Next Article in Journal
Evaluation of Everolimus Activity against Mycobacterium tuberculosis Using In Vitro Models of Infection
Next Article in Special Issue
Phenotypic and Genotypic Antimicrobial Susceptibility Testing of Chlamydia trachomatis Isolates from Patients with Persistent or Clinical Treatment Failure in Spain
Previous Article in Journal
Multidrug-Resistant ESBL-Producing E. coli in Clinical Samples from the UK
Previous Article in Special Issue
Increases in the Macrolide Resistance of Mycoplasma genitalium and the Emergence of the A2058T Mutation in the 23S rRNA Gene: Clonal Spread?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Azithromycin Susceptibility Testing and Molecular Investigation of Neisseria gonorrhoeae Isolates Collected in Russia, 2020–2021

1
Center for Precision Genome Editing and Genetic Technologies for Biomedicine, Engelhardt Institute of Molecular Biology, Russian Academy of Sciences, Moscow 119991, Russia
2
State Research Center of Dermatovenerology and Cosmetology, Russian Ministry of Health, Moscow 107076, Russia
*
Author to whom correspondence should be addressed.
Antibiotics 2023, 12(1), 170; https://doi.org/10.3390/antibiotics12010170
Submission received: 13 December 2022 / Revised: 11 January 2023 / Accepted: 12 January 2023 / Published: 13 January 2023
(This article belongs to the Special Issue Antibiotic Resistance of Sexually Transmitted Pathogens)

Abstract

:
The aim of this work was to study the resistance to macrolides (azithromycin) in the modern Russian population of N. gonorrhoeae with the analysis of genetic resistance determinants. Azithromycin is not used to treat gonococcal infection in Russia. However, among 162 isolates collected in 2020–2021, 22 isolates (13.6%) were phenotypically resistant to azithromycin. Mutations in 23S rRNA genes were found only in two isolates; erm and mefA genes were absent. Azithromycin resistance was shown to be predominantly associated with mutations in the mtrR and mtrD genes of the MtrCDE efflux pump and their mosaic alleles which may have formed due to a horizontal transfer from N. meningitidis. A total of 30 types of mtrR alleles and 10 types of mtrD alleles were identified including mosaic variants. Matching between the mtrR and mtrD alleles was revealed to indicate the cooperative molecular evolution of these genes. A link between the mtrR and mtrD alleles and NG-MAST types was found only for NG-MAST 228 and 807, typical of N. gonorrhoeae in Russia. The high level of resistance to azithromycin in Russia may be related to the spread of multiple transferable resistance to antimicrobials regardless of their use in the treatment of gonococcal infection.

1. Introduction

Gonococcal infection is one of the most dangerous and widespread sexually transmitted infections (STIs). The causative agent of gonococcal infection is N. gonorrhoeae, a Gram-negative diplococcus, that is characterized by a high rate of both phenotypic and genetic adaptive changes. According to WHO data, about 82.4 million cases of gonorrhea were newly detected in 2020 [1]. Of particular concern is the ability of N. gonorrhoeae to rapidly accumulate mutations in the genome and acquire resistance to antimicrobial drugs [2,3].
Currently, most countries of the world use third-generation cephalosporins together with azithromycin as a dual antimicrobial therapy for the treatment of gonococcal infection [4,5,6]. Azithromycin is a macrolide class antibiotic used against a wide range of bacteria, including gonococcus. Macrolides cause the disruption of protein synthesis on the ribosomes of a microbial cell by binding to the 50S ribosomal subunit. The main binding sites are located at positions A2058, A2059, and C2611 in domains II and V of the 23S rRNA [7]. The expediency of the joint use of cephalosporins and azithromycin, which have different mechanisms of action, is explained by the increase in the efficacy of the therapy of gonococcal infection and by the eradication of another, most frequently detected STI pathogen, Chlamydia trachomatis, which is typically found as part of a mixed infection [6]. Gonococci with a reduced susceptibility to cephalosporins are found in the world population sporadically (0.1–2%), while the proportion of azithromycin-resistant N. gonorrhoeae isolates grows every year, and in some countries of Europe and Asia, it is close to 5%, which is the threshold value for the exclusion of this drug from treatment regimens in accordance with WHO recommendations [8,9].
The mechanisms of N. gonorrhoeae resistance to macrolides include specific chromosomal mutations, the presence of genes encoding rRNA methyltransferases, and the overexpression of efflux pumps [10]. Antibiotic-specific genetic determinants of resistance are mutations in the peptidyltransferase loop of the II and V domains of the 23S rRNA [11]. The C2611T substitution leads to the appearance of isolates with moderate resistance (minimum inhibitory concentration, MICAzm up to 4 mg/L), while the A2058G/A2059G substitution leads to highly resistant isolates (MICAzm > 256 mg/L) [12]. It should be especially noted that the level of resistance is also determined by the number of mutant 23S rRNA alleles. The N. gonorrhoeae genome contains four copies of the rrn operon (rrnA, rrnB, rrnC, rrnD), which include the nucleotide sequence encoding 23S rRNA. In particular, isolates with three or all four mutant alleles have been shown to be highly resistant (MICAzm ≥ 256 mg/L and up to 4096 mg/L), while isolates with only one copy of the mutant 23S rRNA allele have an insignificant level of resistance (MICAzm = 0.25–1 mg/L) [13], i.e., below or at the threshold separating resistant and susceptible forms (ECOFF MICAzm ≤ 1 mg/L) according to the criterion of the 2022 European Committee on Antimicrobial Susceptibility Testing (EUCAST) [14].
Other mechanisms of N. gonorrhoeae resistance are associated with efflux systems that ensure the outflow of antimicrobial drugs, including macrolides, from the cell. The MtrCDE efflux system is encoded by the mtrC-mtrD-mtrE operon. Its activation is associated either with changes in the functioning of the MtrCDE transcriptional repressor, the mtrR gene [15,16], or with changes in the configuration of MtrCDE proteins [17]. For the mtrR gene, the Ala39Thr, Arg44His, Gly45Asp, and Ala86Thr substitutions, and the delA deletion at position −35 in the promoter region and/or insertion of the insT/instTT nucleotides at position −10 of mtrR leads to an increase in the expression of the MtrCDE efflux pump and the removal of antimicrobial drugs, including macrolides, out of the cell. Isolates with such mutations have a reduced phenotypic susceptibility to azithromycin. Mosaic alleles of the MtrCDE efflux pump operon appear as a result of a horizontal transfer between closely related Neisseria species, such as N. meningitidis, N. sicca, N. lactamica, N. cinerea, N. flavescence, etc. Mosaic variants of the mtrR and mtrD genes are capable of activating the MtrCDE efflux system and, respectively, the removal of azithromycin from the cell, which leads to an increase in the MICAzm to 8 mg/L or more [17]. Mosaic variants of the promoter region of the mtrR gene (Meningitidis-like (MG-like) promoter) have also been described; as a result of mosaicity, the repression of the entire MtrCDE operon may be impaired, which leads to an increase in the MICAzm up to 8 mg/L [18,19].
The outflow of macrolides from the bacterial cell and a decrease in susceptibility to macrolides is also associated with the gonococcus efflux pump encoded by the mefA conjugative gene. It has been shown that the mef gene, which was originally found in Gram-positive microorganisms, can be transferred and expressed in Gram-negative microorganisms, including N. gonorrhoeae [20,21].
Another factor affecting the susceptibility of N. gonorrhoeae to macrolides is the synthesis of 23S rRNA methyltransferases ErmA, ErmB, ErmC, and ErmF. These enzymes cause methylation of the adenine A2058 in the peptidyltransferase loop of the V domain of the 23S rRNA, that leads to the modification of the antibiotic target, a decrease in binding affinity, and the formation of azithromycin resistance (MICAzm ≥ 8–64 mg/L) [11]. The presence of erm genes is associated with an increase in the level of phenotypic susceptibility of N. gonorrhoeae to macrolides; however, at present, 23S rRNA methylases are rarely found in resistant isolates [22], for example, the PubMLST database includes only three isolates with erm genes.
In the Russian Federation, azithromycin has never been used for the treatment of gonococcal infections because the proportion of resistant isolates has consistently exceeded the WHO recommended threshold of 5% since the starting of large-scale screening for susceptibility to this antibiotic in 2007 [23]. At the same time, the reasons for the decrease in the susceptibility of N. gonorrhoeae to azithromycin in the Russian population when this antibiotic is not used in gonorrhea treatment regimens are unclear, and the mechanisms that determine azithromycin resistance have not been studied in detail.
The goal of this work was to study the resistance of N. gonorrhoeae to macrolides (azithromycin) in the modern Russian population of the pathogen (2020–2021) with the analysis of the genetic determinants of resistance and the reasons for their spread.

2. Results

2.1. Phenotypic Azithromycin Susceptibility and Analysis of Genetic Determinants Associated with Azithromycin Resistance

A total of 162 N. gonorrhoeae clinical isolates collected in different regions of the Russian Federation in 2020–2021 were analyzed; MICAzm and genetic determinants associated with azithromycin resistance were determined. Among these isolates, 22 (13.6%) were found to be azithromycin-resistant (MICAzm > 1 mg/L, up to 8 mg/L) and 140 (86.4%) isolates were azithromycin-susceptible (MICAzm = 0.01–1.0 mg/L).
Only two isolates carried mutations in the 23S rRNA gene. MICAzm of both isolates was 4 mg/L, which was due to C2611T mutations in all four copies of the 23S rRNA gene. In 160 isolates (98.7%), mutations in the MtrCDE efflux pump genes were found, leading to amino acid substitutions both in the mtrR gene and its promoter region, and in the mtrD gene. In some cases, an increase in the expression of the MtrCDE efflux pump and, as a consequence, the formation of resistance (MICAzm ≈ 2 mg/L) were due to the mosaic structure of the mtrR and/or mtrD genes. No isolates containing the erm genes encoding Erm methyltransferases, and the mefA efflux pump gene were found. Characteristics of the studied isolates including mtrR and mtrD allele types, NG-MAST type, MICAzm, mutations in the mtrR, mtrD, and 23S rRNA genes, are provided in Supplementary Materials, Table S1.
Thus, the resistance to azithromycin in the analyzed sample of N. gonorrhoeae isolates was only sporadically associated with mutations in the 23S rRNA gene and was mainly formed as a result of an increase in the expression of the MtrCDE efflux pump, and this was the rationale for carrying out in-depth analyses of the mtrR and mtrD genes.

2.2. Diversity of the mtrR Gene Alleles

The types of mtrR alleles were determined for the first time in the Russian population of the causative agent of gonorrhea. A total of 30 different types of mtrR alleles were identified in 162 clinical isolates (Figure 1a). Among them, there were no wild type (NG-STAR “Allele 353”) alleles.
Most of the alleles had 2–4 amino acid substitutions compared with the wild type sequence. The most common substitutions were Ala39Thr (29.0%), Gly45Asp (6.2%), Gly45Ser (7.4%), Ala86Thr (89.5%), and His105Tyr (17.9%), either separately or in various combinations. In total, 43 isolates with mutations in the promoter region of the mtrR gene were identified; all of these isolates had an adenine deletion at position −35 from the start codon (-35delA), and one isolate additionally had a thymine insertion at position −10 (-10insT).
Of particular note is that 26 isolates (16%) had alleles associated with the mosaic structure of the mtrR gene which were probably formed as a result of horizontal transfer from N. meningitidis (NG-STAR Alleles “485”; “436”; “520”). A mosaic structure in these alleles was found both in the coding sequence of the mtrR gene (from four to six amino acid substitutions) and in its promoter region (MG-like promoter) [18].
Five novel mtrR alleles that were not present in the databases were identified. All new sequences were submitted to the NG-STAR database and assigned the following numbers:
  • allele 525—insertion of nucleotide A at position 38, resulting in a frameshift;
  • allele 526—deletion of two GA nucleotides at position 465 of the gene, resulting in a frameshift;
  • allele 528—T→G substitution at position 531, which does not lead to amino acid substitution;
  • allele 529—T→C substitution at position 132, resulting in the Cys44Arg amino acid substitution;
  • allele 530—T→G substitution at position 531, which does not lead to amino acid substitution, and T→C substitution at position 535, resulting in the Phe178Leu amino acid substitution.

2.3. Diversity of the mtrD Gene Alleles

A total of ten variants of mtrD alleles were identified (Figure 1b); among them, there were no wild type alleles (NEIS1633 “Allele 640”). Some types of alleles differed only in the nucleotide sequence, i.e., they had synonymous substitutions that did not lead to a change in the amino acid sequence.
The most common were alleles 908 (26%), 628 (24%), and 937 (23%), which had 2–15 amino acid substitutions in comparison with the wild type sequence. However, we also found alleles of the mtrD gene that arose as a result of a probable horizontal transfer from N. meningitidis (NEIS1633 Alleles “3353”; “3354”; “3359”). These mosaic allele types had 19–62 amino acid substitutions and were found in 36 isolates, comprising about 20% of the entire sample. Of particular note is that all mosaic types were found to have Ser821Ala and Lys823Glu substitutions which were most often associated with a decrease in susceptibility to macrolides [17].

2.4. Relationship between the mtrR and mtrD Alleles

Our analysis revealed matching between different alleles of the mtrR and mtrD genes. For example, twenty N. gonorrhoeae isolates carrying mtrR allele 122 also carried mtrD allele 937; in twelve isolates, mtrR allele 62 was accompanied by mtrD allele 908 (Figure 2 and Supplementary Table S2). The integration of the results of phylogenetic analysis for the mtrR and mtrD genes with data on the matching of alleles allowed us to identify the matching clades. In total, 25 matched clades with linked mtrR and mtrD alleles were found in the isolate sample; out of them, 7 clades included 5–20 isolates (shown on Figure 2) and 18 clades included 2–4 isolates (not shown on Figure 2 but presented in Supplementary Table S2). Thus, for the first time for the Russian population of N. gonorrhoeae, we demonstrated the co-evolution of the mtrR and mtrD genes.
Of particular note is the clade of isolates with linked mtrR 485 and mtrD 3353 alleles. It includes twenty isolates in which nucleotide sequences from N. meningitidis were found both in the mtrR gene and its promoter, and in the mtrD gene. This cluster was one of the most numerous in terms of the number of isolates. Due to the presence of mosaic structures in the mtrR and mtrD genes, isolates from this cluster proved to be resistant to azithromycin (MICAzm ≈ 2 mg/L).
The second largest clade consisted of fifteen isolates with linked mtrR 219 and mtrD 628 alleles. It is formed exclusively by isolates with NG-MAST sequence types 807 and 228, which are the most common in the Russian population of N. gonorrhoeae [24]. This clade of isolates, although one of the most numerous, has not been associated with azithromycin resistance.

2.5. Effect of Genetic Profile on MICAzm

To systematize the obtained data, 162 studied N. gonorrhoeae isolates were divided into seven groups according to the median MICAzm values and mutations in the mtrR and mtrD alleles, and rrn operon (Table 1):
  • group I: 12 isolates with the single His105Tyr substitution in the mtrR gene;
  • group II: 47 isolates, 98% of which had the Ala86Thr mutation in the mtrR gene;
  • group III: 35 isolates with His105Tyr and Ala39Thr substitutions in the mtrR gene. Several isolates carried the Gly45Ser mutation;
  • group IV: 23 isolates with the His105Tyr substitution (100%) and the -35delA mutation in the promoter region of the mtrR gene (83%). There were also the -10insT, Ala39Thr, and Gly45Asp mutations;
  • group V: 13 isolates with the His105Tyr (100%) and Ala39Thr (54%) substitutions in the mtrR gene and the Ser821Ala, Lys823Glu substitutions in the mtrD gene associated with mosaic structure (92%);
  • group VI: 23 isolates with nucleotide sequences derived from N. meningitides both in the mtrR gene and its promoter (MG-like promoter), and in the mtrD gene (Ser821Ala and Lys823Glu mutations). There was also the -35delA mutation in the promoter region of the mtrR gene and the Ala86Thr mutation in the mtrR gene;
  • group VII: two isolates with the C2611T mutations in all four alleles of the 23S rRNA gene. One isolate contained an MG-like promoter, the -35delA, Ala39Thr, Ala86Thr mutations in the mtrR gene, and the Ser821Ala, Lys823Glu mutations in the mtrD gene;
  • ungrouped (7 isolates).
The median value of MICAzm for isolates from group I was 0.06 mg/L, and the single His105Tyr mutation detected in these isolates, apparently, makes a minimal contribution to the formation of resistance. Isolates from groups II–IV demonstrated only a slight decrease in susceptibility to azithromycin (MICAzm = 0.12 mg/L).
The median value of MICAzm for isolates from group V was 0.25 mg/L. It should be noted, however, that this group includes azithromycin-resistant isolates (five out of thirteen) with MICAzm of 2–8 mg/L. The decreased susceptibility to azithromycin within this group is most likely associated with the Ser821Ala and Lys823Glu mutations in the mtrD gene.
The formation of resistance to azithromycin within group VI with the median value of MICAzm of 2 mg/L is probably associated with the MG-like promoter, the simultaneous presence of -35delA and Ala86Thr mutations in the mtrR gene, and the Ser821Ala and Lys823Glu mutations in the mtrD gene associated with its mosaic structure.
The highest median value of MICAzm = 4 mg/L was observed in group VII and was associated with the C2611T mutations in all four alleles of the 23S rRNA gene. The box plot diagram of changes in MICAzm depending on the genetic profile in groups is shown in Figure 3.

2.6. Phylogenetic Tree for mtrR and mtrD Genes

For alleles of the mtrR and mtrD genes found in the Russian N. gonorrhoeae population, a maximum likelihood phylogenetic tree (phylogram) was constructed (Figure 4). The group colors correspond to those in the box-plot diagram (Figure 3). Isolates on the phylogram form clades in accordance with the division into groups I-VII described above. This tree demonstrates the relative phylogenetic proximity of isolates from groups I-IV, in which no azithromycin-resistant isolates were found. Isolates from groups V and VI are located phylogenetically far away from groups I-IV and form a separate cluster in the tree. This remoteness is due to a large number of substitutions in the mtrR and mtrD genes associated with their mosaic structure, which also affects the resistance to azithromycin within these groups (MICAzm ≈ 2 mg/L) (Table 1, Figure 4).

2.7. MtrR and mtrD Alleles in Isolates Belonging to NG-MAST 228 and 807

The NG-MAST types were determined for all the studied isolates (Table S1). In general, we did not see any relationship between molecular types and mtrR and mtrD gene types for our sample of isolates. The exceptions were isolates belonging to MG-MAST 228 and 807, i.e., those types of sequences that, as has been previously established [24], were most common in the Russian Federation. These sequence types belong to the G807 genogroup, and the combined porB and tbpB sequences of NG-MAST 228 and 807 differ by one nucleotide. The isolates were characterized by susceptibility to azithromycin and belonged to group II (Table 1). Almost all the isolates of these types had the same allele of the mtrD gene, allele 628, and similar alleles of the mtrR gene. To test the assumption about the relationship between the types of mtrR and mtrD alleles and NG-MAST types, we expanded the sample of isolates by adding the isolates of NG-MAST 228 and 807 collected in Russia in 2016–2019, for which the whole genome sequencing was performed by us previously (GenBank, Bioproject PRJNA768989). In total, the sample included 41 Russian isolates of N. gonorrhoeae collected in 2016–2021 (Table 2).
Almost all Russian isolates of NG-MAST 228 and 807 (40 out of 41 isolates) had the mtrD allele 628, and 39 out of 41 isolates carried mtrR alleles 5, 445, 526, and 537, differing from allele 219 by one nucleotide (Table 2). Thus, indeed, a link was revealed between the NG-MAST 228 and 807, typical of the Russian N. gonorrhoeae population, and the types of mtrR and mtrD alleles.

3. Discussion

Despite the fact that azithromycin is not used for the treatment of gonococcal infection in Russia, the resistance of N. gonorrhoeae to this antibiotic is at a high level. According to the EUCAST criteria, only 86.4% of the clinical isolates collected in the Russian Federation in 2020–2021 were phenotypically susceptible to this antimicrobial, which is below the WHO recommended threshold of 95% for its use. This situation is especially paradoxical since the proportion of resistant isolates in neighboring countries remains negligible. In Russia’s neighbor Belarus, where azithromycin is also not used to treat gonorrhea, the proportion of azithromycin-resistant isolates in 2018–2019 was only 2.6% [25]. At the same time, in Kyrgyzstan, for which the similarity of N. gonorrhoeae populations with those in Russia was noted (the presence of a large number of isolates of NG-MAST 1751 and 807), no isolates resistant to azithromycin were found in 2017, although azithromycin, along with ceftriaxone, is recommended for the treatment of gonococcal infections [26].
To clarify the nature of N. gonorrhoeae resistance to azithromycin in Russia, we analyzed the genetic determinants of resistance. We identified polymorphisms in the 23S rRNA, mtrR, and mtrD genes, including their mosaic variants, while the erm and mefA genes were found to be missing in the genomes of the isolates. Antibiotic-specific mutations in all four copies of the 23 rRNA genes were found only in two isolates (1.2% of the analyzed sample). Thus, resistance to azithromycin in the studied sample of isolates was mainly due to nucleotide substitutions in the genes encoding the MtrCDE efflux pump (non-specific mechanism of resistance), namely, mutations in the mtrR and mtrD genes, and their mosaic alleles, which were formed in the course of a probable horizontal gene transfer from N. meningitidis. The obtained results allowed us to conclude that the mechanism of azithromycin resistance associated with the MtrCDE efflux pump arose in the Russian population of N. gonorrhoeae as a means of counteracting other antimicrobial drugs (e.g., penicillin, tetracycline). Thus, the increased level in resistance to azithromycin in the Russian population of N. gonorrhoeae may be associated with the wide spreading of non-specific mechanisms of resistance to antimicrobial drugs that have emerged in the population regardless of the use/non-use of this antibiotic.
An important observation is the matching between the mtrR and mtrD alleles, as a result of which, 25 matching clades with linked alleles were found in the Russian population of N. gonorrhoeae. Of particular note are matching clades which simultaneously contain mosaic variants of both mtrR and mtrD genes. Thus, we revealed the cooperative molecular evolution of the mtrR and mtrD genes in the Russian population of N. gonorrhoeae. The demonstrated linkage shows that recombination events between closely related Neisseria spp. can, apparently, affect both individual genes and entire operons.
Previously, we have shown a functional dependence between the penA allele type and the NG-MAST, i.e., isolates of the same sequence type carried the same penA alleles, although the opposite was not true [27]. In this study, a link between the mtrR and mtrD alleles of the efflux pump and NG-MAST was found only in isolates of the sequence types 228 and 807 that are typical of the Russian population of N. gonorrhoeae. It can be noted that these isolates did not have the determinants of azithromycin resistance and they mainly carried the mtrR alleles 445/219 and the mtrD allele 628. In this work, we did not perform detailed phylogenetic study of N. gonorrhoeae isolates and comparative population analysis with European isolates in order to identify similar clades of azithromycin-resistant isolates, and this will be the subject of a separate study.
The reason for the emerging azithromycin resistance in the Russian gonococcal population and the formation of a separate cluster of resistant isolates can most likely be due to several reasons: (1) the cross-border transfer of a resistant strain with specific resistance determinants from abroad, where such strains are already found everywhere [16] and resistance in such strains has been developed during the treatment of gonococcal infection with azithromycin; (2) a recombination event between N. gonorrhoeae and closely related N. meningitidis followed by its fixation in the population [28,29] as a mechanism of resistance non-specific to azithromycin.

4. Materials and Methods

4.1. Collection and Characterization of N. gonorrhoeae Isolates

For the characterization of the modern Russian population of N. gonorrhoeae, clinical isolates of N. gonorrhoeae (162 samples) were collected in 2020–2021 from 7 subjects of the Russian Federation (Arkhangelsk, Astrakhan, Kaluga, Omsk Oblasts, City of Moscow, Stavropol Krai, Chuvash Republic) by the State Research Center for Dermatovenerology and Cosmetology of the Ministry of Health of Russia. The samples were provided by specialized medical organizations of the dermatovenerological profile, each sample from an individual patient. Sample collection, transportation, culturing, phenotyping, and storage were performed according to the protocols described previously [23,30].

4.2. Testing of N. gonorrhoeae Susceptibility for Azithromycin

Measurement of the minimum inhibitory concentration for azithromycin (MICAzm) was performed by the serial dilutions in chocolate agar. Each strain was characterized in accordance with the EUCAST (ECOFF) criteria for azithromycin: S—susceptible (MICAzm) < 1.0 mg/L), R—resistant (MICAzm) > 1.0 mg/L).

4.3. Study of Genetic Determinants of N. gonorrhoeae Resistance to Azithromycin

The sequences of all four copies of the 23S rRNA gene in N. gonorrhoeae samples were determined by Sanger sequencing as described previously [11]. The mtrR gene sequences were determined by Sanger sequencing using the NG-STAR protocol (https://ngstar.canada.ca/, accessed on 26 September 2022) with the previously described primers [31].
The mtrD gene was amplified using PCR with For1-5′-ACGGCATCTGAAGCCAAA-3′ and Rev1-5′-AAAGTCCTGATGCCGTCTG-3′ primers. The resulting 3304 bp amplicons were Sanger sequenced using additional primers For2-5′-CGGCAACGTCATCCTCCG-3′ and Rev2-5′-TGTAACCGCCGCCCAATT-3′. The obtained nucleotide sequences, two from the forward primer and two from the reverse primer, were concatenated using the Unipro UGENE v.44 program [32].
The study of N. gonorrhoeae isolates for the presence of the ermA/B/C/F methyltransferase genes and the mefA gene, was performed as described previously [20].

4.4. Determination of mtrR and mtrD Allele Types

The resulting nucleotide sequences for each gene were combined into a multiple alignment file in the Bio-Edit program (Ibis Biosciences, Carlsbad, CA, USA) and checked for mutations, stop codons, or frameshift. The wild type sequence PuBMLST “Allele 640” was used as a reference genome for mtrD allele alignment. As noted in [33], there is no consensus about which amino acid, Tyr or His, is present in codon 105 of the wild type mtrR sequence. In this study, we took the sequence NG-STAR “Allele 353”, GenBank accession number KT954121.1, as a wild type for mtrR, i.e., this sequence containing His at codon 105.
For each isolate in the sample (162 isolates), the mtrR and mtrD allele numbers were determined. For the mtrR alleles, the NG-STAR database (https://ngstar.canada.ca, accessed on 26 September 2022) was used; for mtrD alleles, the BIGSdb database (https://pubmlst.org, accessed on 10 October 2022) was used according to the “NEIS1633(mtrD)” typing scheme. If there was not a complete match between the analyzed sequence and the gene sequence in the database, for example, the sequence under study contained a mutation, a stop codon, or a frameshift, we submitted that sequence into the database as a new one.

4.5. NG-MAST Typing

Molecular typing of N. gonorrhoeae isolates was performed using the standard NG-MAST protocol [34]. The variable internal regions of the porB and tbpB genes were PCR-amplified and the resulting products were purified and sequenced using a 3730xl Genetic Analyzer (Applied Biosystems, Waltham, MA, USA). Both the leading and reverse strands were assessed and the sequencing data were processed using the 3730/3730xl Data Collection Software version 3.0. Allele numbers for the porB and tbpB sequences and sequence types were assigned according to the NG MAST v2.0 database (https://pubmlst.org, accessed on 20 October 2022).

4.6. Construction of Phylogenetic Trees

Phylogenetic trees were constructed using the RaxML v.8.2.4 software (https://usegalaxy.eu/, accessed on 17 October 2022) with 1000 rapid bootstrap inferences. Cladograms for the mtrR and mtrD genes were constructed independently from each other according to the sequences of alleles corresponding to each isolate. To construct a phylogram, we used concatenated mtr and mtrD gene sequences obtained using the Bio-Edit alignment tool (Ibis biosciences, Carlsbad, CA, USA).

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/antibiotics12010170/s1, Table S1: Characteristics of N. gonorrhoeae isolates collected in the Russian Federation in 2020–2021; Table S2: Link between mtrR and mtrD allele types of N. gonorrhoeae isolates collected in the Russian Federation in 2020–2021.

Author Contributions

Conceptualization, I.K. and B.S.; investigation, I.K., A.V., S.G. and J.S.; sequencing, D.K. and I.K.; clinical samples collection, V.S.; writing—original draft preparation, I.K. and E.D.; writing—review and editing, E.D., D.D. and D.G.; supervision, funding acquisition, project administration, A.K. and D.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Ministry of Science and Higher Education of the Russian Federation to the EIMB Center for Precision Genome Editing and Genetic Technologies for Biomedicine, agreement number 075-15-2019-1660 (molecular investigation of N. gonorrhoeae isolates), and by the Ministry of Health of the Russian Federation, government assignment No. 056-03-2021-124 (azithromycin susceptibility testing).

Institutional Review Board Statement

Ethical approval/written informed consent was not required for the study of animals/human participants in accordance with the local legislation and institutional requirements.

Data Availability Statement

sequences of the new mtrR alleles were added to the NG-STAR database under accession numbers: allele 525, 526, 528, 529, 530.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Global progress report on HIV, viral hepatitis and sexually transmitted infections, 2021. In The Global Health Sector Strategies 2016–2021: Actions for Impact; World Health Organization: Geneva, Switzerland, 2021.
  2. Golparian, D.; Unemo, M. Antimicrobial resistance prediction in Neisseria gonorrhoeae: Current status and future prospects. Expert Rev. Mol. Diagn. 2022, 22, 29–48. [Google Scholar] [CrossRef]
  3. Shaskolskiy, B.; Kandinov, I.; Dementieva, E.; Gryadunov, D. Antibiotic resistance in Neisseria gonorrhoeae: Challenges in research and treatment. Microorganisms 2022, 10, 1699. [Google Scholar] [CrossRef]
  4. Barbee, L.A.; St Cyr, S.B. Management of Neisseria gonorrhoeae in the United States: Summary of evidence from the development of the 2020 gonorrhea treatment recommendations and the 2021 Centers for Disease Control and Prevention sexually transmitted infection treatment guidelines. Clin. Infect. Dis. 2022, 74 (Suppl. S2), S95–S111. [Google Scholar] [CrossRef]
  5. WHO Guidelines for the Treatment of Neisseria gonorrhoeae; World Health Organization: Geneva, Switzerland, 2016.
  6. Unemo, M.; Ross, J.; Serwin, A.B.; Gomberg, M.; Cusini, M.; Jensen, J.S. 2020 European guideline for the diagnosis and treatment of gonorrhoea in adults. Int. J. STD AIDS 2020, 956462420949126. [Google Scholar] [CrossRef] [PubMed]
  7. Derbie, A.; Mekonnen, D.; Woldeamanuel, Y.; Abebe, T. Azithromycin resistant gonococci: A literature review. Antimicrob. Resist. Infect. Control 2020, 9, 138. [Google Scholar] [CrossRef] [PubMed]
  8. George, C.R.R.; Enriquez, R.P.; Gatus, B.J.; Whiley, D.M.; Lo, Y.R.; Ishikawa, N.; Wi, T.; Lahra, M.M. Systematic review and survey of Neisseria gonorrhoeae ceftriaxone and azithromycin susceptibility data in the Asia Pacific, 2011 to 2016. PLoS ONE 2019, 14, e0213312. [Google Scholar] [CrossRef] [Green Version]
  9. Shimuta, K.; Lee, K.; Yasuda, M.; Furubayashi, K.; Uchida, C.; Nakayama, S.I.; Takahashi, H.; Ohnishi, M. Characterization of 2 Neisseria gonorrhoeae strains with high-level azithromycin resistance isolated in 2015 and 2018 in Japan. Sex Transm. Dis. 2021, 48, e85–e87. [Google Scholar] [CrossRef] [PubMed]
  10. Mlynarczyk-Bonikowska, B.; Kowalewski, C.; Krolak-Ulinska, A.; Marusza, W. Molecular mechanisms of drug resistance and epidemiology of multidrug-resistant variants of Neisseria gonorrhoeae. Int. J. Mol. Sci. 2022, 23, 10499. [Google Scholar] [CrossRef] [PubMed]
  11. Chisholm, S.A.; Dave, J.; Ison, C.A. High-level azithromycin resistance occurs in Neisseria gonorrhoeae as a result of a single point mutation in the 23S rRNA genes. Antimicrob. Agents Chemother. 2010, 54, 3812–3816. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Whiley, D.M.; Kundu, R.L.; Jennison, A.V.; Buckley, C.; Limnios, A.; Hogan, T.; Enriquez, R.; El Nasser, J.; George, C.R.; Lahra, M.M. Azithromycin-resistant Neisseria gonorrhoeae spreading amongst men who have sex with men (MSM) and heterosexuals in New South Wales, Australia, 2017. J. Antimicrob. Chemother. 2018, 73, 1242–1246. [Google Scholar] [CrossRef]
  13. Zhou, Q.; Liu, J.; Chen, S.; Xu, W.; Han, Y.; Yin, Y. The accuracy of molecular detection targeting the mutation C2611T for detecting moderate-level azithromycin resistance in Neisseria gonorrhoeae: A systematic review and meta-analysis. Antibiotics 2021, 10, 1027. [Google Scholar] [CrossRef] [PubMed]
  14. The European Committee on Antimicrobial Susceptibility Testing. Breakpoint Tables for Interpretation of MICs and Zone Diameters. Version 12.0. 2022. Available online: https://www.eucast.org/clinical_breakpoints (accessed on 1 September 2022).
  15. Beggs, G.A.; Ayala, J.C.; Kavanaugh, L.G.; Read, T.D.; Hooks, G.M.; Schumacher, M.A.; Shaferm, W.M.; Brennan, R.G. Structures of Neisseria gonorrhoeae MtrR-operator complexes reveal molecular mechanisms of DNA recognition and antibiotic resistance-conferring clinical mutations. Nucl. Acids Res. 2021, 49, 4155–4170. [Google Scholar] [CrossRef] [PubMed]
  16. Joseph, S.J.; Thomas, J.C.; Schmerer, M.W.; Cartee, J.C.; St Cyr, S.; Schlanger, K.; Kersh, E.N.; Raphael, B.H.; Gernert, K.M. Global emergence and dissemination of Neisseria gonorrhoeae ST-9363 isolates with reduced susceptibility to azithromycin. Genome Biol. Evol. 2022, 14, evab287. [Google Scholar] [CrossRef] [PubMed]
  17. Rouquette-Loughlin, C.E.; Reimche, J.L.; Balthazar, J.T.; Dhulipala, V.; Gernert, K.M.; Kersh, E.N.; Pham, C.D.; Pettus, K.; Abrams, A.J.; Trees, D.L.; et al. Mechanistic basis for decreased antimicrobial susceptibility in a clinical isolate of Neisseria gonorrhoeae possessing a mosaic-like mtr efflux pump locus. mBio 2018, 9, e02281-18. [Google Scholar] [CrossRef] [Green Version]
  18. Demczuk, W.; Martin, I.; Peterson, S.; Bharat, A.; Van Domselaar, G.; Graham, M.; Lefebvre, B.; Allen, V.; Hoang, L.; Tyrrell, G.; et al. Genomic epidemiology and molecular resistance mechanisms of azithromycin-resistant Neisseria gonorrhoeae in Canada from 1997 to 2014. J. Clin. Microbiol. 2016, 54, 1304–1313. [Google Scholar] [CrossRef] [Green Version]
  19. Trembizki, E.; Doyle, C.; Jennison, A.; Smith, H.; Bates, J.; Lahra, M.; Whiley, D. A Neisseria gonorrhoeae strain with a meningococcal mtrR sequence. J. Med. Microbiol. 2014, 63, 1113–1115. [Google Scholar] [CrossRef] [Green Version]
  20. Cousin, S., Jr.; Whittington, W.L.; Roberts, M.C. Acquired macrolide resistance genes in pathogenic Neisseria spp. isolated between 1940 and 1987. Antimicrob. Agents Chemother. 2003, 47, 3877–3880. [Google Scholar] [CrossRef] [Green Version]
  21. Luna, V.A.; Cousin, S., Jr.; Whittington, W.L.; Roberts, M.C. Identification of the conjugative mef gene in clinical Acinetobacter junii and Neisseria gonorrhoeae isolates. Antimicrob. Agents Chemother. 2000, 44, 2503–2506. [Google Scholar] [CrossRef] [Green Version]
  22. Belkacem, A.; Jacquier, H.; Goubard, A.; Mougari, F.; La Ruche, G.; Patey, O.; Micaelo, M.; Semaille, C.; Cambau, E.; Bercot, B. Molecular epidemiology and mechanisms of resistance of azithromycin-resistant Neisseria gonorrhoeae isolated in France during 2013-14. J. Antimicrob. Chemother. 2016, 71, 2471–2478. [Google Scholar] [CrossRef] [Green Version]
  23. Kubanov, A.; Solomka, V.; Plakhova, X.; Chestkov, A.; Petrova, N.; Shaskolskiy, B.; Dementieva, E.; Leinsoo, A.; Gryadunov, D.; Deryabin, D. Summary and trends of the Russian Gonococcal Antimicrobial Surveillance Programme, 2005 to 2016. J. Clin. Microbiol. 2019, 57, e02024-18. [Google Scholar] [CrossRef]
  24. Shaskolskiy, B.; Dementieva, E.; Kandinov, I.; Chestkov, A.; Kubanov, A.; Deryabin, D.; Gryadunov, D. Genetic diversity of Neisseria gonorrhoeae multi-antigen sequence types in Russia and Europe. Int. J. Infect. Dis. 2020, 93, 1–8. [Google Scholar] [CrossRef] [PubMed]
  25. Aniskevich, A.; Shimanskaya, I.; Boiko, I.; Golubovskaya, T.; Golparian, D.; Stanislavova, I.; Jacobsson, S.; Adaskevich, A.; Unemo, M. Antimicrobial resistance in Neisseria gonorrhoeae isolates and gonorrhoea treatment in the Republic of Belarus, Eastern Europe, 2009-2019. BMC Infect. Dis. 2021, 21, 520. [Google Scholar] [CrossRef] [PubMed]
  26. Karymbaeva, S.; Boiko, I.; Jacobsson, S.; Mamaeva, G.; Ibraeva, A.; Usupova, D.; Golparian, D.; Unemo, M. Antimicrobial resistance and molecular epidemiological typing of Neisseria gonorrhoeae isolates from Kyrgyzstan in Central Asia, 2012 and 2017. BMC Infect. Dis. 2021, 21, 559. [Google Scholar] [CrossRef] [PubMed]
  27. Kandinov, I.; Dementieva, E.; Kravtsov, D.; Chestkov, A.; Kubanov, A.; Solomka, V.; Deryabin, D.; Gryadunov, D.; Shaskolskiy, B. Molecular typing of Neisseria gonorrhoeae clinical isolates in Russia, 2018-2019: A link between penA alleles and NG-MAST types. Pathogens 2020, 9, 941. [Google Scholar] [CrossRef]
  28. Sanchez-Buso, L.; Cole, M.J.; Spiteri, G.; Day, M.; Jacobsson, S.; Golparian, D.; Sajedi, N.; Yeats, C.A.; Abudahab, K.; Underwood, A.; et al. Europe-wide expansion and eradication of multidrug-resistant Neisseria gonorrhoeae lineages: A genomic surveillance study. Lancet Microbe 2022, 3, e452–e463. [Google Scholar] [CrossRef]
  29. Tanaka, M.; Furuya, R.; Kobayashi, I.; Kanesaka, I.; Ohno, A.; Katsuse, A.K. Antimicrobial resistance and molecular characterisation of Neisseria gonorrhoeae isolates in Fukuoka, Japan, 1996-2016. J. Glob. Antimicrob. Resist. 2019, 17, 3–7. [Google Scholar] [CrossRef]
  30. Shaskolskiy, B.; Dementieva, E.; Kandinov, I.; Filippova, M.; Petrova, N.; Plakhova, X.; Chestkov, A.; Kubanov, A.; Deryabin, D.; Gryadunov, D. Resistance of Neisseria gonorrhoeae isolates to beta-lactams (benzylpenicillin and ceftriaxone) in Russia, 2015-2017. PLoS ONE 2019, 14, e0220339. [Google Scholar] [CrossRef] [Green Version]
  31. Mavroidi, A.; Tzouvelekis, L.S.; Kyriakis, K.P.; Avgerinou, H.; Daniilidou, M.; Tzelepi, E. Multidrug-resistant strains of Neisseria gonorrhoeae in Greece. Antimicrob. Agents Chemother. 2001, 45, 2651–2654. [Google Scholar] [CrossRef] [Green Version]
  32. Okonechnikov, K.; Golosova, O.; Fursov, M.; UGENE Team. Unipro UGENE: A unified bioinformatics toolkit. Bioinformatics 2012, 28, 1166–1167. [Google Scholar] [CrossRef] [Green Version]
  33. Hall, C.L.; Harrison, M.A.; Pond, M.J.; Chow, C.; Harding-Esch, E.M.; Sadiq, S.T. Genotypic determinants of fluoroquinolone and macrolide resistance in Neisseria gonorrhoeae. Sex Health 2019, 16, 479–487. [Google Scholar] [CrossRef]
  34. Martin, I.M.; Ison, C.A.; Aanensen, D.M.; Fenton, K.A.; Spratt, B.G. Rapid sequence-based identification of gonococcal transmission clusters in a large metropolitan area. J. Infect. Dis. 2004, 189, 1497–1505. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Amino acid sequences of mtrR (a) and mtrD (b) alleles in N. gonorrhoeae isolates collected in the Russian Federation in 2020–2021 in comparison with the wild type (WT) mtrR (NG-STAR “Allele 353”) and mtrD (PubMLST “Allele 640”) sequences. Designations: ‘+’ – presence of deletion/ MG-like promoter; ‘–‘ – absence of deletion/ MG-like promoter.
Figure 1. Amino acid sequences of mtrR (a) and mtrD (b) alleles in N. gonorrhoeae isolates collected in the Russian Federation in 2020–2021 in comparison with the wild type (WT) mtrR (NG-STAR “Allele 353”) and mtrD (PubMLST “Allele 640”) sequences. Designations: ‘+’ – presence of deletion/ MG-like promoter; ‘–‘ – absence of deletion/ MG-like promoter.
Antibiotics 12 00170 g001
Figure 2. Correspondence between mtrR and mtrD cladograms constructed independently for the Russian N. gonorrhoeae isolates collected in 2020–2021. Sample numbers and the mtrR or mtrD allele numbers are indicated on the branches. Matched clades are formed by linked leaves of the phylogenetic tree. Matched clades consisting of five or more isolates are painted in different colors.
Figure 2. Correspondence between mtrR and mtrD cladograms constructed independently for the Russian N. gonorrhoeae isolates collected in 2020–2021. Sample numbers and the mtrR or mtrD allele numbers are indicated on the branches. Matched clades are formed by linked leaves of the phylogenetic tree. Matched clades consisting of five or more isolates are painted in different colors.
Antibiotics 12 00170 g002
Figure 3. Box plot: MICAzm in groups of N. gonorrhoeae isolates with different genetic profiles. A description of the profiles by group is given in Table 1. Thick solid lines in each box indicate the median MICAzm value for each group. The whiskers extend to the boundaries of the interquartile range. The dotted line indicates the EUCAST breakpoint for azithromycin resistance.
Figure 3. Box plot: MICAzm in groups of N. gonorrhoeae isolates with different genetic profiles. A description of the profiles by group is given in Table 1. Thick solid lines in each box indicate the median MICAzm value for each group. The whiskers extend to the boundaries of the interquartile range. The dotted line indicates the EUCAST breakpoint for azithromycin resistance.
Antibiotics 12 00170 g003
Figure 4. Maximum likelihood phylogenetic distance tree for concatenated mtrR and mtrD gene sequences of the N. gonorrhoeae isolates collected in Russia in 2020–2021. Isolate numbers are colored according to groups I-VII (Table 1, Figure 3). Bootstrap values of 70 or more are colored in red.
Figure 4. Maximum likelihood phylogenetic distance tree for concatenated mtrR and mtrD gene sequences of the N. gonorrhoeae isolates collected in Russia in 2020–2021. Isolate numbers are colored according to groups I-VII (Table 1, Figure 3). Bootstrap values of 70 or more are colored in red.
Antibiotics 12 00170 g004
Table 1. Genetic profiles of N. gonorrhoeae isolates.
Table 1. Genetic profiles of N. gonorrhoeae isolates.
Group
(Number of Isolates)
Median MICAzm (mg/L)Genetic Profile mtrRGenetic Profile mtrD23S rRNA
mtr
Allele
MG-Like Promoter-35delA-10insTAla39
Thr
Gly45
Asp
Gly45
Ser
Ala86
Thr
His105
Tyr
mtrD
Allele
Ser821
Ala
Lys823
Glu
C2611T
(4 Alleles)
I (12)0.06299 (42%)
142 (33%)
270 (25%)
100% *908 (75%)
937 (25%)
II (47)0.12219 (33%)
445 (30%)
374 (9%)
274 (6%)
129 (4%)
142 (4%)
526 (4%)
5 (2%)
62 (2%)
455 (2%)
500 (2%)
525 (2%)
11%2%98%628 (74%)
908 (11%)
937 (11%)
670 (4%)
III (35)0.12122 (57%)
62 (31%)
129 (3%)
69 (3%)
432 (3%)
216 (3%)
94%31%100%6%937 (60%)
908 (34%)
628 (6%)
IV (23)0.12288 (36%)
19 (22%)
25 (13%)
528 (9%)
113 (4%)
142 (4%)
232 (4%)
256 (4%)
455 (4%)
83%4%4%39%100%52%908 (61%)
937 (26%)
3888 (9%)
670 (4%)
V (13)0.25432 (45%)
445 (23%)
25 (8%)
122 (8%)
526 (8%)
529 (8%)
8%54%8%100%3353 (54%)
3359 (30%)
937 (8%)
3354(8%)
92%92%
VI (23)2.0485 (83%)
530 (17%)
100%100%100%3353 (100%)100%100%
VII (2)4.0485 (50%)
122 (50%)
50%50%50%100%3353 (50%)
937 (50%)
50%50%100%
Ungrouped (7) 1129 (30%)
142 (14%)
299 (14%)
219 (14%)
189 (14%)
520 (14%)
43%43%908 (33%)
628 (33%)
994 (17%)
3381 (17%)
* Cells, where the percentage of occurrence of a particular mutation in the group exceeded 50%, are highlighted in bold.
Table 2. Characteristics of N. gonorrhoeae isolates belonging to the sequence types most common in the Russian Federation.
Table 2. Characteristics of N. gonorrhoeae isolates belonging to the sequence types most common in the Russian Federation.
NG-MASTNumber of IsolatesmtrR AllelemtrD AlleleComment
2284219628Alleles mtrR 5, 445, 526, 537 differ from allele mtrR 219 by one nucleotide
4445628
25628
3537628
162908
80714219628
3445628
3526628
65628
1500628
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kandinov, I.; Shaskolskiy, B.; Kravtsov, D.; Vinokurova, A.; Gorshkova, S.; Kubanov, A.; Solomka, V.; Shagabieva, J.; Deryabin, D.; Dementieva, E.; et al. Azithromycin Susceptibility Testing and Molecular Investigation of Neisseria gonorrhoeae Isolates Collected in Russia, 2020–2021. Antibiotics 2023, 12, 170. https://doi.org/10.3390/antibiotics12010170

AMA Style

Kandinov I, Shaskolskiy B, Kravtsov D, Vinokurova A, Gorshkova S, Kubanov A, Solomka V, Shagabieva J, Deryabin D, Dementieva E, et al. Azithromycin Susceptibility Testing and Molecular Investigation of Neisseria gonorrhoeae Isolates Collected in Russia, 2020–2021. Antibiotics. 2023; 12(1):170. https://doi.org/10.3390/antibiotics12010170

Chicago/Turabian Style

Kandinov, Ilya, Boris Shaskolskiy, Dmitry Kravtsov, Alexandra Vinokurova, Sofya Gorshkova, Alexey Kubanov, Victoria Solomka, Julia Shagabieva, Dmitry Deryabin, Ekaterina Dementieva, and et al. 2023. "Azithromycin Susceptibility Testing and Molecular Investigation of Neisseria gonorrhoeae Isolates Collected in Russia, 2020–2021" Antibiotics 12, no. 1: 170. https://doi.org/10.3390/antibiotics12010170

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop