Open access peer-reviewed chapter

Carbapenemases

Written By

Branka Bedenić and Sanda Sardelić

Submitted: 15 January 2018 Reviewed: 07 March 2018 Published: 05 November 2018

DOI: 10.5772/intechopen.76282

From the Edited Volume

Growing and Handling of Bacterial Cultures

Edited by Madhusmita Mishra

Chapter metrics overview

2,005 Chapter Downloads

View Full Metrics

Abstract

Carbapenems are usually regarded as the last treatment option for serious infections caused by Gram-negative and Gram-positive microorganisms. Although they are stable to hydrolysis by most β-lactamases, their usage as the last resort antbiotics was seriously compromised by the appearance of carbapenem-hydrolyzing enzymes called carbapenemases. These enzymes are produced mostly by Enterobacteriaceae and Gram-negative nonfermentative bacteria such as Pseudomonas aeruginosa and Acinetobacter baumannii. True carbapenemases belonging to Ambler molecular classes A, B, and D are often encoded by genes embedded in mobile genetic elements like plasmids, integrons, and transposons, which often harbor multiple resistance determinants limiting further the treatment options. At present, large number of nosocomial and community-acquired infections caused by worldwide spread of carbapenem-resistant Gram-negative bacteria have become a major public health problem. Although polymyxins remain active, in vitro reports of benefits of combination schemes favor this strategy against carbapenemase-producing Gram-negative bacilli.

Keywords

  • carbapenems
  • carbapenemases
  • Enterobacteriaceae
  • Acinetobacter baumannii
  • Pseudomonas aeruginosa

1. Introduction

Carbapenems are bactericidal antibiotics often used as a last treatment option for severe infections caused by isolates producing extended-spectrum and/or AmpC β-lactamases. However, their stability to β-lactamases of Gram-negative bacteria is not absolute. From the first report of carbapenemases in the 1980s, the last decade is marked with the rapid spread of carbapenemases-producing organisms, becoming a public health concern worldwide.

Advertisement

2. Carbapenem antibiotics

Carbapenems are β-lactam antibiotics with broad antimicrobial spectrum, which encompasses Gram-positive and Gram-negative aerobic organisms and anaerobic bacteria. They often remain as the last treatment option for infections caused by multiresistant Gram-negative isolates. Their in vitro efficacy is due to good penetration through the outer membrane of Gram-negative bacteria, high affinity for penicillin-binding protein (PBP) molecules and excellent stability against β-lactamases [1]. The carbapenems bind to PBP-1 and PBP-2 and induce spheroplast formation and cell lysis without filament formation. They comprise imipenem, meropenem, ertapenem, and doripenem. In general, imipenem, and doripenem are potent also against Gram-positive bacteria, while meropenem, ertapenem, and doripenem are slightly more effective against Gram-negative organisms. Ertapenem is not active against P. aeruginosa or A. baumannii, while doripenem has lower minimum inhibitory concentrations (MICs) than imipenem and meropenem versus P. aeruginosa and A. baumannii [1] Meropenem has an advantage over imipenem because it is not susceptible to renal dihydropeptidase and can be administered without cilastatin which is neurotoxic and contraindicated in neurological patients. Carbapenems are stable against wide range of serine-based β-lactamases including extended-spectrum β-lactamases (ESBLs), plasmid-mediated AmpC β-lactamases or against strains with inducible and derepressed expression of chromosomal AmpC β-lactamases. Although the stability against these common enzymes is not complete, the rate of hydrolysis is slow and only becomes significant when coupled with reduced outer membrane permeablity. Carbapenems are strong inducers of chromosomal β-lactamases, but β-lactamase induction does not affect their activity [2]. However, induction of chromosomal AmpC beta-lactamase can cause antagonism with penicillins and cephalosporins, and thus carbapenems should not be administed in combination with other β-lactams.

Advertisement

3. Classification, historical background, and properties of carbapenemases

Some β-lactamases, named carbapenemases, can efficiently hydrolyze carbapenems [3]. These include chromosomal enyzmes, which are ubiquitious in some unusual pathogens and mostly environmental isolates (e.g., Stenotrophomonas maltophilia, Aeromonas spp., Flavobacterium spp., Bacteroides fragilis) and acquired carbapenemases of classes A, B, and D [3]. Although several class C carbapenemases have been described (ACT-1, CMY-2, CMY-10, and ADC-68), their producers usually exhibit reduced susceptibility to carbapenems as a consequence of low enzyme’s catalytic efficiency and a permeability defect. Therefore, they are not considered as true carbapenemases [4].

Based on the structure of their active site, carbapenemases are devided into two groups: class A-penicillinase and class D-oxacillinases; they contain serine at their active site and are inhibited by β-lactam inhibitors including clavulanic acid and tazobactam. The second group is metallo-β-lactamases belonging to class B which contain zinc atom at the active site and is inhibited by metal chelators such as ethylenediaminetetraacetic acid (EDTA) [5].

3.1 Class A carbapenemases

Serin carbapenemases of molecular (Ambler) group A are rare, detected mainly in Enterobacteriaceae. They belong to the functional subgroup 2f according to the classification scheme proposed by Bush et al. [6, 7]. The genes encoding them are located on either chromosome or plasmid. This group comprises SME-1 (Serratia marcescens) [8], SME-2 (S. marcescens) [9], SME-3 (S. marcescens) [10], IMI-1 (Enterobacter cloacae) [11], NMC-A (E. cloacae) [12], KPC-1, KPC-2, and KPC-3 (Klebsiella pneumoniae) [13, 14, 15] and GES-2 (P. aeruginosa) [16]. They confer resistance to aminopenicillins, ureidopenicillins, first- and second-generation cephalosporins, aztreonam, and imipenem. Imipenem is hydrolyzed very slowly without causing a clinically relevant resistance [17]. SME β-lactamase was reported for the first time in S. marcescens in the UK in 1982 [8]. Infections caused by SME-positive S. marcescens are usually sporadic without potential of causing outbreaks. IMI and NMC β-lactamases have been sporadically detected in isolates of E. cloacae in the USA, France, and Argentina [9, 10]. These β-lactamase are encoded chromosomally and have never been described on transferable genetic elements explaining their sporadic reports wordwide [17]. KPC-1 is a novel class A enzyme isolated in 1996 from K. pneumoniae in the USA [13]. The discovery of KPC-1 was quickly followed by two new allelic variants: KPC-2 and KPC-3 [14]. Subsequent revision of the blaKPC-1 sequence demonstrated that KPC-1 and KPC-2 are identical enzymes [13, 14, 15]. KPC-3 differs from KPC-2 (the former KPC-1) by one amino acid change (H272Y) [14]. Although, so far, more than 20 different KPC variants have been described, KPC-2 and KPC-3 remain the most commonly identified. Unlike other class A carbapenemases, they hydrolyze expanded-spectrum cephalosporins and all carbapenems. KPC β-lactamases are weakly inhibited by clavulanic acid and tazobactam and are often coharbored with OXA-1 conferring resistance to β-lactam inhibitor combinations. They are not inducible and are encoded on transferable plasmids. A 10-kb Tn3-type transposon Tn4401 with two insertion sequences ISKpn6 and ISKpn7 is repeatedly reported as the main genetic structure enhancing the spread of blaKPC-type genes [5]. Recently, a chromosal integration of the blaKPC gene was shown in K. pneumoniae, thus enabling easier maintenance of the gene in bacterial population [18]. Today, KPC enzymes are highly prevalent, not only found in K. pneumoniae but also in an increasing number of Enterobacteriaceae, P. aeruginosa, and A. baumannii. They are not only found all over the USA but also in Central and South America, while in European countries, KPC-producing isolates are mostly prevalent in Mediterranean countries, especially in Italy and Greece, countries reported as having endemic situation [19, 20].

They are also still present in Middle East where in 2007 a nationwide strategy was implemented to contain the outbreak faced by Israeli hospitals [21].

KPC-producing isolates are also reported in Asia, especially China where a hypervirulent Klebsiella pneumoniae strain K1-producing KPC is concerning [22, 23].

The main transporter in the worldwide dissemination of KPC-producing enzymes is succesfull clone Klebsiella pneumoniae ST258 [24]. In contrast to the majority of other class A carbapenemases, KPCs are typically plasmid mediated and usually carried on IncF plasmid. Plasmid location of blaKPC genes is responsible for the rapid dissemination of those important resistance determinants [23, 24].

GES-2 β-lactamase was derived by point mutation from GES-1 extended-spectrum β-lactamase (with amino acid change G170 N) which was described in K. pneumoniae isolate from French Guinea [16, 25]. GES-2 β-lactamase was first reported from Pseudomonas aeruginosa from South Africa, was plasmid mediated, and conferred resistance or reduced susceptibility to carbapenems [26]. Because the enzymes have broad hydrolysis spectrum that includes penicillins and expanded-spectrum cephalosporins, they were initially classified as ESBLs. Their hydrolysis spectrum was expanded in 2001 to include imipenem. GES β-lactamases are related to integron-borne-cephalosporinase—IBC) ESBLs, but confusing IBC names have been converted to GES [2, 3, 17]. Genes encoding GES family of β-lactamases are located on integrons embedded in plasmids with a possibility of conjugal transfer. However, they appear rarely in contrast to KPC β-lactamases and are not frequently associated with nosocomial outbreaks [3, 17, 26]. At present, the number of GES variants is increasing (with more then 20 described), but only several of them (mainly GES-2, -6, -14, -20) display carbapenemase activity (others are mostly ESBL). GES-type carbapenemases were identified in Pseudomonas aeruginosa, in Acinetobacter baumannii, and Enterobacteriaceae as well (mostly Klebsiella pneumoniae, Serratia marcescens, and Enterobacter spp. isolates) and have worldwide distribution [17, 23]. In spite of the fact that they are plasmid mediated, they do not have a capacity for successful dissemination like KPC [23].

3.2 Class B carbapenemases

Metallo-β-lactamases (MBLs) are clinically the most relevant carbapenemases [23, 27]. This group of enzymes catalyzes the β-lactam hydrolysis by active site containing one or more bounded zinc ions that promote formation of nucleophilic hydoxide further attacking the β-lactam ring. The first descriptions of metallo-β-lactamases included studies on environmental and opportunistic bacteria such as Stenotrophomonas maltophilia, Bacillus cereus, Aeromonas spp., Legionella gormanii, Pseudomonas stutzeri, Shewanella spp., Myroides odoratus, and so on, which carry MBLs on their chromosome but could present potential reservoir of mobile carbapenemase determinants [27, 28].

The metallo-β-lactamases or class B carbapenemases are further divided into subclasses (B1, B2 and B3), but the largest number of clinically relevant MBLs belong to B1 subclass, including the most frequently described Verona integron-encoded MBL (VIM), imipenemase (IMP), and New Delhi MBL (NDM). Those MBLs are usually located within different integron structures, and these integrons are associated with mobile plasmids or transposons facilitating the transfer of resistance genes between bacteria [3, 4, 5]. Acquired MBLs are reported from all over the world but mostly from Europe and Far East [27]. Transferable imipenem resistance was first detected in 1990 in Japan in P. aeruginosa isolate [29]. The β-lactamase-conferring imipenem resistance was designated IMP-1 and was located on a conjugative plasmid [30]. The same enzyme was later found on an integron in S. marcescens and other Enterobacteriaceae also in Japan. The first member of IMP family reported in Europe was IMP-1 found in A. baumannii in Italy [31]. The gene encoding this enzyme was located on class 1 integron [32]. Since that time, the IMP family was found throughout the world, but most reports originate from Far East [4, 22, 27, 33]. IMP-2 β-lactamase was found in an Acinetobacter baumanni isolate from Italy in 1997 and shares 85% of amino acid identity with IMP-1 with similar kinetic parameters, while IMP-3 was reported in Shigella flexneri from Japan. That was the first appearance of MBL in typical community-acquired pathogen [34]. IMP-3 enzyme was previously known as MET-1, but the substrate specificity, that is, preferential hydrolysis of cephalosporins (and not penicilins and carbapenems), was a direct consequence of a single-base mutation (although the genes differ in two bases and two amino acid changes, that is, S262G and E126G). The gene and the enzyme were then renamed to blaIMP-3 or IMP-3 β-lactamase. Since IMP-1 shows broader substrate profile, it seemed that the gene blaIMP-3 was an ancestor of blaIMP-1 [27]. IMP-4 MBL was first reported from A. baumannii associated with nosocomial outbreak in Hong Kong in 2001 [35]. IMP-5 was identified in urinary isolate of A. baumannii from Portugal in 2002, highly resistant to penicillins, broad-spectrum cephalosporins, including ceftazidime, ceftriaxone, cefepime, cefpirome, and to aztreonam, but remained susceptible to ampicillin/sulbactam, aminoglycosides, and quinolones [36]. IMP-5 showed greater homology with IMP-1, IMP-3, and IMP-4 than with IMP-2. BlaIMP-5 was the only gene cassette inserted into a class 1 integron, named In76. This was the first IMP-β-lactamase in Portugal and the second in Europe [36]. There are more than 50 allelic IMP variants reported so far [37]. They are usually devided into IMP-1 and IMP-2 cluster, suggesting their different phylogenetic origin. IMP-12, a very different variant outside the IMP-1 cluster (it has 85% similarity to IMP-1), has a unique structural feature and shows reduced catalytic effieciency toward penicillins [38]. IMP-51 was recently identified in a P. aeruginosa clinical isolate in Vietnam and showed increased doripenem- and meropenem-hydrolyzing activities [39]. IMP-type enzymes have been reported all over the world in Enterobacteriaceae, P. aeruginosa, and A. baumannii but still represent the dominant carbapenemases in Japan and in Southeast Asia [33, 38, 40]. Analysis of the genetic platform of blaIMP genes has shown that most of them are embedded as gene cassettes in class I integron, harboring also other resistance genes, such as aac (mediating resistance to aminoglycosides), blaOXA (different serine oxacillinases), resistance to antiseptics, or chloramphenicol [38, 40, 41].

Another prevalent family of integron-associated metallo-β-lactamases is composed of VIM enzymes. VIM means Verona integron-encoded metallo-β-lactamase because the first enzyme in this family was reported in Pseudomonas aeruginosa isolate from Verona in 1997 [42]. BlaVIM-1 gene was located within class 1 integron repeatedly showing the important role in the dissemination of IMP and VIM genes not only in nonfermentative bacteria but also in Enterobacteriaceae [4]. VIM-2 was described in Pseudomonas aeruginosa from France that was isolated in 1996 and shared 90% similarity with VIM-1 [43]. VIM-3 was reported in P. aeruginosa isolates associated with nosocomial outbreak from Taiwan [44]. The enzyme was chromosomally encoded and differed in two amino acids from VIM-1, namely C178A and A443G. VIM-4 β-lactamases were identified in K. pneumoniae and Enterobacter cloacae isolates from Italy in a patient previously treated with carbapenems and has one amino acid change in comparison to VIM-1 (S228A), being more active against imipenem than VIM-1 [45]. There are more than 40 allelic variants of VIM enzymes reported so far, and they mainly belong to three sublineages—VIM-1-like, VIM-2-like, and VIM-7-like enzymes, representing phylogenetically the most divergent variants [38, 46]. VIM-2-like enzymes have been associated mostly with P. aeruginosa, whereas VIM-1-like enzymes, in particular, VIM-4, have been frequently reported in strains of the Enterobacteriaceae [38, 40, 46]. It is interesting to mention that if VIM-1 or VIM-4 enzymes are found in E. cloacae and K. pneumoniae, mostly of ST147 and 11, they are usually carried on incompatibility group N or A/C plasmids, whereas in E. coli, those genes are often associated to IncFI/II plasmids [38, 41]. VIM-7, a more divergent enzyme, with lower efficiency for cephalosporins, was for the first time reported in the USA in 2001, and it was the first MBL in North America [47, 48]. North America was spared from MBLs until early 2000s. VIM enzymes are today widely distributed in all continents and at present the most common MBL found all over the world. In Mediterranean countries, such as Italy and Greece, VIM-positive P. aeruginosa strains of clonal complex 235 and 111 have been involved in nosocomial outbreaks [38]. The activity of aztreonam, as the only β-lactam not hydrolyzed by MBLs, is strongly compromised by the copresence of ESBLs, plasmid-mediated AmpC, or other serine β-lactamase together with MBLs [49, 50]. BlaVIM genes are typically embedded in class 1 integrons, which can be incorporated in either plasmids or chromosome. Plasmids carrying blaVIM genes in Enterobacterriaceae belong most frequently to IncA/C or IncN group [23, 38, 41].

NDM is a new family of metallo-β-lactamases unrelated to other MBLs. The first NDM-1 positive isolate from Europe was reported in Sweden from a patient who contracted a K. pneumoniae infection in India in late 2007 [51]. At the beginning, NDM enzymes spread in the Indian subcontinent, and they are still endemic in India, Pakistan, and Bangladesh. They were found not only in clinical isolates but also in public tap water and sewage water [52].

The spread soon involved the UK, historically related to India, but until now, it was described as present in all continents. However, it should be noted that Balkan region, Arabian Peninsula, and North African countries are considered as possible additional reservoirs of NDM-producing strains [38, 53, 54].

So far, the NDM enzymes have been found mostly in Enterobacteriaceae (with the importance of epidemic Klebsiella pneumoniae ST147, 11 and 14 and E. coliST131 and 405), P. aeruginosa, and A. baumannii. Enterobacteriaceae-producing NDM β-lactamases pose a serious public health concern because huge conjugative plasmids carrying the blaNDM genes can have up to 14 other antibiotic resistance determinants and can trasfer this resistance to other bacteria, resulting in multidrug resistance or pan-drug resistance phenotypes [55].

The genetic platform of blaNDM gene mostly involves transposon Tn125 with two flanking ISAba125 elements, often truncated in Enterobacteriaceae and part of broad-host-range plasmid including IncA/C (predominant plasmid type), F, R, H, N, L/M, and X types [38, 41, 53, 54]. NDM producers can acquire additional β-lactamases, even carbapenemases, and are recognized as the source of community-acquired infections [54, 55]. The highest distribution of NDM variants (more than 15 so far) is detected in K. pneumoniae and E. coli species from Asian continent with China and India still serving as the major reservoir of NDM producers [38, 54].

A global surveillance study conducted through 2012–2014 enrolled 202 medical centers from 40 countries all over the world and identified 471 MBL-producing isolates of Pseudomonas aeruginosa and Enterobacteriaceae with 32 different MBL variants. Among them, seven novel allelic MBL variants were identified (VIM-42, -43, -44 and =45, IMP-48 and -49 and NDM-16), differing by one amino acid from previously reported enzymes proving still dynamic genetic background of these worldwide disseminated enzymes [56].

Other MBL-encoding genes are mostly limited to not only a single report or to particular geographic regions, including genes encoding SPM-1, GIM-1 and SIM-1, but also more recently described AIM-1, KHM-1, DIM-1, BIC-1, and TMB-1 [38, 57, 58, 59, 60]. GIM–MBL, unlike other class B carbapenemases, has two zinc ions in the active site.

SPM β-lactamases, originally described in P. aeruginosa strain from Brazil, are mostly limited to South America and have been associated with outbreaks at many hospitals. There were two imported cases in Europe until now, one in Switzerland, and one in the UK [57, 61, 62]. Unlike the other MBL-encoding genes, blaSPM gene was not part of gene cassette or integron, but it was located on plasmid and flanked with two expanded form is insertion segment common region (ISCR) elements probably providing its mobilization capacity [57]. The gene was also described in A. baumannii strain from Tehran [63].

3.3 Class D carbapenemases (carbapenem hydrolyzing class D β-lactamases; chdls)

The main characteristic of these enzymes is their hydrolytic activity toward oxacillin. Ambler group D carbapenemases include oxacillinases typical for Acinetobacter baumannii. [64]. OXA enzymes (the name came since they hydrolyze isoxazolylpenicillins much faster than penicillins) were recently divided into 12 subgroups: OXA-23-like, OXA-24/40-like OXA-48, OXA-51-like, OXA-58-like, OXA-143-like, OXA-253, OXA-211, OXA-213, OXA-214, OXA-229, and OXA-235 [65]. The amino acid sequence identities between members of the same subgroup are more than 90%, whereas the identities between enzymes that belong to different subgroups are less than 70% [65]. Enzymes belonging to OXA-51 group are naturally occurring β-lactamases of A. baumannii (often used as marker of identification) and are normally expressed at low levels but can be overexpressed as a consequence of the ISAba1 location upstream of the genes [64]. The clinically significant resistance is usually expressed in combination with other mechanisms like porin loss or upregulated efflux pumps. OXA-23 subgroup has been reported worldwide but most reports are coming from Acinetobacter baumannii isolates from Latin America. OXA-23-encoding gene is usually located on transposon Tn2006, almost always bracketed on both sides by ISAba1 insertion sequence [5, 65, 66, 67]. It could be located on transferable plasmids, but it was also described integrated in the A. baumannii chromosome. OXA-24/40 subgroup was reported less frequently from A. baumannii, mostly from Iberian Peninsula and the USA [68]. It confers high level of carbapenem resistance and is usually chromosomally encoded. OXA-58 and OXA-143 are usually plasmid mediated, first reported in France and Germany, respectively [69, 70]. Similar to OXA-23, blaOXA-58 is located within transposon usually bracketed by two copies of ISAba3. OXA-72 is the most prevalent allelic variant found in Europe, Far East, and South America [65, 71, 72, 73]. OXA-72 and OXA-23 variants have not only been identified in long-term care facilities but also in environmental samples such as municipal wastewater stressing the fact that strains resistant to antimicrobials are constantly circulating between the hospitals, community, and the nature [74, 75, 76].

Newer oxacillinase subgroups were mostly described as naturally occurring enzymes with low carbapenemase activity in Acinetobacter “non-baumannii“ isolates like OXA-134a and OXA-143 in A. lwoffi, OXA-211 in A. johansonii, OXA-213 in A. calcoaceticus, OXA-214 in A. haemolyticus, OXA-229 in Acinetobacter bereziniae, and OXA-235 in Acinetobacter schindleri [65, 66, 70]. However, it is suggested that those enzymes present a reservoir of enzyme variants that could evolve to a more significant level and could be easily transferred to other bacterial species.

OXA-48 β-lactamase was originally described in a Klebsiella pneumoniae isolate from Istanbul, Turkey, in 2001 [77]. It is now widespread not only in Klebsiella pneumoniae but also in other Enterobacteriacea. Turkey was reported as having the highest epidemiologic level called endemic in 2015. The most important reservoirs of OXA-48 carbapenemases are linked with Turkey, India, Middle East, and North African countries [78]. blaOXA-48 gene is usually located in pOXA-48a plasmid, which belongs to the group of IncL/M plasmids. This plasmid has high conjugation rate, and it is self-conjugative, explaining partly the global dissemination of OXA-48 enzymes [53]. The spread of blaOXA-48 gene is further mediated by IS1999 insertion sequence located upstream of the Tn1999 transposon which is usually carried on L/M plasmid. In the majority of European countries, at present, it is the most prevalent type of carbapenemase among Enterobacteriaceae [65, 78].

OXA-48 hydrolyzes carbapenems at a low level, shows very weak activity against expanded-spectrum cephalosporins, and does not significantly hydrolyze ceftazidime and cefepime but in combination with impermeability can lead to high-level resistance to carbapenems [79]. OXA-48 significantly hydrolyze only penicillins and narrow-spectrum cephalosporins and like other class D enyzmes is not inhibited by clavulanic acid, sulbactam, or tazobactam leading to elevated MICs of amoxycillin/clavulanate and piperacillin/tazobactam, as shown in Table 1. Resistance to expanded-spectrum cephalosporins occurs only in strains positive for addtional ESBL or AmpC. Some OXA-48 variants, including OXA-181 (differing by four amino acids from OXA-48) and OXA-162 (with one amino acid substitution T213A) were also found in Enterobacteriaceae [78]. OXA-162 variant is also able to hydrolyze ceftazidime, whereas OXA-48 is not. Interestingly, OXA-181 is often associated with other carbapenemase genes, and many reports from Indian subcontinent describe isolates coproducing OXA-181, NDM-1, and VIM enzymes [80].

Type of carbapenemaseSubstrate profileInhibiton byCountry of originHost speciesReference
PenicillinsNarrow-spectrum cephalosporinsExpanded-spectrum cephalosporinsAztreonamImipenemMeropenem
Class A
IMI+++Clavulanic acidFranceE. cloacae[11]
NMC+++Clavulanic acidFranceE. cloacae[12]
SME-1+++Clavulanic acidUnited StatesS. marcescens[8]
SME-2+++Clavulanic acidUnited StatesS. marcescens[9]
SME-3+++Clavulanic acidUnited StatesS. marcescens[10]
KPC-1++++++PBAUnited StatesK. pneumoniae[13]
KPC-2++++++PBAUnited StatesK. pneumoniae[13, 15]
KPC-3++++++PBAUnited StatesK. pneumoniae[14]
GES-2+++Clavulanic acidFrench GuineaP. aeruginosa[16]
Class B
IMP-1+++++EDTAJapanP. aeruginosa[29]
IMP-2+++++EDTAItalyA. baumannii[32]
IMP-4+++++EDTAJapanA. baumannii[35]
IMP-5+++++EDTAPortugalA. baumannii[36]
VIM-1+++++EDTAItalyP. aeruginosa[42]
VIM-2+++++EDTAFranceP. aeruginosa[43]
VIM-3+++++EDTATaiwanP. aeruginosa[44]
VIM-4+++++EDTAFranceE. cloacae, K. pneumoniae[45]
VIM-7++++EDTAUSAP. aeruginosa[47]
NDM-1+++++EDTASwedenK. pneumoniae[51]
GIM-1+++++EDTAGermanyP. aeruginosa[58]
SPM-1+++++EDTABrasilP. aeruginosa[57]
SIM-1+++++EDTASouth KoreaP. aeruginosa[59]
DIM-1+++++EDTAThe NetherlandsP. aeruginosa[60]
Class D
OXA-48-like+++TurkeyK. pneumoniae[77]
OXA-23 (ARI-1)++++++UKA. baumannii[67]
OXA-24/40-like++++++NaClSpainA. baumannii[68]
OXA-58-like++++++FranceA. baumannii[69]
OXA-143-like++++++GermanyA. baumannii[70]
OXA-253++++++GermanyA. baumannii[65]

Table 1.

Classification, substrate profile, susceptibility to inhibitors, host species, and county of origin of carbapenemases.

The most prevalent allelic variants are shown. Abbreviations: PBA-phenylboronic acid; EDTA-ethylenediaminetetraacetic acid.

In a recent report, environmental species, Shewanella xiamenensis was identified as the progenitor of the blaOXA-181 gene [81]. It is therefore hypothesized that a mobilization of a chromosomal gene to plasmid happened, and then the gene disseminated further in clinically relevant species. This work emphasizes the constant interplay between microorganisms in the environment, community, and clinics. Gram-negative bacteria are ubiquitous in nature and are widely distributed in soil and water where they can survive for a long time posing a possible reservoir of resistance determinants. Main characteristics of the most prevalent allelic variants of the Ambler classes A, B, and D carbapenemases are shown in Table 1.

Advertisement

4. Clinical significance of carbapenemases

Carbapenemase-producing Gram-negative bacteria can cause a wide spectrum of infections including bacteremia, nosocomial pneumonia, wound infections, endocarditis, and urinary tract infections. Risk factors for carbapenem-resistant isolates usually include procedures involving manipulations of the gastrointestinal tract, trauma, orthopedic procedure, or life-threatening conditions like septicemia and septic shock. Those infections are often associated with high hospitalization cost, treatment failures, longer hospital stay, and high mortality rates [82, 83, 84].

Advertisement

5. Laboratory detection

Appropriate and timely laboratory detection of carbapenemase-producing microorganisms is of crucial importance to implement adequate antimicrobial therapy. Screening for carbapenemase production is usually based on reduced inhibition zones around carbapenem disks in routine disk-diffusion testing or elevated MICs for carbapenems. However, the level of resistance depends on the type of carbapenemase, allelic variant, gene expression, and/or the presence of additional resistance mechanisms such as porin loss or upregulation of efflux pumps.

Recommendations for carbapenem susceptibility breakpoints have been established by both Clinical Laboratory Standards Institute (CLSI) and The European Committee on Antimicrobial Susceptibility Testing (EUCAST), but no global consensus exists with EUCAST breakpoints being generally lower [85, 86]. However, some strains with low-level resistance are still being missed. EUCAST has also established epidemiologic breakpoints to distiguish between wildtype isolates and those with carbapenem-resistant determinants [86].

Ideally, methods for identifying carbapenemase should have a short turnaround time to ensure timely implementation of control measures. This could be challenged by difficulties in detecting carbapenemase producers, since MICs to carbapenems could be elevated but within susceptible range or even low, as described in Enterobacteriaceae and A. baumannii. Particularly in OXA-48-producing organisms, the MICs of carbapenemas can be very low. The screening usually relies on the reduction of the inhibition zone around ertapenem disk [87, 88]. However, relevant methodology with specific laboratory test has not yet been standardized. Modified Hodge test is the only test recommended by CLSI for the phenotypic detection of carbapenemase producers but often lacking sensitivity and specificity. There are also several inhibitor-based tests using different inhibitors (EDTA and phenantroline as inhibitors of MBLs, phenylboronic acid as inhibitor of KPC) in combination with carbapenem (e.g., meropenem) or cephalosporin (e.g., ceftazidime) in different formats—disk-diffusion or broth dilution or E-test [88]. There is no specific inhibitor that could be used in detection of class D carbapenemases, but there are reports on using temocillin disk (or combined with avibactam) for this purpose [88]. Carba NP (derived from the names Nordmann-Poirel) test is a simple biochemical test based on hydrolysis of imipenem detectable by a change of color of the indicator due to decrease of pH. It is applicable in most microbiological laboratories, although the reference standard in detection of carbapenemase production is spectrophotometric measurement of carbapenem hydrolysis in the presence or absence of inhibitor, but it is still reserved for reference laboratories [89]. Recently, a new method for the detection of carbapenemases was described [90]. The test was called carbapenem inactivation method (CIM), and it was based on degradation of meropenem by carbapenemase. The water suspension of carbapenem-resistant microorganism is incubated with a 10-ug meropenem disk for 2 h. After incubation, the disk is removed from the suspension and placed onto a Mueller-Hinton agar plate previously inoculated with the carbapenem-sensitive microorganism (usually Escherichia coli). If the carbapenem-resistant microorganism produces carbapenemase, consequently, meropenem will be hydrolyzed and the indicator (carbapenem susceptible) microorganism will grow close to the disk [90]. CIM method has high sensitivity, comparable to that of Carba NP test, as reported recently [91]. Modified CIM test includes preparation of the bacterial suspension in tryptic soy broth and extending the time of incubation to 4 hours improving the detection of some carbapenemases in Enterobacteriaceae [92]. Recently, the use of mass spectrometry (matrix-assisted laser desorption/ionization-time of flight—MALDI-TOFF) based on analysis of degradation of carbapenem-molecule-enabled rapid detection of KPC carbapenemase (in 45 min) or MBL (150 min) [93]. Finally, simplex or multiplex PCR, real-time PCR, or hybridization tests could significantly improve detection of carbapenemase genes in clinical laboratory bypassing the sensitivity and specificity problems with phenotypic tests [89]. However, molecular methods require expensive equipment and trained laboratory stuff.

Advertisement

6. Therapeutic options

Colistin is usually the last resort antibiotic for the treatment of infections associated with carbapenemase-producing Gram-negative bacteria. However, nephrotoxicity and neurotoxicity are of clinical concern [94]. Colistin-resistant isolates of A. baumannii, P. aeruginosa, and K. pneumoniae have emerged recently leading to pan-drug resistant phenotypes with modified outer membrane lipopolysaccharide [95, 96, 97]. The optimal treatment approach for infection caused by carbapenemase-producing Gram-negative bacilli remains a controversy. Still, a combination therapy is strongly recommended and includes two or more antibiotics active in vitro such as colistin, tigecycline, amikacin, aztreonam, or carbapenem in different combination schemes [98, 99]. Colistin is not recommended as monotherapy because of the development of heteroresistance. The combinations of colistin with rifampicin, vancomycin, and meropenem were shown to be synergistic in A. baumannii in vitro by chequerboard, 2-well and time-kill metod, but the results of the randomized, controlled clinical trials are controversial [99, 100, 101]. If the pathogen is suspected to be MBL producer, aztreonam may present the core drug (if the isolate does not possess other extended-spectrum or ampC β-lactamase), but if the patient is critically ill, the combination therapy with three drugs is preferred. If aminoglycosides or fosfomycin retain in vitro activity, for any serious infection, then they should not be using monotherapy. However, oral fosfomycin should not be used for the management of any infection outside the urinary tract, and in critically ill patients, combination therapy should be considered. Tigecycline is recommended for skin and soft-tissue infection with carbapenemase-producing A. baumannii and Enterobacteriaceae. If MICs to meropenem are low or moderately high (up to 16 μg/ml), high-dose meropenem can be administered by prolonged infusion [99, 100]. Dual carbapenem-based regimen for carbapenem-resistant Enterobacteriaceae usually includes high-dose meropenem or high-dose doripenem with ertapenem, and it is also most effective if the third drug is added [99]. In bloodstream and respiratory tract infections colistin could be added, while in gastrointestinal and biliary tract infections colistin and tigecycline should be included. Newer β-lactam-β-lactamase inhibitor combinations such as ceftolozan/tazobactam seem promising in treatment of carbapenem-resistant Pseudomonas aeruginosa infections [102]. However, it is not effective against MBL-producing organisms. Continued reports should be acknowledged in efforts of optimizing therapy for infections caused by carbapenem-resistant Gram-negative bacili. Furthermore, there is an urgent need to develop new antimicrobials. Although antibiotics still represent the mainstream direction for the treatment, the alternatives to them should be taken into account, including vaccines, bacteriocins, and probiotics. The revival of old antibiotics should be supported by expanding the knowledge of their pharmacokinetics and pharmacodynamics in order to avoid side effects and resistance developments.

Advertisement

7. Conclusions

In the last decade, an epidemic spread of carbapenamases among Enterobacteriaceae and Gram-negative nonfermentative bacilli was observed worldwide. The type of the most prevalent carbapenemase in a geographic region differs and might be associated with historical or cultural connections and exchange of people between countries where certain types of carbapenemase are endemic including the transfer of patients or staff across the borders, medical tourism, and migration of refugees. Particularly alarming is the fact that carbapenemases are not restricted only to hospital isolates any more. They have been reported and are continuously circulating between hospitals, long-term care facilities, community, and the environment. In conclusion, facing the global dissemination of carbapenemase-producing Gram-negative bacteria, and awaiting for the new antimicrobials to appear, a reasonable approach including hand hygiene, contact precautions, rational antibiotic usage, and active surveillance should be part of an intensive strategy aimed to reduce the incidence of colonization and infections with carbapenem-resistant microorganisms and to stop the spread of these bacteria into the community or the environment and vice versa.

References

  1. 1. El-Gamal MI, Brahim I, Hisham N, Aladdin R, Mohammed H, Bahaaeldin A. Recent updates of carbapenem antibiotics. European Journal of Medicinal Chemistry. 2017;131:185-195
  2. 2. Queenan AM, Bush K. Carbapenemases: The versatile β-lactamases. Clinical Microbiology Reviews. 2007;20(3):440-458
  3. 3. Nordmann P, Poirel L. Emerging carbapenemases in gram-negative aerobes. Clinical Microbiology and Infection. 2002;8:321-331
  4. 4. Meletis G. Carbapenem resistance: Overview of the problem and future perspectives. Therapeutic Advances in Infectious Disease. 2016;3(1):15-21
  5. 5. Diene SM, Rolain J-M. Carbapenemases genes and genetic platforms in gram-negative bacilli: Enterobacteriaceae, pseudomonas and Acinetobacter species. Clinical Microbiology and Infection. 2014;20:831-838
  6. 6. Bush K, Jacoby GA, Medeiros AA. A functional classification scheme for β-lactamases and its correlation with molecular structure. Antimicrobial Agents and Chemotherapy. 1995;39:1211-1233
  7. 7. Bush K, Jacoby GA. Updated functional classification of β-lactamases. Antimicrobial Agents and Chemotherapy. 2010;54:969-976
  8. 8. Naas T, Vandel W, Sougakoff W, Livermore DM, Nordmann P. Cloning and sequence analysis of the gene for carbapenem hydrolyzing class a β-lactamase, Sme-1 from Serratia marcescens S6. Antimicrobial Agents and Chemotherapy. 1994;38:1262-1270
  9. 9. Gales AC, Biedenbach DJ, Winokur P, Pfaller A, Jones RN. Carbapenem-resistant Serratia marcescens isolates producing group 2F β-lactamase (SME-2) in the unites states: Results from the mystic program. Diagnostic Microbiology and Infectious Disease. 2001;39:125-127
  10. 10. Queenan AM, Shang W, Schreckenberg P, Lolans K, Bush K, Quinn J. SME-3 a novel member of the Serratia marcescens SME family of carbapenem-hydrolyzing β-lactamases. Antimicrobial Agents and Chemotherapy. 2006;50:3485-3487
  11. 11. Nordmann P, Mariotte s NT, Labia R, Nicolas MH. Biochemical properties of a carbapenem-hydrolyzing β-lactamase of Enterobacter cloacae and cloning of the gene into Escherichia coli. Antimicrobial Agents and Chemotherapy. 1993;37:939-946
  12. 12. Prottumarthy S, Moland ES, Jeretschko S, Swanzy R, Thomson KS, Fritsche TR. NMC-A carbapenem-hydrolyzing enyzme in Enterobacter cloacae in North America. Emerging Infectious Diseases. 2003;9:999-1002
  13. 13. Yigit H, Quennan AM, Anderson GJ, et al. Novel carbapenem-hydrolyzing β-lactamase, KPC-1, from a carbapenem-resistant strain of Klebsiella pneumoniae. Antimicrobial Agents and Chemotherapy. 2001;45(4):1151-1161
  14. 14. Woodford NP, Tierno PM, Young K, Tysal M, Palepou E, Ward R, Painter E, et al. Outbreak of Klebsiella pnemoniae producing a new carbapenem hydrolyzing class a β-lactamase KPC-3 in a New York medical center. Antimicrobial Agents and Chemotherapy. 2004;48:4793-4799
  15. 15. Yigit H, Quennan AM, Anderson GJ, et al. Author’s correction. Novel carbapenem-hydrolyzing β-lactamase, KPC-1, from a carbapenem-resistant strain of Klebsiella pneumoniae. Antimicrobial Agents and Chemotherapy. 2008;52:809
  16. 16. Poirel L, Weldhagen GF, Naas T, De Champs C, Dove G, Nordmann P. GES-2 a novel class a β-lactamase with increased hydrolysis of imipenem. Antimicrobial Agents and Chemotherapy. 2001;45:2598-2603
  17. 17. Naas T, Dortet L, Iorga BI. Structural and functional aspects of class A carbapenemases. Current Drug Targets. 2016;17(9):1006-1028
  18. 18. Mathers AJ, Stoesser N, Chai W, et al. Chromosomal integration of the Klebsiella pneumoniae carbapenemase gene blaKPC in Klebsiella species is elusive but not rare. Antimicrobial Agents and Chemotherapy. 2017;61(3):1-11. pii: e01823-16. DOI: 10.1128/AAC.01823-16
  19. 19. Nordmann P, Cuzon G, Naas T. The real threat of Klebsiella pneumoniae carbapenemase-producing bacteria. The Lancet Infectious Diseases. 2009;9(4):228-236
  20. 20. Parisi SG, Bartolini A, Santacatterina E, et al. Prevalence of Klebsiella pneumoniae strains producing carbapenemases and increase of resistance to colistin in an Italian teaching hospital from January 2012 to December 2014. BMC Infectious Diseases. 2015;15:244. DOI: 10.1186/s12879-015-0996-7
  21. 21. Schwaber MJ, Lev B, Israeli A, Solter E, Smollan G, Rubinovitch B, Shalit I, Carmeli Y, Israel Carbapenem-Resistant Enterobacteriaceae Working Group. Containment of a country-wide outbreak of carbapenem-resistant Klebsiella pneumoniae in Israeli hospitals via a nationally implemented intervention. Clinical Infectious Diseases. 2011;52(7):848-855. DOI: 10.1093/cid/cir025
  22. 22. Wei DD, Wan LG, Deng Q , Liu Y. Emergence of KPC-producing Klebsiella pneumoniae hypervirulent clone of capsular serotype K1 that belongs to sequence type 11 in Mainland China. Diagnostic Microbiology and Infectious Disease. 2016;85(2):192-194. DOI: 10.1016/j.diagmicrobio.2015.03.012
  23. 23. Van Duin D, Doi Y. The global epidemiology of carbapenemase-producing Enterobacteriacea. Virulence. 2017;8:460-469
  24. 24. Pitout JD, Nordmann P, Poirel L. Carbapenemase-producing Klebsiella pneumoniae, a key pathogen set for global nosocomial dominance. Antimicrobial Agents and Chemotherapy. 2015;59(10):5873-5884
  25. 25. Poirel L, Thomas IL, Naas T, Karim A, Nordman P. Biochemical sequence analysis of GES-1, a novel class a extended-spectrum β-lactamase and the class I integron In52 from Klebsiella pneumoniae. Antimicrobial Agents and Chemotherapy. 2000;44:622-632
  26. 26. Poirel L, Weldhagen GF, De Champs C, Nordmann P. A nosocomial outbreak of Pseudomonas aeruginosa isolates expressing the extended-spectrum beta-lactamase GES-2 in South Africa. The Journal of Antimicrobial Chemotherapy. 2002;49(3):561-565
  27. 27. Walsh T, Toleman MA, Poirel L, Nordmann P. Metallo-β-lactamases: The quiet before the storm. Clinical Microbiology Reviews. 2005;18(2):306-325
  28. 28. Berglund F, Marathe NP, Österlund T, Bengtsson-Palme J, Kotsakis S, Flach CF, Larsson DGJ, Kristiansson E. Identification of 76 novel B1 metallo-β-lactamases through large-scale screening of genomic and metagenomic data. Microbiome. 2018;5(1):134. DOI: 10.1186/s40168-017-0353-8
  29. 29. Watanabe M, Iyobe S, Inoue M, Mitsuhashi S. Transferable imipenem resistance in Pseudomonas aeruginosa. Antimicrobial Agents and Chemotherapy. 1991;35:147-151
  30. 30. Arakawa Y, Murakami M, Suzuki K, et al. A novel integron like-element carrying the metallo-β-lactamase gene blaIMP. Antimicrobial Agents and Chemotherapy. 1995;39:1612-1615
  31. 31. Riccio CG, ML MA, Lauretti L, Fontana R, Rossolini GM. Appearance of IMP-1 metallo-β-lactamase in Europe. Lancet. 1999;353:899-900
  32. 32. Riccio ML, Franceschini N, Boschi L, et al. Characterization of the metallo-β-lactamase determinant of Acinetobacter baumannii AC-54/97 reveals the existance of blaIMP allelic variant carried by gene cassettes of different phylogeny. Antimicrobial Agents and Chemotherapy; 44:1229-1235
  33. 33. Matsumura Y, Peirano G, Motyl MR, Adams MD, et al. Global molecular epidemiology of IMP-producing enterobacteriaceae. Antimicrobial Agents and Chemotherapy. 2017;61(4). DOI: 10.1128/AAC.02729-16
  34. 34. Iyobe S, Kusadokoro H, Ozaki J, Matsumura N, Minami S, Haruta S, Sawai T, O’Hara K. Amino acid substitutions in a variant of IMP-1 metallo-β-lactamase. Antimicrobial Agents and Chemotherapy. 2000;44(8):2023-2027
  35. 35. Chu YW, Afzal Shah M, Huang E, Palepou MF, Lyon DJ, Woodford N, Livermore DM. IMP-4, a novel metallo-β-lactamase from nosocomial Acinetobacter spp. collected in Hong-Kong between 1994 and 1998. Antimicrobial Agents and Chemotherapy. 2001;45(3):710-714
  36. 36. Da Silva GJ, Correia M, Vital C, Ribero G, Sousa JC, Leitao R, Peixe L, Duarte A. Molecular characterization of bla(IMP-5) a new integron born metallo-β-lactamase gene from an Acinetobacter baumannii nosocomial isolate in Portugal. FEMS Microbiology Letters. 2002;24(215):33-39
  37. 37. Available from: ftp://ftp.ncbi.nlm.nih.gov/pathogen/βlactamases/Allele.tab [Accessed: 2018-03-06]
  38. 38. Mojica MF, Bonomo RA, Fast W. B1-metallo-β-lactamases: Where do we stand? Current Drug Targets. 2016;17(9):1029-1050
  39. 39. Tada T, Nhung PH, Miyoshi-Akiyama T, Shimada K, Phuong DM, Anh NQ , Ohmagari N, Kirikae T. IMP-51, a novel IMP-type metallo-β-lactamase with increased doripenem and meropenem hydrolyzing activities, in a carbapenem-resistant Pseudomonas aeruginosa clinical isolate. Antimicrobial Agents and Chemotherapy. 2015;59:7090-7093
  40. 40. Cornaglia G, Giamarellou H, Rossolini GM. Metallo-β-lactamases: A last frontier for β-lactams? The Lancet Infectious Diseases. 2011;11:381-393
  41. 41. Carattoli A. Resistance plasmid families in Enterobacteriaceae. Antimicrobial Agents and Chemotherapy. 2009;53(6):2227-2238. DOI: 10.1128/AAC.01707-08
  42. 42. Lauretti L, Riccio ML, Mazzariol A, Cornaglia G, Amicosante G, Fontana R, Rossolini GM. Cloning and characterization of blaVIM, a new integron-borne metallo- β-lactamase gene from Pseudomonas aeruginosa clinical isolate. Antimicrobial Agents and Chemotherapy. 1999;43(7):1584-1590
  43. 43. Poirel L, Naas T, Nicolas D, Collet L, Bellais S, Cavallo JD, Nordmann P. Characterization of VIM-2, a carbapenem-hydrolyzing metallo-β-lactamase and its plasmid and integron-borne gene from Pseudomonas aeruginosa clinical isolate in France. Antimicrobial Agents and Chemotherapy. 2000;44:891-897
  44. 44. Yan JH, Po-Ren H, Wen-Chien K, et al. Metallo-β-lactamases in clinical Pseudomonas isolates in Taiwan and identification of VIM-3, a novel variant of the VIM-2 enzyme. Antimicrobial Agents and Chemotherapy. 2001;458:2224-2228
  45. 45. Luzzaro F, Docquier JD, Colinon C, et al. Emergence in Klebsiella pneumoniae and Enterobacter cloacae clinical isolates of the VIM-4 metallo β-lactamase encoded by a conjugative plasmid. Antimicrobial Agents and Chemotherapy. 2004;48:648-650
  46. 46. Zhao W-H, Hu Z-Q. Epidemiology and genetics of VIM-type metallo-β-lactamase in Gram-negative bacilli. Future Microbiology. 2011;6:317-333
  47. 47. Toleman MA, Rolson K, Jones RN, Walsh T. BlaVIM-7 and evolutionary distinct metallo-β-lactamase gene in a Pseudomonas aeruginosa isolate from the United States. Antimicrobial Agents and Chemotherapy. 2004;48:329-332
  48. 48. Samuelsen O, Castanheira M, Walsh TR, Spencer J. Kinetic characterization of VIM-7, a divergent member of the VIM metallo-beta-lactamase family. Antimicrobial Agents and Chemotherapy. 2008;52(8):2905-2908. DOI: 10.1128/AAC.00166-08
  49. 49. Kumar V, Sen MR, Nigam C, Gahlot R, Kumari S. Burden of different beta-lactamase classes among clinical isolates of AmpC-producing Pseudomonas aeruginosa in burn patients: A prospective study. Indian Journal of Critical Care Medicine. 2012;16(3):136-140
  50. 50. Oelschlaeger P, Ai N, Duprez KT, Welsh WJ, Toney JH. Evolving carbapenemases: Can medicinal chemists advance one step ahead of the coming storm? Journal of Medicinal Chemistry. 2010;53:3013-3027
  51. 51. Yong D, Toleman MA, Giske CG, Cho HS, Sundman K, Lee K, Walsh TR. Characterization of a new metallo-β-lactamase gene, blaNDM-1, and a novel erythromycin esterase gene carried on a unique genetic structure in Klebsiella pneumoniae sequence type 14 from India. Antimicrobial Agents and Chemotherapy. 2009;53(12):5046-5054
  52. 52. Walsh TR, Weeks J, Livermore DM, Toleman MA. Dissemination of NDM-1 positive bacteria in New Delhi environment and its implications for human health: An environmental point prevalence study. The Lancet Infectious Diseases. 2011;11:355-362
  53. 53. Lee C-R, Lee JH, Park KS, Kim YB, Jeong BC, Lee SH. Global dissemination of carbapenemase producing Klebsiella pneumoniae: Epidemiology, genetic context, treatment options and detection methods. Frontiers in Microbiology. 2016;7:895. DOI: 10.3389/frmicb.2016.00895
  54. 54. Khan AU, Maryam L, Zarrilli R. Structure, genetics and worldwide spread of New Delhi metallo-β-lactamase (NDM): A treat to public health. BMC Microbiology. 2017;17:101. DOI: 10.1186/s12866-017-1012-8
  55. 55. Kashyap A, Gupta R, Sharma R, Verma VV, Gupta S, Pradeep G. New Delhi metallo beta lactamase: Menace and its challenges. Journal of Molecular and Genetic Medicine. 2017;11:4. DOI: 10.4172/1747-0862.1000299
  56. 56. Kazmiercak KM, Rabine S, Hackel M, McLaughlin RE, Biedenbach DJ, Bouchillon SK, Sahm DF, Bradford PA. Multiyear, multinational survey of the incidence and global distribution of metallo-β-lactamase-producing Enterobacteriaceae and Pseudomonas aeruginosa. Antimicrobial Agents and Chemotherapy. 2015;60(2):1067-1078. DOI: 10.1128/AAC.02379-15
  57. 57. Toleman MA, Simm AM, Murphy TA, Gales AC, Biedenbach DJ, Jones RN, Walsh TR. Molecular characterization of SPM-1, a novel metallo β-lactamase isolated in Latin America: Report from the SENT antimicrobial surveillance programme. The Journal of Antimicrobial Chemotherapy. 2002;50:673-679
  58. 58. Castanheira M, Toleman MA, Jones RN, Schmidt FJ, Walsh TR. Molecular characterization of β-lactamase gene, blaGIM encoding a new subclass of metallo β-lactamase. Antimicrobial Agents and Chemotherapy. 2004;48:4654-4661
  59. 59. Lee K, Yum JH, Yong D, Chong J, Kim JM, Docquirer JD, Rossolini GM, Chong Y. Novel acquired metallo-β-lactamase gene, blaSIM-1, in a class 1 integron from Acinetobacter baumannii clinical isolates, from Korea. Antimicrobial Agents and Chemotherapy. 2005;49:4485-4491
  60. 60. Deshpande LM, Jones RN, Woosley LN, Castanheira M. Retrospective molecular analysis of DIM-1 metallo-β-lactamase discovered in Pseudomonas stutzeri from India in 2000. Antimicrobial Agents and Chemotherapy. 2014;58:596-598
  61. 61. Salabi AE, Toleman MA, Weeks J, Bruderer T, Frei R, Walsh TR. First report of the metallo-beta-lactamase SPM-1 in Europe. Antimicrobial Agents and Chemotherapy. 2010;54:582
  62. 62. Hopkins KL, Meunier D, Findlay J, Mustafa N, Parsons H, Pike R, Wright L, Woodford N. SPM-1 metallo-β-lactamase-producing Pseudomonas aeruginosa ST277 in the UK. Journal of Medical Microbiology. 2016;65(7):696-697. DOI: 10.1099/jmm.0.000269
  63. 63. Shahcheraghi F, Abbasalipour M, Feizabadi M, Ebrahimipour G, Akbari N. Isolation and genetic characterization of metallo-β-lactamase and carbapenamase producing strains of Acinetobacter baumannii from patients at Tehran hospitals. Iranian Journal of Microbiology. 2011;3:68-74
  64. 64. Brown S, Amyes S. OXA β-lactamase in Acinetobacter: The story so far. The Journal of Antimicrobial Chemotherapy. 2006;57:1-3
  65. 65. Evans BA, Amyes SG. OXA β-lactamases. Clinical Microbiology Reviews. 2014;27:241-263. DOI: 10.1128/CMR.00117-13
  66. 66. Da Silva GJ, Domingues S. Insights on the horizontal gene transfer of carbapenemase determinants in the opportunistic pathogen Acinetobacter baumannii. Microorganisms 2016;4:29. doi:10.3390/microorganism4030029
  67. 67. Paton RH, Miles RS, Hood J, Amyes SGB. ARI-1: β-lactamase mediated imipenem resistance in Acinetobacter baumannii. International Journal of Antimicrobial Agents. 1993;2:81-88
  68. 68. Bou G, Oliver A, Martinez-Beltran J. OXA-24, a novel class D β-lactamase with carbapenemase activity in an Acinetobacter baumannii clinical strain. Antimicrobial Agents and Chemotherapy. 2000;44:1556-1561
  69. 69. Marque S, Poirel L, Heritier C, et al. Regional occurrence of plasmid-mediated carbapenem-hydrolyzing oxacillinase OXA-58 in Acinetobacter spp. in Europe. Journal of Clinical Microbiology. 2005;43:4885-4888
  70. 70. Higgins P, Poirel L, Lehmann M, Nordmann P, Seifert H. OXA-143, a novel carbapenem-hydrolyzing class D β-lactamase in Acinetobacter baumannii. Antimicrobial Agents and Chemotherapy. 2009;53:5035-5038
  71. 71. Goić-Barišić I, Towner KJ, Kovačić A, Šiško-Kraljević K, Tonkić M, Novak A, Punda-Polić V. Outbreak in Croatia caused by a new carbapenem-resistant clone of Acinetobacter baumannii producing OXA-72 carbapenemase. Journal of Hospital Infection. 2011;77:368-369
  72. 72. Wang H, Guo P, Sun H, Wang H, Yang Q , Chen M, Xu Y, Zhu Y. Molecular epidemiology of clinical isolates of carbapenem-resistant Acinetobacter spp. from Chinese hospitals. Antimicrobial Agents and Chemotherapy. 2007;51:4022-4028
  73. 73. Werneck JS, Picao RC, Carvalhaes CG, Cardoso JP, Gales A. OXA-72-producing Acinetobacter baumannii in Brazil: A case report. The Journal of Antimicrobial Chemotherapy. 2011;66(2):452-454
  74. 74. Bedenić B, Beader N, Godič-Torkar K, Vranić-Ladavac M, Luxner J, Veir Z, Grisold AJ, Zarfel G. Nursing home as reservoir of carbapenem-resistant Acinetobacter baumannii. Microbial Drug Resistance. 2015;21(3):270-278
  75. 75. Hrenović J, Goić-Barišić I, Kazazic S, Kovačić A, Ganjto M, Tonkić M. Carbapenem-resistant isolates of Acinetobacter baumannii in a municipal wastewater treatment plant, Croatia, 2014. Euro Surveillance. 2016;21(15):pii=30195. DOI: http://dx.doi.org/10.2807/1560-7917.ES.2016.21.15.30195
  76. 76. Goić-Barišić I, Hrenović J, Kovačić A, Šeruga-Musić M. Emergence of oxacillinases in environmental carbapenem-resistant Acinetobacter baumannii associated with clinical isolates. Microbial Drug Resistance. 2016;22(7):559-563
  77. 77. Poirel L, Heritier C, Tolun V, Nordmann P. Emergence of oxacillinases-mediated resistance to imipenem in Klebsiella pneumoniae. Antimicrobial Agents and Chemotherapy. 2004;48(1):15-22
  78. 78. Poirel L, Potron A, Nordmann P. OXA-48-like carbapenemases: The phantom menace. Journal of Antimicrobial Chemotherapy. 2012;67(7):1597-1606. DOI: 10.1093/jac/dks121
  79. 79. Gulmez D, Woodford N, Palepou MF, et al. Carbapenem resistant Escherichia coli and Klebsiella pneumoniae isolates from Turkey with OXA-48 like carbapenemase and outer membrane protein loss. International Journal of Antimicrobial Agents. 2008;31(6):523
  80. 80. Poirel L, Ros A, Carricajo A, et al. Extremely drug-resistant Citrobacter freundii isolate producing NDM-1 and other carbapenemases identified in a patient returning from India. Antimicrobial Agents and Chemotherapy. 2011;55:447-448
  81. 81. Potron A, Poirel L, Nordmann P. Origin of OXA-181, an emerging carbapenem-hydrolyzing oxacillinase, as a chromosomal gene in Shewanella xiamenensis. Antimicrobial Agents and Chemotherapy. 2011;55:4405-4407
  82. 82. Borer A, Saidel-Odes L, Riesenberg K, Eskira S, Peled N, Nativ R, Schlaeffer F, Sherf M. Attributable mortality rate for carbapenem-resistant Klebsiella pneumoniae bacteremia. Infection Control and Hospital Epidemiology. 2009;30:972-976
  83. 83. Zilberberg MD, Nathanson BH, Sulham K, Fan W, Shorr AF. Carbapenem resistance, inappropriate empiric treatment and outcomes among patients hospitalized with Enterobacteriaceae urinary tract infection, pneumonia and sepsis. BMC Infectious Diseases. 2017;17:279. DOI: 10.1186/s12879-017-2383-z
  84. 84. Akova DGL, Tzouvelekis L, Carmeli Y. Interventional strategies and current clinical experience with carbapenemase-producing gram-negative bacteria. Clinical Microbiology and Infection. 2012;18:439-448
  85. 85. Clinical and Laboratory Standards Institute. Performance Standards for Antimicrobial Susceptibility Testing. 27th ed. CLSI Document M100S. Wayne, PA, USA: CLSI; 2017
  86. 86. The European Committee on Antimicrobial Susceptibility Testing. Breakpoint Tables for Interpretation of MICs and Zone Diameters. Version 7.1. 2017. Available online: http://www.eucast.org [Accessed: 2018-03-06]
  87. 87. Miriagou V, Cornaglia G, Edelstein M, et al. Acquired carbapenemases in gram-negative bacterial pathogens: Detection and surveillance issues. Clinical Microbiology and Infection. 2010;16:112-122
  88. 88. Nordmann P, Gniadkowski M, Giske CG, Poirel L, Woodford N, Miriagou V, European Network on Carbapenemases. Identification and screening of carbapenemase-producing enterobacteriaceae. Clinical Microbiology and Infection. 2012;18:432-438
  89. 89. Madkour LA, Soliman MS, Hassan DM, Soliman NS, ElMahdy YA. Detection of carbapenemase-producers: Evaluating the performance of the carbapenem inactivation method and Carba NP test versus multiplex PCR. Journal of Global Antimicrobial Resistance. 2017;9:10-14
  90. 90. Aguirre-Quiñonero A, Cano ME, Gamal D, Calvo J, Martínez-Martínez L. Evaluation of the carbapenem inactivation method (CIM) for detecting carbapenemase activity in enterobacteria. Diagnostic Microbiology and Infectious Disease. 2017;88:214-218
  91. 91. Sun K, Xu X, Yan J, Zhang L. Evaluation of six phenotypic methods for the detection of carbapenemases in gram-negative bacteria with characterized resistance mechanisms. Annals of Laboratory Medicine. 2017;37(4):305-312
  92. 92. Pierce VM, Simner PJ, Lonsway DR, et al. Modified carbapenem inactivation method for phenotypic detection of carbapenemase production among enterobacteriaceae. Journal of Clinical Microbiology. 2017;55(8):2321-2333
  93. 93. Johansson A, Ekelöf JJ, Giske CG, Sundqvist M. The detection and verification of carbapenemases using ertapenem and matrix assisted laser desorption ionization-time of flight. BMC Microbiology. 2014;14:89. DOI: 10.1186/1471-2180-14-89
  94. 94. Michalopoulos A, Falagas ME. Colistin and polymyxin B in critical care. Critical Care Clinics. 2008;466(24):377-391
  95. 95. Cai Y, Chai D, Wang R, Liang B, 638 Bai N. Colistin resistance of Acinetobacter baumannii: Clinical reports, mechanisms and antimicrobial strategies. The Journal of Antimicrobial Chemotherapy. 2012;67:1607-1615
  96. 96. Ko KS, Suh JY, Kwon KT, Jung S-I, et al. High rates of resistance to colistin and polymyxin B in subgroups of Acinetobacter baumannii isolates from Korea. The Journal of Antimicrobial Chemotherapy. 2007;60:1163-1167
  97. 97. Helander IM, Kato Y, Kilpeläinen I, Kostiainen R, Lindner B, Nummila K, et al. Characterization of lipopolysaccharides of polymyxin-resistant and polymyxin-sensitive Klebsiella pneumoniae. European Journal of Biochemistry. 1996;23:272-278
  98. 98. Paul M, Carmeli Y, Durante-Mangoni E, et al. Combination therapy for carbapenem-resistant gram-negative bacteria. The Journal of Antimicrobial Chemotherapy. 2014;69(9):2305
  99. 99. Trecarichi EM, Tumbarello M. Therapeutic options for carbapenem-resistant Enterobacteriaceae infections. Virulence. 2017;8:470-484
  100. 100. Gilad J, Carmely Y. Treatment options for multidrug-resistant Acinetobacter spp. Drugs. 2008;68:165-189
  101. 101. Petrosillo N, Ionnidou E, Falagas ME. Colistin monotherapy vs. combination therapy: Evidence from microbiological, animal and clinical studies. Clinical Microbiology and Infection. 2008;14:816-827
  102. 102. Munita JM, Aitken SL, Miller WR, et al. Multicenter evaluation of ceftolozane/tazobactam for serious infections caused by carbapenem-resistant Pseudomonas aeruginosa. Clinical Infectious Diseases. 2017;65:158-161

Written By

Branka Bedenić and Sanda Sardelić

Submitted: 15 January 2018 Reviewed: 07 March 2018 Published: 05 November 2018