Next Article in Journal
Role of bZIP Transcription Factors in Plant Salt Stress
Next Article in Special Issue
PCR Array Profiling of miRNA Expression Involved in the Differentiation of Amniotic Fluid Stem Cells toward Endothelial and Smooth Muscle Progenitor Cells
Previous Article in Journal
The Evaluation of SHAPE-MaP RNA Structure Probing Protocols Reveals a Novel Role of Mn2+ in the Detection of 2′-OH Adducts
Previous Article in Special Issue
Somaclonal Variation—Advantage or Disadvantage in Micropropagation of the Medicinal Plants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Antimicrobial Resistance in Romania: Updates on Gram-Negative ESCAPE Pathogens in the Clinical, Veterinary, and Aquatic Sectors

by
Ilda Czobor Barbu
1,2,
Irina Gheorghe-Barbu
1,2,*,
Georgiana Alexandra Grigore
1,2,3,
Corneliu Ovidiu Vrancianu
1,2 and
Mariana Carmen Chifiriuc
1,2,4,5
1
Microbiology-Immunology Department, Faculty of Biology, University of Bucharest, 050095 Bucharest, Romania
2
The Research Institute of the University of Bucharest, 050095 Bucharest, Romania
3
National Institute of Research and Development for Biological Sciences, 060031 Bucharest, Romania
4
Academy of Romanian Scientists, 050044 Bucharest, Romania
5
Romanian Academy, 010071 Bucharest, Romania
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(9), 7892; https://doi.org/10.3390/ijms24097892
Submission received: 3 April 2023 / Revised: 20 April 2023 / Accepted: 21 April 2023 / Published: 26 April 2023
(This article belongs to the Special Issue State-of-the-Art Molecular Biology in Romania)

Abstract

:
Multidrug-resistant Gram-negative bacteria such as Acinetobacter baumannii, Pseudomonas aeruginosa, and members of the Enterobacterales order are a challenging multi-sectorial and global threat, being listed by the WHO in the priority list of pathogens requiring the urgent discovery and development of therapeutic strategies. We present here an overview of the antibiotic resistance profiles and epidemiology of Gram-negative pathogens listed in the ESCAPE group circulating in Romania. The review starts with a discussion of the mechanisms and clinical significance of Gram-negative bacteria, the most frequent genetic determinants of resistance, and then summarizes and discusses the epidemiological studies reported for A. baumannii, P. aeruginosa, and Enterobacterales-resistant strains circulating in Romania, both in hospital and veterinary settings and mirrored in the aquatic environment. The Romanian landscape of Gram-negative pathogens included in the ESCAPE list reveals that all significant, clinically relevant, globally spread antibiotic resistance genes and carrying platforms are well established in different geographical areas of Romania and have already been disseminated beyond clinical settings.

1. Introduction

Gram-negative bacteria represent one of the most significant pathogens involved in public health issues. Among them, members of the families Enterobacterales, Moraxellales, and Pseudomonadales have a major clinical significance, being key Gram-negative pathogens listed in the ESKAPE (Enterococcus faecium, Staphylococcus aureus, Klebsiella pneumoniae, Acinetobacter baumannii, Pseudomonas aeruginosa, and Enterobacter species) or later designated ESCAPE group (E. faecium, S. aureus, Clostridioides difficile, A. baumannii, P. aeruginosa, Enterobacterales), as well as in the WHO priority pathogens list for R&D of new antibiotics in Priority 1: CRITICAL list (namely carbapenem-resistant A. baumannii, carbapenem-resistant P. aeruginosa, and carbapenem-resistant, extended-spectrum β-lactamase (ESBL)-producing Enterobacterales) [1]. The Centers for Disease Control and Prevention (CDC) described the dramatic increase of antibiotic resistance (AR) during the last decades as one of the most critical threats to public health, with very few effective antimicrobials left, either novel or old molecules [2,3], therefore requiring concerted research and management efforts [4].
Gram-negative bacteria possess multiple AR mechanisms towards all antibiotic classes: besides intrinsic resistance (mainly due to their outer membrane, efflux pumps, and extra wall structures), they may acquire resistance via chromosomal mutations and horizontal gene transfer (HGT) [5]. Chromosomal mutations may alter some critical enzymes (for example, mutations in genes encoding DNA gyrases and topoisomerases lead to quinolone resistance) and regulatory proteins (as mutations in the mgrB small regulatory protein in K. pneumoniae lead to colistin resistance) that may be involved in the loss, downregulation, or alteration of porins or may lead to an increased expression level of efflux pumps. Nevertheless, the main mechanisms involved in AR in Gram-negative bacteria are represented by hydrolytic enzymes, especially β-lactamases, acquired via HGT [5].
In this context, we aimed to bring together all the available epidemiological studies reported for A. baumannii, P. aeruginosa, and Enterobacterales-resistant strains circulating in Romania in different environments (clinical, community, veterinary settings, wastewater, and surface water networks) in order to take a step further to illuminate the extent of the dissemination of antibiotic resistance genes (ARGs) in our country, taking into account that in Romania there has not yet been implemented a national action plan to combat the dissemination of AR.

2. Main Mechanisms of AR in Gram-Negative Pathogens

2.1. β-Lactam Resistance

Resistance to β-lactam antibiotics (penicillins, cephalosporins, carbapenems, monobactams, and β-lactamase inhibitors) is conferred by the production of antibiotic-modifying enzymes, efflux pumps, porins, protection of the antibiotic target, or biofilm production [6,7,8,9]. However, the primary mechanism of β-lactam resistance in Gram-negative bacilli is represented by the production of β-lactamases (serine-β-lactamases—Ambler classes A, C, and D, and metallo-β-lactamases or MBL—Ambler class B) [10,11,12].
Class A β-lactamases (except for Klebsiella pneumoniae carbapenemase—KPC) hydrolyze penicillins and cephalosporins more efficiently than carbapenems and are inhibited by clavulanic acid [13]. It includes narrow spectrum, or ESBLs, inhibited by clavulanic acid or tazobactam, and they confer penicillin, cephalosporins, monobactams, and carbapenem resistance [14] (Table 1). In addition, it is well known that ESBLs play an essential role in resistance against later-generation cephalosporins such as cefotaxime, ceftazidime, and cefepime [15]. Class B β-lactamases include metallo-β-lactamases (MBLs), broad-spectrum enzymes that require zinc or another heavy metal for hydrolysis β-lactam antibiotics except for monobactams [13]. Class C β-lactamases, also known as AmpC β-lactamases, confer resistance to cephamycins, penicillins, cephalosporins, and β-lactamase inhibitors [13], which can be: (a) inducible resistance via chromosomally encoded ampC genes (e.g., Enterobacter cloacae, Serratia marcescens, Citrobacter freundii, P. aeruginosa), (b) non-inducible chromosomal resistance due to promoter and/or attenuator mutations (e.g., Escherichia coli, Shigella spp., A. baumannii), or (c) plasmid-mediated resistance (e.g., Klebsiella pneumoniae, E. coli, Salmonella spp.). Class D β-lactamases include oxacillinases (OXAs) or carbapenem-hydrolyzing class D β-lactamases (CHDLs) that are serine β-lactamases that hydrolyze all β-lactam antibiotics and are not inhibited by clavulanic acid, sulbactam, or tazobactam [16].
AR caused by β-lactamase production may significantly increase when interacting with outer membrane proteins (OMPs) or efflux pumps. OMPs are monomeric or trimeric porins that mediate the diffusion of small molecules into or out of the periplasmic space of Gram-negative bacteria and can be classified as non-specific or specific according to their activity [17]. Efflux pumps actively control the intracellular concentration of antibiotic molecules with different chemical structures and are involved in multidrug resistance (MDR) mechanisms. The MDR efflux pumps belong to the ATP-binding cassette (ABC), major facilitator superfamily (MFS), small multidrug resistance (SMR), multidrug and toxic compound extrusion (MATE), proteobacterial antimicrobial compound efflux (PACE), and resistance-nodulation-cell division (RND) superfamilies and are the focus of current research [18].
In A. baumannii, β-lactam resistance is encoded by transferable β-lactamases from class A [19], class B [20,21], chromosomal class C β-lactamases (ADC), and chromosomal or plasmidial CHDLs (class D β-lactamases) (Table 1). The most characteristic β-lactamases in A. baumannii are represented by CHDLs: OXA-23, OXA-24/-40, OXA-58, OXA-143, and OXA-235 [14] (Table 1). The first OXA enzyme with carbapenemase activity in A. baumannii, ARI-1, was identified in a clinical strain isolated in Scotland and was subsequently renamed OXA-23 [22] (Table 1). This enzyme is disseminated worldwide due to its association with several transposons (Tn2006, Tn2007, Tn2009, Tn2008, and Tn2008B) [23].
In P. aeruginosa, there were described ESBLs from class A, represented mainly by VEB (Vietnamese extended-spectrum β-lactamase) and PER (Pseudomonas extended resistance), but also CTX-M (Cefotaxime first isolated in Munich), TEM (for Temoneira patient’s name), SHV (Sulfydryl Variable enzyme), and BEL (Belgium ESBLs) families [24,25,26,27]. Class A carbapenemases encountered in P. aeruginosa are represented by GES (Guyana Extended Spectrum β-lactamase) and KPC enzymes [28] (Table 1). Acquired MBLs in P. aeruginosa include the VIM (Verona imipenemase), IMP (active-on-imipenem), SPM (Saõ Paolo metallo-β-lactamase), SIM (Seoul imipenemase), FIM (Florence imipenemase), AIM (Australian imipenemase), and DIM (Dutch imipenemase) enzymes [29,30]. Class D comprises OXA enzymes, the name being derived from their preference for oxacillin and cloxacillin hydrolysis. They are mostly narrow-spectrum β-lactamases that confer resistance to cefotaxime or ceftazidime, with some OXA β-lactamases linked to resistance and/or reduced susceptibility to cefepime and/or aztreonam [30,31,32,33].
In Enterobacterales, the main mechanisms of β-lactam resistance are represented by the expression of class A ESBLs (mainly CTX-M and variants of TEM, SHV), classes A, B, and D carbapenemases [KPC, New Delhi MBL (NDM), OXA-48, IMP, VIM), or class C AmpC chromosomal or plasmidial enzymes [CMY (confer cephamycin resistance), FOX (cefoxitin resistance), MOX (moxalactam resistance), LAT (latamoxef resistance), ACC (Ambler class C), ACT (AmpC type), MIR-1 (Miriam Hospital in Providence), DHA (Dhahran hospital in Saudi Arabia)] [34,35].

2.2. Resistance to Other Antibiotic Classes

2.2.1. Aminoglycoside Resistance

Aminoglycoside resistance in Gram-negative bacilli is mainly encoded by enzymatic resistance mechanisms (by the production of aminoglycoside-modifying enzymes—AMEs) but also by alteration of the ribosome structure (by 16S rRNA methyltransferases) and limited antibiotic uptake (due to the loss of cell membrane permeability or hyperactivity of the efflux pumps) [36,37,38]. AMEs are classified as aminoglycoside phosphotransferases (APH), aminoglycoside acetyltransferases (AAC), aminoglycoside nucleotidyltransferases (ANT), and aminoglycoside adenylyltransferases (AAD) [36]. The most frequently detected AME-encoding genes in Gram-negative strains listed in the ESCAPE group are presented in Table 2.

2.2.2. Tetracycline Resistance

Tetracyclines bind to the 30S ribosomal subunit and inhibit protein synthesis, stopping translation [39]. Several mechanisms cause tetracycline resistance: efflux dependent on ATP (Tet(A), Tet(B), Tet(C), Tet(D), and Tet(G)) [40,41,42], ribosomal protection proteins (Tet(M), Tet(O), and Tet(H)) [14,43], target modification, and enzymatic antibiotic inactivation (Tet(X)) [44,45].
Table 1. The most frequent transmissible β-lactamases and corresponding encoding genes in Gram-negative rods.
Table 1. The most frequent transmissible β-lactamases and corresponding encoding genes in Gram-negative rods.
β-Lactamases Class (Ambler)Resistance MechanismsEncoding GenesSpeciesReferences
Class ANarrow spectrum β-lactamasesblaSCO-1;A. baumannii [46,47,48,49,50,51,52,53,54]
blaPSE-1/CARB-2; blaPSE-4/CARB-1;A. baumannii, P. aeruginosa;
blaCARB-3, -4; blaTEM-1, -2, blaSHV-1;Enterobacterales, A. baumannii, P. aeruginosa
Extended-spectrum β-lactamases (ESBLs)blaCARB-10; A. baumannii; [24,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74]
blaCTX-M-1, -2, -9, -14, -15, -27, -43, -92;
blaGES-1, -11, -13;
Enterobacterales, A. baumannii, P. aeruginosa;
blaTEM-3, -4, -12, -21, -24, -42, -116;
blaSHV-1, -2, -2a, -5, -12, -27, -41, -187;
Enterobacterales, A. baumannii, P. aeruginosa;
blaVEB-1, -2, -9; blaVEB-1;A. baumannii, P. aeruginosa;
blaBEL-1, -2; P. aeruginosa, Enterobacterales;
blaPER-1, -2, -7;A. baumannii, P. aeruginosa;
CarbapenemasesblaKPC-2, -3, -4, -5, -10;Enterobacterales, A. baumannii, P. aeruginosa; [68,75,76,77,78,79,80,81,82,83,84,85,86]
blaGES-2, -5, -6, -14, -15, -20;P. aeruginosa, Enterobacterales;
blaGPC-1;P. aeruginosa;
blaIMI-2, blaIMI-3;
blaSHV-38;
Enterobacterales
Class BCarbapenemasesblaNDM-1, -2, -5;;Enterobacterales, A. baumannii, P. aeruginosa; [20,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126]
blaVIM-1, -2, -3, -11;P. aeruginosa, Enterobacterales, A. baumannii;
blaGIMlike; blaSIM-1;A. baumannii, P. aeruginosa;
blaIMP-1, -2, -4, -5, -6, -8, -9, -11, -14a, -19, -27, -51, -55; Enterobacterales, A. baumannii, P. aeruginosa;
blaSPM-1; blaSMP-1; blaDIM; blaHMB-1; blaCAM; blaAIM-1; blaFIM-1A. baumannii
P. aeruginosa
Class CChromosomally-derived cephalosporinases blaADC-25, -30, 73; blaAmpC-69, -70, -71; blaADC-55, -67, -68, -196A. baumannii [127,128,129,130,131,132,133,134,135,136]
blaCMY, blaFOX, blaACT, blaDHA, blaACC;Enterobacterales, P. aeruginosa; [72,137,138]
Class DNarrow spectrumblaOXA-4, blaOXA-20, blaOXA-47; Enterobacterales, P. aeruginosa, A. baumannii; [139,140,141]
ESBLblaOXA-1, -31, blaOXA-2, -161, blaOXA-5, blaOXA-10, -11, -7; blaOXA-18, blaOXA-31; blaOXA-45, blaOXA-46;P. aeruginosa, Enterobacterales;
[31,32,142,143,144,145]
Carbapenem-hydrolyzing class D β-lactamases (CHDLs)blaOXA-51; blaOXA-23; blaOXA-24; blaOXA-58; blaOXA-143; blaOXA-235;
blaOXA-198;
blaOXA-48 and blaOXA-48-like (blaOXA-181, -162, -232);
A. baumannii;
P. aeruginosa; Enterobacterales
[57,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159]

2.2.3. Fluoroquinolone Resistance

Quinolones or fluoroquinolones are broad-spectrum bactericidal antibiotics that disrupt DNA replication by inhibiting the activity of type II topoisomerase, DNA gyrase, and topoisomerase IV. Fluoroquinolones primarily affect gyrase activity, while toxicity against topoisomerase IV is secondary. Resistance to fluoroquinolones in Gram-negative bacilli is mediated by the underexpression of porins or the overexpression of cellular efflux pumps [160], mutations in gyrase and topoisomerase IV (gyrA, gyrB, and parC) [161], and plasmid-mediated quinolone resistance (PMQR), encoded by qnr and oqx genes [160,161,162,163,164,165,166,167] (Table 2).
Table 2. The most frequent aminoglycosides, tetracyclines, fluoroquinolones, and, polymyxins transmissible resistance genes in Gram-negative rods.
Table 2. The most frequent aminoglycosides, tetracyclines, fluoroquinolones, and, polymyxins transmissible resistance genes in Gram-negative rods.
Antibiotic ClassAntibioticsARGsSpeciesReferences
Aminogly
cosides
AMEs
gentamicin, sisomicin, fortimicinaac(3)-I
a/b/c/e
A. baumannii, P. aeruginosa, E. coli, K. pneumoniae, K. aerogenes, E. cloacae, E. hormachei [14,36,78,168,169,170,171]
aac3-Ib-aac6′-Ib’P. aeruginosa
gentamicin, netilmicin, tobramycin, sisomicin, dibekacinaac(3)-II
a/c/d/e
E. coli, K. pneumoniae, Pseudomonas sp.
A. baumannii, E. cloacae complex
[14,172,173,174,175,176,177]
aac(3)-IVaA. baumannii, K. pneumoniae, E. coli [14,178]
amikacin, gentamicinaac(6′)-Ia
/i/l/q/ae/af/ai/30/33/aacA43
P. aeruginosa, E. coli, E. cloacae, E. hormachei, [14,36,168,179,180,181,182,183,184,185]
aac(6′)-Ib
/aacA4
P. aeruginosa
A. baumannii, K. pneumoniae, E. cloacae, E. coli
kanamycin, tobramycin, amikacin, ciprofloxacin, norfloxacinaac(6′)-Ib-crE. coli, E. cloacae complex, K. pneumoniae, A. baumannii, P. aeruginosa [186]
aac(6′)-If
aac(6′)-Ig/h/j/k/k/r/s/t/u/v/w/x/ad/ae//aa/ad/
aacA29
E. cloacae,
A. baumannii
P. aeruginosa
[36,187,188]
gentamicinaac(6′)-IIa/I31/32/cP. aeruginosa, K. pneumoniae, A. baumannii [36,182,188]
dibekacin, gentamicin, kanamycin, sisomicin, tobramycinant(2″)-IaA. baumannii, K. pneumoniae, E. coli [168,189,190,191,192]
spectinomycin, streptomycinant(3”)-IIA. baumannii [14]
amikacin, tobramycin
isepamicin
ant(4′)-IIa/bP. aeruginosa [191,192,193]
streptomycinaadA1,2,5,11,13,16E. coli, K. pneumoniae, E. cloacae, P. aeruginosa, A. baumannii [14,168,194,195,196]
aadA6/aadA10P. aeruginosa
gentamincin b, kanamycin, neomycin, paromomycin, lividomycin, ribostamycinaph(3′)-Ia/b/c/E. coli, K. pneumoniae, P. aeruginosa, A. baumannii [14,168,197]
kanamycin, neomycin, butirosin, paromomycin, ribostamycinaph(3′)-IIaE. asburiae, E. coli, A. baumannii, P. aeruginosa, K. pneumoniae [14,198]
kanamycin, neomycin, paromomycin, ribostamycin, butirosin, amikacin, isepamycinaph(3′)-VIa, aph(3′)-VIbA. baumannii, K. pneumoniae [14,36,199]
hygromycinaph(4)-IaE. coli, K. pneumoniae, A. baumannii [14]
streptomycinaph(6)-IdA. baumannii, K. pneumoniae, E. coli, P. aeruginosa [168,200,201]
streptomycinaph33ibA. baumannii [168]
Aminoglycosides: 16S
rRNA methylase genes
aminoglycosides (all)armAA. baumannii
P. aeruginosa, E. coli, K. pneumoniae, E. cloacae complex
[100,192,202]
rmtA,
rmtB, rmtC, rmtD, rmtE, rmtF, rmtG, rmtH, rmtF
P. aeruginosa
A. baumannii
P. aeruginosa, E. coli, K. pneumoniae, E. cloacae complex
[14,193,202,203,204,205,206,207,208]
Tetracycli
nes
doxycycline, tetracyclinetet(A), tet(G)A. baumannii
P. aeruginosa
E. coli
E. cloacae complex
K. pneumoniae
[14,41,42,43,45,209,210,211,212]
minocyclinetet(B),
tetracyclinetet(C), tet(D)
oxytetracycline, tetracyclinetet(H)
tetracycline, doxycycline, minocyclinetet(M), tet(O)
tetracyclines (all)tet(X)
Fluoroqui
nolones
ciprofloxacinqnrA
qnrB
qnrS
K. pneumoniae, E. coli, E. cloacae complex, A. baumannii, P. aeruginosa [14,213]
Polymyxinscolistinmcr1-10E. coli, K. pneumoniae, E. cloacae complex, P. aeruginosa, A. baumannii [214,215]

2.2.4. Polymyxin Resistance

Polymyxins (colistin and polymyxin B) are last-resort antibiotics reintroduced into clinical practice to treat infections caused by MDR Gram-negative bacteria resistant to carbapenems. Resistance to polymixins occurs due to mutations in chromosomal genes involved in lipopolysaccharide structure (mgrB, pmrABC, lpxACD), plasmid-mediated LPS modification (mcr genes), or active efflux [216,217,218,219,220,221,222,223].

2.2.5. Fosfomycin Resistance

Fosfomycin is a broad-spectrum antimicrobial agent that inhibits the final step of peptidoglycan biosynthesis. Resistance is attributed to the modification of transporters across the cytoplasmic membrane, amino acid substitution in the MurA active site, which decreases fosfomycin binding affinity, and the production of the fosfomycin-inactivating enzyme FosA [224].

2.2.6. Antifolate Resistance

Antifolate antibiotics inhibit purine metabolism and DNA and RNA synthesis by interfering with folic acid biosynthesis. Sulfonamides bind dihydropteroate synthase (DHPS), a catalytic enzyme in the folic acid biosynthesis pathway, inhibiting dihydrofolic acid formation [225]. Trimethoprim is a dihydrofolate reductase (DHFR) inhibitor. Sulfonamides combined with trimethoprim (such as sulfamethoxazole) are well-known folate inhibitors. Resistance to sulfonamides occurs due to mutations of the folP gene encoding DHPS or the acquisition of alternative DHPS genes (sul1, sul2, sul3, sul4) with low affinity to sulfonamides [226]. Trimethoprim resistance is mediated by dfr genes encoding trimethoprim-resistant dihydrofolate reductases.
Gram-negative pathogens have developed resistance mechanisms to all antibiotics used for therapy (Figure 1).

2.3. Mobile Genetic Elements (MGEs)

MGEs [insertion sequences (IS), transposons (Tn), integrons, resistance islands, and plasmids] play a significant role in the evolution of prokaryotic genomes, conferring adaptative traits, including AR. Thus, MGEs can retain, capture, and disseminate the ARGs between bacterial strains or species responsible for the emergence of MDR [227,228].

2.3.1. Insertion Sequences (IS)

IS are MGE that carry one or two transposase (tnp) genes and are responsible for the intracellular transportation of ARGs. ISs can be found across all prokaryotes, and their role in AR dissemination was well documented, not only by being able to mobilize ARGs but also by activating or inactivating specific genes in the bacterial chromosome, acting as promoters for silent ARGs or enhancing their expression, or, as in the case of certain enterobacteria, by inactivating certain porin-encoding genes or regulator genes [229,230]. In addition, two ISs can also form a composite transposon and sequester the genes between them [231].
In A. baumannii, the insertion of ISAba1/Aba2/Aba3/Aba4/Aba10 enhances the blaOXA-51, blaOXA-23, blaOXA-58 and carO genes expression, leading to carbapenem resistance, or with AmpC, leading to cephalosporin resistance [29,92,232,233,234,235,236,237,238,239]. In P. aeruginosa, the insertion of ISPa133, ISPa1328, ISPa46, ISPa1635, ISPa45, ISPa26, ISPa8, and ISPre2-like in the oprD porin-encoding gene confers carbapenem resistance [240,241,242,243,244,245,246]. In K. pneumoniae strains, the insertion of IS1,5,26,903 in the ompK36 porin-encoding gene confers cefoxitin resistance [247]. Colistin resistance emerges due to the insertion of ISAba11 in the lpxA or lpxC gene in A. baumannii [248] or IS1Flike, IS5like, ISKpn13, ISKpn14 in the mgrB gene in K. pneumoniae [249,250].
The notorious IS26 (or IS6) family is a well-known example of IS capable of sequestering and mobilizing ARGs. Members of the IS family are often found in arrays, in direct and/or inverted repeats, in MDR plasmids described in Gram-negative ESCAPE strains, and are able to capture virtually every ARG [251].

2.3.2. Transposons (Tn)

Transposons are MGE that can harbor additional genes, including ARGs, besides transposases. The Tn3 family is the essential Tn family involved in ARG transmission in Gram-negative and Gram-positive bacteria. Tn3 usually contains β-lactamase resistance genes in Gram-negative bacteria; Tn5 is associated with kanamycin, bleomycin, neomycin, and streptomycin; Tn7 with trimethoprim, streptothricin, spectinomycin, and streptomycin; Tn9 with chloramphenicol; and Tn10 with tetracycline resistance [252,253,254].

2.3.3. Integrons

MGE from the integrons group are located on bacterial chromosomes or plasmids. Four integron classes have been described in the nosocomial Gram-negative ESCAPE group, carrying ARGs responsible for β-lactams, aminoglycosides, and trimethoprim resistance [5]. In addition, different gene cassettes have been revealed: e.g., blaCARB-2, aadA1, aadA2, aadB, dfrA1, dfrA7, dfrA1-gcuF, dfrA1- aadA1, dfr17-aadA5, dfr12-gcuF-aadA2, sat1 responsible for cephalosporins, aminoglycosides, and trimethoprim resistance in A. baumannii [255,256,257]; aadA2, aadB, dfr17-aadA5, dfr12-gcuF- aadA2 associated with aminoglycosides and trimethoprim resistance in P. aeruginosa strains [258]; blaCARB-2, blaGES-1, aadA, aadA1, aadB, dfrA1, dfrA7, dfrA1-gcuF, dfrA1-aadA1a, dfr17-aadA5, dfr12-gcuF-aadA2 with cephalosporins, trimethoprim and aminoglycosides resistance in K. pneumoniae [257,259]; and aadA1, aadA2, aadA5 aadB, dfrA1, dfrA5, dfrA7 dfrA12, dfr14, dfrA17, dfrB2, dfrA1-gcuC, dfrA1-aadA1, dfr17-aadA5, dfr12-gcuF-aadA2, dfrA1-sat1-aadA1, dfrA1-sat2-aadA1, estX-sat2-aadA1, blaOXA-101-aac (6′)–Ib with aminoglycosides, trimethoprim and cephalosporins resistance in E. coli [260]. Different integron structures were also found in strains from various sources. Soufi et al. [261] revealed the presence of different gene cassette arrays in E. coli isolates from poultry meat in Tunisia (dfrA, aadA, sat-psp-aadA2-cmlA1- aadA1-qacH-IS440-sul3), while Su et al. have shown the presence of class 1 and 2 integrons in E. coli from one river in South China [262].

2.3.4. Genomic Resistance Islands

Genomic islands (GIs) represent clusters of genes of probable foreign origin, providing adaptative traits and representing a significant source of variation between bacterial strains. For example, GIs are responsible for forming different pathotypes of E. coli (uropathogenic—UPEC, enteropathogenic—EPEC, and enterohaemorrhagic—EHEC E. coli) [263]. GIs were associated with AR in A. baumannii epidemic clones, for which several A. baumannii Resistance Islands (AbaR) have been described [5], most of them in European clones I and II (AbaR1, AbaR3, AbaR5, AbaR6, AbaR7, AbaR8, AbaR9, and AbaR10) [264]. In K. pneumoniae, the GIE492 carries the blaSHV-190 gene, while in E. cloacae, the MIR17-GI carries blaMIR17 carbapenemase [265].

2.3.5. Plasmids

Plasmids are the main shuttles for ARG dissemination. Complex MGEs may disseminate intra-species, inter-species, inter-genus, and even across more distantly related taxa. Plasmid dissemination was excellently described in the scientific literature [266,267]. In A. baumannii, a plasmid-based replicon typing scheme, currently containing 20 groups encoded GR1-GR20, has been proposed [268,269]. Different resistance plasmids carrying carbapenem resistance genes were reported in P. aeruginosa strains (e.g., IncP-1, IncP-2, IncP-6) [270,271,272]. In Enterobacterales, 26 different compatibility plasmid groups (Inc plasmids) have been described based on their compatibility with other plasmids in the same bacterial host [227]. The plasmid relaxase gene typing (PRaseT) allowed the classification into five relaxase clades designated HIα, HIβ, HIγ, HIδ, and HIɛ of IncHI1 and IncHI2 plasmids, to which IncX1–4 and ColE plasmids were added [273].

3. Clinical Significance of Antibiotic-Resistant Gram-Negative Pathogens

The MDR and virulence potential of A. baumannii are responsible for hospital and community-acquired infections [274]. A. baumannii is recognized as an opportunistic nosocomial pathogen, mainly in immunocompromised patients, and is frequently associated with therapeutic failures, especially during the COVID-19 pandemic. Several countries have reported that COVID-19 was associated with secondary MDR carbapenem-resistant A. baumannii (MDR CRAB) infections of the lower respiratory tract in intensive care unit (ICU) patients, emphasizing the importance of limiting the risk of co-infection and the dissemination of MDR CRAB strains in ICUs [275,276,277].
Carbapenem-resistant P. aeruginosa (CR-PA) is a major healthcare-associated pathogen worldwide [278]. P. aeruginosa is the primary cause of ventilator-associated pneumonia (VAP) in long-term acute care hospitals and hospital wards and the second most common cause of VAP in intensive care units. It is also the third most common cause of catheter-related urinary tract infections [279]. In P. aeruginosa, several mechanisms are responsible for carbapenem resistance. The first mechanism is the efflux pump, which is mediated by overexpression of the MexAB-OprM efflux pump, resulting in resistance to most β-lactam drugs except for imipenem. The second mechanism is the overproduction of AmpC beta-lactamase and the inactivation of the OprD outer membrane protein. This combination can lead to resistance to essentially all antipseudomonal β-lactams. Another resistance mechanism is the production of carbapenemases [280,281], which significantly alter the efficacy of commonly used antipseudomonal agents, including ceftazidime, cefepime, and piperacillin-tazobactam, as well as the newly introduced β-lactam/β-lactamase inhibitor combinations such as ceftolozane-tazobactam, imipenem-relebactam, and ceftazidime-avibactam. The carbapenem resistance determinants carried by P. aeruginosa are often encoded on plasmids, such as IncP type; class I integrons, for example, those carrying the blaVIM gene; and other MGE, such as those associated with insertion sequence common region (ISCR) elements [114]. In addition, these isolates frequently carry additional resistance determinants to fluoroquinolones and aminoglycosides. Carbapenemase-producing P. aeruginosa (CP-PA) is often resistant to these therapeutic options, thus making treatment failure likely. CP-PA has also been associated with nosocomial spread, prompting infection prevention interventions [280].
The Enterobacterales order, as defined by Adelou et al. in 2016, comprises Gram-negative, non-spore-forming, rod-shaped, and facultative anaerobes bacteria. The order contains the families Enterobacteriaceae, Erwiniaceae, Pectobacteriaceae, Yersiniaceae, Hafniaceae, Morganellaceae, and Budviciaceae, some of which are members of the normal microbiota of the mammalian gastrointestinal tract [282]. The drastic rise in the incidence of MDR and extended drug-resistant (XDR) pathogens belonging to the Enterobacteriaceae group is a significant economic problem as these pathogens are prevalent natural residents of the human and animal microbiomes and spread quickly between humans. Moreover, Enterobacterales easily acquire ARGs via MGEs [283].
Most notable in Enterobacterales is the resistance to β-lactam antibiotics due to ESBL production, mainly in E. coli and K. pneumoniae, followed by aminoglycoside and fluoroquinolone resistance (Table 1 and Table 2). These resistance phenotypes are often coupled, leading to MDR and the necessity to use last-resort antibiotics [284].
K. pneumoniae is the causative agent of about one-third of all Gram-negative infections (urinary tract infections, cystitis, pneumonia, surgical wound infections, endocarditis, septicemia, necrotizing pneumonia, pyogenic liver abscesses, and endogenous endophthalmitis), associated with high mortality rates and extended hospitalization, coupled with high economic costs. Due to selective pressure caused by antibiotic usage, K. pneumoniae collects ARGs, which led to the development of XDR strains harboring a ‘super resistome’. These include the emergence of hypervirulent K. pneumoniae (hvKp) or hypermucoviscous K. pneumoniae (HMKP), usually susceptible to last-line antibiotics (carbapenems and colistin) [285,286]. The continuous global dissemination of high-risk MDR and XDR K. pneumoniae highlights their complex evolution, involving the transfer and spread of ARGs and epidemic plasmids [287,288]. Most of the carbapenemase and/or ESBL-producing K. pneumoniae strains, as well as those harboring aminoglycoside resistance, belong to specific clones CC (clonal complex) 258, CC15, and CC14 [289], while colistin-resistant clones mainly belong to CC11, 43, and 258 [290,291,292,293].
Enterobacter spp. are increasingly described as contributing to the dissemination of infections caused by carbapenem resistant strains. Amongst the 22 species of this genus, Enterobacter aerogenes, E. cloacae, and E. hormaechei are the most frequently isolated species in clinical infections, mainly in immunocompromised patients and those hospitalized in ICU, due to their adaptation to the hospital environment and their ability to efficiently acquire numerous genetic mobile elements containing resistance and virulence genes [294].
Enterobacter cloacae complex (E. cloacae, E. asburiae, E. hormaechei, E. kobei, E. ludwigii, E. mori, and E. nimipressuralis) are common nosocomial pathogens involved in a wide variety of infections (pneumonia, UTI, and septicemia). The emergence of MDR clones, including resistance to the last-resort carbapenems, increased interest in these pathogens [295].
E. coli is particularly interesting since it represents a significant part of the normal microbiota, but it can also cause severe infections in humans and animals. In humans, E. coli can cause infections in practically every anatomical site of the human body, involving urinary tract infections, appendicitis, pneumonia, the bloodstream, gastrointestinal infections, skin abscesses, intra-amniotic and puerperal infections in pregnant women, meningitis, and endocarditis. Moreover, E. coli is involved in community-acquired and healthcare-related infections and can cause disease in all age groups [296,297].
E. coli is the second bacteria (after Klebsiella) involved in human infections associated with MDR bacterial infections. Furthermore, the significant increase in the emergence and dissemination of E. coli to the main antibiotic classes (β-lactams, quinolones, aminoglycosides, sulfonamides, and fosfomycin), including the last-resort carbapenems and polymyxins, is correlated with prolonged hospital stays and patient deaths [297,298].

Other Enterobacterales

Citrobacter spp., mainly C. freundii, are inhabitants of the intestinal tract and have been associated with nosocomial infections involving the urinary tract, liver, biliary tract, peritoneum, intestines, bone, respiratory tract, endocardium, wounds, soft tissue, meninges, and the bloodstream. The emergence of MDR Citrobacter strains is an increasing concern due to the production of AmpC, broad-spectrum β-lactamase, ESBL, or even carbapenemase, particularly MBL or KPC types. In addition, quinolone resistance (qnr and aac(6′)-Ib-cr genes), numerous qnrB alleles, and about 40 qnrB variants (located on the chromosome of Citrobacter spp., especially C. freundii) were described [299].
Hafnia alvei is rarely isolated from human samples. When it does, it is responsible for nosocomial infections, including gastroenteritis, urinary tract infections, meningitis, pneumonia, wound infections, soft tissue infections, endophthalmitis, and septicemia. The organism resides in the gastrointestinal tract of humans and many animals. Most infections with H. alvei are identified in patients with severe underlying diseases (e.g., malignancies) or after surgery or trauma. Besides its natural resistance to colistin and expression of AmpC chromosomal β-lactamase, it was described as the emergence of a Hafnia paralvei resistant to carbapenems due to a defective porin [300].
Morganella morganii is ubiquitous and is often associated with stool specimens collected from patients with symptoms of diarrhea. They are normal inhabitants of the gastrointestinal tract. M. morganii has intrinsic resistance to oxacillin, ampicillin, amoxicillin, and most first- and second-generation cephalosporins, macrolides, lincosamides, glycopeptides, fosfomycin, fusidic acid, and colistin. AR in M. morganii has been raised in recent years, mainly due to MGEs, leading to MDR and XDR strains [301].
Providencia spp. are usually isolated from patients with urinary tract infections and diarrhea and are associated with nosocomial outbreaks. Most commonly, P. rettgeri and P. stuartii represent the majority of MDR strains isolated and are intrinsically resistant to penicillins and the first-generation cephalosporins, aminoglycosides, tetracyclines (including tigecycline), and colistin [302,303].
Serratia spp., most commonly S. marcescens, is involved in nosocomial outbreaks and the colonization of diverse healthcare settings. S. marcescens has been associated with meningitis, sepsis, UTIs, skin infections, bloodstream infections, and respiratory infections. The intrinsic resistance to ampicillin, first- and second-generation cephalosporins, macrolides, and antimicrobial peptides, including colistin, is very challenging for clinical management. Moreover, some strains express the SME-1 enzyme, conferring resistance to imipenem, aztreonam, cephalosporins, and penicillins [304,305].
Salmonellae are Gram-negative bacteria that are pathogenic to humans and are traditionally subdivided into two major groups based on their clinical presentation: typhoidal Salmonella and non-typhoidal Salmonella. Typhoidal Salmonella, comprising the S. enterica subspecies enterica (hereafter Salmonella) serovars Typhi and Paratyphi A, B, and C, cause a systemic disease also known as enteric fever [306]. Human-restricted S. Typhi is the dominant cause of typhoid fever, with an estimated number of cases between 21.7 million and 26.9 million per year [307] and an estimated 217,000 deaths per year [308]. S. enterica constitutes a significant public health concern, and it is estimated to cause more than 300,000 annual deaths, mostly in developing countries [309]. This species is classified into hundreds of serovars based on surface antigenic composition. Some serovars (e.g., S. Typhi and S. Paratyphi) are host-adapted to humans, where they cause a systemic infection known as typhoid or paratyphoid fever and are therefore referred to as “typhoidal” serovars. Other serovars, such as S. Typhimurium, have a broad host range and, in humans, most often cause self-limiting gastroenteritis and are referred to as “non-typhoidal” serovars [310].
In Salmonella spp., particularly S. Typhi, antimicrobial resistance could be mediated by plasmid or chromosomal DNA. Usually, resistance is developed by the inactivation of antibacterial agents, alteration of drug targets, and employing various efflux pumps. In addition, external resistance factors may be actively mediated by gene transfer using virulence plasmids, phages, and MGEs [311]. S. Typhi typically has plasmids that contain several virulence factors and ARGs. These plasmids vary in size (50–90 kb) and carry the spv operon, which is significantly involved in causing infection. The genes of this operon are reportedly pivotal for bacterial proliferation in host cells and supposedly enhance the virulence of the pathogen [312]. Considering that most virulence plasmids are not self-transferable, some contain transgenes that enable the transfer of plasmids via conjugation. Incompatible (Inc) plasmids encode multiple antimicrobial resistance genes in S. Typhi and are classified into IncH1, IncH2, and IncH3. In addition, plasmids R27, pHCM1, and pAKU1 comprise a composite transposon that can harbor multidrug resistance in MDR S. Typhi strains [313].
Regarding the production of β-lactamases, TEM, SHV, and CTX-M are the main types of ESBLs in Salmonella spp., conferring resistance to penicillin and cephalosporin [314]. In S. Typhi, the presence of these genes has been attributed to the genetic transfer of resistance genes from other Gram-negative bacterial species [315]. In Salmonella spp., there were described genes encoding resistance to tetracycline (tetA, tetB, tetG), quinolones (qnrA, qnrB, qnrC, qnrS), and chloramphenicol (cat1 and cat2). Genetic elements identifying the mobile gene cassettes that carry multidrug-resistant genes are known as integrons. In S. Typhi, the presence of integrons (classes 1 and 2) equalizes the distribution of antimicrobial resistance, in which class 1 is more dominant [316].

4. Antibiotic Resistance in Romania

4.1. Antibiotic Resistance in Romanian Hospital Settings

Romania is one of the European countries with the highest rates of MDR in A. baumannii clinical isolates (in 2020, the highest resistance percentages were recorded for fluoroquinolones, carbapenems, and aminoglycosides) [317]. Concerning P. aeruginosa, the highest resistance levels in 2020 were recorded for fluoroquinolones, carbapenems, and ceftazidime [318]. To the best of our knowledge, available data on molecular characterization of the non-fermenting strains is sourced from the major geographical areas of the country, including the capital city (Table 3).
Clinical A. baumannii strains from different Romanian regions exhibited CHLD-producing blaOXA-23 (West, North, Central, and South regions), followed by blaOXA-24/72 gene (North, Central, and South regions), and blaOXA-58 gene revealed by one study from the capital city, and respectively MBLs encoding blaVIM-2 and blaIMP-1 genes (South and North). The distribution of CHLDs and MBLs by isolation sources and period highlighted that carbapenem-encoding genes were not correlated over the period with the specific isolation sources, being described in strains isolated from infection sites or anal carriage. Studies from the West, Central, and South regions reported ESBL-encoding genes (blaPER-1; blaTEM-1; blaTEM-12; blaTEM-84 and blaSHV-12) from sterile or non-sterile isolation sources (Table 3). However, a limited number of studies have focused on the MGE carrying the CHLDs and revealed the presence of class 1 integrons, insertion sequences, transposons, or different plasmid types in A. baumannii recovered from South and Western Romanian intrahospital infections. A. baumannii reported from western and southern Romania belonged to high-risk international clones like ST2, ST1, ST636, and ST492.
Between 2008 and 2015, the most common carbapenemase-encoding genes in P. aeruginosa clinical strains isolated from intra-hospital infections or carriages in North, Central, and South Romania were blaVIM-2, blaVIM-4 and blaIMP-13. In addition, ESBL-encoding genes (blaSHVlike; blaGESlike; blaVEBlike; blaTEMlike) were encountered in P. aeruginosa strains isolated in the southern region of Romania.
Clinical enterobacterial isolates harbored mainly the ESBL blaCTX-M(-15) gene, followed by blaSHV and blaTEM (for which the ESBL phenotype depends on the gene variant). blaOXA-48 was the most commonly reported carbapenemase, followed by blaNDM and blaKPC (Table 4). Aminoglycoside resistance was mainly associated with AMEs, and all variants of PMQR have been described (Table 5).
Several studies investigated the mobile genetic platforms carrying the respective ARGs, particularly in the southern part of the country, and highlighted the presence of plasmids and integrons previously associated with AR without being associated with specific isolation sources. In addition, strains isolated from different infection sites of inpatients from hospitals in the south of Romania belonged to widespread international E. coli (e.g., ST131, ST10, and ST5) or K. pneumoniae (ST101) clones.
In 2020, the reported antibiotic resistance levels in Enterobacterales indicated that K. pneumoniae had the highest resistance levels in third generation cephalosporins, fluoroquinolones, aminoglycosides, and E. coli for aminopenicillins, fluoroquinolones, and third generation cephalosporins [317].
Several encoding genes and MGE (class 1 integrons and Inc plasmids) were reported in different parts of Romania (Table 4 and Table 5).

4.2. Community-Acquired Antibiotic Resistance

Generally, infections are classified into two categories: community-acquired and nosocomial (intra-hospital) infections. Healthcare-associated infections are specific to admitted hospital patients and occur after at least 48 h from admission, while community-acquired infections are contracted outside of a healthcare facility and diagnosed within 48 h after admission (community onset) [346,358].
The occurrence of AR in community-acquired infections is increasing due to multiple factors. Antibiotic overuse, for example, can imbalance the composition of the gut microbiota, facilitating the emergence and colonization of the gut with antibiotic-resistant bacteria (ARB) and the proliferation of opportunistic pathogens [359]. On the other hand, antibiotic residues in the environment or in food products could select for resistance [360,361].
Community-acquired AR was scarcely investigated in Romania (Table 6). For E. coli, we have identified five studies reporting carbapenemases, ESBLs, and aminoglycosides, PMQR, trimethoprim, and tetracycline resistance genes in strains isolated from community-acquired UTIs. In A. baumannii, the presence of CHLDs, AMEs, and sulphonamide, tetracycline, and macrolide resistance genes has beeen reported, while in P. aeruginosa, the presence of MBL, AMEs, and PMQR, sulphonamide, trimethoprim, and tetracycline resistance genes has been reported in strains isolated from all geographical regions of the country. Furthermore, class 1 and 2 integrons were involved in disseminating ARGs in these strains.

4.3. Antibiotic Resistance in Veterinary Settings

Several factors can affect the occurrence and dissemination of AR in the animal industry, including antibiotic use and farm management. Many studies have focused on how the use of antibiotics in food-producing animals has led to the expansion of antibiotic resistance. In industrialized countries, the companion animal population has dramatically increased during the last few decades. The increased interaction between animals and humans leads to a higher risk of infections and the cross-transmission of AR traits. Thus, the potential of reverse zoonosis and the creation of animal reservoirs that keep the loop of infection and AR diffusion open are gaining steadily increasing concern. Antimicrobial resistance of pet origin, responsible for both direct and/or indirect threats to human health, involves mainly carbapenemase-producing enterobacteria and ESBL Gram-negative bacteria [366].
The epidemiological scale of AR transmission between humans and animals is not yet well defined, as multiple parameters should be taken into account (population features, geographical location, investigative methods), and the sole abuse or misuse of antibiotics is insufficient for such a massive transmission of resistant microorganisms between humans and pets [367,368]. Therefore, several research lines are being explored, such as human-animal transmission and vice versa, although controversial results are being observed [369]. Moreover, the environment most likely contributes to AR dissemination, intended as the vector connecting the human and animal environments, including anthropic activities. It was also suggested that monitoring non-pathogenic specimens and their potential capability to acquire resistance traits is a promising strategy to predict and prevent future resistant strains.
In Romania, very scarce information is available regarding the isolation, identification, and AR of Gram-negative bacilli from veterinary settings; the ARGs and carrying platforms were investigated only for Enterobacterales species, reported in three studies (Table 7). The other studies reveal only the presence of resistant bacterial strains in different animal isolates without investigating the genetic background of AR.
Thus, gentamycin- and penicillin-resistant Pseudomonadaceae strains were described in samples from boar semen from three artificial insemination centers in the northwest of Romania [367]. Cristina et al. investigated the presence of AR in isolates from pet reptiles (chelonians, snakes, and lizards) and identified the presence of P. aeruginosa, Citrobacter koseri, C. brakii, and K. oxytoca resistant to cephalosporins (up to the fourth generation), tetracyclines, quinolones, aminoglycosides, and others [370]. Tîrziu et al. investigated the prevalence and AR profiles of two major foodborne pathogens (Salmonella spp. and, respectively, Campylobacter spp.) in different food products from two Transylvanian counties of Romania and revealed high levels of resistance to tetracycline, ciprofloxacin, and nalidixic acid in both pathogens [371]. A high level of AR in Campylobacter spp. was also reported in strains isolated from broiler chicken flocks from three north-western Transylvanian counties of Romania [372].
Table 7. Genetic background of AR and carrying platforms in Gram-negative strains isolated in Romanian veterinary settings.
Table 7. Genetic background of AR and carrying platforms in Gram-negative strains isolated in Romanian veterinary settings.
SpeciesLocation/YearIsolation Sourceβ-Lactam Resistance GenesAminoglycoside Resistance GenesQuinolone Resistance GenesOther ARGsMGEsReference
E. coli, K. pneumoniaeNorth-East and South-East Romania
2017–2018
dog faecal samplesblaCTX-M-1, -3, -9, -14, -15;
blaTEM-1;
blasHV-2, -52;
aph(3′)-Ia; aph(6)-Id; ant(3″)-Ib;
aac(3)-IId;
aac(2′)-IIa;
qnrS1mph;
sul1, -2;
tet(A), -(B)
Inc plasmids:
HI2; Y; P1; FIB; F; N;
F2A; L/M
FIB/K; I1
FIB/Y/FIA;
[355]
E. coliTimiș and Arad counties
2019–2020
swine intestinal microbiota--qnrB, -S-- [373]
E. coli
Salmonella spp.
Cluj-Napoca
2012–2013
chicken carcassesblaTEMaadA1-tet(A), sul1, dfrIa- [374]
E. coli Romania, 1980–2016rectal swabbing of calves, foals, and piglet
blaTEM--tet(A), tet(B), tet(C), sul1, dfrA1int1, int2 [375]

4.4. Antibiotic Resistance in Gram-Negative ESCAPE Pathogens in Wastewaters

Antibiotics are among the most popular pharmaceuticals used in human medicine, veterinary care, and farming [376,377,378]. They are also frequent contaminants in wastewater, municipal sewage, and wastewater treatment plants’ influents and effluents [379]. Hospitals generate an impressive amount of wastewater per day; the hospital effluents are loaded with pathogenic microorganisms, antibiotics, and other pharmaceutical or toxic substances, which are only partially removed during wastewater treatments, contributing to the pollution of the natural environment, including the selection and dissemination of AR [380,381]. Wastewater treatment plants (WWTPs) are one of the critical reservoirs of both antibiotic-resistant bacteria and ARGs and represent hotspots for horizontal gene transfer (HGT) via MGEs, such as plasmid integrons, transposons, resistance islands, and insertion sequences, enabling the development and dissemination of ARGs between bacteria [382,383]. Antibiotic-resistant bacteria are collected and mixed with environmental strains, which, in turn, could introduce the newly acquired ARGs into the clinics [384]. Romania had the third-highest consumption of antibacterials for systemic use in the community sector in 2019 [385]. In 2016 and 2018, the most-consumed classes for food-producing animals in Romania, according to the European Surveillance of Veterinary Antimicrobial Consumption (EVSAC), were tetracyclines and penicillins, respectively [386]. This explains the presence of antibiotics in WWTP, representing a high selective pressure for AR [387]. In this context, during the last few years, international authorities have made considerable efforts to improve the monitoring of the circulation of the antibiotic-resistant bacteria in different environments, underscoring the necessity to strengthen intersectional human, animal, and agricultural cooperation, which has been included as a priority in the work plan for the EU Health Programme. One of the priority topics of the Joint Programme Initiative on Antimicrobial Resistance (JPIAMR) is the elucidation of the role of the environment as a source for the selection and dissemination of AR, which is expected to provide essential data for monitoring AR, as the lack of surveillance is considered one of the main contributors to the spread of AR, particularly in developing countries. In this regard, one important goal is mapping the distribution of MDR strains and plasmids and different genomic lineages of critical nosocomial pathogens in different clinical and aquatic compartments. This vital knowledge could be translated into policy measures to control the emergence and spread of antibiotic-resistant bacteria [388,389].
Contrary to clinical studies, there needs to be more information regarding the ARG reservoirs in the wastewater network in Romania. Our research team showed a high repertoire of ARGs and virulence markers in K. pneumoniae ST101 isolated from intra-hospital infections and wastewater samples collected from the influent and, respectively, from the effluent of hospital collecting sewage tanks in the southern regions of Romania; the transmission of MDR, carbapenemase and ESBL-producing K. pneumoniae ST101 from hospital to hospital effluent; and its persistence after the chlorine treatment [339,344]. In the country’s central region, a chlorinated wastewater treatment system from a public hospital revealed the presence of the following carbapenemase and ESBL-encoding genes in the influent: blaPER, blaVIM, blaNDM-1, and blaSHV. In contrast, the chlorinated effluent exhibited blaVIM and blaSHV [390]. Another study performed in Cluj-Napoca on one WWTP and the receiver river Someșul Mic revealed the presence of tetracycline and sulphonamide ARGs [sul1, tet(O), and tet(W)] in wastewater without focusing on total antibiotic-resistant bacteria or ARGs identification [391]. Butiuc-Keul et al., in 2019, revealed the genetic background of AR in Pseudomonas spp. from urban water sources and their environmental impact in north-western Romania [392]. Several carbapenemase-encoding genes have recently shown spatiotemporal variation in wastewater samples from the influent and effluent of three Cluj-Napoca WWTPs [393]. In South Romania, Van et al. recently revealed the efficiency of commercial essential oils against antibiotic-resistant P. aeruginosa clinical and wastewater strains [394]. Gheorghe-Barbu et al. have demonstrated a high repertoire of ARGs in A. baumannii and P. aeruginosa strains isolated two years consecutively from intra-hospital infections, wastewater, and surface water from three geographical regions of Romania and highlighted the importance of screening for acquired antimicrobial resistance in the environment [330] (Table 8).

4.5. Antibiotic Resistance in Other Aquatic Ecosystems

Surface water plays an essential role in AR dissemination by being both a habitat and a dissemination ecosystem for microorganisms. Recently, Banciu et al. demonstrated the dissemination of A. baumannii and P. aeruginosa clinical strains in wastewater or surface water or the presence of E. coli, K. oxytoca, C. freundii, and P. mirabilis resistant to ampicillin and clavulanic acid, strains isolated from the St. Gheorghe branch of the Danube Delta [386]. The Danube River is considered the most critical non-oceanic body of water in Europe and the “future central axis for the European Union,” Its Danube Delta is included in the Biosphere Reserve and Ramsar Site lists. The Danube River crosses ten countries. This basin represents an optimal pool for resistant pathogens and anthropogenic pollutants dissemination and accumulation throughout large and distant areas, being assigned as a reservoir of AR. Previously, it has been demonstrated that Bucharest was at the top of the most polluted sampling locations from twelve WWTPs in nine countries (Romania, Serbia, Hungary, Slovenia, Croatia, Slovakia, Czechia, Austria, and Germany) in the Danube River Basin collected and analyzed for ARGs and MGE presence [395].
Several other authors characterized at the molecular level the Gram-negative rods from the surface water (Dambovița river—south Romania), from the 4 Romanian natural aquatic fishery lowland salted lakes from the Natura 2000 Network located in Buzǎu and Brǎila counties, carrying a high diversity of resistance markers correlated with class one integrons [396] (see Table 8).
Table 8. Genetic background of antibiotic resistance and carrying platforms in Gram-negative strains isolated in Romania from wastewater and surface waters.
Table 8. Genetic background of antibiotic resistance and carrying platforms in Gram-negative strains isolated in Romania from wastewater and surface waters.
SpeciesLocation/PeriodIsolation Sourceβ-lactam Antibiotics GenesAminoglycoside Resistance GenesQuinolone Resistance GenesOther ARGsMGEsReference
A. baumanniiBucharest, Târgoviște, Vâlcea, Iași, Galați, Timișoara, Cluj; 2018–2019hospital sewage tanks, WWTPs—influent, active sludge, effluent; surface water—upstream and downstream region (200 m) of the rivers: Dambovita, Ialomita, Olt, Bahlui, Siret, Bega, SomesblaTEM-1, -12
blaOXA-23;
blaOXA-72;
blaOXA-65, -66, -126, -217;
blaADC-5, -25, -73, -81, -154, -167;
aph(6)-Id;
aph (3′)-VIa;
ant(3″)-Ib, -IIa;
aac(3)I, -Ip;
aadA1, -A2;
armA;
ant(2″)-Ia;
-catA1; catB8; cmlB1; sul1, -2; tet(A), -(B); mph(E); msrE; drfA12;
qacE1 integron associated gene
[330]
P. aeruginosaBucharest; 2018–2019;
Târgoviște, Vâlcea, Iași, Galați, Timișoara, Cluj; 2018–2019
-blaGES-4; blaVEB-9
blaTEM-40; blaIMP-13; blaVIM-2;
aac(6′)-II; aadA1, -A2; aph(3′)-Ia, -IIb;
aac(6)-Id, -II;
ant(2″)-Ia
-fosA;
catB7, bcr1;
tet(A), -(G); mphE; msrE;
qacE1 integron associated gene [330]
P. aeruginosaBucharest
2018–2020
intra-hospital infections, hospital sewage tank, WWTPsblaCTX-M; blaSHV; blaGES; blaVEB; blaVIM---- [393]
Pseudomonas spp.Cluj-Napoca
2015
hospital effluent, municipal WWTP, Somesul Mic river—upstream and downstreamblaTEM-1; blaSHV-1; blaPER-1; blaVIM-1;
blaPstS
aac(6′)-II; aac(3)-IIIaqnrA, -B
ermB;
tet(A), -(B), -(C)
intI [391]
not providedCluj county, 2019–2020influent, effluent samples from 3 WWTPs blaKPC; blaIMP; blaNDM; blaOXA-48---- [390]
S. marcescens,
K. pneumoniae, K. oxytoca, E. coli, E. cloacae complex, A. calcoaceticus, R. orniyhinolytica, E. hermanii, E. cowanii, S. rubidaea, P. ananatis, H. alvei
Buzău, Brăila counties
2016
Four lowland salted lakes included in Natura 2000 networkblaCTX-M (S. marcescens, K. oxytoca, A. calcoaceticus, E. cloacae, E. coli, H. alvei)
blaNDM (E. coli)
blaIMP (E. kobei)
aac-(3)Ia (K. pneumoniae)qnrS (E. cloacae)sul1intI [397]
K. pneumoniaeBucharest
2018–2019
hospital collecting sewage tank (influent, effluent)blaCTX–M–15;
blaOXA–1;
blaOXA–48;
blaSHV-1, -12, -106, –107, -145, -158, -187;
blaKPC-2;
blaTEM–1, -150; (K. pneumoniae)
aac(3)II-a; aph(6)-Id; ant(3″)-Ib; aadA2 (K. pneumoniae)qnrS1; qnrB1; qnrB4tet(A), -(D)
catA1, -A2; sul1, -2;
arr2, -3;
dfrA1, -A12, -A14;
msrE; mphE; cmlA5 (K. pneumoniae);
qacE∆1 [344]
K. pneumoniaeBucharest, Galați and Târgoviște
2018–2019
wastewater—influent, effluent from WWTPsblaCMY-4; blaCTX-M-15
blaDHA-1;
blaKPC-2;
blaNDM-1
blaOXA-1, -9, -10;
blaOXA-48, -162;
blaSHV-1, -11, -12; -100, -107, -145, -158, -161; blaTEM-1, -150;
aac(3)-IIa, -IId
aac(6′)-Ib, -Ibcr, -IId, -Il;
aadA1, -A2, -A5
ant(2″)Ia
aph(3′)-Ia, -Ib;
aph(6)Id;
qnrB1, -B4-B10, -B19, -B36, -B67
qnrD1
qnrS1
tet(A), -(D)
catA1, -A2;
catB3; cmlA5; fosA, -A6;
mphA;mphE; msrE;
arr2, -3;
dfrA1, -A12, -A14;
sul1, -2
armA;ble;
qacEdelta1 [339]
Gram-negative rodsBucharest, 2011–2012hospital sewage; influent, effluent of WWTPs; surface water—Dâmbovița river—upstream, downstream after the WWTP dischargeblaTEM, blaSHV, blaCMY, blaCTX-M,
blaNDM, blaVIM
-qnrB; qnrStet(B), -(M)
sulII; dfrA1-aadA1;
- [398]
not providedCluj
2017
influent, effluent of WWTPs; surface water—Someșul Mic river- upstream, downstream after the WWTP discharge---sul1; tet(O), -(W);intI1 [391]
not providedCluj county
2015
hospital sewage tank (influent, effluent)blaPER; blaVIM; blaNDM-1; blaSHV;---- [390]
Acinetobacter spp., Enterobacteriaceae, Pseudomonas spp.Cluj county
2015
hospital sewage tanksblaVIM; blaSHV;aacC2-sul 1, -2; qacE, tet(A), -(B), -(C), -(W); catA1; floR; - [399]
Enterobacter spp., E. coli, K. pneumoniaeDanube river
2013
surface waterblaCTX-M-15, -3, -9, -27, -55; blaSHV-1,-2,-11,-12; blaKPC-2; blaNDM-1,---- [400]
not providedBucharest
2017
wastewater (WWTP)blaOXA; blaSHVaph(III)aqnrStet(B); tet(M); sul1; ermB; ermF; vanAintI1 [396]
not providedCluj
2017
wastewater (WWTP)blaOXA; blaSHVaph(III)aqnrStet(M); sul1; ermB; ermFintI1 [396]

5. Conclusions

Updated information regarding the genetic background and molecular epidemiology of AR is crucial for tackling the spread of this phenomenon. This review brings together the available data regarding the AR of Gram-negative ESCAPE pathogens circulating in Romania. The big picture for Gram-negative ESCAPE pathogens reveals that all significant, clinically relevant, globally spread ARGs and carrying platforms are well established in different areas of our country and are already disseminated beyond clinical settings.
To constrain the spread of ESCAPE pathogens, it is now well recognized that collaborative efforts are required by policymakers, funders, and those responsible for the treatment and management of ESCAPE pathogens. Aside from novel drug development, these collaborative endeavors will require sustainable stewardship practices to reduce the inappropriate use of antibiotics in both the human health and agricultural sectors. In addition, improvements in factors encompassing AR surveillance, diagnostics, patient education, and patient treatment options will help facilitate AR control.

Author Contributions

M.C.C. conceived, revised, and corrected the manuscript; I.G.-B. and I.C.B. drafted chapters 1, 2, and 3; G.A.G. and C.O.V. drafted the tables. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Romanian Executive Agency for Higher Education, Research, Development, and Innovation (https://uefiscdi.gov.ro/) research projects PN-III-P1-1.1-TE-2021-1515 (TE 112/2022) and PN-III-P4-PCE-2021-1797 (PCE 96/2022) and by the Ministry of Research, Innovation, and Digitalization through Program 1—Development of the national R&D system, Subprogram 1.2—Institutional performance—Financing projects for excellence in RDI research project C1.2.PFE-CDI.2021-587, contract no. 41PFE/30.12.2021. The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper. The funding had no role in the study design, data collection, analysis, the decision to publish, or the preparation of the manuscript.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. World Health Organization. Available online: https://www.who.int/news/item/27-02-2017-who-publishes-list-of-bacteria-for-which-new-antibiotics-are-urgently-needed (accessed on 22 March 2023).
  2. Boucher, H.W.; Talbot, G.H.; Bradley, J.S.; Edwards, J.E.; Gilbert, D.; Rice, L.B.; Scheld, M.; Spellberg, B.; Bartlett, J. Bad bugs, no drugs: No ESKAPE! An update from the Infectious Diseases Society of America. Clin. Infect Dis. 2009, 48, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Mehrad, B.; Clark, N.M.; Zhanel, G.G.; Lynch, J.P. Antimicrobial resistance in hospital-acquired gram-negative bacterial infections. Chest 2015, 147, 1413–1421. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Centers for Disease Control and Prevention. National Vaccine Program Office Workshop on Aluminum in Vaccines. JAMA 2000, 283, 1955. [Google Scholar] [CrossRef] [Green Version]
  5. Breijyeh, Z.; Jubeh, B.; Karaman, R. Resistance of Gram-Negative Bacteria to Current Antibacterial Agents and Approaches to Resolve It. Molecules 2020, 25, 1340. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Zhu, Y.; Huang, W.E.; Yang, Q. Clinical Perspective of Antimicrobial Resistance in Bacteria. Infect Drug Resist. 2022, 15, 735–746. [Google Scholar] [CrossRef] [PubMed]
  7. Halat, D.H.; Moubareck, C.A. The Current Burden of Carbapenemases: Review of Significant Properties and Dissemination among Gram-Negative Bacteria. Antibiotics 2020, 9, 186. [Google Scholar] [CrossRef]
  8. Santajit, S.; Indrawattana, N. Mechanisms of antimicrobial resistance in ESKAPE pathogens. Biomed Res. Int. 2016, 2016, 2475067. [Google Scholar] [CrossRef] [Green Version]
  9. Pachori, P.; Gothalwal, R.; Gandhi, P. Emergence of antibiotic resistance Pseudomonas aeruginosa in intensive care unit; a critical review. Gene Funct. Dis. 2019, 6, 109–119. [Google Scholar] [CrossRef]
  10. Knott-Hunziker, V.; Waley, S.G.; Orlek, B.S.; Sammes, P.G. Penicillinase active sites: Labelling of serine-44 in beta-lactamase I by 6beta-bromopenicillanic acid. FEBS Lett. 1979, 99, 59–61. [Google Scholar] [CrossRef] [Green Version]
  11. Zhang, H.; Hao, Q. Crystal structure of NDM-1 reveals a common β-lactam hydrolysis mechanism. FASEB J. 2011, 25, 2574–2582. [Google Scholar] [CrossRef]
  12. Ambler, R.P. The structure of beta-lactamases. Philos. Trans. R. Soc. Lond. Ser. B Biol. Sci. 1980, 289, 321–331. [Google Scholar] [CrossRef]
  13. Jeon, J.H.; Lee, J.H.; Lee, J.J.; Park, K.S.; Karim, A.M.; Lee, C.R.; Jeong, B.C.; Lee, S.H. Structural basis for carbapenem-hydrolyzing mechanisms of carbapenemases conferring antibiotic resistance. Int. J. Mol. Sci. 2015, 16, 9654–9692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Kyriakidis, I.; Vasileiou, E.; Pana, Z.D.; Tragiannidis, A. Acinetobacter baumannii Antibiotic Resistance Mechanisms. Pathogens 2021, 10, 373. [Google Scholar] [CrossRef] [PubMed]
  15. Zhanel, G.G.; Lawson, C.D.; Adam, H.; Schweizer, F.; Zelenitsky, S.; Lagacé-Wiens, P.R.; Denisuik, A.; Rubinstein, E.; Gin, A.S.; Hoban, D.J.; et al. Ceftazidime-avibactam: A novel cephalosporin/β-lactamase inhibitor combination. Drugs 2013, 73, 159–177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Poirel, L.; Naas, T.; Nordmann, P. Diversity, epidemiology, and genetics of class D beta-lactamases. Antimicrob. Agents Chemother. 2010, 54, 24–38. [Google Scholar] [CrossRef] [Green Version]
  17. Choi, U.; Lee, C.-R. Distinct Roles of Outer Membrane Porins in Antibiotic Resistance and Membrane Integrity in Escherichia coli. Front. Microbiol. 2019, 10, 953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Davin-Regli, A.; Pages, J.-M.; Ferrand, A. Clinical Status of Efflux Resistance Mechanisms in Gram-Negative Bacteria. Antibiotics 2021, 10, 1117. [Google Scholar] [CrossRef]
  19. Poirel, L.; Bonnin, R.A.; Nordmann, P. Genetic basis of antibiotic resistance in pathogenic Acinetobacter species. IUBMB Life 2011, 63, 1061–1067. [Google Scholar] [CrossRef]
  20. Mojica, M.F.; Bonomo, R.A.; Fast, W. B1-metallo-β-lactamases: Where do we stand? Curr. Drug Targets 2016, 17, 1029–1050. [Google Scholar] [CrossRef]
  21. Cornaglia, G.; Giamarellou, H.; Rossolini, G.M. Metallo-β-lactamases: A last frontier for β-lactams? Lancet Infect. Dis. 2011, 11, 381–393. [Google Scholar] [CrossRef]
  22. Donald, H.M.; Scaife, W.; Amyes, S.G.; Young, H.K. Sequence analysis of ARI-1, a novel OXA beta-lactamase, responsible for imipenem resistance in Acinetobacter baumannii 6B92. Antimicrob. Agents Chemother. 2000, 44, 196–199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Nigro, S.J.; Hall, R.M. Structure and context of Acinetobacter transposons carrying the oxa23 carbapenemase gene. J. Antimicrob. Chemother. 2016, 71, 1135–1147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Bush, K.; Jacoby, G.A. Updated functional classification of beta-lactamases. Antimicrob. Agents Chemother. 2010, 54, 969–976. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Sawa, T.; Kooguchi, K.; Moriyama, K. Molecular diversity of extended-spectrum β-lactamases and carbapenemases, and antimicrobial resistance. J. Intensive Care 2020, 8, 13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Hocquet, D.; Plesiat, P.; Dehecq, B.; Mariotte, P.; Talon, D.; Bertrand, X. Nationwide investigation of extended-spectrum β-lactamases, metallo-β-lactamases, and extended-spectrum oxacillinases produced by ceftazidime-resistant Pseudomonas aeruginosa strains in France. Antimicrob. Agents Chemother. 2010, 54, 3512–3515. [Google Scholar] [CrossRef] [Green Version]
  27. Poole, K. Pseudomonas Aeruginosa: Resistance to the Max. Front. Microbiol. 2011, 2, 65. [Google Scholar] [CrossRef] [Green Version]
  28. Zhao, W.H.; Hu, Z.Q. β-lactamases identified in clinical isolates of Pseudomonas aeruginosa. Crit. Rev. Microbiol. 2010, 36, 245–258. [Google Scholar] [CrossRef]
  29. Gheorghe, I.; Barbu, I.C.; Surleac, M.; Sârbu, I.; Popa, L.I.; Paraschiv, S.; Feng, Y.; Lazăr, V.; Chifiriuc, M.C.; Oțelea, D.; et al. Subtypes, resistance and virulence platforms in extended-drug resistant Acinetobacter baumannii Romanian isolates. Sci. Rep. 2021, 11, 1–12. [Google Scholar] [CrossRef]
  30. Gupta, V. Metallo β-lactamases in Pseudomonas aeruginosa and Acinetobacter species. Expert Opin. Investig. Drugs. 2008, 17, 131–143. [Google Scholar] [CrossRef]
  31. Toleman, M.A.; Rolston, K.; Jones, R.N.; Walsh, T.R. Molecular and biochemical characterization of OXA-45, an extended-spectrum class 2d′ β-lactamase in Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 2003, 47, 2859–2863. [Google Scholar] [CrossRef] [Green Version]
  32. Juan, C.; Mulet, X.; Zamorano, L.; Alberti, S.; Perez, J.L.; Oliver, A. Detection of the novel extended-spectrum β-lactamase OXA-161 from a plasmid-located integron in Pseudomonas aeruginosa clinical isolates from Spain. Antimicrob. Agents Chemother. 2009, 53, 5288–5290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Fournier, D.; Hocquet, D.; Dehecq, B.; Cholley, P.; Plesiat, P. Detection of a new extended-spectrum oxacillinase in Pseudomonas aeruginosa. J. Antimicrob. Chemother. 2010, 65, 364–365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Rodríguez-Baño, J.; Gutiérrez-Gutiérrez, B.; Machuca, I.; Pascual, A. Treatment of Infections Caused by Extended-Spectrum-Beta-Lactamase-, AmpC-, and Carbapenemase-Producing Enterobacteriaceae. Clin. Microbiol. Rev. 2018, 31, e00079-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Smet, A.; Martel, A.; Persoons, D.; Dewulf, J.; Heyndrickx, M.; Herman, L.; Haesebrouck, F.; Butaye, P. Broad-spectrum beta-lactamases among Enterobacteriaceae of animal origin: Molecular aspects, mobility and impact on public health. FEMS Microbiol Rev. 2010, 34, 295–316. [Google Scholar] [CrossRef] [Green Version]
  36. Ramirez, M.S.; Tolmasky, M.E. Aminoglycoside modifying enzymes. Drug Resist. Updat. 2010, 13, 151–171. [Google Scholar] [CrossRef] [Green Version]
  37. De Silva, P.M.; Kumar, A. Signal transduction proteins in Acinetobacter baumannii: Role in antibiotic resistance, virulence, and potential as drug targets. Front. Microbiol. 2019, 10, 49. [Google Scholar] [CrossRef] [Green Version]
  38. Ratjen, F.; Brockhaus, F.; Angyalosi, G. Aminoglycoside therapy against Pseudomonas aeruginosa in cystic fibrosis: A review. J. Cyst. Fibros. 2009, 8, 361–369. [Google Scholar] [CrossRef]
  39. Chukwudi, C.U. rRNA binding sites and the molecular mechanism of action of the tetracyclines. Antimicrob. Agents Chemother. 2016, 60, 4433–4441. [Google Scholar] [CrossRef] [Green Version]
  40. Sutcliffe, J.A.; O’Brien, W.; Fyfe, C.; Grossman, T.H. Antibacterial activity of eravacycline (TP-434), a novel fluorocycline, against hospital and community pathogens. Antimicrob. Agents Chemother. 2013, 57, 5548–5558. [Google Scholar] [CrossRef] [Green Version]
  41. Nikolakopoulou, T.L.; Giannoutsou, E.P.; Karabatsou, A.A.; Karagouni, A.D. Prevalence of tetracycline resistance genes in Greek seawater habitats. J. Microbiol. 2008, 46, 633–640. [Google Scholar] [CrossRef]
  42. Lin, L.; Wang, S.F.; Yang, T.Y.; Hung, W.C.; Chan, M.Y.; Tseng, S.P. Antimicrobial resistance and genetic diversity in ceftazidime nonsusceptible bacterial pathogens from ready-to-eat street foods in three Taiwanese cities. Sci. Rep. 2017, 7, 15515. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Coyne, S.; Courvalin, P.; Périchon, B. Efflux-mediated antibiotic resistance in Acinetobacter spp. Antimicrob. Agents Chemother. 2011, 55, 947–953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Nang, S.C.; Morris, F.C.; McDonald, M.J.; Han, M.L.; Wang, J.; Strugnell, R.A.; Velkov, T.; Li, J. Fitness cost of mcr-1-mediated polymyxin resistance in Klebsiella pneumoniae. J. Antimicrob. Chemother. 2018, 73, 1604–1610. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Chen, Q.; Li, X.; Zhou, H.; Jiang, Y.; Chen, Y.; Hua, X.; Yu, Y. Decreased susceptibility to tigecycline in Acinetobacter baumannii mediated by a mutation in trm encoding SAM-dependent methyltransferase. J. Antimicrob. Chemother. 2014, 69, 72–76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Poirel, L.; Corvec, S.; Rapoport, M.; Mugnier, P.; Petroni, A.; Pasteran, F.; Faccone, D.; Galas, M.; Drugeon, H.; Cattoir, V.; et al. Identification of the novel narrow-spectrum beta-lactamase SCO-1 in Acinetobacter spp. from Argentina. Antimicrob Agents Chemother. 2007, 51, 2179–2184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Boissinot, M.; Levesque, R.C. Nucleotide sequence of the PSE-4 carbenicillinase gene and correlations with the Staphylococcus aureus PC1 b-lactamase crystal structure. J. Biol. Chem. 1990, 265, 1225–1230. [Google Scholar] [CrossRef]
  48. Huovinen, P.; Jacoby, G.A. Sequence of the PSE-1 b-lactamase gene. Antimicrob. Agents Chemother. 1991, 35, 2428–2430. [Google Scholar] [CrossRef] [Green Version]
  49. Lachapelle, J.; Dufresne, J.; Levesque, R.C. Characterization of the blaCARB-3 gene encoding the carbenicillinase-3 b-lactamase of Pseudomonas aeruginosa. Gene 1991, 102, 7–12. [Google Scholar] [CrossRef]
  50. Sanschagrin, F.; Bejaoui, N.; Levesque, R.C. Structure of CARB-4 and AER-1 carbenicillin-hydrolyzing b-lactamases. Antimicrob. Agents Chemother. 1998, 42, 1966–1972. [Google Scholar] [CrossRef] [Green Version]
  51. Mavroidi, A.; Tzelepi, E.; Tsakris, A.; Miriagou, V.; Sofianou, D.; Tzouvelekis, L.S. An integron-associated b-lactamase (IBC-2) from Pseudomonas aeruginosa is a variant of the extended-spectrum b-lactamase IBC-1. J. Antimicrob. Chemother. 2001, 48, 627–630. [Google Scholar] [CrossRef] [Green Version]
  52. Dai, X.; Zhou, D.; Xiong, W.; Feng, J.; Luo, W.; Luo, G.; Wang, H.; Sun, F.; Zhou, X. The IncP-6 plasmid p10265-KPC from Pseudomonas aeruginosa carries a novel deltaISEc33-associated blaKPC-2 gene cluster. Front. Microbiol. 2016, 7, 310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. De Wals, P.Y.; Doucet, N.; Pelletier, J.N. High tolerance to simultaneous active-site mutations in TEM-1 beta-lactamase: Distinct mutational paths provide more generalized β-lactam recognition. Protein Sci. 2009, 18, 147–160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Endimiani, A.; Luzzaro, F.; Migliavacca, R.; Mantengoli, E.; Hujer, A.R.; Hujer, K.M.; Pagani, L.; Bonomo, R.A.; Rossolini, G.M.; Toniolo, A. Spread in an Italian hospital of a clonal Acinetobacter baumannii strains producing the TEM-92 extended spectrum β-lactamase. Antimicrob. Agents Chemother. 2007, 51, 2211–2214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Lee, C.-R.; Lee, J.H.; Park, M.; Park, K.S.; Bae, I.K.; Kim, Y.B.; Cha, C.-J.; Jeong, B.C.; Lee, S.H. Biology of Acinetobacter baumannii: Pathogenesis, Antibiotic Resistance Mechanisms, and Prospective Treatment Options. Front. Cell Infect. Microbiol. 2017, 7, 55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Moubareck, C.; Brémont, S.; Conroy, M.-C.; Courvalin, P.; Lambert, T. GES-11, a novel integron-associated GES variant in Acinetobacter baumannii. Antimicrob. Agents Chemother. 2009, 53, 3579–3581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Hammoudi, D.; Moubareck, C.A.; Hakime, N.; Houmani, M.; Barakat, A.; Najjar, Z.; Suleiman, M.; Fayad, N.; Sarraf, R.; Sarkis, D.K. Spread of imipenem-resistant Acinetobacter baumannii co-expressing OXA-23 and GES-11 carbapenemases in Lebanon. Int. J. Infect. Dis. 2015, 36, 56–61. [Google Scholar] [CrossRef] [Green Version]
  58. Celenza, G.; Pellegrini, C.; Caccamo, M.; Segatore, B.; Amicosante, G.; Perilli, M. Spread of bla(CTX-M-type) and bla(PER-2) β-lactamase genes in clinical isolates from Bolivian hospitals. J. Antimicrob. Chemother. 2006, 57, 975–978. [Google Scholar] [CrossRef] [Green Version]
  59. Ghanaie, R.M.; Fallah, F.; Noori, M.; Karimi, A.; Hashemi, A. Prevalence of blaPER-1 and blaVEB-1 genes among ESBL-producing Acinetobacter baumannii isolated from two hospitals of Tehran, Iran. J. Drug Metab. Toxicol. 2016, 7, e8888. [Google Scholar] [CrossRef] [Green Version]
  60. Vahaboglu, H.; Ozturk, R.; Aygun, G.; Coskunkan, F.; Yaman, A.; Kaygusuz, A.; Leblebicioglu, H.; Balik, I.; Aydin, K.; Otkun, M. Widespread detection of PER-1-type extended-spectrum beta-lactamases among nosocomial Acinetobacter and Pseudomonas aeruginosa isolates in Turkey: A nationwide multicenter study. Antimicrob. Agents Chemother. 1997, 41, 2265–2269. [Google Scholar] [CrossRef] [Green Version]
  61. Wolter, D.J.; Lister, P.D. Mechanisms of beta-lactam resistance among Pseudomonas aeruginosa. Curr. Pharm. Des. 2013, 19, 209–222. [Google Scholar] [CrossRef]
  62. Shaikh, S.; Fatima, J.; Shakil, S.; Rizvi, S.M.D.; Kamal, M.A. Antibiotic resistance and extended spectrum beta-lactamases: Types, epidemiology and treatment. Saudi J. Biol. Sci. 2015, 22, 90–101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Croughs, P.D.; Klaassen, C.H.W.; van Rosmalen, J.; Maghdid, D.M.; Boers, S.A.; Hays, J.P.; Goessens, W.H.F. Unexpected mechanisms of resistance in Dutch Pseudomonas aeruginosa isolates collected during 14 years of surveillance. Int. J. Antimicrob. Agents 2018, 52, 407–410. [Google Scholar] [CrossRef] [PubMed]
  64. Jacoby, G.A.; Matthew, M. The distribution of beta-lactamase genes on plasmids found in Pseudomonas. Plasmid 1979, 2, 41–47. [Google Scholar] [CrossRef] [PubMed]
  65. Naas, T.; Philippon, L.; Poirel, L.; Ronco, E.; Nordmann, P. An SHV-derived extended-spectrum β-lactamase in Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 1999, 43, 1281–1284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Bauernfeind, A.; Stemplinger, I.; Jungwirth, R.; Mangold, P.; Amann, S.; Akalin, E.; Anğ, O.; Bal, C.; Casellas, J.M. Characterization of β-lactamase gene blaPER-2, which encodes an extended-spectrum class A β-lactamase. Antimicrob. Agents Chemother. 1996, 40, 616–620. [Google Scholar] [CrossRef] [Green Version]
  67. Girlich, D.; Naas, T.; Leelaporn, A.; Poirel, L.; Fennewald, M.; Nordmann, P. Nosocomial Spread of the Integron-Located veb-1—Like Cassette Encoding an Extended-Spectrum β-Lactamase in Pseudomonas aeruginosa in Thailand. Clin. Infect. Dis. 2002, 34, 603–611. [Google Scholar] [CrossRef] [Green Version]
  68. Laudy, A.E.; Róg, P.; Smolińska-Król, K.; Ćmiel, M.; Słoczyńska, A.; Patzer, J.; Dzierżanowska, D.; Wolinowska, R.; Starościak, B.; Tyski, S. Prevalence of ESBL-producing Pseudomonas aeruginosa isolates in Warsaw, Poland, detected by various phenotypic and genotypic methods. PLoS ONE 2017, 12, e0180121. [Google Scholar] [CrossRef] [Green Version]
  69. Bogaerts, P.; Bauraing, C.; Deplano, A.; Glupczynski, Y. Emergence and dissemination of BEL-1-producing Pseudomonas aeruginosa isolates in Belgium. Antimicrob. Agents Chemother. 2007, 51, 1584–1585. [Google Scholar] [CrossRef] [Green Version]
  70. Poirel, L.; Docquier, J.D.; De Luca, F.; Verlinde, A.; Ide, L.; Rossolini, G.M.; Nordmann, P. BEL-2, an extended-spectrum β-lactamase with increased activity toward expanded-spectrum cephalosporins in Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 2010, 54, 533–535. [Google Scholar] [CrossRef] [Green Version]
  71. Paulitsch-Fuchs, A.H.; Melchior, N.; Haitzmann, T.; Fingerhut, T.; Feierl, G.; Baumert, R.; Kittinger, C.; Zarfel, G. Analysis of Extended Spectrum Beta Lactamase (ESBL) Genes of Non-Invasive ESBL Enterobacterales in Southeast Austria in 2017. Antibiotics 2022, 12, 1. [Google Scholar] [CrossRef]
  72. Legese, M.H.; Asrat, D.; Aseffa, A.; Hasan, B.; Mihret, A.; Swedberg, G. Molecular Epidemiology of Extended-Spectrum Beta-Lactamase and AmpC Producing Enterobacteriaceae among Sepsis Patients in Ethiopia: A Prospective Multicenter Study. Antibiotics 2022, 11, 131. [Google Scholar] [CrossRef] [PubMed]
  73. Liakopoulos, A.; Mevius, D.; Ceccarelli, D. A Review of SHV Extended-Spectrum β-Lactamases: Neglected Yet Ubiquitous. Front. Microbiol. 2016, 7, 1374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Fernandes, R.; Amador, P.; Oliveira, C.; Prudêncio, C. Molecular Characterization of ESBL-Producing Enterobacteriaceae in Northern Portugal. Sci. World J. 2014, 2014, 782897. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Robledo, I.E.; Aquino, E.E.; Santé, M.I.; Santana, J.L.; Otero, D.M.; León, C.F.; Vázquez, G.J. Detection of KPC in Acinetobacter spp. in Puerto Rico. Antimicrob. Agents Chemother. 2010, 54, 1354–1357. [Google Scholar] [CrossRef] [Green Version]
  76. Ribeiro, P.C.S.; Monteiro, A.S.; Marques, S.G.; Monteiro, S.G.; Monteiro-Neto, V.; Coqueiro, M.M.M.; Marques, A.C.G.; de Jesus Gomes Turri, R.; Santos, S.G.; Bomfim, M.R.Q. Phenotypic and molecular detection of the bla KPC gene in clinical isolates from inpatients at hospitals in São Luis, MA, Brazil. BMC Infect. Dis. 2016, 16, 737. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Caneiras, C.; Calisto, F.; Jorge da Silva, G.; Lito, L.; Melo-Cristino, J.; Duarte, A. First Description of Colistin and Tigecycline-Resistant Acinetobacter baumannii Producing KPC-3 Carbapenemase in Portugal. Antibiotics 2018, 7, 96. [Google Scholar] [CrossRef] [Green Version]
  78. Dubois, V.; Poirel, L.; Marie, C.; Arpin, C.; Nordmann, P.; Quentin, C. Molecular characterization of a novel class 1 integron containing bla(GES-1) and a fused product of aac3-Ib/aac6′-Ib’ gene cassettes in Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 2002, 46, 638–645. [Google Scholar] [CrossRef] [Green Version]
  79. Poirel, L.; Weldhagen, G.F.; De Champs, C.; Nordmann, P. A nosocomial outbreak of Pseudomonas aeruginosa isolates expressing the extendedspectrum b-lactamase GES-2 in South Africa. J. Antimicrob. Chemother. 2002, 49, 561–565. [Google Scholar] [CrossRef] [Green Version]
  80. Viedma, E.; Juan, C.; Acosta, J.; Zamorano, L.; Otero, J.R.; Sanz, F.; Chaves, F.; Oliver, A. Nosocomial spread of colistin-only-sensitive sequence type 235 Pseudomonas aeruginosa isolates producing the extended-spectrum beta-lactamases GES-1 and GES-5 in Spain. Antimicrob. Agents Chemother. 2009, 53, 4930–4933. [Google Scholar] [CrossRef] [Green Version]
  81. Garza-Ramos, U.; Barrios, H.; Reyna-Flores, F.; Tamayo-Legorreta, E.; Catalan-Najera, J.C.; Morfin-Otero, R.; Rodríguez-Noriega, E.; Volkow, P.; Cornejo-Juarez, P.; González, A.; et al. Widespread of ESBL- and carbapenemase GES-type genes on carbapenem-resistant Pseudomonas aeruginosa clinical isolates: A multicenter study in Mexican hospitals. Diagn. Microbiol. Infect. Dis. 2015, 81, 135–137. [Google Scholar] [CrossRef]
  82. Botelho, J.; Grosso, F.; Peixe, L. Unravelling the genome of a Pseudomonas aeruginosa isolate belonging to the high-risk clone ST235 reveals an integrative conjugative element housing a blaGES-6 carbapenemase. J. Antimicrob. Chemother. 2018, 73, 77–83. [Google Scholar] [CrossRef] [PubMed]
  83. Schauer, J.; Gatermann, S.G.; Hoffmann, D.; Hupfeld, L.; Pfennigwerth, N. GPC-1, a novel class A carbapenemase detected in a clinical Pseudomonas aeruginosa isolate. J. Antimicrob. Chemother. 2020, 75, 911–916. [Google Scholar] [CrossRef]
  84. Naas, T.; Nordmann, P. OXA-type b-lactamases. Curr. Pharm. Des. 1999, 5, 865–879. [Google Scholar] [CrossRef] [PubMed]
  85. Hall, L.M.; Livermore, D.M.; Gur, D.; Akova, M.; Akalin, H.E. OXA-11, an extended-spectrum variant of OXA-10 (PSE-2) beta-lactamase from Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 1993, 37, 1637–1644. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Aubert, D.; Poirel, L.; Chevalier, J.; Leotard, S.; Pages, J.M.; Nordmann, P. Oxacillinase-mediated resistance to cefepime and susceptibility to ceftazidime in Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 2001, 45, 1615–1620. [Google Scholar] [CrossRef] [Green Version]
  87. Chen, Y.; Zhou, Z.; Jiang, Y.; Yu, Y. Emergence of NDM-1-producing Acinetobacter baumannii in China. J. Antimicrob. Chemother. 2011, 66, 1255–1259. [Google Scholar] [CrossRef] [Green Version]
  88. Jaidane, N.; Naas, T.; Oueslati, S.; Bernabeu, S.; Boujaafar, N.; Bouallegue, O.; Bonnin, R.A. Whole-genome sequencing of NDM-1-producing ST85 Acinetobacter baumannii isolates from Tunisia. Int. J. Antimicrob. Agents 2018, 52, 916–921. [Google Scholar] [CrossRef]
  89. Beigverdi, R.; Sattari-Maraji, A.; Emaneini, M.; Jabalameli, F. Status of carbapenem-resistant Acinetobacter baumannii harboring carbapenemase: First systematic review and meta-analysis from Iran. Infect. Genet. Evol. 2019, 73, 433–443. [Google Scholar] [CrossRef]
  90. Salloum, T.; Tannous, E.; Alousi, S.; Arabaghian, H.; Rafei, R.; Hamze, M.; Tokajian, S. Genomic mapping of ST85 blaNDM-1 and blaOXA-94 producing Acinetobacter baumannii isolates from Syrian Civil War Victims. Int. J. Infect. Dis. 2018, 74, 100–108. [Google Scholar] [CrossRef] [Green Version]
  91. El-Mahdy, T.S.; Al-Agamy, M.H.; Al-Qahtani, A.A.; Shibl, A.M. Detection of blaOXA-23-like and blaNDM-1 in Acinetobacter baumannii from the Eastern Region, Saudi Arabia. Microb. Drug Resist. 2017, 23, 115–121. [Google Scholar] [CrossRef]
  92. Bonnin, R.A.; Poirel, L.; Naas, T.; Pirs, M.; Seme, K.; Schrenzel, J.; Nordmann, P. Dissemination of New Delhi metallo-β-lactamase-1-producing Acinetobacter baumannii in Europe. Clin. Microbiol. Infect. 2012, 18, E362–E365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Hrabák, J.; Štolbová, M.; Študentová, V.; Fridrichová, M.; Chudáčková, E.; Zemlickova, H. NDM-1 producing Acinetobacter baumannii isolated from a patient repatriated to the Czech Republic from Egypt, July 2011. Euro. Surveill. 2012, 17, 1–3. [Google Scholar] [CrossRef]
  94. Kaase, M.; Nordmann, P.; Wichelhaus, T.A.; Gatermann, S.G.; Bonnin, R.A.; Poirel, L. NDM-2 carbapenemase in Acinetobacter baumannii from Egypt. J. Antimicrob. Chemother. 2011, 66, 1260–1262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Ramadan, R.A.; Gebriel, M.G.; Kadry, H.M.; Mosallem, A. Carbapenem-resistant Acinetobacter baumannii and Pseudomonas aeruginosa: Characterization of carbapenemase genes and E-test evaluation of colistin-based combinations. Infect. Drug Resist. 2018, 11, 1261–1269. [Google Scholar] [CrossRef] [Green Version]
  96. Tsakris, A.; Ikonomidis, A.; Pournaras, S.; Tzouvelekis, L.S.; Sofianou, D.; Legakis, N.J.; Maniatis, A.N. VIM-1 metallo-beta-lactamase in Acinetobacter baumannii. Emerg. Infect. Dis. 2006, 12, 981–983. [Google Scholar] [CrossRef] [Green Version]
  97. Lee, K.; Lee, W.G.; Uh, Y.; Ha, G.Y.; Cho, J.; Chong, Y. VIM- and IMP-type metallo-beta-lactamase-producing Pseudomonas spp. and Acinetobacter spp. in Korean hospitals. Emerg. Infect. Dis. 2003, 9, 868–871. [Google Scholar] [CrossRef]
  98. Lee, M.F.; Peng, C.F.; Hsu, H.J.; Chen, Y.H. Molecular characterisation of the metallo-beta-lactamase genes in imipenem-resistant Gram-negative bacteria from a university hospital in southern Taiwan. Int. J. Antimicrob. Agents 2008, 32, 475–480. [Google Scholar] [CrossRef]
  99. Girija, S.A.; Jayaseelan, V.P.; Arumugam, P. Prevalence of VIM- and GIM-producing Acinetobacter baumannii from patients with severe urinary tract infection. Acta Microbiol. Immunol. Hung. 2018, 65, 539–550. [Google Scholar] [CrossRef] [Green Version]
  100. Gholami, M.; Moshiri, M.; Ahanjan, M.; Chirani, A.S.; Bibalan, M.H.; Asadi, A.; Eshaghi, M.; Pournajaf, A.; Abbasian, S.; Kouhsari, E.; et al. The diversity of class B and class D carbapenemases in clinical Acinetobacter baumannii isolates. Infez. Med. 2018, 26, 329–335. [Google Scholar]
  101. Aghamiri, S.; Amirmozafari, N.; Mehrabadi, J.F.; Fouladtan, B.; Abdar, M.H. Antibiotic Resistance Patterns and a Survey of Metallo-β-Lactamase Genes Including bla-IMP and bla-VIM Types in Acinetobacter baumannii Isolated from Hospital Patients in Tehran. Chemotherapy 2016, 61, 275–280. [Google Scholar] [CrossRef]
  102. Cornaglia, G.; Riccio, M.L.; Mazzariol, A.; Lauretti, L.; Fontana, R.; Rossolini, G.M. Appearance of IMP-1 metallo-β-lactamase in Europe. Lancet 1999, 353, 899–900. [Google Scholar] [CrossRef] [PubMed]
  103. Iyobe, S.; Kusadokoro, H.; Ozaki, J.; Matsumura, N.; Minami, S.; Haruta, S.; Sawai, T.; O’Hara, K. Amino acid substitutions in a variant of IMP-1 metallo-β-lactamase. Antimicrob. Agents Chemother. 2000, 44, 2023–2027. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Chu, Y.-W.; Afzal-Shah, M.; Houang, E.T.S.; Palepou, M.-F.I.; Lyon, D.J.; Woodford, N.; Livermore, D.M. IMP-4, a novel metallo-β-lactamase from nosocomial Acinetobacter spp. collected in Hong-Kong between 1994 and 1998. Antimicrob. Agents Chemother. 2001, 45, 710–714. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Da Silva, G.J.; Correia, M.; Vital, C.; Ribeiro, G.; Sousa, J.C.; Leitao, R.; Peixe, L.; Duarte, A. Molecular characterization of bla(IMP-5) a new integron born metallo-β-lactamase gene from an Acinetobacter baumannii nosocomial isolate in Portugal. FEMS Microbiol. Lett. 2002, 24, 33–39. [Google Scholar] [CrossRef] [Green Version]
  106. Ramirez, M.S.; Bonomo, R.A.; Tolmasky, M.E. Carbapenemases: Transforming Acinetobacter baumannii into a Yet More Dangerous Menace. Biomolecules 2020, 10, 720. [Google Scholar] [CrossRef]
  107. Rezaei, A.; Fazeli, H.; Halaji, M.; Moghadampour, M.; Faghri, J. Prevalence of metallo-beta-lactamase producing Acinetobacter baumannii isolated from intensive care unit in tertiary care hospitals. Ann. Ig. Med. Prev. Comunita. 2018, 30, 330–336. [Google Scholar] [CrossRef]
  108. Shakibaie, M.R.; Azizi, O.; Shahcheraghi, F. Insight into stereochemistry of a new IMP allelic variant (IMP-55) metallo-β-lactamase identified in a clinical strain of Acinetobacter baumannii. Infect. Genet. Evol. 2017, 51, 118–126. [Google Scholar] [CrossRef] [PubMed]
  109. Fallah, F.; Borhan, R.S.; Hashemi, A. Detection of bla(IMP) and bla(VIM) metallo-beta-lactamases genes among Pseudomonas aeruginosa strains. Int. J. Burns Trauma. 2013, 3, 122–124. [Google Scholar] [PubMed]
  110. Amin, M.; Navidifar, T.; Shooshtari, F.S.; Goodarzi, H. Association of the genes encoding Metallo-β-Lactamase with the presence of integrons among multidrug-resistant clinical isolates of Acinetobacter baumannii. Infect. Drug Resist. 2019, 12, 1171–1180. [Google Scholar] [CrossRef] [Green Version]
  111. Xiong, J.; Hynes, M.F.; Ye, H.; Chen, H.; Yang, Y.; M’zali, F.; Hawkey, P.M. blaIMP-9 and its association with large plasmids carried by Pseudomonas aeruginosa isolates from the People’s Republic of China. Antimicrob. Agents Chemother. 2006, 50, 355–358. [Google Scholar] [CrossRef] [Green Version]
  112. Yano, H.; Kuga, A.; Okamoto, R.; Kitasato, H.; Kobayashi, T.; Inoue, M. Plasmid-encoded metallo-beta-lactamase (IMP-6) conferring resistance to carbapenems, especially meropenem. Antimicrob. Agents Chemother. 2001, 45, 1343–1348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Tada, T.; Nhung, P.H.; Miyoshi-Akiyama, T.; Shimada, K.; Phuong, D.M.; Anh, N.Q.; Ohmagari, N.; Kirikae, T. IMP-51, a novel IMP-type metallo-β-lactamase with increased doripenem and meropenem hydrolyzing activities, in a carbapenem-resistant Pseudomonas aeruginosa clinical isolate. Antimicrob. Agents Chemother. 2015, 59, 7090–7093. [Google Scholar] [CrossRef] [Green Version]
  114. Yoon, E.-J.; Jeong, S.H. Mobile Carbapenemase Genes in Pseudomonas aeruginosa. Front. Microbiol. 2021, 12, 614058. [Google Scholar] [CrossRef] [PubMed]
  115. Lolans, K.; Queenan, A.M.; Bush, K.; Sahud, A.; Quinn, J.P. First nosocomial outbreak of Pseudomonas aeruginosa producing an integron-borne metallo-b-lactamase (VIM-2) in the United States. Antimicrob. Agents Chemother. 2005, 49, 3538–3540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. 135Cornaglia, G.; Mazzariol, A.; Lauretti, L.; Rossolini, G.M.; Fontana, R. Hospital outbreak of carbapenem-resistant Pseudomonas aeruginosa producing VIM-1, a novel transferable metallo–b-lactamase. Clin. Infect. Dis. 2000, 31, 1119–1125. [Google Scholar] [CrossRef] [Green Version]
  117. Yan, J.J.; Hsueh, P.R.; Ko, W.C.; Luh, K.T.; Tsai, S.H.; Wu, H.M.; Wu, J.J. Metallo-β-lactamases in clinical Pseudomonas isolates in Taiwan and identification of VIM-3, a novel variant of the VIM-2 enzyme. Antimicrob. Agents Chemother. 2001, 45, 2224–2228. [Google Scholar] [CrossRef] [Green Version]
  118. Jovcic, B.; Lepsanovic, Z.; Suljagic, V.; Rackov, G.; Begovic, J.; Topisirovic, L.; Kojic, M. Emergence of NDM-1 metallo-beta-lactamase in Pseudomonas aeruginosa clinical isolates from Serbia. Antimicrob. Agents Chemother. 2011, 55, 3929–3931. [Google Scholar] [CrossRef] [Green Version]
  119. Toleman, M.A.; Simm, A.M.; Murphy, T.A.; Gales, A.C.; Biedenbach, D.J.; Jones, R.N.; Walsh, T.R. Molecular characterization of SPM-1, a novel metallo-beta-lactamase isolated in Latin America: Report from the SENTRY antimicrobial surveillance programme. J. Antimicrob. Chemother. 2002, 50, 673–679. [Google Scholar] [CrossRef] [Green Version]
  120. Cacci, L.C.; Chuster, S.G.; Martins, N.; Do Carmo, P.R.; De Carvalho Girão, V.B.; Nouér, S.A.; De Freitas, W.V.; De Matos, J.A.; De Gouveia Magalhães, A.C.; Ferreira, A.L.P.; et al. Mechanisms of carbapenem resistance in endemic Pseudomonas aeruginosa isolates after an SPM-1 metallo-β-lactamase producing strain subsided in an intensive care unit of a teaching hospital in Brazil. Mem. Inst. Oswaldo Cruz 2016, 111, 551–558. [Google Scholar] [CrossRef] [Green Version]
  121. Tada, T.; Shimada, K.; Satou, K.; Hirano, T.; Pokhrel, B.M.; Sherchand, J.B.; Kirikae, T. Pseudomonas aeruginosa clinical isolates in nepal coproducing metallo-beta-lactamases and 16S rRNA methyltransferases. Antimicrob. Agents Chemother. 2017, 61, e00694-17. [Google Scholar] [CrossRef] [Green Version]
  122. Sun, F.; Zhou, D.; Wang, Q.; Feng, J.; Feng, W.; Luo, W.; Zhang, D.; Liu, Y.; Qiu, X.; Yin, Z.; et al. The first report of detecting the blaSIM-2 gene and determining the complete sequence of the SIM-encoding plasmid. Clin. Microbiol. Infect. 2016, 22, 347–351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Pfennigwerth, N.; Lange, F.; Campos, C.B.; Hentschke, M.; Gatermann, S.G.; Kaase, M. Genetic and biochemical characterization of HMB-1, a novel subclass B1 metallo-beta-lactamase found in a Pseudomonas aeruginosa clinical isolate. J. Antimicrob. Chemother. 2017, 72, 1068–1073. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Boyd, D.A.; Lisboa, L.F.; Rennie, R.; Zhanel, G.G.; Dingle, T.C.; Mulvey, M.R. Identification of a novel metallo-beta-lactamase, CAM-1, in clinical Pseudomonas aeruginosa isolates from Canada. J. Antimicrob. Chemother. 2019, 74, 1563–1567. [Google Scholar] [CrossRef] [PubMed]
  125. Yong, D.; Toleman, M.A.; Bell, J.; Ritchie, B.; Pratt, R.; Ryley, H.; Walsh, T.R. Genetic and biochemical characterization of an acquired subgroup B3 metallo-beta-lactamase gene, blaAIM-1, and its unique genetic context in Pseudomonas aeruginosa from Australia. Antimicrob. Agents Chemother. 2012, 56, 6154–6159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Pollini, S.; Maradei, S.; Pecile, P.; Olivo, G.; Luzzaro, F.; Docquier, J.-D.; Rossolini, G.M. FIM-1, a new acquired metallo-beta-lactamase from a Pseudomonas aeruginosa clinical isolate from Italy. Antimicrob. Agents Chemother. 2013, 57, 410–416. [Google Scholar] [CrossRef] [Green Version]
  127. Wareth, G.; Brandt, C.; Sprague, L.D.; Neubauer, H.; Pletz, M.W. WGS based analysis of acquired antimicrobial resistance in human and non-human Acinetobacter baumannii isolates from a German perspective. BMC Microbiol. 2021, 21, 210. [Google Scholar] [CrossRef]
  128. Ingti, B.; Upadhyay, S.; Hazarika, M.; Khyriem, A.B.; Paul, D.; Bhattacharya, P.; Joshi, S.R.; Bora, D.; Dhar, D.; Bhattacharjee, A. Distribution of carbapenem resistant Acinetobacter baumannii with blaADC-30 and induction of ADC-30 in response to beta-lactam antibiotics. Res. Microbiol. 2020, 171, 128–133. [Google Scholar] [CrossRef]
  129. Peleg, A.Y.; Seifert, H.; Paterson, D.L. Acinetobacter baumannii: Emergence of a successful pathogen. Clin. Microbiol. Rev. 2008, 21, 538–582. [Google Scholar] [CrossRef] [Green Version]
  130. Kumburu, H.H.; Sonda, T.; van Zwetselaar, M.; Leekitcharoenphon, P.; Lukjancenko, O.; Mmbaga, B.T.; Alifrangis, M.; Lund, O.; Aarestrup, F.M.; Kibiki, G.S. Using WGS to identify antibiotic resistance genes and predict antimicrobial resistance phenotypes in MDR Acinetobacter baumannii in Tanzania. J. Antimicrob. Chemother. 2019, 74, 1484–1493. [Google Scholar] [CrossRef] [Green Version]
  131. Jia, H.; Chen, Y.; Wang, J.; Ruan, Z. Genomic characterisation of a clinical Acinetobacter baumannii ST1928 isolate carrying a new ampC allelic variant blaADC-196 gene from China. J. Glob. Antimicrob. Resist. 2019, 19, 43–45. [Google Scholar] [CrossRef]
  132. Mayanskiy, N.; Chebotar, I.; Alyabieva, N.; Kryzhanovskaya, O.; Savinova, T.; Turenok, A.; Bocharova, Y.; Lazareva, A.; Polikarpova, S.; Karaseva, O. Emergence of the Uncommon Clone ST944/ST78 Carrying blaOXA-40-like and blaCTX-M-like Genes Among Carbapenem-Nonsusceptible Acinetobacter baumannii in Moscow, Russia. Microb. Drug Resist. 2017, 23, 864–870. [Google Scholar] [CrossRef] [PubMed]
  133. Ramirez, M.S.; Adams, M.D.; Bonomo, R.A.; Centrón, D.; Tolmasky, M.E. Genomic Analysis of Acinetobacter baumannii A118 by Comparison of Optical Maps: Identification of Structures Related to Its Susceptibility Phenotype. Antimicrob. Agents Chemother. 2011, 55, 1520–1526. [Google Scholar] [CrossRef] [Green Version]
  134. Xiao-min, X.; You-fen, F.; Wei-yun, F.; Xing-bei, W. Antibiotic resistance determinants of a group of multidrug-resistant Acinetobacter baumannii in China. J. Antibiot. 2014, 67, 439–444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Evans, B.A.; Amyes, S.G.B. OXA β-lactamases. Clin. Microbiol. Rev. 2014, 27, 241–263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Jeon, J.H.; Hong, M.K.; Lee, J.H.; Lee, J.J.; Park, K.S.; Karim, A.M.; Jo, J.Y.; Kim, J.H.; Ko, K.S.; Kang, L.W.; et al. Structure of ADC-68, a novel carbapenem-hydrolyzing class C extended-spectrum β-lactamase isolated from Acinetobacter baumannii. Acta Crystallogr. D Biol. Crystallogr. 2014, 70, 2924–2936. [Google Scholar] [CrossRef]
  137. Upadhyay, S.; Mishra, S.; Sen, M.R.; Banerjee, T.; Bhattacharjee, A. Co-existence of Pseudomonas-derived cephalosporinase among plasmid encoded CMY-2 harbouring isolates of Pseudomonas aeruginosa in north India. Indian J. Med. Microbiol. 2013, 31, 257–260. [Google Scholar] [CrossRef]
  138. Fraile-Ribot, P.A.; Del Rosario-Quintana, C.; López-Causapé, C.; Gomis-Font, M.A.; Ojeda-Vargas, M.; Oliver, A. Emergence of resistance to novel β-lactam-β-lactamase inhibitor combinations due to horizontally acquired AmpC (FOX-4) in Pseudomonas aeruginosa sequence type 308. Antimicrob. Agents Chemother. 2019, 64, e02112-19. [Google Scholar] [CrossRef]
  139. Bush, K.; Bradford, P.A. Epidemiology of β-Lactamase-Producing Pathogens. Clin. Microbiol. Rev. 2020, 33, e00047-19. [Google Scholar] [CrossRef]
  140. Segatore, B.; Piccirilli, A.; Cherubini, S.; Principe, L.; Alloggia, G.; Mezzatesta, M.L.; Salmeri, M.; Di Bella, S.; Migliavacca, R.; Piazza, A.; et al. In Vitro Activity of Sulbactam-Durlobactam against Carbapenem-Resistant Acinetobacter baumannii Clinical Isolates: A Multicentre Report from Italy. Antibiotics 2022, 11, 1136. [Google Scholar] [CrossRef]
  141. Poirel, L.; Héritier, C.; Tolün, V.; Nordmann, P. Emergence of oxacillinase-mediated resistance to imipenem in Klebsiella pneumoniae. Antimicrob Agents Chemother. 2004, 48, 15–22. [Google Scholar] [CrossRef] [Green Version]
  142. Sevillano, E.; Gallego, L.; Garcia-Lobo, J.M. First detection of the OXA-40 carbapenemase in P. aeruginosa isolates, located on a plasmid also found in A. baumannii. Pathol. Biol. 2009, 57, 493–495. [Google Scholar] [CrossRef] [PubMed]
  143. Borah, V.V.; Saikia, K.K.; Hazarika, N.K. First report on the detection of OXA-48 beta-lactamase gene in Escherichia coli and Pseudomonas aeruginosa co-infection isolated from a patient in a tertiary care hospital in Assam. Indian J. Med. Microbiol. 2016, 34, 252–253. [Google Scholar] [CrossRef] [PubMed]
  144. Findlay, J.; Hopkins, K.L.; Loy, R.; Doumith, M.; Meunier, D.; Hill, R.; Pike, R.; Mustafa, N.; Livermore, D.M.; Woodford, N. OXA-48-like carbapenemases in the UK: An analysis of isolates and cases from 2007 to 2014. J. Antimicrob. Chemother. 2017, 72, 1340–1349. [Google Scholar] [CrossRef] [PubMed]
  145. Bonnin, R.A.; Bogaerts, P.; Girlich, D.; Huang, T.D.; Dortet, L.; Glupczynski, Y.; Nass, T. Molecular characterization of OXA-198 carbapenemase-producing Pseudomonas aeruginosa clinical isolates. Antimicrob. Agents Chemother. 2018, 62, e02496-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Figueiredo, S.; Poirel, L.; Papa, A.; Koulourida, V.; Nordmann, P. Overexpression of the naturally occurring blaOXA-51 gene in Acinetobacter baumannii mediated by novel insertion sequence ISAba9. Antimicrob. Agents Chemother. 2009, 53, 4045–4047. [Google Scholar] [CrossRef] [Green Version]
  147. Hou, C.; Yang, F. Drug-resistant gene of blaOXA-23, blaOXA-24, blaOXA-51 and blaOXA-58 in Acinetobacter baumannii. Int. J. Clin. Exp. Med. 2015, 8, 13859–13863. [Google Scholar]
  148. Scaife, W.; Young, H.K.; Paton, R.H.; Amyes, S.G. Transferable imipenem-resistance in Acinetobacter species from a clinical source. J. Antimicrob. Chemother. 1995, 36, 585–586. [Google Scholar] [CrossRef]
  149. Mugnier, P.D.; Poirel, L.; Naas, T.; Nordmann, P. Worldwide dissemination of the blaOXA-23 carbapenemase gene of Acinetobacter baumannii. Emerg. Infect. Dis. 2010, 16, 35–40. [Google Scholar] [CrossRef]
  150. Wibberg, D.; Salto, I.P.; Eikmeyer, F.G.; Maus, I.; Winkler, A.; Nordmann, P.; Pühler, A.; Poirel, L.; Schlüter, A. Complete Genome Sequencing of Acinetobacter baumannii Strain K50 Discloses the Large Conjugative Plasmid pK50a Encoding Carbapenemase OXA-23 and Extended-Spectrum β-Lactamase GES-11. Antimicrob. Agents Chemother. 2018, 62, e00212-18. [Google Scholar] [CrossRef] [Green Version]
  151. Santimaleeworagun, W.; Samret, W.; Preechachuawong, P.; Kerdsin, A.; Jitwasinkul, T. Emergence of Co-Carbapenemase Genes, Bla(Oxa23), Bla(Vim) And Bla(Ndm) In Carbapenemresistant Acinetobacter baumannii Clinical Isolates. Southeast Asian J. Trop. Med. Public Health 2016, 47, 1001–1007. [Google Scholar]
  152. Grosso, F.; Quinteira, S.; Poirel, L.; Novais, A.; Peixe, L. Role of common blaOXA-24/OXA-40-carrying platforms and plasmids in the spread of OXA-24/OXA-40 among Acinetobacter species clinical isolates. Antimicrob. Agents Chemother. 2012, 56, 3969–3972. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Shi, X.; Wang, H.; Wang, X.; Jing, H.; Duan, R.; Qin, S.; Lv, D.; Fan, Y.; Huang, Z.; Stirling, K.; et al. Molecular characterization and antibiotic resistance of Acinetobacter baumannii in cerebrospinal fluid and blood. PLoS ONE 2021, 16, e0247418. [Google Scholar] [CrossRef] [PubMed]
  154. Mathlouthi, N.; Lamine, Y.B.; Somai, R.; Bouhalila-Besbes, S.; Bakour, S.; Rolain, J.-M.; Chouchani, C. Incidence of OXA-23 and OXA-58 Carbapenemases Coexpressed in Clinical Isolates of Acinetobacter baumannii in Tunisia. Microb. Drug Resist. 2018, 24, 136–141. [Google Scholar] [CrossRef] [PubMed]
  155. Bertini, A.; Poirel, L.; Bernabeu, S.; Fortini, D.; Villa, L.; Nordmann, P.; Carattoli, A. Multicopy blaOXA-58 gene as a source of high-level resistance to carbapenems in Acinetobacter baumannii. Antimicrob. Agents Chemother. 2007, 51, 2324–2328. [Google Scholar] [CrossRef] [Green Version]
  156. Sarikhani, Z.; Nazari, R.; Rostami, M.N. First report of OXA-143-lactamase producing Acinetobacter baumannii in Qom, Iran. Iran J. Basic Med. Sci. 2017, 20, 1282–1286. [Google Scholar] [CrossRef] [PubMed]
  157. Zander, E.; Bonnin, R.A.; Seifert, H.; Higgins, P.G. Characterization of blaOXA-143 variants in Acinetobacter baumannii and Acinetobacter pittii. Antimicrob. Agents Chemother. 2014, 58, 2704–2708. [Google Scholar] [CrossRef] [Green Version]
  158. Neves, F.C.; Clemente, W.T.; Lincopan, N.; Paião, I.D.; Neves, P.R.; Romanelli, R.M.; Lima, S.S.S.; Paiva, L.F.; Mourão, P.H.O.; Nobre-Junior, V.A. Clinical and microbiological characteristics of OXA-23- and OXA-143-producing Acinetobacter baumannii in ICU patients at a teaching hospital, Brazil. Braz. J. Infect. Dis. 2016, 20, 556–563. [Google Scholar] [CrossRef] [Green Version]
  159. Higgins, P.G.; Pérez-Llarena, F.J.; Zander, E.; Fernández, A.; Bou, G.; Seifert, H. OXA-235, a novel class D β-lactamase involved in resistance to carbapenems in Acinetobacter baumannii. Antimicrob. Agents Chemother. 2013, 57, 2121–2126. [Google Scholar] [CrossRef] [Green Version]
  160. Behzadi, P.; Baráth, Z.; Gajdács, M. It’s not easy being green: A narrative review on the microbiology, virulence and therapeutic prospects of multidrug-resistant Pseudomonas aeruginosa. Antibiotics. 2021, 10, 42. [Google Scholar] [CrossRef]
  161. Jalal, S.; Ciofu, O.; Høiby, N.; Gotoh, N.; Wretlind, B. Molecular mechanisms of fluoroquinolone resistance in Pseudomonas aeruginosa isolates from cystic fibrosis patients. Antimicrob. Agents Chemother. 2000, 44, 710–712. [Google Scholar] [CrossRef] [Green Version]
  162. Aldred, K.J.; Kerns, R.J.; Osheroff, N. Mechanism of quinolone action and resistance. Biochemistry 2014, 53, 1565–1574. [Google Scholar] [CrossRef]
  163. Magnet, S.; Courvalin, P.; Lambert, T. Resistance-nodulation-cell division-type efflux pump involved in aminoglycoside resistance in Acinetobacter baumannii strain BM4454. Antimicrob. Agents Chemother. 2001, 45, 3375–3380. [Google Scholar] [CrossRef] [Green Version]
  164. Yang, H.; Hu, L.; Liu, Y.; Ye, Y.; Li, J. Detection of the plasmid-mediated quinolone resistance determinants in clinical isolates of Acinetobacter baumannii in China. J. Chemother. 2016, 28, 443–445. [Google Scholar] [CrossRef]
  165. Touati, A.; Brasme, L.; Benallaoua, S.; Gharout, A.; Madoux, J.; De Champs, C. First report of qnrB-producing Enterobacter cloacae and qnrA-producing Acinetobacter baumannii recovered from Algerian hospitals. Diagn Microbiol. Infect. Dis. 2008, 60, 287–290. [Google Scholar] [CrossRef] [PubMed]
  166. Mirnejad, R.; Heidary, M.; Bahramian, A.; Goudarzi, M.; Pournajaf, A. Evaluation of polymyxin B susceptibility profile and detection of drug resistance genes among Acinetobacter baumannii clinical isolates in Tehran, Iran during 2015-2016. Mediterr. J. Hematol. Infect. Dis. 2018, 10, e2018044. [Google Scholar] [CrossRef] [PubMed]
  167. Satlin, M.J.; Lewis, J.S.; Weinstein, M.P.; Patel, J.; Humphries, R.M.; Kahlmeter, G.; Giske, C.G.; Turnidge, J. Clinical and Laboratory Standards Institute (CLSI) and European Committee on Antimicrobial Susceptibility Testing (EUCAST) position statements on polymyxin B and colistin clinical breakpoints. Clin. Infect. Dis. 2020, 71, e523–e529. [Google Scholar] [CrossRef] [PubMed]
  168. Wang, H.; Wang, J.; Yu, P.; Ge, P.; Jiang, Y.; Xu, R.; Chen, R.; Liu, X. Identification of antibiotic resistance genes in the multidrug-resistant Acinetobacter baumannii strain, MDR-SHH02, using whole-genome sequencing. Int. J. Mol. Med. 2017, 39, 364–372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Chen, L.; Chavda, K.D.; Fraimow, H.S.; Mediavilla, J.R.; Melano, R.G.; Jacobs, M.R.; Bonomo, R.A.; Kreiswirth, B.N. Complete nucleotide sequences of blaKPC-4- and blaKPC-5-harboring IncN and IncX plasmids from Klebsiella pneumoniae strains isolated in New Jersey. Antimicrob Agents Chemother. 2013, 57, 269–276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Aishwarya, K.V.L.; Geetha, P.V.; Eswaran, S.; Mariappan, S.; Sekar, U. Spectrum of Aminoglycoside Modifying Enzymes in Gram-Negative Bacteria Causing Human Infections. J Lab Physicians. 2020, 12, 27–31. [Google Scholar] [CrossRef]
  171. Levings, R.S.; Partridge, S.R.; Lightfoot, D.; Hall, R.M.; Djordjevic, S.P. New integron-associated gene cassette encoding a 3-N-aminoglycoside acetyltransferase. Antimicrob Agents Chemother. 2005, 49, 1238–1241. [Google Scholar] [CrossRef] [Green Version]
  172. Zhang, Y.; Zhang, N.; Wang, M.; Luo, M.; Peng, Y.; Li, Z.; Xu, J.; Ou, M.; Kan, B.; Li, X.; et al. The prevalence and distribution of aminoglycoside resistance genes. Biosaf. Health 2023, 5, 14–20. [Google Scholar] [CrossRef]
  173. Mancini, S.; Marchesi, M.; Imkamp, F.; Wagner, K.; Keller, P.M.; Quiblier, C.; Bodendoerfer, E.; Courvalin, P.; Böttger, E.C. Population-based inference of aminoglycoside resistance mechanisms in Escherichia coli. EBioMedicine 2019, 46, 184–192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Yossapol, M.; Suzuki, K.; Odoi, J.O.; Sugiyama, M.; Usui, M.; Asai, T. Persistence of extended-spectrum β-lactamase plasmids among Enterobacteriaceae in commercial broiler farms. Microbiol. Immunol. 2020, 64, 712–718. [Google Scholar] [CrossRef] [PubMed]
  175. Ho, P.L.; Wong, R.C.; Lo, S.W.; Chow, K.H.; Wong, S.S.; Que, T.L. Genetic identity of aminoglycoside-resistance genes in Escherichia coli isolates from human and animal sources. J. Med. Microbiol. 2010, 59, 702–707. [Google Scholar] [CrossRef]
  176. Radisic, V.; Grevskott, D.H.; Lunestad, B.T.; Øvreås, L.; Marathe, N.P. Sewage-based surveillance shows presence of Klebsiella pneumoniae resistant against last resort antibiotics in the population in Bergen, Norway. Int. J. Hyg. Environ. Health. 2023, 248, 114075. [Google Scholar] [CrossRef]
  177. Dubois, V.; Arpin, C.; Dupart, V.; Scavelli, A.; Coulange, L.; Andre, C.; Fischer, I.; Grobost, F.; Brochet, J.P.; Lagrange, I.; et al. β-lactam and aminoglycoside resistance rates and mechanisms among Pseudomonas aeruginosa in French general practice (community and private healthcare centres). J. Antimicrob. Chemother. 2008, 62, 316–323. [Google Scholar] [CrossRef] [Green Version]
  178. Plattner, M.; Gysin, M.; Haldimann, K.; Becker, K.; Hobbie, S.N. Epidemiologic, Phenotypic, and Structural Characterization of Aminoglycoside-Resistance Gene aac(3)-IV. Int. J. Mol. Sci. 2020, 21, 6133. [Google Scholar] [CrossRef]
  179. Sekiguchi, J.-I.; Asagi, T.; Miyoshi-Akiyama, T.; Fujino, T.; Kobayashi, I.; Morita, K.; Kikuchi, Y.; Kuratsuji, T.; Kirikae, T. Multidrug-resistant Pseudomonas aeruginosa strain that caused an outbreak in a neurosurgery ward and its aac(6′)-Iae gene cassette encoding a novel aminoglycoside acetyltransferase. Antimicrob. Agents Chemother. 2005, 49, 3734–3742. [Google Scholar] [CrossRef] [Green Version]
  180. Kitao, T.; Miyoshi-Akiyama, T.; Kirikae, T. AAC(6′)-Iaf, a novel aminoglycoside 6′-N-acetyltransferase from multidrug-resistant Pseudomonas aeruginosa clinical isolates. Antimicrob. Agents Chemother. 2009, 53, 2327–2334. [Google Scholar] [CrossRef] [Green Version]
  181. Gutiérrez, O.; Juan, C.; Cercenado, E.; Navarro, F.; Bouza, E.; Coll, P.; Pérez, J.L.; Oliver, A. Molecular epidemiology and mechanisms of carbapenem resistance in Pseudomonas aeruginosa isolates from Spanish hospitals. Antimicrob. Agents Chemother. 2007, 51, 4329–4335. [Google Scholar] [CrossRef] [Green Version]
  182. Mendes, R.; Toleman, M.; Ribeiro, J.; Sader, H.; Jones, R.; Walsh, T. Integron carrying a novel metallo-b-lactamase gene, blaIMP-16, and a fused form of aminoglycoside-resistance gene aac(6′)-30/aac(6′)-Ib: Report from the SENTRY antimicrobial surveillance program. Antimicrob. Agents Chemother. 2004, 48, 4693–4702. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Mabilat, C.; Lourencao-Vital, J.; Goussard, S.; Courvalin, P. A new example of physical linkage between Tn1 and Tn21: The antibiotic multiple-resistance region of plasmid pCFF04 encoding extended-spectrum β-lactamase TEM-3. Mol. Gen. Genet. 1992, 235, 113–121. [Google Scholar] [CrossRef] [PubMed]
  184. Mugnier, P.; Casin, I.; Bouthors, A.T.; Collatz, E. Novel OXA-10-derived extended-spectrum β-lactamases selected in vivo or in vitro. Antimicrob. Agents Chemother. 1998, 42, 3113–3116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Kim, D.W.; Thawng, C.N.; Lee, K.; Cha, C.J. Revisiting Polymorphic Diversity of Aminoglycoside N-Acetyltransferase AAC(6′)-Ib Based on Bacterial Genomes of Human, Animal, and Environmental Origins. Front. Microbiol. 2018, 9, 1831. [Google Scholar] [CrossRef] [PubMed]
  186. Robicsek, A.; Strahilevitz, J.; Jacoby, G.A.; Macielag, M.; Abbanat, D.; Park, C.H.; Bush, K.; Hooper, D.C. Fluoroquinolone-modifying enzyme: A new adaptation of a common aminoglycoside acetyltransferase. Nat. Med. 2005, 12, 83–88. [Google Scholar] [CrossRef] [PubMed]
  187. Doi, Y.; Wachino, J.; Yamane, K.; Shibata, N.; Yagi, T.; Shibayama, K.; Kato, H.; Arakawa, Y. Spread of novel aminoglycoside resistance gene aac(6′)-Iad among Acinetobacter clinical isolates in Japan. Antimicrob. Agents Chemother. 2004, 48, 2075–2080. [Google Scholar] [CrossRef] [Green Version]
  188. Poirel, L.; Lambert, T.; Türkoglü, S.; Ronco, E.; Gaillard, J.; Nordmann, P. Characterization of Class 1 integrons from Pseudomonas aeruginosa that contain the bla(VIM-2) carbapenem-hydrolyzing b-lactamase gene and of two novel aminoglycoside resistance gene cassettes. Antimicrob. Agents Chemother. 2001, 45, 546–552. [Google Scholar] [CrossRef] [Green Version]
  189. Shaw, K.J.; Cramer, C.A.; Rizzo, M.; Mierzwa, R.; Gewain, K.; Miller, G.H.; Hare, R.S. Isolation, characterization, and DNA sequence analysis of an AAC(6′)-II gene from Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 1989, 33, 2052–2062. [Google Scholar] [CrossRef] [Green Version]
  190. Mokhtari, H.; Eslami, G.; Zandi, H.; Dehghan-Banadkouki, A.; Vakili, M. Evaluating the Frequency of aac(6′)-IIa, ant(2″)-I, intl1, and intl2 Genes in Aminoglycosides Resistant Klebsiella pneumoniae Isolates Obtained from Hospitalized Patients in Yazd, Iran. Avicenna J. Med. Biotechnol. 2018, 10, 115–119. [Google Scholar]
  191. Jacoby, G.A.; Blaser, M.J.; Santanam, P.; Hächler, H.; Kayser, F.H.; Hare, R.S.; Miller, G.H. Appearance of amikacin and tobramycin resistance due to 4′-aminoglycoside nucleotidyltransferase [ANT(4′)-II] in gram-negative pathogens. Antimicrob. Agents Chemother. 1990, 34, 2381–2386. [Google Scholar] [CrossRef] [Green Version]
  192. Sabtcheva, S.; Galimand, M.; Gerbaud, G.; Courvalin, P.; Lambert, T. Aminoglycoside resistance gene ant(4′)-IIb of Pseudomonas aeruginosa BM4492, a clinical isolate from Bulgaria. Antimicrob. Agents Chemother. 2003, 47, 1584–1588. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Rahman, M.; Prasad, K.N.; Pathak, A.; Pati, B.K.; Singh, A.; Ovejero, C.M.; Ahmad, S.; Gonzalez-Zorn, B. RmtC and RmtF 16S rRNA Methyltransferase in NDM-1-Producing Pseudomonas aeruginosa. Emerg Infect Dis. 2015, 21, 2059–2062. [Google Scholar] [CrossRef] [PubMed]
  194. Revilla, C.; Garcillán-Barcia, M.P.; Fernández-López, R.; Thomson, N.R.; Sanders, M.; Cheung, M.; Thomas, C.M.; de la Cruz, F. Different pathways to acquiring resistance genes illustrated by the recent evolution of IncW plasmids. Antimicrob. Agents Chemother. 2008, 52, 1472–1480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Yan, J.-J.; Hsueh, P.-R.; Lu, J.-J.; Chang, F.-Y.; Ko, W.-C.; Wu, J.-J. Characterization of acquired β-lactamases and their genetic support in multidrug-resistant Pseudomonas aeruginosa isolates in Taiwan: The prevalence of unusual integrons. J. Antimicrob. Chemother. 2006, 58, 530–536. [Google Scholar] [CrossRef] [PubMed]
  196. Fiett, J.; Baraniak, A.; Mrówka, A.; Fleischer, M.; Drulis-Kawa, Z.; Naumiuk, L.; Samet, A.; Hryniewicz, W.; Gniadkowski, M. Molecular epidemiology of acquired-metallo-β-lactamase-producing bacteria in Poland. Antimicrob. Agents Chemother. 2006, 50, 880–886. [Google Scholar] [CrossRef] [Green Version]
  197. Lee, K.Y.; Hopkins, J.D.; Syvanen, M. Direct involvement of IS26 in an antibiotic resistance operon. J. Bacteriol. 1990, 172, 3229–3236. [Google Scholar] [CrossRef] [Green Version]
  198. Woegerbauer, M.; Kuffner, M.; Domingues, S.; Nielsen, K.M. Involvement of aph(3′)-IIa in the formation of mosaic aminoglycoside resistance genes in natural environments. Front. Microbiol. 2015, 6, 442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Martin, P.; Jullien, E.; Courvalin, P. Nucleotide sequence of Acinetobacter baumannii aphA-6 gene: Evolutionary and functional implications of sequence homologies with nucleotide-binding proteins, kinases and other aminoglycoside-modifying enzymes. Mol. Microbiol. 1988, 2, 615–625. [Google Scholar] [CrossRef]
  200. Desvars-Larrive, A.; Ruppitsch, W.; Lepuschitz, S.; Szostak, M.P.; Spergser, J.; Feßler, A.T.; Schwarz, S.; Monecke, S.; Ehricht, R.; Walzer, C.; et al. Urban brown rats (Rattus norvegicus) as possible source of multidrug-resistant Enterobacteriaceae and meticillin-resistant Staphylococcus spp., Vienna, Austria, 2016 and 2017. Euro. Surveill. 2019, 32, 1900149. [Google Scholar] [CrossRef] [Green Version]
  201. Miller, E.A.; Ponder, J.B.; Willette, M.; Johnson, T.J.; VanderWaal, K.L. Merging Metagenomics and Spatial Epidemiology To Understand the Distribution of Antimicrobial Resistance Genes from Enterobacteriaceae in Wild Owls. Appl. Environ. Microbiol. 2020, 86, e00571-20. [Google Scholar] [CrossRef] [PubMed]
  202. Hu, X.; Xu, B.; Yang, Y.; Liu, D.; Yang, M.; Wang, J.; Shen, H.; Zhou, X.; Ma, X. A high throughput multiplex PCR assay for simultaneous detection of seven aminoglycoside-resistance genes in Enterobacteriaceae. BMC Microbiol. 2013, 13, 58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Doi, Y.; de Oliveira Garcia, D.; Adams, J.; Paterson, D.L. Coproduction of novel 16S rRNA methylase RmtD and metallo-beta-lactamase SPM-1 in a panresistant Pseudomonas aeruginosa isolate from Brazil. Antimicrob. Agents Chemother. 2007, 51, 852–856. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Francisco, G.R.; Nora, S.T.R.; Bueno, M.F.C.; da Silva Filho, L.V.R.F.; de Oliveira Garcia, D. Identification of Aminoglycoside-Resistant Pseudomonas aeruginosa Producing RmtG 16S rRNA Methyltransferase in a Cystic Fibrosis Patient. Antimicrob. Agents Chemother. 2015, 59, 2967–2968. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. El-Far, A.; Samir, S.; El-Gebaly, E.; Omar, M.; Dahroug, H.; El-Shenawy, A.; Soliman, N.S.; Gamal, D. High Rates of Aminoglycoside Methyltransferases Associated with Metallo-Beta-Lactamases in Multidrug-Resistant and Extensively Drug-Resistant Pseudomonas aeruginosa Clinical Isolates from a Tertiary Care Hospital in Egypt. Infect. Drug Resist. 2021, 14, 4849–4858. [Google Scholar] [CrossRef] [PubMed]
  206. Yamane, K.; Rossi, F.; Barberino, M.G.; Adams-Haduch, J.M.; Doi, Y.; Paterson, D.L. 16S ribosomal RNA methylase RmtD produced by Klebsiella pneumoniae in Brazil. J. Antimicrob Chemother. 2008, 61, 746–747. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Davis, M.A.; Baker, K.N.; Orfe, L.H.; Shah, D.H.; Besser, T.E.; Call, D.R. Discovery of a gene conferring multiple-aminoglycoside resistance in Escherichia coli. Antimicrob Agents Chemother. 2010, 54, 2666–2669. [Google Scholar] [CrossRef] [Green Version]
  208. Bueno, M.F.; Francisco, G.R.; O’Hara, J.A.; 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. [Google Scholar] [CrossRef] [Green Version]
  209. Grossman, T.H. Tetracycline antibiotics and resistance. Cold Spring Harbor Perspect. Med. 2016, 6, a025387. [Google Scholar] [CrossRef] [Green Version]
  210. Miranda, C.D.; Kehrenberg, C.; Ulep, C.; Schwarz, S.; Roberts, M.C. Diversity of tetracycline resistance genes in bacteria from Chilean salmon farms. Antimicrob. Agents Chemother. 2003, 47, 883–888. [Google Scholar] [CrossRef] [Green Version]
  211. Nguyen, F.; Starosta, A.L.; Arenz, S.; Sohmen, D.; Dönhöfer, A.; Wilson, D.N. Tetracycline antibiotics and resistance mechanisms. Biol. Chem. 2014, 395, 559–575. [Google Scholar] [CrossRef]
  212. Sheykhsaran, E.; Baghi, H.B.; Soroush, M.H.; Ghotaslou, R. An overview of tetracyclines and related resistance mechanisms. Rev. Med. Microbiol. 2019, 30, 69–75. [Google Scholar] [CrossRef]
  213. Yanat, B.; Rodríguez-Martínez, J.M.; Touati, A. Plasmid-mediated quinolone resistance in Enterobacteriaceae: A systematic review with a focus on Mediterranean countries. EJCMID 2017, 36, 421–435. [Google Scholar] [CrossRef] [PubMed]
  214. Anyanwu, M.U.; Jaja, I.F.; Nwobi, O.C. Occurrence and Characteristics of Mobile Colistin Resistance (mcr) Gene-Containing Isolates from the Environment: A Review. Int. J. Environ. Res. Public Health. 2020, 17, 1028. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Khuntayaporn, P.; Thirapanmethee, K.; Chomnawang, M.T. An Update of Mobile Colistin Resistance in Non-Fermentative Gram-Negative Bacilli. Front. Cell Infect. Microbiol. 2022, 12, 882236. [Google Scholar] [CrossRef]
  216. Karaiskos, I.; Giamarellou, H. Multidrug-resistant and extensively drug-resistant Gram-negative pathogens: Current and emerging therapeutic approaches. Expert. Opin. Pharmacother. 2014, 15, 1351–1370. [Google Scholar] [CrossRef]
  217. Arroyo, L.A.; Herrera, C.M.; Fernandez, L.; Hankins, J.V.; Trent, M.S.; Hancock, R.E. The pmrCAB operon mediates polymyxin resistance in Acinetobacter baumannii ATCC17978 and clinical isolates through phosphoethanolamine modification of lipid A. Antimicrob Agents Chemother. 2011, 55, 3743–3751. [Google Scholar] [CrossRef] [Green Version]
  218. Raetz, C.R.H.; Reynolds, C.M.; Trent, M.S.; Bishop, R.E. Lipid A modification systems in Gram-negative bacteria. Annu. Rev. Biochem. 2007, 76, 295–329. [Google Scholar] [CrossRef] [Green Version]
  219. Martins-Sorenson, N.; Snesrud, E.; Xavier, D.E.; Cacci, L.C.; Iavarone, A.T.; McGann, P.; Riley, L.W.; Moreira, B.M. A novel plasmid-encoded mcr-4.3 gene in a colistin-resistant Acinetobacter baumannii clinical strain. J. Antimicrob. Chemother. 2020, 75, 60–64. [Google Scholar] [CrossRef]
  220. Moffatt, J.H.; Harper, M.; Boyce, J.D. Mechanisms of Polymyxin Resistance. Adv. Exp. Med. Biol. 2019, 1145, 55–71. [Google Scholar] [CrossRef]
  221. Bojkovic, J.; Richie, D.L.; Six, D.A.; Rath, C.M.; Sawyer, W.S.; Hu, Q.; Dean, C.R. Characterization of an Acinetobacter baumannii lptD deletion strain: Permeability defects and response to inhibition of lipopolysaccharide and fatty acid biosynthesis. J. Bacteriol. 2016, 198, 731–741. [Google Scholar] [CrossRef] [Green Version]
  222. Luke, N.R.; Sauberan, S.L.; Russo, T.A.; Beanan, J.M.; Olson, R.; Loehfelm, T.W.; Cox, A.D.; Michael, F.; Vinogradov, E.V.; Campagnari, A.A. Identification and characterization of a glycosyltransferase involved in Acinetobacter baumannii lipopolysaccharide core biosynthesis. Infect. Immun. 2010, 78, 2017–2023. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Paul, D.; Mallick, S.; Das, S.; Saha, S.; Ghosh, A.K.; Mandal, S.M. Colistin induced assortment of antimicrobial resistance in a clinical isolate of Acinetobacter baumannii SD01. Infect. Disord. Drug Targets. 2019, 20, 501–505. [Google Scholar] [CrossRef]
  224. Popovic, M.; Steinort, D.; Pillai, S.; Joukhadar, C. Fosfomycin: An old, new friend? Eur. J. Clin. Microbiol. Infect. Dis. 2010, 29, 127–142. [Google Scholar] [CrossRef] [PubMed]
  225. Skȯld, O. Sulfonamide resistance: Mechanisms and trends. Drug Resist. Updates. 2000, 3, 155–160. [Google Scholar] [CrossRef] [PubMed]
  226. Sánchez-Osuna, M.; Cortés, P.; Llagostera, M.; Barbé, J.; Erill, I. Exploration into the origins and mobilization of di-hydrofolate reductase genes and the emergence of clinical resistance to trimethoprim. Microb. Genom. 2020, 6, mgen000440. [Google Scholar] [CrossRef] [PubMed]
  227. Partridge, S.R.; Kwong, S.M.; Firth, N.; Jensen, S.O. Mobile genetic elements associated with antimicrobial resistance. Clin. Microbiol. Rev. 2018, 31, e00088-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Khedkar, S.; Smyshlyaev, G.; Letunic, I.; Maistrenko, O.M.; Coelho, L.P.; Orakov, A.; Forslund, S.K.; Hildebrand, F.; Luetge, M.; Schmidt, T.S.B.; et al. Landscape of mobile genetic elements and their antibiotic resistance cargo in prokaryotic genomes. Nucleic Acids Res. 2022, 50, 3155–3168. [Google Scholar] [CrossRef] [PubMed]
  229. Fordham, S.M.E.; Mantzouratou, A.; Sheridan, E. Prevalence of insertion sequence elements in plasmids relating to mgrB gene disruption causing colistin resistance in Klebsiella pneumoniae. Microbiologyopen 2022, 11, e1262. [Google Scholar] [CrossRef]
  230. Lipszyc, A.; Szuplewska, M.; Bartosik, D. How Do Transposable Elements Activate Expression of Transcriptionally Silent Antibiotic Resistance Genes? Int. J. Mol. Sci. 2022, 23, 8063. [Google Scholar] [CrossRef]
  231. Babakhani, S.; Oloomi, M. Transposons: The agents of antibiotic resistance in bacteria. J. Basic Microbiol. 2018, 58, 905–917. [Google Scholar] [CrossRef]
  232. Mussi, M.A.; Limansky, A.S.; Viale, A.M. Acquisition of resistance to carbapenems in multidrug-resistant clinical strains of Acinetobacter baumannii: Natural insertional inactivation of a gene encoding a member of a novel family of beta-barrel outer membrane proteins. Antimicrob. Agents Chemother. 2005, 49, 1432–1440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Turton, J.F.; Ward, M.E.; Woodford, N.; Kaufmann, M.E.; Pike, R.; Livermore, D.M.; Pitt, T.L. The role of ISAba1 in expression of OXA carbapenemase genes in Acinetobacter baumannii. FEMS Microbiol. Lett. 2006, 258, 72–77. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Héritier, C.; Poirel, L.; Nordmann, P. Cephalosporinase over-expression resulting from insertion of ISAba1 in Acinetobacter baumannii. Clin. Microbiol. Infect. 2006, 12, 123–130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Villalón, P.; Valdezate, S.; Medina-Pascual, M.J.; Carrasco, G.; Vindel, A.; Saez-Nieto, J.A. Epidemiology of the Acinetobacter-derived cephalosporinase: Carbapenem-hydrolysing oxacillinase and metallo-β-lactamase genes, and of common insertion sequences, in epidemic clones of Acinetobacter baumannii from Spain. J. Antimicrob Chemother. 2013, 68, 550–553. [Google Scholar] [CrossRef] [Green Version]
  236. Fu, Y.; Jiang, J.; Zhou, H.; Jiang, Y.; Fu, Y.; Yu, Y.; Zhou, J. Characterization of a Novel Plasmid Type and Various Genetic Contexts of blaOXA-58 in Acinetobacter spp. from Multiple Cities in China. PLoS ONE 2014, 9, e84680. [Google Scholar] [CrossRef] [Green Version]
  237. Bogaerts, P.; Cuzon, G.; Naas, T.; Bauraing, C.; Deplano, A.; Lissoir, B.; Nordmann, P.; Glupczynski, Y. Carbapenem-resistant Acinetobacter baumannii isolates expressing the blaOXA-23 gene associated with ISAba4 in Belgium. Antimicrob Agents Chemother. 2008, 52, 4205–4206. [Google Scholar] [CrossRef] [Green Version]
  238. Corvec, S.; Poirel, L.; Naas, T.; Drugeon, H.; Nordmann, P. Genetics and expression of the carbapenem-hydrolyzing oxacillinase gene bla(OXA-23) in Acinetobacter baumannii. Antimicrob Agents Chemother. 2007, 51, 1530–1533. [Google Scholar] [CrossRef] [Green Version]
  239. Lee, Y.; Kim, C.K.; Lee, H.; Jeong, S.H.; Yong, D.; Lee, K. A novel insertion sequence: ISAba10, inserted into ISAba1 adjacent to the bla(OXA-23) gene and disrupting the outer membrane protein gene carO in Acinetobacter baumannii. Antimicrob Agents Chemother 2011, 55, 361–363. [Google Scholar] [CrossRef] [Green Version]
  240. Ruiz-Martínez, L.; López-Jiménez, L.; d’Ostuni, V.; Fusté, E.; Vinuesa, T.; Viñas, M. A mechanism of carbapenem resistance due to a new insertion element (ISPa133) in Pseudomonas aeruginosa. Int. Microbiol. 2011, 14, 51–58. [Google Scholar] [CrossRef] [Green Version]
  241. Al-Bayssari, C.; Valentini, C.; Gomez, C.; Reynaud-Gaubert, M.; Rolain, J.M. First detection of insertion sequence element ISPa1328 in the oprD porin gene of an imipenem-resistant Pseudomonas aeruginosa isolate from an idiopathic pulmonary fibrosis patient in Marseille, France. New Microbes New Infect. 2015, 7, 26–27. [Google Scholar] [CrossRef] [Green Version]
  242. Diene, S.M.; L’Homme, T.; Bellulo, S.; Stremler, N.; Dubus, J.C.; Mely, L.; Leroy, S.; Degand, N.; Rolain, J.-M. ISPa46, a novel insertion sequence in the oprD porin gene of an imipenem-resistant Pseudomonas aeruginosa isolate from a cystic fibrosis patient in Marseille, France. Int. J. Antimicrob. Agents 2013, 42, 268–271. [Google Scholar] [CrossRef] [PubMed]
  243. Wolter, D.J.; Hanson, N.D.; Lister, P.D. Insertional inactivation of oprD in clinical isolates of Pseudomonas aeruginosa leading to carbapenem resistance. FEMS Microbiol. Lett. 2004, 236, 137–143. [Google Scholar] [CrossRef] [PubMed]
  244. Evans, J.C.; Segal, H. A novel insertion sequence, ISPa26, in oprD of Pseudomonas aeruginosa is associated with carbapenem resistance. Antimicrob Agents Chemother. 2007, 51, 3776–3777. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Fowler, R.C.; Hanson, N.D. Emergence of carbapenem resistance due to the novel insertion sequence ISPa8 in Pseudomonas aeruginosa. PLoS ONE 2014, 9, e91299. [Google Scholar] [CrossRef]
  246. Al Bayssari, C.; Diene, S.M.; Loucif, L.; Gupta, S.K.; Dabboussi, F.; Mallat, H.; Hamze, M.; Rolain, J.-M. Emergence of VIM-2 and IMP-15 carbapenemases and inactivation of oprD gene in carbapenem-resistant Pseudomonas aeruginosa clinical isolates from Lebanon. Antimicrob Agents Chemother. 2014, 58, 4966–4970. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Hernandez-Alles, S.; Benedi, V.J.; Martinez-Martinez, L.; Pascual, A.; Aguilar, A.; Tomas, J.M.; Alberti, S. Development of resistance during antimicrobial therapy caused by insertion sequence interruption of porin genes. Antimicrob Agents Chemother. 1999, 43, 937–939. [Google Scholar] [CrossRef] [Green Version]
  248. Moffatt, J.H.; Harper, M.; Adler, B.; Nation, R.L.; Li, J.; Boyce, J.D. Insertion sequence ISAba11 is involved in colistin resistance and loss of lipopolysaccharide in Acinetobacter baumannii. Antimicrob Agents Chemother. 2011, 55, 3022–3024. [Google Scholar] [CrossRef] [Green Version]
  249. Cannatelli, A.; Giani, T.; D’Andrea, M.M.; Di Pilato, V.; Arena, F.; Conte, V.; Tryfinopoulou, K.; Vatopoulos, A.; Rossolini, G.M. MgrB inactivation is a common mechanism of colistin resistance in KPC-producing Klebsiella pneumoniae of clinical origin. Antimicrob Agents Chemother. 2014, 58, 5696–5703. [Google Scholar] [CrossRef] [Green Version]
  250. Poirel, L.; Jayol, A.; Bontron, S.; Villegas, M.V.; Ozdamar, M.; Turkoglu, S.; Nordmann, P. The mgrB gene as a key target for acquired resistance to colistin in Klebsiella pneumoniae. J. Antimicrob. Chemother. 2015, 70, 75–80. [Google Scholar] [CrossRef]
  251. Varani, A.; He, S.; Siguier, P.; Ross, K.; Chandler, M. The IS6 family, a clinically important group of insertion sequences including IS26. Mob DNA 2021, 12, 11. [Google Scholar] [CrossRef]
  252. Roberts, A.P.; Chandler, M.; Courvalin, P.; Guédon, G.; Mullany, P.; Pembroke, T.; Rood, J.I.; Smith, C.J.; Summers, A.O.; Tsuda, M.; et al. Revised nomenclature for transposable genetic elements. Plasmid 2008, 60, 167–173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Reznikoff, W.S. Tn5 as a model for understanding DNA transposition. Mol. Microbiol. 2003, 47, 1199–1206. [Google Scholar] [CrossRef] [PubMed]
  254. Iyer, A.; Barbour, E.; Azhar, E.; El Salabi, A.A.; Hassan, H.M.A.; Qadri, I.; Chaudhary, A.; Abuzenadah, A.; Kumosani, T.; Damanhouri, G.; et al. Transposable elements in Escherichia coli antimicrobial resistance. Adv. Biosci. Biotechnol. 2013, 4, 415–423. [Google Scholar] [CrossRef] [Green Version]
  255. Domingues, S.; da Silva, G.J.; Nielsen, K.M. Global dissemination patterns of common gene cassette arrays in class 1 integrons. Microbiology 2015, 161, 1313–1337. [Google Scholar] [CrossRef]
  256. Liu, C.-C.; Tang, C.Y.; Chang, K.-C.; Kuo, H.-Y.; Liou, M.-L. A comparative study of class 1 integrons in Acinetobacter baumannii. Gene 2014, 544, 75–82. [Google Scholar] [CrossRef]
  257. Akrami, F.; Shahandashti, E.F.; Yahyapour, Y.; Sadeghi, M.; Khafri, S.; Pournajaf, A.; Rajabnia, R. Integron types, genecassettes and antimicrobial resistance profile of Acinetobacter baumannii isolated from BAL samples in Babol, North of Iran. Microb. Pathog. 2017, 109, 35–38. [Google Scholar] [CrossRef]
  258. Ghaly, T.M.; Geoghegan, J.L.; Tetu, S.G.; Gillings, M.R. The peril and promise of integrons: Beyond antibiotic resistance. Trends Microbiol. 2019, 28, 455–464. [Google Scholar] [CrossRef]
  259. Wu, Y.-W.; Doak, T.G.; Ye, Y. The gain and loss of chromosomal integron systems in the treponema species. BMC Evol. Biol. 2013, 13, 16–24. [Google Scholar] [CrossRef] [Green Version]
  260. Kargar, M.; Mohammadalipour, Z.; Doosti, A.; Lorzadeh, S.; Japoni-Nejad, A. High prevalence of class 1 to 3 integrons among multidrug-resistant diarrheagenic Escherichia coli in Southwest of Iran. Osong Public Health Res. Perspect. 2014, 5, 193–198. [Google Scholar] [CrossRef] [Green Version]
  261. Soufi, L.; Abbassi, M.S.; Sáenz, Y.; Vinué, L.; Somalo, S.; Zarazaga, M.; Abbas, A.; Dbaya, R.; Khanfir, L.; Ben Hassen, A.; et al. Prevalence and diversity of integrons and associated resistance genes in Escherichia coli isolates from poultry meat in Tunisia. Foodborne Pathog Dis. 2009, 6, 1067–1073. [Google Scholar] [CrossRef]
  262. Su, H.-C.; Ying, G.-G.; Tao, R.; Zhang, R.-Q.; Zhao, J.-L.; Liu, Y.-S. Class 1 and 2 integrons, sul resistance genes and antibiotic resistance in Escherichia coli isolated from Dongjiang River, South China. Environ. Pollut. 2012, 169, 42–49. [Google Scholar] [CrossRef] [PubMed]
  263. Juhas, M.; van der Meer, J.R.; Gaillard, M.; Harding, R.M.; Hood, D.W.; Crook, D.W. Genomic islands: Tools of bacterial horizontal gene transfer and evolution. FEMS Microbiol. Rev. 2009, 33, 376–393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Pagano, M.; Martins, A.F.; Barth, A.L. Mobile genetic elements related to carbapenem resistance in Acinetobacter baumannii. Braz J. Microbiol. 2016, 47, 785–792. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Moglad, E.; Alanazi, N.; Altayb, H.N. Genomic Study of Chromosomally and Plasmid-Mediated Multidrug Resistance and Virulence Determinants in Klebsiella Pneumoniae Isolates Obtained from a Tertiary Hospital in Al-Kharj, KSA. Antibiotics 2022, 11, 1564. [Google Scholar] [CrossRef] [PubMed]
  266. von Wintersdorff, C.J.H.; Penders, J.; Van Niekerk, J.M.; Mills, N.D.; Majumder, S.; Van Alphen, L.B.; Savelkoul, P.H.M.; Wolffs, P.F.G. Dissemination of Antimicrobial Resistance in Microbial Ecosystems through Horizontal Gene Transfer. Front. Microbiol. 2016, 7, 173. [Google Scholar] [CrossRef] [Green Version]
  267. Redondo-Salvo, S.; Fernández-López, R.; Ruiz, R.; Vielva, L.; DE Toro, M.; Rocha, E.P.C.; Garcillán-Barcia, M.P.; De La Cruz, F. Pathways for horizontal gene transfer in bacteria revealed by a global map of their plasmids. Nat. Commun. 2020, 11, 3602. [Google Scholar] [CrossRef]
  268. Bertini, A.; Poirel, L.; Mugnier, P.D.; Villa, L.; Nordmann, P.; Carattoli, A. Characterization and PCR-based replicon typing of resistance plasmids in Acinetobacter baumannii. Antimicrob Agents Chemother. 2010, 54, 4168–4177. [Google Scholar] [CrossRef] [Green Version]
  269. Lean, S.S.; Yeo, C.C. Small, enigmatic plasmids of the nosocomial pathogen, Acinetobacter baumannii: Good, bad, who knows? Front. Microbiol. 2017, 8, 1547. [Google Scholar] [CrossRef] [Green Version]
  270. Shintani, M.; Sanchez, Z.K.; Kimbara, K. Genomics of microbial plasmids: Classification and identification based on replication and transfer systems and host taxonomy. Front. Microbiol. 2015, 6, 242. [Google Scholar] [CrossRef] [Green Version]
  271. Botelho, J.; Grosso, F.; Quinteira, S.; Mabrouk, A.; Peixe, L. The complete nucleotide sequence of an IncP-2 megaplasmid unveils a mosaic architecture comprising a putative novel blaVIM-2-harbouring transposon in Pseudomonas aeruginosa. J. Antimicrob. Chemother. 2017, 72, 2225–2229. [Google Scholar] [CrossRef] [Green Version]
  272. Liu, J.; Yang, L.; Chen, D.; Peters, B.M.; Li, L.; Li, B.; Xu, Z.; Shirtliff, M.E. Complete sequence of pBM413, a novel multidrug resistance megaplasmid carrying qnrVC6 and blaIMP-45 from Pseudomonas aeruginosa. Int. J. Antimicrob Agents. 2018, 51, 145–150. [Google Scholar] [CrossRef] [PubMed]
  273. Compain, F.; Poisson, A.; Le Hello, S.; Branger, C.; Weill, F.-X.; Arlet, G.; Decré, D. Targeting relaxase genes for classification of the predominant plasmids in Enterobacteriaceae. Int. J. Med. Microbiol. 2014, 304, 236–242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  274. Liu, P.Y.; Lee, Y.L.; Lu, M.C.; Shao, P.L.; Lu, P.L.; Chen, Y.H.; Cheng, S.-H.; Ko, W.-C.; Lin, C.-Y.; Wu, T.-S.; et al. National surveillance of antimicrobial susceptibility of bacteremic Gram-negative bacteria with emphasis on community-acquired resistant isolates: Report from the 2019 Surveillance of Multicenter Antimicrobial Resistance in Taiwan (SMART). Antimicrob. Agents Chemother. 2020, 64, e01089-20. [Google Scholar] [CrossRef]
  275. Lima, W.G.; Brito, J.C.M.; da Cruz Nizer, W.S. Ventilator-associated pneumonia (VAP) caused by carbapenem-resistant Acinetobacter baumannii in patients with COVID-19: Two problems, one solution? Med. Hypotheses. 2020, 144, 110139. [Google Scholar] [CrossRef] [PubMed]
  276. Blair, J.M.A.; Webber, M.A.; Baylay, A.J.; Ogbolu, D.O.; Piddock, L.J.V. Molecular mechanisms of antibiotic resistance. Nat. Rev. Microbiol. 2015, 13, 42–51. [Google Scholar] [CrossRef] [PubMed]
  277. Manchanda, V.; Sanchaita, S.; Singh, N.P. Multidrug Resistant Acinetobacter. J. Glob. Infect. Dis. 2010, 2, 291–304. [Google Scholar] [CrossRef]
  278. Horcajada, J.P.; Montero, M.; Oliver, A.; Sorlí, L.; Luque, S.; Gómez-Zorrilla, S.; Benito, N.; Grau, S. Epidemiology and treatment of multidrug-resistant and extensively drug-resistant Pseudomonas aeruginosa infections. Clin. Microbiol. Rev. 2019, 32, e00031-19. [Google Scholar] [CrossRef]
  279. Weiner-Lastinger, L.M.; Abner, S.; Edwards, J.R.; Kallen, A.J.; Karlsson, M.; Magill, S.S.; Pollock, D.; See, I.; Soe, M.M.; Walters, M.S.; et al. Antimicrobial-resistant pathogens associated with adult healthcare-associated infections: Summary of data reported to the National Healthcare Safety Network, 2015–2017. Infect. Control Hosp. Epidemiol. 2020, 41, 1–18. [Google Scholar] [CrossRef] [Green Version]
  280. Tenover, F.C.; Nicolau, D.P.; Gill, C.M. Carbapenemase-producing Pseudomonas aeruginosa—An emerging challenge. Emerg. Microbes. Infect. 2022, 11, 811–814. [Google Scholar] [CrossRef]
  281. Wang, M.-G.; Liu, Z.-Y.; Liao, X.-P.; Sun, R.-Y.; Li, R.-B.; Liu, Y.; Fang, L.-X.; Sun, J.; Liu, Y.-H.; Zhang, R.-M. Retrospective data insight into the global distribution of carbapenemase-producing Pseudomonas aeruginosa. Antibiotics 2021, 10, 548. [Google Scholar] [CrossRef]
  282. Adeolu, M.; Alnajar, S.; Naushad, S.; Gupta, R. Genome-based phylogeny and taxonomy of the Enterobacteriales: Proposal for Enterobacterales ord. nov. divided into the families Enterobacteriaceae, Erwiniaceae fam. nov., Pectobacteriaceae fam. nov., Yersiniaceae fam. nov., Hafniaceae fam. nov., Morganellaceae fam. nov., and Budviciaceae fam. nov. Int. J. Syst. Evol. 2016, 66, 5575–5599. [Google Scholar]
  283. Lynch, J.P., 3rd; Clark, N.M.; Zhanel, G.G. Escalating antimicrobial resistance among Enterobacteriaceae: Focus on carbapenemases. Expert Opin Pharmacother. 2021, 22, 1455–1473. [Google Scholar] [CrossRef] [PubMed]
  284. Iredell, J.; Brown, J.; Tagg, K. Antibiotic resistance in Enterobacteriaceae: Mechanisms and clinical implications. BMJ 2016, 352, h6420. [Google Scholar] [CrossRef] [PubMed]
  285. Martin, R.M.; Bachman, M.A. Colonization, Infection, and the Accessory Genome of Klebsiella pneumoniae. Front. Cell Infect. Microbiol. 2018, 8, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  286. Effah, C.Y.; Sun, T.; Liu, S.; Wu, Y. Klebsiella pneumoniae: An increasing threat to public health. Ann. Clin. Microbiol. Antimicrob. 2020, 19, 1. [Google Scholar] [CrossRef]
  287. Navon-Venezia, S.; Kondratyeva, K.; Carattoli, A. Klebsiella pneumoniae: A major worldwide source and shuttle for antibiotic resistance. FEMS Microbiol. Rev. 2017, 41, 252–275. [Google Scholar] [CrossRef]
  288. Mathers, A.J.; Peirano, G.; Pitout, J.D. The role of epidemic resistance plasmids and international high-risk clones in the spread of multidrug-resistant Enterobacteriaceae. Clin. Microbiol. Rev. 2015, 28, 565–591. [Google Scholar] [CrossRef] [Green Version]
  289. Hennequin, C.; Robin, F. Correlation between antimicrobial resistance and virulence in Klebsiella pneumoniae. Eur. J. Clin. Microbiol. Infect. Dis. 2015, 35, 333–341. [Google Scholar] [CrossRef]
  290. Tietgen, M.; Sedlaczek, L.; Higgins, P.G.; Kaspar, H.; Ewers, C.; Göttig, S. Colistin Resistance Mechanisms in Human and Veterinary Klebsiella pneumoniae Isolates. Antibiotics 2022, 11, 1672. [Google Scholar] [CrossRef]
  291. Longo, L.G.A.; de Sousa, V.S.; Kraychete, G.B.; Justo-da-Silva, L.H.; Rocha, J.A.; Superti, S.V.; Bonelli, R.R.; Martins, I.S.; Moreira, B.M. Colistin resistance emerges in pandrug-resistant Klebsiella pneumoniae epidemic clones in Rio de Janeiro, Brazil. Int. J. Antimicrob Agents. 2019, 54, 579–586. [Google Scholar] [CrossRef]
  292. Lalaoui, R.; Bakour, S.; Livnat, K.; Assous, M.V.; Diene, S.M.; Rolain, J.-M. Spread of Carbapenem and Colistin-Resistant Klebsiella pneumoniae ST512 Clinical Isolates in Israel: A Cause for Vigilance. Microb. Drug Resist. 2019, 25, 63–71. [Google Scholar] [CrossRef] [PubMed]
  293. Azam, M.; Gaind, R.; Yadav, G.; Sharma, A.; Upmanyu, K.; Jain, M.; Singh, R. Colistin Resistance Among Multiple Sequence Types of Klebsiella pneumoniae Is Associated With Diverse Resistance Mechanisms: A Report From India. Front. Microbiol. 2021, 12, 609840. [Google Scholar] [CrossRef] [PubMed]
  294. Davin-Regli, A.; Lavigne, J.P.; Pagès, J.M. Enterobacter spp.: Update on Taxonomy, Clinical Aspects, and Emerging Antimicrobial Resistance. Clin. Microbiol. Rev. 2019, 32, e00002-19. [Google Scholar] [CrossRef] [PubMed]
  295. Annavajhala, M.K.; Gomez-Simmonds, A.; Uhlemann, A.-C. Multidrug-Resistant Enterobacter cloacae Complex Emerging as a Global, Diversifying Threat. Front. Microbiol. 2019, 10, 44. [Google Scholar] [CrossRef] [Green Version]
  296. Poirel, L.; Madec, J.Y.; Lupo, A.; Schink, A.K.; Kieffer, N.; Nordmann, P.; Schwarz, S. Antimicrobial Resistance in Escherichia coli. Microbiol. Spectr. 2018, 6, 4–6. [Google Scholar] [CrossRef] [Green Version]
  297. Galindo-Méndez, M. Antimicrobial Resistance in Escherichia Coli. E. Coli Infections—Importance of Early Diagnosis and Efficient Treatment; IntechOpen: London, UK, 2020; 234p. [Google Scholar] [CrossRef]
  298. MacKinnon, M.C.; Sargeant, J.M.; Pearl, D.L.; Reid-Smith, R.J.; Carson, C.A.; Parmley, E.J.; McEwen, S.A. Evaluation of the health and healthcare system burden due to antimicrobial-resistant Escherichia coli infections in humans: A systematic review and meta-analysis. Antimicrob Resist. Infect. Control 2020, 9, 200. [Google Scholar] [CrossRef]
  299. Liu, L.; Zhang, L.; Zhou, H.; Yuan, M.; Hu, D.; Wang, Y.; Sun, H.; Xu, J.; Lan, R. Antimicrobial Resistance and Molecular Characterization of Citrobacter spp. Causing Extraintestinal Infections. Front. Cell Infect. Microbiol. 2021, 11, 737636. [Google Scholar] [CrossRef]
  300. Skurnik, D.; Nucci, A.; Ruimy, R.; Lasocki, S.; Muller-Serieys, C.; Montravers, P.; Andremont, A.; Courvalin, P. Emergence of carbapenem-resistant Hafnia: The fall of the last soldier. Clin. Infect. Dis. 2010, 50, 1429–1431. [Google Scholar] [CrossRef] [Green Version]
  301. Liu, H.; Zhu, J.; Hu, Q.; Rao, X. Morganella morganii, a non-negligent opportunistic pathogen. Int. J. Infect. Dis. 2016, 50, 10–17. [Google Scholar] [CrossRef] [Green Version]
  302. Yuan, C.; Wei, Y.; Zhang, S.; Cheng, J.; Cheng, X.; Qian, C.; Wang, Y.; Zhang, Y.; Yin, Z.; Chen, H. Comparative Genomic Analysis Reveals Genetic Mechanisms of the Variety of Pathogenicity, Antibiotic Resistance, and Environmental Adaptation of Providencia Genus. Front. Microbiol. 2020, 11, 572642. [Google Scholar] [CrossRef]
  303. Guan, J.; Bao, C.; Wang, P.; Jing, Y.; Wang, L.; Li, X.; Mu, X.; Li, B.; Zhou, D.; Guo, X.; et al. Genetic Characterization of Four Groups of Chromosome-Borne Accessory Genetic Elements Carrying Drug Resistance Genes in Providencia. Infect. Drug Resist. 2022, 15, 2253–2270. [Google Scholar] [CrossRef] [PubMed]
  304. Tavares-Carreon, F.; De Anda-Mora, K.; Rojas-Barrera, I.C.; Andrade, A. Serratia marcescens antibiotic resistance mechanisms of an opportunistic pathogen: A literature review. PeerJ 2023, 11, e14399. [Google Scholar] [CrossRef] [PubMed]
  305. Sandner-Miranda, L.; Vinuesa, P.; Cravioto, A.; Morales-Espinosa, R. The Genomic Basis of Intrinsic and Acquired Antibiotic Resistance in the Genus Serratia. Front. Microbiol. 2018, 9, 828. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  306. Takaya, A.; Yamamoto, T.; Tokoyoda, K. Humoral Immunity vs. Salmonella. Front. Immunol. 2020, 10, 3155. [Google Scholar] [CrossRef] [PubMed]
  307. Buckle, G.C.; Walker, C.L.; Black, R.E. Typhoid fever and paratyphoid fever: Systematic review to estimate global morbidity and mortality for 2010. J. Glob Health 2012, 2, 10401. [Google Scholar] [CrossRef]
  308. Crump, J.A.; Luby, S.P.; Mintz, E.D. The global burden of typhoid fever. Bull. World Health Organ. 2004, 82, 346–353. [Google Scholar]
  309. Majowicz, S.E.; Musto, J.; Scallan, E.; Angulo, F.J.; Kirk, M.; O’Brien, S.J.; Jones, T.F.; Fazil, A.; Hoekstra, R.M. International Collaboration on Enteric Disease ‘Burden of Illness’ Studies. The global burden of nontyphoidal Salmonella gastroenteritis. Clin. Infect. Dis. 2010, 50, 882–889. [Google Scholar] [CrossRef] [Green Version]
  310. Galan, J.E. Salmonella Typhimurium and inflammation: A pathogen-centric affair. Nat. Rev. Microbiol. 2021, 19, 716–725. [Google Scholar] [CrossRef]
  311. Carattoli, A. Plasmids and the spread of resistance. Int. J. Med. Microbiol. 2013, 303, 298–304. [Google Scholar] [CrossRef]
  312. Lobato-Márquez, D.; Molina-García, L.; Moreno-Córdoba, I.; García-del Portillo, F.; Díaz-Orejas, R. Stabilization of the virulence plasmid pSLT of Salmonella Typhimurium by three maintenance systems and its evaluation by using a new stability test. Front. Mol. Biosci. 2016, 3, 66. [Google Scholar] [CrossRef] [Green Version]
  313. Klemm, E.J.; Shakoor, S.; Page, A.J.; Qamar, F.N.; Judge, K.; Saeed, D.K.; Wong, V.K.; Dallman, T.J.; Nair, S.; Baker, S.; et al. Emergence of an Extensively Drug-Resistant Salmonella enterica Serovar Typhi clone harboring a promiscuous plasmid encoding resistance to fluoroquinolones and third-generation cephalosporins. MBio 2018, 9, e00105-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  314. Al-Gallas, N.; Belghouthi, K.; Barratt, N.A.; Ghedira, K.; Hotzel, H.; Tomaso, H.; El-Adawy, H.; Neubauer, H.; Laouini, D.; Zarrouk, S.; et al. Identification and characterization of multidrug-resistant ESBL-producing Salmonella enterica serovars Kentucky and Typhimurium isolated in Tunisia CTX-M-61/TEM-34, a novel cefotaxime-hydrolysing β-lactamase of Salmonella. J. Appl. Microbiol. 2022, 132, 279–289. [Google Scholar] [CrossRef] [PubMed]
  315. Riyaaz, A.A.A.; Perera, V.; Sivakumaran, S.; de Silva, N. Typhoid fever due to extended spectrum β-lactamase-producing Salmonella enterica serovar Typhi: A case report and literature review. Case Rep. Infect. Dis. 2018, 2018, 4610246. [Google Scholar]
  316. Kim, C.; Latif, I.; Neupane, D.P.; Lee, G.Y.; Kwon, R.S.; Batool, A.; Ahmed, Q.; Qamar, M.U.; Song, J. The molecular basis of extensively drug-resistant Salmonella Typhi isolates from pediatric septicemia patients. PLoS ONE 2021, 16, e0257744. [Google Scholar] [CrossRef] [PubMed]
  317. Radu-Popescu, M.A.; Dumitriu, S.; Enache-Soare, S.; Bancescu, G.; Udristoiu, A.; Cojocaru, M.; Vagu, C. Phenotypic and genotypic characterization of antibiotic resistance patterns in Acinetobacter baumannii strains isolated in a Romanian hospital. Farmacia 2010, 58, 362–367. [Google Scholar]
  318. Mereuţă, A.I.; Docquier, J.D.; Rossolini, G.M.; Buiuc, D. Detection of metallo-beta-lactamases in Gram-negative bacilli isolated in hospitals from Romania-research fellowship report. Bacteriol. Virusol. Parazitol. Epidemiol. 2007, 52, 45–49. [Google Scholar]
  319. Mereuţă, A.I.; Bădescu, A.C.; Dorneanu, O.S.; Iancu, L.S.; Tuchiluş, C.G. Spread of VIM-2 metallo-beta-lactamase in Pseudomonas aeruginosa and Acinetobacter baumannii clinical isolates from Iaşi, Romania. Rev. Romana Med. Lab. 2013, 21, 423–430. [Google Scholar] [CrossRef] [Green Version]
  320. European Centre for Disease Prevention and Control. Available online: https://atlas.ecdc.europa.eu/public/index.aspx. (accessed on 22 March 2023).
  321. Naas, T.; Nordmann, P.; Heidt, A. Intercountry transfer of PER-1 extended-spectrum beta-lactamase-producing Acinetobacter baumannii from Romania. Int. J. Antimicrob. Agents. 2007, 29, 226–228. [Google Scholar] [CrossRef]
  322. Bonnin, R.A.; Poirel, L.; Licker, M.; Nordmann, P. Genetic diversity of carbapenem-hydrolysing β-lactamases in Acinetobacter baumannii from Romanian hospitals. Clin. Microbiol. Infect. 2011, 17, 1524–1528. [Google Scholar] [CrossRef] [Green Version]
  323. Timofte, D.; Panzaru, C.V.; Maciuca, I.E.; Dan, M.; Mare, A.D.; Man, A.; Toma, F. Active surveillance scheme in three Romanian hospitals reveals a high prevalence and variety of carbapenamase-producing Gram-negative bacteria: A pilot study, December 2014 to May 2015. Eurosurveillance 2016, 21, 30262. [Google Scholar] [CrossRef] [Green Version]
  324. Gheorghe, I.; Novais, A.; Grosso, F.; Rodrigues, C.; Chifiriuc, M.C.; Lazar, V.; Peixe, L. 2015, Snapshot on carbapenemase-producing Pseudomonas aeruginosa and Acinetobacter baumannii in Bucharest hospitals reveals unusual clones and novel genetic surroundings for blaOXA-23. J. Antimicrob. Chemother. 2015, 70, 1016–1020. [Google Scholar] [CrossRef] [Green Version]
  325. Czobor, I.; Gheorghe, I.; Banu, O.; Velican, A.; Lazăr, V.; Mihăescu, G.; Chifiriuc, M. ESBL genes in Multi Drug Resistant Gram-negative strains isolated in a one year survey from an Intensive Care Unit in Bucharest, Romania. Rom. Biotechnol. Lett. 2014, 19, 9553. [Google Scholar]
  326. Gheorghe, I.; Czobor, I.; Chifiriuc, M.C.; Borcan, E.; Ghiţă, C.; Banu, O.; Lazăr, V.; Mihăescu, G.; Mihăilescu, D.F.; Zhiyong, Z. Molecular screening of carbapenemase-producing Gram-negative strains in Romanian intensive care units during a one year survey. J. Med. Microbiol. 2014, 63, 1303–1310. [Google Scholar] [CrossRef]
  327. Saviuc, C.; Gheorghe, I.; Coban, S.; Drumea, V.; Chifiriuc, M.C.; Banu, O.; Bezirtzoglou, E.; Lazăr, V. Rosmarinus officinalis essential oil and eucalyptol act as efflux pumps inhibitors and increase ciprofloxacin efficiency against Pseudomonas aeruginosa and Acinetobacter baumannii MDR Strains. Rom. Biotechnol. Lett. 2016, 21, 11782–11790. [Google Scholar]
  328. Georgescu, M.; Gheorghe, I.; Dudu, A.; Czobor, I.; Costache, M.; Cristea, V.-C.; Lazăr, V.; Chifiriuc, M.C. First report of OXA-72 producing Acinetobacter baumannii in Romania. New Microbes New Infect. 2016, 13, 87–88. [Google Scholar] [CrossRef] [Green Version]
  329. Mihai, D.C.; Chifiriuc, M.C.; Gheorghe, I.; Diţu, L.M.; Mihăescu, G. Resistance and virulence gene profiles in Klebsiella pneumoniae and Acinetobacter baumannii strains isolated from hospitalized patients. In Proceedings of the 18th International Multidisciplinary Scientific GeoConference SGEM, Sofia, Bulgaria, 3–6 December 2018; SGEM: Sofia, Bulgaria, 2018; pp. 265–274. [Google Scholar] [CrossRef]
  330. Pirii, L.E.; Friedrich, A.W.; Rossen, J.; Vogels, W.; Beerthuizen, G.; Nieuwenhuis, M.K.; Kooistra-Smid, A.M.D.; Bathoorn, E. Extensive colonization with carbapenemase-producing microorganisms in Romanian burn patients: Infectious consequences from the Colectiv fire disaster. Eur. J. Clin. Microbiol. Infect. Dis. 2018, 37, 175–183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  331. Al Shaikhli, N.; Al Ali, O.; Chifiriuc, M.C.; Mihaescu, G.; Alshwaikh, R.M.A.; Mohsin, S.; Mohammed, H.B.; Pircalabioru, G.G.; Gheorghe, I. Antibiotic resistance determinants of Acinetobacter baumannii strains isolated from nosocomial infections. Rom. Biotechnol. Lett. 2020, 25, 1658–1665. [Google Scholar] [CrossRef]
  332. Yusuf, E.; Tompa, M.; Strepis, N.; Klaassen, C.H.W.; Goessens, W.H.F. High Prevalence of ST502 Carrying an OXA-24 Carbapenemase gene in Carbapenem-Nonsusceptible Acinetobacter baumannii-calcoaceticus Isolates in Romania. Microb. Drug Resist. 2022, 28, 636–644. [Google Scholar] [CrossRef] [PubMed]
  333. Gheorghe, I.; Cristea, V.C.; Marutescu, L.; Popa, M.; Murariu, C.; Trusca, B.S.; Borcan, E.; Ghita, M.C.; Lazar, V.; Chifiriuc, M.C. Resistance and virulence features in carbapenem-resistant Acinetobacter baumannii community acquired and nosocomial isolates in Romania. Rev. Chim. 2019, 70, 3502–3507. [Google Scholar] [CrossRef]
  334. Gheorghe-Barbu, I.; Barbu, I.C.; Popa, L.I.; Pîrcălăbioru, G.G.; Popa, M.; Măruțescu, L.; Niță-Lazar, M.; Banciu, A.; Stoica, C.; Gheorghe, I.; et al. Temporo-spatial variations in resistance determinants and clonality of Acinetobacter baumannii and Pseudomonas aeruginosa strains from Romanian hospitals and wastewaters. Antimicrob Resist. Infect. Control 2022, 11, 115. [Google Scholar] [CrossRef] [PubMed]
  335. Craciunas, C.; Butiuc-Keul, A.; Flonta, M.; Brad, A.; Sigarteu, A. Application of molecular techniques to the study of Pseudomonas aeruginosa clinical isolate in Cluj-Napoca, Romania. Analele Univ. din Oradea Fasc. Biol. 2010, XVII, 243–247. [Google Scholar]
  336. Dortet, L.; Flonta, M.; Boudehen, Y.M.; Creton, E.; Bernabeu, S.; Vogel, A.; Naas, T. Dissemination of carbapenemase-producing Enterobacteriaceae and Pseudomonas aeruginosa in Romania. Antimicrob. Agents. Chemother. 2015, 59, 7100–7103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  337. Georgescu, M.; Gheorghe, I.; Curutiu, C.; Lazar, V.; Bleotu, C.; Chifiriuc, M.C. Virulence and resistance features of Pseudomonas aeruginosa strains isolated from chronic leg ulcers. BMC Infect. Dis. 2016, 16, 92. [Google Scholar] [CrossRef] [Green Version]
  338. Porumbel, I.; Gheorghe, I.; Popa, M.; Banu, O.; Lazăr, V.; Chifiriuc, M.C. Virulence and resistance profiles among Pseudomonas aeruginosa isolated from patients with cardiovascular diseases in Bucharest. Biointerface Res. Appl. Chem. 2016, 6, 1176–1179. [Google Scholar]
  339. Matroș, L.; Krausz, T.L.; Pandrea, S.L.; Ciontea, M.I.; Chiorean, E.; Pepelea, L.S.; Berar, A.M.; Junie, L.M. Phenotypic and genotypic study of carbapenem-resistant strains isolated from hospitalized patients. Rev. Romana. Med. Lab. 2016, 24, 2. [Google Scholar] [CrossRef] [Green Version]
  340. Tatu, L.A.; Merezeanu, N.; Pantea, O.; Gheorghe, I.; Popa, M.; Banu, O.; Cristea, V.C.; Chifiriuc, M.C.; Lazar, V.; Marutescu, L. Resistance features of Pseudomonas aeruginosa strains isolated from patients with infectious complications of cardiovascular surgery. Biointerface Res. Appl. Chem. 2017, 7, 2004–2008. [Google Scholar]
  341. Mitache, M.M.; Gheorghe, I.; Totea, G.; Bleotu, C.; Curutiu, C.; Cochior, D.; Rusu, E.; Chifiriuc, M.C. Biochemical, virulence and resistance features in bacterial strains recovered from hospital surfaces after decontamination with quaternary ammonium compounds, triclosan and iodine disinfectants. Rev. Chim. 2017, 68, 997–1001. [Google Scholar] [CrossRef]
  342. Miftode, E.; Dorneanu, O.; Badescu, A.; Ghibu, L.; Leca, D.; Vremera, T.; Mereuţă, A. Emergence of a new group CTX-M enzyme in Romania and risk factors for extended spectrum beta-lactamase producing E. coli infections. Rev. Med. Chir. Soc. Med. Nat. Iasi. 2012, 116, 477–480. [Google Scholar]
  343. Székely, E.; Damjanova, I.; Jánvári, L.; Vas, K.E.; Molnár, S.; Bilca, D.V.; Lőrinczi, L.K.; Tóth, A. First description of bla(NDM-1), bla(OXA-48), bla(OXA-181) producing Enterobacteriaceae strains in Romania. Int. J. Med. Microbiol. 2013, 303, 697–700. [Google Scholar] [CrossRef]
  344. Lixandru, B.E.; Cotar, A.I.; Straut, M.; Usein, C.R.; Cristea, D.; Ciontea, S.; Tatu-Chiţoiu, D.; Codiţă, I.; Rafilă, A.; Nica, M.; et al. Carbapenemase-producing Klebsiella pneumoniae in Romania: A six-month survey. PLoS ONE 2015, 10, e0143214. [Google Scholar] [CrossRef]
  345. Ghenea, A.E.; Zlatian, O.M.; Cristea, O.M.; Ungureanu, A.; Mititelu, R.R.; Balasoiu, A.T.; Vasile, C.M.; Salan, A.-I.; Iliuta, D.; Popescu, M.; et al. TEM, CTX-M, SHV genes in ESBL-producing Escherichia coli and Klebsiella pneumoniae isolated from clinical samples in a county Clinical emergency hospital Romania-predominance of CTX-M-15. Antibiotics 2022, 11, 503. [Google Scholar] [CrossRef] [PubMed]
  346. Friedman, N.D.; Kaye, K.S.; Stout, J.E.; McGarry, S.A.; Trivette, S.L.; Briggs, J.P. Healthcare-associated bloodstream infections in adults: A reason to change the accepted definition of community-acquired infections. Ann. Intern. Med. 2002, 137, 791–797. [Google Scholar] [CrossRef] [PubMed]
  347. Farkas, A.; Tarco, E.; Butiuc-Keul, A. Antibiotic resistance profiling of pathogenic Enterobacteriaceae from Cluj-Napoca, Romania. Germs. 2019, 9, 17–27. [Google Scholar] [CrossRef] [PubMed]
  348. Trușcă, B.-S.; Gheorghe, I.; Marutescu, L.; Curutiu, C.; Marinescu, F.; Ghiță, C.M.; Borcan, E.; Țuică, L.; Minciuna, V.; Gherghin, H.-E.; et al. Beta-lactam and quinolone resistance markers in uropathogenic strains isolated from renal transplant recipients. Rev. Romana Med. Lab. 2017, 25, 365–373. [Google Scholar] [CrossRef] [Green Version]
  349. Chelariu, M.; Grosu, M.; Gheorghe, I.; Gradisteanu, G.; Picu, A.; Petcu, L.; Finlay, B.B. Host metabolic syndrome can disrupt the intestinal microbiota and promote the acquisition of resistance and virulence genes in Enterobacteriaceae stains. Rom. Biotechnol. Lett. 2017, 22, 12643–12650. [Google Scholar]
  350. Al Shaikhli, N.H.; Cristea, V.C.; Gheorghe, I.; Irrayif, S.M.; Hamzah Basil, M.; Sadik, O.; Mihaescu, G. Antibiotic resistance features in Klebsiella pneumoniae and Escherichia coli strains isolated from hospital and community acquired urinary tract infections. Rom. Biotechnol. Lett. 2021, 26, 2244–2248. [Google Scholar]
  351. Popa, L.I.; Gheorghe, I.; Barbu, I.C.; Surleac, M.; Paraschiv, S.; Măruţescu, L.; Popa, M.; Pîrcălăbioru, G.G.; Talapan, D.; Niţă, M.; et al. Klebsiella pneumoniae ST101 clone survival chain from inpatients to hospital effluent after chlorine treatment. Front. Microbiol. 2021, 11, 610296. [Google Scholar] [CrossRef]
  352. Rayyif, S.M.; Alwan, W.N.; Mohammed, H.B.; Barbu, I.C.; Holban, A.M.; Gheorghe, I.; Banu, O.; Shalal, O.S.; Chifiriuc, M.C.; Mihaescu, G. Snapshot of resistance and virulence features in ESCAPE strains frequently isolated from surgical wound infections in a Romanian hospital. Rev. Romana. Med. Lab. 2022, 30, 215–226. [Google Scholar] [CrossRef]
  353. Usein, C.R.; Papagheorghe, R.; Oprea, M.; Condei, M.; Strãuţ, M. Molecular characterization of bacteremic Escherichia coli isolates in Romania. Folia Microbiol. 2016, 61, 221–226. [Google Scholar] [CrossRef]
  354. Molnár, S.; Eszter Vas, K.; Székely, E. Carbapenemase Producing Enterobacterales in Romania: Investigating the Origins. Rev. Rom. Med. Lab. 2020, 28, 341–348. [Google Scholar] [CrossRef]
  355. Cozma, A.P.; Rimbu, C.M.; Zendri, F.; Maciuca, I.E.; Timofte, D. Clonal Dissemination of Extended-Spectrum Cephalosporin-Resistant Enterobacterales between Dogs and Humans in Households and Animal Shelters of Romania. Antibiotics 2022, 11, 1242. [Google Scholar] [CrossRef] [PubMed]
  356. Trușcă, B.S.; Ianculescu, E.; Maruțescu, L.; Gheorghe-Barbu, I.; Chifiriuc, M.-C.; Lazăr, V. Comparative Performance of Two Immunochromatographic Tests for the Rapid Detection of PCR Confirmed, Carbapenemase Producing Enterobacterales. Bointerface Res. Appl. Chem. 2022, 13, 322. [Google Scholar] [CrossRef]
  357. Surleac, M.; Barbu, I.C.; Paraschiv, S.; Popa, L.I.; Gheorghe, I.; Marutescu, L.; Popa, M.; Sarbu, I.; Talapan, D.; Nita, M.; et al. Whole genome sequencing snapshot of multi-drug resistant Klebsiella pneumoniae strains from hospitals and receiving wastewater treatment plants in Southern Romania. PLoS ONE 2020, 15, e0228079. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  358. Morin, C.A.; Hadler, J.L. Population-based incidence and characteristics of community-onset Staphylococcus aureus infections with bacteremia in 4 metropolitan Connecticut areas. J. Infect. Dis. 2001, 184, 1029–1034. [Google Scholar] [CrossRef] [Green Version]
  359. Cerini, P.; Meduri, F.R.; Tomassetti, F.; Polidori, I.; Brugneti, M.; Nicolai, E.; Bernardini, S.; Pieri, M.; Broccolo, F. Trends in Antibiotic Resistance of Nosocomial and Community-Acquired Infections in Italy. Antibiotics 2023, 12, 651. [Google Scholar] [CrossRef]
  360. Langdon, A.; Program, F.T.C.P.E.; Schwartz, D.J.; Bulow, C.; Sun, X.; Hink, T.; Reske, K.A.; Jones, C.; Burnham, C.-A.D.; Dubberke, E.R.; et al. Microbiota restoration reduces antibiotic-resistant bacteria gut colonization in patients with recurrent Clostridioides difficile infection from the open-label PUNCH CD study. Genome Med. 2021, 13, 28. [Google Scholar] [CrossRef]
  361. Wang, X.; Ryu, D.; Houtkooper, R.H.; Auwerx, J. Antibiotic use and abuse: A threat to mitochondria and chloroplasts with impact on research, health, and environment. Bioessays 2015, 37, 1045–1053. [Google Scholar] [CrossRef]
  362. Oprea, M.; Militaru, M.C.; Ciontea, A.S.; Cristea, D.; Cristea, V.; Usein, C.R. Characterization of antibiotic resistance integrons harbored by Romanian Escherichia coli uropathogenic strains. Rev. Romana Med. Lab. 2020, 28, 331–340. [Google Scholar] [CrossRef]
  363. Cristea, V.C.; Gheorghe, I.; Czobor Barbu, I.; Popa, L.I.; Ispas, B.; Grigore, G.A.; Bucatariu, I.; Popa, G.L.; Angelescu, M.C.; Velican, A.; et al. Snapshot of phylogenetic groups, virulence, and resistance markers in Escherichia coli uropathogenic strains isolated from outpatients with urinary tract infections in Bucharest, Romania. Bio. Med. Res. Int. 2019, 8, 5712371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  364. Porumbel, I.; Chelariu, M.; Gheorghe, I.; Curutiu, C.; Popa, M.; Delcaru, C.; Grosu, M.; Grigore, R.; Bertesteanu, S.; Lazar, V.; et al. Genotypic characterization of virulence and resistance markers in ESBL positive E. coli strains isolated from ambulatory urinary tract infections. Rom. Biotechnol. Lett. 2018, 22, 3. [Google Scholar]
  365. Mladin, C.; Chifiriuc, M.C.; Palade, A.; Usein, C.R.; Slavu, C.L.; Rucoreanu, A.; Damian, M. Phenotypic and genetic analysis of the antibiotic resistance patterns in uropathogenic Escherichia coli strains. Rom. Biotechnol. Lett. 2010, 15, 2. [Google Scholar]
  366. Pomba, C.; Rantala, M.; Greko, C.; Baptiste, K.E.; Catry, B.; Van Duijkeren, E.; Mateus, A.; Moreno, M.A.; Pyörälä, S.; Ružauskas, M.; et al. Public health risk of antimicrobial resistance transfer from companion animals. J. Antimicrob. Chemother. 2017, 72, 957–968. [Google Scholar] [CrossRef] [PubMed]
  367. Rendle, D.I.; Page, S.W. Antimicrobial resistance in companion animals. Equine. Vet. J. 2018, 50, 147–152. [Google Scholar] [CrossRef] [PubMed]
  368. Loeffler, A.; Lloyd, D.H. Companion animals: A reservoir for methicillin-resistant Staphylococcus aureus in the community? Epidemiol. Infect. 2010, 138, 595–605. [Google Scholar] [CrossRef] [PubMed]
  369. Davis, M.F.; Iverson, S.A.; Baron, P.; Vasse, A.; Silbergeld, E.K.; Lautenbach, E.; Morris, D.O. Household transmission of methicillin-resistant Staphylococcus aureus and other staphylococci. Lancet Infect. Dis. 2012, 12, 703–716. [Google Scholar] [CrossRef] [PubMed]
  370. Cristina, R.T.; Kocsis, R.; Dégi, J.; Muselin, F.; Dumitrescu, E.; Tirziu, E.; Herman, V.; Darău, A.P.; Oprescu, I. Pathology and Prevalence of Antibiotic-Resistant Bacteria: A Study of 398 Pet Reptiles. Animals 2022, 12, 1279. [Google Scholar] [CrossRef] [PubMed]
  371. Tîrziu, E.; Bărbălan, G.; Morar, A.; Herman, V.; Cristina, R.T.; Imre, K. Occurrence and Antimicrobial Susceptibility Profile of Salmonella spp. in Raw and Ready-To-Eat Foods and Campylobacter spp. in Retail Raw Chicken Meat in Transylvania, Romania. Foodborne Pathog Dis. 2020, 17, 479–484. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  372. Popa, S.A.; Morar, A.; Ban-Cucerzan, A.; Tîrziu, E.; Herman, V.; Sallam, K.I.; Morar, D.; Acaroz, U.; Imre, M.; Florea, T.; et al. Occurrence of Campylobacter spp. and Phenotypic Antimicrobial Resistance Profiles of Campylobacter jejuni in Slaughtered Broiler Chickens in North-Western Romania. Antibiotics 2022, 11, 1713. [Google Scholar] [CrossRef]
  373. Doma, A.O.; Popescu, R.; Mitulețu, M.; Muntean, D.; Dégi, J.; Boldea, M.V.; Radulov, I.; Dumitrescu, E.; Muselin, F.; Puvača, N.; et al. Comparative evaluation of qnrA, qnrB, and qnrS genes in Enterobacteriaceae ciprofloxacin-resistant cases, in swine units and a hospital from Western Romania. Antibiotics 2020, 9, 698. [Google Scholar] [CrossRef]
  374. Dan, S.D.; Tabaran, A.; Mihaiu, L.; Mihaiu, M. Antibiotic susceptibility and prevalence of foodborne pathogens in poultry meat in Romania. J. Infect. Dev. Ctries 2015, 9, 35. [Google Scholar] [CrossRef] [Green Version]
  375. Chirila, F.; Tabaran, A.; Fit, N.; Nadas, G.; Mihaiu, M.; Tabaran, F.; Cătoi, C.; Reget, O.L.; Dan, S.D. Concerning Increase in Antimicrobial Resistance in Shiga Toxin-Producing Escherichia coli Isolated from Young Animals during 1980-2016. Microbes Environ. 2017, 32, 252–259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  376. Costinar, L.; Herman, V.; Pitoiu, E.; Iancu, I.; Degi, J.; Hulea, A.; Pascu, C. Boar semen contamination: Identification of Gram-negative bacteria and antimicrobial resistance profile. Animals 2021, 12, 43. [Google Scholar] [CrossRef] [PubMed]
  377. Baquero, F.; Martínez, J.L.; Cantón, R. Antibiotics and antibiotic resistance in water environments. Curr. Opin. Biotechnol. 2008, 19, 260–265. [Google Scholar] [CrossRef] [PubMed]
  378. Bergeron, S.; Boopathy, R.; Nathaniel, R.; Corbin, A.; LaFleur, G. Presence of antibiotic resistant bacteria and antibiotic resistance genes in raw source water and treated drinking water. Int. Biodeterior. Biodegrad. 2015, 102, 370–374. [Google Scholar] [CrossRef]
  379. Kümmerer, K. Antibiotics in the aquatic environment–a review–part I. Chemosphere 2009, 75, 417–434. [Google Scholar] [CrossRef] [PubMed]
  380. Kim, S.; Aga, D.S. Potential ecological and human health impacts of antibiotics and antibiotic-resistant bacteria from wastewater treatment plants. J. Toxicol. Environ. Health B Crit. Rev. 2007, 10, 559–573. [Google Scholar] [CrossRef] [PubMed]
  381. Alrhmoun, M. Hospital wastewaters treatment: Upgrading water systems plans and impact on purifying biomass Doctoral dissertation, Université de Limoges. Environ. Eng. Univ. De Limoges 2014. [Google Scholar]
  382. Karkman, A.; Johnson, T.A.; Lyra, C.; Stedtfeld, R.D.; Tamminen, M.; Tiedje, J.M.; Virta, M. High-throughput quantification of antibiotic resistance genes from an urban wastewater treatment plant. FEMS Microbiol. Ecol. 2016, 92, fiw014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  383. Rizzo, L.; Manaia, C.; Merlin, C.; Schwartz, T.; Dagot, C.; Ploy, M.C.; Michael, I.; Fatta-Kassinos, D. Urban wastewater treatment plants as hotspots for antibiotic resistant bacteria and genes spread into the environment: A review. Sci. Total Environ. 2013, 447, 345–360. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  384. Krishnasamy, S.; Vijayan, J.V.; Ramya Srinivasan, G.D.; Indumathi, M.N. Antibiotic usage, residues and resistance genes from food animals to human and environment: An Indian scenario. J. Environ. Chem. Eng. 2018, 8, 102221. [Google Scholar] [CrossRef]
  385. Available online: https://www.ecdc.europa.eu/en/antimicrobial-consumption/database/country-overview (accessed on 26 February 2022).
  386. Available online: https://www.ema.europa.eu/en/documents/report/sales-veterinary-antimicrobial-agents-31-european-countries-2018-trends-2010-2018-tenth-esvac-report_en.pdf (accessed on 28 June 2022).
  387. Schlüter, A.; Szczepanowski, R.; Pühler, A.; Top, E.M. Genomics of IncP-1 antibiotic resistance plasmids isolated from wastewater treatment plants provides evidence for a widely accessible drug resistance gene pool. FEMS Microbiol. Rev. 2007, 31, 449–477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  388. Available online: https://www.jpiamr.eu/wp-content/uploads/2019/05/JPIAMR_SRIA_final.pdf (accessed on 28 June 2022).
  389. Chokshi, A.; Sifri, Z.; Cennimo, D.; Horng, H. Global contributors to antibiotic resistance. J. Glob. Infect. Dis. 2019, 11, 36. [Google Scholar] [CrossRef]
  390. Ţugui, C.G.; Vlădăreanu, I.; Baricz, A.; Coman, C. Detection of beta-lactamase resistance genes in a hospital chlorinated wastewater treatment system. Stud. Univ. Babes. Bolyai Biol. 2015, 60, 33–38. [Google Scholar]
  391. Lupan, I.; Carpa, R.; Oltean, A.; Kelemen, B.S.; Popescu, O. Release of antibiotic resistant bacteria by a waste treatment plant from Romania. Microbes Environ. 2017, 32, 219–225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  392. Butiuc-Keul, A.; Carpa, R.; Podar, D.; Szekeres, E.; Muntean, V.; Iordache, D.; Farkas, A. Antibiotic Resistance in Pseudomonas spp. Through the Urban Water Cycle. Curr. Microbiol. 2021, 78, 1227–1237. [Google Scholar] [CrossRef]
  393. Teban-Man, A.; Szekeres, E.; Fang, P.; Klümper, U.; Hegedus, A.; Baricz, A.; Berendonk, T.U.; Pârvu, M.; Coman, C. Municipal Wastewaters Carry Important Carbapenemase Genes Independent of Hospital Input and Can Mirror Clinical Resistance Patterns. Microbiol. Spectr. 2022, 10, e02711-21. [Google Scholar] [CrossRef] [PubMed]
  394. Van, L.T.; Hagiu, I.; Popovici, A.; Marinescu, F.; Gheorghe, I.; Curutiu, C.; Ditu, L.M.; Holban, A.-M.; Sesan, T.E.; Lazar, V. Antimicrobial Efficiency of Some Essential Oils in Antibiotic-Resistant Pseudomonas aeruginosa Isolates. Plants 2022, 11, 2003. [Google Scholar] [CrossRef] [PubMed]
  395. Banciu, A.R.; Ionica, D.L.; Vaideanu, M.A.; Radulescu, D.M.; Nita-Lazar, M.; Covaliu, C.I. The occurrence of potentially pathogenic and antibiotic resistant Gram-negative bacteria isolated from the Danube Delta ecosystem. Sustainability 2021, 13, 3955. [Google Scholar] [CrossRef]
  396. Alygizakis, N.A.; Besselink, H.; Paulus, G.K.; Oswald, P.; Hornstra, L.M.; Oswaldova, M.; Medema, G.; Thomaidis, N.S.; Behnisch, P.A.; Slobodnik, J. Characterization of wastewater effluents in the Danube River Basin with chemical screening, in vitro bioassays and antibiotic resistant genes analysis. Environ. Int. 2019, 127, 420–429. [Google Scholar] [CrossRef]
  397. Lazăr, V.; Gheorghe, I.; Curutiu, C.; Savin, I.; Marinescu, F.; Cristea, V.C.; Dobre, D.; Popa, G.L.; Chifiriuc, M.C.; Popa, M.I. Antibiotic resistance profiles in cultivable microbiota isolated from some romanian natural fishery lakes included in Natura 2000 network. BMC Vet. Res. 2021, 17, 52. [Google Scholar] [CrossRef] [PubMed]
  398. Marinescu, F.; Marutescu, L.; Savin, I.; Lazar, V. Antibiotic resistance markers among Gram-negative isolates from wastewater and receiving rivers in South Romania. Rom Biotechnol. Lett. 2015, 20, 10055–10069. [Google Scholar]
  399. Szekeres, E.; Baricz, A.; Chiriac, C.M.; Farkas, A.; Opris, O.; Soran, M.-L.; Andrei, A.-S.; Rudi, K.; Balcázar, J.L.; Dragos, N.; et al. Abundance of antibiotics, antibiotic resistance genes and bacterial community composition in wastewater effluents from different Romanian hospitals. Environ. Pollut. 2017, 225, 304–315. [Google Scholar] [CrossRef] [PubMed]
  400. Kittinger, C.; Lipp, M.; Folli, B.; Kirschner, A.; Baumert, R.; Galler, H.; Grisold, A.J.; Luxner, J.; Weissenbacher, M.; Farnleitner, A.H.; et al. Enterobacteriaceae isolated from the River Danube: Antibiotic resistances, with a focus on the presence of ESBL and carbapenemases. PLoS ONE 2016, 11, e0165820. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Schematic representation of mechanisms of actions and resistance to different antibiotic classes (Created with https://biorender.com/ accessed on 10 March 2023).
Figure 1. Schematic representation of mechanisms of actions and resistance to different antibiotic classes (Created with https://biorender.com/ accessed on 10 March 2023).
Ijms 24 07892 g001
Table 3. Transferable β-lactamase encoding genes described in A. baumannii and P. aeruginosa nosocomial strains isolated in Romania.
Table 3. Transferable β-lactamase encoding genes described in A. baumannii and P. aeruginosa nosocomial strains isolated in Romania.
Species/CloneCity and
Year of Isolation
Isolation SourceARGsMGEsReference
A. baumanniiBucharest
2001–2003
not providedblaIPM-1, blaVIM-2, blaOXA-24, blaOXA-51, blaOXA58class 1 integrons [319]
A. baumanniiIași
2003–2007
not providedblaVIM-2- [320]
A. baumanniiIași
2008–2012
urine, pus, sputum, tracheal aspirate, blood, cerebrospinal fluidblaVIM-2- [316]
A. baumanniiPetroșani
2004
rectal swabsblaPER-1
blaTEM-1
Tn1213 [321]
A. baumannii ST1; ST2Timișoara, Arad and Reșita
2009–2010
bronchial aspirates, wound secretions, catheters, stool, blood, urine cultures, tracheal secretionsblaOXA-23
blaPER-1
blaTEM-1
ISAba1 [322]
A. baumanniiIași and Tg.-Mureș
2014–2015
stool, blood, urine, tracheal secretions, stool samplesblaOXA-23, blaOXA-24/72
blaSHV-12
- [323]
A. baumannii ST437, ST764, ST765Bucharest
2011–2012
tracheal secretions, wounds, blood, urine, catheters, stoolblaOXA-23aci6 (pABKp1-like) [324]
A. baumanniiBucharest
2012–2013
not providedblaSHV-like- [325]
A. baumanniiBucharest
2011–2012
wound secretions, nasal–pharyngeal exudates, sputum, bronchoalveolar lavage, stool, urine, blood, pleural fluid, cathetersblaOXA-23- [326]
A. baumanniiBucharest
2014–2015
catheter, tracheal secretion, nasal exudateblaOXA-23, blaTEM- [327]
A. baumanniiBucharest
2015
chronic leg ulcer samplesblaOXA-72- [328]
A. baumanniiBucharest
2017
not providedblaOXA-23,- [329]
A. baumannii ST231Bucharest
2015
wound secretions left earblaOXA-23, blaOXA-64- [330]
A. baumanniiBucharest
2015–2016
nasal secretions, tracheal secretionsblaOXA-23- [331]
A. baumannii ST502Romania
2015–2017
lower respiratory tract, blood, skin or soft tissue, urine, intra-abdominal fluid, woundblaOXA-24, blaOXA-23- [332]
A. baumanniiBucharest
2017
not providedblaIPM, blaVIM-2, blaOXA-23, blaOXA-51Aci1, pACICU2,
rep135040, p3S18,
Aci6
[333]
A. baumannii ST636, ST492, ST1, ST2, ST642, ST312Bucharest
2017–2018
blood, urine, pharyngeal exudate, stool, anal carriage, tracheal secretion, catheter, cerebrospinal fluid, sputumblaOXA-23, blaOXA-24,
blaTEM-84 blaTEM-12 blaPER-1
class 1 integrons (5′-CS—dfrA12—aadA2—3′-CS; 5′-CS—aac(3)-I—aadA1—3′CS); plasmids (pACICU2-like; pMAL-1 like—blaOXA-72 [29]
A. baumanniiBucharest, Târgoviște, Vâlcea, Iași, Galați, Timișoara and Cluj
2018–2019
intra-hospital infections; not provided sourcesblaOXA-72;
blaOXA-23; blaOXA-66;
blaOXA-65;
blaTEM-12;
blaADC-30
qacE∆1 integron associated gene [334]
P. aeruginosa ST233, ST364, ST1074Bucharest
2012–2013
tracheal secretions, wounds, blood, urine, catheter, stool samplesblaVIM-2- [324]
P. aeruginosaCluj Napoca
2010
respiratory, urinary tract, and postoperative wound infectionsblaOXA-50, blaOXA-2- [335]
P. aeruginosaIași
2008–2012
urine, pus, sputum, tracheal aspirate, blood, cerebrospinal fluid, catheterblaVIM-2class 1 integrons (IntI1-aacA7-blaVIM-2-qacE1 and IntI1-aacA7-blaVIM-cmlA1-qacE1) [320]
P. aeruginosaIași and Tg.-Mureș
2014–2015
stool samplesblaVIM-2- [323]
P. aeruginosaBucharest
2012–2013
not providedblaSHVlike
blaGESlike blaVEBlike
- [325]
P. aeruginosaBucharest
2011–2012
wound secretions, nasal–pharyngeal exudates, sputum, bronchoalveolar lavage, stool, urine, blood, pleural fluid, and cathetersblaVIM-4 blaSPM-likeblaGESlikeclass 1 integrons (aacA7- blaVIM-4 aadB) [326]
P. aeruginosa ST2026, ST1982Cluj-Napoca
2011–2013
blood, urine, earblaVIM-2, blaIMP-13IncFIC [336]
P. aeruginosaBucharest
2014
chronic leg ulcersblaIMP- [337]
P. aeruginosaBucharest
2015
not providedblaIMP- [338]
P. aeruginosaCluj-Napoca
2016
pus, tracheal secretion, bile, sputum, blood, central venous catheters, urine, nasal secretion, stoolblaVIM
blaIMP
- [339]
P. aeruginosaBucharest
2014–2015
stool samples, tracheal secretionblaTEM- [327]
P. aeruginosaBucharest
2014–2015
wound secretions, bloodblaIMP- [340]
P. aeruginosaBucharest
2016
hospital surfacesblaTEM [341]
P. aeruginosaBucharest, Târgoviște, Vâlcea, Iași, Galați, Timișoara and Cluj
2018–2019
intra-hospital infections (sources not provided)blaIMP-13;
blaVIM-2;
blaVEB-9;
blaTEM-40;
[334]
Table 4. Genetic background and carrying platforms of β-lactam resistance in Enterobacterales clinical strains isolated in Romania.
Table 4. Genetic background and carrying platforms of β-lactam resistance in Enterobacterales clinical strains isolated in Romania.
Species/CloneCity and
Year of Isolation
Isolation SourceESBL GenesCarba-
Penemases
MGEsReference
E. coliIași
2012
urinary catheterblaCTX-M-15-- [342]
K. pneumoniae ST525, E. cloacae, E. coli ST131, K. pneumoniae ST101, Serratia marcescens, K. pneumoniae, S. marcescensTârgu Mureș
2013
blood, pus (not provided for all strains)blaCTX-M-15blaNDM-1, blaOXA-48, -181 [343]
K. pneumoniae, E. cloacaeBucharest
2014
wound secretions, nasal–pharyngeal exudates, sputum, bronchoalveolar lavage, stool, urine, blood, pleural fluid, and catheters-blaOXA-48, blaNDM-1- [326]
E. coliBucharest
2014
not provided-blaNDM-1- [325]
EnterobacteriaceaeBucharest
2014
not providedblaCTX-M, blaTEM, blaSHV-- [325]
K. pneumoniaeBacău
2015
not provided-blaVIM-1- [344]
K. pneumoniaeIași
2015
pharyngeal secretion, stool samples, feeding tub, bedside, food jarblaCTX-M-15, blaCTX-M-55/79-- [323]
K. pneumoniae ST147, ST395; E. cloacae ST114, P. stuartiiBucharest
2015
perineum, rectumblaCTX-M-15,
blaOXA-1, -10,
blaCMY-4, blaACT-16
blaOXA-48, blaNDM-1 [330]
E. coli, Enterobacter, ProteusCluj-Napoca
2016
urinary tract infections, wound infections, patients with persistent diarrheablaCTX-M intI1 (not related to blaCTX-M) [345]
K. pneumoniae, E. coliBucharest
2016
hospital surfacesblaTEM, blaCTX-MblaNDM [342]
E. coli ST5;
E. hormachei ST74, ST171;
K. pneumoniae ST101, ST395
Bucharest
2016
stool, bronchial secretion, pleural fluid, tracheal secretion, peritoneal fluidblaCTX-MblaNDM, blaOXA-48, -181IncFIIγ, IncL, IncR [346]
K. pneumoniaeIași
2017
urine, lower respiratory tract, blood, wound, puncture, and peritoneum sources-blaNDM-1- [347]
K. pneumoniaeBucharest
2017
urine, lower respiratory tract, blood, wound, puncture, and peritoneum sources-blaNDM-1, blaKPC-2, blaVIM-1- [344]
K. pneumoniae, M. morgannii, E. tardaBucharest
2017
urineblaCTX-M, blaTEM, blaSHV-- [348]
E. coli, K. pneumoniae, E. asburiae, Citrobacter freundiiBucharest

2017
stool samples, bloodblaTEM, blaCTX-MblaOXA-48- [349]
K. pneumoniaeBucharest
2018
not providedblaCTX-M, blaTEMblaNDM, blaOXA-48- [329]
E. coliCraiova
2021
purulent secretion, tracheal aspirate, catheterblaCTXM-15; blaTEM-1-- [345]
K. pneumoniaeCraiova
2021
sputum, peritoneal fluid, purulent secretion, tracheal aspirate, catheter, wound secretionblaCTX-M-15, blaSHV-1; blaTEM-1 [345]
K. pneumoniaeBucharest
2021
urinary tract infectionsblaTEMblaOXA-48 [350]
K. pneumoniae ST101Bucharest
2021
not provided-blaNDM-1 [351]
E. coli, K. pneumoniae, P. mirabilis, Citrobacter sp., M. morganii, P. stuarti, E. cloacae, S. marcescens, P. rettgeriBucharest
2022
surgical wound infectionsblaTEM, blaSHV, blaCTX-MblaOXA-48 [352]
E. coli ST131, ST10, ST131, ST167, ST410, ST540, ST1275, ST10, ST167Bucharest
2010–2012
bloodblaCTX-M-15-IncF [353]
Enterobacterales (K. pneumoniae, E. cloacae, E. coli; P. mirabilis; S. marcescens; S. liquefaciens; S. plymuthica)Tg.-Mureș
2012–2013
respiratory tract infections, skin, and soft tissue infections, urine, blood, stool samples, catheter tips, bile, cerebrospinal fluid, pleural fluid, peritoneal fluid-blaOXA-48, blaNDM-1IncR, L, FIIK, FII, FIB KN, HI2, M, A/C [354]
E. coliIași
2017–2018
stool samplesblaCTX-M-1, -3, -14, -15 blaTEM-1, -55blaSHV-134
blaSHV-like;
-Inc FIB/FIA; F; F/I1; I1; L/M; HI2; P1; N; Y; HI2 [355]
EnterobacteralesBucharest 2017–2018
2017–2018
urine, bronchial secretions, blood, ascites fluids, abscesses, catheters blaKPC; blaOXA-48; blaNDM [356]
K. pneumoniae ST101, ST219Bucharest, Galați and Târgoviște
2018–2019
not providedblaCTX-M-15;
blaTEM-1, -150; blaSHV-11, -12, -33, -100, -101, -106, -107, -145, -158, -161, 187
blaNDM-1; blaKPC-2 blaOXA-48qacE1 [357]
Table 5. Genetic background and carrying platforms of non-β-lactam ARGs in Enterobacterales clinical strains isolated in Romania.
Table 5. Genetic background and carrying platforms of non-β-lactam ARGs in Enterobacterales clinical strains isolated in Romania.
SpeciesCity and
Year of Isolation
Isolation SourceAminoglycoside ResistanceQuinolone Resistance GenesOther ARGsMGEsReference
E. coliIași
2017–2018
faecal sampleaph(4)-Ia, aac(3)-IV, aadA, -A2; aph(3′)-IIa; aac(2′)-IIaqnrB, -Ssul3, tet(A)Inc plasmids [346]
K. pneumoniaeBucharest, Galați, Târgoviște, 2018–2020intra-hospital infectionsaac(3)IIa, -IId;
aac(6′)Ib,—Ibcr, -IId; aadA1, -A2;
ant(2′’)Ia; aph(3′)Ia; aph(3′’)Ib; aph(6)Id; rmtC
qnrB1, -4, -10, -19, -36, -67;
qnrD1; qnrS1;
tet(A), -(D)
catA1, -B3; cmlA5;
fosA6, -7
mphA; dfrA1, -7, -12, 14; sul1, -2; ble
qacE1 [357]
E. coli, Citrobacter spp., Enterobacter spp.Cluj
2016
urinary tract infections, wounds, diarrheaaac(3)-IIIa
aac(6′)-II (
aac(6′)-Ie-aph(2”)
qnrS (E. coli, Enterobacter)sul1, -2, -3intI1 [347]
E. coliTimișoara
2019–2020
blood-qnrA, -B, -S-- [358]
K. pneumoniae,
E. coli
Bucharest
2016
hospital surfaces qnrA [341]
K. pneumoniae, M. morgannii, E. tarda, Bucharest
2017
urine-qnrB-- [348]
K. pneumoniaeBucharest
2017
not provided-qnrA, -B, -S-- [329]
K. pneumoniaeBucharest
2017
not provided--tet(A), tet(D)- [350]
K. pneumoniae, E. coli, P. mirabilis, Citrobacter spp.Bucharest
2022
surgical wound infectionsaphAI; aadA1, -2qnrB, -S-- [352]
Table 6. Genetic background and carrying platforms of AR in Gram-negative strains isolated from community-acquired infections in Romania.
Table 6. Genetic background and carrying platforms of AR in Gram-negative strains isolated from community-acquired infections in Romania.
SpeciesCity
Year of Isolation
Isolation SourceESBL/CarbapenemaseOther ARGsMGEsReferences
E. coliBucharest
2017
urineblaTEM-tet(A), tet(D)- [350]
E. coliBucharest
2014–2015
urine-aadA1, -A1a, -A2, -A5, A-22, aadB
dfrA1, -A5, -A7, -A12, -A14, -A16, -A17
class 1 and 2 integrons [362]
E. coliBucharest
2019
urineblaCTX-M, blaTEM, blaSHV--- [363]
E. coliBucharest
2017
urineblaNDM
blaTEMlike
blaCTX-Mlike-
-- [364]
E. coliBucharest
2015
urineblaTEMlike blaCTX-M
blaNDM; blaOXA-48
qnrA, qnrB, qnrS- [365]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Barbu, I.C.; Gheorghe-Barbu, I.; Grigore, G.A.; Vrancianu, C.O.; Chifiriuc, M.C. Antimicrobial Resistance in Romania: Updates on Gram-Negative ESCAPE Pathogens in the Clinical, Veterinary, and Aquatic Sectors. Int. J. Mol. Sci. 2023, 24, 7892. https://doi.org/10.3390/ijms24097892

AMA Style

Barbu IC, Gheorghe-Barbu I, Grigore GA, Vrancianu CO, Chifiriuc MC. Antimicrobial Resistance in Romania: Updates on Gram-Negative ESCAPE Pathogens in the Clinical, Veterinary, and Aquatic Sectors. International Journal of Molecular Sciences. 2023; 24(9):7892. https://doi.org/10.3390/ijms24097892

Chicago/Turabian Style

Barbu, Ilda Czobor, Irina Gheorghe-Barbu, Georgiana Alexandra Grigore, Corneliu Ovidiu Vrancianu, and Mariana Carmen Chifiriuc. 2023. "Antimicrobial Resistance in Romania: Updates on Gram-Negative ESCAPE Pathogens in the Clinical, Veterinary, and Aquatic Sectors" International Journal of Molecular Sciences 24, no. 9: 7892. https://doi.org/10.3390/ijms24097892

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop