 µg/ml) and novobiocin (640 µg/ml) were unchanged in the mutants indicating that the I36V and A141S changes are functionally silent.

Two qnr-positive strains had frameshift mutations, along with point mutations, in the gmhB gene, and one of these strains also had a mutation in surA, but novobiocin hypersensitivity was not observed (Table 3). The reason to sequence the surA gene was that SurA is the primary periplasmic molecular chaperone that facilitates the folding and assembling of outer membrane proteins in Gram-negative bacteria, and SurA-deficient cells are more susceptible to hydrophobic drugs [[21]].

4. Discussion

Many studies have evaluated the gyrA and parC mutations that accompany qnr in clinical isolates [[1], [2], [3], [4], [22], [23], [24], [25]] and a few investigations have examined expression of efflux genes in quinolone-resistant strains without qnr [[26], [27], [28], [29]], but this study is the first to evaluate both topoisomerase mutations and pump expression in qnr-containing clinical isolates.

The strains were selected by the presence or absence of qnr genes, but other PMQR genes were also evaluated, with the finding that aac(6’)Ib-cr was present in 5 of 13 qnr-positive but none of the qnr-negative strains. In both groups non-qnr PMQR determinants were found in the most ciprofloxacin-resistant isolates and were associated with mutations in GyrA and ParC. Interestingly, three different mutations in ParE were also detected in this study, only one of which (D420N) was in a known QRDR region [[17]]. The D420N mutation has been described to increase the ciprofloxacin MIC in Vibrio cholerae, [[30]] but to our knowledge, this is the first time that it has been found in E. coli. The ParE S458A and S458T mutations have been described before, and the S458A alteration has been associated with high levels of resistance to fluoroquinolones in E. coli [[31], [32], [33]]. Such topoisomerase mutations were the main contributors to the high ciprofloxacin MICs of the clinical strains. Outcross of the plasmid from E. coli strain 76 with a ciprofloxacin MIC of 1024 µg/ml produced a transconjugant with a ciprofloxacin MIC of only 4 µg/ml, indicating the contribution of chromosomal mutations to the high level ciprofloxacin resistance observed. The plasmid carried qnrA1, aac(6’)-Ib-cr, oqxAB, and qepA1 and thus demonstrated that a strain with the combination of four PMQR genes can reach the current CLSI breakpoint for ciprofloxacin resistance (≥ 4µg/ml) in the absence of topoisomerase mutations or efflux pump overexpression.

Eleven of the qnr-negative strains had elevated marA expression, but only 4 of the qnr-positive strains; in both sets of strains, AcrA was the predominant overexpressed efflux pump. Increased MarA levels not only increase ciprofloxacin MICs but also protect bacteria from the bactericidal effect of the fluoroquinolones [[34]]. Most of the mutations found in marR in this study were already published as not contributing to resistance. Marcusson et al., [[35]] reported that mutations in acrR and marR were associated with a significant fitness cost. Therefore, increased AcrA levels might have arisen to take advantage of MarA–regulated bactericidal protection by mechanisms other than alteration in MarR or AcrR because of their associated cost. Existing data also support a model in which AcrA plays a role in quorum sensing by emitting quorum-sensing signals, so it is also possible that clinical strains increase levels of AcrA because doing so confers a fitness advantage not directly related to drug efflux [[26]].

There were only 3 qnr-negative strains that did not overexpress any efflux pump studied (their ciprofloxacin MICs ranged from 0.06 to 1 µg/ml). The rest of the qnr-negative and qnr-positive strains overexpressed acrA, mdtE, and ydhE separately or in combination. It has been seen before that overexpression of each of the three pumps separately resulted in roughly similar levels of quinolone resistance, whereas simultaneous overexpression of mdtE or ydhE in combination with acrA gave synergistic increases in quinolone resistance [[28], [29]]. In our data, there was only one qnr-positive strain that overexpressed mdtE by itself (ciprofloxacin MIC of 0.5 µg/ml) and only one qnr-negative strain that overexpressed the single pump ydhE but had a borderline increase in expression of acrA (a 1.9-fold change). This finding suggests that all the combinations of expression of the three efflux pumps contribute to the different levels of resistance detected.

Regarding the LPS defects and novobiocin susceptibility, we identified several new mutations in rfaE, rfaF and gmhB genes. Heptose biosynthesis in E. coli is a process which involves several enzymes (GmhA, RfaE, GmhB and RfaD) that act in different steps in the pathway. Interruption of biosynthesis or transport of heptose causes a heptose-less phenotype called “deep-rough” that is unstable and leads to increase susceptibility to hydrophobic compounds like novobiocin [[36]]. We recreated the most frequent point mutations detected (I136V and A141S in the rfaF gene) but found no associated phenotype. A stop codon was also detected in gmhB in two strains that nonetheless lacked novobiocin hypersensitivity. In a laboratory strain of E. coli deletion of gmhB encoding D-α,β-D-heptose 1,7-bisphosphate phosphatase did not confer a complete heptose-less LPS core phenotype, suggesting the presence of another as yet unidentified phosphatase activity that can partially compensate in the synthesis of a complete core [[37]].

5. Conclusions

Many qnr-containing clinical isolates of E. coli had in addition to topoisomerase mutations increased expression of efflux pump genes, especially acrA. Similar efflux gene overexpression was also seen in control strains lacking qnr. Disabling mutations in core LPS synthesis, such as were detected on selecting more ciprofloxacin-resistant derivatives from a qnr-containing laboratory strain, were not found in the clinical isolates. Other PMQR genes often accompanied qnr in clinical strains. In one such E. coli strain with gyrA, parC and parE topoisomerase mutations, overexpression of AcrA, and four PMQR genes, the ciprofloxacin MIC reached 1024 µg/ml. Bacteria evidently have many ways to achieve protection from a lethal agent such as fluoroquinolone, and a diversity of resistance mechanisms occurs in clinical isolates with and without qnr genes.

Supplementary Material

supplement

Table 2.

Mutations in MarR, SoxR and AcrR and relative expression of regulators and efflux pumps in clinical strains.

Strains Mutated protein (mutation detected) Mean fold change in expression relative to J53 strain (SEM)b

marA soxS acrA ydhE mdtE
6–42 MarR (S3N, G103S, Y137H); SoxR (G74R) 4.18 (1.89) 10.30 (1.95) 15.55 (4.18) 5.73 (1.80) 5.49 (1.75)
EC35 MarR (G103S, Y137H); SoxR (G74R) 1.03 (0.17) 1.26 (0.34) 2.61 (0.47) 1.69 (0.47) 1.39 (0.43)
UAB4 MarR (S3N, G103S, Y137H); SoxR (G74R) 2.35 (0.15) 1.47 (0.13) 1.89 (0.005) 1.44 (0.49) 2.70 (1.01)
6–75 MarR (G103S, Y137H) 4.08 (2.28) 2.63 (0.82) 3.87 (0.94) 2.69 (0.65) 2.35 (0.71)
3–41 - 0.86 (0.16) 1.41 (0.41) 3.61 (0.63) 1.45 (0.54) 2.19 (0.53)
3–48 MarR (G103S, Y137H) 1.77 (0.35) 2.47 (0.77) 4.20 (0.68) 2.61 (0.63) 2.29 (0.65)
6–74 MarR (G103S, Y137H) 8.53 (3.65) 4.50 (0.46) 6.82 (1.70) 3.30 (0.85) 2.04 (0.47)
4 MarR (K62R, G103S, Y137H); SoxR (T38S, G74R); AcrR (L109fsa) 0.89 (0.31) 1.67 (0.63) 5.14 (1.55) 0.65 (0.12) 0.38 (0.13)
29 MarR (K62R, G103S, Y137H); SoxR (T38S, G74R); AcrR (L109fs) 0.63 (0.11) 2.83 (0.50) 9.59 (2.59) 4.05 (0.24) 0.75 (0.09)
7 MarR (K62R, G103S, Y137H); SoxR (T38S, G74R); AcrR (E190fs) 0.53 (0.15) 1.92 (0.53) 2.10 (0.18) 0.30 (0.18) 0.40 (0.10)
10 MarR (K62R, G103S, Y137H); SoxR (T38S, G74R); AcrR (L109fs) 0.50 (0.13) 1.83 (0.44) 5.52 (1.14) 1.15 (0.32) 1.19 (0.53)
12 MarR (K62R, G103S, Y137H); SoxR (T38S, G74R); AcrR (L109fs) 1.75 (0.65) 4.15 (0.14) 11.92 (1.08) 3.08 (0.84) 4.38 (1.06)
76 SoxR (A111T) 0.79 (0.36) 1.01 (0.28) 2.01 (0.50) 1.00 (0.39) 0.64 (0.15)
3–62 - 2.59 (0.37) 3.11 (1.10) 8.15 (4.43) 6.76 (3.15) 3.62 (1.62)
6–66 MarR (3SN, G103S, Y137H); SoxR (G74R) 19.56 (8.09) 8.95 (3.73) 27.20 (0.85) 21.43 (6.05) 2.99 (1.56)
3–25 MarR (G103S, Y137H); SoxR (T38S, G74R); AcrR (T213I, N214T) 2.96 (0.64) 2.88 (0.65) 6.21 (1.93) 6.67 (1.19) 1.57 (0.73)
3–7 MarR (K62R, G103S, Y137H); SoxR (T38S, G74R); AcrR (I113V, T213I, N214T) 2.72 (0.06) 1.40 (0.28) 1.27 (0.70) 1.03 (0.86) 0.79 (0.15)
3–33 MarR (G103S, Y137H); SoxR (I40V, G74R) 6.65 (1.63) 3.31 (1.76) 1.68 (1.37) 0.68 (0.51) 1.17 (0.11)
6–52 MarR (K62R, G103S, Y137H); SoxR (T38S, G74R) 6.18 (2.70) 2.71 (0.75) 6.53 (2.57) 0.91 (0.16) 0.70 (0.28)
4–76 MarR (G103S, Y137H); SoxR (T38S, G74R); AcrR (T213I, N214T) 5.95 (0.66) 4.74 (0.75) 6.37 (1.09) 3.66 (0.15) 1.06 (0.50)
5–58 MarR (G103S, Y137H); SoxR (T38S, G74R); AcrR (T213I, N214T) 2.78 (0.99) 1.55 (0.560) 6.95 (2.63) 1.56 (1.35) 0.65 (0.18)
5–66 MarR (K62R, G103S, Y137H); SoxR (T38S, G74R) 3.25 (0.57) 1.96 (0.30) 1.92 (0.67) 0.88 (0.08) 0.35 (0.06)
5–65 MarR (K62R, G103S, Y137H); SoxR (T38S, G74R) 9.06 (4.21) 4.09 (0.73) 6.43 (3.33) 6.97 (4.52) 1.31 (0.72)
5–81 MarR (D76G, G103S, Y137H) 2.30 (0.67) 1.09 (0.19) 2.05 (0.85) 3.72 (2.16) 1.612 (0.52)
6–13 MarR (G103S, Y137H) 1.17 (0.84) 3.70 (1.04) 2.52 (1.42) 1.36 (0.31) 1.47 (1.13)
5–33 SoxR (A111T); AcrR (T32P) 0.96 (0.60) 2.39 (0.32) 3.11 (0.70) 1.34 (0.43) 0.95 (0.48)
5–59 MarR (G103S, Y137H) 0.73 (0.51) 1.25 (0.36) 1.98 (1.05) 2.26 (1.42) 1.12 (0.24)
a

fs, frameshift

b

Values in boldface type differ from values for the parent by at least 2-fold.

SEM= error standard of the mean.

Highlights.

  • Elevated expression of efflux pumps accompanies topoisomerase mutations.

  • AcrA is the predominant overexpressed efflux pump.

  • Other PMQR genes often accompanied qnr in clinical strains.

  • In one E. coli clinical strain the ciprofloxacin MIC reached 1024 µg/ml.

  • A diversity of resistance mechanisms occurs in clinical isolates with and without qnr genes.

Acknowledgments

Funding: This work was supported by grant 5R01AI057576-13 (to David C. Hooper) from the National Institutes of Health, U.S. Public Health Service.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Competing Interests: No conflict of interest

Ethical Approval: Not applicable

References

  • 1.Lascols C, Robert J, Cattoir V, Bébéar C, Cavallo JD, Podglajen I, Ploy MC, Bonnet R, Soussy CJ, Cambau E. Type II topoisomerase mutations in clinical isolates of Enterobacter cloacae and other enterobacterial species harbouring the qnrA gene. Int J Antimicrob Agents. 2007;29:402–409. doi: 10.1016/j.ijantimicag.2006.11.008. [DOI] [PubMed] [Google Scholar]
  • 2.Briales A, Rodriguez-Martinez JM, Velasco C, de Alba PD, Rodriguez-Bano J, Martinez-Martinez L, Pascual A. Prevalence of plasmid-mediated quinolone resistance determinants qnr and aac(6')-Ib-cr in Escherichia coli and Klebsiella pneumoniae producing extended-spectrum β-lactamases in Spain. Int J Antimicrob Agents. 2012;39:431–434. doi: 10.1016/j.ijantimicag.2011.12.009. [DOI] [PubMed] [Google Scholar]
  • 3.Silva-Sánchez J, Cruz-Trujillo E, Barrios H, Reyna-Flores F, Sánchez-Pérez A, Consortium BR, Garza-Ramos U. Characterization of plasmid-mediated quinolone resistance (PMQR) genes in extended-spectrum β-lactamase-producing Enterobacteriaceae pediatric clinical isolates in Mexico. PLoS One. 2013;8:e77968. doi: 10.1371/journal.pone.0077968. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Xue G, Li J, Feng Y, Xu W, Li S, Yan C, Zhao H, Sun H. High prevalence of plasmid-mediated quinolone resistance determinants in Escherichia coli and Klebsiella pneumoniae isolates from pediatric patients in China. Microb Drug Resist. 2017;23:107–114. doi: 10.1089/mdr.2016.0004. [DOI] [PubMed] [Google Scholar]
  • 5.Martínez-Martínez L, Pascual A, García I, Tran J, Jacoby GA. Interaction of plasmid and host quinolone resistance. J Antimicrob Chemother. 2003;51:1037–1039. doi: 10.1093/jac/dkg157. [DOI] [PubMed] [Google Scholar]
  • 6.Cesaro A, Bettoni RR, Lascols C, Merens A, Soussy CJ, Cambau E. Low selection of topoisomerase mutants from strains of Escherichia coli harbouring plasmid-borne qnr genes. J Antimicrob Chemother. 2008;61:1007–1015. doi: 10.1093/jac/dkn077. [DOI] [PubMed] [Google Scholar]
  • 7.Goto K, Kawamura K, Arakawa Y. Contribution of QnrA, plasmid-mediated quinolone resistance peptide, to survival of Escherichia coli exposed to lethal ciprofloxacin concentration. Jpn J Infect Dis. 2015;68:196–202. doi: 10.7883/yoken.JJID.2014.153. [DOI] [PubMed] [Google Scholar]
  • 8.Vinué L, Corcoran MA, Hooper DC, Jacoby GA. Mutations that enhance the ciprofloxacin resistance of Escherichia coli with qnrA1. Antimicrob Agents Chemother. 2016;60:1537–1545. doi: 10.1128/AAC.02167-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Jacoby GA, Chow N, Waites KB. Prevalence of plasmid-mediated quinolone resistance. Antimicrob Agents Chemother. 2003;47:559–562. doi: 10.1128/AAC.47.2.559-562.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Wang M, Tran JH, Jacoby GA, Zhang Y, Wang F, Hooper DC. Plasmid-mediated quinolone resistance in clinical isolates of Escherichia coli from Shanghai, China. Antimicrob Agents Chemother. 2003;47:2242–2248. doi: 10.1128/AAC.47.7.2242-2248.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Jacoby GA, Walsh KE, Mills DM, Walker VJ, Oh H, Robicsek A, Hooper DC. qnrB, another plasmid-mediated gene for quinolone resistance. Antimicrob Agents Chemother. 2006;50:1178–1182. doi: 10.1128/AAC.50.4.1178-1182.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Kim HB, Park CH, Kim CJ, Kim EC, Jacoby GA, Hooper DC. Prevalence of plasmid-mediated quinolone resistance determinants over a 9-year period. Antimicrob Agents Chemother. 2009;53:639–645. doi: 10.1128/AAC.01051-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Jacoby GA, Han P. Detection of extended-spectrum β-lactamases in clinical isolates of Klebsiella pneumoniae and Escherichia coli. J Clin Microbiol. 1996;34:908–911. doi: 10.1128/jcm.34.4.908-911.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.CLSI. CLSI document M07-A9. Wayne, PA: Clinical and Laboratory Standards Institute; 2012. Methods for dilution antimicrobial susceptibility tests for bacteria that grow aerobically; approved standard - ninth edition. [Google Scholar]
  • 15.Pietsch F, Bergman JM, Brandis G, Marcusson LL, Zorzet A, Huseby DL, Hughes D. Ciprofloxacin selects for RNA polymerase mutations with pleiotropic antibiotic resistance effects. J Antimicrob Chemother. 2017;72:75–84. doi: 10.1093/jac/dkw364. [DOI] [PubMed] [Google Scholar]
  • 16.Vila J, Ruiz J, Marco F, Barcelo A, Goñi P, Giralt E, Jimenez de Anta T. Association between double mutation in gyrA gene of ciprofloxacin-resistant clinical isolates of Escherichia coli and MICs. Antimicrob Agents Chemother. 1994;38:2477–2479. doi: 10.1128/aac.38.10.2477. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Ruiz J, Casellas S, Jimenez de Anta MT, Vila J. The region of the parE gene, homologous to the quinolone-resistant determining region of the gyrB gene, is not linked with the acquisition of quinolone resistance in Escherichia coli clinical isolates. J Antimicrob Chemother. 1997;39:839–840. doi: 10.1093/jac/39.6.839. [DOI] [PubMed] [Google Scholar]
  • 18.Oethinger M, Podglajen I, Kern WV, Levy SB. Overexpression of the marA or soxS regulatory gene in clinical topoisomerase mutants of Escherichia coli. Antimicrob Agents Chemother. 1998;42:2089–2094. doi: 10.1128/aac.42.8.2089. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Park YH, Yoo JH, Huh DH, Cho YK, Choi JH, Shin WS. Molecular analysis of fluoroquinolone-resistance in Escherichia coli on the aspect of gyrase and multiple antibiotic resistance (mar) genes. Yonsei Med J. 1998;39:534–540. doi: 10.3349/ymj.1998.39.6.534. [DOI] [PubMed] [Google Scholar]
  • 20.Koutsolioutsou A, Pena-Llopis S, Demple B. Constitutive soxR mutations contribute to multiple-antibiotic resistance in clinical Escherichia coli isolates. Antimicrob Agents Chemother. 2005;49:2746–2752. doi: 10.1128/AAC.49.7.2746-2752.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Zhong M, Ferrell B, Lu W, Chai Q, Wei Y. Insights into the function and structural flexibility of the periplasmic molecular chaperone SurA. J Bacteriol. 2013;195:1061–1067. doi: 10.1128/JB.01143-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Musumeci R, Rausa M, Giovannoni R, Cialdella A, Bramati S, Sibra B, Giltri G, Viganò F, Cocuzza CE. Prevalence of plasmid-mediated quinolone resistance genes in uropathogenic Escherichia coli isolated in a teaching hospital of northern Italy. Microb Drug Resist. 2012;18:33–41. doi: 10.1089/mdr.2010.0146. [DOI] [PubMed] [Google Scholar]
  • 23.Kao CY, Wu HM, Lin WH, Tseng CC, Yan JJ, Wang MC, Teng CH, Wu JJ. Plasmid-mediated quinolone resistance determinants in quinolone-resistant Escherichia coli isolated from patients with bacteremia in a university hospital in Taiwan, 2001–2015. Sci Rep. 2016;6:32281. doi: 10.1038/srep32281. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Albornoz E, Lucero C, Romero G, Quiroga MP, Rapoport M, Guerriero L, Andres P, Rodriguez C, Galas M, Centron D, Corso A, Petroni A. Prevalence of plasmid-mediated quinolone resistance genes in clinical enterobacteria from Argentina. Microb Drug Resist. 2017;23:177–187. doi: 10.1089/mdr.2016.0033. [DOI] [PubMed] [Google Scholar]
  • 25.Nazir H, Cao S, Hasan F, Hughes D. Can phylogenetic type predict resistance development? J Antimicrob Chemother. 2011;66:778–787. doi: 10.1093/jac/dkq505. [DOI] [PubMed] [Google Scholar]
  • 26.Singh R, Swick MC, Ledesma KR, Yang Z, Hu M, Zechiedrich L, Tam VH. Temporal interplay between efflux pumps and target mutations in development of antibiotic resistance in Escherichia coli. Antimicrob Agents Chemother. 2012;56:1680–1685. doi: 10.1128/AAC.05693-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Morgan-Linnell SK, Becnel Boyd L, Steffen D, Zechiedrich L. Mechanisms accounting for fluoroquinolone resistance in Escherichia coli clinical isolates. Antimicrob Agents Chemother. 2009;53:235–241. doi: 10.1128/AAC.00665-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Yang S, Clayton SR, Zechiedrich EL. Relative contributions of the AcrAB, MdfA and NorE efflux pumps to quinolone resistance in Escherichia coli. J Antimicrob Chemother. 2003;51:545–556. doi: 10.1093/jac/dkg126. [DOI] [PubMed] [Google Scholar]
  • 29.Swick MC, Morgan-Linnell SK, Carlson KM, Zechiedrich L. Expression of multidrug efflux pump genes acrAB-tolC, mdfA and norE in Escherichia coli clinical isolates as a function of fluoroquinolone and multidrug resistance. Antimicrob Agents Chemother. 2011;55:921–924. doi: 10.1128/AAC.00996-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Zhou Y, Yu L, Li J, Zhang L, Tong Y, Kan B. Accumulation of mutations in DNA gyrase and topoisomerase IV genes contributes to fluoroquinolone resistance in Vibrio cholerae O139 strains. Int J Antimicrob Agents. 2013;42:72–75. doi: 10.1016/j.ijantimicag.2013.03.004. [DOI] [PubMed] [Google Scholar]
  • 31.Moon DC, Seol SY, Gurung M, Jin JS, Choi CH, Kim J, Lee YC, Cho DT, Lee JC. Emergence of a new mutation and its accumulation in the topoisomerase IV gene confers high levels of resistance to fluoroquinolones in Escherichia coli isolates. Int J Antimicrob Agents. 2010;35:76–79. doi: 10.1016/j.ijantimicag.2009.08.003. [DOI] [PubMed] [Google Scholar]
  • 32.Komp Lindgren P, Karlsson A, Hughes D. Mutation rate and evolution of fluoroquinolone resistance in Escherichia coli isolates from patients with urinary tract infections. Antimicrob Agents Chemother. 2003;47:3222–3232. doi: 10.1128/AAC.47.10.3222-3232.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Sorlozano A, Gutierrez J, Jimenez A, de Dios Luna J, Martínez JL. Contribution of a new mutation in parE to quinolone resistance in extended-spectrum-β-lactamase-producing Escherichia coli isolates. J Clin Microbiol. 2007;45:2740–2742. doi: 10.1128/JCM.01093-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Goldman JD, White DG, Levy SB. Multiple antibiotic resistance (mar) locus protects Escherichia coli from rapid cell killing by fluoroquinolones. Antimicrob Agents Chemother. 1996;40:1266–1269. doi: 10.1128/aac.40.5.1266. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Marcusson LL, Frimodt-Møller N, Hughes D. Interplay in the selection of fluoroquinolone resistance and bacterial fitness. PLoS Pathog. 2009;5:e1000541. doi: 10.1371/journal.ppat.1000541. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Linkevicius M, Sandegren L, Andersson DI. Mechanisms and fitness costs of tigecycline resistance in Escherichia coli. J Antimicrob Chemother. 2013;68:2809–2819. doi: 10.1093/jac/dkt263. [DOI] [PubMed] [Google Scholar]
  • 37.Kneidinger B, Marolda C, Graninger M, Zamyatina A, McArthur F, Kosma P, Valvano MA, Messner P. Biosynthesis pathway of ADP-L-glycero-beta-D-manno-heptose in Escherichia coli. J Bacteriol. 2002;184:363–369. doi: 10.1128/JB.184.2.363-369.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

supplement

RESOURCES