Skip to main content
Antimicrobial Agents and Chemotherapy logoLink to Antimicrobial Agents and Chemotherapy
. 1999 Dec;43(12):2823–2830. doi: 10.1128/aac.43.12.2823

Nomenclature for Macrolide and Macrolide-Lincosamide-Streptogramin B Resistance Determinants

Marilyn C Roberts 1,*, Joyce Sutcliffe 2, Patrice Courvalin 3, Lars Bogo Jensen 4, Julian Rood 5, Helena Seppala 6
PMCID: PMC89572  PMID: 10582867

Macrolides are composed of 14 (erythromycin and clarithromycin)-, 15 (azithromycin)-, or 16 (josamycin, spiramycin, and tylosin)-membered lactones to which are attached amino and/or neutral sugars via glycosidic bonds. Erythromycin was introduced in 1952 as the first macrolide antibiotic. Unfortunately, within a year, erythromycin-resistant (Emr) staphylococci from the United States, Europe, and Japan were described (101). Erythromycin is produced by Saccharopolyspora erythraea, while the newer macrolides are semisynthetic molecules with substitutions on the lactone. The newer derivatives, such as clarithromycin and azithromycin, have improved intracellular and tissue penetration, are more stable, are better absorbed, have a lower incidence of gastrointestinal side effects, and are less likely to interact with other drugs. They are useable against a wider range of infectious bacteria, such as Legionella, Chlamydia, Haemophilus, and some Mycobacterium species (not M. tuberculosis), and their pharmacokinetics provide for less frequent dosing than erythromycin (21, 47, 96, 97). As a result, the usage of the newer macrolides has increased dramatically over the last few years, which has led to increased exposure of bacterial populations to macrolides (101103, 107).

Macrolides inhibit protein synthesis by stimulating dissociation of the peptidyl-tRNA molecule from the ribosomes during elongation (101, 103). This results in chain termination and a reversible stoppage of protein synthesis. The first mechanism of macrolide resistance described was due to posttranscriptional modification of the 23S rRNA by the adenine-N6 methyltransferase (101103). These enzymes add one or two methyl groups to a single adenine (A2058 in Escherichia coli) in the 23S rRNA moiety. Over the last 30 years, a number of adenine-N6-methyltransferases from different species, genera, and isolates have been described. In general, genes encoding these methylases have been designated erm (erythromycin ribosome methylation), although there are exceptions, especially in the antibiotic-producing organisms (see Tables 1 and 3) (103). As the number of erm genes described has grown, the nomenclature for these genes has varied and has been inconsistent (Table 1). In some cases, unrelated genes have been given the same letter designation, while in other cases, highly related genes (>90% identity) have been given different names.

TABLE 1.

rRNA methylase genes involved in MLSB resistance

Class Protein Gene name Gene(s) included % Homology
Plasmid transposonc GenBank no. or referencee
DNA Amino acid
A Erm(A) erm(A) erm(A) 83 81 Tn554 X03216
erm(TR) AF002716
B Erm(B) erm(B) erm(AM) 98–100 98–100 Tn1545 X52632
erm(AM) pAMβ-1 Y00116
erm(AM) pAM77 K00551
erm(B) Tn917 M36722
erm(B) pAD2 M11180
erm(B) pIP501 X72021
erm(AMR) U86375
erm(BC) pIP1527 M19270
erm(P), erm(BP)a pIP402 U18931
erm(IP) pIP501 U00453
erm(Z), erm(BZ1), erm(BZ2)a Tn5398 AF109075
erm pLEM3 U48430, X82819
erm(2) pBT233, pMD101 X64695, X66468
C Erm(C) erm(C) erm(C) 99–100 98–100 pE194 J01755, V01278
erm(C) pT48 M19652
erm(C) pE5 M17990
erm(C) pJR5 L04687
erm(C) pA22 X54338
erm(C) pSES6 X82668
erm(C) pSES5 Y09001
erm(C) pSES4a Y09002
erm(C) pSES21 Y09003
erm(C) J3356::pOX7 U36911
erm(IM) pIM13 M13761
erm(M) pNE131 M12730
erm(M) pPV141 U82607
erm(M) pPV142 AF019140
Dd Erm(D) erm(D) erm(D) 97–99  97–99  pBD90 M29832
erm(J) pBA423 L08389
erm(K) M77505
E Erm(E) erm(E) erm(E) 99 96 pUC31, pIJ43 X51891
erm(E2) M11200
F Erm(F) erm(F) erm(F) 98–100 97–100 pBF4 M14730
erm(F) Tn4351 M17124
erm(FS) pBI106, Tn4551 M17808
erm(FU) Chromosomal M62487
G Erm(G) erm(G) erm(G) 99 99 pBD370 M15332
erm(G) Tn7853 L42817
Hb Erm(H) erm(H) car(B) pOJ159 P13079, M16503
Ib Erm(I) erm(I) mdm(A) 34
Nb Erm(N) erm(N) tlr(D) X97721
Ob Erm(O) erm(O) lrm 84 84 pLST391 M74717
srm(A) AJ223970
Q Erm(Q) erm(Q) erm(Q) Chromosomal L22689
R Erm(R) erm(R) erm(R) M11276
Sb Erm(S) erm(S) erm(SF) 100 100 pET23 M19269
tlr(A) P45439
Tb Erm(T) erm(T) erm(GT) pGT633 M64090
Ub Erm(U) erm(U) lmr(B) pPZ303 X62867
Vb Erm(V) erm(V) erm(SV) U59450
Wb Erm(W) erm(W) myr(B) M77505, P43433
Xb Erm(X) erm(X) erm(CD), erm(A)b 99–100 99–100 pNG2 M36726, X51472
erm(CX) Tn5432 U21300
Yb Erm(Y) erm(Y) erm(GM) pMS97 57
Zb Erm(2) erm(2) srm(D) 69
Unclassified clr pJI702 18
a

When two or three gene names are listed under “Genes included,” it means that the same gene was designated by two or three different names in the literature; for erm(B), erm(C), and erm(M), multiple related genes have been sequenced, and a selection of these are listed, although they are not complete lists, since some of them have not been completely sequenced. 

b

Newly designated classes, proteins, and genes given a single-letter designation rather than the two-letter or different three-letter designations currently used in the literature. 

c

The list of plasmids includes both recombinant and natural plasmids. 

d

Class D includes genes labeled in the literature as erm(D), erm(J), and erm(K); because they share between 97 and 99% DNA and amino acid homology, they have been grouped under class D and erm(D). 

e

Information was provided by GenBank and references 1012, 18, 25, 29, 31, 32, 34, 35, 37, 38, 42, 46, 5557, 6062, 65, 69, 71, 77, 88, 89, 91, 95, and 108110

TABLE 3.

Location of antibiotic resistance genesa

Gene  Genus or genera
Methylases
erm(A) Actinobacillus, Staphylococcus, Streptococcus
erm(B) Actinobacillus, Clostridium, Escherichia Enterococcus, Klebsiella, Neisseria, Pediococcus, Staphylococcus, Streptococcus, Wolinella
erm(C) Actinobacillus, Bacillus, Eubacterium, Lactobacillus, Neisseria, Staphylococcus, Streptococcus, Wolinella
erm(D) Bacillus
erm(E) Streptomyces
erm(F) Actinobacillus, Actinomyces, Bacteroides, Clostridium, Eubacterium, Fusobacterium, Gardnerella, Haemolphilus, Neisseria, Porphyromonas, Prevotella, Peptosteptococcus, Selenomonas, Streptococcus, Treponema, Veillonella, Wolinella
erm(G) Bacillus, Bacteroides
erm(H) Streptomyces
erm(I) Streptomyces
erm(N) Strepomyces
erm(O) Streptomyces
erm(Q) Actinobacilus, Clostridium, Streptococcus, Wolinella
erm(R) Aeromicrobium
erm(S) Streptomyces
erm(T) Lactobacillus
erm(U) Streptomyces
erm(V) Streptomyces
erm(W) Micromonospora
erm(X) Corynebacterium
erm(Y) Staphylococcus
clr Streptomyces
ATP-binding transporters
car(A) Strepomyces
msr(A) Staphylococcus
ole(B) Streptomyces
ole(C) Streptomyces
srm(B) Streptomyces
tlr(C) Streptomyces
vga(A) Staphylococcus
vga(B) Staphylococcus
Major facilitators
lmr(A) Streptomyces
mef(A) Corynebacterium, Enterococcus, Micrococcus Staphylococcus, Streptococcus
Esterases
ere(A) Enterobacter, Escherichia, Klebsiella, Citrobacter
ere(B) Escherichia, Klebsiella, Proteus
Hydrolases
vgb(A) Enterococcus, Staphylococcus
vgb(B) Staphylococcus
Transferases
lnu(A) Staphylococcus
lnu(B) Enterococcus
vat(A) Staphylococcus
vat(B) Staphylococcus
vat(C) Staphylococcus
vat(D) Enterococcus
vat(E) Enterococcus
Phosphorylases
mph(A) Escherichia
mph(B) Escherichia
mph(C) Staphylococcus
a

Data were taken from references 1, 2, 6, 8, 9, 14, 16, 19, 22, 27, 33, 34, 3741, 43, 44, 47, 48, 5357, 5967, 6985, 88, 90, 93, 94, 98, 100, 105, and 108110

The binding site in the 50S ribosomal subunit for erythromycin overlaps the binding site of the newer macrolides, as well as the structurally unrelated lincosamides and streptogramin B antibiotics. The modification by methylase(s) reduces the binding of all three classes of antibiotics, which results in resistance against macrolides, lincosamides, and streptogramin B antibiotics (MLSB). The rRNA methylases are the best studied among macrolide resistance mechanisms (47, 101103). However, a variety of other mechanisms have been described which also confer resistance (Table 2). Many of these alternative mechanisms of resistance confer resistance to only one or two of the antibiotic classes of the MLSB complex.

TABLE 2.

Efflux and inactivating genes

Resistance profile Protein name Gene included Genes % Homology
Plasmidb GenBank no. or referenced
DNA Amino acid
ATP-binding transporters
 Lincomycin Car(A) car(A) car(A) pOJ158 M80346
 Erythromycin Msr(A) msr(A) msr(A) 98 98 pUL5054 X52085
pSR1 AF167161
 Streptogramin B msr(SA) pEP2104 57
msr(SA′) pMS97 AB013298
msr(B) pCH200 M81802
 Oleandomycin Ole(B) ole(B) ole(B) pALOR26E L36601
 Oleandomycin Ole(C) ole(C) ole(C) L06249
 Spiramycin Srm(B) srm(B) srm(B) pKC514 X63451
 Tylosin Tlr(C) tlr(C) tlr(C) M57437
 Streptogramin A Vga(A) vga(A) vga pIP524 M90056
 Streptogramin A Vga(B) vga(B) vga(B) pIP1633 U82085
Major facilitators
 Lincomycin Lmr(A) lmr(A) lmr(A) pLST21 X59926
 Erythromycin Mef(A) mef(A) mef(A) 90 91 p53-6 U70055
 Erythromycin mef(E) pAT15-5 U83667
Esterases
 Erythromycin Ere(A) ere(A) ere(A) pI1100, pAT63 M11277
 Erythromycin Ere(B) ere(B) ere(B) 99 99 pIP1527 A15097
ere(B) pAT72 X03988
Hydrolases
 Streptogramin B Vgb(A) vgb(A) vgb pIP524 M20129
 Streptogramin B Vgb(B) vgb(B) vgb(B) pIP1714 AF015628
Transferases
 Lincomycin Lnu(A)a lnu(A)c lin(A′) pIP856 J03947
 Lincomycin lin(A) pIP855 M14039
 Lincomycin Lnu(B)a lnu(B)c lin(B) pVMM25 AJ238249
 Streptogramin A Vat(A) vat(A) vat pIP680 L07778
 Streptogramin A Vat(B) vat(B) vat(B) pIP524 L38809
 Streptogramin A Vat(C) vat(C) vat(C) pIP1714 AF015628
 Streptogramin A Vat(D)a vat(D) sat(A) pAT15, pAT421 L12033
 Streptogramin A Vat(E)a vat(E) sat(G) AF139725
Phosphorylases
 Macrolides Mph(A) mph(A) mph(A) 99 93–99 pTZ3519 D16251
 Macrolides mph(K) pGE64 U36578
 Macrolides Mph(B) mph(B) mph(B) pTZ3721, pTZ3723 D85892
 Macrolides Mph(C) mph(C) mph(BM) 100 100 pMS97 Q9ZNK8
 Macrolides mph(C) pSR1 AF167161
a

These are newly designated classes, proteins, and genes. 

b

The list of plasmids includes both recombinant and natural plasmids. 

c

New gene designation because the lin designation has been used for other genes. No genes use the lnu designation. 

d

Information provide by GenBank and references 18, 14, 16, 17, 19, 23, 28, 30, 33, 34, 3740, 42, 45, 4750, 54, 55, 60, 63, 64, 66, 67, 69, 70, 72, 73, 83, 85, 87, 90, 93, 94, 99, 105, 110

In this review, we suggest a new nomenclature for naming MLS genes and propose to use the rules developed for identifying and naming new tetracycline resistance genes (51, 52). This system, with a few recent modifications, was originally designed because of the ability of two genes to be distinguished uniquely by DNA-DNA probe methodology (51). It was generally found that two genes with <80% amino acid sequence identity provided enough variability in nucleotide sequence to permit distinct probes to be designed. Although many investigators are likely to sequence new genes, the use of probe technology allows rapid identification of isolates containing potentially new genes, as well as a reliable way to screen populations and determine the frequency of any one resistant determinant. Therefore, we continued this paradigm by assigning two genes of ≥80% amino acid identity to the same class and same letter designation, while two genes that show ≤79% amino acid identity are given a different letter designation. Table 1 shows the results of the classification, with some classes having members with little variability, while others, like classes A and O, show a greater range of homology at both the DNA and amino acid levels. As new gene sequences emerge, ideally they will need to be compared by oligonucleotide probe hybridization and/or sequence analysis against the bank of known genes before a new designation is assigned. If multiple genes are available in any one class, especially when there is a range as in class A, then all representative members of the class should be examined, not just one. To confirm that the proposed name or number for the newly discovered resistance determinant has not been used by another investigator, please contact M. C. Roberts for this information. A similar request has been made for new tet genes (52).

RRNA METHYLASES

Over the last 30 years, a large number of different rRNA methylase genes (erm) have been isolated from a variety of bacteria that range from E. coli to Haemophilus influenzae in gram-negative species and from Streptococcus pneumoniae to Corynebacterium spp. in gram-positive species (Table 3). In addition, a variety of gram-positive and gram-negative anaerobes, and even spirochetes such as Borrelia burgdorferi and Treponema denticola, have all been shown to carry erm genes (Table 3) (36, 77, 78). All erm enzymes methylate the same adenine residue, resulting in an MLSB phenotype (9, 100103). This adenine (A2058) or one of the adjacent residues in the peptidyltransferase region (A2057 or A2059) is changed to another nucleotide by mutation in macrolide-resistant Mycobacterium intracellulare, Mycobacterium avium, Propionibacterium spp., and Helicobacter pylori (58, 84, 100103).

Differences between the various erm genes are seen in the regulation of expression of the phenotype. Some of the enzymes are inducibly regulated by translational attenuation of a mRNA leader sequence; in the absence of erythromycin, the mRNA is in an inactive conformation due to a sequestered Shine-Dalgarno sequence, preventing efficient initiation of translation of the erm transcripts. Mutational analyses of the erm(C) leader peptide suggested that the peptide, (FS)IFVI, is critical for induction (103). However, when the erm peptides from the erm genes are compared, little sequence similarity is apparent (103). Recently, a second mechanism of regulation has been described in which the lack of erythromycin prevents the complete synthesis of the mRNA due to rho factor-independent termination. This type of regulation has been described for the erm(K) system (20), and by homology, we hypothesize that it may also exist for erm(D), as well as erm(J), because they are highly related and have been grouped together under class D (Table 1). In either system, inducible isolates, when tested, may appear to be susceptible or intermediately resistant to macrolides and susceptible to lincosamides. Erythromycin is generally a good inducer in most species; in animal or human streptococcal isolates, lincosamides and/or streptogramin B may be good inducers (47, 76). Good overviews of regulation of the erm genes can be found in recent reviews by Weisblum (100103).

Inducible strains predominated in the 1960s to 1970s. However, today it is more common in many geographical areas to find isolates that constitutively produce the rRNA methylase without preexposure to antibiotics. Constitutive erm gene expression is usually associated with structural alterations in the erm translational attenuator, including deletions, duplications, and point mutations in erm(C) (104). They can be distinguished from inducible isolates by the stable MICs for them regardless of whether they are pregrown with or without an inducer (76, 102).

Many of the erm genes are associated with conjugative or nonconjugative transposons which tend to reside on the chromosomes, although some have been found in plasmids. They are often associated with other antibiotic resistance genes, especially tetracycline resistance genes. The erm(F) gene is often linked with the tet(Q) gene, while the erm(B) gene is often linked with the tet(M) gene (24, 86, 95). These conjugative transposons can have a wide host range, which may explain why clinical isolates of many different bacterial species have been found to carry these erm genes (Table 3). The erm genes in general have low G+C contents (31 to 34%), while the overall chromosomal G+C contents found in gram-negative species are ≥50% and ∼35% in gram-positive species.

It has been common practice for investigators to give their erm gene a new name regardless of the DNA and predicted amino acid sequence similarity to previously characterized erm genes and without regard to whether the gene resides in a different isolate, species, or genus. The result has been that, over the years, the names of these erm genes have become confusing, and often a complex table is required to remember which genes are closely related (Table 1). In the worst cases, genes for unrelated enzymes have been given the same name (erm(A), causing confusion in the literature and GenBank listings (Table 1). The opposite also has occurred where very closely related or virtually identical enzymes have been given a variety of different names. For example, erm(F) (GenBank no. M14730) is found on the Bacteroides transposons Tn4351 and Tn4000 (71), erm(FS) (no. M17808) is on Bacteroides transposon Tn4551 (91), and erm(FU) (no. M62487) (32) is also from Bacteriodes. All three enzymes share ≥97% DNA and amino acid identity (Table 1). Since there are no phenotypic differences between the three erm(F) genes and distinguishing them by any method other than sequencing is problematic, we propose that all three should be known as class F: the Erm(F) protein and the erm(F) gene (Table 1).

The situation is even worse with class B, which is composed of a larger number of genes, including erm(AM), erm(B), erm(BC), erm(BP), and erm(Z), whose sequences share ≥98% homology (Table 1). Because the normal gene designation is to use a single letter (26) and the possibility of confusion between erm(A) and erm(AM), we propose that this group be known as class B: the erm(B) genes and the Erm(B) protein (Table 1). Recent dendrograms of many of the erm genes can be found in articles by Seppälä et al. (88) and Matsuoka et al. (56) and support this grouping of all of these genes within the class B designation.

To help those in the field, GenBank numbers or references for sequences that have not been deposited are listed in Table 1. If a new gene sequence shows ≥80% amino acid homology to any member of a gene class and confers a similar phenotype to the host, we propose that the new gene be placed in the existing group and not be given a new letter or number designation. Thus, with classes that show a wide range of homologies, like class A (81% amino acid homology) or class O (84% amino acid homology), multiple members must be compared to the new gene. Note that the class designation is based on the amino acid sequence of the structural gene only and does not include the various regulatory sequences that can occur upstream of the gene. These guidelines are intended to apply to all of the N-methyltransferases, regardless of whether the gene was originally identified in pathogenic, opportunistic, normal flora bacteria or an antibiotic-producing species. Once all of the single capital letters have been used to identify new erm genes, we recommend naming genes as follows: erm(30), erm(31), etc. This system has been proposed for naming of new tet genes [tet(30), etc.] (52). Furthermore, a similar set of guidelines should be adopted for the genes that encode other mechanisms of resistance to any of the MLS antibiotics (Table 1). Class Y for gene erm(GM), class S for gene erm(SF), class T for gene erm(GT), class V for gene erm(SV), class X for genes erm(CD), erm(CX), and erm(A), and class 2 for gene srm(D) are new class designations that conform to the single-letter designation (Table 1).

There are a number of other methylase genes, most often found in methylase-producing organisms which have not been given erm designations, such as tlr(D), car(B), myr(B), and smr(A). All are from species which confer resistance to a 16-membered ring macrolide (Table 1). We have grouped and renamed them classes H for car(B), I for mdm(A), N for tlr(D), O for genes lrm and srm(A), U for lmr(B), and W for myr(B). The clr gene could not be classified, because there is no sequence in the database or literature available. Less work has been done to determine if these genes are found outside their respective antibiotic producers (Table 3). erm genes are often linked with tet genes, and since genes conferring resistance to oxytetracycline, originally found in antibiotic-producing streptomycetes, are now found in some clinical Mycobacterium isolates, it is certainly possible that some erm genes have also moved into Mycobacterium spp. and other genera (68).

To prevent two unrelated genes from being given the same designation, we propose to establish a reference center, as has recently been recommended for tetracycline resistance genes. By using the guideline presented above in governing the identification of new erm genes, surveys can be conducted in bacterial populations to examine the spread of particular MLSB-resistant determinants. A single internal DNA fragment or oligonucleotide probe or a PCR assay that detects all members of a gene class can be established to screen large numbers of isolates. Not only will the adoption of a uniform naming system reduce the number of new erm gene names, but it will hopefully prevent confusion over unrelated genes being given the same designation and also prevent highly related genes from having different gene designations.

EFFLUX SYSTEMS

A number of different antibiotic resistance genes code for transport (efflux) proteins. These do not modify either the antibiotic or the antibiotic target, but instead pump the antibiotic out of the cell or the cellular membrane, keeping intracellular concentrations low and ribosomes free from antibiotic. Many of these proteins [mef(A), mef(E), and lmr(A)] have homology to the major facilitator superfamily (MFS) of efflux proteins. Others [car(A), msr(A), msr(B), ole(B), ole(C) srm(B), tlr(C), vga, and vga(B)] are putative members of the ABC transporter superfamily (70). In early years, most macrolide resistance was mediated by the presence of erm genes. However, more recently, other mechanisms of macrolide resistance have been found in increasing frequency in certain gram-positive populations (23, 27, 41, 43, 44, 92, 93, 106). Three different efflux systems which confer resistance have been described for gram-positive cocci [msr(A) (macrolide and streptogramin B resistant), mef(A) (macrolide efflux), and vga and vga(B) (virginiamycin factor A)] (4) (Table 2). Besides the academic interest in these genes, their presence in an erythromycin-resistant bacterial pathogen of interest may also have implications in terms of therapeutic choices. If an isolate carries a mef gene, clindamycin can be considered, whereas the presence of an erm(B) gene would preclude consideration of a lincosamide. Recently, we and others have identified Streptococcus pneumoniae strains which carry both mef and erm(B) genes and, as expected, have the MLSB phenotype (41, 53).

The mef genes have been found in a variety of gram-positive genera, including corynebacteria, enterococci, micrococci, and a variety of streptococcal species (30, 43, 53, 90) (Table 3), suggesting a much wider distribution of this group of genes than originally imagined. Many of these genes are associated with conjugative elements located in the chromosome and are readily transferred conjugally across species and genus barriers (43, 53).

Two mef genes have been characterized in the literature: mef(A) (23) and mef(E) (94). The mef(A) gene was described in Streptococcus pyogenes, while the mef(E) gene was found in S. pneumoniae. Since the two genes share 90% DNA and 91% amino acid homology (Table 2), we recommended that these two genes be considered a single class, A: mef(A) gene and Mef(A) protein (Table 2).

The msr(A), msr(SA), msr(SA)′, and msr(B) group differs from the mef genes because they confer resistance to both macrolide and streptogramin B antibiotics (MS) (13, 5557). The msr(B) gene is roughly half the size of msr(A), but very homologous to it. Though this gene is significantly shorter than the msr(A) gene sequence, we placed it with the other msr genes (Table 2).

In antibiotic producers, there are efflux pumps specific for MLSB antibiotics that generally belong to the ABC transporter superfamily (87). They include car(A) from Streptomyces thermotolerans (87), ole(B) from Streptomyces antibioticus (7, 80), srm(B) from Streptomyces ambofaciens (73), lmr(C) from Streptomyces lincolnensis (70), and tlr(C) from Streptomyces fradiae (87). In addition to the msr(A) efflux pumps, there are two efflux systems identified in staphylococci that confer resistance to streptogramin A antibiotics, vga and vga(B) (4). Besides mef(A), other efflux proteins that appear to be fueled by the proton motive force have been described for MLSB antibiotics. A lincomycin-specific efflux pump encoded by lmr(A) has been described in S. lincolnensis (110).

OTHER MECHANISMS

A variety of other mechanisms which usually confer resistance to only one of the three classes (M, L, or S) or one component such as streptogramin A, but not streptogramin B, have been described (103) (Table 2). These proteins modify the antibiotic rather than the rRNA target or serve as pumps that shuttle the antibiotic out of the bacterial cell. Enzymes which hydrolyze streptogramin B [vgb (virginiamycin factor B hydrolase), vgb(B) genes] or modify the antibiotic by adding an acetyl group (acetyltransferases) to streptogramin A [vat (virginiamycin, factor A acetylation), vat(B), vat(C), sat(A), and sat(G) genes] have been described (16) (Table 2). Many of these genes are plasmid borne, and often these vat-related genes [vat, vat(B), and vat(C) genes] are downstream of other genes encoding resistance to streptogramins [vgb, vga(B), and vgb(B) genes, respectively] in staphylococci (2), but not in enterococci (72). The acetyltransferase genes are related, in the active site region, to a novel chloramphenicol acetyltransferase family of enzymes. We have renamed sat(A) as vat(D) and sat(G) as vat(E) to simplify the nomenclature (Table 2).

Unlike most of the other genes described in this review, both the ere (erythromycin esterification) and mph (macrolide phosphotransferase) genes (Table 2) were first described in E. coli rather than gram-positive cocci (8, 63, 64, 66, 67). According to our guidelines, mph(K) has been reassigned to mph(A), because there are only 10 amino acid (1%) differences between the two proteins. mph(BM) and mph(C) (66a) are grouped under Mph(C), because these genes are nearly identical to each other and distinct from mph(A) and mph(B) (Table 2). Several lincomycin nucleotidyltransferases have been identified: lin(A) in Staphylococcus haemolyticus (16), lin(A)′ in Staphylococcus aureus (17), and lin(B) in Enterococcus faecium (14). We propose changing lin(A) and lin(B) to lnu(A) and lnu(B) (for lincomycin nucleotidyltransferase), because the former letters have already been used for gamma BHC dehydrochlorinase and cyclohexadiene hydrolase genes. It is suggested that prior to naming a new gene class, it is necessary to determine if the proposed three-letter designation has been used for other previously characterized genes.

CONCLUSIONS

With the introduction of the newer, more stable macrolides with enhanced properties, there has been a significant increase in macrolide usage. Macrolides like azithromycin and clarithromycin are recommended for prophylactic use to prevent Mycobacterium avium complex disease in human immunodeficiency virus patients. As macrolide use increases, so does its exposure to bacterial populations, increasing the opportunity for bacteria to acquire macrolide or MLS resistance. Given that intragenic transfer of macrolide-resistant determinants is possible (15), it is likely that all of the genes described in this review will spread into new species and that new genes will be identified. Therefore, it is important to clarify the nomenclature of these resistance genes for their expanding audience.

ACKNOWLEDGMENTS

We thank M. Matsuoka for providing unpublished material; B. Weisblum for discussions; J. Davies, C. J. Smith, and S. Schwarz for reading the manuscript; and S. Lerner for doing sequence comparisons.

REFERENCES

  • 1.Allignet J, El Solh N. Diversity among the gram-positive acetyltransferases inactivating streptogramin A and structurally related compounds and characterization of a new staphylococcal determinant, vatB. Antimicrob Agents Chemother. 1995;39:2027–2036. doi: 10.1128/aac.39.9.2027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Allignet J, Liassine N, El Solh N. Characterization of a staphylococcal plasmid related to pUB110 and carrying two novel genes, vatC and vgbB, encoding resistance to streptogramins A and B and similar antibiotics. Antimicrob Agents Chemother. 1998;42:1794–1798. doi: 10.1128/aac.42.7.1794. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Allignet J, Loncle V, Simenel C, Delepierre M, El Solh N. Sequence of a staphylococcal gene, vat, encoding an acetyltransferase inactivating the A-type compounds of virginiamycin-like antibiotics. Gene. 1993;130:91–98. doi: 10.1016/0378-1119(93)90350-c. [DOI] [PubMed] [Google Scholar]
  • 4.Allignet J, Loncle V, El Solh N. Sequence of a staphylococcal plasmid gene, vga, encoding a putative ATP-binding protein involved in resistance to virginianmycin A-like antibiotics. Gene. 1992;117:45–51. doi: 10.1016/0378-1119(92)90488-b. [DOI] [PubMed] [Google Scholar]
  • 5.Allignet J, Loncle V, Mazodier P, El Solh N. Nucleotide sequence of a staphylococcal plasmid gene, vgb, encoding a hydrolase inactivating the B components of virginiamycin-like antibiotics. Plasmid. 1988;20:271–275. doi: 10.1016/0147-619x(88)90034-0. [DOI] [PubMed] [Google Scholar]
  • 6.Allignet J, El Solh N. Diversity among the gram-positive acetyltransferases inactivating streptogramin A and structurally related compounds and characterization of a new staphylococcal determinant, vatB. Antimicrob Agents Chemother. 1995;39:2027–2036. doi: 10.1128/aac.39.9.2027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Aparicio G, Buche A, Mendez C, Salas J-A. Characterization of the ATPase activity of the N-terminal nucleotide binding domain of an ABC transporter involved in oleandomycin secretion by Streptomyces antibioticus. FEMS Microbiol Lett. 1996;141:157–162. doi: 10.1111/j.1574-6968.1996.tb08378.x. [DOI] [PubMed] [Google Scholar]
  • 8.Arthur M, Andremont A, Courvalin P. Distribution of erythromycin esterase and rRNA methylase genes in members of the family Enterobacteriaceae highly resistant to erythromycin. Antimicrob Agents Chemother. 1987;31:404–409. doi: 10.1128/aac.31.3.404. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Arthur M, Brisson-Noel A, Courvalin P. Origin and evolution of genes specifying resistance to macrolides, lincosamides and streptogramin antibiotics: data and hypothesis. J Antimicrob Chemother. 1987;20:783–802. doi: 10.1093/jac/20.6.783. [DOI] [PubMed] [Google Scholar]
  • 10.Berryman D I, Rood J I. Cloning and hybridization analysis of ermP, a macrolide-lincosamide-streptogramin B resistance determinant from Clostridium perfringens. Antimicrob Agents Chemother. 1989;33:1346–1353. doi: 10.1128/aac.33.8.1346. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Berryman D I, Rood J I. The closely related ermB-ermAM genes from Clostridium perfringens, Enterococcus faecalis (pAMβ1), and Streptococcus agalactiae (pIP501) are flanked by variants of a directly repeated sequence. Antimicrob Agents Chemother. 1995;39:1830–1834. doi: 10.1128/aac.39.8.1830. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Berryman D I, Lyristis M, Rood J I. Cloning and sequence analysis of ermQ, the predominant macrolide-lincosamide-streptogramin B resistance gene in Clostridium perfringens. Antimicrob Agents Chemother. 1994;38:1041–1046. doi: 10.1128/aac.38.5.1041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Beyer D, Pepper K. The streptogramin antibiotics: update on their mechanism of action. Exp Opin Investig Drugs. 1998;7:591–599. doi: 10.1517/13543784.7.4.591. [DOI] [PubMed] [Google Scholar]
  • 14.Bozdogan B, Berrezouga L, Kuo M-S, Yurek D A, Farley K A, Stockman B J, LeClercq R. A new resistance gene, linB, conferring resistance to lincosamides by nucleotidylation in Enterococcus faecium HM1025. Antimicrob Agents Chemother. 1999;43:925–929. doi: 10.1128/aac.43.4.925. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Brisson-Noël A, Arthur M, Courvalin P. Evidence for natural gene transfer from gram-positive cocci to Escherichia coli. J Bacteriol. 1988;170:1739–1745. doi: 10.1128/jb.170.4.1739-1745.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Brisson-Noel A, Courvalin P. Nucleotide sequence of gene linA encoding resistance to lincosamides in Staphylococcus haemolyticus. Gene. 1986;43:247–253. doi: 10.1016/0378-1119(86)90213-1. [DOI] [PubMed] [Google Scholar]
  • 17.Brisson-Noel A, Delrieu P, Samain D, Courvalin P. Inactivation of lincosaminide antibiotics in Staphylococcus. Identification of lincosaminide O-nucleotidyltransferases and comparison of the corresponding resistance genes. J Biol Chem. 1988;263:15880–15887. [PubMed] [Google Scholar]
  • 18.Calcutt M J, Cundliffe E. Cloning of a lincosamide resistance determinant from Streptomyces caelestis, the producer of celesticetin, and characterization of the resistance mechanism. J Bacteriol. 1990;172:4710–4714. doi: 10.1128/jb.172.8.4710-4714.1990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Cheng J, Grebe T, Wondrack L, Courvalin P, Sutcliffe J. Program and abstracts of the 39th Interscience Conference on Antimicrobial Agents and Chemotherapy. Washington, D.C.: American Society for Microbiology; 1999. Characterization of genes involved in erythromycin resistance in a clinical strain of Staphylococcus aureus, abstr. 837; p. 114. [Google Scholar]
  • 20.Choi S-S, Kim S-K, Oh T-G, Choi E-C. Role of mRNA termination in regulation of ermK. J Bacteriol. 1997;179:2065–2067. doi: 10.1128/jb.179.6.2065-2067.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Chu D. Recent developments in 14- and 15-membered macrolides. Exp Opin Investig Drugs. 1995;4:65–94. [Google Scholar]
  • 22.Chung W O, Werckenthin C, Schwarz S, Roberts M C. Host range of the ermF rRNA methylase gene in human and animal bacteria. J Antimicrob Chemother. 1999;43:5–14. doi: 10.1093/jac/43.1.5. [DOI] [PubMed] [Google Scholar]
  • 23.Clancy J, Petitpas J W, Dib-Hajj F, Yuan W, Cronan M, Kamath A, Bergeron J, Retsema J. Molecular cloning and functional analysis of a novel macrolide-resistance determinant mefA from Streptococcus pyogenes. Mol Microbiol. 1996;22:867–879. doi: 10.1046/j.1365-2958.1996.01521.x. [DOI] [PubMed] [Google Scholar]
  • 24.Clewell D B, Flannagan S E, Jaworski D D. Unconstrained bacterial promiscuity: the Tn916-Tn1545 family of conjugative transposons. Trends Microbiol. 1995;3:229–236. doi: 10.1016/s0966-842x(00)88930-1. [DOI] [PubMed] [Google Scholar]
  • 25.Cooper A J, Shoemaker N B, Salyers A A. The erythromycin resistance gene from the Bacteroides conjugal transposon Tcr Emr 7853 is nearly identical to ermG from Bacillus sphaericus. Antimicrob Agents Chemother. 1996;40:506–508. doi: 10.1128/aac.40.2.506. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Council of Biology Editors, Inc. CBE style manual: a guide for authors, editors, and publishers in the biological sciences. 5th ed. Bethesda, Md: Council of Biology, Editors, Inc.; 1983. [Google Scholar]
  • 27.Eady E A, Ross J I, Tipper J L, Walters C E, Cove J H, Noble W C. Distribution of genes encoding erythromycin ribosomal methylases and an erythromycin efflux pump in epidemiologically distinct groups of staphylococci. J Antimicrob Chemother. 1993;31:211–217. doi: 10.1093/jac/31.2.211. [DOI] [PubMed] [Google Scholar]
  • 28.Epp J K, Burgett S G, Schoner G E. Cloning and nucleotide sequence of a carbomycin-resistance gene from Streptomyces thermotolerans. Gene. 1987;53:73–83. doi: 10.1016/0378-1119(87)90094-1. [DOI] [PubMed] [Google Scholar]
  • 29.Farrow, K. A., D. Lyras, and J. I. Rood. GenBank accession no. AF109075.
  • 30.Fraimow H, Knob C. Abstracts of the 98th General Meeting of the American Society for Microbiology. Washington, D.C.: American Society for Microbiology; 1997. Amplification of macrolide efflux pumps msr and mef from Enterococcus faecium by polymerase chain reaction, abstr. A-125; p. 22. [Google Scholar]
  • 31.Hächler H, Kayser F H, Berger-Bächi B. Homology of a transferable tetracycline resistance determinant of Clostridium difficile with Streptococcus (Enterococcus) faecalis transposon Tn916. Antimicrob Agents Chemother. 1987;31:1033–1038. doi: 10.1128/aac.31.7.1033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Halula M, Manning S, Macrina F L. Nucleotide sequence of ermFU, macrolide-lincosamide-streptogramin (MLS) resistance gene encoding an RNA methylase from the conjugal element of Bacteroides fragilis V503. Nucleic Acids Res. 1991;19:3453. doi: 10.1093/nar/19.12.3453. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Hammerum A M, Jensen L, Bogo L, Aarestrup F M. Detection of the satA gene and transferability of virginiamycin resistance in Enterococcus faecium from food-animals. FEMS Microbiol Lett. 1998;168:145–151. doi: 10.1111/j.1574-6968.1998.tb13267.x. [DOI] [PubMed] [Google Scholar]
  • 34.Hara Q, Hutchinson C R. Cloning of midecamycin (MLS)-resistance genes from Streptomyces mycarofaciens, Streptomyces lividans and Streptomyces coelicolor A3(2) J Antibiot (Tokyo) 1990;43:977–991. doi: 10.7164/antibiotics.43.977. [DOI] [PubMed] [Google Scholar]
  • 35.Hodgson A L M, Krywult J, Radford A J. Nucleotide sequence of the erythromycin resistance gene from Corynebacterium plasmid pNG2. Nucleic Acids Res. 1990;18:1891. doi: 10.1093/nar/18.7.1891. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Hudson C R, Roberts M C, Gherardini F C. Abstracts of the 98th Annual Meeting of the American Society for Microbiology. Washington, D.C.: American Society for Microbiology; 1998. Evidence of conjugal transfer of an erythromycin-resistance determinant in Borrelia burgdorferi, abstr. D-2; p. 223. [Google Scholar]
  • 37.Inouye M, Morohoshi T, Horinouchi S, Beppu T. Cloning and sequences of two macrolides-resistance-encoding genes from mycinamicin-producing Micromonospora griseorubida. Gene. 1994;141:39–46. doi: 10.1016/0378-1119(94)90125-2. [DOI] [PubMed] [Google Scholar]
  • 38.Jenkins G, Zalacain M, Cundliffe E. Inducible ribosomal RNA methylation in Streptomyces lividans, conferring resistance to lincomycin. J Gen Microbiol. 1989;129:2703–2714. doi: 10.1099/00221287-135-12-3281. [DOI] [PubMed] [Google Scholar]
  • 39.Jensen L B, Hammerum A M, Aarestrup F M, van den Bogaard A E, Stobberingh E E. Occurrence of satA and vgb genes in streptogramin-resistant Enterococcus faecium isolates of animal and human origins in The Netherlands. Antimicrob Agents Chemother. 1998;42:3330–3331. doi: 10.1128/aac.42.12.3330. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Jensen L B, Frimodt-Moller N, Aarestrup F M. Presence of erm gene classes in Gram-positive bacteria of animal and human origin in Denmark. FEMS Microbiol Lett. 1999;170:151–158. doi: 10.1111/j.1574-6968.1999.tb13368.x. [DOI] [PubMed] [Google Scholar]
  • 41.Johnston N J, de Azavedo J C, Kellner J D, Low D E. Prevalence and characterization of the mechanisms of macrolide, lincosamide, and streptogramin resistance in isolates of Streptococcus pneumoniae. Antimicrob Agents Chemother. 1998;42:2425–2426. doi: 10.1128/aac.42.9.2425. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Kamimiya S, Weisblum B. GenBank deposit: Streptomyces viridochromogenes rRNA (adenine-N6-) methyltransferase, ermSV gene. Accession no. U59450. 1996. [Google Scholar]
  • 43.Kataja J, Seppälä H, Skurnik M, Sarkkinen H, Huovinen P. Different erythromycin resistance mechanisms in group C and group G streptococci. Antimicrob Agents Chemother. 1998;42:1493–1494. doi: 10.1128/aac.42.6.1493. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Kataja J, Huovinen P, Skurnik M, Seppälä H the Finnish Study Group for Antimicrobial Resistance. Erythromycin resistance genes in group A streptococci in Finland. Antimicrob Agents Chemother. 1999;43:48–52. doi: 10.1128/aac.43.1.48. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Kim S-K, Baek M-C, Choi S-S, Kim B-K, Choi E-C. Nucleotide sequence, expression and transcriptional analysis of the Escherichia coli mphK gene encoding macrolide-phosphotransferase K. Mol Cells. 1996;6:153–160. [Google Scholar]
  • 46.Kovalic D, Giannattasio R B, Jin H-J, Weisblum B. 23S rRNA domain V, a fragment that can be specifically methylated in vitro by the ErmSF (TlrA) methyltransferase. J Bacteriol. 1994;176:6992–6998. doi: 10.1128/jb.176.22.6992-6998.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Leclercq R, Courvalin P. Bacterial resistance to macrolide, lincosamide, and streptogramin antibiotics by target modification. Antimicrob Agents Chemother. 1991;35:1267–1272. doi: 10.1128/aac.35.7.1267. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Leclercq R, Courvalin P. Intrinsic and unusual resistance to macrolide, lincosamide, and streptogramin antibiotics in bacteria. Antimicrob Agents Chemother. 1991;35:1273–1276. doi: 10.1128/aac.35.7.1273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Le Goffic F, Capmau M L, Bonnet M L, Cerceau C, Soussy C J, Dublanchet A, Duval J. Plasmid-mediated pristinamycin resistance: PH1A, a pristinamycin 1A hydrolase. Ann Inst Pasteur. 1977;128:471–474. [PubMed] [Google Scholar]
  • 50.Le Goffic F, Capmau M L, Bonnet M L, Cerceau C, Soussy C J, Dublanchet A, Duval J. Plasmid-mediated pristinamycin resistance: PCIIA, a new enzyme which modifies pristinamycin IIA. J Antibiot. 1977;30:665–669. doi: 10.7164/antibiotics.30.665. [DOI] [PubMed] [Google Scholar]
  • 51.Levy S B, McMurry L M, Burdett V, Courvalin P, Hillen W, Roberts M C, Taylor D E. Nomenclature for tetracycline resistance determinants. Antimicrob Agents Chemother. 1989;33:1373–1374. doi: 10.1128/aac.33.8.1373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Levy S B, McMurry L M, Barbosa T M, Burdett V, Courvalin P, Hillen W, Roberts M C, Rood J I, Taylor D E. Nomenclature for new tetracycline resistance determinants. Antimicrob Agents Chemother. 1999;43:1523–1524. doi: 10.1128/aac.43.6.1523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Luna V A, Coates P, Eady E A, Cove J, Nguyen T T H, Roberts M C. A variety of Gram-positive bacteria carry mobile mef genes. J Antimicrob Chemother. 1999;44:19–25. doi: 10.1093/jac/44.1.19. [DOI] [PubMed] [Google Scholar]
  • 54.Matsuoka M, Inoue M, Nakajima Y. A mechanism of resistance to partial macrolide and streptogramin B antibiotics in Staphylococcus aureus clinically isolated in Hungary. Biol Pharm Bull. 1995;18:1482–1486. doi: 10.1248/bpb.18.1482. [DOI] [PubMed] [Google Scholar]
  • 55.Matsuoka M, Inoue M, Nakajima Y. A dyadic plasmid that shows MLS and PMS resistance in Staphylococcus aureus. FEMS Microbiol Lett. 1997;148:91–96. doi: 10.1111/j.1574-6968.1997.tb10272.x. [DOI] [PubMed] [Google Scholar]
  • 56.Matsuoka M, Inoue M, Nakajima Y. Abstracts of the 38th Interscience Conference on Antimicrobial Agents and Chemotherapy. Washington, D.C.: American Society for Microbiology; 1998. A new class of erm genes mediating MLS-coresistance in Staphylococcus aureus: it resides on plasmid pMS97 together with msrSA′ gene coding for an active efflux pump, abstr. C-35; p. 78. [Google Scholar]
  • 57.Matsuoka M, Endou K, Kobayashi H, Inoue M, Nakajima Y. A plasmid that encodes three genes for resistance to macrolide antibiotics in Staphylococcus aureus. FEMS Microbiol Lett. 1998;167:221–227. doi: 10.1111/j.1574-6968.1998.tb13232.x. [DOI] [PubMed] [Google Scholar]
  • 58.Meier A, Kirschner P, Springer B, Steingrube V A, Brown B A, Wallace R J, Jr, Böttger E C. Identification of mutations in the 23S rRNA gene of clarithromycin-resistant Mycobacterium intracellulare. Antimicrob Agents Chemother. 1994;38:381–384. doi: 10.1128/aac.38.2.381. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Miller E S, Woese C R, Brenner S. Description of the erythromycin-producing bacterium Arthrobacter sp. strain NRRL B-3381 as Aeromicrobium erythreum gen. nov., sp. nov. Int J Syst Bacteriol. 1991;41:363–368. doi: 10.1099/00207713-41-3-363. [DOI] [PubMed] [Google Scholar]
  • 60.Milton I D, Hewitt C L, Harwood C R. Cloning and sequencing of a plasmid-mediated erythromycin resistance determinant from Staphylococcus xylosus. FEMS Microbiol Lett. 1992;76:141–147. doi: 10.1016/0378-1097(92)90377-z. [DOI] [PubMed] [Google Scholar]
  • 61.Monod M, Denoya C, Dubnau D. Sequence and properties of pIM13, a macrolide-lincosamide-streptogramin B resistance plasmid from Bacillus subtilis. J Bacteriol. 1986;167:138–147. doi: 10.1128/jb.167.1.138-147.1986. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Monod M, Mohan S, Dubnau D. Cloning and analysis of ermG, a new macrolide-lincosamide-streptogramin B resistance element from Bacillus sphaericus. J Bacteriol. 1987;169:340–350. doi: 10.1128/jb.169.1.340-350.1987. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Noguchi N, Emura A, Matsuyama H, O’Hara K, Sasatsu M, Kono M. Nucleotide sequence and characterization of erythromycin resistance determinant that encodes macrolide 2′-phosphotransferase I in Escherichia coli. Antimicrob Agents Chemother. 1995;39:2359–2363. doi: 10.1128/aac.39.10.2359. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Noguchi N, Katayama J, O’Hara K. Cloning and nucleotide sequence of the mphB gene for macrolide 2′-phosphotransferase II in Escherichia coli. FEMS Microbiol Lett. 1996;144:197–202. doi: 10.1111/j.1574-6968.1996.tb08530.x. [DOI] [PubMed] [Google Scholar]
  • 65.Oh T-G, Kwon A-R, Choi E-C. Induction of ermAMR from a clinical strain of Enterococcus faecalis by 16-membered-ring macrolide antibiotics. J Bacteriol. 1998;180:5788–5791. doi: 10.1128/jb.180.21.5788-5791.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.O’Hara K, Kanda T, Ohmiya K, Ebisu T, Kono M. Purification and characterization of macrolide 2′-phosphotransferase from a strain of Escherichia coli that is highly resistant to erythromycin. Antimicrob Agents Chemother. 1989;33:1354–1357. doi: 10.1128/aac.33.8.1354. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66a.O’Hara, K. Personal communication.
  • 67.Ounissi H, Courvalin P. Nucleotide sequence of the gene ereA encoding the erythromycin esterase in Escherichia coli. Gene. 1985;35:271–278. doi: 10.1016/0378-1119(85)90005-8. [DOI] [PubMed] [Google Scholar]
  • 68.Pang Y, Brown B A, Steingrube V A, Wallace R J, Jr, Roberts M C. Tetracycline resistance determinants in Mycobacterium and Streptomyces species. Antimicrob Agents Chemother. 1994;38:1408–1412. doi: 10.1128/aac.38.6.1408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Pernodet J L, Blondelet-Rouault M H, Guerineau M. Resistance to spiramycin in Streptomyces ambofaciens, the producer organism, involves at least two different mechanisms. J Gen Microbiol. 1993;139:1003–1011. doi: 10.1099/00221287-139-5-1003. [DOI] [PubMed] [Google Scholar]
  • 70.Peschke U, Schmidt H, Zhang H-Z, Piepersberg W. Molecular characterization of the lincomycin-production gene cluster of Streptomyces lincolnensis 78-11. Mol Microbiol. 1995;16:1137–1156. doi: 10.1111/j.1365-2958.1995.tb02338.x. [DOI] [PubMed] [Google Scholar]
  • 71.Rasmussen J L, Odelson D A, Macrina F L. Complete nucleotide sequence and transcription of ermF, a macrolide-lincosamide-streptogramin B resistance determinant from Bacteroides fragilis. J Bacteriol. 1986;168:523–533. doi: 10.1128/jb.168.2.523-533.1986. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Rende-Fournier R, LeClercq R, Galimand M, Duval J, Courvalin P. Identification of the satA gene encoding a streptogramin A acetyltransferase in Enterococcus faecium BM4145. Antimicrob Agents Chemother. 1993;37:2119–2125. doi: 10.1128/aac.37.10.2119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Richardson M A, Kuhstoss S, Solenberg P, Schaus N A, Rao R N. A new shuttle cosmid vector, pKC505, for streptomycetes: its use in the cloning of three different spiramycin-resistance genes from a Streptomyces ambovaciens library. Gene. 1987;61:231–241. doi: 10.1016/0378-1119(87)90187-9. [DOI] [PubMed] [Google Scholar]
  • 74.Roberts A N, Hudson G S, Brenner S. An erythromycin-resistance gene from an erythromycin-producing strain of Arthrobacter sp. Gene. 1985;35:259–270. doi: 10.1016/0378-1119(85)90004-6. [DOI] [PubMed] [Google Scholar]
  • 75.Roberts M C. Distribution of tetracycline and macrolides-lincosamides-streptogramin B (MLS) genes in anaerobic bacteria. Clin Infect Dis. 1995;20:S367–S369. doi: 10.1093/clinids/20.supplement_2.s367. [DOI] [PubMed] [Google Scholar]
  • 76.Roberts M C, Brown M B. Macrolide-lincosamide resistance determinants in streptococcal species isolated from the bovine mammary gland. Vet Microbiol. 1994;40:253–261. doi: 10.1016/0378-1135(94)90114-7. [DOI] [PubMed] [Google Scholar]
  • 77.Roberts M C, Chung W O, Roe D E. Characterization of tetracycline and erythromycin determinants in Treponema denticola. Antimicrob Agents Chemother. 1996;40:1690–1694. doi: 10.1128/aac.40.7.1690. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Roberts M C, Chung W O, Roe D, Xia M, Marquez C, Borthagaray G, Whittington W L, Holmes K K. Erythromycin-resistant Neisseria gonorrhoeae and oral commensal Neisseria spp. carry known rRNA methylase genes. Antimicrob Agents Chemother. 1999;43:1367–1372. doi: 10.1128/aac.43.6.1367. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Roberts M C, McFarland L V, Mullany P, Mulligan M E. Characterization of the genetic basis of antibiotic resistance in Clostridium difficile. J Antimicrob Chemother. 1994;33:419–429. doi: 10.1093/jac/33.3.419. [DOI] [PubMed] [Google Scholar]
  • 80.Rodriguez A M, Olano C, Vilches C, Mendez C, Salas J A. Streptomyces antibioticus contains at least three olendomycin resistance determinants, one of which shows homology with proteins of the ABC-transporter superfamily. Mol Microbiol. 1993;8:571–582. doi: 10.1111/j.1365-2958.1993.tb01601.x. [DOI] [PubMed] [Google Scholar]
  • 81.Roe D E, Weinberg A, Roberts M C. Mobility of rRNA methylase genes in Campylobacter (Wolinella) rectus. J Antimicrob Chemother. 1995;36:738–740. doi: 10.1093/jac/36.4.738. [DOI] [PubMed] [Google Scholar]
  • 82.Roe D E, Weinberg A, Roberts M C. Mobile rRNA methylase genes in Actinobacillus actinomycetemcomitans. J Antimicrob Chemother. 1996;37:457–464. doi: 10.1093/jac/37.3.457. [DOI] [PubMed] [Google Scholar]
  • 83.Ross J I, Eady E A, Cove J H, Baumberg S. Minimal functional system required for expression of erythromycin resistance by msrA in Staphylococcus aureus RN4220. Gene. 1996;183:143–148. doi: 10.1016/s0378-1119(96)00541-0. [DOI] [PubMed] [Google Scholar]
  • 84.Ross J I, Eady E A, Cove J H, Jones C E, Ratyal A H, Miller Y W, Vyakrnam S, Cunliffe W J. Clinical resistance to erythromycin and clinidamycin in cutaneous propionibacteria isolated from acne patients is associated with mutations in 23S rRNA. Antimicrob Agents Chemother. 1989;41:1162–1165. doi: 10.1128/aac.41.5.1162. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Rosteck P R, Jr, Reynolds P A, Hershberger C L. Homology between proteins controlling Streptomyces fradiae tylosin resistance and ATP-binding transport. Gene. 1991;102:27–32. doi: 10.1016/0378-1119(91)90533-h. [DOI] [PubMed] [Google Scholar]
  • 86.Salyers A A, Shoemaker N B, Stevens A M, Li L-Y. Conjugative transposons: an unusual and diverse set of integrated gene transfer elements. Microbiol Rev. 1995;59:579–590. doi: 10.1128/mr.59.4.579-590.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Schoner B, Geistlich M, Rosteck P I, Jr, Rao R N, Seno E, Reynolds P, Cox K, Burgett S, Hershberger C. Sequence similarity between macrolide-resistance determinants and ATP-binding transport proteins. Gene. 1992;115:93–96. doi: 10.1016/0378-1119(92)90545-z. [DOI] [PubMed] [Google Scholar]
  • 88.Seppälä H, Skurnik M, Soini H, Roberts M C, Huovinen P. A novel erythromycin resistance methylase gene (ermTR) in Streptococcus pyogenes. Antimicrob Agents Chemother. 1998;42:257–262. doi: 10.1128/aac.42.2.257. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Shoemaker N B, Barber R D, Salyers A A. Cloning and characterization of a Bacteroides conjugal tetracycline-erythromycin resistance element by using a shuttle cosmid vector. J Bacteriol. 1989;171:1294–1302. doi: 10.1128/jb.171.3.1294-1302.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Shortridge V D, Flamm R K, Ramer N, Beyer J, Tanaka S K. Novel mechanism of macrolide resistance in Streptococcus pneumoniae. Diagn Microbiol Infect Dis. 1996;26:73–78. doi: 10.1016/s0732-8893(96)00183-6. [DOI] [PubMed] [Google Scholar]
  • 91.Smith C J. Nucleotide sequence analysis of Tn4551: use of ermFS operon fusions to detect promoter activity in Bacteroides fragilis. J Bacteriol. 1987;169:4589–4596. doi: 10.1128/jb.169.10.4589-4596.1987. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Sutcliffe J, Grebe T, Tait-Kamradt A, Wondrack L. Detection of erythromycin-resistant determinants by PCR. Antimicrob Agents Chemother. 1996;40:2562–2566. doi: 10.1128/aac.40.11.2562. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Sutcliffe J, Tait-Kamradt A, Wondrack L. Streptococcus pneumoniae and Streptococcus pyogenes resistant to macrolides but sensitive to clindamycin: a common resistance pattern mediated by an efflux system. Antimicrob Agents Chemother. 1996;40:1817–1824. doi: 10.1128/aac.40.8.1817. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Tait-Kamradt A, Clancy J, Cronan M, Dib-Hajj F, Wondrack L, Yuan W, Sutcliffe J. mefE is necessary for the erythromycin-resistant M phenotype in Streptococcus pneumoniae. Antimicrob Agents Chemother. 1997;41:2251–2255. doi: 10.1128/aac.41.10.2251. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Trieu-Cuot P, Poyart-Salmeron C, Carlier C, Courvalin P. Nucleotide sequence of the erythromycin resistance gene of the conjugative transposon Tn1545. Nucleic Acids Res. 1990;18:3660. doi: 10.1093/nar/18.12.3660. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Vergis E N, Yu V L. Macrolides are ideal for empiric therapy of community-acquired pneumonia in the immunocompetent host. Semin Respir Infect. 1997;12:322–328. [PubMed] [Google Scholar]
  • 97.Vergis E N, Yu V L. New macrolides or new quinolones as monotherapy for patients with community-acquired pneumonia; our cup runneth over? Chest. 1998;113:1158–1159. doi: 10.1378/chest.113.5.1158. [DOI] [PubMed] [Google Scholar]
  • 98.Wasteson Y, Robe D E, Falk K, Roberts M C. Characterization of tetracycline and erythromycin resistance in Actinobacillus pleuropneumoniae. Vet Microbiol. 1996;48:41–50. doi: 10.1016/0378-1135(95)00130-1. [DOI] [PubMed] [Google Scholar]
  • 99.Weber J M, Leung J O, Main G T, Potenz R H B, Paulus T J, DeWitt J P. Organization of a cluster of erythromycin genes in Saccharopolyspora erythraea. J Bacteriol. 1990;172:2372–2383. doi: 10.1128/jb.172.5.2372-2383.1990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Weisblum B. Resistance to macrolide-lincosamide-streptogramin antibiotics. In: Fischetti V A, editor. Gram-positive pathogens. Washington, D.C.: American Society for Microbiology; 1999. pp. 682–698. [Google Scholar]
  • 101.Weisblum B. Erythromycin resistance by ribosome modification. Antimicrob Agents Chemother. 1995;39:577–585. doi: 10.1128/AAC.39.3.577. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Weisblum B. Insights into erythromycin action from studies of its activity as inducer of resistance. Antimicrob Agents Chemother. 1995;39:797–805. doi: 10.1128/aac.39.4.797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Weisblum B. Macrolide resistance. Drug Resist Update. 1998;1:29–41. doi: 10.1016/s1368-7646(98)80212-4. [DOI] [PubMed] [Google Scholar]
  • 104.Werckenthin C, Schwarz S, Westh H. Structural alterations in the translational attenuator of constitutively expressed ermC genes. Antimicrob Agents Chemother. 1999;43:1681–1685. doi: 10.1128/aac.43.7.1681. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Werner G, Witte W. Characterization of a new enterococcal gene, satG, encoding a putative acetyltransferase conferring resistance to streptogramin A compounds. Antimicrob Agents Chemother. 1999;43:1813–1814. doi: 10.1128/aac.43.7.1813. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Widdowson C A, Klugman K P. Emergence of the M phenotype of erythromycin-resistant pneumococci in South Africa. Emerg Infect Dis. 1998;4:277–281. doi: 10.3201/eid0402.980216. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Young H, Moyes A, McMillan A. Azithromycin and erythromycin resistant Neisseria gonorrhoeae following treatment with azithromycin. Int J Sex Transm Dis AIDS. 1997;8:299–302. doi: 10.1258/0956462971920127. [DOI] [PubMed] [Google Scholar]
  • 108.Zalacain M, Cundliffe E. Methylation of 23S rRNA caused by tlrA (ermSF), a tylosin resistance determinant from Streptomyces fradiae. J Bacteriol. 1989;171:4254–4260. doi: 10.1128/jb.171.8.4254-4260.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Zalacain M, Cundliffe E. Cloning of tlrD, a fourth resistance gene, from the tylosin producer, Streptomyces fradiae. Gene. 1991;97:137–142. doi: 10.1016/0378-1119(91)90021-3. [DOI] [PubMed] [Google Scholar]
  • 110.Zhang H-Z, Schmidt H, Piepersberg W. Molecular cloning and characterization of two lincomycin-resistance genes, lmrA and lmrB, from Streptomyces lincolnensis 78-11. Mol Microbiol. 1992;6:2147–2157. doi: 10.1111/j.1365-2958.1992.tb01388.x. [DOI] [PubMed] [Google Scholar]

Articles from Antimicrobial Agents and Chemotherapy are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES