Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2015 Apr 7.
Published in final edited form as: Diagn Microbiol Infect Dis. 2013 Oct 3;77(3):179–194. doi: 10.1016/j.diagmicrobio.2013.06.001

Non-Phenotypic Tests to Detect and Characterize Antibiotic Resistance Mechanisms in Enterobacteriaceae

Agnese Lupo 1, Krisztina M Papp-Wallace 2,3, Parham Sendi 1,4, Robert A Bonomo 2,3,5, Andrea Endimiani 1,*
PMCID: PMC4388032  NIHMSID: NIHMS604684  PMID: 24091103

Abstract

In the past two decades, we have observed a rapid increase of infections due to multidrug-resistant Enterobacteriaceae. Regrettably, these isolates possess genes encoding for extended-spectrum β-lactamases (e.g., blaCTX-M, blaTEM, blaSHV) or plasmid-mediated AmpCs (e.g., blaCMY) that confer resistance to last-generation cephalosporins. Furthermore, other resistance traits against quinolones (e.g., mutations in gyrA and parC, qnr elements) and aminoglycosides (e.g., aminoglycosides modifying enzymes and 16S rRNA methylases) are also frequently co-associated. Even more concerning is the rapid increase of Enterobacteriaceae carrying genes conferring resistance to carbapenems (e.g., blaKPC, blaNDM). Therefore, the spread of these pathogens puts in peril our antibiotic options. Unfortunately, standard microbiological procedures require several days to isolate the responsible pathogen and to provide correct antimicrobial susceptibility test results. This delay impacts the rapid implementation of adequate antimicrobial treatment and infection control countermeasures. Thus, there is emerging interest in the early and more sensitive detection of resistance mechanisms. Modern non-phenotypic tests are promising in this respect, and hence, can influence both clinical outcome and healthcare costs. In this review, we present a summary of the most advanced methods (e.g., next-generation DNA sequencing, multiplex PCRs, real-time PCRs, microarrays, MALDITOF MS, and PCR/ESI MS) presently available for the rapid detection of antibiotic resistance genes in Enterobacteriaceae. Taking into account speed, manageability, accuracy, versatility, and costs, the possible settings of application (research, clinic, and epidemiology) of these methods and their superiority against standard phenotypic methods are discussed.

Keywords: PCR, multiplex, real-time, microarray, ESBL, carbapenemases, MALDI-TOF

1. INTRODUCTION

Microorganisms belonging to the family of Enterobacteriaceae (Ent) are important causes of community- and hospital-acquired infections. Escherichia coli is a frequent cause of urinary tract infections (UTIs), Klebsiella spp. and Enterobacter spp. are important causes of nosocomial pneumonia, and all of the Ent have been implicated in bloodstream (BSIs) and intra-abdominal infections (Donnenberg, 2005; Gaynes & Edwards, 2005).

Serious infections due to Ent are usually treated with extended-spectrum cephalosporins (ESCs) included in the third- and fourth-generations (Endimiani & Paterson, 2007; Michalopoulos & Falagas, 2010). These drugs target a broad-spectrum of pathogens, possess low toxicity, and demonstrate favorable pharmacokinetics and pharmacodynamics (Grayson, 2010). Unfortuantely, there has been a rapid increase of infections due to ESC-resistant (ESC-R) Ent (Coque, Baquero, & Canton, 2008; Meyer, Schwab, Schroeren-Boersch, & Gastmeier, 2010). As a consequence, quinolones and aminoglycosides are the alternative antibiotics considered, but high co-associated resistance rates are also reported (Giamarellou & Poulakou, 2009; Hawser, et al., 2010). Although it is understandable from a clinical perspective to use then carbapenems, it is equally worrisome that this practice leads to the rapid selection of carbapenem-resistant Ent (Walsh, 2010). As a result, only a few antimicrobial agents (e.g., colistin, fosfomycin, tigecycline) with an uncertain in vivo efficacy and/or reported toxicity are usually left to treat infections due to multidrug-resistant Ent (MDR-Ent) (Livermore, et al., 2011). Thus, the rapid characterization of the resistance mechanism(s) is essential to establish a correct antibiotic therapy and to implement measures to contain the spread of these life-threatening pathogens.

1.1 Why should we rapidly detect antibiotic resistance mechanisms in Ent

Standard microbiological procedures require several days to isolate the causing pathogen(s) responsible for the syndrome and to provide the antimicrobial susceptibility test (AST) results. For instance, for the diagnosis of bacteremia clinical laboratories usually implement automated systems capable of monitoring microbial growth in blood culture bottles (Riedel & Carroll, 2010). When the bottles are positive (most Ent grow within 2 days after incubating), the sample is plated on standard agar media (Endimiani, Tamborini, Luzzaro, Lombardi, & Toniolo, 2002). If colonies grow (usually after overnight incubation), ASTs are performed. Generally, these tests require at least additional 24 hrs to yield results. Hence, for non-MDR-Ent, a total time from primary incubation to test results (TTR) of 3-4 days is often required. However, ASTs can be inaccurate for certain MDR-Ent(Livermore, et al., 2012)causing mischaracterization of carbapenemase (Poirel, Potron, & Nordmann, 2012), extended-spectrum β-lactamase (ESBL) (Luzzaro, et al., 2006), or plasmid-mediated AmpC (pAmpC) producers (Doi & Paterson, 2007) (see also section 1.2). These difficulties can be partially overcome by several adjunctive phenotypic tests that may include: i) clavulanate for ESBLs detection (Drieux, Brossier, Sougakoff, & Jarlier, 2008), ii) boronic acid for pAmpCs and KPCs detection (Pournaras, Poulou, & Tsakris, 2010), or iii) modified Hodge test (mHT) for detecting carbapenemases (Doyle, et al., 2012). Unfortunately, these tests require at least an additional day. Thus, up to 4-6 days are required to provide physicians with an accurate phenotype for MDR-Ent.

The delay in identifying and reporting AST results has important clinical implications. Patients with MDR-Ent infections have less favorable outcomes than those infected by non-MDR isolates (Schwaber, et al., 2006; Tumbarello, et al., 2007; Tumbarello, et al., 2010). Many studies have shown that: i) mortality attributable to BSI is significantly higher in patients infected with ESBL producers than those with non-ESBL producers; and ii) treatment failure rate for patients infected with ESBL-producing Ent (ESBL-Ent) is almost twice as high as that of the patients infected with non-ESBL-producing organisms (Endimiani, et al., 2005; Endimiani, et al., 2004). Hence, a rapid detection of MDR-Ent can help to identify patients at risk for poor outcome.

The delay has also economic implications. A rapid reporting system of antibiotic resistance patterns reduces the duration of hospitalization and, as consequence, the healthcare costs. For instance, the number of days using broad-spectrum antibiotics can be shortened (e.g., avoiding the use of the more expensive carbapenems). Lee et al. showed that the mean length of hospital stay (MLHS) for patients with ESBL producers was longer than that for patients infected with non-ESBL producers (21 vs. 11 days). The additional costs for an infection due to ESBL producers were $16,450 per patient (S. Y. Lee, Kotapati, Kuti, Nightingale, & Nicolau, 2006). Schwaber et al. noted that patients with BSIs due to ESBL-Ent had a MLHS post-infection of 11 days vs. 5 days in those with non-ESBL producers. The average hospital cost for BSIs caused by ESBL producers was $46,970 vs. $16,877 for those due to non-ESBL producers (Schwaber, et al., 2006).

The rapid identification of patients colonized with MDR-Ent can also assure a prompt implementation of hospital hygiene precautions to prevent the spread (e.g., outbreaks) of these pathogens. However, standard laboratory methods are too slow and may lack the required sensitivity. For instance, screening of intestinal carriers of ESBL- or carbapenemase-producing Ent is usually performed with commercially available selective agar plates that have a lower sensitivity compared to PCR-based methods. This is especially true for those isolates expressing low-level MICs for the antibiotics added to the agar (e.g., OXA-48-producing E. coli) (Nordmann, Gniadkowski, et al., 2012). In a clinical study, Singh et al. showed that a real-time PCR designed to detect blaKPC in rectal swabs demonstrated a sensitivity of 97%, whereas CHROMagar plates were only of 77%; overall, 13% of colonized subjects were not identified with standard cultures (Singh, et al., 2012). Naas T et al. demonstrated that a real-time PCR for blaNDM had a limit of detection of 10 CFU/ml, whereas for ChromID ESBL and CHROMAgar KPC plates it was 102-103 CFU/ml of stool (Naas, Ergani, Carrer, & Nordmann, 2011).

Based on the above examples, there is no doubt that early and more sensitive detection of resistance mechanisms by non-phenotypic tests has benefits. However, we also emphasize the importance of implementing advanced and rapid methodologies in other settings, such as the research context and/or when large collection of isolates should be screened before further and more specific molecular analyses.

Overview of the main mechanisms of antibiotic resistance in Ent

1.2.1 Resistance to extended-spectrum cephalosporins (ESCs).

Resistance to β-lactams in Ent may be due to mutations in the penicillin-binding proteins (PBPs) or reduced permeability of the cell wall (e.g., disruption of outer membrane proteins, OMPs). However, production of β-lactamases is the most frequent mechanism encountered in Ent. These periplasmic enzymes can be grouped into four classes (A to D) on the basis of their amino acid sequence homology (Bush & Jacoby, 2009).

The most clinically important class A enzymes are the ESBLs. They confer resistance to penicillins, ESCs and aztreonam but not to carbapenems, and are usually inhibited by the commercially available β-lactamase inhibitors (i.e., clavulanate, tazobactam, sulbactam). TEM, SHV, and CTX-M are the three main families of ESBLs described. While TEM and SHV ESBLs arise via substitutions in strategically positioned amino acids from the natural narrow-spectrum TEM-1/-2 or SHV-1 β-lactamase, all CTX-M variants demonstrate an ESBL phenotype (Gniadkowski, 2008; Paterson & Bonomo, 2005). Until the 1990s, most ESBLs identified were of SHV- and TEM-types. At present, the CTX-M (especially CTX-M-15) are the most prevalent ESBLs worldwide (Livermore, et al., 2007; Rossolini, D'Andrea, & Mugnaioli, 2008; Seiffert, Hilty, Perreten, & Endimiani, 2013). With a limited global, but a significant local impact, other ESBLs (e.g., PER, VEB, and GES-1) have also been reported (Akinci & Vahaboglu, 2010).

Several Ent possess genes encoding for class C chromosomal AmpCs (cAmpCs; e.g., Citrobacter freundii, Enterobacter spp., Serratia spp.). Such β-lactamases are under the control of a complex regulon (ampD, ampR) and confer resistance to third-generation cephalosporins and β-lactam/β-lactamase inhibitor combinations, but not to carbapenems (Bush & Jacoby, 2009; Harris & Ferguson, 2012; Hilty, et al., 2013; Jacoby, 2009). Also, E. coli possess a chromosomal blaAmpC but only mutations in the promoter/attenuator region can lead to constitutive hyper-expression of the gene resulting in cephalosporins resistance. However, these strains are only occasionally reported (Jorgensen, Nielsen, Friis-Moller, Fjeldsoe-Nielsen, & Schonning, 2010).

An increasing number of blapAmpC genes have been observed among Ent. These enzymes belong to several families (i.e., CMY, FOX, LAT, MIR, ACT, DHA, ACC, MOX) and are derived from those possessed by the chromosomal producers. So far, CMY-2 is the most prevalent pAmpC in Ent, but DHA and FOX have high prevalence in certain geographic regions (Jacoby, 2009; Seiffert, et al., 2013). Unlike class A enzymes, cAmpCs and pAmpCs are poorly inhibited by the commercially available β-lactamase inhibitors but the fourth-generation cephalosporins (e.g., cefepime) usually remain in the susceptible ranges (Harris & Ferguson, 2012; Hilty, et al., 2013).

1.2.2. Resistance to carbapenems

Carbapenemases are β-lactamases able to hydrolize β-lactam antibiotics, including carbapenems (Papp-Wallace, Endimiani, Taracila, & Bonomo, 2011). Class A carbapenemases can be chromosomally encoded (e.g., SME) or plasmid-encoded (e.g., KPC, GES-5). KPC-types are the most clinically common carbapenemases found in Ent and are responsible for hospital outbreaks in many countries (Rapp & Urban, 2012; Walther-Rasmussen & Hoiby, 2007). Class B carbapenemases (metallo-β-lactamases, MBLs) are usually of VIM- and IMP-types, but the recently emerged NDM-types are becoming the most threatening carbapenemases and have spread rapidly among Ent in all continents (Nordmann, Naas, & Poirel, 2011; Walsh, 2010). The GIM and SPM MBLs have less global impact, but their prevalence in some countries deserve attention (e.g., South America and Germany, respectively) (Walsh, 2010). In Ent, class D carbapenemases are mainly represented by the OXA-48-like enzymes (e.g., OXA-48, -162, and -181). These genes are extensively reported among E. coli and K. pneumoniae isolates in the European and African Mediterranean countries (Poirel, et al., 2012). Recently, OXA-48 producers have been reported in North America (Lascols, Peirano, Hackel, Laupland, & Pitout, 2013; Mathers, et al., 2013).

It should be noted that carbapenem resistance in Ent may also be due to disruption of OMPs. This phenomenon may be observed under carbapenem treatment in K. pneumoniae (OmpK-35 and -36) (Endimiani, et al., 2009; Tsai, et al., 2011), E. coli (OmpF/C) (Oteo, et al., 2008; Tangden, Adler, Cars, Sandegren, & Lowdin, 2013), and Enterobacter spp. (OmpF/C) (Doumith, Ellington, Livermore, & Woodford, 2009).

1.2.3. Resistance to quinolones

Quinolones resistance among Ent is usually mediated by chromosomal mutations in the quinolone-resistance determining region (QRDR) that encode DNA gyrase (gyrA and parC) genes (Hooper, 2001; Jacoby, 2005). Nevertheless, low-level resistance can also arise from the expression of plasmid-mediated quinolone resistance (PMQR) determinants such as:: i) qnrA, -B, -S, -C, -D genes that encode proteins protecting the DNA gyrase from the quinolones’ action; ii) an aminoglycoside acetyltransferase encoded by the aac(6’)-Ib-cr gene that also acetylates quinolones; and iii) plasmid-mediated quinolone efflux-pumps (qepA) (Strahilevitz, Jacoby, Hooper, & Robicsek, 2009). The prevalence of qnr genes in ESBL-Ent is estimated around 10% (Karah, et al., 2010; Robicsek, Strahilevitz, Sahm, Jacoby, & Hooper, 2006), whereas that of aac(6’)-Ib-cr is much higher (15-50%) (Ambrozic Avgustin, Keber, Zerjavic, Orazem, & Grabnar, 2007; C. H. Park, Robicsek, Jacoby, Sahm, & Hooper, 2006; K. S. Park, et al., 2012; Pitout, Wei, Church, & Gregson, 2008). Notably, Hansen et al. characterized oqxAB (Hansen, Johannesen, Burmolle, Sorensen, & Sorensen, 2004), a plasmid-mediated gene encoding for an efflux pump conferring resistance to chloramphenicol and quinolones (Hansen, Jensen, Sorensen, & Sorensen, 2007) (Kim, et al., 2009; K. S. Park, et al., 2012).

1.2.4. Resistance to aminoglycosides

Aminoglycosides resistance in Ent is generally due to enzymatic inactivation, which is mediated by 3 different classes of aminoglycoside-modifying enzymes (AMEs): acetyltransferases, nucleotidyltransferases, and phosphotransferases (Magnet & Blanchard, 2005). More recently, a new aminoglycosides resistance mechanism that consists of ribosomal protection through enzymatic methylation of specific residues within the 16S rRNA (impeding binding of drugs to the 30S ribosomal subunits) has been described. These 16S rRNA methylases (ArmA, RmtA, RmtB, RmtC, RmtD, RmtF, RmtG, and NpmA) confer high-levels of resistance to aminoglycosides and can be mobilized among different species (Bueno, Francisco, O'Hara, de Oliveira Garcia, & Doi, 2013; Doi & Arakawa, 2007; Hidalgo, et al., 2013). For instance, the armA (the most prevalent methylase gene) can be associated with the blaESBLs (Galimand, Sabtcheva, Courvalin, & Lambert, 2005; Yan, et al., 2004). More importantly, the blaNDM genes are usually linked with 16S rRNA methylases (Nordmann, Poirel, Walsh, & Livermore, 2011).

1.2.5. Resistance to other antibiotic classes

Ent non-susceptible to ESCs and/or carbapenems are also frequently co-resistant to other classes of antibiotics. The most frequent genes conferring such resistance traits are as follows: tetracyclines (tet genes), sulfonamides (sul genes), trimethoprim (dfr genes), and phenicols (cml, cat, and floR genes) (Endimiani, Rossano, Kunz, Overesch, & Perreten, 2012).

1.3 Ideal characteristics of rapid detection methods

On the basis of epidemiological data (section 1.2), some antibiotic resistance traits deserve more attention than others when we design a rapid diagnostic tool (Table 1). In particular, we believe that blaCTX-M, blaTEM, blaSHV, blaCMY, blaKPC, blaNDM, blaOXA-48, the 16S rRNA methylases genes and mutations in gyrA and parC are the most important targets that should be always tested for Ent. However, systems capable to detect further genes (e.g., blaPER, blaVEB, blaSPM, blaGIM, blaSME, and PMQR determinants) with less global epidemiological and/or clinical impact would be welcome. This is important because in some geographic areas the spread of specific resistance genes (e.g., blaSPM and blaSME in America or blaGIM in central Europe) has reached worrisome levels (Bush, et al., 2013; Hamprecht, et al., 2013; Walsh, 2010).

Table 1.

Most important antibiotic resistance traits that molecular methods should be able to detect in Ent. The list is based on: i) the frequency of observation during epidemiological studies conducted worldwide; ii) the clinical importance of the antimicrobials affected by the resistance mechanism(s)

Antibiotics Mechanism of resistance Most frequent resistance traits Importance Ideal level of detection/characterization
Extended-spectrum cephalosporins Extended-spectrum β-lactamases (ESBLs) CTX-M ++++ Generic
TEM and SHV ++++ Distinguishing ESBLs from non-ESBLs a
PER ++ Generic
GES ++ Distinguishing ESBLs (e.g., GES-1) from carbapenemases (e.g., GES-5) b
SFO + Generic
Plasmid-mediated AmpCs (pAmpCs) CMY ++++ Generic
DHA +++ Generic
FOX ++ Generic
LAT, MIR, ACT, MOX + Generic
Chromosomal AmpCs (cAmpCs) Mutations in the promoter (E. coli) ++ Identifying substitution (e.g., at -32/-42) and insertions (between -10 and -35)
Mutations in ampD (Enterobacter spp.) + Unknown specific target
Carbapenems Carbapenemases
    Class A KPC ++++ Generic
GES ++ Distinguishing ESBLs (e.g., GES-1) from carbapenemases (e.g., GES-5) b
    Class B (MBL) NDM ++++ Generic
SME + Generic
VIM ++++ Generic
IMP +++ Generic
SPM and GIM ++ Generic
AIM, SIM + Generic
    Class D OXA-48 ++++ Specific
OXA-181, −162, −163 +++ Specific
Impermeability (porins) OmpK-35 and −36 (K. pneumoniae) ++ Identifying insertions and/or frame-shift mutation(s)
OmpF/C (E. coli and Enterobacter spp.) + Unknown specific mutation/target
Quinolones Amino acid substitutions in the QRDR GyrA ++++ Mutation at position -248, -259/260 c
ParC ++++ Mutation at position 238/239 and 250/251 d
PMQR determinants QnrA B/S/D + Generic
Aac(6)-Ib-cr ++ Distinguishing the aac(6')-Ib-cr from the AME aac(6')-Ib
QepA ++ Generic
OqxAB + Generic
Aminoglycosides AMEs Aac(6)-I-like, Ant(3)-like, Aph(3)-like +++ Specific
16S rRNA methylases ArmA, RmtA/B/C/D/E, NpmA ++++ Generic

QRDR, quinolone resistance determining region; AMEs, aminoglycoside modifying enzymes.

a

Amino acidic substitutions E104K, R164S/H, G238S and E240K for TEM; D179N, G238S and E240K for SHV (Gniadkowski, 2008; Perilli, et al., 2008; Randegger, Keller, Irla, Wada, & Hachler, 2000);

b

Amino acidic substitution G170A/S (Girlich, Poirel, Szczepanowski, Schluter, & Nordmann, 2012);

c

Lead to modification of S83 and D87 (Jacoby, 2005; Qiang, et al., 2002);

d

Lead to modification of S80 and E84 (Qiang, et al., 2002) (Jacoby, 2005)

Ideally, a rapid system should investigate all resistance traits in one reaction and should provide easy to interpret results the same day the test is performed. Moreover, the assay should be able to identify specific variants of alleles that encode for proteins with a different impact on the antibiotic resistance phenotype (e.g., TEM and SHV with ESBL spectrum; OXA-48 rather than non-OXA carbapenemases). Finally, the methodology should be easy to perform, relatively cheap, accurate (sensitive and specific), and versatile enough to be regularly updated according to the evolution of the antibiotic resistance traits.

2. END-POINT PCRs

2.1 Single end-point PCRs

The single PCR is the ancestor of the molecular methods used in diagnostic and epidemiological studies for the detection of antibiotic resistance genes. It begins with the amplification of a target gene supposed to confer resistance and then includes an endpoint visualization of the amplification product(s) (Mullis & Faloona, 1987; Predari, Ligozzi, & Fontana, 1991; Zhou, Bordon, Sirot, Kitzis, & Gutmann, 1994). The single PCR allows identification of only one gene and requires the design of specific primers for the intended target. This approach is especially useful for the screening of a limited number of isolates and only if the detection of a gene at the generic level is sufficient (Table 1). This method can be appropriate in a known outbreak context when early detection of new cases is required.

Furthermore, by designing degenerate primers, single end-point PCRs can be used to detect specific genes (Table 2). For instance, they allow the rapid detection of all variants of the CTX-M-types enzymes (Boyd, et al., 2004). However, for specific gene characterization, discriminatory analysis of the PCR products (e.g., DNA sequencing) is necessary (see section 3). In some cases, digestion by endonucleases of the relative PCR product may discriminate among variants of a given gene, preventing sequencing. For example, the variant aac(6)-Ib-cr lacks the cleavage site for the BstF5I enzyme which is present in the ancestor gene (C. H. Park, et al., 2006).

Table 2.

Single and multiplex end-point PCRs used for the detection of the main antibiotic resistance genes in Enterobacteriaceae

Approach Detected resistance target Nr. primers sets Cycling conditions TTR (hrs) b Comments Reference
Single PCRs a CTX-M 4 1 2 Identification of CTX-M at group level (Pitout, Hossain, & Hanson, 2004)
CTX-M 1 1 2 Generic detection of CTX-M (Boyd, et al., 2004)
TEM, SHV, PER 1, 1, 1 1, 1, 1 2 Generic detection (De Champs, et al., 2004)
VEB, GES, SFO 1, 1, 1 1, 1, 1 2 Generic detection (Poirel, Le Thomas, Naas, Karim, & Nordmann, 2000)
KPC 1 1 2 Generic detection (Bradford, et al., 2004)
IMP and VIM 1, 1 1, 1 2 Generic detection (Pitout, et al., 2005)
NDM 1 1 2 Generic detection (Nordmann, Poirel, Carrer, Toleman, & Walsh, 2011)
OmpK-35, OmpK-36 (K. pneumoniae) 2 1 2 Sequencing always required (C. H. Lee, et al., 2007)
OmpC (E. cloacae) 1 1 2 Sequencing always required (Lou, et al., 2011)
OmpF (E. coli) 1 1 2 Sequencing always required (Kishii & Takei, 2009)
AmpD 1 1 2 Sequencing always required (Naas, Massuard, Garnier, & Nordmann, 2001)
GyrA 1 1 2 Sequencing always required (Weigel, Steward, & Tenover, 1998)
ParC 1 1 2 Sequencing always required (McDonald, et al., 2001)
QnrA, QnrB 1 1 2 Generic detection (Wang, et al., 2003)
QnrS 1 1 2 Generic detection (Robicsek, et al., 2006)
Aac(6)-Ib-cr 1 1 2 Sequencing always required (C. H. Park, et al., 2006)
QepA 1 1 2 Specific detection (Perichon, Courvalin, & Galimand, 2007)
OqxAB 1 1 2 Specific detection (Kim, et al., 2009)
AMEs 3 1 2 Specific detection (Miro, et al., 2012)
Multiplex PCRs TEM, SHV, CTX-M NA 1 2.5 Commercial kit (Hyplex) NA
CMY, DHA, ACC, FOX, MOX, LAT, ACT, MIR 6 1 2-4 Generic detection. DNA sequencing is required for CMY. Cross reaction with chromosomal genes (Perez-Perez & Hanson, 2002)
TEM, SHV, CTX-M, GES, VEB, PER
KPC, IMP, VIM
CMY, DHA, ACC, FOX, MOX, LAT, ACT, MIR
OXA (including OXA-48-like)
21 3 3 NDM-types not included
Cross reaction with chromosomal genes
(Dallenne, et al., 2010)
TEM, SHV, CTX-M, GES, VEB, PER
KPC, SME, IMI, NMC-A, IMP, VIM, GIM, SIM, NDM, SPM
CMY, DHA, MOX, ACC, ACT, MIR, FOX
OXA (including OXA-48-like)
20 1 3 Partial cross reaction with chromosomal genes (Voets, et al., 2011)
KPC, BIC
SPM, VIM, GIM, AIM, IMP, NDM, SIM, DIM
OXA-48, OXA-163, OXA-181
11 1 <4 Limited to carbapenemases detection (Poirel, et al., 2011)
KPC, IMP, VIM, NDM, OXA-48-like NA 1 2.5 Commercial kit (Hyplex) (Kaase, et al., 2012)
SME, NMC-A, KPC, GES 4 1 <3 No detection of NDM (Hong, et al., 2012)
KPC, NDM, VIM, IMP, OXA-48-like 5 1 <3 It uses previously reported primers (Doyle, et al., 2012)
QnrA, QnrB, QnrS 3 1 <3 No detection of other PMQR determinants (Cattoir, et al., 2007)
ArmA, RmtA, RmtB, RmtC, RmtD, NpmA 6 1 <4 Generic detection (Bercot, et al., 2011)

Note. NA, not available

a

Numerous single PCRs have been designed to detect the listed antibiotic resistance genes. In this table, we present the most frequently implemented PCR schemes;

b

The TTR is calculated starting from DNA template

2.2 Multiplex end-point PCRs

The multiplex end-point PCRs are based on the same principle of the single end-point PCRs, but use multiple primer sets allowing the amplification of different targets in the same reaction. The designing of a multiplex PCR presents some complex requirements: i) high specificity of the primer sets to the intended targets; ii) no self-reaction of the primers; iii) ideally, identical cycling condition for all primer sets; and iv) different sizes for the intended amplicons to allow a direct differentiation (Markoulatos, Siafakas, & Moncany, 2002). The use of positive controls is also required, especially when similar PCR product sizes are generated for different targets and when standard agarose gels are implemented for visualization.

To date, several multiplex PCRs able to detect antimicrobial resistance traits have been proposed (Table 2). One example is the multiplex designed more than 10 years ago by Perez-Perez and Hanson for the detection of blapAmpCs. The DNA template was obtained by boiling the colonies and the TTR was estimated at around 3 hrs. However, cross reaction with the cAmpCs occurred. Moreover, the size of different gene amplicons was too similar to be discriminated by standard gels; thus WAVE DNA analysis was applied for the discrimination (Perez-Perez & Hanson, 2002). Dallenne et al. developed a multiplex PCR able to detect a wide panel of bla genes. The setup, besides being laborious (i.e., seven multiplex reactions and three different cycling conditions), presents cross hybridization with the chromosomal class A -lactamases of certain producers (e.g., Kluyvera ascorbata, K. georgiana, and K. pneumoniae) and with the cAmpC of Hafnia alvei, Enterobacter asburiae and Morganella morganii. However, direct DNA sequencing of obtained amplicons allowed correct gene identification. The sequencing did not discriminate CTX-M-15 from CTX-M-28 and some VEB and CMY variants, but was able to discern between TEM and SHV ESBLs from non-ESBL variants (Dallenne, Da Costa, Decre, Favier, & Arlet, 2010). This method has found application in numerous studies for the screening of ESBLs and carbapenemase producing strains (Hornsey, Phee, & Wareham, 2011). Voets et al. used part of the above setup and implemented it by including the detection of further carbapenemases. They optimized the PCR process to reduce cross reaction with chromosomal β-lactamases and simplified the procedure to only one cycling condition (Voets, Fluit, Scharringa, Cohen Stuart, & Leverstein-van Hall, 2011). Ellington et al. published a multiplex PCR for the detection of MBLs. DNA extracts obtained by boiling the colonies were implemented with the method and proved to amplify different IMP and VIM variants (Ellington, Kistler, Livermore, & Woodford, 2007). Later, Poirel et al. included this approach in a multiplex PCR with a wider range of carbapenemase detection and using DNA templates obtained by alkaline lysis of colonies. The method was specific when primers sets were used either in single or multiplex PCR reactions (Poirel, Walsh, Cuvillier, & Nordmann, 2011). This protocol is easily and rapidly executed (<4 hrs) and is applicable for the screening of large collections (Capone, et al., 2013). More recently, Hong et al. proposed a multiplex PCR for the detection of class A carbapenemases. Although the method was tested on a limited number of control strains, which did not comprise several variants of the targeted genes, the procedure is simple and rapid and it could find application in an opportune epidemiological context. This multiplex PCR did not produce false positive results, suggesting high specificity (Hong, et al., 2012) and it has been used for the molecular confirmation of phenotypic analysis (Rojo-Bezares, Martin, Lopez, Torres, & Saenz, 2012; Ruiz, et al., 2012). Doyle et al. designed an in-house multiplex PCR for the simultaneous detection of blaKPC, blaNDM, blaVIM, blaIMP, and blaOXA-48-like. The methodology was evaluated against 142 Ent obtained from the SMART study (2008-2009) and compared to the performance of phenotypic tests (i.e. mHT, Mastdiscs ID inhibitor combination disks, Rosco Diagnostica Neo-Sensitabs, and MBL Etest). The PCR had 100% sensitivity and specificity, whereas the other methods were 58% and 93%, 78% and 93%, 80% and 93%, and 55% and 100%, respectively (Doyle, et al., 2012).

As anticipated, monitoring of co-resistance to other antibiotics (especially aminoglycosides and quinolones) among ESBL- and/or carbapenemase-producing Ent is an important task (Table 1). Cattoir et al. designed a multiplex PCR for the detection of several variants of qnrA/B/S genes (Cattoir, Poirel, Rotimi, Soussy, & Nordmann, 2007). However, we note that there is no multiplex PCR able to detect all relevant PMQR determinants. Thus, the method proposed by Cattoir et al. should be used in association with further single PCRs for the detection of qepA, aac(6′)-Ib-cr and oqxAB (Yamane, Wachino, Suzuki, & Arakawa, 2008; Yue, et al., 2008). Berçot et al. developed a protocol for the detection of 16S rRNA methylase genes in NDM-possessing Ent. The template for the multiplex reaction was obtained utilizing a commercial DNA extraction kit and the TTR starting from bacterial culture was estimated to be less than 4 hrs. This proposed multiplex PCR is rapid, specific, sensitive, of simple execution, and it has been applied for the screening of strain collections (Bercot, Poirel, & Nordmann, 2011; Williamson, et al., 2012). Diaz et al. established a multiplex PCR limited to the detection of the AMEs aac(3)-IIa and aac(3)-IVa in Ent of environmental origin (Diaz, Cooper, Cloeckaert, & Siebeling, 2006). Except for this latter report, no multiplex PCRs detecting AMEs in Ent have been engineered, so far. Probably the high diversity of such enzymes and their corresponding sequences hinders the determination of an efficient multiplex protocol. Of note, several investigators have designed single PCRs for the detection of AMEs genes (Table 2).

2.3 Commercially available multiplex PCRs

Amplex Diagnostics (Germany) has designed three commercial PCR-ELISA kits to rapidly detect clinically important bla genes (www.hyplex.info/). The assays are based on specific multiplex PCRs for target genes, reverse hybridization with labeled PCR products in a 96-well plate, addition of conjugate (peroxidase), addition of substrate (tetra-methyl-benzidin), and photometric measurement at 450 nm. In particular, hyplex ESBL ID, hyplex SuperBug ID, and hyplex CarbOXA ID detect respectively: i) blaTEM, blaSHV, blaCTX-M, and blaOXA; ii) blaKPC, blaIMP, blaVIM, blaNDM, blaOXA-48-like; and iii) blaOXA-23, blaOXA-24, and blaOXA-51-like. Hyplex SuperBug ID was recently evaluated by testing 132 carbapenem-resistant Ent which included KPC-, VIM-, IMP-, NDM-, and OXA-48-like-producing isolates. Overall, the agreement with the PCR/sequencing results was 97% (128/132), but three VIM producers were not detected. Starting from buffer lysis of colonies, the TTR is usually of 2.5-4 hrs (Kaase, Szabados, Wassill, & Gatermann, 2012). A prototype of the above platform (i.e., hyplex MBL ID: able to detect only blaIMP and blaVIM) was also evaluated against 326 clinical samples (including blood, urine, bronchial secretions, and wound swabs) of which 73 were positive for VIM-producing Ent or nonfermenters. Overall, sensitivity and specificity were 98.6% and 98%, respectively (Avlami, et al., 2010).

2.4 Advantage and disadvantages of PCRs

Both single and multiplex PCRs require a high copy number of the intended target to produce a detectable amplification product. Thus, preliminary bacterial isolation, cultivation and DNA extraction are necessary. In this regard, amplification from colonies represents a rapid alternative to genomic DNA isolation. In contrast, clinical samples may harbor inhibitors (e.g. hemoglobin and heparin in the blood; urates in the urine; polysaccarides in the feces) for the PCR reaction making the test less sensitive (Schrader, Schielke, Ellerbroek, & Johne, 2012).

Depending on the level of details required, results from single and multiplex PCRs can be obtained in less than 4 hrs (i.e., amplification results) to more than 24 hrs (determination of the specific DNA sequence or further analyses). Typically, multiplex PCRs provide a generic identification of targets genes. However, such PCRs detect numerous genes simultaneously and, even if the setup of such protocols is laborious, the execution is simple. The optimization of the method to a few reactions and/or one cycling condition, and the relatively low costs of performance, make multiplex PCRs more appropriate and useful for epidemiological studies focusing on the molecular characterization of the drug resistance genes (Table 4).

Table 4.

Main characteristics of the rapid molecular tools so far available for the detection and characterization of the antibiotic resistance genes in multidrug-resistant Enterobacteriaceae (MDR-Ent)

Molecular tool Time Costs c Main setting(s) of application Comments
Single PCR coupled by standard DNA sequencing 2-3 hrs (PCR) a
1-3 or more days (sequencing)
+
+++
Research/Epidemiology Detect all genes but labor intensive, especially for multiple genes
    Next-generation sequencing methodologies 3 hrs a +++ Clinical Much more rapid than standard sequencing, especially if coupled to real-time PCR. However, they are still unavailable in most laboratories
Multiplex end-point PCR 3-4 hrs a + Epidemiology PCR reaction should be optimized. Moderately labor intensive
    Hyplex platforms (commercial) 2.5-4 hrs a +++ Clinical Helpful for treatment and screening of carriers
Real-time PCR 2-3 hrs a ++ Clinical Helpful for treatment and screening of carriers. However, only one target gene is tested
    NucliSENS EasyQ KPC platform (commercial) 2 hrs a +++ Clinical It detects only blaKPC
Multiplex real-time PCR 2-3 hrs a ++ Clinical Reactions should be optimized. Moderately labor intensive
    Check-Points (commercial) 4.5 hrs b +++ Clinical Helpful for treatment and screening of carriers of ESBL-Ent and/or carbapenemase producers
LAMP 1-1.5 hrs a + Clinical Very rapid and cheap. Good for low income countries
Microarray 6-8 hrs or more a ++ Epidemiology In-house platforms are difficult to be implemented for clinical applications
    Check-Points (commercial kits) 5-6 hrs a +++ Epidemiology/Clinical Very versatile and easy to use
    Identibac (commercial) 5-6 hrs a +++ Epidemiology/Research Capable to detect many genes (including those for non-β-lactams) but not yet validated testing large and well-characterized collections of isolates
MALDI-TOF MS (for detection of catalytic activity) 3 hrs b + Clinical/Research It detects only the spectrum of β-lactamases without distinguishing the specific kind of enzymes
MALDI-TOF MS (for detection of atomic mass) 3 hrs b + Clinical/Research It can detect the class of β-lactamases. The specific kind of enzymes can be distinguished if the accuracy and the databases are improved
MALDI-TOF MS (for mini-sequencing) 3 hrs a ++ Research Difficult to be implemented when compared to the other easier and more rapid systems
PCR/ESI-MS (PLEX-ID, commercial) 4-6 hrs (starting from sample) ++++ (Clinical)/Research Excellent for clinical applications. However, so far available only for research
a

TTR calculated starting from DNA template

b

TTR calculated starting from colonies

c

+, lowest cost; ++++, highest cost

3. NEXT-GENERATION DNA SEQUENCING METHODOLOGIES

As previously mentioned, characterizing variants of certain genes is important, thus their DNA sequence must be determined. Two de novo sequencing methods were developed in the ′70s, one by Maxam and Gilbert (Maxam & Gilbert, 1977), and the other one by Sanger (Sanger, Nicklen, & Coulson, 1977). The first method is based on the enzymatic digestion of the DNA template. Four digestions in independent reactions are carried out on the PCR product, each of them performing cleavages at one specific base. The resolution by electrophoresis of the obtained fragments provides with the sequence of the product. The Sanger method consists of a reaction of DNA synthesis by a DNA polymerase using dNTPs (deoxy-nucleotides tri-phosphates) and ddNTPs (dideoxy-nucleotides tri-phosphates) labeled with four different fluorophores (one fluorophore for each base). The incorporation of ddNTPs, which are terminators of the synthesis, produces fragments of different sizes that can be separated by capillary electrophoresis or analyzed by mass spectrometry (Murray, 1996). So far, this methodology is the most used in clinics for the sequencing of antibiotic resistance genes. However, this method is laborious, requiring several post-PCR procedures such as purification of the products, sequencing reaction and resolution of the sequence products. A common alternative is out-sourcing the process which in turn results in a delay of the final data.

Many efforts have been made in the last years to overcome these limitations giving rise to the so-called “next-generation sequencing” (NGS) methodologies (L. Liu, et al., 2012). Mostly, these methods are based on the “real-time” fluorescence/light detection occurring during the sequencing process. The methods differ for the chemistry of the underling reactions. Ronaghi et al. published a sequencing methodology based on the action of four enzymes: a DNA polymerase adds dNTPs to a growing DNA strand, the pyrophosphate group PPi released during the polymerization is the substrate of a sulfurylase which in turn synthetizes ATP, this latter is used by a luciferase for light emission. An apyrase removes excess of dNTPs and ATP improving fidelity of the sequencing, also ensured by the DNA strand conformation. The intensity of light is specific to the base that has been incorporated (Ronaghi, Karamohamed, Pettersson, Uhlen, & Nyren, 1996). This methodology, famous as pyrosequencing, and commercially as “454” (Roche), has found application for antibiotic resistance gene characterization in several studies and is becoming a bench-top tool. For instance, Poirel et al. implemented it to characterize blaGES variants encoding carbapenemases from K. pneumoniae and Pseudomonas aeruginosa (Poirel, Naas, & Nordmann, 2006). Naas et al. characterized CTX-M variants in clinical ESBL-producing E. coli using pyrosequencing after real-time PCR of blaCTX-M. The five CTX-M groups were discriminated with this technique by 13 bp DNA region analysis and this result was confirmed by the standard DNA sequencing approach (Naas, Oxacelay, & Nordmann, 2007). In one study, the application of pyrosequencing and analysis of single nucleotide polymorphisms (SNPs) for blaSHV and blaTEM detected ESBLs variants of the respective genes in a collection of Ent. Notably, this method also enabled the identification of new bla variants (Jones, et al., 2009). Interestingly, Haanpera et al. used the pyrosequencing for the discrimination of chromosomal SHV from acquired ESBLs types on a collection of K. pneumoniae confirming the powerful resolution of this technique (Haanpera, Forssten, Huovinen, & Jalava, 2008). Pyrosequencing was also efficiently applied for the discrimination of the aac(6)-Ib from the variant aac(6)-Ib-cr (Guillard, et al., 2010), and for the detection of gyrA and parC mutations conferring quinolones resistance in a Salmonella enterica (Hopkins, Arnold, & Threlfall, 2007).

Although yet not applied for the antibiotic resistance characterization, other NGS methods are available. The sequencing by synthesis, commercially referred to as “Illumina”, consists of the neo-synthesis of DNA by a DNA polymerase incorporating dNTPs labeled with fluorescent terminators which are detected in real-time (L. Liu, et al., 2012). The sequencing by ligation, implemented by Life-technology with the commercial name “SOLiD”, is based on the action of a DNA ligase which adds fluorescent labeled dNTPs to an anchor probe hybridizing a single strand DNA in proximity of the region to be sequenced (L. Liu, et al., 2012).

The Ion torrent sequencing is another powerful NGS approach recently developed and which might become a bench-top NGS tool. The sequencing by synthesis reaction is performed on a semiconductor chip and is based on the detection of the hydrogen ions released at each addition of a nucleic base. Electrical pulses are then directly converted into a DNA sequence (L. Liu, et al., 2012). This technology has been recently applied to the investigation of an outbreak of E. coli infections, allowing not only the detection of CTX-M-15, but providing the genome sequences of these strains (Sherry, et al., 2013). The authors described this method as fast (five days from a positive culture to completion of sequencing) and reasonably cheap (~ US $300 per strain, reagents only). Although development of further bioinformatics for an easier data interpretation is necessary, the advent of NGS methodologies and their commercialization promise to be revolutionary for the improvement of diagnosis and epidemiology of infections due to MDR-Ent.

4. REAL-TIME PCRs

The real-time PCR consists of an amplification reaction coupled with the simultaneous detection of the exponentially amplified target and visualization of the reaction phases (Higuchi, Fockler, Dollinger, & Watson, 1993). The number of cycles necessary to reach this exponential phase often depends on the initial quantity of target present in the reaction. The cycle at which the reaction assumes an exponential state is referred to as the quantitative cycle (Cq) because it allows the quantification of the initial DNA target (qPCR). Such quantification can be absolute or relative, the first requires the comparison of the target quantification to a standard with known concentration. This quantification is useful to calculate the number of copies of a target in a genome and/or to compare the amount of a target from two different samples. The relative quantification, using mRNA as template or a retro-transcriptional template (RT-PCR), allows defining the level of expression of a target compared to a constantly expressed (housekeeping) gene (Heid, Stevens, Livak, & Williams, 1996). Furthermore, real-time PCR apparatus are able to quantify the amount of denatured DNA (single strand DNA) at a certain temperature (Tm), thereby providing the melting curve of a DNA fragment. The denaturing temperature is specific to the composition of the DNA fragment, thus a precise measure of the melting curve(s) gives information on even small variations in the DNA sequence (the so called “High Resolution Melting Analysis”) (Erali, Voelkerding, & Wittwer, 2008; Tong & Giffard, 2012).

Monitoring of fluorescence emission occurring during the amplification is the mechanism on which the dynamics of real-time PCR is based (Higuchi, Dollinger, Walsh, & Griffith, 1992). The fluorescence emission can be obtained according to four approaches (Arya, et al., 2005): i) generic, implementing a dye (e.g., SYBR Green, FAM, Cal Fluor Red, Hex) which fluoresces when introduced in the double-strand DNA; ii) specific, using TaqMan probes harboring a fluorophore at the 5′ and a quencher (preventing fluorescence emission) at the 3′. TaqMan probes hybridize on a specific DNA target and are depredated by the Taq polymerase exo-nuclease activity during amplification. The removal of the fluorophore from the quencer allows fluorescence emission. Different TaqMan probes labeled with diverse fluorophores can be used in the same PCR mixture; iii) specific, utilizing HYB probes, which harbor two different fluorescent labels. They hybridize on the target only during the annealing step in a head-to-tail direction and when they are close to each other emit fluorescence. The probes are removed during the elongation, due to the increase in temperature of the reaction; iv) specific, using molecular beacons consisting of the probe flanked by two complementary sequences carrying a fluorophore and a quencer at their respective extremities. This match creates a stem-loop structure where the quencer, close to the fluorophore, prevents fluorescence emission. Appropriate Tm and presence of a complementary sequence for the probe open the stem-loop structure allowing the removal of the fluorophore from the quencer and favoring fluorescence emission. Scorpion probes are designed using a similar principle (Arya, et al., 2005).

The loop-mediated isothermal amplification of DNA (LAMP), based on the detection of fluorescence using SYBR Green and a peculiar engineering of the primer set, allows DNA target amplification at a unique and constant temperature (Notomi, et al., 2000). This method, being cheaper, could represent an alternative to PCR (due to its higher sensitivity) and to real-time PCR. 4.1 Single real-time PCRs

Single real-time PCRs

One of the first applications of real-time PCR for detecting antibiotic resistance in Ent was performed by Hammond et al. where a SNP of blaSHV was analyzed to distinguish ESBL from non-ESBL variants in K. pneumoniae (Table 3). The methodology also enabled the quantification of copy numbers of the allele and to elucidate the effect of a mobile genetic element on the amplification of the blaSHV in the genome of the strain (Hammond, Schooneveldt, Nimmo, Huygens, & Giffard, 2005). Mroczkowska and Barlow used molecular beacons for detecting single nucleotide polymorphism of blaTEM (Mroczkowska & Barlow, 2008). Using TaqMan probes specific for CTX-M variants, Birkett et al. implemented a highly specific method that avoids sequencing costs and provides a faster alternative to conventional end-point PCR (Birkett, et al., 2007). Several other studies have highlighted the higher sensitivity of real-time PCR compared to culture based detection of carbapenemase producers (Table 3). Raghunathan et al. compared real-time PCR using the dye SYBR Green with the mHT for the detection of blaKPC in K. pneumoniae isolates. Although sensitivity and specificity were comparable, real-time PCR was more rapid (Raghunathan, Samuel, & Tibbetts, 2011). Kruettgen et al. designed real-time PCR for the detection of blaNDM and proposed a methodology for synthetic positive controls preparation (Kruttgen, Razavi, Imohl, & Ritter, 2011). Detection of blaNDM using TaqMan showed a detection limit of 10 copies of the target gene per reaction (Manchanda, et al., 2011). Oxacelay et al. engineered a HYB probe for the detection of CTX-M variants directly from urine samples. The method was quite sensitive (detection limit of 102-103 cells/ml of urine) (Oxacelay, Ergani, Naas, & Nordmann, 2009). Naas et al. engineered a TaqMan probe for the detection of blaNDM-1 from spiked fecal stools. This method had a detection limit of 101 CFU/100 mg of feces, while selective media such as ChromoID ESBL and CHROMAgar KPC were 101-3×101 and 101-4×103 CFU/100 mg of feces, respectively. The method demonstrated 100% sensitivity and specificity, and the protocol was rapidly peformed (4 hrs). Therefore, this method is a valuable tool for rapid detection of NDM-1 producers in clinical samples (Naas, Ergani, et al., 2011). Alternatively, the LAMP methodology has been applied for NDM-1 detection with a very rapid protocol (1 h) and a detection limit of 10.7 pg/ml (W. Liu, et al., 2012). Recently, real-time PCR using a 100% sensitive TaqMan probe for the detection of blaOXA-48 in spiked stool samples was implemented. The method had a detection limit of 10-50 CFU/100 mg of feces, while for ChromID ESBL and SUPERCARBA medium were at 101-102 and 1-3×101 CFU/100 mg of feces, respectively (Naas, Cotellon, Ergani, & Nordmann, 2013).

Table 3.

Single and multiplex real-time PCRs used for the detection of the main antibiotic resistance genes in Enterobacteriaceae

Approach Detected resistance target Nr. primers sets/probes Cycling conditions TTR (hrs) Comments Reference
Single real-time PCR SHV (ESBLs variants) 2 1 2 SNPs are detected by the specificity of the primers (Hammond, et al., 2005)
TEM (ESBLs variants) 2/2 1 2 SNPs are detected by the probes (Mroczkowska & Barlow, 2008)
CTX-M 1/1 1 2 Specific characterization (Oxacelay, et al., 2009)
CTX-M 1/4 1 2 Specific detection (Birkett, et al., 2007)
KPC NA 1 2 Commercial kit (EasyQ KPC test, bioMerieux) (Spanu, et al., 2012)
KPC 1/1 1 <4 Clinical implementation using enriched rectal swabs (Singh, et al., 2012)
KPC 1/SG 1 2 Used previously published primers (Raghunathan, et al., 2011)
NDM-1 2/3 1 3 Synthesized a positive control (Kruttgen, et al., 2011)
NDM-1 1/1 1 2 Good performance of the TaqMan probe for copy number determination (Manchanda, et al., 2011)
NDM-1 1/1 1 < 4 From spiked stools (Naas, Ergani, et al., 2011)
OXA-48-like 1/1 1 < 4 From spiked stools (Naas, et al., 2013)
CMY-2 1/SG 1 2 Quantification purpose (Kurpiel & Hanson, 2011)
OmpK35, OmpK36 2/2 1 2 Quantification purpose (Landman, et al., 2009)
Multiplex real-time PCR TEM and SHV (ESBL variants), CTX-M NA 1 3.5 Commercial kit (CheckPoints) (Nijhms, et al., 2012)
CIT, MOX, FOX, DHA, ACC, EBC 6/SG 1 3 Performed with previous published primers (Brolund, et al., 2010)
CMY, ACT, DHA, MOX, ACC, FOX, 16S rRNA 7/7 1 3 No cross reaction with cAmpC of E. coli (Geyer, et al., 2012)
KPC, 16S rRNA 2/2 1 < 4 From over-night cultures (Singh, et al., 2012)
IMP, VIM 2/4 1 < 1 Specific detection (Bisiklis, et al., 2007)
KPC, RNaseP 2/2 1 3 Clinical implementation from blood cultures (Hindiyeh, et al., 2011)
KPC, TonB, GapA 4/4 1 4 Identifies peculiar K. pneumoniae sequencing types (Chen, et al., 2012)
GES, IMI/NMC, KPC, SME, OXA-48 5/5 1 3 It does not discriminate the carbapenemase GES variants (Swayne, et al., 2011)
KPC, NDM, GES, OXA-48, IMP, VIM 6/EG 1 3 Specific identification by Tm analysis (Monteiro, et al., 2012)
KPC, NDM, VIM, IMP, OXA-48 NA 1 3.5 Commercial kit (CheckPoints). Not yet evaluated NA
QnrA/B/S/C/D, QepA 6 2 < 2 Specific identification (Guillard, et al., 2011)
QnrA/B/ S 3 1 < 2 Coupled with High resolution melting curve analysis (Guillard, et al., 2012)
Aac(6)-Ib-cr 1/1 1 < 1 Coupled with high resolution melting curve analysis, specific detection (Bell, et al., 2010)

Note. NA, not available; SB, SYBR Green; EG, EvaGreen dye

Real-time PCR can also be used for copy number determination of target genes. Kurpiel and Hanson used SYBR Green for the copy number determination of blaCMY-2, harbored by clinical E. coli isolates, demonstrating tandem duplication of the gene (Kurpiel & Hanson, 2011). Importantly, Landman et al. designed two TaqMan probes for the quantification of OmpK35 and -36 in carbapenem-resistant K. pneumoniae isolates, confirming the role of OmpK36 expression in the susceptibility profile of these organisms (Landman, Bratu, & Quale, 2009).

4.2 Multiplex real-time PCRs

The analysis of the melting temperature curves allows a specific and easy discrimination of amplified products obtained in a multiplex real-time PCR (Table 3). Bisiklis et al. used HYB probe to amplify blaVIM and blaIMP. The analysis of the Tm provided discrimination between the two genes and allowed the detection of gene variants (Bisiklis, Papageorgiou, Frantzidou, & Alexiou-Daniel, 2007). Adapting a previously developed method to detect pAmpCs (Perez-Perez & Hanson, 2002), Brolund et al. designed two triplex real-time PCRs using SYBR Green dye with the same cycling conditions. Starting from DNA templates of boiled colonies, the method was rapid (<3 hrs) and the analysis of the melting curves demonstrated high specificity (Brolund, et al., 2010). More recently, Geyer et al. developed a real-time TaqMan assay for pAmpCs showing specificity and sensitivity of 100% (Geyer, Reisbig, & Hanson, 2012). Swyne et al. designed a TaqMan real-time PCR for the detection of class A and D carbapenemases. The method was specific and the amplicons could be distinguished by melting curve analysis (Swayne, Ludlam, Shet, Woodford, & Curran, 2011). Highly accurate detection of different classes of carbapenemases was also obtained by Monteiro et al. by implementing Tm analysis (Monteiro, Widen, Pignatari, Kubasek, & Silbert, 2012). Interestingly, Chen et al. combined MLST and blaKPC detection in K. pneumoniae using molecular beacons in a unique multiplex mixture. The method was rapid (1 h) and had a detection limit of 4 CFU per reaction (Chen, et al., 2012). The same authors applied a similar technique to distinguish KPC variants (Chen, et al., 2011).

Guillard et al. elaborated the first real-time PCR for the detection of PMQR genes using the ResoLight dye for the detection of qnr genes and SYBR Green for qepA detection. The method was rapid (< 2 hrs) and melting curve analysis guaranteed discrimination of the qnr alleles (Guillard, et al., 2011). The same authors developed a single and multiplex PCR coupled with high resolution melting analysis for the detection of PMQR determinants (Guillard, et al., 2012). Taking advantage of the high resolution melting curve analysis, Bell et al. were able to detect the aac(6’)-Ib-cr variant after the real-time PCR assay (Bell, Turnidge, & Andersson, 2010).

4.3 Commercially available real-time PCRs

The New NucliSENS EasyQ KPC (bioMérieux) has been shown to be a highly sensitive and specific commercial real-time PCR system for the detection of blaKPC in K. pneumoniae. Spanu et al. evaluated the performance of the kit on a large collection of Ent and compared the results with those obtained with conventional amplification methods. Besides the high sensitivity and specificity, the method was very rapid (2 hrs) making it valuable for surveillance studies and rapid identification of carriers (Spanu, et al., 2012). Check-Points Health BV (the Netherlands) has developed a rapid (3.5 hrs) ligation mediated multiplex real-time PCR (Check-MDR ESBL kit) to detect SHV/TEM variants (ESBL and non-ESBL) and CTX-M groups. Nijhuis et al. compared this platform with the phenotypic combination disc method showing sensitivity and specificity of 99% and 92.2%, respectively (Nijhuis, van Zwet, Stuart, Weijers, & Savelkoul, 2012). The company has also developed a multiplex real-time PCR (Check-MDR Carba) to rapidly detect blaKPC, blaNDM, blaVIM, blaIMP, and blaOXA-48. However, the performance and accurateness of the kit has not been evaluated yet.

4.4 Clinical experiences with real-time PCRs

A TaqMan probe design enabled blaKPC detection in DNA extracted directly from rectal/perianal swabs using two automated systems (i.e., NucliSENS easyMAG and MagNA Pure LC). The method was rapid and had a superior analytical sensitivity compared to culture dependent methods (Hindiyeh, et al., 2008). Similarly, Singh et al. implemented a single PCR using two TaqMan probes, one designed for blaKPC and one for the 16S gene, for the rectal screening of KPC carriers. The method used overnight cultures enriched with ceftazidime and DNA template was obtained by boiling. The assay had both sensitivity and a specificity of 97% demonstrating a higher sensitivity compared to selective media CHROMAgar (77.3%) (Singh, et al., 2012). Hindiyeh et al. adapted the above TaqMan probe protocol for the detection of blaKPC with a multiplex real-time PCR including the RNaseP gene as control. This method was able to detect blaKPC from blood culture bottles with a sensitivity and specificity of 100%. The automated extraction of the DNA improved the performance of the assay and the method was performed in less than 4 hrs (Hindiyeh, et al., 2011).

4.5 Advantage and disadvantages of real-time PCR

Real-time PCR avoids time consuming steps (e.g., electrophoresis analysis) in comparison to end-point PCRs and in many cases it does not require DNA sequencing of the obtained products. For these reasons, the method is rapid and excludes experimental errors linked to post amplification procedures. Moreover, real-time PCR is very sensitive, especially when TaqMan/HYB probes or molecular beacons are used instead of the general fluorescent dyes. Indeed, the latter might intercalate into primer dimers producing ambiguous signals. The selective hybridization of these probes, and the possibility to label them with different fluorophores, allows greater specificity in multiplex conditions and easier resolution of amplicons compared to end-point approaches. On the other hand, the setup of a real-time PCR is more complicated because the Tm of primers and probes needs to be taken into consideration. However, probe sequences available from literature and implementation of commercial kits could facilitate the procedure.

Generally, the reproducibility of the results obtained with real-time PCR is the major limitation of this technique (Bustin, et al., 2009). However, some precautions in the experimental setup (especially standardizing the DNA preparation) and in the interpretation of the data make real-time PCR a powerful, sensitive, reliable alternative to end-point PCRs, and a cost effective choice compared to the other currently available molecular approaches.

The overall characteristics of real-time PCR make this molecular approach particularly suitable for diagnostic purposes. This is especially true for multiplex real-time PCRs able to simultaneously detect clinically important drug-resistance traits. In the future, this application could lead to: i) the early detection of antibiotic resistance genes in clinical samples, dramatically influencing the effectiveness of therapy; ii) a less extensive use of broad-spectrum antibiotics in the empirical treatment of serious infections; and iii) the rapid screening of ESBL and/or carbapenemase producers carriers. . This application enormously contributes to the implementation of hygienic measures, necessary for combating the spread of such mechanisms among Ent and other MDR Gram-negative pathogens.

5. MICROARRAY

Microarray is an ideal methodology that allows for simultaneous identification and partial characterization of a very large number of genes (even >1000) of interest. This technology consists of pre-designated oligonucleotide probes that are specifically bound and immobilized on the solid surface of an array. If the targeted allele of a pathogen is present, it is labeled and subsequently hybridized to the immobilized probe present on the matrix of the array; such successful reactions are then measured with dedicated scanners (Miller & Tang, 2009; Sibley, Peirano, & Church, 2012).

Several microarrays for the genotyping of bla genes have been developed in the past. For instance, Grimm et al. designed a rapid DNA microarray (i.e., TTR of ~6 hrs) for genotyping of TEM β-lactamases. The assay was able to identify SNPs of 96% of TEM variants which were related to ESBL phenotypes (Grimm, et al., 2004). The same authors further developed an array for blaTEM, blaSHV and blaCTX-M implementing 618 probes capable of covering the mutations responsible for 156 key amino acid substitutions. The validity of the tests was demonstrated detecting 58 out of 60 isolates previously characterized as ESBL producers by phenotypic tests (Leinberger, et al., 2010). In an assay designed to study blaTEM, blaSHV and blaCTX-M, Robtsova et al. implemented biotin and streptavidin horseradish peroxidase to label the bla genes and to detect (by colorimetric reaction) their presence on the matrix, respectively. The system correctly detected 90 ESBL producers previously characterized with phenotypic and molecular methods (Rubtsova, Ulyashova, Edelstein, & Egorov, 2010).

Although the above in-house microarrays demonstrated excellent ability to detect and characterize the main blaESBL genes, the complete lack of probes able to detect further genes (e.g., those for carbapenemases) makes their use insufficient for the present epidemiology of MDR-Ent. More importantly, the implementation of those assays in other laboratories appears difficult, especially because of problems related to the standardization of the procedures. The advent of commercially available microarrays addresses this concern.

5.1 Commercially available microarrays

Check-Points has developed several rapid DNA microarray tests based on a ligation-mediated amplification reaction to detect the most important bla genes carried by Gram-negative pathogens (www.check-points.com). The results are automatically interpreted by the software provided by the manufacturer and the overall TTR starting from fresh colonies is very rapid (e.g., 50 strains in ~8 hrs) (Endimiani, Hujer, Hujer, Gatta, et al., 2010).

Cohen Stuart et al. evaluated the first microarray platform designed by the company to detect blaTEM, blaSHV, and blaCTX-M in a collection of 212 Ent (of which 106 were ESBL producers). In particular, specific probes were designed to distinguish non-ESBL and ESBL variants of TEM and SHV (i.e., SNPs determining substitutions E104K, R164S/C/H, G238S for TEMs and D179A/N/G, G238S/A, E240K for SHVs) and the CTX-M-1, -2, -9, and -8/25 groups. After comparison with the PCR/DNA sequencing data, the sensitivity and specificity of the microarray resulted 95% and 100%, respectively (Cohen Stuart, et al., 2010). In the same period, we evaluated an updated version (Check KPC/ESBL) of the microarray able to further detect the blaKPC. A total of 106 Gram-negative strains (including 57 KPC producers) were tested showing an overall agreement of 91.5% with standard DNA sequencing. The following sensitivity and specificity results were also recorded, respectively: blaSHV (98.8% and 100%), blaTEM (100% and 96.4%), blaCTX-M and blaKPC (both 100% and 100%) (Endimiani, Hujer, Hujer, Gatta, et al., 2010). Later, consistent results (i.e., sensitivity of 97% and specificity of 98%) were also obtained for 346 clinical Ent collected from 31 clinical laboratories in the Netherlands in 2009 (Platteel, et al., 2011), and for 125 Gram-negatives collected in one French hospital (i.e., 95% of agreement) (Naas, Cuzon, Truong, Bernabeu, & Nordmann, 2010). The Check KPC/ESBL was recently implemented to perform the molecular epidemiology of 1,051 Ent collected over the world during the longitudinal Study for Monitoring Antimicrobial Resistance Trends (SMART; 2008-2009). The assay demonstrated to be a rapid and very accurate system to screen for blaESBLs and blaKPC genes and to guide further PCR/sequencing analyses (Lascols, et al., 2012).

More recently, new Check-Points platforms have been designed to detect further clinically relevant bla genes. The Check-MDR CT101 assay detects blaSHV, blaTEM, blaCTX-M and blaKPC as Check KPC/ESBL, but also blapAmpCs (blaCMY, blaDHA, blaACT/MIR, blaACC) and blaNDM. Bogaerts et al. evaluated this microarray testing 207 clinical and reference strains of Ent showing an agreement with PCR/DNA sequencing of 100% for both blapAmpCs and blaNDM. As expected, false positive pAmpC results were recorded for cAmpC producers (e.g., Enterobacter spp. and C. freundii) (Bogaerts, et al., 2011). The Check-MDR CT102 platform detects blaSHV, blaTEM, and blaCTX-M, but also possesses the ability to detect more carbapenemase genes (i.e., blaVIM, blaIMP, and the blaOXA-48-like) rather than only blaKPC and blaNDM. However, it does not contain probes to detect the blapAmpCs. Naas et al. tested 144 Gram-negative strains possessing various bla genes indicating that this assay enables accurate identification of common ESBL and carbapenemase genes (i.e., both sensitivity and specificity of 100%; agreement with standard sequencing of 99%) (Naas, Cuzon, Bogaerts, Glupczynski, & Nordmann, 2011). So far, the Check-MDR CT103 is the most complete microarray platform provided by Check-Points. It merges the characteristics of CT101 and CT102 making the platform able to identify the most important non-ESBL, ESBL, pAmpCs, and carbapenemase genes spreading among Ent. Cuzon et al. tested 187 well-characterized Gram-negative bacilli showing sensitivity and specificity of both 100% for each single gene tested; moreover, all blaCTX-M were classified into the appropriate family group (Cuzon, Naas, Bogaerts, Glupczynski, & Nordmann, 2012).

Batchelor et al. developed a miniaturized microarray for the detection of 47 resistance genes in Gram-negative bacteria. In particular, this assay was able to detect traits encoding resistance to aminoglycosides [AMEs: aadA1/2/4-like, aac(3)-Ia/-IVa, aac(6’)-Ib, ant(2’’)-Ia], quinolones (qnr), trimethoprim (dfrA1/7/14/17/19, dfr12), sulphonamides (sul1-3), chloramphenicol (cmlA1-like, catA1, catIII, catB3-like, floR), tetracyclines [tet(A)-(G)], β-lactams (blaPSE, blaDHA, blaACC, blaMOX, blaCMY, blaFOX, blaSHV, blaLEN, blaTEM, blaOXA, blaCTX-M) and integrons (intI1-2). The performance of the assay for 87 Ent was compared to the PCR analysis showing a correlation of 98.8% (Batchelor, et al., 2008). This microarray was the ancestor for other improved platforms that are provided by Identibac (http://identibac.com). The most updated assay contains probes to detect numerous virulence factors of E. coli and further important resistance alleles (e.g., 16S rRNA methylases, qepA, blaPER, blaVEB, blaOXA-23/-40/-48/-51/-58, blaSPM, blaNDM, blaIMP, blaVIM, and blaKPC). However, though the system detects an enormous number of clinically important genes (including those for non-β-lactam antibiotics), it is still unable to distinguish between ESBLs and non-ESBLs in the SHV and TEM families. Moreover, a careful validation of the assay testing large collections of well-characterized isolates is still lacking.

5.2 Clinical implementation of the microarrays

The Check KPC/ESBL was recently used by Fishbain et al. to detect bla genes in 60 positive blood culture bottles with Ent (n=40; of which 20 possessing blaESBLs and/or blaKPC) or non-fermenters (n=20; negative controls). Aliquots of 0.5 ml were withdrawn from the bottles and, after removing erythrocytes by centrifugation, DNA extraction was performed. The overall concordance rate between phenotypic tests and microarray was 98.8% (59/60) when the water lysis method was implemented to perform DNA extraction. This approach could reduce the notification time of ESBL and/or KPC production by 18-20 hrs (Fishbain, Sinyavskiy, Riederer, Hujer, & Bonomo, 2012). More recently, Wintermans et al. showed that the implementation of the first Check-Points microarray (i.e., for detecting blaSHV, blaTEM, and blaCTX-M) instead of phenotypic tests significantly reduced the costs for unnecessary isolation for patients (i.e. those suspected to be colonized with ESBL producers) (Wintermans, Reuland, Wintermans, Bergmans, & Kluytmans, 2012). Peter et al. evaluated an in-house DNA microarray assay to directly detect all variants of blaKPC from simulated urine samples. The system was able to distinguish the different KPC-types with a TTR of 5-6 hrs (Peter, et al., 2012).

5.3 Advantage and disadvantages of microarrays

Microarrays possess greater analytical capacity compared to most of the other molecular methods (e.g., those PCR-based) because they can analyze a larger number of target genes in one reaction (Cuzon, et al., 2012). Moreover, assays designed to detect different SNPs in the same genes may be able to suggest the simultaneous presence of allelic variants (e.g., blaSHV-11 and blaSHV-12 in K. pneumoniae) (Endimiani, Hujer, Hujer, Gatta, et al., 2010). This makes their implementation to screen large collections of isolates very appealing (Lascols, et al., 2012). However, the system can also find application in a clinical practice (e.g., for BSIs), especially in a context with a high prevalence of MDR-Ent (Fishbain, et al., 2012). The methodology is easy to perform (especially for the commercially available platforms) and to be updated. On the other hand, the TTR is moderately high (6-8 hrs including DNA extraction) and the cross reaction of the pAmpCs probes with the cAmpCs of some organisms is a problem that is difficult to solve (Batchelor, et al., 2008; Bogaerts, et al., 2011). Devices and commercial kits are also quite expensive but, due to the rapid commercialization of this promising methodology, more affordable prices are expected in the near future (Table 4).

6. MASS SPECTROMETRY-BASED METHODOLOGIES

6.1 MALDI-TOF MS

Matrix-Assisted Laser Desorption Ionization-Time of Flight Mass Spectroscopy (MALDI-TOF MS) allows clinical microbiologists to dissect unique protein signatures (e.g., ribosomal proteins) of bacteria in order to identify the pathogen (Wieser, Schneider, Jung, & Schubert, 2012); in addition, there is a potential for the detection of resistance mechanisms. However, the use of MALDI-TOF MS for resistance mechanisms detection is in its early evaluation state (Hrabak, Chudackova, & Walkova, 2013). Currently, multiple applications of MALDI-TOF MS are being pursued for the detection of resistance mechanisms within Ent. These applications include: identification of the antibiotic itself and its modified/degradated counterparts (Burckhardt & Zimmermann, 2011; Hooff, et al., 2012; Hrabak, et al., 2012; Hrabak, Walkova, Studentova, Chudackova, & Bergerova, 2011; Sparbier, Schubert, Weller, Boogen, & Kostrzewa, 2012), detection of the resistance proteins within the cell (Cai, Hu, Zhang, Zhou, & Chen, 2012; Camara & Hays, 2007; Schaumann, et al., 2012), and discovery of mutations within resistance genes through mini-sequencing (Ikryannikova, et al., 2008).

Several studies have assessed the utility of MALDI-TOF MS for the identification of β-lactams and β-lactam degradation products (Burckhardt & Zimmermann, 2011; Hooff, et al., 2012; Hrabak, et al., 2012; Hrabak, et al., 2011; Sparbier, et al., 2012). Typically, β-lactams are incubated with bacterial cultures, the cultures are centrifuged, and the supernatants are analyzed for β-lactam degradation or modified products by MALDI-TOF MS; the length of this assay (not including growth of the cells) is about 1-4 hrs. This method is effective for identifying carbapenemase-producing Ent (Burckhardt & Zimmermann, 2011; Hrabak, et al., 2012; Hrabak, et al., 2011). In addition, β-lactamase inhibitors can be added to the incubation to determine their efficacy against the strains being tested (Hooff, et al., 2012; Sparbier, et al., 2012). However, this method can only detect the presence of β-lactamases as a resistance mechanism. Other methods implementing MALTI-TOF MS are required to detect the presence of efflux pumps or changes in porins or PBPs, which also result in β-lactams resistance.

Detection of the actual proteins conferring resistance in Ent by MALDI-TOF MS is more problematic because the protein signatures of bacteria are very complex. Camara and Hays established a proof of principle for this method in their study in which they dissected the differences between a wild-type E. coli and an ampicillin-resistant strain (Camara & Hays, 2007). They identified a 29,000 atomic mass unit peak in the ampicillin-resistant E. coli that was not present in the wild-type strain and confirmed that this peak represented a β-lactamase. On the other hand, in the study of Schaumann et al., discrimination between β-lactamase producing strains and non-producing strains was found to be difficult as the accuracy rate of the method was just 70% (Schaumann, et al., 2012). Using less complex protein samples, such as OMP preparations from five carbapenem-susceptible and seven carbapenem-resistant K. pneumoniae isolates, MALDI-TOF MS showed loss of porin expression in several β-lactam resistant isolates (Cai, et al., 2012). Currently, MALDI-TOF MS of resistance proteins is more effective in less complex backgrounds or when using partially purified protein preparations.

Mini-sequencing using MALDI-TOF MS may detect SNPs within resistance genes. This method requires PCR of the resistance gene, extension, and purification, followed by MALDI-TOF MS analysis. As a proof of concept, Ikryannikova et al. focused on three nucleotide positions that correspond to the amino acid positions 104, 164, and 238, as mutations in these gene regions can confer an ESBL phenotype in TEM-type β-lactamases. The system identified polymorphisms accurately at these three sites (Ikryannikova, et al., 2008). In another investigation, Stürenburg et al. implemented the MALDI-TOF MS to analyze fragment masses of base-specific cleavage of blaSHV PCR products after in vitro-generated RNA transcription. Twenty-one Ent possessing different SHV-types were identified unambiguously and in perfect accordance to the PCR/sequencing results (Sturenburg, et al., 2006).

6.1.1 Advantage and disadvantages of MALDI-TOF MS

The unique advantage of MALDI-TOF MS is that it is fast and inexpensive (Emonet, Shah, Cherkaoui, & Schrenzel, 2010; Ho & Reddy, 2011; Lavigne, et al., 2012). However, even if more rapid, its implementation to detect β-lactam degradation or modified products is just in addition to the second-line phenotypic tests (e.g., synergy with clavulanate or the mHT). In fact, this approach defines only if an enzyme with ESBL or carbapenemase hydrolytic spectrum is expressed by the tested pathogens (Burckhardt & Zimmermann, 2011; Hrabak, et al., 2011). We note that this information can be easily obtained implementing rapid (<2 hrs) and cost-effective phenotypic tests, such as those recently developed using phenol red in combination with different β-lactams and β-lactam/β-lactamase inhibitor combinations (Nordmann, Dortet, & Poirel, 2012; Nordmann, Poirel, & Dortet, 2012). Moreover, it is still unclear how sensitive and specific MALDI-TOF MS is for enzymes with poor affinity (Km) and very low turnover numbers (kcat). Finally, the use of MALDI TOF MS to perform mini-sequencing is, so far, too difficult and time consuming to be a candidate for replacing standard DNA sequencing methods. On the other hand, implementation of MALDI-TOF MS for detecting the specific atomic mass peak of β-lactamases may be a future application. However, the mass of the detected enzymes should be defined with very high resolution and accuracy guaranteeing distinction of important variants (e.g., the non-ESBL SHV-1: 28,872 vs. the ESBL SHV-2: 28,900). These data should also be coupled by databases able to match molecular weights with corresponding β-lactamases expressed by the tested pathogen (Table 4).

6.2 PCR/ESI MS

The PCR/Electrospray Ionization Mass Spectrometry (PCR/ESI MS) is a methodology that performs mini-sequencing of small PCR products (100-450 bp) measuring their exact molecular mass and interpreting such data with an advanced software. This system has demonstrated success in the identification of pathogens (Emonet, et al., 2010; Ho & Reddy, 2011; Lavigne, et al., 2012; Wolk, Kaleta, & Wysocki, 2012), and unlike MALDI-TOF MS has been developed into a fully automated system (PLEX-ID; Abbott Biosciences) with results available within 4-6 hrs starting from clinical samples. Like MALDI-TOF MS, the application of PCR/ESI MS for the detection of resistance mechanisms and SNPs within resistance genes in Ent is in its early stages (Endimiani, et al., 2011). An important study revealed the utility of PCR/ESI MS for detecting blaKPC expressing Ent. One hundred and ten previously characterized clinical isolates of which 69 carried blaKPC were used in this study. PCR/ESI MS correctly detected 100% of the blaKPC producing isolates. Control strains carrying blaKPC-4, blaKPC-5, blaKPC-6, blaKPC-7, and blaKPC-8 were also correctly identified, however the investigators could not discriminate between blaKPC-6, blaKPC-7, and blaKPC-8 (Endimiani, Hujer, Hujer, Sampath, et al., 2010).

6.2.1. Advantage and disadvantages of PCR/ESI MS

PCR/ESI MS represents a very promising technology. The rapid and fully automated ability to detect multiple genes (not only those conferring antibiotic resistance but also certain pathogenic factors) in one reaction may represent a revolution for the diagnosis of infectious diseases (D. J. Ecker, et al., 2008). This platform also offers the ability to detect pathogens at very low copy numbers and determine genetic relatedness based upon an MLST-like approach (J. A. Ecker, et al., 2006), thus providing “real-time” data essential to prevent outbreaks and spread of hyper-epidemic clones of MDR-Ent. Unlike other MS type applications, the detection of DNA from organisms that are non-cultured is a major step forward in the diagnosis of “culture negative infections”. Unfortunately, PCR/ESI MS is still under development and currently only available for research applications. Moreover, due to its enormous technological complexity, this instrument and its diagnostic kits are very expensive (Table 4).

7. MISCELLANEOUS

Faria-Ramos et al. developed a rapid detection assay of ESBL producers based on flow cytometric analysis. Bacterial cells were incubated with ceftazidime or cefotaxime for 1-2 hrs in presence or absence of clavulanate and then stained with a fluorescent dye able to diffuse across depolarized membranes. After incubating, non-ESBL isolates displayed an increased fluorescence, whereas those producing ESBLs showed this phenomenon only if incubated with clavulanate. The assay is rapid (~3 hrs) and correlates well with standard methods but it cannot be performed with bacteria in stationary growth phase. Moreover, it performs only a generic identification of ESBL and non-ESBL producers similarly to the phenotypic tests (Faria-Ramos, et al., 2013). Fujita et al. evaluated the usefulness of microchip gel electrophoresis (MGE) to analyze the results of a single PCR for blaCTX-M performed on positive blood cultures. All CTX-M producers tested were correctly identified, suggesting that this approach is rapid (1.5 hrs starting from DNA template) and cost-effective (US $6.50 per specimen). However, this system indicates only the presence of blaCTX-M (similarly to the classic single PCR) but it is unable to distinguish allelic variants (Fujita, et al., 2011). The denaturing high-performance liquid chromatography (dHPLC) is a powerful methodology that has been extensively used to detect allelic variation (Xiao & Oefner, 2001). It consists of separated multiplex PCRs, heteroduplex formation between reference and sample PCR amplicons, and analysis of heteroduplex DNA by dHPLC. Xu et al. implemented this technique to genotype members of all blaCTX-M groups. All of the 73 CTX-M-producing strains tested during the study were correctly characterized with results consistent with standard DNA sequencing (Xu, Evans, Ling, Nye, & Hawkey, 2007). Idaho Technology developed the FilmArray technique, a closed, very rapid (1 hr) and fully automated system that combines DNA extraction from clinical sample, nested multiplex PCRs, post-PCR amplicons melt curve analysis and automated readout of results (Poritz, et al., 2011). This method was recently evaluated in blood culture to rapidly identify the causative pathogens and to detect several antibiotic resistance genes, including blaKPC. However, data regarding the ability to detect KPC producers was not provided (Blaschke, et al., 2012).

8. CONCLUSIONS

This review provides a summary of the existing technologies that can be applied to the molecular diagnosis of infections due to MDR-Ent. On the other hand, this work also shows that we are at the threshold of a “new dawn” in clinical microbiology. Each method described above promises early and more sensitive detection. This knowledge inspires us to believe that the results obtained will have an impact on clinical outcomes (reduce mortality and morbidity, minimize resistance by stewardship) and healthcare costs. Key applications here will be for the early detection of MDR organisms in severe infections (e.g., sepsis) and the recognition of colonized patients (e.g., consequences for hospital hygiene). More importantly, the methods described above have the potential to improve the sensitivity of current automated systems by detecting pathogens in settings that were previously reported to be “culture negative”. The field of clinical microbiology has long awaited this promise as the techniques we use today still reflect in large part the methods developed by Louis Pasteur. Our challenge will be to accurately and carefully use this knowledge to show that these promises can be realized. In addition, these novel molecular technologies will also lead to new insights into disease pathogenesis. It is true that we will need a considerable amount of expertise to work with these advanced approaches. However, it is also imperative for us to accept these challenges and to enact wise choices in the selection of diagnostics. We are not certain which method will best suit our needs, but, perhaps, “chance will favor a prepared mind”.

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 errorsmaybe discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

REFERENCES

  1. Akinci E, Vahaboglu H. Minor extended-spectrum β-lactamases. Expert Rev Anti Infect Ther. 2010;8:1251–1258. doi: 10.1586/eri.10.119. [DOI] [PubMed] [Google Scholar]
  2. Ambrozic Avgustin J, Keber R, Zerjavic K, Orazem T, Grabnar M. Emergence of the quinolone resistance-mediating gene aac(6′)-Ib-cr in extended-spectrum-beta-lactamase-producing Klebsiella isolates collected in Slovenia between 2000 and 2005. Antimicrob Agents Chemother. 2007;51:4171–4173. doi: 10.1128/AAC.01480-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Arya M, Shergill IS, Williamson M, Gommersall L, Arya N, Patel HR. Basic principles of real-time quantitative PCR. Expert Rev Mol Diagn. 2005;5:209–219. doi: 10.1586/14737159.5.2.209. [DOI] [PubMed] [Google Scholar]
  4. Avlami A, Bekris S, Ganteris G, Kraniotaki E, Malamou-Lada E, Orfanidou M, Paniara O, Pantazatou A, Papagiannitsis CC, Platsouka E, Stefanou I, Tzelepi E, Vagiakou H, Miriagou V. Detection of metallo-β-lactamase genes in clinical specimens by a commercial multiplex PCR system. J Microbiol Methods. 2010;83:185–187. doi: 10.1016/j.mimet.2010.08.014. [DOI] [PubMed] [Google Scholar]
  5. Batchelor M, Hopkins KL, Liebana E, Slickers P, Ehricht R, Mafura M, Aarestrup F, Mevius D, Clifton-Hadley FA, Woodward MJ, Davies RH, Threlfall EJ, Anjum MF. Development of a miniaturised microarray-based assay for the rapid identification of antimicrobial resistance genes in Gram-negative bacteria. Int J Antimicrob Agents. 2008;31:440–451. doi: 10.1016/j.ijantimicag.2007.11.017. [DOI] [PubMed] [Google Scholar]
  6. Bell JM, Turnidge JD, Andersson P. aac(6′)-Ib-cr genotyping by simultaneous high-resolution melting analyses of an unlabeled probe and full-length amplicon. Antimicrob Agents Chemother. 2010;54:1378–1380. doi: 10.1128/AAC.01476-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Bercot B, Poirel L, Nordmann P. Updated multiplex polymerase chain reaction for detection of 16S rRNA methylases: high prevalence among NDM-1 producers. Diagn Microbiol Infect Dis. 2011;71:442–445. doi: 10.1016/j.diagmicrobio.2011.08.016. [DOI] [PubMed] [Google Scholar]
  8. Birkett CI, Ludlam HA, Woodford N, Brown DF, Brown NM, Roberts MT, Milner N, Curran MD. Real-time TaqMan PCR for rapid detection and typing of genes encoding CTX-M extended-spectrum β-lactamases. J Med Microbiol. 2007;56:52–55. doi: 10.1099/jmm.0.46909-0. [DOI] [PubMed] [Google Scholar]
  9. Bisiklis A, Papageorgiou F, Frantzidou F, Alexiou-Daniel S. Specific detection of blaVIM and blaIMP metallo-β-lactamase genes in a single real-time PCR. Clin Microbiol Infect. 2007;13:1201–1203. doi: 10.1111/j.1469-0691.2007.01832.x. [DOI] [PubMed] [Google Scholar]
  10. Blaschke AJ, Heyrend C, Byington CL, Fisher MA, Barker E, Garrone NF, Thatcher SA, Pavia AT, Barney T, Alger GD, Daly JA, Ririe KM, Ota I, Poritz MA. Rapid identification of pathogens from positive blood cultures by multiplex polymerase chain reaction using the FilmArray system. Diagn Microbiol Infect Dis. 2012;74:349–355. doi: 10.1016/j.diagmicrobio.2012.08.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Bogaerts P, Hujer AM, Naas T, de Castro RR, Endimiani A, Nordmann P, Glupczynski Y, Bonomo RA. Antimicrob Agents Chemother. 2011;55:4457–4460. doi: 10.1128/AAC.00353-11. Multicenter evaluation of a new DNA microarray for rapid detection of clinically relevant bla genes from β-lactam-resistant gram-negative bacteria. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Boyd DA, Tyler S, Christianson S, McGeer A, Muller MP, Willey BM, Bryce E, Gardam M, Nordmann P, Mulvey MR. Complete nucleotide sequence of a 92-kilobase plasmid harboring the CTX-M-15 extended-spectrum β-lactamase involved in an outbreak in long-term-care facilities in Toronto, Canada. Antimicrob Agents Chemother. 2004;48:3758–3764. doi: 10.1128/AAC.48.10.3758-3764.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Bradford PA, Bratu S, Urban C, Visalli M, Mariano N, Landman D, Rahal JJ, Brooks S, Cebular S, Quale J. Emergence of carbapenem-resistant Klebsiella species possessing the class A carbapenem-hydrolyzing KPC-2 and inhibitor-resistant TEM-30 β-lactamases in New York City. Clin Infect Dis. 2004;39:55–60. doi: 10.1086/421495. [DOI] [PubMed] [Google Scholar]
  14. Brolund A, Wisell KT, Edquist PJ, Elfstrom L, Walder M, Giske CG. Development of a real-time SYBR Green PCR assay for rapid detection of acquired AmpC in Enterobacteriaceae. J Microbiol Methods. 2010;82:229–233. doi: 10.1016/j.mimet.2010.06.006. [DOI] [PubMed] [Google Scholar]
  15. Bueno MF, Francisco GR, O'Hara JA, de Oliveira Garcia D, Doi Y. Coproduction of 16S rRNA Methyltransferase RmtD or RmtG with KPC-2 and CTX-M Group Extended-Spectrum β-Lactamases in Klebsiella pneumoniae. Antimicrob Agents Chemother. 2013;57:2397–2400. doi: 10.1128/AAC.02108-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Burckhardt I, Zimmermann S. Using matrix-assisted laser desorption ionization-time of flight mass spectrometry to detect carbapenem resistance within 1 to 2.5 hours. J Clin Microbiol. 2011;49:3321–3324. doi: 10.1128/JCM.00287-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Bush K, Jacoby GA. Updated functional classification of β-lactamases. Antimicrob Agents Chemother. 2009;54:969–976. doi: 10.1128/AAC.01009-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Bush K, Pannell M, Lock JL, Queenan AM, Jorgensen JH, Lee RM, Lewis JS, Jarrett D. Detection systems for carbapenemase gene identification should include the SME serine carbapenemase. Int J Antimicrob Agents. 2013;41:1–4. doi: 10.1016/j.ijantimicag.2012.08.008. [DOI] [PubMed] [Google Scholar]
  19. Bustin SA, Benes V, Garson JA, Hellemans J, Huggett J, Kubista M, Mueller R, Nolan T, Pfaffl MW, Shipley GL, Vandesompele J, Wittwer CT. The MIQE guidelines: minimum information for publication of quantitative real-time PCR experiments. Clin Chem. 2009;55:611–622. doi: 10.1373/clinchem.2008.112797. [DOI] [PubMed] [Google Scholar]
  20. Cai JC, Hu YY, Zhang R, Zhou HW, Chen GX. Detection of OmpK36 porin loss in Klebsiella spp. by matrix-assisted laser desorption ionization-time of flight mass spectrometry. J Clin Microbiol. 2012;50:2179–2182. doi: 10.1128/JCM.00503-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Camara JE, Hays FA. Discrimination between wild-type and ampicillin-resistant Escherichia coli by matrix-assisted laser desorption/ionization time-of-flight mass spectrometry. Anal Bioanal Chem. 2007;389:1633–1638. doi: 10.1007/s00216-007-1558-7. [DOI] [PubMed] [Google Scholar]
  22. Capone A, Giannella M, Fortini D, Giordano A, Meledandri M, Ballardini M, Venditti M, Bordi E, Capozzi D, Balice MP, Tarasi A, Parisi G, Lappa A, Carattoli A, Petrosillo N. High rate of colistin resistance among patients with carbapenem-resistant Klebsiella pneumoniae infection accounts for an excess of mortality. Clin Microbiol Infect. 2013;19:E23–30. doi: 10.1111/1469-0691.12070. [DOI] [PubMed] [Google Scholar]
  23. Cattoir V, Poirel L, Rotimi V, Soussy CJ, Nordmann P. Multiplex PCR for detection of plasmid-mediated quinolone resistance qnr genes in ESBL-producing enterobacterial isolates. J Antimicrob Chemother. 2007;60:394–397. doi: 10.1093/jac/dkm204. [DOI] [PubMed] [Google Scholar]
  24. Chen L, Chavda KD, Mediavilla JR, Zhao Y, Fraimow HS, Jenkins SG, Levi MH, Hong T, Rojtman AD, Ginocchio CC, Bonomo RA, Kreiswirth BN. Multiplex real-time PCR for detection of an epidemic KPC-producing Klebsiella pneumoniae ST258 clone. Antimicrob Agents Chemother. 2012;56:3444–3447. doi: 10.1128/AAC.00316-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Chen L, Mediavilla JR, Endimiani A, Rosenthal ME, Zhao Y, Bonomo RA, Kreiswirth BN. Multiplex real-time PCR assay for detection and classification of Klebsiella pneumoniae carbapenemase gene (blaKPC) variants. J Clin Microbiol. 2011;49:579–585. doi: 10.1128/JCM.01588-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Cohen Stuart J, Dierikx C, Al Naiemi N, Karczmarek A, Van Hoek AH, Vos P, Fluit AC, Scharringa J, Duim B, Mevius D, Leverstein-Van Hall MA. Rapid detection of TEM, SHV and CTX-M extended-spectrum β-lactamases in Enterobacteriaceae using ligation-mediated amplification with microarray analysis. J Antimicrob Chemother. 2010;65:1377–1381. doi: 10.1093/jac/dkq146. [DOI] [PubMed] [Google Scholar]
  27. Coque TM, Baquero F, Canton R. Increasing prevalence of ESBL-producing Enterobacteriaceae in Europe. Euro Surveill. 2008;13 [PubMed] [Google Scholar]
  28. Cuzon G, Naas T, Bogaerts P, Glupczynski Y, Nordmann P. Evaluation of a DNA microarray for the rapid detection of extended-spectrum β-lactamases (TEM, SHV and CTX-M), plasmid-mediated cephalosporinases (CMY-2-like, DHA, FOX, ACC-1, ACT/MIR and CMY-1-like/MOX) and carbapenemases (KPC, OXA-48, VIM, IMP and NDM). J Antimicrob Chemother. 2012;67:1865–1869. doi: 10.1093/jac/dks156. [DOI] [PubMed] [Google Scholar]
  29. Dallenne C, Da Costa A, Decre D, Favier C, Arlet G. Development of a set of multiplex PCR assays for the detection of genes encoding important β-lactamases in Enterobacteriaceae. J Antimicrob Chemother. 2010;65:490–495. doi: 10.1093/jac/dkp498. [DOI] [PubMed] [Google Scholar]
  30. De Champs C, Chanal C, Sirot D, Baraduc R, Romaszko JP, Bonnet R, Plaidy A, Boyer M, Carroy E, Gbadamassi Mc, Laluque S, Oules O, Poupart MC, Villemain M, Sirot J. Frequency and diversity of Class A extended-spectrum β-lactamases in hospitals of the Auvergne, France: a 2 year prospective study. J Antimicrob Chemother. 2004;54:634–639. doi: 10.1093/jac/dkh395. [DOI] [PubMed] [Google Scholar]
  31. Diaz MA, Cooper RK, Cloeckaert A, Siebeling RJ. Plasmid-mediated high-level gentamicin resistance among enteric bacteria isolated from pet turtles in Louisiana. Applied and environmental microbiology. 2006;72:306–312. doi: 10.1128/AEM.72.1.306-312.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Doi Y, Arakawa Y. 16S ribosomal RNA methylation: emerging resistance mechanism against aminoglycosides. Clin Infect Dis. 2007;45:88–94. doi: 10.1086/518605. [DOI] [PubMed] [Google Scholar]
  33. Doi Y, Paterson DL. Detection of plasmid-mediated class C β-lactamases. Int J Infect Dis. 2007;11:191–197. doi: 10.1016/j.ijid.2006.07.008. [DOI] [PubMed] [Google Scholar]
  34. Donnenberg M. Enterobacteriaceae. In: Mandell GL, Bennett JE, Dolin R, editors. Principles and practice of infectious diseases. Sixth Ed. Churchill Livingstone; 2005. pp. 2567–2586. [Google Scholar]
  35. Doumith M, Ellington MJ, Livermore DM, Woodford N. Molecular mechanisms disrupting porin expression in ertapenem-resistant Klebsiella and Enterobacter spp. clinical isolates from the UK. J Antimicrob Chemother. 2009;63:659–667. doi: 10.1093/jac/dkp029. [DOI] [PubMed] [Google Scholar]
  36. Doyle D, Peirano G, Lascols C, Lloyd T, Church DL, Pitout JD. Laboratory detection of Enterobacteriaceae that produce carbapenemases. J Clin Microbiol. 2012;50:3877–3880. doi: 10.1128/JCM.02117-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Drieux L, Brossier F, Sougakoff W, Jarlier V. Phenotypic detection of extended-spectrum β-lactamase production in Enterobacteriaceae: review and bench guide. Clin Microbiol Infect. 2008;14(Suppl 1):90–103. doi: 10.1111/j.1469-0691.2007.01846.x. [DOI] [PubMed] [Google Scholar]
  38. Ecker DJ, Sampath R, Massire C, Blyn LB, Hall TA, Eshoo MW, Hofstadler SA. Ibis T5000: a universal biosensor approach for microbiology. Nat Rev Microbiol. 2008;6:553–558. doi: 10.1038/nrmicro1918. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Ecker JA, Massire C, Hall TA, Ranken R, Pennella TT, Agasino Ivy C, Blyn LB, Hofstadler SA, Endy TP, Scott PT, Lindler L, Hamilton T, Gaddy C, Snow K, Pe M, Fishbain J, Craft D, Deye G, Riddell S, Milstrey E, Petruccelli B, Brisse S, Harpin V, Schink A, Ecker DJ, Sampath R, Eshoo MW. Identification of Acinetobacter species and genotyping of Acinetobacter baumannii by multilocus PCR and mass spectrometry. J Clin Microbiol. 2006;44:2921–2932. doi: 10.1128/JCM.00619-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Ellington MJ, Kistler J, Livermore DM, Woodford N. Multiplex PCR for rapid detection of genes encoding acquired metallo-β-lactamases. J Antimicrob Chemother. 2007;59:321–322. doi: 10.1093/jac/dkl481. [DOI] [PubMed] [Google Scholar]
  41. Emonet S, Shah HN, Cherkaoui A, Schrenzel J. Application and use of various mass spectrometry methods in clinical microbiology. Clin Microbiol Infect. 2010;16:1604–1613. doi: 10.1111/j.1469-0691.2010.03368.x. [DOI] [PubMed] [Google Scholar]
  42. Endimiani A, Depasquale JM, Forero S, Perez F, Hujer AM, Roberts-Pollack D, Fiorella PD, Pickens N, Kitchel B, Casiano-Colon AE, Tenover FC, Bonomo RA. Emergence of blaKPC-containing Klebsiella pneumoniae in a long-term acute care hospital: a new challenge to our healthcare system. J Antimicrob Chemother. 2009;64:1102–1110. doi: 10.1093/jac/dkp327. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Endimiani A, Hujer AM, Hujer KM, Gatta JA, Schriver AC, Jacobs MR, Rice LB, Bonomo RA. Evaluation of a commercial microarray system for detection of SHV-, TEM-, CTX-M-, and KPC-type β-lactamase genes in Gram-negative isolates. J Clin Microbiol. 2010;48:2618–2622. doi: 10.1128/JCM.00568-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Endimiani A, Hujer KM, Hujer AM, Kurz S, Jacobs MR, Perlin DS, Bonomo RA. Are we ready for novel detection methods to treat respiratory pathogens in hospital-acquired pneumonia? Clin Infect Dis. 2011;52(Suppl 4):S373–383. doi: 10.1093/cid/cir054. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Endimiani A, Hujer KM, Hujer AM, Sampath R, Ecker DJ, Bonomo RA. Rapid identification of blaKPC-possessing Enterobacteriaceae by PCR/electrospray ionization-mass spectrometry. J Antimicrob Chemother. 2010;65:1833–1834. doi: 10.1093/jac/dkq207. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Endimiani A, Luzzaro F, Brigante G, Perilli M, Lombardi G, Amicosante G, Rossolini GM, Toniolo A. Proteus mirabilis bloodstream infections: risk factors and treatment outcome related to the expression of extended-spectrum β-lactamases. Antimicrob Agents Chemother. 2005;49:2598–2605. doi: 10.1128/AAC.49.7.2598-2605.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Endimiani A, Luzzaro F, Perilli M, Lombardi G, Coli A, Tamborini A, Amicosante G, Toniolo A. Bacteremia due to Klebsiella pneumoniae isolates producing the TEM-52 extended-spectrum β-lactamase: treatment outcome of patients receiving imipenem or ciprofloxacin. Clin Infect Dis. 2004;38:243–251. doi: 10.1086/380645. [DOI] [PubMed] [Google Scholar]
  48. Endimiani A, Paterson DL. Optimizing therapy for infections caused by Enterobacteriaceae producing extended-spectrum β-lactamases. Semin Respir Crit Care Med. 2007;28:646–655. doi: 10.1055/s-2007-996411. [DOI] [PubMed] [Google Scholar]
  49. Endimiani A, Rossano A, Kunz D, Overesch G, Perreten V. First countrywide survey of third-generation cephalosporin-resistant Escherichia coli from broilers, swine, and cattle in Switzerland. Diagn Microbiol Infect Dis. 2012;73:31–38. doi: 10.1016/j.diagmicrobio.2012.01.004. [DOI] [PubMed] [Google Scholar]
  50. Endimiani A, Tamborini A, Luzzaro F, Lombardi G, Toniolo A. Epidemiology of bloodstream infections and time to detection of positive blood cultures: an evaluation of the automated BacT/Alert and BACTEC 9240 systems. New Microbiol. 2002;25:9–16. [PubMed] [Google Scholar]
  51. Erali M, Voelkerding KV, Wittwer CT. High resolution melting applications for clinical laboratory medicine. Exp Mol Pathol. 2008;85:50–58. doi: 10.1016/j.yexmp.2008.03.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Faria-Ramos I, Espinar MJ, Rocha R, Santos-Antunes J, Rodrigues AG, Canton R, Pina-Vaz C. A novel flow cytometric assay for rapid detection of extended-spectrum β-lactamases. Clin Microbiol Infect. 2013;19:E8–E15. doi: 10.1111/j.1469-0691.2012.03986.x. [DOI] [PubMed] [Google Scholar]
  53. Fishbain JT, Sinyavskiy O, Riederer K, Hujer AM, Bonomo RA. Detection of extended-spectrum β-lactamase and Klebsiella pneumoniae Carbapenemase genes directly from blood cultures by use of a nucleic acid microarray. J Clin Microbiol. 2012;50:2901–2904. doi: 10.1128/JCM.01023-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Fujita S, Yosizaki K, Ogushi T, Uechi K, Takemori Y, Senda Y. Rapid identification of gram-negative bacteria with and without CTX-M extended-spectrum β-lactamase from positive blood culture bottles by PCR followed by microchip gel electrophoresis. J Clin Microbiol. 2011;49:1483–1488. doi: 10.1128/JCM.01976-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Galimand M, Sabtcheva S, Courvalin P, Lambert T. Worldwide disseminated armA aminoglycoside resistance methylase gene is borne by composite transposon Tn1548. Antimicrob Agents Chemother. 2005;49:2949–2953. doi: 10.1128/AAC.49.7.2949-2953.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Gaynes R, Edwards JR. Overview of nosocomial infections caused by gram-negative bacilli. Clin Infect Dis. 2005;41:848–854. doi: 10.1086/432803. [DOI] [PubMed] [Google Scholar]
  57. Geyer CN, Reisbig MD, Hanson ND. Development of a TaqMan multiplex PCR assay for detection of plasmid-mediated ampC β-lactamase genes. J Clin Microbiol. 2012;50:3722–3725. doi: 10.1128/JCM.02038-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Giamarellou H, Poulakou G. Multidrug-resistant Gram-negative infections: what are the treatment options? Drugs. 2009;69:1879–1901. doi: 10.2165/11315690-000000000-00000. [DOI] [PubMed] [Google Scholar]
  59. Girlich D, Poirel L, Szczepanowski R, Schluter A, Nordmann P. Carbapenemhydrolyzing GES-5-encoding gene on different plasmid types recovered from a bacterial community in a sewage treatment plant. Applied and environmental microbiology. 2012;78:1292–1295. doi: 10.1128/AEM.06841-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Gniadkowski M. Evolution of extended-spectrum β-lactamases by mutation. Clin Microbiol Infect. 2008;14(Suppl 1):11–32. doi: 10.1111/j.1469-0691.2007.01854.x. [DOI] [PubMed] [Google Scholar]
  61. Grayson M,L. Kocers’ the Use of Antibiotics. 6th ed ASM Press; 2010. Cephalosporins and Related Drugs. [Google Scholar]
  62. Grimm V, Ezaki S, Susa M, Knabbe C, Schmid RD, Bachmann TT. Use of DNA microarrays for rapid genotyping of TEM β-lactamases that confer resistance. J Clin Microbiol. 2004;42:3766–3774. doi: 10.1128/JCM.42.8.3766-3774.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Guillard T, de Champs C, Moret H, Bertrand X, Scheftel JM, Cambau E. High-resolution melting analysis for rapid characterization of qnr alleles in clinical isolates and detection of two novel alleles, qnrB25 and qnrB42. J Antimicrob Chemother. 2012;67:2635–2639. doi: 10.1093/jac/dks292. [DOI] [PubMed] [Google Scholar]
  64. Guillard T, Duval V, Moret H, Brasme L, Vernet-Garnier V, de Champs C. Rapid detection of aac(6′)-Ib-cr quinolone resistance gene by pyrosequencing. J Clin Microbiol. 2010;48:286–289. doi: 10.1128/JCM.01498-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Guillard T, Moret H, Brasme L, Carlier A, Vernet-Garnier V, Cambau E, de Champs C. Rapid detection of qnr and qepA plasmid-mediated quinolone resistance genes using real-time PCR. Diagn Microbiol Infect Dis. 2011;70:253–259. doi: 10.1016/j.diagmicrobio.2011.01.004. [DOI] [PubMed] [Google Scholar]
  66. Haanpera M, Forssten SD, Huovinen P, Jalava J. Typing of SHV extended-spectrum β-lactamases by pyrosequencing in Klebsiella pneumoniae strains with chromosomal SHV β-lactamase. Antimicrob Agents Chemother. 2008;52:2632–2635. doi: 10.1128/AAC.01259-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Hammond DS, Schooneveldt JM, Nimmo GR, Huygens F, Giffard PM. bla(SHV) Genes in Klebsiella pneumoniae: different allele distributions are associated with different promoters within individual isolates. Antimicrob Agents Chemother. 2005;49:256–263. doi: 10.1128/AAC.49.1.256-263.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Hamprecht A, Poirel L, Gottig S, Seifert H, Kaase M, Nordmann P. Detection of the carbapenemase GIM-1 in Enterobacter cloacae in Germany. J Antimicrob Chemother. 2013;68:558–561. doi: 10.1093/jac/dks447. [DOI] [PubMed] [Google Scholar]
  69. Hansen LH, Jensen LB, Sorensen HI, Sorensen SJ. Substrate specificity of the OqxAB multidrug resistance pump in Escherichia coli and selected enteric bacteria. J Antimicrob Chemother. 2007;60:145–147. doi: 10.1093/jac/dkm167. [DOI] [PubMed] [Google Scholar]
  70. Hansen LH, Johannesen E, Burmolle M, Sorensen AH, Sorensen SJ. Plasmid-encoded multidrug efflux pump conferring resistance to olaquindox in Escherichia coli. Antimicrob Agents Chemother. 2004;48:3332–3337. doi: 10.1128/AAC.48.9.3332-3337.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Harris PN, Ferguson JK. Antibiotic therapy for inducible AmpC β-lactamase-producing Gram-negative bacilli: what are the alternatives to carbapenems, quinolones and aminoglycosides? Int J Antimicrob Agents. 2012;40:297–305. doi: 10.1016/j.ijantimicag.2012.06.004. [DOI] [PubMed] [Google Scholar]
  72. Hawser SP, Bouchillon SK, Hoban DJ, Badal RE, Canton R, Baquero F. Incidence and antimicrobial susceptibility of Escherichia coli and Klebsiella pneumoniae with extended-spectrum β-lactamases in community- and hospital-associated intra-abdominal infections in Europe: results of the 2008 Study for Monitoring Antimicrobial Resistance Trends (SMART). Antimicrob Agents Chemother. 2010;54:3043–3046. doi: 10.1128/AAC.00265-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Heid CA, Stevens J, Livak KJ, Williams PM. Real time quantitative PCR. Genome Res. 1996;6:986–994. doi: 10.1101/gr.6.10.986. [DOI] [PubMed] [Google Scholar]
  74. Hidalgo L, Hopkins KL, Gutierrez B, Ovejero CM, Shukla S, Douthwaite S, Prasad KN, Woodford N, Gonzalez-Zorn B. Association of the novel aminoglycoside resistance determinant RmtF with NDM carbapenemase in Enterobacteriaceae isolated in India and the UK. J Antimicrob Chemother. 2013 doi: 10.1093/jac/dkt078. [DOI] [PubMed] [Google Scholar]
  75. Higuchi R, Dollinger G, Walsh PS, Griffith R. Simultaneous amplification and detection of specific DNA sequences. Biotechnology (N Y) 1992;10:413–417. doi: 10.1038/nbt0492-413. [DOI] [PubMed] [Google Scholar]
  76. Higuchi R, Fockler C, Dollinger G, Watson R. Kinetic PCR analysis: real-time monitoring of DNA amplification reactions. Biotechnology (N Y) 1993;11:1026–1030. doi: 10.1038/nbt0993-1026. [DOI] [PubMed] [Google Scholar]
  77. Hilty M, Sendi P, Seiffert SN, Droz S, Perreten V, Hujer AM, Bonomo RA, Muhlemann K, Endimiani A. Characterisation and clinical features of Enterobacter cloacae bloodstream infections occurring at a tertiary care university hospital in Switzerland: is cefepime adequate therapy? Int J Antimicrob Agents. 2013;41:236–249. doi: 10.1016/j.ijantimicag.2012.10.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Hindiyeh M, Smollan G, Grossman Z, Ram D, Robinov J, Belausov N, Ben-David D, Tal I, Davidson Y, Shamiss A, Mendelson E, Keller N. Rapid detection of blaKPC carbapenemase genes by internally controlled real-time PCR assay using bactec blood culture bottles. J Clin Microbiol. 2011;49:2480–2484. doi: 10.1128/JCM.00149-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Hindiyeh M, Smollen G, Grossman Z, Ram D, Davidson Y, Mileguir F, Vax M, Ben David D, Tal I, Rahav G, Shamiss A, Mendelson E, Keller N. Rapid detection of blaKPC carbapenemase genes by real-time PCR. J Clin Microbiol. 2008;46:2879–2883. doi: 10.1128/JCM.00661-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Ho YP, Reddy PM. Advances in mass spectrometry for the identification of pathogens. Mass Spectrom Rev. 2011;30:1203–1224. doi: 10.1002/mas.20320. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Hong SS, Kim K, Huh JY, Jung B, Kang MS, Hong SG. Multiplex PCR for rapid detection of genes encoding class A carbapenemases. Ann Lab Med. 2012;32:359–361. doi: 10.3343/alm.2012.32.5.359. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Hooff GP, van Kampen JJ, Meesters RJ, van Belkum A, Goessens WH, Luider TM. Characterization of β-lactamase enzyme activity in bacterial lysates using MALDI-mass spectrometry. J Proteome Res. 2012;11:79–84. doi: 10.1021/pr200858r. [DOI] [PubMed] [Google Scholar]
  83. Hooper DC. Mechanisms of action of antimicrobials: focus on fluoroquinolones. Clin Infect Dis. 2001;32(Suppl 1):S9–S15. doi: 10.1086/319370. [DOI] [PubMed] [Google Scholar]
  84. Hopkins KL, Arnold C, Threlfall EJ. Rapid detection of gyrA and parC mutations in quinolone-resistant Salmonella enterica using Pyrosequencing technology. J Microbiol Methods. 2007;68:163–171. doi: 10.1016/j.mimet.2006.07.006. [DOI] [PubMed] [Google Scholar]
  85. Hornsey M, Phee L, Wareham DW. A novel variant, NDM-5, of the New Delhi metallo-β-lactamase in a multidrug-resistant Escherichia coli ST648 isolate recovered from a patient in the United Kingdom. Antimicrob Agents Chemother. 2011;55:5952–5954. doi: 10.1128/AAC.05108-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Hrabak J, Chudackova E, Walkova R. Matrix-assisted laser desorption ionization-time of flight (maldi-tof) mass spectrometry for detection of antibiotic resistance mechanisms: from research to routine diagnosis. Clin Microbiol Rev. 2013;26:103–114. doi: 10.1128/CMR.00058-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Hrabak J, Studentova V, Walkova R, Zemlickova H, Jakubu V, Chudackova E, Gniadkowski M, Pfeifer Y, Perry JD, Wilkinson K, Bergerova T. Detection of NDM-1, VIM-1, KPC, OXA-48, and OXA-162 carbapenemases by matrix-assisted laser desorption ionization-time of flight mass spectrometry. J Clin Microbiol. 2012;50:2441–2443. doi: 10.1128/JCM.01002-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Hrabak J, Walkova R, Studentova V, Chudackova E, Bergerova T. Carbapenemase activity detection by matrix-assisted laser desorption ionization-time of flight mass spectrometry. J Clin Microbiol. 2011;49:3222–3227. doi: 10.1128/JCM.00984-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Ikryannikova LN, Shitikov EA, Zhivankova DG, Il'ina EN, Edelstein MV, Govorun VM. A MALDI TOF MS-based minisequencing method for rapid detection of TEM-type extended-spectrum β-lactamases in clinical strains of Enterobacteriaceae. J Microbiol Methods. 2008;75:385–391. doi: 10.1016/j.mimet.2008.07.005. [DOI] [PubMed] [Google Scholar]
  90. Jacoby GA. Mechanisms of resistance to quinolones. Clin Infect Dis. 2005;41(Suppl 2):S120–126. doi: 10.1086/428052. [DOI] [PubMed] [Google Scholar]
  91. Jacoby GA. AmpC β-lactamases. Clin Microbiol Rev. 2009;22:161–182. doi: 10.1128/CMR.00036-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Jones CH, Ruzin A, Tuckman M, Visalli MA, Petersen PJ, Bradford PA. Pyrosequencing using the single-nucleotide polymorphism protocol for rapid determination of TEM- and SHV-type extended-spectrum β-lactamases in clinical isolates and identification of the novel β-lactamase genes blaSHV-48, blaSHV-105, and blaTEM-155. Antimicrob Agents Chemother. 2009;53:977–986. doi: 10.1128/AAC.01155-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Jorgensen RL, Nielsen JB, Friis-Moller A, Fjeldsoe-Nielsen H, Schonning K. Prevalence and molecular characterization of clinical isolates of Escherichia coli expressing an AmpC phenotype. J Antimicrob Chemother. 2010;65:460–464. doi: 10.1093/jac/dkp484. [DOI] [PubMed] [Google Scholar]
  94. Kaase M, Szabados F, Wassill L, Gatermann SG. Detection of carbapenemases in Enterobacteriaceae by a commercial multiplex PCR. J Clin Microbiol. 2012;50:3115–3118. doi: 10.1128/JCM.00991-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Karah N, Poirel L, Bengtsson S, Sundqvist M, Kahlmeter G, Nordmann P, Sundsfjord A, Samuelsen O. Plasmid-mediated quinolone resistance determinants qnr and aac(6′)-Ib-cr in Escherichia coli and Klebsiella spp. from Norway and Sweden. Diagn Microbiol Infect Dis. 2010;66:425–431. doi: 10.1016/j.diagmicrobio.2009.12.004. [DOI] [PubMed] [Google Scholar]
  96. Kim HB, Wang M, Park CH, Kim EC, Jacoby GA, Hooper DC. oqxAB encoding a multidrug efflux pump in human clinical isolates of Enterobacteriaceae. Antimicrob Agents Chemother. 2009;53:3582–3584. doi: 10.1128/AAC.01574-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Kishii R, Takei M. Relationship between the expression of ompF and quinolone resistance in Escherichia coli. J Infect Chemother. 2009;15:361–366. doi: 10.1007/s10156-009-0716-6. [DOI] [PubMed] [Google Scholar]
  98. Kruttgen A, Razavi S, Imohl M, Ritter K. Real-time PCR assay and a synthetic positive control for the rapid and sensitive detection of the emerging resistance gene New Delhi Metallo-β-lactamase-1 (blaNDM-1). Med Microbiol Immunol. 2011;200:137–141. doi: 10.1007/s00430-011-0189-y. [DOI] [PubMed] [Google Scholar]
  99. Kurpiel PM, Hanson ND. Association of IS5 with divergent tandem blaCMY-2 genes in clinical isolates of Escherichia coli. J Antimicrob Chemother. 2011;66:1734–1738. doi: 10.1093/jac/dkr212. [DOI] [PubMed] [Google Scholar]
  100. Landman D, Bratu S, Quale J. Contribution of OmpK36 to carbapenem susceptibility in KPC-producing Klebsiella pneumoniae. J Med Microbiol. 2009;58:1303–1308. doi: 10.1099/jmm.0.012575-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Lascols C, Hackel M, Hujer AM, Marshall SH, Bouchillon SK, Hoban DJ, Hawser SP, Badal RE, Bonomo RA. Using nucleic acid microarrays to perform molecular epidemiology and detect novel β-lactamases: a snapshot of extended-spectrum β-lactamases throughout the world. J Clin Microbiol. 2012;50:1632–1639. doi: 10.1128/JCM.06115-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Lascols C, Peirano G, Hackel M, Laupland KB, Pitout JD. Surveillance and molecular epidemiology of Klebsiella pneumoniae isolates that produce Carbapenemases: first report of OXA-48-like enzymes in North America. Antimicrob Agents Chemother. 2013;57:130–136. doi: 10.1128/AAC.01686-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Lavigne JP, Espinal P, Dunyach-Remy C, Messad N, Pantel A, Sotto A. Mass spectrometry: a revolution in clinical microbiology? Clin Chem Lab Med. 2012;0:1–14. doi: 10.1515/cclm-2012-0291. [DOI] [PubMed] [Google Scholar]
  104. Lee CH, Chu C, Liu JW, Chen YS, Chiu CJ, Su LH. Collateral damage of flomoxef therapy: in vivo development of porin deficiency and acquisition of blaDHA-1 leading to ertapenem resistance in a clinical isolate of Klebsiella pneumoniae producing CTX-M-3 and SHV-5 β-lactamases. J Antimicrob Chemother. 2007;60:410–413. doi: 10.1093/jac/dkm215. [DOI] [PubMed] [Google Scholar]
  105. Lee SY, Kotapati S, Kuti JL, Nightingale CH, Nicolau DP. Impact of extended-spectrum β-lactamase-producing Escherichia coli and Klebsiella species on clinical outcomes and hospital costs: a matched cohort study. Infect Control Hosp Epidemiol. 2006;27:1226–1232. doi: 10.1086/507962. [DOI] [PubMed] [Google Scholar]
  106. Leinberger DM, Grimm V, Rubtsova M, Weile J, Schroppel K, Wichelhaus TA, Knabbe C, Schmid RD, Bachmann TT. Integrated detection of extended-spectrum-β-lactam resistance by DNA microarray-based genotyping of TEM, SHV, and CTX-M genes. J Clin Microbiol. 2010;48:460–471. doi: 10.1128/JCM.00765-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Liu L, Li Y, Li S, Hu N, He Y, Pong R, Lin D, Lu L, Law M. Comparison of next-generation sequencing systems. J Biomed Biotechnol. 2012;2012:251364. doi: 10.1155/2012/251364. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Liu W, Zou D, Li Y, Wang X, He X, Wei X, Shao C, Li X, Shang W, Yu K, Liu D, Guo J, Yin Z, Yuan J. Sensitive and rapid detection of the new Delhi metallo-β-lactamase gene by loop-mediated isothermal amplification. J Clin Microbiol. 2012;50:1580–1585. doi: 10.1128/JCM.06647-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Livermore DM, Andrews JM, Hawkey PM, Ho PL, Keness Y, Doi Y, Paterson D, Woodford N. Are susceptibility tests enough, or should laboratories still seek ESBLs and carbapenemases directly? J Antimicrob Chemother. 2012;67:1569–1577. doi: 10.1093/jac/dks088. [DOI] [PubMed] [Google Scholar]
  110. Livermore DM, Canton R, Gniadkowski M, Nordmann P, Rossolini GM, Arlet G, Ayala J, Coque TM, Kern-Zdanowicz I, Luzzaro F, Poirel L, Woodford N. CTX-M: changing the face of ESBLs in Europe. J Antimicrob Chemother. 2007;59:165–174. doi: 10.1093/jac/dkl483. [DOI] [PubMed] [Google Scholar]
  111. Livermore DM, Warner M, Mushtaq S, Doumith M, Zhang J, Woodford N. What remains against carbapenem-resistant Enterobacteriaceae? Evaluation of chloramphenicol, ciprofloxacin, colistin, fosfomycin, minocycline, nitrofurantoin, temocillin and tigecycline. Int J Antimicrob Agents. 2011;37:415–419. doi: 10.1016/j.ijantimicag.2011.01.012. [DOI] [PubMed] [Google Scholar]
  112. Lou H, Chen M, Black SS, Bushell SR, Ceccarelli M, Mach T, Beis K, Low AS, Bamford VA, Booth IR, Bayley H, Naismith JH. Altered antibiotic transport in OmpC mutants isolated from a series of clinical strains of multi-drug resistant E. coli. PLoS One. 2011;6:e25825. doi: 10.1371/journal.pone.0025825. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Luzzaro F, Gesu G, Endimiani A, Ortisi G, Malandrin S, Pagani L, Rossolini GM. Performance in detection and reporting β-lactam resistance phenotypes in Enterobacteriaceae: a nationwide proficiency study in Italian laboratories. Diagn Microbiol Infect Dis. 2006;55:311–318. doi: 10.1016/j.diagmicrobio.2006.01.017. [DOI] [PubMed] [Google Scholar]
  114. Magnet S, Blanchard JS. Molecular insights into aminoglycoside action and resistance. Chem Rev. 2005;105:477–498. doi: 10.1021/cr0301088. [DOI] [PubMed] [Google Scholar]
  115. Manchanda V, Rai S, Gupta S, Rautela RS, Chopra R, Rawat DS, Verma N, Singh NP, Kaur IR, Bhalla P. Development of TaqMan real-time polymerase chain reaction for the detection of the newly emerging form of carbapenem resistance gene in clinical isolates of Escherichia coli, Klebsiella pneumoniae, and Acinetobacter baumannii. Indian J Med Microbiol. 2011;29:249–253. doi: 10.4103/0255-0857.83907. [DOI] [PubMed] [Google Scholar]
  116. Markoulatos P, Siafakas N, Moncany M. Multiplex polymerase chain reaction: a practical approach. J Clin Lab Anal. 2002;16:47–51. doi: 10.1002/jcla.2058. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Mathers AJ, Hazen KC, Carroll J, Yeh AJ, Cox HL, Bonomo RA, Sifri CD. First clinical cases of OXA-48-producing carbapenem-resistant Klebsiella pneumoniae in the United States: the “menace” arrives in the new world. J Clin Microbiol. 2013;51:680–683. doi: 10.1128/JCM.02580-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Maxam AM, Gilbert W. A new method for sequencing DNA. Proc Natl Acad Sci U S A. 1977;74:560–564. doi: 10.1073/pnas.74.2.560. [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. McDonald LC, Chen FJ, Lo HJ, Yin HC, Lu PL, Huang CH, Chen P, Lauderdale TL, Ho M. Emergence of reduced susceptibility and resistance to fluoroquinolones in Escherichia coli in Taiwan and contributions of distinct selective pressures. Antimicrob Agents Chemother. 2001;45:3084–3091. doi: 10.1128/AAC.45.11.3084-3091.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Meyer E, Schwab F, Schroeren-Boersch B, Gastmeier P. Dramatic increase of third-generation cephalosporin-resistant E. coli in German intensive care units: secular trends in antibiotic drug use and bacterial resistance, 2001 to 2008. Crit Care. 2010;14:R113. doi: 10.1186/cc9062. [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Michalopoulos A, Falagas ME. Treatment of Acinetobacter infections. Expert Opin Pharmacother. 2010;11:779–788. doi: 10.1517/14656561003596350. [DOI] [PubMed] [Google Scholar]
  122. Miller MB, Tang YW. Basic concepts of microarrays and potential applications in clinical microbiology. Clin Microbiol Rev. 2009;22:611–633. doi: 10.1128/CMR.00019-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Miro E, Grunbaum F, Gomez L, Rivera A, Mirelis B, Coll P, Navarro F. Characterization of Aminoglycoside-Modifying Enzymes in Enterobacteriaceae Clinical Strains and Characterization of the Plasmids Implicated in Their Diffusion. Microb Drug Resist. 2012 doi: 10.1089/mdr.2012.0125. [DOI] [PubMed] [Google Scholar]
  124. Monteiro J, Widen RH, Pignatari AC, Kubasek C, Silbert S. Rapid detection of carbapenemase genes by multiplex real-time PCR. J Antimicrob Chemother. 2012;67:906–909. doi: 10.1093/jac/dkr563. [DOI] [PubMed] [Google Scholar]
  125. Mroczkowska JE, Barlow M. Fitness trade-offs in blaTEM evolution. Antimicrob Agents Chemother. 2008;52:2340–2345. doi: 10.1128/AAC.00018-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Mullis KB, Faloona FA. Specific synthesis of DNA in vitro via a polymerase-catalyzed chain reaction. Methods Enzymol. 1987;155:335–350. doi: 10.1016/0076-6879(87)55023-6. [DOI] [PubMed] [Google Scholar]
  127. Murray KK. DNA sequencing by mass spectrometry. J Mass Spectrom. 1996;31:1203–1215. doi: 10.1002/(SICI)1096-9888(199611)31:11<1203::AID-JMS445>3.0.CO;2-3. [DOI] [PubMed] [Google Scholar]
  128. Naas T, Cotellon G, Ergani A, Nordmann P. Real-time PCR for detection of blaOXA-48 genes from stools. J Antimicrob Chemother. 2013;68:101–104. doi: 10.1093/jac/dks340. [DOI] [PubMed] [Google Scholar]
  129. Naas T, Cuzon G, Bogaerts P, Glupczynski Y, Nordmann P. Evaluation of a DNA microarray (Check-MDR CT102) for rapid detection of TEM, SHV, and CTX-M extended-spectrum β-lactamases and of KPC, OXA-48, VIM, IMP, and NDM-1 carbapenemases. J Clin Microbiol. 2011;49:1608–1613. doi: 10.1128/JCM.02607-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Naas T, Cuzon G, Truong H, Bernabeu S, Nordmann P. Evaluation of a DNA microarray, the check-points ESBL/KPC array, for rapid detection of TEM, SHV, and CTX-M extended-spectrum β-lactamases and KPC carbapenemases. Antimicrob Agents Chemother. 2010;54:3086–3092. doi: 10.1128/AAC.01298-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Naas T, Ergani A, Carrer A, Nordmann P. Real-time PCR for detection of NDM-1 carbapenemase genes from spiked stool samples. Antimicrob Agents Chemother. 2011;55:4038–4043. doi: 10.1128/AAC.01734-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Naas T, Massuard S, Garnier F, Nordmann P. AmpD is required for regulation of expression of NmcA, a carbapenem-hydrolyzing β-lactamase of Enterobacter cloacae. Antimicrob Agents Chemother. 2001;45:2908–2915. doi: 10.1128/AAC.45.10.2908-2915.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Naas T, Oxacelay C, Nordmann P. Identification of CTX-M-type extended-spectrum--lactamase genes using real-time PCR and pyrosequencing. Antimicrob Agents Chemother. 2007;51:223–230. doi: 10.1128/AAC.00611-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Nijhuis R, van Zwet A, Stuart JC, Weijers T, Savelkoul P. Rapid molecular detection of extended-spectrum β-lactamase gene variants with a novel ligation-mediated real-time PCR. J Med Microbiol. 2012;61:1563–1567. doi: 10.1099/jmm.0.047910-0. [DOI] [PubMed] [Google Scholar]
  135. Nordmann P, Dortet L, Poirel L. Rapid detection of extended-spectrum-β-lactamase-producing Enterobacteriaceae. J Clin Microbiol. 2012;50:3016–3022. doi: 10.1128/JCM.00859-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Nordmann P, Gniadkowski M, Giske CG, Poirel L, Woodford N, Miriagou V. Identification and screening of carbapenemase-producing Enterobacteriaceae. Clin Microbiol Infect. 2012;18:432–438. doi: 10.1111/j.1469-0691.2012.03815.x. [DOI] [PubMed] [Google Scholar]
  137. Nordmann P, Naas T, Poirel L. Global spread of Carbapenemase-producing Enterobacteriaceae. Emerg Infect Dis. 2011;17:1791–1798. doi: 10.3201/eid1710.110655. [DOI] [PMC free article] [PubMed] [Google Scholar]
  138. Nordmann P, Poirel L, Carrer A, Toleman MA, Walsh TR. How to detect NDM-1 producers. J Clin Microbiol. 2011;49:718–721. doi: 10.1128/JCM.01773-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  139. Nordmann P, Poirel L, Dortet L. Rapid detection of carbapenemase-producing Enterobacteriaceae. Emerg Infect Dis. 2012;18:1503–1507. doi: 10.3201/eid1809.120355. [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Nordmann P, Poirel L, Walsh TR, Livermore DM. The emerging NDM carbapenemases. Trends Microbiol. 2011;19:588–595. doi: 10.1016/j.tim.2011.09.005. [DOI] [PubMed] [Google Scholar]
  141. Notomi T, Okayama H, Masubuchi H, Yonekawa T, Watanabe K, Amino N, Hase T. Loop-mediated isothermal amplification of DNA. Nucleic Acids Res. 2000;28:E63. doi: 10.1093/nar/28.12.e63. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Oteo J, Delgado-Iribarren A, Vega D, Bautista V, Rodriguez MC, Velasco M, Saavedra JM, Perez-Vazquez M, Garcia-Cobos S, Martinez-Martinez L, Campos J. Emergence of imipenem resistance in clinical Escherichia coli during therapy. Int J Antimicrob Agents. 2008;32:534–537. doi: 10.1016/j.ijantimicag.2008.06.012. [DOI] [PubMed] [Google Scholar]
  143. Oxacelay C, Ergani A, Naas T, Nordmann P. Rapid detection of CTX-M-producing Enterobacteriaceae in urine samples. J Antimicrob Chemother. 2009;64:986–989. doi: 10.1093/jac/dkp336. [DOI] [PubMed] [Google Scholar]
  144. Papp-Wallace KM, Endimiani A, Taracila MA, Bonomo RA. Carbapenems: past, present, and future. Antimicrob Agents Chemother. 2011;55:4943–4960. doi: 10.1128/AAC.00296-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Park CH, Robicsek A, Jacoby GA, Sahm D, Hooper DC. Prevalence in the United States of aac(6′)-Ib-cr encoding a ciprofloxacin-modifying enzyme. Antimicrob Agents Chemother. 2006;50:3953–3955. doi: 10.1128/AAC.00915-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Park KS, Kim MH, Park TS, Nam YS, Lee HJ, Suh JT. Prevalence of the plasmid-mediated quinolone resistance genes, aac(6′)-Ib-cr, qepA, and oqxAB in clinical isolates of extended-spectrum β-lactamase (ESBL)-producing Escherichia coli and Klebsiella pneumoniae in Korea. Ann Clin Lab Sci. 2012;42:191–197. [PubMed] [Google Scholar]
  147. Paterson DL, Bonomo RA. Extended-spectrum β-lactamases: a clinical update. Clin Microbiol Rev. 2005;18:657–686. doi: 10.1128/CMR.18.4.657-686.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  148. Perez-Perez FJ, Hanson ND. Detection of plasmid-mediated AmpC β-lactamase genes in clinical isolates by using multiplex PCR. J Clin Microbiol. 2002;40:2153–2162. doi: 10.1128/JCM.40.6.2153-2162.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Perichon B, Courvalin P, Galimand M. Transferable resistance to aminoglycosides by methylation of G1405 in 16S rRNA and to hydrophilic fluoroquinolones by QepA-mediated efflux in Escherichia coli. Antimicrob Agents Chemother. 2007;51:2464–2469. doi: 10.1128/AAC.00143-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Perilli M, Celenza G, De Santis F, Pellegrini C, Forcella C, Rossolini GM, Stefani S, Amicosante G. E240V substitution increases catalytic efficiency toward ceftazidime in a new natural TEM-type extended-spectrum β-lactamase, TEM-149, from Enterobacter aerogenes and Serratia marcescens clinical isolates. Antimicrob Agents Chemother. 2008;52:915–919. doi: 10.1128/AAC.01028-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Peter H, Berggrav K, Thomas P, Pfeifer Y, Witte W, Templeton K, Bachmann TT. Direct detection and genotyping of Klebsiella pneumoniae carbapenemases from urine by use of a new DNA microarray test. J Clin Microbiol. 2012;50:3990–3998. doi: 10.1128/JCM.00990-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Pitout JD, Gregson DB, Poirel L, McClure JA, Le P, Church DL. Detection of Pseudomonas aeruginosa producing metallo-β-lactamases in a large centralized laboratory. J Clin Microbiol. 2005;43:3129–3135. doi: 10.1128/JCM.43.7.3129-3135.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Pitout JD, Hossain A, Hanson ND. Phenotypic and molecular detection of CTX-M-β-lactamases produced by Escherichia coli and Klebsiella spp. J Clin Microbiol. 2004;42:5715–5721. doi: 10.1128/JCM.42.12.5715-5721.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  154. Pitout JD, Wei Y, Church DL, Gregson DB. Surveillance for plasmid-mediated quinolone resistance determinants in Enterobacteriaceae within the Calgary Health Region, Canada: the emergence of aac(6′)-Ib-cr. J Antimicrob Chemother. 2008;61:999–1002. doi: 10.1093/jac/dkn068. [DOI] [PubMed] [Google Scholar]
  155. Platteel TN, Stuart JW, Voets GM, Scharringa J, van de Sande N, Fluit AC, Leverstein-Van Hall MA. Evaluation of a commercial microarray as a confirmation test for the presence of extended-spectrum β-lactamases in isolates from the routine clinical setting. Clin Microbiol Infect. 2011;17:1435–1438. doi: 10.1111/j.1469-0691.2011.03567.x. [DOI] [PubMed] [Google Scholar]
  156. Poirel L, Le Thomas I, Naas T, Karim A, Nordmann P. Biochemical sequence analyses of GES-1, a novel class A extended-spectrum β-lactamase, and the class 1 integron In52 from Klebsiella pneumoniae. Antimicrob Agents Chemother. 2000;44:622–632. doi: 10.1128/aac.44.3.622-632.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  157. Poirel L, Naas T, Nordmann P. Pyrosequencing as a rapid tool for identification of GES-type extended-spectrum β-lactamases. J Clin Microbiol. 2006;44:3008–3011. doi: 10.1128/JCM.02576-05. [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Poirel L, Potron A, Nordmann P. OXA-48-like carbapenemases: the phantom menace. J Antimicrob Chemother. 2012;67:1597–1606. doi: 10.1093/jac/dks121. [DOI] [PubMed] [Google Scholar]
  159. Poirel L, Walsh TR, Cuvillier V, Nordmann P. Multiplex PCR for detection of acquired carbapenemase genes. Diagn Microbiol Infect Dis. 2011;70:119–123. doi: 10.1016/j.diagmicrobio.2010.12.002. [DOI] [PubMed] [Google Scholar]
  160. Poritz MA, Blaschke AJ, Byington CL, Meyers L, Nilsson K, Jones DE, Thatcher SA, Robbins T, Lingenfelter B, Amiott E, Herbener A, Daly J, Dobrowolski SF, Teng DH, Ririe KM. FilmArray, an automated nested multiplex PCR system for multi-pathogen detection: development and application to respiratory tract infection. PLoS One. 2011;6:e26047. doi: 10.1371/journal.pone.0026047. [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Pournaras S, Poulou A, Tsakris A. Inhibitor-based methods for the detection of KPC carbapenemase-producing Enterobacteriaceae in clinical practice by using boronic acid compounds. J Antimicrob Chemother. 2010;65:1319–1321. doi: 10.1093/jac/dkq124. [DOI] [PubMed] [Google Scholar]
  162. Predari SC, Ligozzi M, Fontana R. Genotypic identification of methicillin-resistant coagulase-negative staphylococci by polymerase chain reaction. Antimicrob Agents Chemother. 1991;35:2568–2573. doi: 10.1128/aac.35.12.2568. [DOI] [PMC free article] [PubMed] [Google Scholar]
  163. Qiang YZ, Qin T, Fu W, Cheng WP, Li YS, Yi G. Use of a rapid mismatch PCR method to detect gyrA and parC mutations in ciprofloxacin-resistant clinical isolates of Escherichia coli. J Antimicrob Chemother. 2002;49:549–552. doi: 10.1093/jac/49.3.549. [DOI] [PubMed] [Google Scholar]
  164. Raghunathan A, Samuel L, Tibbetts RJ. Evaluation of a real-time PCR assay for the detection of the Klebsiella pneumoniae carbapenemase genes in microbiological samples in comparison with the modified Hodge test. Am J Clin Pathol. 2011;135:566–571. doi: 10.1309/AJCPPVNUI3O9JHJW. [DOI] [PubMed] [Google Scholar]
  165. Randegger CC, Keller A, Irla M, Wada A, Hachler H. Contribution of natural amino acid substitutions in SHV extended-spectrum β-lactamases to resistance against various β-lactams. Antimicrob Agents Chemother. 2000;44:2759–2763. doi: 10.1128/aac.44.10.2759-2763.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  166. Rapp RP, Urban C. Klebsiella pneumoniae carbapenemases in Enterobacteriaceae: history, evolution, and microbiology concerns. Pharmacotherapy. 2012;32:399–407. doi: 10.1002/j.1875-9114.2012.01035.x. [DOI] [PubMed] [Google Scholar]
  167. Riedel S, Carroll KC. Blood cultures: key elements for best practices and future directions. J Infect Chemother. 2010;16:301–316. doi: 10.1007/s10156-010-0069-1. [DOI] [PubMed] [Google Scholar]
  168. Robicsek A, Strahilevitz J, Sahm DF, Jacoby GA, Hooper DC. qnr prevalence in ceftazidime-resistant Enterobacteriaceae isolates from the United States. Antimicrob Agents Chemother. 2006;50:2872–2874. doi: 10.1128/AAC.01647-05. [DOI] [PMC free article] [PubMed] [Google Scholar]
  169. Rojo-Bezares B, Martin C, Lopez M, Torres C, Saenz Y. First detection of blaIMI-2 gene in a clinical Escherichia coli strain. Antimicrob Agents Chemother. 2012;56:1146–1147. doi: 10.1128/AAC.05478-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. Ronaghi M, Karamohamed S, Pettersson B, Uhlen M, Nyren P. Real-time DNA sequencing using detection of pyrophosphate release. Anal Biochem. 1996;242:84–89. doi: 10.1006/abio.1996.0432. [DOI] [PubMed] [Google Scholar]
  171. Rossolini GM, D'Andrea MM, Mugnaioli C. The spread of CTX-M-type extended-spectrum β-lactamases. Clin Microbiol Infect. 2008;14(Suppl 1):33–41. doi: 10.1111/j.1469-0691.2007.01867.x. [DOI] [PubMed] [Google Scholar]
  172. Rubtsova MY, Ulyashova MM, Edelstein MV, Egorov AM. Oligonucleotide microarrays with horseradish peroxidase-based detection for the identification of extended-spectrum β-lactamases. Biosens Bioelectron. 2010;26:1252–1260. doi: 10.1016/j.bios.2010.06.053. [DOI] [PubMed] [Google Scholar]
  173. Ruiz E, Ocampo-Sosa AA, Rezusta A, Revillo MJ, Roman E, Torres C, Martinez-Martinez L. Acquisition of carbapenem resistance in multiresistant Klebsiella pneumoniae strains harbouring blaCTX-M-15, qnrS1 and aac(6′)-Ib-cr genes. J Med Microbiol. 2012;61:672–677. doi: 10.1099/jmm.0.038083-0. [DOI] [PubMed] [Google Scholar]
  174. Sanger F, Nicklen S, Coulson AR. DNA sequencing with chain-terminating inhibitors. Proc Natl Acad Sci U S A. 1977;74:5463–5467. doi: 10.1073/pnas.74.12.5463. [DOI] [PMC free article] [PubMed] [Google Scholar]
  175. Schaumann R, Knoop N, Genzel GH, Losensky K, Rosenkranz C, Stingu CS, Schellenberger W, Rodloff AC, Eschrich K. A step towards the discrimination of β-lactamase-producing clinical isolates of Enterobacteriaceae and Pseudomonas aeruginosa by MALDI-TOF mass spectrometry. Med Sci Monit. 2012;18:MT71–77. doi: 10.12659/MSM.883339. [DOI] [PMC free article] [PubMed] [Google Scholar]
  176. Schrader C, Schielke A, Ellerbroek L, Johne R. PCR inhibitors - occurrence, properties and removal. J Appl Microbiol. 2012;113:1014–1026. doi: 10.1111/j.1365-2672.2012.05384.x. [DOI] [PubMed] [Google Scholar]
  177. Schwaber MJ, Navon-Venezia S, Kaye KS, Ben-Ami R, Schwartz D, Carmeli Y. Clinical and economic impact of bacteremia with extended- spectrum-β-lactamase-producing Enterobacteriaceae. Antimicrob Agents Chemother. 2006;50:1257–1262. doi: 10.1128/AAC.50.4.1257-1262.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  178. Seiffert SN, Hilty M, Perreten V, Endimiani A. Extended-spectrum cephalosporin-resistant gram-negative organisms in livestock: An emerging problem for human health? Drug Resist Updat. 2013 doi: 10.1016/j.drup.2012.12.001. [DOI] [PubMed] [Google Scholar]
  179. Sherry NL, Porter JL, Seemann T, Watkins A, Stinear TP, Howden BP. Outbreak investigation using high-throughput genome sequencing within a diagnostic microbiology laboratory. J Clin Microbiol. 2013 doi: 10.1128/JCM.03332-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  180. Sibley CD, Peirano G, Church DL. Molecular methods for pathogen and microbial community detection and characterization: current and potential application in diagnostic microbiology. Infect Genet Evol. 2012;12:505–521. doi: 10.1016/j.meegid.2012.01.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  181. Singh K, Mangold KA, Wyant K, Schora DM, Voss B, Kaul KL, Hayden MK, Chundi V, Peterson LR. Rectal screening for Klebsiella pneumoniae carbapenemases: comparison of real-time PCR and culture using two selective screening agar plates. J Clin Microbiol. 2012;50:2596–2600. doi: 10.1128/JCM.00654-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  182. Spanu T, Fiori B, D'Inzeo T, Canu G, Campoli S, Giani T, Palucci I, Tumbarello M, Sanguinetti M, Rossolini GM. Evaluation of the New NucliSENS EasyQ KPC test for rapid detection of Klebsiella pneumoniae carbapenemase genes (blaKPC). J Clin Microbiol. 2012;50:2783–2785. doi: 10.1128/JCM.00284-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  183. Sparbier K, Schubert S, Weller U, Boogen C, Kostrzewa M. Matrix-assisted laser desorption ionization-time of flight mass spectrometry-based functional assay for rapid detection of resistance against β-lactam antibiotics. J Clin Microbiol. 2012;50:927–937. doi: 10.1128/JCM.05737-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  184. Strahilevitz J, Jacoby GA, Hooper DC, Robicsek A. Plasmid-mediated quinolone resistance: a multifaceted threat. Clin Microbiol Rev. 2009;22:664–689. doi: 10.1128/CMR.00016-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Sturenburg E, Storm N, Sobottka I, Horstkotte MA, Scherpe S, Aepfelbacher M, Muller S. Detection and genotyping of SHV β-lactamase variants by mass spectrometry after base-specific cleavage of in vitro-generated RNA transcripts. J Clin Microbiol. 2006;44:909–915. doi: 10.1128/JCM.44.3.909-915.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  186. Swayne RL, Ludlam HA, Shet VG, Woodford N, Curran MD. Real-time TaqMan PCR for rapid detection of genes encoding five types of non-metallo- (class A and D) carbapenemases in Enterobacteriaceae. Int J Antimicrob Agents. 2011;38:35–38. doi: 10.1016/j.ijantimicag.2011.03.010. [DOI] [PubMed] [Google Scholar]
  187. Tangden T, Adler M, Cars O, Sandegren L, Lowdin E. Frequent emergence of porin-deficient subpopulations with reduced carbapenem susceptibility in ESBL-producing Escherichia coli during exposure to ertapenem in an in vitro pharmacokinetic model. J Antimicrob Chemother. 2013 doi: 10.1093/jac/dkt044. [DOI] [PubMed] [Google Scholar]
  188. Tong SY, Giffard PM. Microbiological applications of high-resolution melting analysis. J Clin Microbiol. 2012;50:3418–3421. doi: 10.1128/JCM.01709-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  189. Tsai YK, Fung CP, Lin JC, Chen JH, Chang FY, Chen TL, Siu LK. Klebsiella pneumoniae outer membrane porins OmpK35 and OmpK36 play roles in both antimicrobial resistance and virulence. Antimicrob Agents Chemother. 2011;55:1485–1493. doi: 10.1128/AAC.01275-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Tumbarello M, Sanguinetti M, Montuori E, Trecarichi EM, Posteraro B, Fiori B, Citton R, D'Inzeo T, Fadda G, Cauda R, Spanu T. Predictors of mortality in patients with bloodstream infections caused by extended-spectrum-β-lactamase-producing Enterobacteriaceae: importance of inadequate initial antimicrobial treatment. Antimicrob Agents Chemother. 2007;51:1987–1994. doi: 10.1128/AAC.01509-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  191. Tumbarello M, Spanu T, Di Bidino R, Marchetti M, Ruggeri M, Trecarichi EM, De Pascale G, Proli EM, Cauda R, Cicchetti A, Fadda G. Costs of bloodstream infections caused by Escherichia coli and influence of extended-spectrum-β-lactamase production and inadequate initial antibiotic therapy. Antimicrob Agents Chemother. 2010;54:4085–4091. doi: 10.1128/AAC.00143-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  192. Voets GM, Fluit AC, Scharringa J, Cohen Stuart J, Leverstein-van Hall MA. A set of multiplex PCRs for genotypic detection of extended-spectrum β-lactamases, carbapenemases, plasmid-mediated AmpC β-lactamases and OXA β-lactamases. Int J Antimicrob Agents. 2011;37:356–359. doi: 10.1016/j.ijantimicag.2011.01.005. [DOI] [PubMed] [Google Scholar]
  193. Walsh TR. Emerging carbapenemases: a global perspective. Int J Antimicrob Agents. 2010;36(Suppl 3):S8–14. doi: 10.1016/S0924-8579(10)70004-2. [DOI] [PubMed] [Google Scholar]
  194. Walther-Rasmussen J, Hoiby N. Class A carbapenemases. J Antimicrob Chemother. 2007;60:470–482. doi: 10.1093/jac/dkm226. [DOI] [PubMed] [Google Scholar]
  195. Wang M, Tran JH, Jacoby GA, Zhang Y, Wang F, Hooper DC. Plasmid-mediated quinolone resistance in clinical isolates of Escherichia coli from Shanghai, China. Antimicrob Agents Chemother. 2003;47:2242–2248. doi: 10.1128/AAC.47.7.2242-2248.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  196. Weigel LM, Steward CD, Tenover FC. gyrA mutations associated with fluoroquinolone resistance in eight species of Enterobacteriaceae. Antimicrob Agents Chemother. 1998;42:2661–2667. doi: 10.1128/aac.42.10.2661. [DOI] [PMC free article] [PubMed] [Google Scholar]
  197. Wieser A, Schneider L, Jung J, Schubert S. MALDI-TOF MS in microbiological diagnostics-identification of microorganisms and beyond (mini review). Appl Microbiol Biotechnol. 2012;93:965–974. doi: 10.1007/s00253-011-3783-4. [DOI] [PubMed] [Google Scholar]
  198. Williamson DA, Sidjabat HE, Freeman JT, Roberts SA, Silvey A, Woodhouse R, Mowat E, Dyet K, Paterson DL, Blackmore T, Burns A, Heffernan H. Identification and molecular characterisation of New Delhi metallo-β-lactamase-1 (NDM-1)- and NDM-6-producing Enterobacteriaceae from New Zealand hospitals. Int J Antimicrob Agents. 2012;39:529–533. doi: 10.1016/j.ijantimicag.2012.02.017. [DOI] [PubMed] [Google Scholar]
  199. Wintermans BB, Reuland EA, Wintermans RG, Bergmans AM, Kluytmans JA. The cost-effectiveness of ESBL detection: towards molecular detection methods? Clin Microbiol Infect. 2012 doi: 10.1111/j.1469-0691.2012.03998.x. [DOI] [PubMed] [Google Scholar]
  200. Wolk DM, Kaleta EJ, Wysocki VH. PCR-electrospray ionization mass spectrometry: the potential to change infectious disease diagnostics in clinical and public health laboratories. J Mol Diagn. 2012;14:295–304. doi: 10.1016/j.jmoldx.2012.02.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  201. Xiao W, Oefner PJ. Denaturing high-performance liquid chromatography: A review. Hum Mutat. 2001;17:439–474. doi: 10.1002/humu.1130. [DOI] [PubMed] [Google Scholar]
  202. Xu L, Evans J, Ling T, Nye K, Hawkey P. Rapid genotyping of CTX-M extended-spectrum β-lactamases by denaturing high-performance liquid chromatography. Antimicrob Agents Chemother. 2007;51:1446–1454. doi: 10.1128/AAC.01088-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  203. Yamane K, Wachino J, Suzuki S, Arakawa Y. Plasmid-mediated qepA gene among Escherichia coli clinical isolates from Japan. Antimicrob Agents Chemother. 2008;52:1564–1566. doi: 10.1128/AAC.01137-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Yan JJ, Wu JJ, Ko WC, Tsai SH, Chuang CL, Wu HM, Lu YJ, Li JD. Plasmid-mediated 16S rRNA methylases conferring high-level aminoglycoside resistance in Escherichia coli and Klebsiella pneumoniae isolates from two Taiwanese hospitals. J Antimicrob Chemother. 2004;54:1007–1012. doi: 10.1093/jac/dkh455. [DOI] [PubMed] [Google Scholar]
  205. Yue L, Jiang HX, Liao XP, Liu JH, Li SJ, Chen XY, Chen CX, Lu DH, Liu YH. Prevalence of plasmid-mediated quinolone resistance qnr genes in poultry and swine clinical isolates of Escherichia coli. Vet Microbiol. 2008;132:414–420. doi: 10.1016/j.vetmic.2008.05.009. [DOI] [PubMed] [Google Scholar]
  206. Zhou XY, Bordon F, Sirot D, Kitzis MD, Gutmann L. Emergence of clinical isolates of Escherichia coli producing TEM-1 derivatives or an OXA-1 β-lactamase conferring resistance to β-lactamase inhibitors. Antimicrob Agents Chemother. 1994;38:1085–1089. doi: 10.1128/aac.38.5.1085. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES