Review

Expert Review of Anti-infective Therapy Downloaded from informahealthcare.com by CDL-UC San Diego on 01/01/15 For personal use only.

Antibiotic resistance in Enterococcus faecium clinical isolates Expert Rev. Anti Infect. Ther. 12(2), 239–248 (2014)

Vincent Cattoir*1,2 and Jean-Christophe Giard2 1 CHU de Caen, Service de Microbiologie & Centre National de Re´fe´rence sur la Re´sistance aux Antibiotiques (Laboratoire Associe´ Ente´rocoques), F-14033 Caen, France 2 Universite´ de Caen Basse-Normandie, EA 4655 U2RM, F-14032 Caen, France *Author for correspondence: Tel.: +33 231 064 572 Fax: +33 231 064 573 [email protected]

The worldwide ratio of Enterococcus faecalis-Enterococcus faecium infections is currently changing in favor of E. faecium. Intrinsic and acquired antimicrobial resistance traits of this latter species can explain this evolution as well as the diffusion of hospital-adapted strains belonging to the clonal complex CC17. Like other enterococci, E. faecium is naturally resistant to cephalosporins and aminoglycosides (at low level). Because of its high genome plasticity, it can also acquire numerous other resistances. It is noteworthy that most modern isolates of E. faecium are highly resistant to ampicillin while a non-negligible proportion of them (depending on geographical locations) are resistant to glycopeptides (especially in the USA). Even if resistance to newer antimicrobial agents (linezolid, daptomycin, tigecycline) is still uncommon, some clinical isolates with reduced susceptibility or resistance have already been reported and better understanding of resistance mechanisms is needed for prediction and prevention of their dissemination. KEYWORDS: antimicrobial resistance • CC17 • E. faecium • glycopeptide resistance • VREF

Although they are part of the normal intestinal flora of humans, enterococci are a leading cause of hospital-acquired infections with the emergence of multidrug-resistant (MDR) isolates, especially vancomycin-resistant enterococci (VRE) [1]. The spread of VRE isolates has been associated with significant increases in mortality, length of hospital stay and healthcare costs [2]. Due to the development of MDR isolates and the paucity of newer antimicrobial agents, therapeutic options for the treatment of enterococcal infections has become challenging. Approximately 80% of enterococcal infections are caused by Enterococcus faecalis while ca. 20% are due to Enterococcus faecium, which is increasingly reported [3]. Notably, E. faecium is part of the ‘no ESKAPE’ pathogens besides Staphylococcus aureus, Klebsiella pneumoniae, Acinetobacter baumannii, Pseudomonas aeruginosa and Enterobacter spp., which comprise major MDR opportunistic organisms [4]. In addition, there has been since the end of the 1980s, the worldwide dissemination of a subpopulation of E. faecium highly resistant to ampicillin and fluoroquinolones that acquired, at a later stage, resistance to vancomycin, commonly referred to as vancomycin-resistant E. faecium (VREF). These

www.expert-reviews.com

10.1586/14787210.2014.870886

hospital-adapted isolates belong to the so-called clonal complex CC17, and are responsible for numerous hospital outbreaks due to their important colonization and persistence capacities [5]. Beside its selective advantage provided by intrinsic and acquired antibiotic resistance traits (TABLE 1), E. faecium is well equipped with many virulence factors such as adhesion and secreted proteins that are involved in their epidemiological success in hospital settings [3]. The aim of this review is to update the knowledge on mechanistic and epidemiological traits of resistance recently published for E. faecium. The text is mainly focused on the most recent findings, particularly on new antimicrobial agents and emerging resistance mechanisms. b-lactams

b-lactam antibiotics inhibit the last steps of the peptidoglycan synthesis by binding to the high-molecular-weight penicillin-binding proteins (PBPs) [6]. Enterococci are intrinsically resistant to cephalosporins by production of low-affinity PBPs. By contrast, penicillins (especially ampicillin) alone or combined with an aminoglycoside represent the drugs of choice for the treatment of enterococcal

 2014 Informa UK Ltd

ISSN 1478-7210

239

Review

Cattoir & Giard

Table 1. Main mechanisms of antimicrobial resistance in Enterococcus faecium. Antibiotics

Mechanism of resistance

Expert Review of Anti-infective Therapy Downloaded from informahealthcare.com by CDL-UC San Diego on 01/01/15 For personal use only.

Intrinsic resistance Cephalosporins

Low-affinity PBP

Aminoglycosides (low level)

Low-level uptake

Tobramycin (moderate level)

Enzymatic inactivation (AAC[6´]-Ii) Ribosomal methylation (EfmM)

Acquired resistance Ampicillin (high level)

Overproduction/alterations of PBP5

Gentamicin (high level)

Enzymatic inactivation (AAC[6´]-Ie-APH[2´´]-Ia)

Streptomycin (high level)

Alterations of S12 ribosomal protein

Glycopeptides

Precursor modification (VanA, VanB)

Fluoroquinolones

Alterations in DNA gyrase and topoisomerase IV

Macrolides

Ribosomal methylation (Erm)

Linezolid

Ribosomal mutations (23S rRNA)

Daptomycin

Complex changes in cell envelope

PBP: Penicillin-binding protein.

infections. However, ampicillin resistance, which is exceptional in E. faecalis, has become very frequent in E. faecium, representing 95–100% and 70–95% of VREF and E. faecium clinical isolates, respectively [7–14]. In this latter species, the resistance is mainly due to overexpression and/or mutations of a gene coding for the low-affinity PBP5 [15–17]. An alternate mechanism of high-level ampicillin resistance has been selected in vitro, being due to the production of b-lactamresistant L,D-transpeptidase and D,D-carboxypeptidase bypassing the D,D-transpeptidase activity of the PBPs [18]. As opposed to PBP5-mediated ampicillin resistance, this L,Dtranspeptidase is inhibited by imipenem [19]. Fortunately, this novel mechanism of resistance has not yet been reported in clinical isolates. Finally, other resistance determinants have been recently identified using genome-wide analysis, especially ddcP coding for a low-molecular-weight PBP with D, D-carboxypeptidase activity [20]. Aminoglycosides

Aminoglycosides are fast-acting bactericidal antimicrobial agents inhibiting the protein biosynthesis through the binding to the conserved A-site of 16S rRNA on the 30S small subunit of the bacterial ribosome [21]. Although enterococci present a lowlevel resistance to these antibiotics, the combination of cell 240

wall-active agents (e.g., penicillins, glycopeptides) and aminoglycosides results in a significant bactericidal synergy. Indeed, the alteration of the parietal structure allows a better penetration of aminoglycosides [22]. Moreover, it has been proposed that the synergistic effect is likely due to the formation of reactive oxygen species contributing to cell death [23]. The intrinsic resistance of enterococci (MICs from 4 to 256 mg/l) is due to a limited drug uptake [24]. In addition, acquired high-level resistance to aminoglycosides (usually MICs ‡2000 mg/l) is commonly found among enterococcal clinical isolates [24,25]. The main mechanism of resistance is the production of aminoglycoside-modifying enzymes (AMEs): aminoglycoside acetyltransferases (AACs), aminoglycoside phosphotransferases (APHs) and aminoglycoside nucleotidyltransferases (ANTs) [21]. Noteworthy, E. faecium is intrinsically resistant to tobramycin and kanamycin (preventing their combination with penicillins or glycopeptides) by production of two chromosomal enzymes: AAC(6´)-Ii and EfmM, a 16S rRNA methyltransferase [26,27]. Therefore, gentamicin and streptomycin constitute the drugs of choice for the treatment of severe infections caused by E. faecium, and testing for high-level resistance to these compounds should be performed [24]. High-level resistance to streptomycin results from single-step mutations in the S12 ribosomal protein or the acquisition of genes encoding ANT(3´´)-Ia or ANT(6´)-Ia enzymes [25]. High-level gentamicin resistance is usually due to the acquisition of a transposon harboring the bifunctional enzyme AAC(6´)-Ie-APH(2´´)-Ia, responsible for cross-resistance to all clinically available aminoglycosides, except streptomycin [25]. Also, several other determinants have been occasionally described in high-level gentamicin resistance, such as APH(2´´)-Ib, APH(2´´)-Ic or APH(2´´)-Id [25]. Finally, two other AMEs, APH(3´)-IIIa and ANT(4´´)-Ia, have been identified in E. faecium, responsible for the low-level resistance to kanamycin and to tobramycin-kanamycin-amikacin, respectively [25]. Around 50–60% of recent E. faecium clinical isolates are highly resistant to streptomycin whereas high-level resistance to gentamicin ranges from 20 to 80% depending on the country [7–9,28] Glycopeptides

Although E. faecalis has caused almost all enterococcal infections since the early 1990s, E. faecium is the most difficult species to treat due to its multidrug resistance. Furthermore, E. faecium slowly but progressively replaces E. faecalis in human infections [29]. The epidemiological success of E. faecium can be partially explained by its great genomic plasticity, which enables it to cope with numerous environmental stresses including antimicrobial pressure. Recently, the comparative analysis of the genomic sequence of 51 E. faecium strains reveals that the epidemic hospital-adapted lineage emerged approximately 75 years ago, which coincides with the beginning of antibiotics area [30]. This illustrates the spectacular ability of E. faecium to remodel its genome contents. For proof, the clonal complex CC17 (based on multilocus sequence typing – MLST – scheme), Expert Rev. Anti Infect. Ther. 12(2), (2014)

Expert Review of Anti-infective Therapy Downloaded from informahealthcare.com by CDL-UC San Diego on 01/01/15 For personal use only.

Antibiotic resistance in Enterococcus faecium

has been well characterized as a subpopulation of E. faecium particularly well hospital-adapted, highly pathogenic and resistant to ampicillin and fluoroquinolones but not always to vancomycin [5]. Moreover, it has been reported that between 2002 and 2006, the increased number of infections caused by enterococci in Danish hospitals was mainly due to E. faecium isolates belonging to the CC17 [31]. The CC17 strains belong to the hospital-associated clade that is genetically distant from the community-associated clade. They harbor the insertion sequence 16 (IS16) and are highly resistant to ampicillin and fluoroquinolones [32–34]. The first VREF has been isolated in 1986 in Europe (after 30 years of glycopeptide use) and since were disseminated worldwide especially in hospital environments [1,30,35–37]. Data on enterococcal infections collected from several US hospitals (84,050 isolates) between 2010 and 2012 revealed that 20.6% of the isolates were resistant to vancomycin. While E. faecium correspond to 24% of the isolates, one-third are vancomycin resistant that represents 75% of the glycopeptide-resistant enterococci strains [38]. This is in accordance with a previous study showing that up to 80% of E. faecium clinical isolates recovered from US hospitals are vancomycin resistant [39]. In Europe, the prevalence of vancomycin resistance among E. faecium isolates highly diverges according to the country and ranged from more than 30% (i.e., Ireland, Greece) to less than 1% (Nordic countries) [37]. Vancomycin and teicoplanin are the two glycopeptidic antibiotics used to treat serious human infections due to Grampositive bacteria. These high-molecular-weight molecules do not enter into the cytoplasm but interact with the D-Ala-D-Ala terminus of the pentapeptide precursors of the peptidoglycan [40]. The formation of a stable complex involving five hydrogen bonds leads to the blockage of the transglycosylation and transpeptidation reactions in cell wall synthesis. Consequently, the precursors accumulate inside the cell, the cell wall loses its integrity and the bacteria die. Like with penicillins, glycopeptides are usually combined with an aminoglycoside because of the bactericidal synergistic effect. The very sophisticated mechanism of resistance to glycopeptides is based on the presence of operons encoding enzymes that synthesize new precursors with low affinity (where the last D-Ala is changed by D-Lac or D-Ser) (ligases), and that eliminate or prevent the formation of the native precursor with high affinity (D,D-dipeptidases and carboxypeptidases). Precursors with the D-Ala-D-Lac terminus have 1000-fold less affinity to vancomycin than those ending in D-Ala-D-Ala resulting in a high level of resistance (MIC >16 mg/l). On the other hand, the precursors ending with D-Ala-D-Ser, those are sevenfold less affine to vancomycin leads to a low level of resistance (MIC of 8–16 mg/l) [41]. Eight different operons involved in the acquired resistance to glycopeptides (vanA, B, D, E, G, L, M, N) and an intrinsic one (vanC1/C2/C3/C4) have been characterized in enterococci. Among them, only vanA, B, D, M and N were found in E. faecium (TABLE 2) [1,42,43]. Of note, the vanN is the www.expert-reviews.com

Review

sole operon leading to the synthesis of D-Ala-D-Ser ending precursors in E. faecium and may be specific for this strain [44,45]. VanA and VanB types are the most frequently retrieved in VREF and are both borne by transposons (Tn1546 and Tn1547, respectively) [41]. VanA isolates are inducible resistant to high level of vancomycin and teicoplanin (MIC >64 mg/l). The expression of genes of the operon is controlled by a two-component regulatory system vanRS. On the other hand, VanB-resistant strains of E. faecium exhibit various levels of vancomycin resistance (MIC from 4 to 1000 mg/l) but remain susceptible to teicoplanin that does not induce the vanB operon. It is also worth noting that ampicillin-resistant mutants producing an L,D-transpeptidase activity also appears to be resistant to glycopeptides [46]. Fluoroquinolones

Fluoroquinolones are bactericidal agents acting against a large panel of Gram-negative and Gram-positive bacteria. However, they show a limited antimicrobial activity against enterococci, although newer compounds (i.e., levofloxacin and moxifloxacin) are more active. High-level acquired resistance results from point mutations in gyrA and parC genes encoding subunits A of DNA gyrase and topoisomerase IV, respectively [47]. Interestingly, high-level fluoroquinolone resistance appears to be increasingly distributed among hospital-adapted E. faecium clinical isolates, especially those belonging to the clonal lineage CC17 [48,49]. Indeed, around 80% of E. faecium and 95% of VREF (all belonging to the CC17) were resistant to levofloxacin [7–14]. Moreover, some findings strongly suggest that a NorA-like efflux pump exists in E. faecium, which may be involved in resistance to hydrophilic fluoroquinolones [50]. Quinupristin–dalfopristin

Quinupristin–dalfopristin is an injectable streptogramin that was the first drug approved by the US FDA (1999) for the treatment of VRE infections. It actually consists of two semisynthetic compounds combined in a 30:70 ratio (w/w): quinupristin (streptogramin B) and dalfopristin (streptogramin A) [51]. Individually, these compounds are bacteriostatic against most Gram-positive bacteria, but they synergistically become bactericidal when combined [52]. Streptogramins A and B act by interfering with bacterial protein synthesis by binding to adjacent regions within the P site of 23S rRNA of the 50S ribosomal subunit [53]. A variety of mechanisms confer resistance to streptogramins A or B, including modifying enzymes, drug efflux and modification of the ribosomal target [54]. The methylation of the 23S rRNA is the most common resistance mechanism to streptogramins, which is due to methylases encoded by erm genes (especially erm[B]), leading to cross-resistance to macrolides–lincosamides–streptogramins B (MLSB phenotype). This latter phenotype is frequently found in E. faecium [55,56], and around 95% of VREF clinical isolates have been found to be resistant to erythromycin and clindamycin in France [9]. As opposed to E. faecalis that is intrinsically resistant to lincosamides and streptogramins A (LSA phenotype) due to the 241

Review

Cattoir & Giard

Table 2. Types of resistance to glycopeptides described in Enterococcus faecium. Type of resistance VanA

Expert Review of Anti-infective Therapy Downloaded from informahealthcare.com by CDL-UC San Diego on 01/01/15 For personal use only.

VanM

VanB

VanD

VanN

Level of resistance

High

High

Variable

Moderate

Low

MIC (mg/l) Vancomycin

64–1000

>256

4–1000

64–128

16

Teicoplanin

16–512

48–128

0.5–1

4–64

1

Expression

Ind.

?

Ind.

Const.

Const.

Location

Plasmid (chromosome)

Plasmid (chromosome)

Chromosome (plasmid)

Chromosome (plasmid)

Chromosome

Conjugative transfer

+

+

+

-

+

Precursors end

D-Ala-D-Lac

D-Ala-D-Lac

D-Ala-D-Lac

D-Ala-D-Lac

D-Ala-D-Ser

Prevalence

Very high

Very low

High

Low

Very low

Const.: Constitutive; Ind.: Inducible. Data taken from [1,36].

synthesis of the ABC protein Lsa(A), E. faecium is naturally susceptible. However, LSA-resistant E. faecium can be selected in vitro and in vivo, bringing into play a unique point mutation in the domestic gene eat(A) coding for an ABC protein related to Lsa(A) [57]. Interestingly, this novel resistance phenotype seems nonetheless common in VREF, since it has been detected among 23% of the 60 VREF clinical isolates collected in France [57]. Streptogramins B can also be enzymatically inactivated by lyases encoded by the vgb(A) gene while acetyltransferases [encoded by vat(D), vat(E), and vat(H) genes] degrade streptogramins A [56]. Note that vat(D) and vat(E) genes have also been detected in animals due to the use of the growth promoter virginiamycin [58]. Finally, a putative efflux mechanism, mediated by the ABC protein VgaD, has been detected in some E. faecium clinical isolates [25]. Linezolid

Linezolid, first oxazolidinone launched since April 2000, is indicated in the treatment of infections caused by Grampositive bacteria, including VRE (600 mg b.i.d.). Linezolid alters protein synthesis by binding to the 23S RNA and preventing the assembly of the functional 70S initiation complex [59]. Like macrolides, lincosamides, tetracyclines and chloramphenicol, oxazolidinones are essentially bacteriostatic, and they exhibit a ca. 2-h post-antibiotic effect. Notably, linezolid is equally active against vancomycin-susceptible and resistant E. faecium isolates [60], with MIC50 and MIC90 both at 1–2 and 2–4 mg/l, respectively [7,10,11,13,61–65]. Despite the difficulty of in vitro selection (frequency ~10-9), linezolid resistance can emerge during therapy [66], especially but not exclusively after a prolonged therapy [67,68]. In enterococci, it is generally associated with a point mutation (G2576T) in the central loop of domain V of the 23S rRNA, also described in linezolid-resistant staphylococci [64,69–73]. Interestingly, the level of resistance is correlated with the proportion of mutated 23S rDNA alleles. E. faecium harbors 242

6 copies, and linezolid MIC is 2, 8–16, 32 and 64 mg/l in the presence of 0, 1, 2 or 3, and 4 or 5 mutated copies, respectively [74]. Another mutation (G2505A) was also identified in an in vitro selected resistant E. faecium mutant [75]. MICs of resistant enterococcal strains range from 8 to 128 mg/l, the CLSI susceptibility breakpoint being £4 mg/l [67,70,71,76]. In staphylococci, other point mutations in domain V of 23S rRNA have also been described as well as mutations in ribosomal proteins L3 and L4, but they have not been yet reported in enterococci [77–82]. More recently, another mechanism of resistance has been described, which is due to the production of the chloramphenicol florfenicol resistance (Cfr) protein that specifically methylates the 23S rRNA at the A2503 residue [83]. Initially identified in staphylococcal isolates from animal sources, it has since been detected in some human S. aureus and one E. faecalis clinical isolates, but not in E. faecium so far [84–86]. Interestingly, Cfr causes cross-resistance to five different antibiotic families: phenicols, lincosamides, oxazolidinones, pleuromutilins and streptogramins A (the so-called PhLOPSA phenotype) [87]. Except for Cfr, none of the mechanisms of resistance to other protein synthesis inhibitors confers cross-resistance to linezolid [88]. A large majority of E. faecium clinical isolates remain susceptible to linezolid with a prevalence of resistance usually lower than 2% irrespective of the resistance phenotype to vancomycin [7,8,10–14,62,64,89]. However, some hospitals have reported notable rates of resistance (5–10%) among VRE isolates, particularly in the USA [11,62], while hospital outbreaks caused by linezolidresistant VREF have been also reported [69]. Daptomycin

Daptomycin is a cyclic lipopeptide antibiotic that exerts a potent and rapid bactericidal activity against Gram-positive bacteria, including MDR E. faecium isolates [90]. It has been available in the USA since September 2003 and it is commonly employed for the treatment of VRE infections, even if it does Expert Rev. Anti Infect. Ther. 12(2), (2014)

Expert Review of Anti-infective Therapy Downloaded from informahealthcare.com by CDL-UC San Diego on 01/01/15 For personal use only.

Antibiotic resistance in Enterococcus faecium

not have FDA approval in this indication [8]. This off-label use is supported by retrospective studies that have shown similar outcomes for patients with severe enterococcal infections treated by daptomycin (median dosage at 6 mg/kg/day intravenously) or linezolid [91–94]. Daptomycin irreversibly interacts with the bacterial cell membrane in a calcium-dependent manner, leading to the disruption in its integrity with an efflux of potassium ions and the dissipation of the ion concentration gradient. Cell death occurs as a result in DNA, RNA and protein synthesis [59]. The development of resistance to daptomycin remains uncommon in enterococci, with spontaneous in vitro resistance frequencies ~10-9 [95]. However, several failures of daptomycin therapy for E. faecium bacteremia have been reported with the emergence of high-level resistance (MIC >32 mg/l) [96–98]. This resistance can rapidly emerge (after less than 10 days of exposure), especially in patients with disorders of calcium homeostasis, low doses of daptomycin or end-stage renal failure [98]. Note that daptomycin activity against E. faecium is independent of the presence of van operons or linezolid resistance [93]. As previously described in E. faecalis, some chromosomal mutations in genes involved in the structure of the cell envelope and biophysical properties of the cell membrane are responsible for resistance in E. faecium [99,100]. Notably, several in-frame deletions have been identified either in genes coding for enzymes involved in phospholipid metabolism, such as GdpD (glycerophosphoryl diester phosphodiesterase) and Cls (cardiolipin synthetase) or in a gene coding for a putative membrane protein named LiaF, which is part of the three-component system LiaFSR (lipide-II interacting antibiotics) known to regulate the response of the cell envelope to antimicrobial agents in Bacillus subtilis [101,102]. Other mutations have been identified in other genes putatively involved in daptomycin resistance, but their role need to be confirmed [103,104]. Particularly, amino acid changes were found in the essential two-component system YycFG (analogous to LiaSFR) that regulates cell envelope homeostasis and cell division [104]. Daptomycin-resistant enterococcal isolates have marked ultrastructural changes in the cell envelope with cell wall thickening, increased net positive surface charge and decreased cell membrane fluidity as well as significant reduction in the ability of daptomycin to depolarize and permeabilize the cell membrane [99,105]. These modifications are associated with biochemical and biophysical alterations in cell membrane lipid metabolism, particularly in phospholipid content [105]. It is worth noting that some therapy failures have been reported in patients suffering from bacteremia caused by E. faecium isolates exhibiting MICs (3–4 mg/l) close to the CLSI susceptibility breakpoint (£4 mg/l) [106,107]. Interestingly, there is a strong association between elevated daptomycin MICs and the presence of mutations in the liaFSR locus. These mutations are a prerequisite for high-level resistance in combination with subsequent mutations in gdpD and/or cls genes [99,100]. Therefore, a daptomycin MIC of 3–4 mg/l may be an indicator of possible treatment failure, corresponding to a first-step mutant predisposed to develop in vivo high-level resistance. www.expert-reviews.com

Review

All surveillance studies conducted in Europe and North America have shown daptomycin susceptibility rates close to 100% in E. faecium with MIC50 and MIC90 around 1 and 2 mg/l, respectively, regardless of the resistance to vancomycin [64,65,89,93]. For instance, in a recent study conducted in France, 100% of 60 epidemiologically unrelated VREF clinical isolates were categorized as susceptible to daptomycin with a range of MICs from 0.25 to 4 mg/l [65]. Tigecycline

Tigecycline is the first representative of the novel glycylcycline class, which exhibit a potent activity against a broad spectrum of Gram-positive and Gram-negative bacteria, including VRE isolates [108]. It is currently approved by the FDA for the treatment of complicated skin and skin-structure infections and complicated intra-abdominal infections (June 2005) as well as community-acquired bacterial pneumonia (May 2009). Tigecycline must be administered parentally at the dosage of 50 mg/ kg every 12 h after an initial dose of 100 mg/kg [108]. This bacteriostatic antibiotic acts through binding to the bacterial 30S ribosomal subunit and blocking entry of the amino-acyl tRNAs into the A site of the ribosome, leading to the inhibition of protein biosynthesis [59]. Interestingly, tigecycline is not affected by the two major mechanisms of tetracycline resistance, that is, active efflux (e.g., tet[K] in enterococci) and ribosomal protection (e.g., tet[M] in enterococci) [109]. Also, common mechanisms of resistance to other antibiotic classes do not affect tigecycline activity. Resistance is difficult to select in vitro (frequency ~10-9) mainly related to overexpression of efflux pump systems in Gram-negatives whereas acquired resistance remains very uncommon in Gram-positives [110]. Indeed, enterococcal clinical isolates with MIC values above the EUCAST susceptible breakpoint (£0.25 mg/l) are exceptional, while high-level resistant strains (MIC >1 mg/l) have not been reported yet. All TEST (Tigecycline Evaluation and Surveillance Trial) international studies have reported tigecycline susceptibility rates close to 100% in E. faecium with MIC50 and MIC90 of 0.06–0.12 and 0.12–0.25 mg/l, respectively, regardless of the resistance to vancomycin [10–13,62,63,111]. In a recent French study, only 2 of 60 epidemiologically unrelated VREF clinical isolates were categorized as non-susceptible to tigecycline with an MIC at 0.5 mg/l [65]. However, molecular mechanism of this reduced susceptibility phenotype remains unknown. Expert commentary & five-year view

Over the years, E. faecium has demonstrated its potential to acquire a variety of resistance determinants and to adapt to different hostile environments, leading to its emergence and diffusion in the hospital settings. This is particularly true for ampicillin- and vancomycin-resistant isolates belonging to the CC17. Toward new antimicrobial agents (such as linezolid and daptomycin), acquired resistance is also emerging in E. faecium, mainly due to mutations in chromosomal genes. In addition, there is a potential risk of emergence of plasmid-mediated 243

Expert Review of Anti-infective Therapy Downloaded from informahealthcare.com by CDL-UC San Diego on 01/01/15 For personal use only.

Review

Cattoir & Giard

linezolid resistance, since the cfr gene has been identified several times in E. faecalis [112]. Several new anti-Gram-positive antibiotics have been very recently commercialized. For instance, ceftaroline is a novel cephalosporin with activity against methicillin-resistant S. aureus and penicillin-resistant pneumococci, but enterococci remain resistant (MIC >32 mg/l) to this drug [113]. Some other drugs (e.g., lipoglycopeptides) are going into development as potential options for the VREF treatment but their number is limited and their activity is variable [114]. Plazomicin is a next-generation aminoglycoside that retains activity against both Gram-negative and Gram-positive bacterial strains expressing AMEs, but its activity against enterococci does not seem to be clinically relevant [115]. Besides classical antimicrobial therapies, other alternative therapeutic options are promising and should be encouraged, such as immunotherapy and bacteriophage therapy [116,117]. To conclude, large efforts

must be pursued in order to prevent development and spread of antibiotic resistance in E. faecium through infection control policies and antibiotic stewardship programs. Acknowledgement

The authors warmly thank A Hartke for critical reading of the manuscript and insightful comments. Financial & competing interests disclosure

The authors have no relevant affiliations or financial involvement with any organization or entity with a financial interest in or financial conflict with the subject matter or materials discussed in the manuscript. This includes employment, consultancies, honoraria, stock ownership or options, expert testimony, grants or patents received or pending or royalties. No writing assistance was utilized in the production of this manuscript.

Key issues • Enterococcus faecium is becoming one of the leading causes of hospital-acquired infections. There is currently a shift in clinical significance from Enterococcus faecalis to this species. • The high genome plasticity of E. faecium is one of the major characteristics that may explain why it successfully adapt to harsh conditions such as the hospital environment and how it can cope with antibiotic and antiseptic stresses. • Its inherent resistance to antimicrobial agents such as cephalosporins and aminoglycosides (low level) as well as to its ability to acquire and share new antibiotic resistance traits (especially against ampicillin, gentamicin [high level] and vancomycin), causes more and more therapeutic problems. • For the treatment of infections caused by vancomycin-resistant E. faecium (VREF) isolates, a limited arsenal is currently available but some molecules (e.g., linezolid, daptomycin or tigecycline) can be alternative options. However, acquired resistance to these novel drugs is possible and has been already detected in vitro and in vivo. • It is necessary to maintain the vigilance face to the dissemination of well hospital-adapted E. faecium strains (in particular, VREF belonging to the clonal complex CC17) and to have rigorous policies of hygiene and antimicrobial use.

References

4.

Rice LB. Federal funding for the study of antimicrobial resistance in nosocomial pathogens: no ESKAPE. J Infect Dis 2008; 197(8):1079-81

9.

5.

Top J, Willems R, Bonten M. Emergence of CC17 Enterococcus faecium: from commensal to hospital-adapted pathogen. FEMS Immunol Med Microbiol 2008; 52(3):297-308

Bourdon N, Fines-Guyon M, Thiolet JM, et al. Changing trends in vancomycin-resistant enterococci in French hospitals. 2001-08. J Antimicrob Chemother 2011;66(4):713-21

10.

6.

Rice LB. Mechanisms of resistance and clinical relevance of resistance to beta-lactams, glycopeptides, and fluoroquinolones. Mayo Clin Proc 2012; 87(2):198-208

Dowzicky MJ, Chmelarova E. Global in vitro activity of tigecycline and linezolid against Gram-positive organisms collected between. 2004 and 2009. Int J Antimicrob Agents 2011;37(6):562-6

11.

7.

Pfaller MA, Sader HS, Jones RN. Evaluation of the in vitro activity of daptomycin against 19,615 clinical isolates of Gram-positive cocci collected in North American hospitals. 2002Diagn Microbiol Infect Dis 2007;57(4):459-65

Dowzicky MJ. Susceptibility to tigecycline and linezolid among gram-positive isolates collected in the United States as part of the tigecycline evaluation and surveillance trial (TEST) between. 2004 and 2009. Clin Ther 2011;33(12):1964-73

12.

8.

Putnam SD, Sader HS, Moet GJ, et al. Worldwide summary of telavancin spectrum and potency against Gram-positive

Balode A, Punda-Polic V, Dowzicky MJ. Antimicrobial susceptibility of Gram-negative and Gram-positive bacteria collected from countries in Eastern Europe:

Papers of special note have been highlighted as: • of interest •• of considerable interest 1.

Cattoir V, Leclercq R. Twenty-five years of shared life with vancomycin-resistant enterococci: is it time to divorce? J Antimicrob Chemother 2013;68(4):731-42

2.

DiazGranados CA, Zimmer SM, Klein M, Jernigan JA. Comparison of mortality associated with vancomycin-resistant and vancomycin-susceptible enterococcal bloodstream infections: a meta-analysis. Clin Infect Dis 2005;41(3):327-33

3.

Arias CA, Murray BE. The rise of the Enterococcus: beyond vancomycin resistance. Nat Rev Microbiol 2012;10(4): 266-78

••

Key review showing the enterococcal determinants involved in virulence as well as mechanisms of antibiotic resistance.

244

pathogens. 2007 to 2008 surveillance results. Diagn Microbiol Infect Dis 2010; 67(4):359-68

Expert Rev. Anti Infect. Ther. 12(2), (2014)

Antibiotic resistance in Enterococcus faecium

results from the Tigecycline Evaluation and Surveillance Trial (T.E.S.T.). 2004-2010. Int J Antimicrob Agents 2013;41(6):527-35 13.

Expert Review of Anti-infective Therapy Downloaded from informahealthcare.com by CDL-UC San Diego on 01/01/15 For personal use only.

14.

15.

16.

17.

18.

19.

Brandon M, Dowzicky MJ. Antimicrobial susceptibility among Gram-positive organisms collected from pediatric patients globally between. 2004 and 2011: results from the Tigecycline Evaluation and Surveillance Trial. J Clin Microbiol 2013; 51(7):2371-8 Sader HS, Flamm RK, Jones RN. Antimicrobial activity of daptomycin tested against Gram-positive pathogens collected in Europe, Latin America. and selected countries in the Asia-Pacific Region (2011). Diagn Microbiol Infect Dis 2013;75(4): 417-22 Fontana R, Grossato A, Rossi L, et al. Transition from resistance to hypersusceptibility to beta-lactam antibiotics associated with loss of a low-affinity penicillin-binding protein in a Streptococcus faecium mutant highly resistant to penicillin. Antimicrob Agents Chemother 1985;28(5): 678-83 Williamson R, le Bouguenec C, Gutmann L, Horaud T. One or two low affinity penicillin-binding proteins may be responsible for the range of susceptibility of Enterococcus faecium to benzylpenicillin. J Gen Microbiol 1985;131(8):1933-40

23.

Barnes AI, Herrero IL, Albesa I. New aspect of the synergistic antibacterial action of ampicillin and gentamicin. Int J Antimicrob Agents 2005;26(2):146-51

24.

Chow JW. Aminoglycoside resistance in enterococci. Clin Infect Dis 2000;31(2): 586-9

25.

Hollenbeck BL, Rice LB. Intrinsic and acquired resistance mechanisms in enterococcus. Virulence 2012;3(5):421-33



Key review of mechanisms underlying antibiotic resistance in Enterococcus faecium and Enterococcus faecalis clinical isolates.

26.

Costa Y, Galimand M, Leclercq R, et al. Characterization of the chromosomal AAC (6’)-Ii gene specific for Enterococcus faecium. Antimicrob Agents Chemother 1993;37(9): 1896-903

27.

28.

European Centre for Disease Prevention and Control. Available from www.ecdc.europa. eu/en/healthtopics/antimicrobial_resistance/ database/Pages/database.aspx.

29.

Hidron AI, Edwards JR, Patel J, et al. National Healthcare Safety Network Team; Participating National Healthcare Safety Network Facilities. NHSN annual update: antimicrobial-resistant pathogens associated with healthcare-associated infections: annual summary of data reported to the National Healthcare Safety Network at the Centers for Disease Control and Prevention, 2006-2007. Infect Control Hosp Epidemiol 2008;29(11):996-1011

Fontana R, Aldegheri M, Ligozzi M, et al. Overproduction of a low-affinity penicillinbinding protein and high-level ampicillin resistance in Enterococcus faecium. Antimicrob Agents Chemother 1994;38(9): 1980-3 Mainardi JL, Legrand R, Arthur M, et al. Novel mechanism of beta-lactam resistance due to bypass of D,D-transpeptidation in Enterococcus faecium. J Biol Chem 2000; 275(22):16490-6 Mainardi JL, Hugonnet JE, Rusconi F, et al. Unexpected inhibition of peptidoglycan L,D-transpeptidase from Enterococcus faecium by the beta-lactam imipenem. J Biol Chem 2007;282(42): 30414-22

20.

Zhang X, Paganelli FL, Bierschenk D, et al. Genome-wide identification of ampicillin resistance determinants in Enterococcus faecium. PLoS Genet 2012;8(6):e1002804

21.

Becker B, Cooper MA. Aminoglycoside antibiotics in the 21st century. ACS Chem Biol 2013;8(1):105-15

22.

Zembower TR, Noskin GA, Postelnick MJ, et al. The utility of aminoglycosides in an era of emerging drug resistance. Int J Antimicrob Agents 1998;10(2):95-105

www.expert-reviews.com

Galimand M, Schmitt E, Panvert M, et al. Intrinsic resistance to aminoglycosides in Enterococcus faecium is conferred by the 16S rRNA m5C1404-specific methyltransferase EfmM. RNA 2011;17(2):251-62

30.

Lebreton F, van Schaik W, McGuire AM, et al. Emergence of epidemic multidrug-resistant Enterococcus faecium from animal and commensal strains. MBio 2013;4(4):e00534-13



Based on genomic analysis, this study shows changes in the E. faecium genomes that paralleled changes in human behavior especially the urbanization of humans and the introduction of antibiotics in medicine and agriculture.

31.

Lester CH, Sandvang D, Olsen SS, et al. Emergence of ampicillin-resistant Enterococcus faecium in Danish hospitals. J Antimicrob Chemother 2008;62(6):1203-6

32.

Werner NL, Hecker MT, Sethi AK, Donskey CJ. Unnecessary use of fluoroquinolone antibiotics in hospitalized patients. BMC Infect Dis 2011;5(11):187

Review

33.

Galloway-Pen˜a JR, Rice LB, Murray BE. Analysis of. PBP5 of early U.S. isolates of Enterococcus faecium: sequence variation alone does not explain increasing ampicillin resistance over time. Antimicrob Agents Chemother 2011;55(7):3272-7

34.

Galloway-Pen˜a J, Roh JH, Latorre M, et al. Genomic and SNP analyses demonstrate a distant separation of the hospital and community-associated clades of Enterococcus faecium. PLoS ONE 2012;7(1):e30187

35.

Uttley AH, Collins CH, Naidoo J, George RC. Vancomycin-resistant enterococci. Lancet 1988;1(8575-6):57-8

36.

Leclercq R, Derlot E, Duval J, Courvalin P. Plasmid-mediated resistance to vancomycin and teicoplanin in Enterococcus faecium. N Engl J Med 1988;319(3):157-61

37.

Werner G, Coque TM, Hammerum AM, et al. Emergence and spread of vancomycin resistance among enterococci in Europe. Euro Surveill. 2008;13(47):pii : 19046

38.

Gilmore MS, Lebreton F, van Schaik W. Genomic transition of enterococci from gut commensals to leading causes of multidrug-resistant hospital infection in the antibiotic era. Curr Opin Microbiol 2013; 16(1):10-16

39.

Arias CA, Murray BE. Emergence and management of drug-resistant enterococcal infections. Expert Rev Anti Infect Ther 2008;6(5):637-55

40.

Gold HS. Vancomycin-resistant enterococci: mechanisms and clinical observations. Clin Infect Dis 2001;33(2):210-19

41.

Courvalin P. Vancomycin resistance in Gram-positive cocci. Clin Infect Dis 2006; 42(Suppl 1):S25-34

42.

Hegstad K, Mikalsen T, Coque TM, et al. Mobile genetic elements and their contribution to the emergence of antimicrobial resistant Enterococcus faecalis and Enterococcus faecium. Clin Microbiol Infect 2010;16(6):541-54

43.

Watanabe S, Kobayashi N, Quin˜ones D, et al. Genetic diversity of the low-level vancomycin resistance gene vanC-2/vanC3 and identification of a novel vanC subtype (vanC-4) in Enterococcus casseliflavus. Microb Drug Resist 2009; 15(1):1-9

44.

Lebreton F, Depardieu F, Bourdon N, et al. D-Ala-D-Ser VanN-type transferable vancomycin resistance in Enterococcus faecium. Antimicrob Agents Chemother 2011;55(10):4606-12

45.

Nomura T, Tanimoto K, Shibayama K, et al. Identification of VanN-type vancomycin resistance in an Enterococcus

245

Review

Cattoir & Giard

faecium isolate from chicken meat in Japan. Antimicrob Agents Chemother 2012;56(12): 6389-92 46.

Expert Review of Anti-infective Therapy Downloaded from informahealthcare.com by CDL-UC San Diego on 01/01/15 For personal use only.

47.

48.

49.

50.

Cremniter J, Mainardi JL, Josseaume N, et al. Novel mechanism of resistance to glycopeptide antibiotics in Enterococcus faecium. J Biol Chem 2006;281(43): 32254-62

Leavis HL, Willems RJ, Top J, Bonten MJ. High-level ciprofloxacin resistance from point mutations in gyrA and parC confined to global hospital-adapted clonal lineage CC17 of Enterococcus faecium. J Clin Microbiol 2006;44(3):1059-64 Werner G, Fleige C, Ewert B, et al. High-level ciprofloxacin resistance among hospital-adapted Enterococcus faecium (CC17). Int J Antimicrob Agents 2010; 35(2):119-25 Oyamada Y, Ito H, Fujimoto K, et al. Combination of known and unknown mechanisms confers high-level resistance to fluoroquinolones in Enterococcus faecium. J Med Microbiol 2006;55(Pt 6):729-36 Canu A, Leclercq R. Overcoming bacterial resistance by dual target inhibition: the case of streptogramins. Curr Drug Targets Infect Disord 2001;1(2):215-25

52.

Cocito C, Di Giambattista M, Nyssen E, Vannuffel P. Inhibition of protein synthesis by streptogramins and related antibiotics. J Antimicrob Chemother 1997;39(Suppl A): 7-13

53.

Poehlsgaard J, Douthwaite S. The bacterial ribosome as a target for antibiotics. Nat Rev Microbiol 2005;3(11):870-81

55.

56.

58.

Hooper DC. Fluoroquinolone resistance among Gram-positive cocci. Lancet Infect Dis 2002;2(9):530-8

51.

54.

57.

Johnston NJ, Mukhtar TA, Wright GD. Streptogramin antibiotics: mode of action and resistance. Curr Drug Targets 2002; 3(4):335-44 De Graef EM, Decostere A, De Leener E, et al. Prevalence and mechanism of resistance against macrolides, lincosamides, and streptogramins among Enterococcus faecium isolates from food-producing animals and hospital patients in Belgium. Microb Drug Resist 2007;13(2):135-41 Lo´pez F, Culebras E, Betriu´ C, et al. Antimicrobial susceptibility and macrolide resistance genes in Enterococcus faecium with reduced susceptibility to quinupristindalfopristin: level of quinupristin-dalfopristin resistance is not dependent on erm(B) attenuator region sequence. Diagn Microbiol Infect Dis 2010; 66(1):73-7

246

Isnard C, Malbruny B, Leclercq R, Cattoir V. Genetic basis for in vitro and in vivo resistance to lincosamides, streptogramins A, and pleuromutilins (LSAP phenotype) in Enterococcus faecium. Antimicrob Agents Chemother 2013;57(9): 4463-9 Jackson CR, Fedorka-Cray PJ, Barrett JB, et al. Prevalence of streptogramin resistance in enterococci from animals: identification of vatD from animal sources in the USA. Int J Antimicrob Agents 2007;30(1):60-6

59.

Hancock RE. Mechanisms of action of newer antibiotics for Gram-positive pathogens. Lancet Infect Dis 2005;5(4): 209-18

60.

Livermore DM. Linezolid in vitro: mechanism and antibacterial spectrum. J Antimicrob Chemother 2003;51(Suppl 2): ii9-16

61.

Bozdogan B, Appelbaum PC. Oxazolidinones: activity, mode of action, and mechanism of resistance. Int J Antimicrob Agents 2004;23(2):113-19

62.

Denys GA, Koch KM, Dowzicky MJ. Distribution of resistant Gram-positive organisms across the census regions of the United States and in vitro activity of tigecycline, a new glycylcycline antimicrobial. Am J Infect Control 2007; 35(8):521-6

63.

Norskov-Lauritsen N, Marchandin H, Dowzicky MJ. Antimicrobial susceptibility of tigecycline and comparators against bacterial isolates collected as part of the TEST study in Europe. (2004-2007). Int J Antimicrob Agents 2009;34(2):121-30

64.

Jones RN, Ross JE, Castanheira M, Mendes RE. United States resistance surveillance results for linezolid (LEADER Program for. 2007). Diagn Microbiol Infect Dis 2008;62(4):416-26

65.

Berenger R, Bourdon N, Auzou M, et al. In vitro activity of new antimicrobial agents against glycopeptide-resistant Enterococcus faecium clinical isolates from France between. 2006 and 2008. Med Mal Infect 2011;41(8):405-9

66.

Meka VG, Gold HS. Antimicrobial resistance to linezolid. Clin Infect Dis 2004; 39(7):1010-15

67.

Gonzales RD, Schreckenberger PC, Graham MB, et al. Infections due to vancomycin-resistant Enterococcus faecium resistant to linezolid. Lancet 2001; 357(9263):1179

68.

Seedat J, Zick G, Klare I, et al. Rapid emergence of resistance to linezolid during linezolid therapy of an Enterococcus faecium

infection. Antimicrob Agents Chemother 2006;50(12):4217-19 69.

Herrero IA, Issa NC, Patel R. Nosocomial spread of linezolid-resistant, vancomycin-resistant Enterococcus faecium. N Engl J Med 2002;346(11):867-9



Key paper describing the first outbreak of linezolid-resistant and vancomycinresistant Enterococcus faecium.

70.

Johnson AP, Tysall L, Stockdale MV, et al. Emerging linezolid-resistant Enterococcus faecalis and Enterococcus faecium isolated from two Austrian patients in the same intensive care unit. Eur J Clin Microbiol Infect Dis 2002;21(10):751-4

71.

Rahim S, Pillai SK, Gold HS, et al. Linezolid-resistant, vancomycin-resistant Enterococcus faecium infection in patients without prior exposure to linezolid. Clin Infect Dis 2003;36(11):E146-8

72.

Werner G, Bartel M, Wellinghausen N, et al. Detection of mutations conferring resistance to linezolid in Enterococcus spp. by fluorescence in situ hybridization. J Clin Microbiol 2007;45(10):3421-3

73.

Schnitzler P, Schulz K, Lampson C, et al. Molecular analysis of linezolid resistance in clinical Enterococcus faecium isolates by polymerase chain reaction and pyrosequencing. Eur J Clin Microbiol Infect Dis 2011;30(1):121-5

74.

Marshall SH, Donskey CJ, Hutton-Thomas R, et al. Gene dosage and linezolid resistance in Enterococcus faecium and Enterococcus faecalis. Antimicrob Agents Chemother 2002;46(10):3334-6

75.

Prystowsky J, Siddiqui F, Chosay J, et al. Resistance to linezolid: characterization of mutations in rRNA and comparison of their occurrences in vancomycin-resistant enterococci. Antimicrob Agents Chemother 2001;45(7):2154-6

76.

Auckland C, Teare L, Cooke F, et al. Linezolid-resistant enterococci: report of the first isolates in the United Kingdom. J Antimicrob Chemother 2002;50(5):743-6

77.

Meka VG, Pillai SK, Sakoulas G, et al. Linezolid resistance in sequential Staphylococcus aureus isolates associated with a T2500A mutation in the 23S rRNA gene and loss of a single copy of rRNA. J Infect Dis 2004;190(2):311-17

78.

Livermore DM, Warner M, Mushtaq S, et al. In vitro activity of the oxazolidinone RWJ-416457 against linezolid-resistant and -susceptible staphylococci and enterococci. Antimicrob Agents Chemother 2007;51(3): 1112-14

Expert Rev. Anti Infect. Ther. 12(2), (2014)

Antibiotic resistance in Enterococcus faecium

79.

Expert Review of Anti-infective Therapy Downloaded from informahealthcare.com by CDL-UC San Diego on 01/01/15 For personal use only.

80.

81.

82.

83.

84.



85.

86.

87.

88.

Livermore DM, Mushtaq S, Warner M, Woodford N. Activity of oxazolidinone TR-700 against linezolid-susceptible and resistant staphylococci and enterococci. J Antimicrob Chemother 2009;63(4):713-15 Locke JB, Hilgers M, Shaw KJ. Novel ribosomal mutations in Staphylococcus aureus strains identified through selection with the oxazolidinones linezolid and torezolid (TR700). Antimicrob Agents Chemother 2009; 53(12):5265-74

89.

90.

Locke JB, Hilgers M, Shaw KJ. Mutations in ribosomal protein. L3 are associated with oxazolidinone resistance in staphylococci of clinical origin. Antimicrob Agents Chemother 2009;53(12):5275-8

91.

Endimiani A, Blackford M, Dasenbrook EC, et al. Emergence of linezolid-resistant Staphylococcus aureus after prolonged treatment of cystic fibrosis patients in Cleveland. Ohio. Antimicrob Agents Chemother 2011;55(4):1684-92

92.

Kehrenberg C, Schwarz S, Jacobsen L, et al. A new mechanism for chloramphenicol, florfenicol and clindamycin resistance: methylation of 23S ribosomal RNA at A2503. Mol Microbiol 2005;57(4):1064-73 Schwarz S, Werckenthin C, Kehrenberg C. Identification of a plasmid-borne chloramphenicol-florfenicol resistance gene in Staphylococcus sciuri. Antimicrob Agents Chemother 2000;44(9):2530-3 First description of the plasmid-mediated cfr gene from a Staphylococcus sciuri isolate of animal origin. Toh SM, Xiong L, Arias CA, et al. Acquisition of a natural resistance gene renders a clinical strain of methicillin-resistant Staphylococcus aureus resistant to the synthetic antibiotic linezolid. Mol Microbiol 2007;64(6):1506-14 Diaz L, Kiratisin P, Mendes RE, et al. Transferable plasmid-mediated resistance to linezolid due to cfr in a human clinical isolate of Enterococcus faecalis. Antimicrob Agents Chemother 2012;56(7):3917-22 Long KS, Poehlsgaard J, Kehrenberg C, et al. The Cfr rRNA methyltransferase confers resistance to Phenicols, Lincosamides, Oxazolidinones, Pleuromutilins, and Streptogramin A antibiotics. Antimicrob Agents Chemother 2006;50(7):2500-5 Fines M, Leclercq R. Activity of linezolid against Gram-positive cocci possessing genes conferring resistance to protein synthesis inhibitors. J Antimicrob Chemother 2000; 45(6):797-802

www.expert-reviews.com

93.

94.

95.

Review

treatment of infections caused by vancomycin-resistant enterococci.

Sader HS, Farrell DJ, Jones RN. Antimicrobial activity of daptomycin tested against Gram-positive strains collected in European hospitals: results from 7 years of resistance surveillance. (2003-2009). J Chemother 2011;23(4):200-6

100.

Kanafani ZA, Corey GR. Daptomycin: a rapidly bactericidal lipopeptide for the treatment of Gram-positive infections. Expert Rev Anti Infect Ther 2007;5(2): 177-84

Munita JM, Panesso D, Diaz L, et al. Correlation between mutations in liaFSR of Enterococcus faecium and MIC of daptomycin: revisiting daptomycin breakpoints. Antimicrob Agents Chemother 2012;56(8):4354-9

101.

Mascher T, Zimmer SL, Smith TA, Helmann JD. Antibiotic-inducible promoter regulated by the cell envelope stress-sensing two-component system LiaRS of Bacillus subtilis. Antimicrob Agents Chemother 2004;48(8):2888-96

102.

Wolf D, Kalamorz F, Wecke T, et al. In-depth profiling of the LiaR response of Bacillus subtilis. J Bacteriol 2010;192(18): 4680-93

103.

Humphries RM, Kelesidis T, Tewhey R, et al. Genotypic and phenotypic evaluation of the evolution of high-level daptomycin nonsusceptibility in vancomycin-resistant Enterococcus faecium. Antimicrob Agents Chemother 2012;56(11):6051-3

104.

Tran TT, Panesso D, Gao H, et al. Whole-genome analysis of a daptomycin-susceptible Enterococcus faecium strain and its daptomycin-resistant variant arising during therapy. Antimicrob Agents Chemother 2013;57(1):261-8

105.

Mishra NN, Bayer AS, Tran TT, et al. Daptomycin resistance in enterococci is associated with distinct alterations of cell membrane phospholipid content. PLoS ONE 2012;7(8):e43958

106.

Arias CA, Torres HA, Singh KV, et al. Failure of daptomycin monotherapy for endocarditis caused by an Enterococcus faecium strain with vancomycin-resistant and vancomycin-susceptible subpopulations and evidence of in vivo loss of the vanA gene cluster. Clin Infect Dis 2007;45(10):1343-6

107.

Schwartz BS, Ngo PD, Guglielmo BJ. Daptomycin treatment failure for vancomycin-resistant Enterococcus faecium infective endocarditis: impact of protein binding? Ann Pharmacother 2008;42(2): 289-90

108.

Noskin GA. Tigecycline: a new glycylcycline for treatment of serious infections. Clin Infect Dis 2005;41(Suppl 5):S303-14

109.

Borbone S, Lupo A, Mezzatesta ML, et al. Evaluation of the in vitro activity of tigecycline against multiresistant Gram-positive cocci containing tetracycline resistance determinants. Int J Antimicrob Agents 2008;31(3):209-15

Mave V, Garcia-Diaz J, Islam T, Hasbun R. Vancomycin-resistant enterococcal bacteraemia: is daptomycin as effective as linezolid? J Antimicrob Chemother 2009; 64(1):175-80 Crank CW, Scheetz MH, Brielmaier B, et al. Comparison of outcomes from daptomycin or linezolid treatment for vancomycin-resistant enterococcal bloodstream infection: a retrospective, multicenter, cohort study. Clin Ther 2010; 32(10):1713-19 Canton R, Ruiz-Garbajosa P, Chaves RL, Johnson AP. A potential role for daptomycin in enterococcal infections: what is the evidence? J Antimicrob Chemother 2010;65(6):1126-36 Twilla JD, Finch CK, Usery JB, et al. Vancomycin-resistant Enterococcus bacteremia: an evaluation of treatment with linezolid or daptomycin. J Hosp Med 2012; 7(3):243-8 Silverman JA, Oliver N, Andrew T, Li T. Resistance studies with daptomycin. Antimicrob Agents Chemother 2001;45(6): 1799-802

96.

Lewis JS II, Owens A, Cadena J, et al. Emergence of daptomycin resistance in Enterococcus faecium during daptomycin therapy. Antimicrob Agents Chemother 2005;49(4):1664-5

97.

Montero CI, Stock F, Murray PR. Mechanisms of resistance to daptomycin in Enterococcus faecium. Antimicrob Agents Chemother 2008;52(3):1167-70

98.

Kelesidis T, Tewhey R, Humphries RM. Evolution of high-level daptomycin resistance in Enterococcus faecium during daptomycin therapy is associated with limited mutations in the bacterial genome. J Antimicrob Chemother 2013;68(8):1926-8

99.

Arias CA, Panesso D, McGrath DM, et al. Genetic basis for in vivo daptomycin resistance in enterococci. N Engl J Med 2011;365(10):892-900

••

Key paper demonstrating mutations associated with the development of resistance to daptomycin during the

247

Review 110.

Rubinstein E, Vaughan D. Tigecycline: a novel glycylcycline. Drugs 2005;65(10): 1317-36

111.

Aznar J, Lepe JA, Dowzicky MJ. Antimicrobial susceptibility among E. faecalis and E. faecium from France, Germany, Italy, Spain and the UK (T.E.S. T. Surveillance Study. 2004-2009). J Chemother 2012;24(2):74-80

112.

Expert Review of Anti-infective Therapy Downloaded from informahealthcare.com by CDL-UC San Diego on 01/01/15 For personal use only.

Cattoir & Giard

Shen J, Wang Y, Schwarz S. Presence and dissemination of the multiresistance gene cfr in Gram-positive and Gram-negative bacteria. J Antimicrob Chemother 2013; 68(8):1697-706

248

113.

Zhanel GG, Sniezek G, Schweizer F, et al. Ceftaroline: a novel broad-spectrum cephalosporin with activity against methicillin-resistant Staphylococcus aureus. Drugs 2009;69(7):809-31

116.

Roux D, Pier GB, Skurnik D. Magic bullets for the 21st century: the reemergence of immunotherapy for multi- and pan-resistant microbes. J Antimicrob Chemother 2012; 67(12):2785-7

114.

Arias CA, Contreras GA, Murray BE. Management of multidrug-resistant enterococcal infections. Clin Microbiol Infect 2010;16(6):555-62

117.

Deresinski S. Bacteriophage therapy: exploiting smaller fleas. Clin Infect Dis 2009;48(8):1096-101

115.

Zhanel GG, Lawson CD, Zelenitsky S, et al. Comparison of the next-generation aminoglycoside plazomicin to gentamicin, tobramycin and amikacin. Expert Rev Anti Infect Ther 2012;10(4):459-73

Expert Rev. Anti Infect. Ther. 12(2), (2014)

Antibiotic resistance in Enterococcus faecium clinical isolates.

The worldwide ratio of Enterococcus faecalis-Enterococcus faecium infections is currently changing in favor of E. faecium. Intrinsic and acquired anti...
345KB Sizes 0 Downloads 0 Views