Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2017 Nov 1.
Published in final edited form as: Drug Resist Updat. 2016 Sep 19;29:30–46. doi: 10.1016/j.drup.2016.09.002

The rapid spread of carbapenem-resistant Enterobacteriaceae

Robert F Potter 1,*, Alaric W D’Souza 1,*, Gautam Dantas 1,2,3,4,
PMCID: PMC5140036  NIHMSID: NIHMS817718  PMID: 27912842

Abstract

Carbapenems, our one-time silver bullet for multidrug resistant bacterial infections, are now threatened by widespread dissemination of carbapenem-resistant Enterobacteriaceae (CRE). Successful expansion of Enterobacteriaceae clonal groups and frequent horizontal gene transfer of carbapenemase expressing plasmids are causing increasing carbapenem resistance. Recent advances in genetic and phenotypic detection facilitate global surveillance of CRE diversity and prevalence. In particular, whole genome sequencing enabled efficient tracking, annotation, and study of genetic elements colocalized with carbapenemase genes on chromosomes and on plasmids. Improved characterization helps detail the co-occurrence of other antibiotic resistance genes in CRE isolates and helps identify pan-drug resistance mechanisms. The novel β-lactamase inhibitor, avibactam, combined with ceftazidime or aztreonam, is a promising CRE treatment compared to current colistin or tigecycline regimens. To halt increasing CRE-associated morbidity and mortality, we must continue quality, cooperative monitoring and urgently investigate novel treatments.

Introduction

Carbapenems are the favored last resort drugs for treatment of multidrug resistant (MDR) bacterial infections. Accordingly, spreading carbapenem resistance is a global health crisis. Carbapenems are β-lactam antibiotics that bind to penicillin binding proteins and inhibit cell wall synthesis. Meropenem, doripenem, ertapenem, and imipenem are the four most clinically used carbapenems [Nicolau 2008].

Enterobacteriaceae are a family of diverse Gammaproteobacteria which include common (Klebsiella pneumoniae, Escherichia coli, Salmonella enterica) and rare (Proteus mirabilis, Raoultella planticola, Citrobacter freundii) human pathogens with increasing antibiotic resistance. The Clinical Laboratory Standards Institute defines Enterobacteriaceae as carbapenem-resistant if they have minimum inhibitory concentrations (MICs) of ≥2 μg/ml against ertapenem and ≥4 μg/ml against doripenem, meropenem, or imipenem [Patel 2015]. In a 2013 US Centers for Disease Control and Prevention (CDC) report, carbapenem-resistant Enterobacteriaceae (CRE) were listed as one of the three most urgent antimicrobial resistant threats. CREs received this highest threat level due to rapidly increasing global spread, propensity for multidrug resistance, and high mortality during blood stream infections (BSI) [Bell 2013].

Carbapenem resistance can arise from β-lactam ring hydrolysis by dedicated carbapenemase enzymes or from changes in membrane permeability via mutations in efflux pumps or porins coupled with extended spectrum β-lactamase (ESBL) expression. Carbapenemases come from ambler class A or D serine β-lactamases and ambler class B metallo-β-lactamases (MBLs). K. pneumoniae carbapenemase (KPC-Class A), New Delhi metallo-β-lactamase (NDM-Class B), Verona integron-encoded metallo-β-lactamase (VIM-Class B), Imipenemase metallo-β-lactamase (IMP-Class B), and Oxacillinase-48 (OXA-48-Class D) variants are the most common carbapenemases in carbapenemase producing Enterobacteriaceae (CPE) and thus our focus in this review, but several emerging Class A enzymes are also described [Meletis 2016]. Carbapenemase production is concerning given high occurrence of carbapenemase genes on MDR plasmids transferred both within Enterobacteriaceae and to other bacterial families [Harmer and Hall 2015].

This review highlights recent work documenting CRE prevalence, detecting CRE, and describing MDR patterns to optimize therapeutic strategy.

Class A β-lactamases

The first class A carbapenemase was reported in 1990 in a Serratia marcescens isolate from the United Kingdom [Yang 1990]. Many class A carbapenemases were identified after the discovery of S. marcescens enzyme one (blaSME-1), including Guiana extended-spectrum two (blaGES-2), K. pneumoniae carbapenemase (blaKPC), imipenemase/non-metallocarbapenemase-A (blaIMI/NMC), Serratia fonticola carbapenemase one (blaSFC-1), and blaSHV-38 [Henriques 2004, Poirel 2003, Poirel 2001, Rasmussen 1996, Yigit 2001]. Many blaGES variants are only ESBLs but frequent point mutations confer increased hydrolytic properties against carbapenems for specific variants [Dang 2016]. Phylogenetic alignment of class A carbapenemase genes determined that blaSME, blaGES, blaKPC, and blaIMI/NMC constitute a separate group from blaSFC-1 and blaSHV-38 [Walther-Rasmussen and Hoiby 2007]. Similar to other Class A β-lactamases, class A carbapenemases are inhibited by clavulanic acid and tazobactam [Bush and Jacoby 2010]. While KPC is the most common Class A carbapenemase in Enterobacteriaceae, sporadic cases of previously discovered enzymes such as SME and NMC and new gene discoveries necessitate molecular characterization to identify potentially pandemic carbapenemases before they spread [Nordmann 2011, Poirel 2007].

KPC

As reported in 2001, blaKPC was discovered in carbapenem resistant K. pneumoniae from the United States and it is largely responsible for conferring carbapenem resistance in CRE infections in the western hemisphere [Gaiarsa 2015, Yigit 2001]. Unlike the previously characterized Class A carbapenemases (blaNMC and blaSME), blaKPC is plasmid borne which facilitated its rapid appearance in non-K. pneumoniae Enterobacteriaceae [Hossain 2004, Miriagou 2003, Yigit 2003]. KPC’s efficient carbapenem hydrolysis confers high levels of carbapenem resistance in the absence of complementary membrane permeability changes or ESBL expression, setting it apart from other serine β-lactamases [Poirel 2007]. In the UCLA Hospital system from January 2011 to January 2013, a few (0.73%) Enterobacteriaceae isolates were CRE, but an overwhelming proportion of those isolates (78.3%) had KPC associated carbapenem resistance [Pollett 2014].

Regardless of location, 97% of whole-genome sequenced, KPC-producing K. pneumoniae belong to the clonal complex 258, demonstrating the global dissemination of member sequence types [Gaiarsa 2015]. Analysis of ST258 isolates identified two clades due to divergence in capsular polysaccharide synthesis (cps) genes [Deleo 2014]. Association of cps-1 with blaKPC-2 and cps-2 with blaKPC-3 suggests that variation in KPC is due to point mutations following cps recombination, rather than horizontal gene transfer via Tn4401 [Bowers 2015]. However, a two year-long study of admitted patients in five Colombian hospitals found a majority (62%, 120/193) of carbapenem-resistant K. pneumoniae isolates were non-258 group members, but overall, KPC was the predominant carbapenemase [Ocampo 2016]. ST258 was identified for the first time in a collection of CPE Isolates from Taiwan. Although KPC was the most common (85.8%, 157/183) carbapenemase found, it was largely associated with K. pneumoniae ST11 [Tseng 2015]. Since ST258 is a hybrid strain of ~80% ST11, there is potential for rapid proliferation of the clone throughout the region [Chen 2014]. Overall percentage of CPE dropped in Israeli post-acute-care hospitals from 2008 (16%, 184/1147) to 2013 (9.9%, 127/1287), but ST258 increased compared to all KPC producing K. pneumoniae (65%, 120/184 in 2008 to 80%, 91/113 in 2013) [Adler 2015].

WGS identified a novel KPC encoding mosaic plasmid, pKp28, associated with an outbreak from an endoscope at a single hospital in the United States from January 2011 to July 2013. blaKPC-2 was found on a Tn4401 transposon with 99.99% nucleotide identity to pKpQIL, while a downstream region had 99.99% identity to pHg, another plasmid associated with ST258 [Marsh 2015]. blaKPC-2 was also discovered in the canonical hypervirulent K. pneumoniae ST23 [Cejas 2014]. Focused genomic analyses demonstrated that CG258 isolates cluster separately from ST23, and broader results confirm that multidrug resistant isolates have little overlap with hyper-virulent ones. However, this does not preclude the possibility of plasmid exchange leading to hypervirulent and multidrug resistant K. pneumoniae expressing KPC [Bialek-Davenet 2014, Struve 2015]. blaKPC-2 was also identified in K. pneumoniae ST1797, a new hypervirulent strain associated with surgical wound infection in China in 2013 [Zhang 2016]. Successful transfer of blaKPC containing plasmids throughout established nosocomial Enterobacteriaceae strains led to outbreaks in the United States in 2002 to 2003, and in Nepal in 2012 [Bratu 2005, Chung The 2015]. A comparative study of hospitals in the United States and Pakistan from February 2012 to March 2013 showed that carriage of both blaKPC and blaNDM in the same isolate is rare, but similarity in genetic background between both locations make the possibility of horizontal gene transfer likely [Pesesky 2015].

Active surveillance methods should be undertaken to quickly identify KPC producers and reduce infection prevalence.

Sequence-divergent Class A Carbapenemases

A novel Class A carbapenemase with 63% sequence identity to an environmental β-lactamase and encoding broad activity against the penicillins, cephalosporins, carbapenems, and monobactams was discovered in three carbapenem-resistant K. pneumoniae ST1781 isolates from Brazil in 2008 [Nicoletti 2015]. The Brazilian Klebsiella Carbapenemase-1 (blaBKC-1) gene was found on a non-conjugative IncQ plasmid with aph3A-VI, a novel variant of the aminoglycoside resistance gene aph3 [Nicoletti 2015]. In 2015 another novel Class A carbapenemase was identified in France from a carbapenem-resistant E. cloacae that produced a tazobactam-inhibitable carbapenemase negative for blaKPC, blaGES, blaSFC-1, blaIMI, and blaNMC-A. The identified French imipenemase-1 (FRI-1) has significant activity against carbapenems and aztreonam, but not broad-spectrum cephalosporins; the low GC content (39%) of blaFRI-1 compared to the E. cloacae genome (55%) suggests acquisition via horizontal gene transfer [Dortet 2015]. Surveillance of CPE carriage at a hospital in Italy identified an Enterobacter ludwigii in August 2014 with a chromosomally integrated blaNMC-A complexed by a novel Xer-dependent integrative mobile element. Examination of publically available whole genome sequencing (WGS) data found that the element was associated with blaIMI and blaNMC-A in other members of the E. cloacae clonal complex, indicating the possibility of dissemination throughout Enterobacteriaceae [Antonelli 2015].

Class B β-lactamases

MBLs can be differentiated from Class A and D β-lactamases by their use of zinc instead of a catalytic serine in the active site mediated hydrolysis of β-lactams [Pitout 2015]. Accordingly, MBLs are inhibited by metal chelators such as Ethylenediaminetetraacetic acid (EDTA) and dipicolinic acid (DPA) but not clavulanic acid or other clinically used β-lactamase inhibitors [Pitout 2015]. They often have hydrolysis profiles against all β-lactams except monobactams and can confer high level of resistance when combined with changes in membrane permeability and ESBL production [Pitout 2015].

NDM

NDM is endemic to parts of Asia and is responsible for sporadic outbreaks around the globe [Zmarlicka 2015]. As of August 2016 when we accessed the NCBI database, 16 variants of NDM were identified on a variety of plasmid types, corresponding to the high diversity of Enterobacteriaceae species shown to express it [NCBI 2016, Zmarlicka 2015]. A comprehensive study from 2012 to 2014 of 38,266 Enterobacteriaceae isolates collected from 40 countries spanning every inhabited continent found that global incidence of MBL presence is low (0.5%, 163/38,266), but dissemination is high (85%, 34/40 countries). blaNDM-1 was the most common gene (36.8%, 60/163) in Enterobacteriaceae [Kazmierczak 2015]. The Study for Monitoring Antimicrobial Resistance Trends identified blaNDM-1 (96.3%, 130/135) as the most common variant in NDM associated infections from 2008 to 2012 across geographically diverse countries. Population migration will undoubtedly contribute to future global spread of CRE as evidenced by the prevalence of blaNDM (59.3%, 19/32) among K. pneumoniae and E. coli isolates from Syrian patients displaced by civil strife in Europe [Lerner 2016]. Co-presence of ESBLs and NDM in the same isolate was high (78.5%, 106/135), but dual carbapenemase prevalence was fortunately low (1/135 with blaVIM-1 + blaNDM, 2/135 with blaOXA-181 + blaNDM) [Biedenbach 2015]. A nosocomial outbreak declared in March 2014 of 19 K. pneumoniae ST10 in the neonatal intensive care unit of a Chinese hospital found that all isolates harbored blaNDM-1 on conjugative IncFI plasmids and that co-production of blaIMP-4 was common (36.8%, 7/19) but fortunately so was tigecycline, amikacin, and ciprofloxacin susceptibility [Zheng 2016]. K. pneumoniae was overwhelmingly responsible for NDM production in CRE colonized patients (370/374 patients) from Poland, a non-endemic area, during a 2012–2014 outbreak [Baraniak 2016]. ST11 was the dominant sequence type and blaNDM-1 was the most common variant, while a mix of IncR, IncFII, and undetermined plasmids were responsible for spread [Baraniak 2016]. Tn3000, a novel transposon, was responsible for proliferation of blaNDM-1 among IncX3 and IncFIIk plasmids extracted from E. coli and Enterobacter hormaechei, respectively, in 2013 [Campos 2015]. Whole plasmid sequencing revealed high levels of intrapatient diversity for blaNDM-1 plasmid types and host species for Enterobacteriaceae and Acinetobacter baumannii isolates obtained in Pakistan during 2010 [Wailan 2015].

Recently, blaNDM-13, a chromosomally integrated NDM variant discovered in E. coli ST101 from Nepal in 2013, was found with a spectrum of action similar to blaNDM-1 but with increased cefotaxime affinity due to D95N and M154L mutations [Shrestha 2015]. Carbapenemase production in Salmonella enterica is rare but documented for KPC and NDM [Miriagou 2003, Savard 2011]. blaNDM-1 was found on a conjugative IncA/C plasmid with blaCMY (Class C plasmid-mediated AmpC beta-lactamase) in an extensively drug resistant S. enterica serovar Senftenberg isolate obtained from a pediatric outpatient in India in 2012. The plasmid had multiple mobile gene elements, complicating precise determination of the genetic events that led to blaNDM-1 acquisition [Sarkar 2015]. A study published in August 2014 identified two blaNDM-1 positive S. enterica serovar Agona isolates obtained from pediatric gastroenteritis infections in Pakistan [Irfan 2015]. These examples demonstrate need for carbapenemase gene surveillance in S. enterica serovars, especially for MBL endemic regions.

IMP

IMP, the first identified acquired MBL, was found in a S. marcescens isolate from Japan in 1991 [Ito 1995]. They are most common in China, Japan, and Australia [Nordmann 2011]. Contrasting global trends, blaIMP-4 was the most common carbapenemase (82.7%, 48/58) found in July 2009 to March 2014 from hospitals in Australia. 10 Enterobacteriaceae species produced blaIMP-4, but E. cloacae was the most frequent (60.4%, 29/48) and possessed blaIMP-4 on conjugatable HI2 (22/29 or L/M (7/29) plasmids demonstrating a plausible mechanism for blaIMP-4 proliferation throughout the region [Sidjabat 2015]. A pediatric patient in China in June 2014 had seven closely related blaIMP-4 producing Raoultella ornithinolytica isolated from an infected surgical wound, including the first observation of an R. ornithinolytica in China producing blaIMP-4 and blaKPC-2 [Zheng 2015]. Analysis of mobile genetic elements surrounding MBLs in Enterobacteriaceae collected in Spain from February to July 2009 found that all 14 blaVIM-1 genes were associated with class 1 integrons but a mixture of genetic backgrounds (2/14 blaVIM-1, IncU and 12/14 unidentified plasmid type) [Zamorano 2015].

VIM

VIM is largely found in Italy and Greece [Pitout 2015]. VIM-1 has the closest amino acid sequence identity (32.4%) to NDM-1 [Pitout 2015]. A multinational European survey of MBL-producing Enterobacteriaceae isolates collected between 2008 and 2011 found universal expression of blaVIM-1 type members (98.9%, 93/94) in nine different Enterobacteriaceae species but disproportionately high representation of K. pneumoniae ST147, ST36, and ST383 (60.6%, 57/94) from Greece [Papagiannitsis 2015]. VIM-39 is a novel blaVIM-1 type member identified in the previous study; it was found to hydrolyze meropenem, doripenem, and imipenem more efficiently than VIM-1 and conferred greater resistance to those carbapenems than blaVIM-1 in E. coli DH5α [Papagiannitsis 2015]. A pediatric patient from the United States was colonized by a MBL producing K. pneumoniae and later infected by a clone of the isolate in May 2014. The isolate transferred an IncA/C replicon 188-kb plasmid containing blaVIM-4 and blaCMY-4 to a recipient E. coli strain [Tamma 2016]. Additionally, eight VIM-producing CRE isolates were found on six patients in the same tertiary care hospital in the United States during August 2015. Fortunately, VIM colonization was not linked with infection, but the diversity of Enterobacteriaceae species (one Raoultella spp, four E. cloacae, one E. coli, and two K. pneumoniae) and patient location (6 neonatal intensive care unit, 2 adult intensive care unit) combined with the unidentified VIM-producing source warrant further investigation of blaVIM containing plasmids [Yaffee 2016].

Class D β-lactamases

Class D β-lactamases were named oxacillinases (OXA) because they cleave oxacillin in addition to penicillin, distinguishing them from class A β-lactamases. Although first characterized as extended spectrum β-lactamases, OXA-23 was found to confer resistance against imipenem in an A. baumannii isolate from the United Kingdom in 1985 [Donald 2000]. Additional carbapenem-hydrolyzing class D β-lactamases (CHDLs) have also been discovered. Since 2001 OXA-23-like, OXA-40-like, OXA-51-like, OXA-58-like, and OXA-48-like family members have been found in the Enterobacteriaceae [Evans and Amyes 2014]. CHDLs are increasing in global frequency more than the Class A and B carbapenemases [Bakthavatchalam 2016]. blaOXA-48 was identified in a K. pneumoniae isolate from Turkey in 2001 and remains the most common CHDL detected [Poirel 2004]. A chromosomally encoded CHDL, blaOXA-54, from the environmental bacteria Shewanella oneidensis had a 92% amino acid similarity with blaOXA-48 and comparable catalytic activity against imipenem [Poirel 2004]. Analysis of the flanking sequences between blaOXA-48 and blaOXA-54 suggest that blaOXA-48 originated in Shewanella spp [Poirel 2004]. blaOXA-48-like family members cleave carbapenems weakly, which can lead to phenotypic susceptibility against carbapenems and therefore complicate treatment options [Bakthavatchalam 2016]. Excepting OXA-163, OXA-48-like family members have little to no activity against Ceftazidine and Cefotaxime but phenotypic resistance to cephalosporins can occur in isolates co-expressing CHDLs and cephalosporin cleaving β-lactamases [Oueslati 2015].

OXA-48-like family members

A survey of CRE in Moscow from January 2013 to October 2014 found examples of blaOXA-48 spreading from K. pneumoniae to Proteus mirabilis, with both pathogens having nearly universal resistance (99% and 100% respectively) to three or more functional drug classes [Fursova 2015]. The majority (71.9%, 87/121) of carbapenem-resistant, carbapenemase producing E. coli obtained from 2012 to 2014 in Spain possess blaOXA-48 but in a less diverse population compared to blaVIM-1 expressing isolates [Ortega 2016]. Conjugatable blaOXA-48 (9/11) was the most prevalent carbapenemase detected in CPE from a hospital in Morocco collected in June to August 2011 [Barguigua 2015]. The first documented instance of a community acquired UTI due to an blaOXA-48-like producing Enterobacteriaceae was in Greece in January 2010 when a recovered K. pneumoniae ST11 isolate was found with carbapenem resistance due to blaOXA-162 [Voulgari 2015]. Multiple studies identified blaOXA-48 as the most prevalent carbapenemase in Spain, and this is associated with successful spread of K. pneumoniae, particularly ST11 [Branas 2015, Oteo 2015]. OXA-48 in the absence of other ESBLs is difficult to detect using traditional phenotypic methods due to low level of resistance against cephalosporins or carbapenems. Molecular characterization of blaOXA-48-like variants is recommended, as expression of blaOXA-48 in E. coli does not always yield carbapenem resistance in vitro, but it can prevent clearance of infections [Haverkate 2015]. Emerging cases of blaOXA-48 in Taiwan from January 2012 to May 2014 suggest expanding ability to confer carbapenem resistance in the Enterobacteriaceae [Ma 2015]. blaOXA-48 was responsible for carbapenem resistance in a collection of Raoultella planticola isolates from January 2011 to December 2015 at a Turkish hospital [Demiray 2016]. A plasmid borne blaOXA-48 contained on Tn1999.2 was reported in a R. ornithinolytica clinical isolate from Lebanon in 2016 [Al-Bayssari 2016].

Within blaOXA-48-like members, carbapenemase activity varies significantly despite few sequence changes. Purified OXA-163 was shown to hydrolyze carbapenems at efficiencies far lower than several other OXA-48-like members. When blaOXA-163 was expressed in E. coli TOP10, this correlated with MICs below the clinical breakpoint for Enterobacteriaceae carbapenem susceptibility [Oueslati 2015]. However, MIC values for E. coli lacking OmpF and OmpC were above the susceptibility cut off for meropenem, doripenem, and ertapenem [Oueslati 2015]. OXA-405 is a blaOXA-48 type carbapenemase without carbapenemase activity discovered in a S. marcescens isolate from France in January 2011 [Dortet 2015]. Similar to OXA-163, it conferred increased resistance to extended-spectrum cephalosporins and aztreonam compared to OXA-48. Rapid global spread of blaOXA-48-like expressing Enterobacteriaceae necessitates molecular detection for informed patient treatment options and documentation of epidemiological occurrences to prevent nosocomial outbreaks [Poirel 2012].

Non-carbapenemase resistance mechanisms

Resistance-nodulation-division (RND) efflux pumps including the AcrAB-TolC system are a common resistance mechanism for Enterobacteriaceae and other Gram-negative bacteria against multiple antibiotic classes. Counterintuitively, KPC-2 production coupled with a loss-of-function AcrAB mutant in K. pneumoniae ECL-8 and E. coli BW25113 resulted in increased resistance against ertapenem and meropenem. Similarly, E. coli, Klebsiella spp., and Enterobacter spp. treated with the efflux pump inhibitor, phenylalanine-arginine β-naphthylamide (PABN), down-regulated ompF expression and had increased ertapenem resistance [Saw 2016]. RND efflux pump inhibition holds promise as synergistic treatment against MDR Gram-negative bacteria, but these results indicate that pump inhibition cannot be broadly applied across antibiotic classes [Aparna 2014]. Single nucleotide polymorphism (SNP) analysis between carbapenem-resistant, non-carbapenemase producing Enterobacter cloacae found that most SNPs are attributable to natural phylogenetic variation. However, SNPs in ampD leading to increased blaampC expression combined with mutations lowering expression of ompC, ompF, or both, thus increasing carbapenem resistance as well [Babouee Flury 2016]. As reported in 2015, K. pneumoniae clinical isolates from the United States and Israel with intermediate carbapenem susceptibility but lacking carbapenemase genes had more frequent porin nonsense mutations (ompK35, ompK36, and ompK37) and greater ESBL expression compared to K. pneumoniae with carbapenemases [Adler 2015].

As reported in 2013, loss of OmpK35 and OmpK36 porins in a K. pneumoniae isolate from Taiwan resulted in a 32-, 8-, and 4-fold increase in the MIC to ertapenem, meropenem, and doripenem respectively, but only ertapenem was beyond the clinical resistance breakpoint [Tsai 2013]. However, OmpK35/36 loss combined with expression of blaCTX-M-15 ESBL or the plasmid-mediated β-lactamase blaDHA-1 and its regulator ampR led to clinical resistance against ertapenem, meropenem, doripenem, and imipenem [Tsai 2013]. This combination of ESBL production and porin loss is believed to have contributed to dissemination of carbapenem-resistant K. pneumoniae within a tertiary-care hospital in South Korea from 2007 to 2008 [Shin 2012].

Stepwise-carbapenem resistance in the absence of a carbapenemase or β-lactamase was reported in Sweden in 2016 when an E. coli ompC and ompF double-deletion mutant was passaged in meropenem or ertapenem for ~60 generations [Adler 2016]. Meropenem but not ertapenem passaging caused mutations in stringent response regulator enzymes, spoT, thrS, and tufA [Adler 2016]. Increased stringent response activity is believed to contribute to carbapenem resistance but at a substantial fitness cost [Adler 2016].

Multiple antibiotic resistance protein A (marA) is an AraC-type transcriptional regulator (TR) responsible for antibiotic resistance in the Enterobacteriaceae primarily by controlling outer membrane permeability in coordination with two other AraC-type TRs, rob and soxS [Alekshun and Levy 1999]. Overexpression of an additional AraC-type TR, ramA, in the K. pneumoniae NCTC 5055 reference strain increased growth around an imipenem Kirby-Bauer disk, suggesting a role in regulating carbapenem resistance [Jimenez-Castellanos 2016]. As reported in 2013, an E. coli isolate from China lacked OmpF and OmpC porins and had a marA mutation which resulted in expression of the previously classified pseudogene yedS [Warner 2013]. G42RMarA facilitated expression of the YedS porin which is believed to contribute to E. coli isolate survival in the absence of OmpF and OmpC [Warner 2013]. These examples indicate that changing membrane permeability by altering efflux pumps and porin expression can confer carbapenem resistance alone or synergistically with ESBLs in the absence of a carbapenemase.

Detection

Initial CRE detection is often by analysis of susceptibility testing results from automated systems, broth microdilution assays, and Kirby-Bauer disk diffusion [Nordmann 2012]. Further characterization is done to differentiate carbapenemase producers from non-producers [Nordmann 2012]. The poor hydrolytic ability of some carbapenemases for carbapenems, most notably OXA-48-like enzymes and KPC variants, makes the appearance of carbapenem susceptible, carbapenemase producers also possible [Gagetti 2016, Poirel 2012]. PCR-based and phenotypic assays are the two detection methods for identifying CRE and CPE [Nordmann 2012].

Screening breakpoints for CPE detection are recommended by CLSI and EUCAST, but there is no consensus on the optimal breakpoints [CLSI 2016, EUCAST 2013]. A 2016 study suggested that CLSI screening breakpoints for CPE detection are missing carbapenem susceptible carbapenemase producers in carbapenemase isolates 14% (26/188) including 25% (14/63) of KPC producers and 40% (12/30) of OXA-48-like producers [Fattouh 2016].

The diversity of OXA-48-like carbapenemases makes precise molecular determination difficult [Hemarajata 2015]. However, a PCR-based assay using high-resolution melt time analysis can differentiate seven blaOXA-48-like family members as well as blaKPC, blaNDM-1, blaIMP, blaVIM, and blaSME [Hemarajata 2015]. A multiplex PCR assay using peptide nucleic acid probes reported high sensitivity and specificity (both above 99.0%) in identifying blaKPC, blaOXA-48, blaGES, blaIMP, blaVIM, and blaNDM from a mixture of Enterobacteriaceae isolates [Jeong 2015]. One assay combined nested PCR, rtPCR, and microfluidics to obtain class level identification of carbapenemases (blaKPC, blaVIM, blaOXA-23, blaOXA-48, blaOXA-51) and ESBLs for rapid detection of diverse multidrug resistant organisms including CRE [Walker 2016]. A PCR-based assay for rectal swabs had high sensitivity (96.6%) and specificity (98.6%) at identifying blaKPC, blaNDM, blaVIM, blaOXA-48, and blaIMP-1 on a rapid time-scale (32–48 minutes) [Tato 2016].

A comparison of disk diffusion assay detection of OXA-48-like carbapenemase producing Enterobacteriaceae from 2014 demonstrated that temocillin disk detection had similar sensitivity (all above 98%) to meropenem and piperacillin/tazobactam disk detection [Huang 2014]. Use of temocillin disk detection dramatically increased specificity (92.0% for K. pneumoniae and 89.3% for other species with temocillin compared to 46.0% for meropenem and 60.5% for piperacillin/tazobactam) using modified detection cut-offs of 29mm, 16mm, and 12mm for meropenem, piperacillin/tazobactam, and temocillin respectively [Huang 2014]. Temocillin can also be used for detection of MBL and KPC using Disks or tablets if synergy is not detected [EUCAST 2013].

The Modified Hodge Test (MHT) is a phenotypic assay for detection of CPE by inactivation of carbapenems allowing increased growth of an indicator strain [Girlich 2012]. However, the MHT has poor sensitivity for class B carbapenemases (sensitivity <50 %) which is marginally improved by addition of ZnSO4 (sensitivity 87%) [Girlich 2012]. As reported in Argentina in 2016, addition of 0.2% Triton X-100 [vol/vol] to Mueller Hinton Agar plates improved sensitivity (97–100%) for detecting class B carbapenemase expressing CPE [Pasteran 2016]. The proposed mechanism relies on the anionic detergent disturbing the hydrophobic interactions between the lipoprotein carbapenemases and the outer membrane [Gonzalez 2016, Pasteran 2016].

Matrix-assisted laser desorption ionization-time of flight mass spectrometry (MALDI-TOF MS) has become an essential tool within Clinical microbiology laboratories for rapid identification of bacterial pathogens as well as their resistance mechanisms by comparing a clinical isolate’s peptide mass fingerprint with known databases [Bohme 2012]. Additionally, MALDI-TOF MS can be adapted to study small molecules. As reported in France in 2015, a MALDI-TOF MS analysis of imipenem hydrolysis by cell supernatants had 100% specificity and sensitivity at differentiating CPE (OXA-48 variants, KPC, NDM, VIM, IMI, IMP, and NMC-A) from CRE that lack carbapenemases [Lasserre 2015]. The low cost per test (< 0.10 USD) and rapid protocol (~20 minutes) holds promise for widespread adoption by clinical microbiology labs [Lasserre 2015].

The easily interpretable Rapidec Carba NP Test showed high sensitivity and specificity (both 96%) at detecting CPE (in addition to A. baumannii and P. aeruginosa) with an under 30-minute incubation time [Poirel and Nordmann 2015]. A phenotypic assay using imipenem, EDTA, and EDTA plus phenylboronic acid had high sensitivity (96.3%) and specificity (97.7%) in identifying OXA-48-like family members within genetically characterized CRE [Tsakris 2015]. A novel colorimetric assay for carbapenemase detection showed high sensitivity (97.8%) and specificity (98.5%) at differentiating carbapenem-resistant, carbapenemase producers from non-producers [Kabir 2016]. An electrochemical assay using imipenem hydrolysis to drive redox changes was capable of detecting CPE in less than 35 minutes with high sensitivity (95%) and perfect specificity (100%) [Bogaerts 2016]. The carbapenem inactivation method is an efficient phenotypic assay that has high concordance with PCR-based detection of carbapenemase production in Enterobacteriaceae as well as Acinetobacter and Pseudomonas species. The simple protocol and required materials makes it promising for widespread adoption by clinical microbiology labs [van der Zwaluw 2015]. An immunochromatographic assay had perfect sensitivity (100%) and specificity (100%) at detecting OXA-48-like CPE compared to isolates possessing OXA-48-like variants lacking carbapenemase activity or Class A and B carbapenemases [Dortet 2016]. Accurate detection methods are important for developing ideal patient treatment options and surveillance of carbapenemases. In a study done between June 2010 to January 2014, an Italian hospital demonstrated that identification of carbapenemases by disk diffusion synergy and quarantine of CRE carriers significantly reduces incidence of high mortality BSIs [Viale 2015].

Plasmid Replicons

Plasmids are extrachromosomal DNA elements that often contain genes which confer fitness to specific extracellular stressors [Carattoli 2013]. Certain plasmids are conjugative and can be passed between members of the Enterobacteriaceae and with other gram negative pathogens such as A. baumannii and Pseudomonas aeruginosa [Vilacoba 2014]. Class A, B, and D carbapenemases contained on mobile gene elements within these plasmids is a core factor in the rapid spread of CRE [Mathers 2015].

IncL/M

IncL/M plasmids are self-transmissible replicons common amongst Enterobacteriaceae [Carattoli 2015]. Genomic characterization of IncL and IncM plasmid identified high nucleotide conservation (~94%) but two distinct branches when clustered for sequence similarity [Carattoli 2015]. Of 20 IncL/M plasmids sequenced by 2015, 18 harbored β-lactamases from ambler classes A–D including KPC, NDM, VIM, and OXA carbapenemases [Adamczuk 2015]. These plasmid types were originally endemic to the Mediterranean region but have become globally disseminated [Adamczuk 2015]. IncL/M plasmids containing blaOXA-48 were demonstrated to have a 40-fold higher transfer efficiency compared to plasmids with blaNDM-1 due to insertion of blaOXA-48 and its flanking transposon Tn1999 into the transfer inhibition protein gene [Potron 2014]. Genetic analysis of the plasmid containing blaOXA-48 determined the absence of other antibiotic resistance genes but the presence of a Tn1999 composite transposon on the common IncL/M type backbone [Poirel 2012]. Given widespread dissemination of the IncL/M type plasmids among Enterobacteriaceae and presence of multiple mobile genetic elements, threat of additional antibiotic resistance genes associating with blaOXA-48 is constant but currently undocumented [Manageiro 2015].

IncX3

A single IncX3 plasmid with two blaKPC-3 copies, both associated with Tn4401, was recovered from an XDR K. pneumoniae ST512 isolate in Italy in 2014. The MICs for carbapenems were identical between the IncX3-K. pneumoniae ST512 and a K. pneumoniae ST512 with KPC on the canonical pkpQIL plasmid. IncX3 E. coli transformants had greater carbapenem resistance than the pkpQIL, but this could be due to differences in plasmid copy number or KPC expression in addition to the two blaKPC-3 genes [Fortini 2016]. An E. coli ST 3835 isolate obtained in November 2012 from China contained blaNDM-1 and blaSHV-12 on the IncX3 plasmid backbone [Feng 2015]. This narrow-spectrum plasmid was found harboring blaNDM-1 in several Enterobacteriaceae isolates in China, suggesting a possible role in global dissemination of blaNDM-1 [Feng 2015]. A comparison of CPE collected from 2013 to 2014 in a Canada found that blaNDM-7 was harbored on an identical IncX3 plasmid across K. pneumoniae, E. coli, Enterobacter hormaechei, and S. marcescens, indicating frequent horizontal gene transfer events within Enterobacteriaceae [Chen 2015]. Five CRE collected at a hospital in China from January to September 2013 possessed blaNDM-1 on an IncX3 plasmid with high similarity in the immediate region surrounding blaNDM-1, but one plasmid was twice as large (104.5 to 138.9 kb) as the other four (33.3 to 54.7 kb), indicating genetic exchanges during horizontal gene transmission between Enterobacteriaceae [Yang 2015].

IncF

IncFII-type plasmids (7/12) were the most common background carrying NDM in a cohort of NDM positive isolates obtained in Australia from 2012 to 2014. IncX3 (4/12) and IncHI1B (1/12) were also identified, but every NDM variant was within Tn125 [Wailan 2016]. blaNDM-5 on an IncFII plasmid was found on two K. pneumoniae ST147 isolates in a suspected nosocomial transmission event at a hospital in South Korea in 2014 [Shin 2016]. A K. pneumoniae isolate obtained from China in October 2010 expressed blaKPC-2 on a IncFIIK conjugative plasmid in association with a different transposon, Tn1772, indicating carriage plasticity by mobile gene elements [Wang 2015]. As reported in 2015, a patient from Spain with no recent travel history had a urine culture positive for E. coli ST448 with blaNDM-5 on a conjugatable IncFII-type plasmid [Pitart 2015]. The first MBL expressing Enterobacteriaceae isolate obtained from Poland in August 2011, an E. coli ST410, was found to carry blaNDM-1 on a IncFII plasmid [Fiett 2014]. An IncFIIy type plasmid containing NDM and three ESBLs (blaTEM, blaCTX-M-1, and blaOXA-1) together was isolated from the first instance of a carbapenem-resistant Leclercia adecarboxylata from China in August 2012 [Sun 2015]. In Europe, clonal group 258 is the primary source of KPC producing K. pneumoniae, with the eponymous ST258 being the most common (69.1%, 76/110) followed by ST512 (19.1%, 21/110) [Baraniak 2016]. Both predominantly possess IncFIIB and IncFIBK, contrasting with non-CG258 K. pneumoniae and other CPE species which mostly possess IncN plasmids [Baraniak 2015].

IncA/C

Chromosomally integrated blaNDM-1 expressing P. aeruginosa and blaNDM-1 on IncA/C plasmid in E. coli were obtained from the same patient [Mataseje 2016]. As reported in 2016, IncA/C2 and IncFIIy plasmids carrying NDM in Enterobacteriaceae from Australia and New Zealand all had segments with high sequence similarity to Acinetobacter plasmids, further documenting frequent genetic exchange with non-Enterobacteriaceae members [Wailan 2016]. Multidrug resistant IncA/C plasmids containing integrons In4863 (blaVIM-19) were identified in two carbapenem-resistant K. pneumoniae ST383 isolates with suspected origin from Greece in 2009–2011 [Papagiannitsis 2016]. Multidrug resistant, conjugative IncA/C plasmids were recovered from two K. pneumoniae ST11 isolates obtained from a hospital in Bulgaria in 2014. The larger (176-kb) plasmid possessed blaNDM-1 with blaCMY-4, blaCTX-M-15, blaTEM-1b, and qnrB while the smaller (86-kb) plasmid had blaNDM-1 with blaTEM-1, and aac(6′)-Ib [Todorova 2016]. blaOXA-48 has additionally been identified on an IncA/C plasmid in Taiwan (2015) and Tunisia (2011) [Ktari 2011].

Other

A K. pneumoniae ST25 isolate obtained from a Portuguese ICU in 2009 possessed blaIMP-8 on a ColE1-like plasmid in association with a novel class 3 integron that also contained the weak carbapenemase blaGES-5, aacA4, and blaBEL-1, an ESBL only identified in P. aeruginosa previously. Formerly, these genes were found solely on class 1 integrons; this suggests a resistance cassette exchange between class 1 and 3 integrons, with potential IncQ plasmid involvement, may be responsible [Papagiannitsis 2015].

blaNDM-1 was discovered on a BJ01-like plasmid in an E. aerogenes isolate from China in June 2012 [Chen 2015]. Since previous reports of pNDM-BJ01 have exclusively been in Acinetobacter, this identification confirms that horizontal gene transfer between Gammaproteobacteria is a continual occurrence [Chen 2015].

The first instance of the narrow-spectrum plasmid, IncT, containing blaNDM-1 was from a Providencia rettgeri isolate colonizing a patient in Canada who received medical care in India in 2015 [Mataseje 2016].

Non-infectious Surveillance

Prediction of carbapenemase dissemination is complicated by the discovery of new reservoirs for antibiotic resistance genes. An E. coli CMUL 64 specimen isolated from domesticated chickens in Lebanon in December 2013 contained blaOXA-48 in addition to two ESBLs [Al Bayssari 2015]. Surveillance of Silver Gull (Chroicocephalus novaehollandiae) excrement from South-East Australia in October 2012 found that CPE carriage was alarmingly high (71.8%, 120/167) and almost completely due to blaIMP-4 (96.7%, 116/120), but blaIMP-26 (1.67%, 2/120) and blaIMP-38 (5.83%, 7/120; 5/7 in E. coli ST58 co-producing blaIMP-4) were also detected [Dolejska 2016]. This agrees with clinical observations reporting blaIMP-4 as the most prevalent gene from CPE clinical isolates in Australia, indicating either high transmission frequency of blaIMP-4 to human pathogens or that these CPE environmental isolates can be infectious agents [Dolejska 2016]. Contrasting with the seemingly high carriage of CPE in birds, CPE are absent (0.6%, 1/160) from studied companion dog fecal microbiota in Madrid, Spain from October 2014 to January 2015 excepting one K. pneumoniae ST2090 expressing blaVIM-1 on an untyped plasmid [Gonzalez-Torralba 2016]. Similarly, WGS analysis of fecal microbiota collected from dairy cows in the southwestern United States from May to July 2014 failed to detect any CRE, though it did find other carbapenem-resistant Gram-negative bacteria [Webb 2016].

Assessment of individual carriage potential demonstrated that recent antibiotic therapy, immunosuppression, and a Charlson Comorbidity Index >4 are all significant predictors of CRE infection [Miller and Johnson 2016]. A US case report from February 2016 indicates that the Middle East may be another reservoir for blaNDM genes in addition to the Indian subcontinent and the Balkans [Li 2016]. Identifying intra-hospital carriage of CRE is important given the high fatality rate from CRE infections in patients with solid organ transplants from endemic areas [Satlin 2014].

A study published in 2016 identified blaIMI-3 on a IncFIIY plasmid obtained from sediment at the Haihe River in China showed high genetic similarity with IncF plasmids from pathogenic Enterobacteriaceae [Dang 2016]. As reported in 2016, an IncP-1β plasmid also obtained from sediment at the Haihe River contained a blaGES-5 variant with several amino acid substitutions that led to loss of carbapenemase activity [Dang 2016]. Two E. coli isolates expressing blaIMP-8 and two E. coli isolates expressing blaVIM-1 and blaVIM-34 were obtained in February 2015 from the Ave river in northern Portugal. blaIMP-8 and blaVIM-1 were found on IncFIB type plasmids capable of transconjugation while blaVIM-34 was suspected to be chromosomally integrated [Kieffer 2016]. As reported in 2016, water collected from Rio De Janeiro yielded many blaKPC producing Enterobacteriaceae as well as blaGES, blaOXA-48-like, and blaNDM genes amplified directly from aquatic microbial communities [de Araujo 2016]. Routine environmental sampling could prove advantageous in anticipating spread of novel antibiotic resistance genes into clinical pathogens.

Contribution of Carbapenemases to MDR/XDR Enterobacteriaceae

Dual Carbapenemases

Presence of multiple carbapenemase genes in one isolate is concerning given the possibility of different carbapenemase genes crossing onto the same plasmid. A K. pneumoniae isolate obtained in the United States in 2013 from the high risk ST258 clade expressed plasmid borne blaKPC-2 and blaVIM-4 [Castanheira 2016]. An E. cloacae ST231 from China in 2012 was found to harbor blaNDM-1 on IncA/C2 mosaic plasmid and blaKPC-3 on a novel IncX6 plasmid [Du 2016]. These same resistance genes were found on two different IncF plasmids in an E. cloacae isolate from China in April 2014, further increasing the likelihood that a recombination event could create a dual carbapenemase plasmid [Wu 2015]. In April 2014 a K. pneumoniae ST147 strain possessing blaNDM-5 and blaOXA-181 was isolated at a hospital in South Korea from a patient native to the United Arab Emirates [Cho 2015]. Four months later, a K. pneumoniae isolate with identical ST, blaOXA-181, and blaNDM-5 genes was obtained from a patient native to South Korea with no history of travel abroad [Cho 2015]. The patients did not overlap temporally or spatially, which suggests extended surface survival of K. pneumoniae [Cho 2015]. This occurrence is also documented outside of K. pneumoniae; in 2016 an E. coli ST410 isolate from Egypt was found to have blaNDM-5 and blaCTX-M-15 on a transmissible ~100 kb plasmid with 99% similarity to pHC105 and blaOXA-181 on a 48.5 kb plasmid that did not transform a recipient strain [Gamal 2016]. A C. freundii isolate from China in July 2013 possessed blaNDM-1 on an IncX3 type plasmid and blaKPC-2 on an untypable plasmid that is notable for a region with genes conferring resistance to macrolides (mphA-mrx-mphR operon), quinolones (aac(6′)-lb-cr), rifampin (arr-3), chloramphenicol (catB3), sulfonamides (sul1), cephalosporins (blaOXA-1), quaternary ammonium compounds (qacEΔl), chromate (chrA), and fosfomycin (fosA3) [Feng 2015].

Tigecycline Resistance

Tigecycline is a tetracycline derivative and one of the few remaining drugs for treating CRE infections. 56 tigecycline non-susceptible K. pneumoniae from South Korean hospitals in 2012 showed that co-resistance with carbapenems is low (2/56) compared to ceftazidime, ceftriaxone, and aztreonam (37/56, 36/56, and 35/56 respectively) [Ahn 2016]. But prior admission to a skilled nursing facility was significantly associated with tigecycline non-susceptible, carbapenem-resistant K. pneumoniae cultures in hospitalized patients [van Duin 2015]. Additionally, tigecycline resistance in carbapenem-resistant K. pneumoniae bacteriuria occurred significantly following tigecycline monotherapy in a multicenter, prospective study from December 2011 to October 2013 in the United States [van Duin 2014]. Clinical resistance to Tigecycline was associated with increased expression of AcrAB by marA mutations in KPC-producing K. pneumoniae isolates from China collected between January 2010 and December 2013 [Hemarajata 2015].

Aminoglycoside Resistance

Gentamicin, amikacin, and tobramycin are aminoglycosides occasionally used to treat CRE [Morrill 2015]. rmtB is a 16s rRNA methylase conferring high level aminoglycoside resistance that is commonly found on IncF, IncA/C, IncK, and IncN plasmids [Kang 2008, Yu 2010]. These plasmids are common in Enterobacteriaceae and indeed rmtB was found in 34% (25/74) of K. pneumoniae ST11 isolates from China in 2012 to 2014 with 97% (72/74) positive for blaKPC-2 and the remainder (2/74) positive for blaNDM-1 [Cheng 2016]. Quinolone resistance genes oqxA, oqxB, and qnrB were also common in 81% (60/74), 76% (56/74), and 8 % (6/74) of isolates, respectively [Cheng 2016]. rmtF is a novel 16S rRNA methyltransferase which was found to confer high aminoglycoside resistance in Enterobacteriaceae isolates collected from 2009 to 2011 in India and the United Kingdom [Hidalgo 2013]. 20 of the 34 rmtF expressing Enterobacteriaceae also contained NDM-1 [Hidalgo 2013]. Two ST11 K. pneumoniae isolated from a tertiary care hospital in Egypt from September 2013 to December 2014 possessed blaNDM-1 and rmtF on the same untypeable 170 kb plasmid [Gamal 2016].

Colistin Resistance

Colistin is an old and rarely used (due to severe nephrotoxicity) polymyxin antibiotic that has emerged as a CRE treatment, making co-resistance concerning. Until the discovery of the plasmid-mediated colistin resistance gene, mcr-1, resistance was attributed to chromosomal modification of the LPS biosynthesis pathway [Liu 2016]. Mcr-2, a colistin resistance gene with 76.7% nucleotide identity to mcr-1, was discovered on a conjugatable IncX4 plasmid in E. coli ST10 and ST167 from Belgium in July 2016 [Xavier]. The IncX4 plasmid did not contain any additional resistance genes, but mcr-2 was contained within an IS element from the IS1595 superfamily, which was associated with the carbapenemase blaOXA-23 in Acinetobacter radioresistens [Higgins]. A third (32/97) of carbapenem-resistant K. pneumoniae found in Italy from December 2010 to May 2011 displayed a colistin MIC of 16 μg/mL, which was associated with higher mortality compared to infection by colistin susceptible, carbapenem-resistant K. pneumoniae [Capone 2013]. The first characterization of a blaIMI-1 producing, colistin resistant E. cloacae from the United States in February 2015 showed absence of mcr-1 and canonical chromosomal modifications, suggesting that Enterobacteriaceae contain other colistin resistance mechanisms [Norgan 2016]. A retroactive study on colistin-resistant Enterobacteriaceae isolates obtained from January 2013 to November 2015 at a hospital in China found two K. pneumoniae harboring mcr-1 and blaNDM-5 [Du 2016]. Four blaKPC-2 producing K. pneumoniae ST101 from a cohort of nosocomial CRE obtained in Italy from November 2013 to August 2014 had colistin resistance (MIC 16–125 μg/ml) through an unidentified mechanism [Del Franco 2015].

A K. pneumoniae ST147 isolated from patient urine in the United Arab Emirates in April 2014 was phenotypically pan drug resistant by the Vitek 2 semi-automated system [Zowawi 2015]. Resistance to carbapenems is derived from the OXA-48-like family member, blaOxa-181, with contributions from an inactivating mutation in OmpK36 [Zowawi 2015]. The isolate was resistant to many antibiotics by acquired genes, but colistin resistance was conferred through a chromosomal copy of the blaOXA-181 transposon disrupting the mgrB allele. Tigecycline resistance was attributed to an inactivating mutation in ramR allowing for increased expression of acrAB [Zowawi 2015].

Treatment Options

Antibiotic Combinations

Therapeutic potency of antibiotics can be increased (synergy), decreased (antagonism), or unaffected (additivity) when combined with other each other as determined by their fractional inhibitory concentration index [Doern 2014]. Most widely used combinations were formulated from clinical observations of effectiveness but not heavily scrutinized for mechanistic basis of synergy [Baym 2016]. Understanding how commonly used CRE infection antibiotic treatments affect each other and identifying new combinations is necessary to optimize CRE treatment.

As reported in 2015, a regimen of intravenous colistin, meropenem, and ertapenem successfully treated a bacteremic, multidrug resistant, KPC expressing K. pneumoniae infection in an elderly patient from Italy. In vitro analysis indicated synergy of colistin/ertapenem/meropenem when each antibiotic is 0.5 × MIC, with MICs of 32, 128, and 256 μg/ml respectively [Oliva 2015]. A retrospective study of patients who received ertapenem-containing double-carbapenem therapy from October 2013 to November 2014 at a hospital in the United States also supported the efficacy of dual carbapenem treatment regimens, finding that ertapenem treatment followed by doripenem or meropenem had promising microbiological success (79%, 11/14) [Cprek and Gallagher 2016].

Resurgence of polymyxins as a CRE treatment motivated further study of other antibiotics with historically limited use, such as chloramphenicol [Abdul Rahim 2015]. Colistin and chloramphenicol were synergistic in 89% (25/28) of cases against three NDM-producing MDR K. pneumoniae clinical isolates and the K. pneumoniae NDM reference strain BAA-ATCC-2146 in vitro. All chloramphenicol resistant isolates were determined by PCR to use efflux pumps, so further research is needed to see if the combination is synergistic against bacteria with carbapenemases and chloramphenicol acetyltransferases [Abdul Rahim 2015]. Polymyxin B and tigecycline together and in triple combination with meropenem significantly increased rat survival during systemic infection with blaKPC-2 expressing K. pneumoniae, but the double combination had superior in vitro activity. This could be explained by antagonistic effects between polymyxin B and meropenem [Toledo 2015]. While empiric antibiotic therapy has led to development of clinically useful combinations, most notably trimethoprim-sulfamethoxazole and β-lactam/β-lactamase inhibitor combinations, immediate benefits from synergy may be countered by quicker evolution of resistance to these pairings [Yeh 2009]. Applying systems biology to understand how antibiotic combinations interact with isolate resistomes is necessary to mechanistically understand the increased effectiveness and design collaterally sensitive formulations [Gonzales 2015, Pal 2015, Roemer and Boone 2013].

Avibactam

β-lactamase inhibitors have extended the spectrum and efficacy of several β-lactam antibiotics, and a few β-lactam/β-lactamase inhibitor combinations have gained widespread clinical use [Drawz and Bonomo 2010]. Unfortunately, our current β-lactamase inhibitors are ineffective against Class B β-lactamases; this is especially concerning given the global dissemination of NDM positive Enterobacteriaceae and regional prevalence of VIM and IMP enzymes [Watkins 2013].

Avibactam is a recent non-β-lactam, β-lactamase inhibitor with broad efficacy against serine β-lactamases, particularly KPC [Krishnan 2015]. Structural analysis showed that avibactam covalently binds to the S70 residue in the active site of two class A β-lactamases, KPC-2 and SHV-1 [Krishnan 2015]. Combining avibactam with ceftazidime, a third generation cephalosporin, shows promise as treatment for non-MBL expressing CRE [Castanheira 2015, Zhanel 2013]. KPC-2 engineered to be avibactam resistant had increased ceftazidime susceptibility, suggesting collateral sensitivity as an explanation for the combination’s clinical efficacy [Papp-Wallace 2015].

However, in a United States hospital in 2015, K. pneumoniae isolated from a patient not previously exposed to ceftazidime-avibactam treatment showed resistance to the combination (MIC 32/4 μg/ml). The only identified carbapenemase was a blaKPC-3 with 100% identity to previously sequenced variants. Additionally, the ceftazidime-avibactam resistance mechanism does not appear efflux based because treatment with PABN did not decrease the MIC. The isolate was multidrug resistant, but was susceptible to aminoglycosides, colistin, and trimethoprim-sulfamethoxazole [Humphries 2015]. In vitro studies suggest decreased membrane permeability via OmpK36 mutation, and the expression of an ESBL in conjunction with KPC-2, or expression of KPC-3 instead of KPC-2 correlated with increased ceftazidime-avibactam resistance [Shields 2015].

The monobactam aztreonam is resistant to inactivation by NDM, but not serine β-lactamases. Accordingly, high levels of NDM and serine β-lactamase co-occurrence reduces the clinical utility of aztreonam monotherapy [Shakil 2011]. An aztreonam-avibactam combination inhibited 99.9% (n=23,516) of Enterobacteriaceae species at 4/4 μg/mL. The combination was particularly effective against Enterobacteriaceae isolates possessing a MBL in addition to serine β-lactamases [Biedenbach 2015]. Aztreonam-avibactam activity had broader efficacy than ceftazidime-avibactam against a panel of CPE species possessing a variety of Class A, B, and D carbapenemases [Vasoo 2015]. An E. cloacae ST88 isolate from a hospital in the United States with chromosomally integrated blaKPC-18 and blaVIM-1 on a 58-kb multidrug resistant IncN plasmid was reported in 2016. The E. cloacae was highly resistant to ceftazidime-avibactam (MIC->256/4) but not aztreonam-avibactam as (MIC-0.5/4) [Thomson 2016].

In 2015, avibactam-ceftazidime combined with ertapenem successfully treated a patient infected with blaNDM-1 expressing pan drug resistant K. pneumoniae at a hospital in the United States. Avibactam has not demonstrated activity against class B β-lactamases, but it showed in vitro synergy with ceftazidime and carbapenems [Camargo 2015]. These current investigations demonstrate the utility of avibactam to extend the effectiveness of aztreonam and ceftazidime against CRE, judicious use will hopefully preserve this efficacy.

Emerging Options

In addition to exploring combinations of antibiotic and resistance inhibitors, urgent work is needed to develop novel antimicrobial therapeutics to treat CRE infections and eliminate colonization. Medicinal chemistry to develop new antibiotic classes and modify current compounds to evade resistance mechanisms is key to revitalizing our arsenal against CRE [Pucci and Bush 2013]. Immune based therapies augmenting endogenous innate and adaptive effector mechanisms are being investigated as alternatives to traditional broad spectrum antibiotics [DiGiandomenico and Sellman 2015, Li 2014]. Restoration of diverse gut microbiota by fecal microbiota transplant (FMT) is being used to restore colonization resistance against MDR enteric pathogens following antibiotic treatment [Allen 2014]. Emerging synthetic tools are being used to identify genetic cytotoxicity loci that can augment or replace antibiotic treatment [Vercoe 2013].

A novel cephalosporin with a 3-position catechol moiety that acts as a siderophore to sequester ferric iron displayed low MIC values against a diverse group of CRE species and carbapenemase genes [Kohira 2015]. The next generation aminoglycoside, plazomicin, had an MIC90 value of 1 μg/ml against 164 Enterobacteriaceae isolates expressing Class A (blaKPC-2, n =34), Class B (blaVIM-1, n=125; blaIMP-22, n=1), or Class D (blaOXA-48, n=4) carbapenemases compared to MIC90 of 256, 64, and 16 μg/ml for gentamicin, tobramycin, and amikacin respectively [Rodriguez-Avial 2015].

Monoclonal antibodies targeting poly-(β-1,6)-N-acetyl glucosamine, a polysaccharide implicated as an E. cloacae and K. pneumoniae virulence factor, showed significant in vivo protection against NDM-1 producing Enterobacteriaceae in a mouse peritonitis model. This target is a promising new vaccine candidate against invasive infection [Skurnik 2016]. Cationic antimicrobial peptides (AMPs) are an ancient feature of multicellular immune systems that hold potential for novel bacterial infection treatments. Two piscidin family AMPs isolated from tilapia (Oreochromis niloticus) in Taiwan in 2012 exhibited in vivo bacteriostatic effects yielding superior survival outcomes compared to tigecycline and imipenem in a mouse sepsis model infected with blaNDM-1 producing K. pneumoniae [Pan 2015].

FMTs are a promising alternative to antibiotics for treatment of inflammatory bowel disorders and active Clostridium difficile infections [Surawicz 2013, Wang 2014]. Commensal gut flora is believed to outcompete enteric pathogens and then establish further colonization resistance [Wang 2014]. As reported in 2015 from the United States, antibiotic administration to mice perturbs the intestinal microbiota enough to lose colonization resistance against vancomycin resistant E. faecium and carbapenem resistant K. pneumoniae [Caballero 2015]. Pre-colonization with one pathogen does not prevent the colonization by the other, but FMT administration reduced levels of both species [Caballero 2015]. This suggests the use of fecal microbiota transplants as a putative clearance method for non-infected, CRE colonized patients [Caballero 2015].

Transduction of carbapenem resistant blaNDM-1 E. coli with CRISPR/Cas9 based RNA guided nucleases for blaNDM-1 resulted in a three-log10 reduction of bacterial CFU [Citorik 2014]. The proposed mechanism relies on an increase in dsDNA breaks leading to activation of the SOS response and antibiotic independent cell death [Citorik 2014]. A combinatorial TR overexpression screen in blaNDM-1 E. coli identified many gene pairs that induced lethality in the presence or absence of ceftriaxone [Cheng 2014]. These results support the notion that antibiotic cytotoxicity is a result of target specific inhibition coupled to increased redox activity via a global stress response [Dwyer 2014]. Further work is needed to characterize additional vulnerabilities in CRE.

Conclusion

Although individual resistance gene presence varies by geography, CRE represents a worsening global threat that ignores national borders. Within the United States, CRE are suspected to cause over 9,000 cases of healthcare-associated Enterobacteriaceae infections [Yaffee 2016]. Predictions indicate that infections caused by MDR bacteria will increase substantially, with MDR E. coli expected to cause 3 million deaths each year by 2050 [O’Neill 2016]. In-depth characterization of clinical CRE isolates is necessary to identify carbapenemase burden and distribution in endemic and non-endemic areas [Lutgring and Limbago 2016]. Additionally, surveillance of CRE in patients, the environment, and animals should be continued to identify reservoirs for carbapenemase genes [Guerra 2014, Viau 2016]. The diversity of plasmids containing carbapenemase genes and propensity of these plasmids to contain multiple antibiotic resistance genes and mobile gene elements foreshadows increasing incidence of extensively drug resistant Enterobacteriaceae [Tangden and Giske 2015]. Though avibactam with ceftazidime or aztreonam is a promising CRE treatment, developing novel antibiotic combinations and revamping the antibiotic development pipeline is required to suppress the worsening CRE threat.

Table 1.

Newly reported class A carbapenemases.

Gene Enterobacteriaceae Year Reported Country Plasmid Accession number Activity Reference
blaBKC-1 K. pneumoniae 2015 Brazil IncQ KP689347 Carbapenems, monobactams, penicillins, cephalosporins 120
blaFRI-1 E. cloacae 2015 France Untypeable KT192551 Carbapenems, monobactams, aztreonam, penicillins 61

Table 2.

Environmental CRE surveillance.

Source Enterobacteriaceae Carbapenem Resistance Year Reported Country Reference
Chicken E. coli blaOXA-48 2015 Lebanon 8
Silver Gull E. coli, Escherichia fergusonii, K. pneumoniae, Kluyvera georgiana, E. aerogenes, E. cloace, C. freundii, Citrobacter braakii, P. mirabilis, Proteus penneri blaIMP-4, blaIMP-26, blaIMP-38 2016 Australia 57
Companion dog K. pneumoniae blaVIM-1 2016 Spain 81
Dairy Cow E. coli, Aeromonass veronii, Aeromonas allosaccharophila ESBL+Porin mutation 2016 United States 197
Ave River E. coli blaVIM-1, blaVIM-34, blaIMP-8 2016 Portugal 99
Carioca River E. cloace/asburiae, K. pneumoniae, Klebsiella spp, E. kobei blaKPC 2016 Brazil 51

Table 3.

Co-occurrence of antibiotic resistance genes in CPE isolates.

Isolate Carbapenemase Identified Antibiotic resistance genes Method Year Reported Country Reference
K. pneumoniae blaOXA-181 rmtF, aac(6′)-lb-cr, blaCTX-M-15, gyrA, parC, qnrB, afrA12, drfA14, acrAB, blaSHV-36, catB1, fosA, tetC WGS 2015 United Arab Emirates 213
K. pneumoniae blaNDM-5, blaOXA-181 blaTEM-1, blaSHV-11, blaCTX-M-15, rmtB PCR 2015 South Korea 44
E. coli blaNDM-5, blaOXA-181 blaCTX-M-15, blaCMY-2, aac(3)-IIa, aac(6′)-lb-cr PCR 2016 Egypt 76
C. freundii blaKPC-2, blaNDM-1 aac(6′)-lb-cr, blaOXA-1, catB3, arr-3, sul1, ampR, qnr, blaCTX-M-14, fosA3, blaSHV-12 PCR 2015 China 69
E. cloacae blaIMI-1 blaAmpC, ampR, ampD, fosA, cat, emrB, macB, mexE, mexX, acrAB, tolC, robA, msbA, ompL, oprD, ompC, phoP, phoQ WGS 2015 United States 123

Footnotes

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

References

  1. Abdul Rahim N, Cheah SE, Johnson MD, Yu H, Sidjabat HE, Boyce J, et al. Synergistic killing of NDM-producing MDR Klebsiella pneumoniae by two ‘old’ antibiotics-polymyxin B and chloramphenicol. J Antimicrob Chemother. 2015;70(9):2589–2597. doi: 10.1093/jac/dkv135. http://dx.doi.org/10.1093/jac/dkv135, Epub 2015/05/30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Adamczuk M, Zaleski P, Dziewit L, Wolinowska R, Nieckarz M, Wawrzyniak P, et al. Diversity and global distribution of IncL/M plasmids enabling horizontal dissemination of beta-lactam resistance genes among the Enterobacteriaceae. BioMed Res Int. 2015;2015:414681. doi: 10.1155/2015/414681. http://dx.doi.org/10.1155/2015/414681, Epub 2015/08/04. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Adler A, Ben-Dalak M, Chmelnitsky I, Carmeli Y. Effect of resistance mechanisms on the inoculum effect of carbapenem in Klebsiella pneumoniae isolates with borderline carbapenem resistance. Antimicrob Agents Chemother. 2015a;59(8):5014–5017. doi: 10.1128/AAC.00533-15. http://dx.doi.org/10.1128/AAC.00533-15, Epub 2015/05/20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Adler A, Hussein O, Ben-David D, Masarwa S, Navon-Venezia S, Schwaber MJ, et al. Persistence of Klebsiella pneumoniae ST258 as the predominant clone of carbapenemase-producing Enterobacteriaceae in post-acute-care hospitals in Israel, 2008–13. J Antimicrob Chemother. 2015b;70(1):89–92. doi: 10.1093/jac/dku333. http://dx.doi.org/10.1093/jac/dku333, Epub 2014/09/11. [DOI] [PubMed] [Google Scholar]
  5. Adler M, Anjum M, Andersson DI, Sandegren L. Combinations of mutations in envZ, ftsI, mrdA, acrB and acrR can cause high-level carbapenem resistance in Escherichia coli. J Antimicrob Chemother. 2016;71(5):1188–1198. doi: 10.1093/jac/dkv475. http://dx.doi.org/10.1093/jac/dkv475, Epub 2016/02/13. [DOI] [PubMed] [Google Scholar]
  6. Ahn C, Yoon SS, Yong TS, Jeong SH, Lee K. The resistance mechanism and clonal distribution of tigecycline-nonsusceptible Klebsiella pneumoniae isolates in Korea. Yonsei Med J. 2016;57(3):641–646. doi: 10.3349/ymj.2016.57.3.641. http://dx.doi.org/10.3349/ymj.2016.57.3.641, Epub 2016/03/22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Al Bayssari C, Olaitan AO, Dabboussi F, Hamze M, Rolain JM. Emergence of OXA-48-producing Escherichia coli clone ST38 in fowl. Antimicrob Agents Chemother. 2015;59(1):745–746. doi: 10.1128/AAC.03552-14. http://dx.doi.org/10.1128/AAC.03552-14, Epub 2014/10/29. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Al-Bayssari C, Olaitan AO, Leangapichart T, Okdah L, Dabboussi F, Hamze M, et al. Whole-genome sequence of a blaOXA-48-harboring Raoultella ornithinolytica clinical isolate from Lebanon. Antimicrob Agents Chemother. 2016;60(4):2548–2550. doi: 10.1128/AAC.02773-15. http://dx.doi.org/10.1128/AAC.02773-15, Epub 2016/02/03. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Alekshun MN, Levy SB. The mar regulon: multiple resistance to antibiotics and other toxic chemicals. Trends Microbiol. 1999;7(10):410–413. doi: 10.1016/s0966-842x(99)01589-9. Epub 1999/09/28. [DOI] [PubMed] [Google Scholar]
  10. Allen HK, Trachsel J, Looft T, Casey TA. Finding alternatives to antibiotics. Ann N Y Acad Sci. 2014;1323:91–100. doi: 10.1111/nyas.12468. http://dx.doi.org/10.1111/nyas.12468, Epub 2014/06/24. [DOI] [PubMed] [Google Scholar]
  11. Antonelli A, D’Andrea MM, Di Pilato V, Viaggi B, Torricelli F, Rossolini GM. Characterization of a novel putative Xer-dependent integrative mobile element carrying the bla(NMC-A) carbapenemase gene, inserted into the chromosome of members of the Enterobacter cloacae complex. Antimicrob Agents Chemother. 2015;59(10):6620–6624. doi: 10.1128/AAC.01452-15. http://dx.doi.org/10.1128/AAC.01452-15, Epub 2015/08/08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Aparna V, Dineshkumar K, Mohanalakshmi N, Velmurugan D, Hopper W. Identification of natural compound inhibitors for multidrug efflux pumps of Escherichia coli and Pseudomonas aeruginosa using in silico high-throughput virtual screening and in vitro validation. PLoS One. 2014;9(7):e101840. doi: 10.1371/journal.pone.0101840. http://dx.doi.org/10.1371/journal.pone.0101840, Epub 2014/07/16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Babouee Flury B, Ellington MJ, Hopkins KL, Turton JF, Doumith M, Loy R, et al. Association of novel nonsynonymous single nucleotide polymorphisms in ampD with cephalosporin resistance and phylogenetic variations in ampC, ampR, ompF, and ompC in Enterobacter cloacae isolates that are highly resistant to carbapenems. Antimicrob Agents Chemother. 2016;60(4):2383–2390. doi: 10.1128/AAC.02835-15. http://dx.doi.org/10.1128/AAC.02835-15, Epub 2016/02/10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Bakthavatchalam YD, Anandan S, Veeraraghavan B. Laboratory detection and clinical implication of oxacillinase-48 like carbapenemase: the hidden threat. J Global Infect Dis. 2016;8(1):41–50. doi: 10.4103/0974-777X.176149. http://dx.doi.org/10.4103/0974-777X.176149, Epub 2016/03/26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Baraniak A, Izdebski R, Fiett J, Herda M, Derde LP, Bonten MJ, et al. KPC-like carbapenemase-producing Enterobacteriaceae colonizing patients in Europe and Israel. Antimicrob Agents Chemother. 2015;60(3):1912–1917. doi: 10.1128/AAC.02756-15. http://dx.doi.org/10.1128/AAC.02756-15, Epub 2015/12/30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Baraniak A, Izdebski R, Fiett J, Gawryszewska I, Bojarska K, Herda M, et al. NDM-producing Enterobacteriaceae in Poland, 2012–14: interregional outbreak of Klebsiella pneumoniae ST11 and sporadic cases. J Antimicrob Chemother. 2016a;71(1):85–91. doi: 10.1093/jac/dkv282. http://dx.doi.org/10.1093/jac/dkv282, Epub 2015/09/21. [DOI] [PubMed] [Google Scholar]
  17. Baraniak A, Izdebski R, Fiett J, Herda M, Derde LP, Bonten MJ, et al. KPC-like carbapenemase-producing Enterobacteriaceae colonizing patients in Europe and Israel. Antimicrob Agents Chemother. 2016b;60(3):1912–1917. doi: 10.1128/AAC.02756-15. http://dx.doi.org/10.1128/AAC.02756-15, Epub 2015/12/30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Barguigua A, Zerouali K, Katfy K, El Otmani F, Timinouni M, Elmdaghri N. Occurrence of OXA-48 and NDM-1 carbapenemase-producing Klebsiella pneumoniae in a Moroccan university hospital in Casablanca, Morocco. Infect Genet Evol. 2015;31:142–148. doi: 10.1016/j.meegid.2015.01.010. http://dx.doi.org/10.1016/j.meegid.2015.01.010, Epub 2015/01/27. [DOI] [PubMed] [Google Scholar]
  19. Baym M, Stone LK, Kishony R. Multidrug evolutionary strategies to reverse antibiotic resistance. Science. 2016;351(6268):aad3292. doi: 10.1126/science.aad3292. http://dx.doi.org/10.1126/science.aad3292, Epub 2016/01/02. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Bell B, Bell M, Bowen A, Braden C, Brandt M, Brown A, et al. In: Antibiotic Resistance Threats. Services USDoHaH, editor. Center for Disease Control and Prevention; 2013. [Google Scholar]
  21. Bialek-Davenet S, Criscuolo A, Ailloud F, Passet V, Jones L, Delannoy-Vieillard AS, et al. Genomic definition of hypervirulent and multidrug-resistant Klebsiella pneumoniae clonal groups. Emerg Infect Dis. 2014;20(11):1812–1820. doi: 10.3201/eid2011.140206. http://dx.doi.org/10.3201/eid2011.140206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Biedenbach D, Bouchillon S, Hackel M, Hoban D, Kazmierczak K, Hawser S, et al. Dissemination of NDM metallo-beta-lactamase genes among clinical isolates of Enterobacteriaceae collected during the SMART global surveillance study from 2008 to 2012. Antimicrob Agents Chemother. 2015a;59(2):826–830. doi: 10.1128/AAC.03938-14. http://dx.doi.org/10.1128/AAC.03938-14, Epub 2014/11/19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Biedenbach DJ, Kazmierczak K, Bouchillon SK, Sahm DF, Bradford PA. In vitro activity of aztreonam-avibactam against a global collection of gram-negative pathogens from 2012 and 2013. Antimicrob Agents Chemother. 2015b;59(7):4239–4248. doi: 10.1128/AAC.00206-15. http://dx.doi.org/10.1128/AAC.00206-15, Epub 2015/05/13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Bogaerts P, Yunus S, Massart M, Huang TD, Glupczynski Y. Evaluation of the BYG Carba test, a new electrochemical assay for rapid laboratory detection of carbapenemase-producing Enterobacteriaceae. J Clin Microbiol. 2016;54(2):349–358. doi: 10.1128/JCM.02404-15. http://dx.doi.org/10.1128/JCM.02404-15, Epub 2015/12/08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Bohme K, Fernandez-No IC, Barros-Velazquez J, Gallardo JM, Canas B, Calo-Mata P. SpectraBank: an open access tool for rapid microbial identification by MALDI-TOF MS fingerprinting. Electrophoresis. 2012;33(14):2138–2142. doi: 10.1002/elps.201200074. http://dx.doi.org/10.1002/elps.201200074, Epub 2012/07/24. [DOI] [PubMed] [Google Scholar]
  26. Bowers JR, Kitchel B, Driebe EM, MacCannell DR, Roe C, Lemmer D, et al. Genomic analysis of the emergence and rapid global dissemination of the clonal group 258 Klebsiella pneumoniae pandemic. PLoS One. 2015;10(7):e0133727. doi: 10.1371/journal.pone.0133727. http://dx.doi.org/10.1371/journal.pone.0133727, Epub 2015/07/22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Branas P, Villa J, Viedma E, Mingorance J, Orellana MA, Chaves F. Molecular epidemiology of carbapenemase-producing Klebsiella pneumoniae in a hospital in Madrid: successful establishment of an OXA-48 ST11 clone. Int J Antimicrob Agents. 2015;46(1):111–116. doi: 10.1016/j.ijantimicag.2015.02.019. http://dx.doi.org/10.1016/j.ijantimicag.2015.02.019, Epub 2015/04/29. [DOI] [PubMed] [Google Scholar]
  28. Bratu S, Landman D, Haag R, Recco R, Eramo A, Alam M, et al. Rapid spread of carbapenem-resistant Klebsiella pneumoniae in New York City: a new threat to our antibiotic armamentarium. Arch Intern Med. 2005;165(12):1430–1435. doi: 10.1001/archinte.165.12.1430. http://dx.doi.org/10.1001/archinte.165.12.1430, Epub 2005/06/29. [DOI] [PubMed] [Google Scholar]
  29. CLSI. 26th Informational Supplement M100 S26:2016. Clinical and Laboratory Standards Institute; 2016. Performance standards for antimicrobial susceptibility testing. [Google Scholar]
  30. Caballero S, Carter R, Ke X, Susac B, Leiner IM, Kim GJ, et al. Distinct but spatially overlapping intestinal niches for vancomycin-resistant Enterococcus faecium and carbapenem-resistant Klebsiella pneumoniae. PLoS Pathog. 2015;11(9):e1005132. doi: 10.1371/journal.ppat.1005132. http://dx.doi.org/10.1371/journal.ppat.1005132, Epub 2015/09/04. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Camargo JF, Simkins J, Beduschi T, Tekin A, Aragon L, Perez-Cardona A, et al. Successful Treatment of carbapenemase-producing pandrug-resistant Klebsiella pneumoniae bacteremia. Antimicrob Agents Chemother. 2015;59(10):5903–5908. doi: 10.1128/AAC.00655-15. http://dx.doi.org/10.1128/AAC.00655-15, Epub 2015/09/20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Campos JC, da Silva MJ, dos Santos PR, Barros EM, Pereira Mde O, Seco BM, et al. Characterization of Tn3000, a transposon responsible for blaNDM-1 dissemination among Enterobacteriaceae in Brazil, Nepal, Morocco, and India. Antimicrob Agents Chemother. 2015;59(12):7387–7395. doi: 10.1128/AAC.01458-15. http://dx.doi.org/10.1128/AAC.01458-15, Epub 2015/09/24. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Capone A, Giannella M, Fortini D, Giordano A, Meledandri M, Ballardini M, et al. High rate of colistin resistance among patients with carbapenem-resistant Klebsiella pneumoniae infection accounts for an excess of mortality. Clin Microbiol Infect. 2013;19(1):E23–E30. doi: 10.1111/1469-0691.12070. http://dx.doi.org/10.1111/1469-0691.12070, Epub 2012/11/10. [DOI] [PubMed] [Google Scholar]
  34. Carattoli A, Seiffert SN, Schwendener S, Perreten V, Endimiani A. Differentiation of IncL and IncM plasmids associated with the spread of clinically relevant antimicrobial resistance. PLoS One. 2015;10(5):e0123063. doi: 10.1371/journal.pone.0123063. http://dx.doi.org/10.1371/journal.pone.0123063, Epub 2015/05/02. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Carattoli A. Plasmids and the spread of resistance. Int J Med Microbiol. 2013;303(6–7):298–304. doi: 10.1016/j.ijmm.2013.02.001. http://dx.doi.org/10.1016/j.ijmm.2013.02.001, Epub 2013/03/19. [DOI] [PubMed] [Google Scholar]
  36. Castanheira M, Mills JC, Costello SE, Jones RN, Sader HS. Ceftazidime-avibactam activity tested against Enterobacteriaceae isolates from U.S. hospitals (2011 to 2013) and characterization of beta-lactamase-producing strains. Antimicrob Agents Chemother. 2015;59(6):3509–3517. doi: 10.1128/AAC.00163-15. http://dx.doi.org/10.1128/AAC.00163-15, Epub 2015/04/08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Castanheira M, Deshpande LM, Mills JC, Jones RN, Soave R, Jenkins SG, et al. Klebsiella pneumoniae isolate from a New York City Hospital belonging to sequence type 258 and carrying blaKPC-2 and blaVIM-4. Antimicrob. Agents Chemother. 2016;60(3):1924–1927. doi: 10.1128/AAC.01844-15. http://dx.doi.org/10.1128/AAC.01844-15, Epub 2016/01/06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Cejas D, Fernandez Canigia L, Rincon Cruz G, Elena AX, Maldonado I, Gutkind GO, et al. First isolate of KPC-2-producing Klebsiella pneumonaie sequence type 23 from the Americas. J Clin Microbiol. 2014;52(9):3483–3485. doi: 10.1128/JCM.00726-14. http://dx.doi.org/10.1128/JCM.00726-14, Epub 2014/07/18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Chen L, Mathema B, Pitout JD, DeLeo FR, Kreiswirth BN. Epidemic Klebsiella pneumoniae ST258 is a hybrid strain. mBio. 2014;5(3):e01355–14. doi: 10.1128/mBio.01355-14. http://dx.doi.org/10.1128/mBio.01355-14. Epub 2014/06/26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Chen L, Peirano G, Lynch T, Chavda KD, Gregson DB, Church DL, et al. Molecular characterization by using next-generation sequencing of plasmids containing blaNDM-7 in Enterobacteriaceae from Calgary, Canada. Antimicrob Agents Chemother. 2015a;60(3):1258–1263. doi: 10.1128/AAC.02661-15. http://dx.doi.org/10.1128/AAC.02661-15, Epub 2015/12/09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Chen Z, Li H, Feng J, Li Y, Chen X, Guo X, et al. NDM-1 encoded by a pNDM-BJ01-like plasmid p3SP-NDM in clinical Enterobacter aerogenes. Front Microbiol. 2015b;6:294. doi: 10.3389/fmicb.2015.00294. http://dx.doi.org/10.3389/fmicb.2015.00294, Epub 2015/05/01. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Cheng AA, Ding H, Lu TK. Enhanced killing of antibiotic-resistant bacteria enabled by massively parallel combinatorial genetics. Proc Natl Acad Sci U S A. 2014;111(34):12462–12467. doi: 10.1073/pnas.1400093111. http://dx.doi.org/10.1073/pnas.1400093111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Cheng L, Cao XL, Zhang ZF, Ning MZ, Xu XJ, Zhou W, et al. Clonal dissemination of KPC-2 producing Klebsiella pneumoniae ST11 clone with high prevalence of oqxAB and rmtB in a tertiary hospital in China: results from a 3- year period. Ann Clin Microbiol Antimicrob. 2016;15:1. doi: 10.1186/s12941-015-0109-x. http://dx.doi.org/10.1186/s12941-015-0109-x, Epub 2016/01/21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Cho SY, Huh HJ, Baek JY, Chung NY, Ryu JG, Ki CS, et al. Kleb-siella pneumoniae co-producing NDM-5 and OXA-181 carbapenemases, South Korea. Emerg Infect Dis. 2015;21(6):1088–1089. doi: 10.3201/eid2106.150048. http://dx.doi.org/10.3201/eid2106.150048, Epub 2015/05/20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Chung The H, Karkey A, Pham Thanh D, Boinett CJ, Cain AK, Ellington M, et al. A high-resolution genomic analysis of multidrug-resistant hospital outbreaks of Klebsiella pneumoniae. EMBO Mol Med. 2015;7(3):227–239. doi: 10.15252/emmm.201404767. http://dx.doi.org/10.15252/emmm.201404767, Epub 2015/02/26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Citorik RJ, Mimee M, Lu TK. Sequence-specific antimicrobials using efficiently delivered RNA-guided nucleases. Nat Biotechnol. 2014;32(11):1141–1145. doi: 10.1038/nbt.3011. http://dx.doi.org/10.1038/nbt.3011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Cprek JB, Gallagher JC. Ertapenem-containing double-carbapenem therapy for treatment of infections caused by carbapenem-resistant Klebsiella pneumoniae. Antimicrob Agents Chemother. 2016;60(1):669–673. doi: 10.1128/AAC.01569-15. http://dx.doi.org/10.1128/AAC.01569-15, Epub 2015/11/11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Dang B, Mao D, Luo Y. Complete nucleotide sequence of IncP-1beta plasmid pDTC28 reveals a non-functional variant of the blaGES-type gene. PLoS One. 2016a;11(5):e0154975. doi: 10.1371/journal.pone.0154975. http://dx.doi.org/10.1371/journal.pone.0154975, Epub 2016/05/07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Dang B, Mao D, Luo Y. Complete nucleotide sequence of pGA45, a 140,698-bp IncFIIY plasmid encoding bla IMI-3-mediated carbapenem resistance, from river sediment. Front Microbiol. 2016b;7:188. doi: 10.3389/fmicb.2016.00188. http://dx.doi.org/10.3389/fmicb.2016.00188, Epub 2016/03/05. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. de Araujo CF, Silva DM, Carneiro MT, Ribeiro S, Fontana-Maurell M, Alvarez P, et al. Detection of carbapenemase genes in aquatic environments in Rio de Janeiro, Brazil. Antimicrob Agents Chemother. 2016;60(7):4380–4383. doi: 10.1128/AAC.02753-15. http://dx.doi.org/10.1128/AAC.02753-15, Epub 2016/05/04. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Del Franco M, Paone L, Novati R, Giacomazzi CG, Bagattini M, Galotto C, et al. Molecular epidemiology of carbapenem resistant Enterobacteriaceae in Valle d’Aosta region, Italy, shows the emergence of KPC-2 producing Klebsiella pneumoniae clonal complex 101 (ST101 and ST1789) BMC Microbiol. 2015;15(1):260. doi: 10.1186/s12866-015-0597-z. http://dx.doi.org/10.1186/s12866-015-0597-z, Epub 2015/11/11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Deleo FR, Chen L, Porcella SF, Martens CA, Kobayashi SD, Porter AR, et al. Molecular dissection of the evolution of carbapenem-resistant multilocus sequence type 258 Klebsiella pneumoniae. Proc Natl Acad Sci U S A. 2014;111(13):4988–4993. doi: 10.1073/pnas.1321364111. http://dx.doi.org/10.1073/pnas.1321364111, Epub 2014/03/19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Demiray T, Koroglu M, Ozbek A, Altindis M. A rare cause of infection, Raoultella planticola: emerging threat and new reservoir for carbapenem resistance. Infection. 2016 doi: 10.1007/s15010-016-0900-4. http://dx.doi.org/10.1007/s15010-016-0900-4, Epub 2016/05/06. [DOI] [PubMed]
  54. DiGiandomenico A, Sellman BR. Antibacterial monoclonal antibodies: the next generation? Curr Opin Microbiol. 2015;27:78–85. doi: 10.1016/j.mib.2015.07.014. http://dx.doi.org/10.1016/j.mib.2015.07.014, Epub 2015/08/25. [DOI] [PubMed] [Google Scholar]
  55. Doern CD. When does 2 plus 2 equal 5? A review of antimicrobial synergy testing. J Clin Microbiol. 2014;52(12):4124–4128. doi: 10.1128/JCM.01121-14. http://dx.doi.org/10.1128/JCM.01121-14, Epub 2014/06/13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Dolejska M, Masarikova M, Dobiasova H, Jamborova I, Karpiskova R, Havlicek M, et al. High prevalence of Salmonella and IMP-4-producing Enterobacteriaceae in the silver gull on Five Islands, Australia. J Antimicrob Chemother. 2016;71(1):63–70. doi: 10.1093/jac/dkv306. http://dx.doi.org/10.1093/jac/dkv306, Epub 2015/10/17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Donald HM, Scaife W, Amyes SG, Young HK. Sequence analysis of ARI-1, a novel OXA beta-lactamase, responsible for imipenem resistance in Acinetobacter baumannii 6B92. Antimicrob Agents Chemother. 2000;44(1):196–199. doi: 10.1128/aac.44.1.196-199.2000. Epub 1999/12/22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Dortet L, Oueslati S, Jeannot K, Tande D, Naas T, Nordmann P. Genetic and biochemical characterization of OXA-405, an OXA-48-type extended-spectrum beta-lactamase without significant carbapenemase activity. Antimicrob Agents Chemother. 2015a;59(7):3823–3828. doi: 10.1128/AAC.05058-14. http://dx.doi.org/10.1128/AAC.05058-14, Epub 2015/04/15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Dortet L, Poirel L, Abbas S, Oueslati S, Nordmann P. Genetic and biochemical characterization of FRI-1, a carbapenem-hydrolyzing class a beta-lactamase from Enterobacter cloacae. Antimicrob Agents Chemother. 2015b;59(12):7420–7425. doi: 10.1128/AAC.01636-15. http://dx.doi.org/10.1128/AAC.01636-15, Epub 2015/09/24. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Dortet L, Jousset A, Sainte-Rose V, Cuzon G, Naas T. Prospective evaluation of the OXA-48 K-SeT assay, an immunochromatographic test for the rapid detection of OXA-48-type carbapenemases. J Antimicrob Chemother. 2016;71:1834–1840. doi: 10.1093/jac/dkw058. http://dx.doi.org/10.1093/jac/dkw058, Epub 2016/03/13. [DOI] [PubMed] [Google Scholar]
  61. Drawz SM, Bonomo RA. Three decades of beta-lactamase inhibitors. Clin Microbiol Rev. 2010;23(1):160–201. doi: 10.1128/CMR.00037-09. http://dx.doi.org/10.1128/CMR.00037-09, Epub 2010/01/13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Du H, Chen L, Chavda KD, Pandey R, Zhang H, Xie X, et al. Genomic characterization of Enterobacter cloacae isolates from China that coproduce KPC-3 and NDM-1 carbapenemases. Antimicrob Agents Chemother. 2016a;60(4):2519–2523. doi: 10.1128/AAC.03053-15. http://dx.doi.org/10.1128/AAC.03053-15, Epub 2016/01/21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Du H, Chen L, Tang YW, Kreiswirth BN. Emergence of the mcr-1 colistin resistance gene in carbapenem-resistant Enterobacteriaceae. Lancet Infect Dis. 2016b;16(3):287–288. doi: 10.1016/S1473-3099(16)00056-6. http://dx.doi.org/10.1016/S1473-3099(16)00056-6, Epub 2016/02/05. [DOI] [PubMed] [Google Scholar]
  64. Dwyer DJ, Belenky PA, Yang JH, MacDonald IC, Martell JD, Takahashi N, et al. Antibiotics induce redox-related physiological alterations as part of their lethality. Proc Natl Acad Sci U S A. 2014;111(20):E2100–E2109. doi: 10.1073/pnas.1401876111. http://dx.doi.org/10.1073/pnas.1401876111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. EUCAST. European Committee on Antimicrobial Susceptibility Testing; 2013. EUCAST guidelines for detection of resistance mechanisms and specific resistances of clinical and/or epidemiological importance. http://www.eucast.org/resistance_mechanisms/ [Google Scholar]
  66. Evans BA, Amyes SG. OXA beta-lactamases. Clin Microbiol Rev. 2014;27(2):241–263. doi: 10.1128/CMR.00117-13. http://dx.doi.org/10.1128/CMR.00117-13, Epub 2014/04/04. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Fattouh R, Tijet N, McGeer A, Poutanen SM, Melano RG, Patel SN. What is the appropriate meropenem MIC for screening of carbapenemase-producing Enterobacteriaceae in low-prevalence settings? Antimicrob Agents Chemother. 2016;60(3):1556–1559. doi: 10.1128/AAC.02304-15. http://dx.doi.org/10.1128/AAC.02304-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Feng J, Qiu Y, Yin Z, Chen W, Yang H, Yang W, et al. Coexistence of a novel KPC-2-encoding MDR plasmid and an NDM-1-encoding pNDM-HN380-like plasmid in a clinical isolate of Citrobacter freundii. J Antimicrob Chemother. 2015a;70(11):2987–2991. doi: 10.1093/jac/dkv232. http://dx.doi.org/10.1093/jac/dkv232, Epub 2015/08/12. [DOI] [PubMed] [Google Scholar]
  69. Feng Y, Yang P, Xie Y, Wang X, McNally A, Zong Z. Escherichia coli of sequence type 3835 carrying bla NDM-1, bla CTX-M-15, bla CMY-42 and bla SHV-12. Sci Rep. 2015b;5:12275. doi: 10.1038/srep12275. http://dx.doi.org/10.1038/srep12275, Epub 2015/07/22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Fiett J, Baraniak A, Izdebski R, Sitkiewicz I, Zabicka D, Meler A, et al. The first NDM metallo-beta-lactamase-producing Enterobacteriaceae isolate in Poland: evolution of IncFII-type plasmids carrying the bla(NDM-1) gene. Antimicrob Agents Chemother. 2014;58(2):1203–1207. doi: 10.1128/AAC.01197-13. http://dx.doi.org/10.1128/AAC.01197-13, Epub 2013/11/20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Fortini D, Villa L, Feudi C, Pires J, Bonura C, Mammina C, et al. Double copies of bla(KPC-3)::Tn4401a on an IncX3 Plasmid in Klebsiella pneumoniae successful clone ST512 from Italy. Antimicrob Agents Chemother. 2016;60(1):646–649. doi: 10.1128/AAC.01886-15. http://dx.doi.org/10.1128/AAC.01886-15, Epub 2015/11/04. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Fursova NK, Astashkin EI, Knyazeva AI, Kartsev NN, Leonova ES, Ershova ON, et al. The spread of bla OXA-48 and bla OXA-244 carbapenemase genes among Klebsiella pneumoniae, Proteus mirabilis and Enterobacter spp. isolated in Moscow, Russia. Ann Clin Microbiol Antimicrob. 2015;14:46. doi: 10.1186/s12941-015-0108-y. http://dx.doi.org/10.1186/s12941-015-0108-y, Epub 2015/11/04. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Gagetti P, Pasteran F, Martinez MP, Fatouraei M, Gu J, Fernandez R, et al. Modeling meropenem treatment, alone and in combination with daptomycin, for KPC-producing Klebsiella pneumoniae strains with unusually low carbapenem MICs. Antimicrob Agents Chemother. 2016;60(8):5047–5050. doi: 10.1128/AAC.00168-16. http://dx.doi.org/10.1128/AAC.00168-16, Epub 2016/05/25. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Gaiarsa S, Comandatore F, Gaibani P, Corbella M, Dalla Valle C, Epis S, et al. Genomic epidemiology of Klebsiella pneumoniae in Italy and novel insights into the origin and global evolution of its resistance to carbapenem antibiotics. Antimicrob Agents Chemother. 2015;59(1):389–396. doi: 10.1128/AAC.04224-14. http://dx.doi.org/10.1128/AAC.04224-14, Epub 2014/11/05. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Gamal D, Fernandez-Martinez M, El-Defrawy I, Ocampo-Sosa AA, Martinez-Martinez L. First identification of NDM-5 associated with OXA-181 in Escherichia coli from Egypt. Emerg Microb Infect. 2016a;5:e30. doi: 10.1038/emi.2016.24. http://dx.doi.org/10.1038/emi.2016.24, Epub 2016/04/07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Gamal D, Fernandez-Martinez M, Salem D, El-Defrawy I, Montes LA, Ocampo-Sosa AA, et al. Carbapenem-resistant Klebsiella pneumoniae isolates from Egypt containing blaNDM-1 on IncR plasmids and its association with rmtF. Int J Infect Dis. 2016b;43:17–20. doi: 10.1016/j.ijid.2015.12.003. http://dx.doi.org/10.1016/j.ijid.2015.12.003, Eub 2015/12/22. [DOI] [PubMed] [Google Scholar]
  77. Girlich D, Poirel L, Nordmann P. Value of the modified Hodge test for detection of emerging carbapenemases in Enterobacteriaceae. J Clin Microbiol. 2012;50(2):477–479. doi: 10.1128/JCM.05247-11. http://dx.doi.org/10.1128/JCM.05247-11, Epub 2011/11/26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Gonzales PR, Pesesky MW, Bouley R, Ballard A, Biddy BA, Suckow MA, et al. Synergistic, collaterally sensitive beta-lactam combinations suppress resistance in MRSA. Nat Chem Biol. 2015;11(11):855–861. doi: 10.1038/nchembio.1911. http://dx.doi.org/10.1038/nchembio.1911. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Gonzalez LJ, Bahr G, Nakashige TG, Nolan EM, Bonomo RA, Vila AJ. Membrane anchoring stabilizes and favors secretion of New Delhi metallo- beta-lactamase. Nat Chem Biol. 2016;12(7):516–522. doi: 10.1038/nchembio.2083. http://dx.doi.org/10.1038/nchembio.2083, Epub 2016/05/18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Gonzalez-Torralba A, Oteo J, Asenjo A, Bautista V, Fuentes E, Alos JI. Survey of carbapenemase-producing Enterobacteriaceae in companion dogs in Madrid, Spain. Antimicrob Agents Chemother. 2016;60(4):2499–2501. doi: 10.1128/AAC.02383-15. http://dx.doi.org/10.1128/AAC.02383-15, Epub 2016/01/30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Guerra B, Fischer J, Helmuth R. An emerging public health problem: acquired carbapenemase-producing microorganisms are present in food-producing animals, their environment, companion animals and wild birds. Vet Microbiol. 2014;171(3–4):290–297. doi: 10.1016/j.vetmic.2014.02.001. http://dx.doi.org/10.1016/j.vetmic.2014.02.001, Epub 2014/03/19. [DOI] [PubMed] [Google Scholar]
  82. Harmer CJ, Hall RM. The A to Z of A/C plasmids. Plasmid. 2015;80:63–82. doi: 10.1016/j.plasmid.2015.04.003. http://dx.doi.org/10.1016/j.plasmid.2015.04.003, Epub 2015/04/26. [DOI] [PubMed] [Google Scholar]
  83. Haverkate MR, Dautzenberg MJ, Ossewaarde TJ, van der Zee A, den Hollander JG, Troelstra A, et al. Within-host and population: transmission of blaOXA-48 in K. pneumoniae and E. coli. PLoS One. 2015;10(10):e0140960. doi: 10.1371/journal.pone.0140960. http://dx.doi.org/10.1371/journal.pone.0140960, Epub 2015/10/21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Hemarajata P, Yang S, Hindler JA, Humphries RM. Development of a novel real-time PCR assay with high-resolution melt analysis to detect and differentiate OXA-48-like beta-lactamases in carbapenem-resistant Enterobacteriaceae. Antimicrob Agents Chemother. 2015;59(9):5574–5580. doi: 10.1128/AAC.00425-15. http://dx.doi.org/10.1128/AAC.00425-15, Epub 2015/07/01. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Henriques I, Moura A, Alves A, Saavedra MJ, Correia A. Molecular characterization of a carbapenem-hydrolyzing class A beta-lactamase, SFC-1, from Serratia fonticola UTAD54. Antimicrob Agents Chemother. 2004;48(6):2321–2324. doi: 10.1128/AAC.48.6.2321-2324.2004. http://dx.doi.org/10.1128/AAC.48.6.2321-2324.2004, Epub 2004/05/25. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Hidalgo L, Hopkins KL, Gutierrez B, Ovejero CM, Shukla S, Douthwaite S, et al. Association of the novel aminoglycoside resistance determinant RmtF with NDM carbapenemase in Enterobacteriaceae isolated in India and the UK. J Antimicrob Chemother. 2013;68(7):1543–1550. doi: 10.1093/jac/dkt078. http://dx.doi.org/10.1093/jac/dkt078, Epub 2013/04/13. [DOI] [PubMed] [Google Scholar]
  87. Higgins PG, Zander E, Seifert H. Identification of a novel insertion sequence element associated with carbapenem resistance and the development of fluoroquinolone resistance in Acinetobacter radioresistens. J Antimicrob Chemother. 2013;68(3):720–722. doi: 10.1093/jac/dks446. http://dx.doi.org/10.1093/jac/dks446, Epub 2012/11/10. [DOI] [PubMed] [Google Scholar]
  88. Hossain A, Ferraro MJ, Pino RM, Dew RB, Jr, Moland ES, Lockhart TJ, et al. Plasmid-mediated carbapenem-hydrolyzing enzyme KPC-2 in an Enterobacter sp. Antimicrob Agents Chemother. 2004;48(11):4438–4440. doi: 10.1128/AAC.48.11.4438-4440.2004. http://dx.doi.org/10.1128/AAC.48.11.4438-4440.2004, Epub 2004/10/27. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Huang TD, Poirel L, Bogaerts P, Berhin C, Nordmann P, Glupczynski Y. Temocillin and piperacillin/tazobactam resistance by disc diffusion as antimicrobial surrogate markers for the detection of carbapenemase-producing Enterobacteriaceae in geographical areas with a high prevalence of OXA-48 producers. J Antimicrob Chemother. 2014;69(2):445–550. doi: 10.1093/jac/dkt367. http://dx.doi.org/10.1093/jac/dkt367, Epub 2013/09/24. [DOI] [PubMed] [Google Scholar]
  90. Humphries RM, Yang S, Hemarajata P, Ward KW, Hindler JA, Miller SA, et al. First report of ceftazidime-avibactam resistance in a KPC-3- expressing Klebsiella pneumoniae isolate. Antimicrob Agents Chemother. 2015;59(10):6605–6607. doi: 10.1128/AAC.01165-15. http://dx.doi.org/10.1128/AAC.01165-15, Epub 2015/07/22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Irfan S, Khan E, Jabeen K, Bhawan P, Hopkins KL, Day M, et al. Clinical isolates of Salmonella enterica serovar agona producing NDM-1 metallo-beta- lactamase: first report from Pakistan. J Clin Microbiol. 2015;53(1):346–348. doi: 10.1128/JCM.02396-14. http://dx.doi.org/10.1128/JCM.02396-14, Epub 2014/11/08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Ito H, Arakawa Y, Ohsuka S, Wacharotayankun R, Kato N, Ohta M. Plasmid-mediated dissemination of the metallo-beta-lactamase gene blaIMP among clinically isolated strains of Serratia marcescens. Antimicrob Agents Chemother. 1995;39(4):824–829. doi: 10.1128/aac.39.4.824. Epub 1995/04/01. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Jeong S, Kim JO, Jeong SH, Bae IK, Song W. Evaluation of peptide nucleic acid-mediated multiplex real-time PCR kits for rapid detection of carbapenemase genes in gram-negative clinical isolates. J Microbiol Methods. 2015;113:4–9. doi: 10.1016/j.mimet.2015.03.019. http://dx.doi.org/10.1016/j.mimet.2015.03.019, Epub 2015/03/31. [DOI] [PubMed] [Google Scholar]
  94. Jimenez-Castellanos JC, Wan Ahmad Kamil WN, Cheung CH, Tobin MS, Brown J, Isaac SG, et al. Comparative effects of overproducing the AraC-type transcriptional regulators MarA, SoxS, RarA and RamA on antimicrobial drug susceptibility in Klebsiella pneumoniae. J Antimicrob Chemother. 2016;71(7):1820–1825. doi: 10.1093/jac/dkw088. http://dx.doi.org/10.1093/jac/dkw088, Eub 2016/04/01. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Kabir MH, Meunier D, Hopkins KL, Giske CG, Woodford N. A two-centre evaluation of RAPIDEC(R) CARBA NP for carbapenemase detection in Enterobacteriaceae, Pseudomonas aeruginosa and Acinetobacter spp. J Antimicrob Chemother. 2016;71(5):1213–1216. doi: 10.1093/jac/dkv468. http://dx.doi.org/10.1093/jac/dkv468, Epub 2016/01/16. [DOI] [PubMed] [Google Scholar]
  96. Kang HY, Kim KY, Kim J, Lee JC, Lee YC, Cho DT, et al. Distribution of conjugative-plasmid-mediated 16S rRNA methylase genes among amikacin-resistant Enterobacteriaceae isolates collected in 1995 to 1998 and 2001 to 2006 at a university hospital in South Korea and identification of conjugative plasmids mediating dissemination of 16S rRNA methylase. J Clin Microbiol. 2008;46(2):700–706. doi: 10.1128/JCM.01677-07. http://dx.doi.org/10.1128/jcm.01677-07, Eub 2007/12/21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Kazmierczak KM, Rabine S, Hackel M, McLaughlin RE, Biedenbach DJ, Bouchillon SK, et al. Multiyear, multinational survey of the incidence and global distribution of metallo-beta-lactamase-producing Enterobacteriaceae and Pseudomonas aeruginosa. Antimicrob Agents Chemother. 2015;60(2):1067–1078. doi: 10.1128/AAC.02379-15. http://dx.doi.org/10.1128/AAC.02379-15, Epub 2015/12/09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Kieffer N, Poirel L, Bessa LJ, Barbosa-Vasconcelos A, da Costa PM, Nordmann P. VIM-1, VIM-34, and IMP-8 carbapenemase-producing Escherichia coli strains recovered from a Portuguese River. Antimicrob Agents Chemother. 2016;60(4):2585–2586. doi: 10.1128/AAC.02632-15. http://dx.doi.org/10.1128/AAC.02632-15, Epub 2016/01/27. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Kohira N, West J, Ito A, Ito-Horiyama T, Nakamura R, Sato T, et al. In vitro antimicrobial activity of a siderophore cephalosporin, S-649266, against Enterobacteriaceae clinical isolates, including carbapenem-resistant strains. Antimicrob Agents Chemother. 2015;60(2):729–734. doi: 10.1128/AAC.01695-15. http://dx.doi.org/10.1128/AAC.01695-15, Epub 2015/11/18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Krishnan NP, Nguyen NQ, Papp-Wallace KM, Bonomo RA, van den Akker F. Inhibition of Klebsiella beta-lactamases (SHV-1 and KPC-2) by avibactam: a structural study. PLoS One. 2015;10(9):e0136813. doi: 10.1371/journal.pone.0136813. http://dx.doi.org/10.1371/journal.pone.0136813, Epub 2015/09/05. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Ktari S, Mnif B, Louati F, Rekik S, Mezghani S, Mahjoubi F, et al. Spread of Klebsiella pneumoniae isolates producing OXA-48 beta-lactamase in a Tunisian university hospital. J Antimicrob Chemother. 2011;66(7):1644–1646. doi: 10.1093/jac/dkr181. http://dx.doi.org/10.1093/jac/dkr181, Epub 2011/05/14. [DOI] [PubMed] [Google Scholar]
  102. Lasserre C, De Saint Martin L, Cuzon G, Bogaerts P, Lamar E, Glupczynski Y, et al. Efficient detection of carbapenemase activity in Enterobacteriaceae by matrix-assisted laser desorption ionization-time of flight mass spectrometry in less than 30 minutes. J Clin Microbiol. 2015;53(7):2163–2171. doi: 10.1128/JCM.03467-14. http://dx.doi.org/10.1128/JCM.03467-14, Epub 2015/05/01. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Lerner A, Solter E, Rachi E, Adler A, Rechnitzer H, Miron D, et al. Detection and characterization of carbapenemase-producing Enterobacteriaceae in wounded Syrian patients admitted to hospitals in northern Israel. Eur J Clin Microbiol Infect Dis. 2016;35(1):149–154. doi: 10.1007/s10096-015-2520-9. http://dx.doi.org/10.1007/s10096-015-2520-9, Epub 2015/11/20. [DOI] [PubMed] [Google Scholar]
  104. Li W, Tailhades J, O’Brien-Simpson NM, Separovic F, Otvos L, Jr, Hossain MA, et al. Proline-rich antimicrobial peptides: potential therapeutics against antibiotic-resistant bacteria. Amino Acids. 2014;46(10):2287–2294. doi: 10.1007/s00726-014-1820-1. http://dx.doi.org/10.1007/s00726-014-1820-1, Epub 2014/08/22. [DOI] [PubMed] [Google Scholar]
  105. Li JJ, Munoz-Price LS, Spychala CN, DePascale D, Doi Y. New Delhi metallo-beta-lactamase-1-producing Klebsiella pneumoniae, Florida, USA. Emerg Infect Dis. 2016;22(4):744–746. doi: 10.3201/eid2204.151176. http://dx.doi.org/10.3201/eid2204.151176, Epub 2016/03/18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Liu YY, Wang Y, Walsh TR, Yi LX, Zhang R, Spencer J, et al. Emergence of plasmid-mediated colistin resistance mechanism MCR-1 in animals and human beings in China: a microbiological and molecular biological study. Lancet Infect Dis. 2016;16(2):161–168. doi: 10.1016/S1473-3099(15)00424-7. http://dx.doi.org/10.1016/S1473-3099(15)00424-7, Epub 2015/11/26. [DOI] [PubMed] [Google Scholar]
  107. Lutgring JD, Limbago BM. The problem of carbapenemase-producing-carbapenem-resistant-Enterobacteriaceae detection. J Clin Microbiol. 2016;54(3):529–534. doi: 10.1128/JCM.02771-15. http://dx.doi.org/10.1128/JCM.02771-15, Epub 2016/01/08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Ma L, Wang JT, Wu TL, Siu LK, Chuang YC, Lin JC, et al. Emergence of OXA-48-producing Klebsiella pneumoniae in Taiwan. PLoS One. 2015;10(9):e0139152. doi: 10.1371/journal.pone.0139152. http://dx.doi.org/10.1371/journal.pone.0139152, Epub 2015/09/29. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Manageiro V, Pinto M, Canica M. Complete sequence of a blaOXA- 48-harboring IncL plasmid from an Enterobacter cloacae clinical isolate. Genome Announc. 2015;3(5) doi: 10.1128/genomeA.01076-15. http://dx.doi.org/10.1128/genomeA.01076-15, Epub 2015/09/19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Marsh JW, Krauland MG, Nelson JS, Schlackman JL, Brooks AM, Pasculle AW, et al. Genomic epidemiology of an endoscope-associated outbreak of Klebsiella pneumoniae carbapenemase (KPC)-producing K. pneumoniae. PLoS One. 2015;10(12):e0144310. doi: 10.1371/journal.pone.0144310. http://dx.doi.org/10.1371/journal.pone.0144310, Epub 2015/12/05. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Mataseje LF, Peirano G, Church DL, Conly J, Mulvey M, Pitout JD. Colistin-nonsusceptible Pseudomonas aeruginosa sequence type 654 with blaNDM-1 arrives in North America. Antimicrob Agents Chemother. 2016;60(3):1794–1800. doi: 10.1128/AAC.02591-15. http://dx.doi.org/10.1128/AAC.02591-15, Epub 2016/01/30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Mathers AJ, Peirano G, Pitout JD. The role of epidemic resistance plasmids and international high-risk clones in the spread of multidrug-resistant Enterobacteriaceae. Clin Microbiol Rev. 2015;28(3):565–591. doi: 10.1128/CMR.00116-14. http://dx.doi.org/10.1128/CMR.00116-14, Epub 2015/05/01. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Meletis G. Carbapenem resistance: overview of the problem and future perspectives. Ther Adv Infect Dis. 2016;3(1):15–21. doi: 10.1177/2049936115621709. http://dx.doi.org/10.1177/2049936115621709, Epub 2016/02/11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Miller BM, Johnson SW. Demographic and infection characteristics of patients with carbapenem-resistant Enterobacteriaceae in a community hospital: development of a bedside clinical score for risk assessment. Am J Infect Control. 2016;44(2):134–137. doi: 10.1016/j.ajic.2015.09.006. http://dx.doi.org/10.1016/j.ajic.2015.09.006, Epub 2015/10/24. [DOI] [PubMed] [Google Scholar]
  115. Miriagou V, Tzouvelekis LS, Rossiter S, Tzelepi E, Angulo FJ, Whichard JM. Imipenem resistance in a Salmonella clinical strain due to plasmid-mediated class A carbapenemase KPC-2. Antimicrob Agents Chemother. 2003;47(4):1297–1300. doi: 10.1128/AAC.47.4.1297-1300.2003. Epub 2003/03/26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Morrill HJ, Pogue JM, Kaye KS, LaPlante KL. Treatment options for carbapenem-resistant Enterobacteriaceae infections. Open Forum Infect Dis. 2015:fv050. doi: 10.1093/ofid/ofv050. http://dx.doi.org/10.1093/ofid/ofv050, Epub 2015/07/01. [DOI] [PMC free article] [PubMed]
  117. NCBI. Beta-Lactamase Database [Internet] National Center for Biotechnology Information; 2016. [Google Scholar]
  118. Nicolau DP. Carbapenems: a potent class of antibiotics. Expert Opin Pharmacother. 2008;9(1):23–37. doi: 10.1517/14656566.9.1.23. http://dx.doi.org/10.1517/14656566.9.1.23, Epub 2007/12/14. [DOI] [PubMed] [Google Scholar]
  119. Nicoletti AG, Marcondes MF, Martins WM, Almeida LG, Nicolas MF, Vasconcelos AT, et al. Characterization of BKC-1 class A carbapenemase from Klebsiella pneumoniae clinical isolates in Brazil. Antimicrob Agents Chemother. 2015;59(9):5159–5164. doi: 10.1128/AAC.00158-15. http://dx.doi.org/10.1128/AAC.00158-15, Epub 2015/06/10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Nordmann P, Naas T, Poirel L. Global spread of varbapenemase-producing Enterobacteriaceae. Emerg Infect Dis. 2011;17(10):1791–1798. doi: 10.3201/eid1710.110655. http://dx.doi.org/10.3201/eid1710.110655, Epub 2011/10/18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Nordmann P, Gniadkowski M, Giske CG, Poirel L, Woodford N, Miriagou V, et al. Identification and screening of carbapenemase-producing Enter-obacteriaceae. Clin Microbiol Infect. 2012;18(5):432–438. doi: 10.1111/j.1469-0691.2012.03815.x. http://dx.doi.org/10.1111/j.1469-0691.2012.03815.x, Epub 2012/04/18. [DOI] [PubMed] [Google Scholar]
  122. Norgan AP, Freese JM, Tuin PM, Cunningham SA, Jeraldo PR, Patel R. Carbapenem- and colistin-resistant Enterobacter cloacae from Delta, Colorado, in 2015. Antimicrob Agents Chemother. 2016;60(5):3141–3144. doi: 10.1128/AAC.03055-15. http://dx.doi.org/10.1128/AAC.03055-15, Epub 2016/02/18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. O’Neill J. Tackling drug-resistant infections globally: final report and recommendations. Rev Antimicrob Resist 2016 [Google Scholar]
  124. Ocampo AM, Chen L, Cienfuegos AV, Roncancio G, Chavda KD, Kreiswirth BN, et al. A two-year surveillance in five colombian tertiary care hospitals reveals high frequency of non-CG258 clones of carbapenem-resistant Klebsiella pneumoniae with distinct clinical characteristics. Antimicrob Agents Chemother. 2016;60(1):332–342. doi: 10.1128/AAC.01775-15. http://dx.doi.org/10.1128/AAC.01775-15, Epub 2015/10/28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Oliva A, Mascellino MT, Cipolla A, D’Abramo A, De Rosa A, Savinelli S, et al. Therapeutic strategy for pandrug-resistant Klebsiella pneumoniae severe infections: short-course treatment with colistin increases the in vivo and in vitro activity of double carbapenem regimen. Int J Infect Dis. 2015;33:132–134. doi: 10.1016/j.ijid.2015.01.011. http://dx.doi.org/10.1016/j.ijid.2015.01.011, Eub 2015/01/20. [DOI] [PubMed] [Google Scholar]
  126. Ortega A, Saez D, Bautista V, Fernandez-Romero S, Lara N, Aracil B, et al. Carbapenemase-producing Escherichia coli is becoming more prevalent in Spain mainly because of the polyclonal dissemination of OXA-48. J Antimicrob Chemother. 2016;71:2131–2138. doi: 10.1093/jac/dkw148. http://dx.doi.org/10.1093/jac/dkw148, Epub 2016/05/06. [DOI] [PubMed] [Google Scholar]
  127. Oteo J, Ortega A, Bartolome R, Bou G, Conejo C, Fernandez-Martinez M, et al. Prospective multicenter study of carbapenemase-producing Enterobacteriaceae from 83 hospitals in Spain reveals high in vitro susceptibility to colistin and meropenem. Antimicrob Agents Chemother. 2015;59(6):3406–3412. doi: 10.1128/AAC.00086-15. http://dx.doi.org/10.1128/AAC.00086-15, Epub 2015/04/01. [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Oueslati S, Nordmann P, Poirel L. Heterogeneous hydrolytic features for OXA-48-like beta-lactamases. J Antimicrob Chemother. 2015;70(4):1059–1063. doi: 10.1093/jac/dku524. http://dx.doi.org/10.1093/jac/dku524, Epub 2015/01/15. [DOI] [PubMed] [Google Scholar]
  129. Pal C, Papp B, Lazar V. Collateral sensitivity of antibiotic-resistant microbes. Trends Microbiol. 2015;23(7):401–407. doi: 10.1016/j.tim.2015.02.009. http://dx.doi.org/10.1016/j.tim.2015.02.009, Epub 2015/03/31. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Pan CY, Chen JC, Chen TL, Wu JL, Hui CF, Chen JY. Piscidin is highly active against carbapenem-resistant Acinetobacter baumannii and NDM-1-producing Klebsiella pneumonia in a systemic septicaemia infection mouse model. Mar Drugs. 2015;13(4):2287–2305. doi: 10.3390/md13042287. http://dx.doi.org/10.3390/md13042287, Epub 2015/04/16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Papagiannitsis CC, Dolejska M, Izdebski R, Dobiasova H, Studentova V, Esteves FJ, et al. Characterization of pKP-M1144, a novel ColE1-like plasmid encoding IMP-8, GES-5, and BEL-1 beta-lactamases, from a Klebsiella pneumoniae sequence type 252 isolate. Antimicrob Agents Chemother. 2015a;59(8):5065–5068. doi: 10.1128/AAC.00937-15. http://dx.doi.org/10.1128/AAC.00937-15, Epub 2015/06/03. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Papagiannitsis CC, Izdebski R, Baraniak A, Fiett J, Herda M, Hrabak J, et al. Survey of metallo-beta-lactamase-producing Enterobacteriaceae colonizing patients in European ICUs and rehabilitation units, 2008–11. J Antimicrob Chemother. 2015b;70(7):1981–1988. doi: 10.1093/jac/dkv055. http://dx.doi.org/10.1093/jac/dkv055, Epub 2015/03/12. [DOI] [PubMed] [Google Scholar]
  133. Papagiannitsis CC, Pollini S, De Luca F, Rossolini GM, Docquier JD, Hrabak J. Biochemical characterization of VIM-39, a VIM-1-like metallo-beta-lactamase variant from a multidrug-resistant Klebsiella pneumoniae isolate from Greece. Antimicrob. Agents Chemother. 2015c;59(12):7811–7814. doi: 10.1128/AAC.01935-15. http://dx.doi.org/10.1128/AAC.01935-15, Epub 2015/09/16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Papagiannitsis CC, Dolejska M, Izdebski R, Giakkoupi P, Skalova A, Chudejova K, et al. Characterisation of IncA/C2 plasmids carrying an In416-like integron with the blaVIM-19 gene from Klebsiella pneumoniae ST383 of Greek origin. Int J Antimicrob Agents. 2016;47(2):158–162. doi: 10.1016/j.ijantimicag.2015.12.001. http://dx.doi.org/10.1016/j.ijantimicag.2015.12.001, Epub 2016/01/23. [DOI] [PubMed] [Google Scholar]
  135. Papp-Wallace KM, Winkler ML, Taracila MA, Bonomo RA. Variants of beta-lactamase KPC-2 that are resistant to inhibition by avibactam. Antimicrob Agents Chemother. 2015;59(7):3710–3717. doi: 10.1128/AAC.04406-14. http://dx.doi.org/10.1128/AAC.04406-14, Epub 2015/02/11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Pasteran F, Gonzalez LJ, Albornoz E, Bahr G, Vila AJ, Corso A. Triton hodge test: improved protocol for modified hodge test for enhanced detection of NDM and other carbapenemase producers. J Clin Microbiol. 2016;54(3):640–649. doi: 10.1128/JCM.01298-15. http://dx.doi.org/10.1128/JCM.01298-15, Epub 2016/01/01. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Patel JB. Methods for Dilution Antimicrobial Susceptibility Tests for Bacteria that Grow Aerobically. In: Institute CLaS, editor. Approved Standard. 10 2015. [Google Scholar]
  138. Pesesky MW, Hussain T, Wallace M, Wang B, Andleeb S, Burnham CA, et al. KPC and NDM-1 genes in related Enterobacteriaceae strains and plasmids from Pakistan and the United States. Emerg Infect Dis. 2015;21(6):1034–1037. doi: 10.3201/eid2106.141504. http://dx.doi.org/10.3201/eid2106.141504, Epub 2015/05/20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  139. Pitart C, Sole M, Roca I, Roman A, Moreno A, Vila J, et al. Molecular characterization of blaNDM-5 carried on an IncFII plasmid in an Escherichia coli isolate from a nontraveler patient in Spain. Antimicrob Agents Chemother. 2015;59(1):659–662. doi: 10.1128/AAC.04040-14. http://dx.doi.org/10.1128/AAC.04040-14, Epub 2014/10/15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Pitout JD, Nordmann P, Poirel L. Carbapenemase-producing Klebsiella pneumoniae, a key pathogen set for global nosocomial dominance. Antimicrob Agents Chemother. 2015;59(10):5873–5884. doi: 10.1128/AAC.01019-15. http://dx.doi.org/10.1128/AAC.01019-15, Epub 2015/07/15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  141. Poirel L, Nordmann P. Rapidec carba NP test for rapid detection of carbapenemase producers. J Clin Microbiol. 2015;53(9):3003–3008. doi: 10.1128/JCM.00977-15. http://dx.doi.org/10.1128/JCM.00977-15, Epub 2015/06/19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Poirel L, Weldhagen GF, Naas T, De Champs C, Dove MG, Nordmann P. GES-2, a class A beta-lactamase from Pseudomonas aeruginosa with increased hydrolysis of imipenem. Antimicrob Agents Chemother. 2001;45(9):2598–2603. doi: 10.1128/AAC.45.9.2598-2603.2001. Epub 2001/08/15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Poirel L, Heritier C, Podglajen I, Sougakoff W, Gutmann L, Nordmann P. Emergence in Klebsiella pneumoniae of a chromosome-encoded SHV beta- lactamase that compromises the efficacy of imipenem. Antimicrob Agents Chemother. 2003;47(2):755–758. doi: 10.1128/AAC.47.2.755-758.2003. Epub 2003/01/25. [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Poirel L, Heritier C, Nordmann P. Chromosome-encoded ambler class D beta-lactamase of Shewanella oneidensis as a progenitor of carbapenem-hydrolyzing oxacillinase. Antimicrob Agents Chemother. 2004a;48(1):348–351. doi: 10.1128/AAC.48.1.348-351.2004. Epub 2003/12/25. [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Poirel L, Heritier C, Tolun V, Nordmann P. Emergence of oxacillinase-mediated resistance to imipenem in Klebsiella pneumoniae. Antimicrob Agents Chemother. 2004b;48(1):15–22. doi: 10.1128/AAC.48.1.15-22.2004. Epub 2003/12/25. [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Poirel L, Pitout JD, Nordmann P. Carbapenemases: molecular diversity and clinical consequences. Future Microbiol. 2007;2(5):501–512. doi: 10.2217/17460913.2.5.501. http://dx.doi.org/10.2217/17460913.2.5.501, Epub 2007/10/12. [DOI] [PubMed] [Google Scholar]
  147. Poirel L, Bonnin RA, Nordmann P. Genetic features of the widespread plasmid coding for the carbapenemase OXA-48. Antimicrob Agents Chemother. 2012a;56(1):559–562. doi: 10.1128/AAC.05289-11. http://dx.doi.org/10.1128/AAC.05289-11, Epub 2011/11/16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  148. Poirel L, Potron A, Nordmann P. OXA-48-like carbapenemases: the phantom menace. J Antimicrob Chemother. 2012b;67(7):1597–1606. doi: 10.1093/jac/dks121. http://dx.doi.org/10.1093/jac/dks121, Epub 2012/04/14. [DOI] [PubMed] [Google Scholar]
  149. Pollett S, Miller S, Hindler J, Uslan D, Carvalho M, Humphries RM. Phenotypic and molecular characteristics of carbapenem-resistant Enterobacteriaceae in a health care system in Los Angeles, California, from 2011 to 2013. J Clin Microbiol. 2014;52(11):4003–4009. doi: 10.1128/JCM.01397-14. http://dx.doi.org/10.1128/jcm.01397-14, Eub 2014/09/12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Potron A, Poirel L, Nordmann P. Derepressed transfer properties leading to the efficient spread of the plasmid encoding carbapenemase OXA-48. Antimicrob Agents Chemother. 2014;58(1):467–471. doi: 10.1128/AAC.01344-13. http://dx.doi.org/10.1128/AAC.01344-13, Epub 2013/11/06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Pucci MJ, Bush K. Investigational antimicrobial agents of 2013. Clin Microbiol Rev. 2013;26(4):792–821. doi: 10.1128/CMR.00033-13. http://dx.doi.org/10.1128/CMR.00033-13, Epub 2013/10/05. [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Rasmussen BA, Bush K, Keeney D, Yang Y, Hare R, O’Gara C, et al. Characterization of IMI-1 beta-lactamase, a class A carbapenem-hydrolyzing enzyme from Enterobacter cloacae. Antimicrob Agents Chemother. 1996;40(9):2080–2086. doi: 10.1128/aac.40.9.2080. Epub 1996/09/01. [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Rodriguez-Avial I, Pena I, Picazo JJ, Rodriguez-Avial C, Culebras E. In vitro activity of the next-generation aminoglycoside plazomicin alone and in combination with colistin, meropenem, fosfomycin or tigecycline against carbapenemase-producing Enterobacteriaceae strains. Int J Antimicrob Agents. 2015;46(6):616–621. doi: 10.1016/j.ijantimicag.2015.07.021. http://dx.doi.org/10.1016/j.ijantimicag.2015.07.021, Epub 2015/09/24. [DOI] [PubMed] [Google Scholar]
  154. Roemer T, Boone C. Systems-level antimicrobial drug and drug synergy discovery. Nat Chem Biol. 2013;9(4):222–231. doi: 10.1038/nchembio.1205. http://dx.doi.org/10.1038/nchembio.1205, Epub 2013/03/20. [DOI] [PubMed] [Google Scholar]
  155. Sarkar A, Pazhani GP, Chowdhury G, Ghosh A, Ramamurthy T. Attributes of carbapenemase encoding conjugative plasmid pNDM-SAL from an extensively drug-resistant Salmonella enterica serovar senftenberg. Front Microbiol. 2015;6:969. doi: 10.3389/fmicb.2015.00969. http://dx.doi.org/10.3389/fmicb.2015.00969, Epub 2015/10/07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Satlin MJ, Jenkins SG, Walsh TJ. The global challenge of carbapenem-resistant Enterobacteriaceae in transplant recipients and patients with hematologic malignancies. Clin Infect Dis. 2014;58(9):1274–1283. doi: 10.1093/cid/ciu052. http://dx.doi.org/10.1093/cid/ciu052, Epub 2014/01/28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  157. Savard P, Gopinath R, Zhu W, Kitchel B, Rasheed JK, Tekle T, et al. First NDM-positive Salmonella sp. strain identified in the United States. Antimicrob Agents Chemother. 2011;55(12):5957–5958. doi: 10.1128/AAC.05719-11. http://dx.doi.org/10.1128/AAC.05719-11, Epub 2011/10/05. [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Saw HT, Webber MA, Mushtaq S, Woodford N, Piddock LJ. Inactivation or inhibition of AcrAB-TolC increases resistance of carbapenemase-producing Enterobacteriaceae to carbapenems. J Antimicrob Chemother. 2016;71(6):1510–1519. doi: 10.1093/jac/dkw028. http://dx.doi.org/10.1093/jac/dkw028, Epub 2016/03/08. [DOI] [PubMed] [Google Scholar]
  159. Shakil S, Azhar EI, Tabrez S, Kamal MA, Jabir NR, Abuzenadah AM, et al. New Delhi metallo-beta-lactamase (NDM-1): an update. J Chemother. 2011;23(5):263–265. doi: 10.1179/joc.2011.23.5.263. http://dx.doi.org/10.1179/joc.2011.23.5.263, Epub 2011/10/19. [DOI] [PubMed] [Google Scholar]
  160. Shields RK, Clancy CJ, Hao B, Chen L, Press EG, Iovine NM, et al. Effects of Klebsiella pneumoniae carbapenemase subtypes, extended-spectrum beta-lactamases, and porin mutations on the in vitro activity of ceftazidime-avibactam against carbapenem-resistant K. pneumoniae. Antimicrob Agents Chemother. 2015;59(9):5793–5797. doi: 10.1128/AAC.00548-15. http://dx.doi.org/10.1128/AAC.00548-15, Epub 2015/07/15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Shin SY, Bae IK, Kim J, Jeong SH, Yong D, Kim JM, et al. Resistance to carbapenems in sequence type 11 Klebsiella pneumoniae is related to DHA-1 and loss of OmpK35 and/or OmpK36. J Med Microbiol. 2012;61(pt 2):239–245. doi: 10.1099/jmm.0.037036-0. http://dx.doi.org/10.1099/jmm.0.037036-0, Epub 2011/09/24. [DOI] [PubMed] [Google Scholar]
  162. Shin J, Baek JY, Cho SY, Huh HJ, Lee NY, Song JH, et al. blaNDM-5-bearing IncFII-Type plasmids of Klebsiella pneumoniae sequence type 147 transmitted by cross-border transfer of a patient. Antimicrob Agents Chemother. 2016;60(3):1932–1934. doi: 10.1128/AAC.02722-15. http://dx.doi.org/10.1128/AAC.02722-15, Epub 2016/01/30. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  163. Shrestha B, Tada T, Miyoshi-Akiyama T, Shimada K, Ohara H, Kirikae T, et al. Identification of a novel NDM variant, NDM-13, from a multidrug-resistant Escherichia coli clinical isolate in Nepal. Antimicrob Agents Chemother. 2015;59(9):5847–5850. doi: 10.1128/AAC.00332-15. http://dx.doi.org/10.1128/AAC.00332-15, Epub 2015/07/15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Sidjabat HE, Townell N, Nimmo GR, George NM, Robson J, Vohra R, et al. Dominance of IMP-4-producing enterobacter cloacae among carbapenemase-producing Enterobacteriaceae in Australia. Antimicrob Agents Chemother. 2015;59(7):4059–4066. doi: 10.1128/AAC.04378-14. http://dx.doi.org/10.1128/AAC.04378-14, Epub 2015/04/29. [DOI] [PMC free article] [PubMed] [Google Scholar]
  165. Skurnik D, Roux D, Pons S, Guillard T, Lu X, Cywes-Bentley C, et al. Extended-spectrum antibodies protective against carbapenemase-producing Enterobacteriaceae. J Antimicrob Chemother. 2016;71(4):927–935. doi: 10.1093/jac/dkv448. http://dx.doi.org/10.1093/jac/dkv448, Epub 2016/01/10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  166. Struve C, Roe CC, Stegger M, Stahlhut SG, Hansen DS, Engelthaler DM, et al. Mapping the evolution of hypervirulent Klebsiella pneumoniae. mBio. 2015;6(4):e00630. doi: 10.1128/mBio.00630-15. http://dx.doi.org/10.1128/mBio.00630-15, Epub 2015/07/23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  167. Sun F, Yin Z, Feng J, Qiu Y, Zhang D, Luo W, et al. Production of plasmid-encoding NDM-1 in clinical Raoultella ornithinolytica and Leclercia adecarboxylata from China. Front Microbiol. 2015;6:458. doi: 10.3389/fmicb.2015.00458. http://dx.doi.org/10.3389/fmicb.2015.00458, Epub 2015/06/09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Surawicz CM, Brandt LJ, Binion DG, Ananthakrishnan AN, Curry SR, Gilligan PH, et al. Guidelines for diagnosis, treatment, and prevention of Clostridium difficile infections. Am J Gastroenterol. 2013;108(4):478–498. doi: 10.1038/ajg.2013.4. http://dx.doi.org/10.1038/ajg.2013.4, quiz 99. Epub 2013/02/27. [DOI] [PubMed] [Google Scholar]
  169. Tamma PD, Suwantarat N, Rudin SD, Logan LK, Simner PJ, Coy LR, et al. First report of a verona integron-encoded metallo-beta-lactamase-producing Klebsiella pneumoniae infection in a child in the United States. J Pediatr Infect Dis Soc. 2016;5:e24–e27. doi: 10.1093/jpids/piw025. http://dx.doi.org/10.1093/jpids/piw025, Epub 2016/05/06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. Tangden T, Giske CG. Global dissemination of extensively drug-resistant carbapenemase-producing Enterobacteriaceae: clinical perspectives on detection, treatment and infection control. J Intern Med. 2015;277(5):501–512. doi: 10.1111/joim.12342. http://dx.doi.org/10.1111/joim.12342, Epub 2015/01/06. [DOI] [PubMed] [Google Scholar]
  171. Tato M, Ruiz-Garbajosa P, Traczewski M, Dodgson A, McEwan A, Humphries R, et al. Multisite evaluation of Cepheid Xpert-Carba-R assay for the detection of carbapenemase-producing organisms in rectal swabs. J Clin Microbiol. 2016 doi: 10.1128/JCM.00341-16. http://dx.doi.org/10.1128/JCM.00341-16, Epub 2016/04/29. [DOI] [PMC free article] [PubMed]
  172. Thomson GK, Snyder JW, McElheny CL, Thomson KS, Doi Y. Coproduction of KPC-18 and VIM-1 carbapenemases by Enterobacter cloacae: implications for newer beta-lactam-beta-lactamase inhibitor combinations. J Clin Microbiol. 2016;54(3):791–794. doi: 10.1128/JCM.02739-15. http://dx.doi.org/10.1128/JCM.02739-15, Epub 2016/01/01. [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Todorova B, Sabtcheva S, Ivanov IN, Lesseva M, Chalashkanov T, Ioneva M, et al. First clinical cases of NDM-1-producing Klebsiella pneumoniae from two hospitals in Bulgaria. J Infect Chemother. 2016 doi: 10.1016/j.jiac.2016.03.014. http://dx.doi.org/10.1016/j.jiac.2016.03.014, Epub 2016/05/01. [DOI] [PubMed]
  174. Toledo PV, Aranha Junior AA, Arend LN, Ribeiro V, Zavascki AP, Tuon FF. Activity of antimicrobial combinations against KPC-2-producing Klebsiella pneumoniae in a rat model and time-kill assay. Antimicrob Agents Chemother. 2015;59(7):4301–4304. doi: 10.1128/AAC.00323-15. http://dx.doi.org/10.1128/AAC.00323-15, Epub 2015/04/22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  175. Tsai YK, Liou CH, Fung CP, Lin JC, Siu LK. Single or in combination antimicrobial resistance mechanisms of Klebsiella pneumoniae contribute to varied susceptibility to different carbapenems. PLoS One. 2013;8(11):e79640. doi: 10.1371/journal.pone.0079640. http://dx.doi.org/10.1371/journal.pone.0079640, Epub 2013/11/23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  176. Tsakris A, Poulou A, Bogaerts P, Dimitroulia E, Pournaras S, Glupczynski Y. Evaluation of a new phenotypic OXA-48 disk test for differentiation of OXA-48 carbapenemase-producing Enterobacteriaceae clinical isolates. J Clin Microbiol. 2015;53(4):1245–1251. doi: 10.1128/JCM.03318-14. http://dx.doi.org/10.1128/JCM.03318-14, Epub 2015/02/06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Tseng IL, Liu YM, Wang SJ, Yeh HY, Hsieh CL, Lu HL, et al. Emergence of carbapenemase producing Klebsiella pneumonia and spread of KPC-2 and KPC-17 in Taiwan: a nationwide study from 2011 to 2013. PLoS One. 2015;10(9):e0138471. doi: 10.1371/journal.pone.0138471. http://dx.doi.org/10.1371/journal.pone.0138471, Epub 2015/09/19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  178. van der Zwaluw K, de Haan A, Pluister GN, Bootsma HJ, de Neeling AJ, Schouls LM. The carbapenem inactivation method (CIM), a simple and low-cost alternative for the Carba NP test to assess phenotypic carbapenemase activity in gram-negative rods. PLoS One. 2015;10(3):e0123690. doi: 10.1371/journal.pone.0123690. http://dx.doi.org/10.1371/journal.pone.0123690, Epub 2015/03/24. [DOI] [PMC free article] [PubMed] [Google Scholar]
  179. van Duin D, Cober ED, Richter SS, Perez F, Cline M, Kaye KS, et al. Tigecycline therapy for carbapenem-resistant Klebsiella pneumoniae (CRKP) bacteriuria leads to tigecycline resistance. Clin Microbiol Infect. 2014;20(12):O1117–O1120. doi: 10.1111/1469-0691.12714. http://dx.doi.org/10.1111/1469-0691.12714, Epub 2014/06/17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  180. van Duin D, Cober E, Richter SS, Perez F, Kalayjian RC, Salata RA, et al. Residence in skilled nursing facilities is associated with tigecycline non-susceptibility in carbapenem-resistant Klebsiella pneumoniae. Infect Control Hosp Epidemiol. 2015;36(8):942–948. doi: 10.1017/ice.2015.118. http://dx.doi.org/10.1017/ice.2015.118, Epub 2015/05/21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  181. Vasoo S, Cunningham SA, Cole NC, Kohner PC, Menon SR, Krause KM, et al. In vitro activities of ceftazidime-avibactam, aztreonam-avibactam, and a panel of older and contemporary antimicrobial agents against carbapenemase-producing gram-negative bacilli. Antimicrob Agents Chemother. 2015;59(12):7842–7846. doi: 10.1128/AAC.02019-15. http://dx.doi.org/10.1128/AAC.02019-15, Epub 2015/09/24. [DOI] [PMC free article] [PubMed] [Google Scholar]
  182. Vercoe RB, Chang JT, Dy RL, Taylor C, Gristwood T, Clulow JS, et al. Cytotoxic chromosomal targeting by CRISPR/Cas systems can reshape bacterial genomes and expel or remodel pathogenicity islands. PLoS Genet. 2013;9(4):e1003454. doi: 10.1371/journal.pgen.1003454. http://dx.doi.org/10.1371/journal.pgen.1003454. [DOI] [PMC free article] [PubMed] [Google Scholar]
  183. Viale P, Tumietto F, Giannella M, Bartoletti M, Tedeschi S, Ambretti S, et al. Impact of a hospital-wide multifaceted programme for reducing carbapenem-resistant Enterobacteriaceae infections in a large teaching hospital in northern Italy. Clin Microbiol Infect. 2015;21(3):242–247. doi: 10.1016/j.cmi.2014.10.020. http://dx.doi.org/10.1016/j.cmi.2014.10.020, Epub 2015/02/07. [DOI] [PubMed] [Google Scholar]
  184. Viau R, Frank KM, Jacobs MR, Wilson B, Kaye K, Donskey CJ, et al. Intestinal carriage of carbapenemase-producing organisms: current status of surveillance methods. Clin Microbiol Rev. 2016;29(1):1–27. doi: 10.1128/CMR.00108-14. http://dx.doi.org/10.1128/CMR.00108-14, Epub 2015/10/30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Vilacoba E, Quiroga C, Pistorio M, Famiglietti A, Rodriguez H, Kovensky J, et al. A blaVIM-2 plasmid disseminating in extensively drug-resistant clinical Pseudomonas aeruginosa and Serratia marcescens isolates. Antimicrob Agents Chemother. 2014;58(11):7017–7018. doi: 10.1128/AAC.02934-14. http://dx.doi.org/10.1128/AAC.02934-14, Epub 2014/09/04. [DOI] [PMC free article] [PubMed] [Google Scholar]
  186. Voulgari E, Poulou A, Dimitroulia E, Politi L, Ranellou K, Gennimata V, et al. Emergence of OXA-162 carbapenemase- and DHA-1 AmpC cephalosporinase-producing sequence type 11 Klebsiella pneumoniae causing community-onset infection in Greece. Antimicrob Agents Chemother. 2015;60(3):1862–1864. doi: 10.1128/AAC.01514-15. http://dx.doi.org/10.1128/AAC.01514-15, Epub 2015/12/17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  187. Wailan AM, Sartor AL, Zowawi HM, Perry JD, Paterson DL, Sidjabat HE. Genetic contexts of blaNDM-1 in patients carrying multiple NDM-producing strains. Antimicrob Agents Chemother. 2015;59(12):7405–7410. doi: 10.1128/AAC.01319-15. http://dx.doi.org/10.1128/AAC.01319-15, Epub 2015/09/24. [DOI] [PMC free article] [PubMed] [Google Scholar]
  188. Wailan AM, Paterson DL, Kennedy K, Ingram PR, Bursle E, Sidjabat HE. Genomic characteristics of NDM-producing Enterobacteriaceae isolates in Australia and their blaNDM genetic contexts. Antimicrob Agents Chemother. 2016a;60(1):136–141. doi: 10.1128/AAC.01243-15. http://dx.doi.org/10.1128/AAC.01243-15, Epub 2015/10/21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  189. Wailan AM, Sidjabat HE, Yam WK, Alikhan NF, Petty NK, Sartor AL, et al. Mechanisms involved in acquisition of blaNDM genes by IncA/C2 and IncFIIY plasmids. Antimicrob Agents Chemother. 2016b;60:4082–4088. doi: 10.1128/AAC.00368-16. http://dx.doi.org/10.1128/AAC.00368-16, Epub 2016/04/27. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Walker GT, Rockweiler TJ, Kersey RK, Frye KL, Mitchner SR, Toal DR, et al. Analytical performance of multiplexed screening test for 10 antibiotic resistance genes from perianal swab samples. Clin Chem. 2016;62(2):353–359. doi: 10.1373/clinchem.2015.246371. http://dx.doi.org/10.1373/clinchem.2015.246371, Epub 2015/12/08. [DOI] [PubMed] [Google Scholar]
  191. Walther-Rasmussen J, Hoiby N. Class A carbapenemases. J Antimicrob Chemother. 2007;60(3):470–482. doi: 10.1093/jac/dkm226. http://dx.doi.org/10.1093/jac/dkm226, Epub 2007/06/28. [DOI] [PubMed] [Google Scholar]
  192. Wang ZK, Yang YS, Chen Y, Yuan J, Sun G, Peng LH. Intestinal microbiota pathogenesis and fecal microbiota transplantation for inflammatory bowel disease. World J Gastroenterol. 2014;20(40):14805–14820. doi: 10.3748/wjg.v20.i40.14805. http://dx.doi.org/10.3748/wjg.v20.i40.14805, Epub 2014/10/31. [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. Wang L, Fang H, Feng J, Yin Z, Xie X, Zhu X, et al. Complete sequences of KPC-2-encoding plasmid p628-KPC and CTX-M-55-encoding p628-CTXM coexisted in Klebsiella pneumoniae. Front Microbiol. 2015;6:838. doi: 10.3389/fmicb.2015.00838. http://dx.doi.org/10.3389/fmicb.2015.00838, Epub 2015/09/09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  194. Warner DM, Yang Q, Duval V, Chen M, Xu Y, Levy SB. Involvement of MarR and YedS in carbapenem resistance in a clinical isolate of Escherichia coli from China. Antimicrob Agents Chemother. 2013;57(4):1935–1937. doi: 10.1128/AAC.02445-12. http://dx.doi.org/10.1128/AAC.02445-12, Epub 2013/01/16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  195. Watkins RR, Papp-Wallace KM, Drawz SM, Bonomo RA. Novel beta-lactamase inhibitors: a therapeutic hope against the scourge of multidrug resistance. Front Microbiol. 2013;4:392. doi: 10.3389/fmicb.2013.00392. http://dx.doi.org/10.3389/fmicb.2013.00392, Epub 2014/01/09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  196. Webb HE, Bugarel M, den Bakker HC, Nightingale KK, Granier SA, Scott HM, et al. Carbapenem-resistant bacteria recovered from faeces of dairy cattle in the high plains region of the USA. PLoS One. 2016;11(1):e0147363. doi: 10.1371/journal.pone.0147363. http://dx.doi.org/10.1371/journal.pone.0147363, Epub 2016/01/30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  197. Wu W, Feng Y, Carattoli A, Zong Z. Characterization of an Enterobacter cloacae strain producing both KPC and NDM carbapenemases by whole-genome sequencing. Antimicrob Agents Chemother. 2015;59(10):6625–6628. doi: 10.1128/AAC.01275-15. http://dx.doi.org/10.1128/AAC.01275-15, Epub 2015/08/08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  198. Xavier BB, Lammens C, Ruhal R, Kumar-Singh S, Butaye P, Goossens H, et al. Identification of a novel plasmid-mediated colistin-resistance gene, mcr- 2, in Escherichia coli, Belgium, June 2016. Euro Surveill.: bulletin Europeen sur les maladies transmissibles = Eur Commun Dis Bull. 2016;21(27):s1. doi: 10.2807/1560-7917.ES.2016.21.27.30280. http://dx.doi.org/10.2807/1560-7917.ES.2016.21.27.30280, Epub 2016/07/16. [DOI] [PubMed] [Google Scholar]
  199. Yaffee AQ, Roser L, Daniels K, Humbaugh K, Brawley R, Thoroughman D, et al. Notes from the field: verona integron-encoded metallo-beta-lactamase-producing carbapenem-resistant Enterobacteriaceae in a neonatal and adult intensive care unit—Kentucky, 2015. MMWR Morb Mortal Wkly Rep. 2016;65(7):190. doi: 10.15585/mmwr.mm6507a5. http://dx.doi.org/10.15585/mmwr.mm6507a5, Epub 2016/02/26. [DOI] [PubMed] [Google Scholar]
  200. Yang YJ, Wu PJ, Livermore DM. Biochemical characterization of a beta-lactamase that hydrolyzes penems and carbapenems from two Serratia marcescens isolates. Antimicrob Agents Chemother. 1990;34(5):755–758. doi: 10.1128/aac.34.5.755. Epub 1990/05/01. [DOI] [PMC free article] [PubMed] [Google Scholar]
  201. Yang Q, Fang L, Fu Y, Du X, Shen Y, Yu Y. Dissemination of NDM- 1-producing Enterobacteriaceae mediated by the IncX3-Type plasmid. PLoS One. 2015;10(6):e0129454. doi: 10.1371/journal.pone.0129454. http://dx.doi.org/10.1371/journal.pone.0129454, Epub 2015/06/06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  202. Yeh PJ, Hegreness MJ, Aiden AP, Kishony R. Drug interactions and the evolution of antibiotic resistance. Nat Rev Microbiol. 2009;7(6):460–466. doi: 10.1038/nrmicro2133. http://dx.doi.org/10.1038/nrmicro2133, Epub 2009/05/16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  203. Yigit H, Queenan AM, Anderson GJ, Domenech-Sanchez A, Biddle JW, Steward CD, et al. Novel carbapenem-hydrolyzing beta-lactamase, KPC-1, from a carbapenem-resistant strain of Klebsiella pneumoniae. Antimicrob Agents Chemother. 2001;45(4):1151–1161. doi: 10.1128/AAC.45.4.1151-1161.2001. http://dx.doi.org/10.1128/AAC.45.4.1151-1161.2001, Epub 2001/03/21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Yigit H, Queenan AM, Rasheed JK, Biddle JW, Domenech-Sanchez A, Alberti S, et al. Carbapenem-resistant strain of Klebsiella oxytoca harboring carbapenem-hydrolyzing beta-lactamase KPC-2. Antimicrob Agents Chemother. 2003;47(12):3881–3889. doi: 10.1128/AAC.47.12.3881-3889.2003. Epub 2003/11/26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  205. Yu FY, Yao D, Pan JY, Chen C, Qin ZQ, Parsons C, et al. High prevalence of plasmid-mediated 16S rRNA methylase gene rmtB among Escherichia coli clinical isolates from a Chinese teaching hospital. BMC Infect Dis. 2010;10:184. doi: 10.1186/1471-2334-10-184. http://dx.doi.org/10.1186/1471-2334-10-184, Epub 2010/06/25. [DOI] [PMC free article] [PubMed] [Google Scholar]
  206. Zamorano L, Miro E, Juan C, Gomez L, Bou G, Gonzalez-Lopez JJ, et al. Mobile genetic elements related to the diffusion of plasmid-mediated AmpC beta-lactamases or carbapenemases from Enterobacteriaceae: findings from a multicenter study in Spain. Antimicrob Agents Chemother. 2015;59(9):5260–5266. doi: 10.1128/AAC.00562-15. http://dx.doi.org/10.1128/AAC.00562-15, Epub 2015/06/17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  207. Zhanel GG, Lawson CD, Adam H, Schweizer F, Zelenitsky S, Lagace-Wiens PR, et al. Ceftazidime-avibactam: a novel cephalosporin/beta-lactamase inhibitor combination. Drugs. 2013;73(2):159–177. doi: 10.1007/s40265-013-0013-7. http://dx.doi.org/10.1007/s40265-013-0013-7, Epub 2013/02/02. [DOI] [PubMed] [Google Scholar]
  208. Zhang R, Lin D, Chan EW, Gu D, Chen GX, Chen S. Emergence of carbapenem-resistant serotype K1 hypervirulent Klebsiella pneumoniae strains in China. Antimicrob Agents Chemother. 2016;60(1):709–711. doi: 10.1128/AAC.02173-15. http://dx.doi.org/10.1128/AAC.02173-15, Epub 2015/11/18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  209. Zheng B, Zhang J, Ji J, Fang Y, Shen P, Ying C, et al. Emergence of Raoultella ornithinolytica coproducing IMP-4 and KPC-2 carbapenemases in China. Antimicrob Agents Chemother. 2015;59(11):7086–7089. doi: 10.1128/AAC.01363-15. http://dx.doi.org/10.1128/AAC.01363-15, Epub 2015/08/19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  210. Zheng R, Zhang Q, Guo Y, Feng Y, Liu L, Zhang A, et al. Outbreak of plasmid-mediated NDM-1-producing Klebsiella pneumoniae ST105 among neonatal patients in Yunnan, China. Ann Clin Microbiol Antimicrob. 2016;15:10. doi: 10.1186/s12941-016-0124-6. http://dx.doi.org/10.1186/s12941-016-0124-6, Epub 2016/02/21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  211. Zmarlicka MT, Nailor MD, Nicolau DP. Impact of the New Delhi metallo-beta-lactamase on beta-lactam antibiotics. Infect Drug Resist. 2015;8:297–309. doi: 10.2147/IDR.S39186. http://dx.doi.org/10.2147/IDR.S39186, Epub 2015/09/09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  212. Zowawi HM, Forde BM, Alfaresi M, Alzarouni A, Farahat Y, Chong TM, et al. Stepwise evolution of pandrug-resistance in Klebsiella pneumoniae. Sci Rep. 2015;5:15082. doi: 10.1038/srep15082. http://dx.doi.org/10.1038/srep15082, Epub 2015/10/20. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES