Next Article in Journal
Extracellular Matrix-Based Approaches in Cardiac Regeneration: Challenges and Opportunities
Previous Article in Journal
The NF-Y Transcription Factor Family in Watermelon: Re-Characterization, Assembly of ClNF-Y Complexes, Hormone- and Pathogen-Inducible Expression and Putative Functions in Disease Resistance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Role of Efflux Pumps on Antimicrobial Resistance in Pseudomonas aeruginosa

by
Andre Bittencourt Lorusso
1,2,
João Antônio Carrara
1,
Carolina Deuttner Neumann Barroso
3,
Felipe Francisco Tuon
4 and
Helisson Faoro
1,5,*
1
Laboratory for Applied Science and Technology in Health, Carlos Chagas Institute, Fiocruz, Curitiba 81350-010, Brazil
2
School of Medicine and Life Sciences, Pontifícia Universidade Católica do Paraná, Curitiba 80215-901, Brazil
3
Laboratory of Clinical Veterinary Parasitology, Federal University of Paraná, Curitiba 80035-050, Brazil
4
Laboratory of Emerging Infectious Diseases, Pontifícia Universidade Católica do Paraná, Curitiba 80215-901, Brazil
5
CHU de Quebec Research Center, Department of Microbiology, Infectious Disease and Immunology, University Laval, Quebec, QC G1V 0A6, Canada
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(24), 15779; https://doi.org/10.3390/ijms232415779
Submission received: 2 November 2022 / Revised: 2 December 2022 / Accepted: 5 December 2022 / Published: 13 December 2022
(This article belongs to the Section Molecular Microbiology)

Abstract

:
Antimicrobial resistance is an old and silent pandemic. Resistant organisms emerge in parallel with new antibiotics, leading to a major global public health crisis over time. Antibiotic resistance may be due to different mechanisms and against different classes of drugs. These mechanisms are usually found in the same organism, giving rise to multidrug-resistant (MDR) and extensively drug-resistant (XDR) bacteria. One resistance mechanism that is closely associated with the emergence of MDR and XDR bacteria is the efflux of drugs since the same pump can transport different classes of drugs. In Gram-negative bacteria, efflux pumps are present in two configurations: a transmembrane protein anchored in the inner membrane and a complex formed by three proteins. The tripartite complex has a transmembrane protein present in the inner membrane, a periplasmic protein, and a porin associated with the outer membrane. In Pseudomonas aeruginosa, one of the main pathogens associated with respiratory tract infections, four main sets of efflux pumps have been associated with antibiotic resistance: MexAB-OprM, MexXY, MexCD-OprJ, and MexEF-OprN. In this review, the function, structure, and regulation of these efflux pumps in P. aeruginosa and their actions as resistance mechanisms are discussed. Finally, a brief discussion on the potential of efflux pumps in P. aeruginosa as a target for new drugs is presented.

1. The Silent Pandemic of Antimicrobial-Resistant Bacteria

Antibiotics revolutionized medicine in the 1930s–1940s, allowing millions of lives to be saved due to their ability to end infections caused by pathogenic bacteria. The first antibiotic used to treat an infectious disease was arsphenamine, also known by the trade name Salvarsan. This synthetic molecule, discovered by Paul Ehrlich, was widely used for the treatment of syphilis [1]. In time, the first commercially available broad-spectrum antibiotic was penicillin, discovered by Ian Flemming in 1929 [2]. Penicillin was introduced and widely used in the 1940s, during World War II, and helped to save many soldiers suffering from war wounds. However, its indiscriminate use quickly led to the emergence of the first penicillin-resistant bacterial strains [3].
Antimicrobial resistance (AMR) is an old pandemic that has become a major public health crisis. According to the World Health Organization (WHO), the topic is considered one of the most important threats to the human species in the 21st century [4]. It is a natural phenomenon shaped by natural selection and evolution. However, in recent decades there has been an acceleration in the emergence of AMR bacteria due to selective pressures caused by excessive or inappropriate use of antibiotics [5]. A relationship between antibiotic consumption, incorrect prescriptions, and the spread of resistant strains of bacteria has already been reported, as well as a high rate (30–50%) of incorrect prescriptions of these drugs [6]. In addition, there is a lack of interest on the part of the largest pharmaceutical companies in developing new antibiotics due to the high cost and low financial return, resulting in a significant innovation gap in the development of antibiotics [7]. Incorrect prescriptions, extensive antibiotic use in agriculture and livestock, and the low availability of new antimicrobials are some of the reasons for the worsening antibiotic resistance crisis. A global study that evaluated the impact of AMR on public health showed that, in 2019, most deaths in the world resulted directly from or were associated with resistant bacteria, which were a contributing factor in the loss of 4.95 million lives [8]. To put this in perspective, estimates of the number of deaths caused by the SARS-CoV-2 virus in the first year of the pandemic (2020) range between 1.8 and 3 million [9]. The effects of the SARS-CoV-2 pandemic slowed down due to mass vaccination and, to a large extent, socio-educational measures. In the long term, it is possible to envisage that the same type of measures will be applied to the use of antibiotics. Unfortunately, the AMR pandemic is not slowing down. In 2017, WHO published a list containing the priority antibiotic-resistant pathogens for the development of new drugs, which considered bacteria with resistance to multiple drugs to be of critical importance and to pose a threat to hospitals and patients who require devices such as ventilators and blood catheters [10]. The number of species of bacteria capable of causing infections is quite large; however, a group of bacteria called the ESKAPE pathogens stands out due to its high prevalence in nosocomial infections [11]. The ESKAPE group is formed by the bacteria Enterococcus faecium, Staphylococcus aureus, Klebsiella pneumoniae, Acinetobacter baumannii, Pseudomonas aeruginosa, and species of the genus Enterobacter.
Bacteria are known to employ a range of different resistance mechanisms. According to CARD database [12], the antimicrobial resistance mechanisms can be classified as: (i) the production of enzymes that modify the antibiotic molecule, leading to inactivation or destruction of the drug; (ii) changes in membrane permeability, which prevents the absorption of external substances; (iii) protection or alteration of target; (iv) target superexpression; (v) target replacement; and, (vi) efflux pumps that expel toxic components [13,14]. Over the years, the selective pressure caused by the continuous use of various antimicrobial agents has led microorganisms to accumulate various resistance mechanisms. Whether by horizontal transfer of resistance genes or selection of resistant mutants, the acquisition of several resistance factors has led to the emergence of multidrug-resistant (MDR) bacteria.

2. Pseudomonas Aeruginosa

P. aeruginosa is a Gram-negative bacillus commonly found in many environments, notably in soil and water exposed to intense human activity (e.g., wastewater-, hydrocarbon-, and pesticide-contaminated soil) [15]. It is an opportunistic bacteria related to healthcare infections, including ventilator-associated pneumonia (VAP), intensive care unit infections, central line-related bloodstream infections, surgical site infections, urinary tract infections, burn wound infections, keratitis, and otitis media [16]. This pathogen affects immunocompromised patients, in part due to its ability to evade both innate and acquired immune defenses through adhesion, colonization, and biofilm formation, and to produce various virulence factors that cause significant tissue damage. It also causes diseases with a high mortality rate in patients diagnosed with cystic fibrosis, neonatal infections, cancer, and severe burns [17,18,19]. Patients with cystic fibrosis present a higher incidence of colonization and infection by P. aeruginosa. Epithelial cells of the lung can ingest the invading P. aeruginosa, followed by desquamation, thus, protecting lungs from infection. However, patients with cystic fibrosis phagocytose less P. aeruginosa [20]. This reduced phagocytosis is associated with internal system defect and protection of specific P. aeruginosa LPS present in the external membrane of the bacteria [21].
P. aeruginosa frequently displays both multidrug resistance (MDR), when the bacterium is resistant to at least one agent in three or more antimicrobial categories, and extensive drug resistance (XDR), when the bacteria remain susceptible to only one or two categories [22,23,24]. According to data from the CDC, in 2017, 33,600 cases of infection and 2700 deaths from P. aeruginosa MDR were identified, associated with an expenditure of USD 767 million [25]. A strong aggravating factor in an infection by P. aeruginosa is its ability to produce biofilms. A biofilm is a structure formed by a bacterial community aggregated by components of the extracellular matrix secreted by the cells that compose it. The biofilm presents a protective environment for cells against abiotic stresses, immune system, and antibiotic action, which makes it difficult to eliminate the infection [16].
This bacterium was isolated for the first time in 1882 from cutaneous wounds of two patients with bluish–green pus, reported by Carle Gessard [26]. After that, several reports associated P. aeruginosa as the cause of blue–green purulence in patients’ wounds. The characteristic pigment, called pyocyanin, caught the researchers’ attention years before the bacterium was isolated. It was extracted by Fordos in 1860 and was associated with rod-shaped organisms in 1862 [27]. P. aeruginosa grows on most culture media used in routine analysis, and it can grow at 42 °C, a characteristic that distinguishes it from other Pseudomonas species. Additionally, colonies are often green due to the production of pyocyanin. Pyocyanin is an important virulence factor in infections caused by P. aeruginosa due to its toxic effects, putting cellular systems under increased oxidative stress [28]. Moreover, high concentrations of pyocyanin are present in the sputum of patients with chronic pulmonary disorders resulting from bacteremia [29]. Rudolf Emmerich and Oscar Löw showed in the 1890s that green bacteria (P. aeruginosa) isolated from patient bandages inhibited the growth of other bacteria [27,30]. Later, it was discovered that this antibiotic action was due to an enzyme called pyocyanase, which is considered to be the first clinic antibiotic used to treat human infections in hospitals [27].
The P. aeruginosa reference strain for genetics and functional analyses of physiology and metabolism is PAO1. P. aeruginosa PAO1 is a spontaneous chloramphenicol-resistant mutant of the original PAO strain that was isolated in 1954 from a wound in Melbourne, Australia [31,32]. A genetic map of its chromosome was constructed using the transduction and conjugation mechanisms of gene exchange in bacteria [33]. A physical map of the PAO1 genome was created using pulsed-field gel electrophoresis (PFGE) and afterward combined with the genetic map data [34,35]. The PAO1 strain was fully sequenced by 2000, and the genome is relatively large (5.5–7 Mbp) compared with other sequenced bacteria [36]. This indicates that the P. aeruginosa genome encodes a large proportion of regulatory enzymes important for metabolism, transportation, and efflux of organic compounds, which contributes to its intrinsic resistance to antibiotics and high adaptability to environmental changes. The contribution of efflux pumps to the emergence of MDR bacteria is significant. This is directly related to the ability of an efflux pump to export different classes of drugs [37]. Multidrug resistance has been demonstrated in all ESKAPE group bacteria, including P. aeruginosa [11].

3. Function and Structure of Efflux Pumps

Six families of efflux pumps have been described in the literature as having multidrug efflux capacity (Table 1): the small multidrug resistance (SMR) family; the major facilitator superfamily (MFS); the resistance/nodulation/cell division (RND) family; the ATP-binding cassette (ABC) superfamily; the multidrug and toxic compound extrusion (MATE) family [38,39], and the recently described proteobacterial antimicrobial compound efflux (PACE) family [40]. Active efflux systems may be responsible for resistance to several chemically distinct antibiotics and bactericides, with alarming numbers of occurrences in environmental and clinical isolates [41,42,43]. Such systems have demonstrated a practically ubiquitous presence in all kingdoms [44,45], sharing mechanisms [39,46], and helping to increase the dynamics of resistance profiles [47,48].
Protein transport systems received much attention in the 1990s, with structural, functional, and phylogenetic studies that aimed to uncover their distinct evolutionary origins [39,43,44,49,50]. These first studies, focused on the RND (super)family, enabled the identification of three common protein units, an inner transmembrane transporter, a periplasmic protein, and an outer membrane channel protein, which combine to form a trans-periplasmic channel [45,51,52] that allows the capture of substrates in the periplasm or cytoplasm [45]. After the assembly of the protein complex, the system stabilizes in two different stages: a resting stage, and a transport stage [53,54]. Driven by the proton motive force, the RND substrate efflux goes through cycles of site access, substrate binding, and extrusion, all mediated by a strictly coordinated rotational mechanism [55] that allows the system to work as a pump mechanism. As other families began to be discovered and more effort started to be put in the study of efflux pump mechanisms, different structures and driving forces began to be described, such as single protein mechanisms [56,57,58,59,60], ion gradients [56,57], and ATP [59] dependent pumps.
Generally, the genes that encode efflux pumps are chromosomal, with a smaller number present on plasmids; the proteins have conserved motifs containing glycine, proline, aspartate, and other hydrophobic residues in addition to tandem-repeat structures linked to secondary structures and coiled-coil and β-strands [51,61] (reviewed elsewhere [39]). Taken together, these characteristics distinguish the class from other outer membrane proteins, helping the complex to function as a barrier to prevent drug access to its target [42,51,62].
Table 1. Major families of bacterial efflux pumps in P. aeruginosa.
Table 1. Major families of bacterial efflux pumps in P. aeruginosa.
Family 1Efflux PumpGene Ids 2Substrates 3References
ABCTtg2 (Mla)PA4456-PA4455-PA4454-PA4453-PA4452CHL, CIP, COL, DMF, DOX, LVX, MIN, OFX, TET, TGC, TOB, TMP[63,64,65]
PA1874-77PA1874-PA1875-PA1876-PA1877CIP, GEN, NOR, TOB[66,67]
PA3228PA3228CAR, LVX, NOR[68]
MFSMfs1PA1262PQT[69]
Mfs2/SmvAPA1282OCT, PQT[69,70]
CmlA1GNT62_RS22140CHL[71,72]
SMRPASmr/EmrEPaePA4990ACR, EtBr, GEN, KAN, NEO[73,74]
SugE subfamily SMRPA1882Further research is needed[75]
MATEPmpMPA1361ACR, BZK, CIP, EtBr, NOR, OFX, TPPCL[76]
PACEPA2880PA2880CHX[77]
RNDMexAB-OprMPA0425-PA0426-PA0427AMI, AMX, ATM, CAR, CR, MA, FEP, CFP, CFSL, CTX, FOX, CZOP, CPO, CES, CAZ, CZX, CRO, CXM, CHL, CTET, CIN, CIP, CLX, DP, DOR, ENX, ERY, FMOX, GEN, IPM, LVX, CLM, MEM, MOX, NAF, NAL, NOR, NOV, OFX, OMC, OTC, PG, PPA, PIP, PTZ, PMA, SPX, SPI, STN, SUL, TZB, TET, TIC, TOS[78,79,80,81]
MexCD-OprJPA4599-PA4598-PA4597AMX, MA, FEP, CFP, CFSL, CTX, FOX, CZOP, CPO, CES, CZX, CRO, CXM, CHL, CHX, CTET, CIN, CIP, CLX, DOR, ENX, ERY, FMOX, LVX, CLM, MEM, NAF, NAL, NOR, NOV, OFX, OMC, OTC, PG, PPA, PIP, PMA, SPX, SPI, TET, TOS[79,80,82,83]
TMexCD-TOprJLSG45_RS29735-LSG45_RS29740-LSG45_RS29745FEP, CEQ, CAZ, CTET, CIP, DOX, ERV, FLO, GEN, MIN, NAL, OTC, STR, TET, TGC[84,85]
MexEF-OprNPA2493-PA2494-PA2495CHL, QN, TET, TMP[86]
MexGHI-OpmDPA4205-PA4206-PA4207-PA42085-Me-PCA, ACR, EtBr, NOR, R6G, TET, V[87,88,89]
MexJK-OprMPA3677-PA3676-PA0427ERY, TET, TCS[90]
MexMN-OprMPA1435-PA1436-PA0427BAL30072, ATM, BIPM, CAR, CMN, CAZ, CFT, CHL, MEM, MET, MOX, NOV, PIP, SUL, TMC, TP, TIC[91,92]
MexPQ-OpmEPA3523-PA3522-PA3521Hoechst 33342, CHL, CIP, ERY, KIT, NOR, RKM, TET, TPPCL[91]
MexVW-OprMPA4374-PA4375-PA0427ACR, CPO, CHL, ERY, EtBr, NOR, OFX, TET[93]
MexXY-OprM (-OprA)PA2019-PA2018-PA0427 (-PSPA7_3271)ACR, AMI, AMX, MA, FEP, CFP, CFSL, CTX, FOX, CZOP, CPO, CAZ, CZX, CRO, CXM, CHL, CTET, CIN, CIP, CLX, DOR, ENX, ERY, EtBr, FMOX, GEN, IPM, KAN, LVX, CLM, MEM, NAF, NAL, NEO, NOR, OFX, OMC, OTC, PG, PPA, PIP, PTZ, PMA, SPX, SPI, STR, TZB, TET, TIC, TOB, TOS, CAR (only with OprA), SUL (only with OprA)[79,80,94,95,96]
MuxABC-OpmBPA2528-PA2527-PA2526-PA2525ATM, ERY, KIT, NOV, RKM, TET[97]
TriABC-OpmHPA0156-PA0157-PA0158-PA4974TCS[98]
1: ABC: ATP-binding cassette transporters; MFS: major facilitator superfamily; SMR: small multidrug resistance; MATE: multidrug and toxic-compound extrusion; PACE: proteobacterial antimicrobial compound efflux; RND: resistance-nodulation-cell division. 2: IDs obtained from the Pseudomonas Genome Database [99]. 3: ACR: acriflavine; AMI: amikacin; AMX: amoxicillin; ATM: aztreonam; BIPM: biapenem; BZK: benzalkonium chloride; CAR: carbenicillin; CAZ: ceftazidime; CEQ: cefquinome; CES: cefsulodin; CFP: cefoperazone; CFSL: cefoselis; CFT: ceftolozane; CHL: chloramphenicol; CHX: chlorhexidine; CIN: cinoxacin; CIP: ciprofloxacin; CLX: cloxacillin; CMN: carumonam; COL: colistin; CPO: cefpirome; CR: carvacrol; CRO: ceftriaxone; CTET: chlortetracycline; CTX: cefotaxime; CXM: cefuroxime; CZOP: cefozopran; CZX: ceftizoxime; DMF: dimethylformamide; DOR: doripenem; DOX: doxycycline; DP: dipyridyl; ENX: enoxacin; ERV: eravacycline; ERY: erythromycin; EtBr: ethidium bromide; FEP: cefepime; FLO: florfenicol; FMOX: flomoxef; FOX: cefoxitin; GEN: gentamicin; IPM: imipenem; KAN: kanamycin; KIT: kitasamycin; LCM: lincomycin; LVX: levofloxacin; MA: cefamandole; MEM: meropenem; MET: methicillin; MIN: minocycline; MOX: moxalactam; NAF: nafcillin; NAL: nalidixic acid; NEO: neomycin; NOR: norfloxacin; NOV: novobiocin; OCT: octenidine; OFX: ofloxacin; OMC: oleandomycin; OTC: oxytetracycline; PG: penicillin G; PIP: piperacillin; PMA: piromidic acid; PPA: pipemidic acid; PQT: paraquat; PTZ: piperacillin-tazobactam; QN: quinolones; R6G: rhodamine 6G; RKM: rokitamycin; SPI: spiramycin; SPX: sparfloxacin; STN: streptonigrin; STR: streptomycin; SUL: sulbenicillin; TCS: triclosan; TET: tetracycline; TGC: tigecycline; TIC: ticarcillin; TMC: temocillin; TMP: trimethoprim; TOB: tobramycin; TOS: tosufloxacin; TP: thiamphenicol; TPPCL: tetraphenylphosphonium chloride; TZB: tazobactam; V: vanadium.

3.1. General Role of Efflux Pumps in Bacteria in a Natural Environment

Efflux pumps evolved as a way for bacteria to interact with the environment [100]. Their ubiquitous presence in nature [101] allowed bacteria to survive in their ecological niches [38], protecting them from toxic compounds produced by other species of bacteria or by the host, antimicrobial molecules, reactive oxygen species (ROS), and toxic byproducts of biochemical degradation pathways [38,43,52,102,103,104]. Efflux pumps can be specific for only one substrate or can export a range of molecules [38,62,105]. This difference in substrate selection is directly associated with the physicochemical characteristics of the substrate and the molecular interactions between the substrate, protein binding site, and environment, possibly reflecting the ability of the efflux pumps to capture their substrates [106]. As such, some of the main substrates of efflux pumps cited in the literature are detergents [52,62,107,108], antibiotics [52,62,102,108,109], organic solvents [102,110], dyes [62,102,107,108,109], disinfectants, aromatic hydrocarbons [109], cationic biocides [111], free fatty acids [112], lipophilic drugs, amphiphilic agents [107], antiseptics [62,108], anticancer drugs, uncouplers [108], and anti-virulence compounds [43].
The transcriptional regulation of efflux systems depends directly on the concentration of external signals [113], together with the action of internal transcription factors such as the TetR family proteins [114], LysR-type transcriptional regulators (LTTRs) [115], the multidrug regulator OstR [113], and activators MarA, Rob, and SoxS [112]. Cross-talking between efflux systems [113] and the need for chaperone-assisted folding increases its complexity, as observed in the downregulation of OmpF porin, which reduces cell permeability [52]. As observed, the presence of efflux pumps contributes to the maintenance of cell viability, virulence, and quorum sensing [43,102], assisting the bacterial cell in tolerating rapidly changing environments [101]. In this case, mechanisms of virulence and resistance evolved to fulfill different functions in the natural environment earlier than and not directly related to their role in multidrug resistance [43].

3.2. Efflux Pumps Involved in AMR in Bacteria

The first study that directly evaluated the action of efflux pumps on antibiotic resistance dates to the 1980s and revealed the action of plasmid-transferred energy-dependent efflux pumps on tetracycline resistance in Escherichia coli [46]. McMurry and colleagues demonstrated the ability of efflux pumps to shift the drug even against a concentration gradient [46], showing a strict relationship between the efflux pump together with outer membrane permeability and the acquisition of drug resistance [42,52].
Efflux systems have been found in comparable numbers among pathogenic and nonpathogenic bacteria. Its emergence earlier than the extensive exposure to antimicrobial agents suggested that their evolution as a resistance mechanism occurred independently of the use of antimicrobials [44]. The efflux system allows the bacteria to survive at low levels of drugs, conferring time to develop a more specialized mechanism of resistance [116]. When combined with the outer membrane barrier, the efflux system guarantees cellular protection from a range of compounds [106], favoring the dissemination of gene coding for such systems. Some isolates have genes for different efflux pumps in their genome, acting as genetic reservoirs [48,52]. This organization can also occur from gene duplication events [117] and gene cassettes [116].
Efflux pump specificity can follow a nonlinear behavior depending on the variety of compounds it can transport, leading to efficient extrusion even in cases of low substrate affinity for the binding site [106]. At the same time, such bonding can lead to the blocking of rotational movement and impediment of extrusion [53]. In this way, the structural changes observed can allow the binding of two or more substrate molecules in the same system, such as the extrusion of hydrophobic and hydrophilic compounds, expanding the specificity of the efflux pump [55].
The role of efflux pumps in AMR can be characterized as a byproduct of physiological functions that are beneficial to bacteria [38], and their efficiency depends on several factors, both internal and external to the bacterium. The presence of efflux pump genes can be considered the first step in the acquisition of resistance to antibiotics [118].

4. Efflux Pumps as a Mechanism of Antimicrobial Resistance in P. aeruginosa

P. aeruginosa has a naturally low susceptibility to antibiotics, explained by its extensive intrinsic resistome [119], low membrane permeability [120], and ability to form biofilms [16]. In its core genome, there are several antibiotic resistance genes, such as blaampC, and the genes encoding the multidrug efflux pumps (Mex) MexXY, MexAB-OprM, MexCD-OprJ, and MexEF-OprN [121], all of which are part of the RND superfamily.
These efflux pumps are tripartite protein complexes, which are drug/proton antiporters that catalyze the extrusion of their specific substrates from the periplasm through the outer membrane [122]. They are composed of three different proteins: a periplasmic adaptor protein, commonly known as periplasmic membrane fusion protein (PMFP), such as MexA, MexX, MexC, or MexE; a resistance-nodulation-cell division transporter (RNDt), such as MexB, MexY, MexD, or MexF, and a channel-forming outer membrane factor (OMF), such as OprM, OprJ, or OprN [123]. Structural analyses of the MexAB-OprM efflux pump show that the PMFP forms a stable complex associated with the inner membrane and the RNDt, both of which are recruited by the OMF, which resides in the outer membrane [124,125] (Figure 1).
As discussed before, efflux pumps are major factors in developing multidrug resistance in bacteria. The four multidrug efflux pumps in P. aeruginosa are involved in the extrusion of toxic molecules and contribute to reduced antibiotic susceptibility [126]. Furthermore, overexpression of these multidrug efflux pumps in P. aeruginosa is directly associated with resistance to most anti-pseudomonal drugs [80,127] and can potentially reduce the efficiency of new classes of drugs developed against P. aeruginosa and other Gram-negative pathogens, such as LpxC inhibitors [128,129] and cefiderocol [130]. These results highlight the importance of taking efflux mechanisms into consideration when developing antipseudomonal drugs.
Other efflux pumps identified in P. aeruginosa, such as MexJK and MexGHI-OpmD, have been mainly associated with quorum-sensing modulation and cell-to-cell communication [88,131] and will not be discussed further in this review.

4.1. MexAB-OprM

The operon MexAB-OprM was the first multidrug efflux pump reported in P. aeruginosa [78] and is considered the main contributor to antibiotic resistance, with MexA as the PMFP, MexB as the RNDt, and OprM as the OMF [132]. When the drug concentration increases near the pump, MexB goes through a conformational change and can eject active molecules toward a tunnel formed by MexA and OprM across the periplasm and outer membrane [133].
MexAB-OprM is constitutively expressed in wild-type strains, and its expression is controlled mainly by the repressor genes mexR [134], nalC [135], and nalD [136]. The use of antibiotics selects P. aeruginosa strains with increased MexAB-OprM expression [137,138]. Several mutations in repressor genes have been associated with this phenomenon [139], in particular mutations causing translational disruption, such as nonsense substitutions and frameshifts [138,140], disruption by insertion sequences [139,141], and non-synonymous substitutions that alter the molecular structure of the repressors [137,142]. However, there are commonly reported non-synonymous substitutions, such as V126E in mexR, which are not associated with an increase in efflux pump expression [143,144].
MexAB-OprM can export several antibiotics, including quinolones, macrolides, tetracyclines, lincomycin, chloramphenicol, novobiocin, and most β-lactams [79], and it has been shown to be a major factor in the resistance to the herbal antimicrobial compound carvacrol [81]. Overexpression of this efflux pump is associated with resistance to most antipseudomonal antibiotics, except for colistin [80], and carbapenemase-producing carbapenem-resistant P. aeruginosa strains often show an increased mexAB-oprM expression, which could be contributing to its resistance to carbapenems [145]. Moreover, it has been shown that overexpression of MexAB-OprM and AmpC, a chromosomally encoded class C β-lactamase that contributes to the resistance to many penicillins [146,147], has synergistic effects on the resistance of P. aeruginosa to most antipseudomonal β-lactams, except for ceftolozane/tazobactam, imipenem, and imipenem/relebactam [148].

4.2. MexXY

The MexXY efflux pump is formed by the PMFP MexX and the RNDt MexY and is the only multidrug efflux pump operon in P. aeruginosa that does not contain a coding sequence for an outer membrane factor. However, it does form a multidrug efflux pump together with OprM from the MexAB-OprM operon [94]. In some strains, such as P. aeruginosa PA7 [149], the MexXY operon contains a coding sequence for another outer membrane factor called OprA, which can also form a complex with MexXY [95,150].
Its expression is antimicrobial-inducible and is controlled mainly by the repressors MexZ [151], the two-component regulatory system ParRS [152], and by the aminoglycoside-inducible anti-repressor ArmZ [153,154]. Several inactivating or single amino acid mutations in mexZ have been associated with an increased expression of MexXY [144,155,156], particularly in cystic fibrosis isolates [125,157], a phenomenon that has also been reported for parR [125,152,158] and parS [125,152].
MexXY is known to contribute to aminoglycoside resistance [159], and its overexpression is commonly observed in strains bearing aminoglycoside-modifying enzymes (AMEs), in which MexXY and the AME work in synergy to promote aminoglycoside resistance [156]. Moreover, it has been shown that MexXY is also associated with resistance to most antipseudomonal antibiotics, similarly to MexAB [80], and it is particularly important in cystic fibrosis lung isolates of P. aeruginosa that harbor a defective MexAB pump [160]. In the OprA-harboring strain P. aeruginosa PA7, MexXY is also able to expel and confer resistance to carbenicillin and sulbenicillin, but only when partnering with the OMF OprA [95]. Moreover, specific single amino acid substitutions in MexY have been linked to increased resistance to aminoglycosides, cefepime, and fluoroquinolones, highlighting the importance of the MexXY system for antibiotic resistance in P. aeruginosa [160].

4.3. MexCD-OprJ

The MexCD-OprJ operon encodes an efflux pump that is usually silent or expressed at a low level in P. aeruginosa [161], which normally does not contribute to this pathogen’s natural resistance to antibiotics; however, it is associated with resistance to several classes of antibiotics when overexpressed [82,162]. Its expression is mainly regulated by the repressor nfxB, which is transcribed divergently from the mexCD-oprJ operon [82]. Several mutations in nfxB have been associated with an increased expression of MexCD-OprJ, such as nucleotide deletions [163], missense, and nonsense mutations [162].
MexCD-OprJ is mainly associated with resistance to fluoroquinolones such as levofloxacin and ciprofloxacin [80,161], but can also extrude other antimicrobial agents such as macrolides, novobiocin [79], tetracyclines, chloramphenicol [162,164,165], zwitterionic cephalosporins, such as cefepime and cefpirome [166], and the biocide chlorhexidine [83]. Additionally, mutations in mexD may change the substrate specificity of the efflux pump and have been associated with resistance to carbenicillin [167], ceftolozane-tazobactam, and ceftazidime-avibactam [163]. However, strains that overexpress MexCD-OprJ commonly show greater susceptibility to aminoglycosides and some β-lactams [166,168,169], while the overexpression of this efflux pump may lead to an increased susceptibility to complement-mediated killing [170]. Recently, Sanz-García and colleagues showed that exposure to subinhibitory concentrations of ciprofloxacin led to a selection of mutants with increased expression of MexCD and cross-resistance to other antibiotics [100].
Interestingly, variations of mexCD-oprJ genes have been found in mobile genetic elements in P. aeruginosa and several other bacteria [84,171]. A group of these mobile efflux pump operons have been called tmexCD-toprJ (“t” for transferable), and data suggest that they originated from the chromosome of a Pseudomonas spp. [85]. Several variations of the original tmexCD-toprJ have been identified since, either chromosomally or plasmid located, and have been directly associated with resistance to tetracyclines and reduced susceptibility to several other classes of antibiotics under laboratory conditions [84,85].

4.4. MexEF-OprN

As for MexCD-OprJ, the MexEF-OprN system is generally inactive under normal conditions but has been associated with resistance to chloramphenicol, quinolones, and trimethoprim when overexpressed [86]. The pump is regulated by MexT, a LysR-like transcriptional activator [172], and MexS, a putative oxidoreductase of unknown function [172,173], both encoded by genes located upstream of the mexEF-oprN operon (Figure 1). Interestingly, mexT is commonly found with inactivating mutations in wild-type P. aeruginosa strains, and overexpression of MexEF-OprN arises from the reversion of these mutations [174]. The mexS gene expression is also activated by MexT [172], and several mutations in mexS, commonly observed in clinical strains, are known to cause overexpression of MexEF-OprN [173,175,176,177].
Several other genes have also been associated with the modulation of mexEF-oprN expression. Iftikhar et al. [178] reported that a P. aeruginosa mutant with a defective pvcB gene from the pvc paerucumarin synthesis operon showed significant repression in mRNA levels of mexEF-oprN, mexT, and mexS, suggesting the involvement of the pvc operon in the transcriptional regulation of mexEF-oprN. It has also been shown that mvaT and ampR, global expression regulators in P. aeruginosa, negatively regulate MexEF-OprN [179,180], and data suggest the existence of still unknown MexEF-OprN regulators [177].
Overexpression of MexEF-OprN is also associated with decreased expression of the porin OprD because of the MexEF-OprN activator MexT acting as an oprD repressor [172,181]. Inactivating mutations and downregulation of oprD have been associated with resistance to carbapenems [139,182] and colistin [183], which explains why strains overexpressing MexEF-OprN are commonly resistant to antibiotics not extruded by this efflux pump.

5. Efflux Pumps and Biofilm Formation

Bacterial biofilms consist of microbial communities structured and organized in a matrix of exopolysaccharides (EPS), produced and secreted by the organisms themselves, adhered to a surface [184]. Efflux pumps play an important role in bacterial biofilm formation. Most studies in this area show that efflux pump expression is upregulated in biofilms, conferring greater antibiotic resistance [16]. There is an important positive feedback relationship between biofilms and efflux pumps. Tests of efflux pump inhibitors showed a significant reduction in biofilm in different bacteria [185]. The mechanism that underpins the relationship between efflux pumps and biofilm formation differs between bacterial species, but three main mechanisms have been identified [186]. Firstly, biofilm formation requires several molecules to be transported to the extracellular environment, a function that depends partially on efflux pumps [187]. Secondly, intercellular communication for biofilm formation is called quorum sensing, a type of molecule-mediated communication that requires certain molecules to be released into the extracellular environment. Without these molecules, biofilms cannot obtain the structure of a mature biofilm, expand, or release planktonic cells [188]. Thirdly, some antibiotics act on the regulation and synthesis of elements for the biofilm; therefore, efflux pumps that expel antibiotics are related to the maintenance of the biofilm [189].
Acyl-homoserine lactones (AHL) are important molecules for biofilm production in P. aeruginosa, as they facilitate quorum sensing [190]. Pearson et al. showed that the inhibition of efflux pumps led to an increase in the intracellular concentration of AHL and a reduction in the extracellular concentration, suggesting a clear association between efflux pumps and elements in biofilm formation [191]. The MexAB-OprM pump plays an important role in P. aeruginosa quorum sensing by delivering AHLs to the extracellular environment [187]. Furthermore, the overexpression of efflux pumps on P. aeruginosa can improve biofilm formation, with effects including denser structure, as demonstrated by Sánchez et al. [192]. P. aeruginosa with mutations on nalB and nfxB overexpresses MexAB-OprM and MexCD-OprJ, respectively, resulting in a denser biofilm formation compared with wild-type P. aeruginosa.
The MexEF-OprN efflux pump drives the efflux of 4-hydroxy-2-heptylquinoline (Pseudomonas quinolone signal), which is used by P. aeruginosa to facilitate quorum sensing [193]. Interestingly, studies comparing the transcriptome of planktonic and sessile cells (basal cells of the biofilm) demonstrate an upregulation of several genes associated with efflux pumps such as mexAB-oprM and mexCD-oprJ [36,194,195]. The work by Harrington et al. evaluated the P. aeruginosa transcriptome in an ex vivo pig lung model and a sputum model, showing a repression in MexGHI-OmpD expression in the lung model-associated biofilm [196]. Thus, it is possible to conclude that pump efflux interferes with biofilms but also that biofilms interfere with efflux pump regulation.

6. Efflux Pumps as Targets for New Drugs

Since the role of efflux pumps as a mechanism of resistance to several antimicrobials became known, significant resources have been devoted to developing molecules that can inhibit this system. Studies in this area began by attempting to find antibiotic-like molecules expelled from the bacteria to inhibit these pumps, such as derivatives of tetracyclines. Despite several in vitro studies, these drugs have never been clinically tested, mostly for reasons of serum instability and toxicity [197]. The first efflux pump inhibitors studied were verapamil, 2,4-dinitrophenol, lansoprazole, protonophore carbonyl cyanide m-chlorophenylhydrazone, and Phe-Arg-ß-naphthylamide. The latter is a dipeptide, a class of molecules that has been studied for their efflux pump inhibitory properties [198,199,200,201,202]. In P. aeruginosa, Phe-Arg-ß-naphthylamide increases the activity of quinolones, decreasing their expulsion from the bacterial cell and lowering their minimum inhibitory concentrations [203]. After these, several other similar molecules were studied, such as MC-207,110, MC-04,112, MC-02,434, MC-510,051, MC-04,124, and MC-02,595 [204]. MC-04,124 showed similar results to the other compounds in the series; however, the authors considered it to be a less toxic drug [205,206]. Peptides have also been successfully employed to inhibit SMR family efflux pumps. Mitchell and collaborators synthesized several peptides that targeted the TM4 region of P. aeruginosa SMR. The TM4 domain is important for homodimer assembly and for a functional SMR pump. They showed that the synthesized peptides were able not only to decrease the efflux via SMR but also to improve the biocidal activity of other compounds [207]. In addition to synthetic components, dozens of natural and herbal products have been tested for use as efflux pump inhibitors; while some have shown inhibitory activity, clinical studies have not yet been carried out [208].
Revisiting existing drugs with safe clinical use is an alternative to the development of new molecules. In vitro studies have shown that mefloquine, a well-known molecule widely used in the treatment of malaria [209], acts as an efflux pump inhibitor in P. aeruginosa [210]. Trimethoprim, a drug used in combination with sulfamethoxazole to treat various infections, also has efflux pump inhibitory activity. The use of trimethoprim together with ciprofloxacin has a synergistic effect, increasing sensitivity in P. aeruginosa to ciprofloxacin, although the mechanism is still unclear [211]. A further study with trimethoprim and sertraline, using Galleria mellonella as an in vivo infection model, showed the effectiveness of both molecules as synergists in levofloxacin treatment compared with levofloxacin alone. Significant increases in P. aeruginosa-infected G. mellonella survival were seen with treatment with levofloxacin and trimethoprim, or levofloxacin and sertraline, compared with monotherapies [212]. On the other hand, Laborda and collaborators tested a library of 1243 natural products for molecules capable of both increasing the expression of efflux pumps and providing them with a substrate without leading to an increase in antibiotic resistance in P. aeruginosa. The authors hypothesized that increasing the expression of these genes and the efflux of the inducing molecule could overcome the efflux of antibiotics while decreasing the efflux of virulent molecules to the outside. They found four compounds that were able to increase the expression of the mexCD-oprJ or mexAB-oprM genes while also decreasing the virulence of P. aeruginosa in tests with Caenorhabditis elegans: two coumarin-like compounds, one 1,4-benzodioxan-like compound, and one 4-chloroindole compound [213]. Additionally, using a drug library selection strategy, Tambat and collaborators found that the ethyl 4-bromopyrrole-2-carboxylate molecule can act synergistically with antibiotics, lowering their minimum inhibitory concentration in strains that overexpress RND family transporters [214]. Finally, comparative and structural studies between MexB and MexY from P. aeruginosa have shown that the presence of bulky tryptophan residues in the hydrophobic pit prevents binding of the efflux pump inhibitor ABI-PP [215].
Despite several potential candidates for efflux pump inhibitors in P. aeruginosa, obtaining an effective drug is still the subject of several studies. P. aeruginosa constitutively expresses several efflux pumps with different substrate specificities, different physicochemical properties and different efflux constants (KE). Understanding these variables and how the compounds correlate with these different penetration mechanisms is crucial to effectively combat P. aeruginosa antibiotic resistance. It is also interesting to report that these efflux pump inhibitors can also affect biofilm production since there is a close relationship between the efflux of signaling molecules for biofilm formation and quorum sensing, as discussed earlier [216]. The action on quorum sensing molecules has been shown in transcriptome studies, where Phe-Arg-β-naphthylamide was used to reduce the expression of genes related to biofilms [217].

7. Concluding Remarks

P. aeruginosa is an opportunistic pathogen of extreme clinical importance that is responsible for most secondary respiratory tract infections in patients with cystic fibrosis or undergoing intubation. The emergence of MDR and XDR strains of P. aeruginosa is a serious aggravating factor to this problem. Efflux pumps are among the most important AMR mechanisms that lead to the emergence of MDR and XDR strains. These protein complexes evolved from a system of interaction with other bacteria, the environment, and hosts with a survival resource. Efflux pumps can export dozens of different compounds into the extracellular environment, such as quorum sensing molecules, virulence factors, and toxic compounds such as antibiotics. This primary mechanism of antibiotic resistance is responsible for reducing the intracellular concentration of the drug to a sub-inhibitory concentration at which other more robust and specific resistance mechanisms can evolve and be selected for. As such, efflux pumps in P. aeruginosa have received significant attention in relation to the development of new drugs. Although several genes have been associated with the transcriptional regulation of efflux pump genes, P. aeruginosa has a complex transcriptional regulatory network in which many genes can be affected by unique global expression modulators. In addition, putative transcription factors with no known targets have been identified through in silico approaches. Additional studies are needed to accurately identify and characterize all trans-acting elements and substrates involved in efflux pump regulation, which may be crucial for the design of new antipseudomonal drugs. Unfortunately, the big pharmaceutical companies have abandoned investments in the development of new drugs due to the high associated cost. In this scenario, it may be necessary to establish public–private partnerships between governments, pharmaceutical companies and the WHO, aiming at the discovery or development of new antibiotic molecules.

Author Contributions

Conceptualization, A.B.L. and H.F.; investigation, A.B.L., J.A.C., C.D.N.B., F.F.T. and H.F.; data curation, A.B.L. and J.A.C.; writing—original draft preparation, A.B.L., J.A.C., C.D.N.B., F.F.T. and H.F.; visualization, A.B.L. and J.A.C.; supervision, H.F.; project administration, A.B.L. and H.F.; funding acquisition, H.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Council for Scientific and Technological Development (CNPq), CAPES, and Fiocruz. CNPq grant number 424410/2018-4; InovaFiocruz/Fundação Oswaldo Cruz grant number VPPCB- 07-FIO-18-2-38.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank the Carlos Chagas Institute, Fiocruz Paraná, and the postgraduate program in Biosciences and Biotechnology of the Carlos Chagas Institute for their support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Williams, K.J. The Introduction of “chemotherapy” Using Arsphenamine—The First Magic Bullet. J. R. Soc. Med. 2009, 102, 343–348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Fleming, A. On the Antibacterial Action of Cultures of a Penicillium, with Special Reference to Their Use in the Isolation of B. Influenzæ. Br. J. Exp. Pathol. 1929, 10, 226–236. [Google Scholar] [CrossRef]
  3. Sengupta, S.; Chattopadhyay, M.K.; Grossart, H.-P. The Multifaceted Roles of Antibiotics and Antibiotic Resistance in Nature. Front. Microbiol. 2013, 4, 47. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. World Health Organization. Global Antimicrobial Resistance and Use Surveillance System (GLASS) Report 2021; World Health Organization: Geneva, Switzerland, 2021; p. 180. [Google Scholar]
  5. Bassetti, M.; Kanj, S.S.; Kiratisin, P.; Rodrigues, C.; Van Duin, D.; Villegas, M.V.; Yu, Y. Early Appropriate Diagnostics and Treatment of MDR Gram-Negative Infections. JAC-Antimicrob. Resist. 2022, 4, dlac089. [Google Scholar] [CrossRef] [PubMed]
  6. Yang, Y.-H.; Buttery, J. Antimicrobial Resistance: A Global One-Health Problem for All Ages. World J. Pediatr. 2018, 14, 521–522. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Plackett, B. Why Big Pharma Has Abandoned Antibiotics. Nature 2020, 586, S50–S52. [Google Scholar] [CrossRef]
  8. Murray, C.J.; Ikuta, K.S.; Sharara, F.; Swetschinski, L.; Robles Aguilar, G.; Gray, A.; Han, C.; Bisignano, C.; Rao, P.; Wool, E.; et al. Global Burden of Bacterial Antimicrobial Resistance in 2019: A Systematic Analysis. The Lancet 2022, 399, 629–655. [Google Scholar] [CrossRef]
  9. World Health Organization. The True Death Toll of COVID-19: Estimating Global Excess Mortality; World Health Organization: Geneva, Switzerland, 2021. [Google Scholar]
  10. WHO. WHO Publishes List of Bacteria for Which New Antibiotics Are Urgently Needed. 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 12 September 2022).
  11. De Oliveira, D.M.P.; Forde, B.M.; Kidd, T.J.; Harris, P.N.A.; Schembri, M.A.; Beatson, S.A.; Paterson, D.L.; Walker, M.J. Antimicrobial Resistance in ESKAPE Pathogens. Clin. Microbiol. Rev. 2020, 33, e00181-19. [Google Scholar] [CrossRef]
  12. Alcock, B.P.; Raphenya, A.R.; Lau, T.T.Y.; Tsang, K.K.; Bouchard, M.; Edalatmand, A.; Huynh, W.; Nguyen, A.-L.V.; Cheng, A.A.; Liu, S.; et al. CARD 2020: Antibiotic Resistome Surveillance with the Comprehensive Antibiotic Resistance Database. Nucleic Acids Res. 2020, 48, D517–D525. [Google Scholar] [CrossRef]
  13. Van Camp, P.-J.; Haslam, D.B.; Porollo, A. Bioinformatics Approaches to the Understanding of Molecular Mechanisms in Antimicrobial Resistance. Int. J. Mol. Sci. 2020, 21, 1363. [Google Scholar] [CrossRef]
  14. Darby, E.M.; Trampari, E.; Siasat, P.; Gaya, M.S.; Alav, I.; Webber, M.A.; Blair, J.M.A. Molecular Mechanisms of Antibiotic Resistance Revisited. Nat. Rev. Microbiol. 2022. [Google Scholar] [CrossRef] [PubMed]
  15. Crone, S.; Vives-Flórez, M.; Kvich, L.; Saunders, A.M.; Malone, M.; Nicolaisen, M.H.; Martínez-García, E.; Rojas-Acosta, C.; Catalina Gomez-Puerto, M.; Calum, H.; et al. The Environmental Occurrence of Pseudomonas Aeruginosa. APMIS Acta Pathol. Microbiol. Immunol. Scand. 2020, 128, 220–231. [Google Scholar] [CrossRef] [PubMed]
  16. Tuon, F.F.; Dantas, L.R.; Suss, P.H.; Tasca Ribeiro, V.S. Pathogenesis of the Pseudomonas Aeruginosa Biofilm: A Review. Pathogens 2022, 11, 300. [Google Scholar] [CrossRef] [PubMed]
  17. Lambert, M.-L.; Suetens, C.; Savey, A.; Palomar, M.; Hiesmayr, M.; Morales, I.; Agodi, A.; Frank, U.; Mertens, K.; Schumacher, M.; et al. Clinical Outcomes of Health-Care-Associated Infections and Antimicrobial Resistance in Patients Admitted to European Intensive-Care Units: A Cohort Study. Lancet Infect. Dis. 2011, 11, 30–38. [Google Scholar] [CrossRef] [PubMed]
  18. Reynolds, D.; Kollef, M. The Epidemiology and Pathogenesis and Treatment of Pseudomonas Aeruginosa Infections: An Update. Drugs 2021, 81, 2117–2131. [Google Scholar] [CrossRef] [PubMed]
  19. Turkina, M.V.; Vikström, E. Bacteria-Host Crosstalk: Sensing of the Quorum in the Context of Pseudomonas Aeruginosa Infections. J. Innate Immun. 2019, 11, 263–279. [Google Scholar] [CrossRef]
  20. Pier, G.B.; Grout, M.; Zaidi, T.S. Cystic Fibrosis Transmembrane Conductance Regulator Is an Epithelial Cell Receptor for Clearance of Pseudomonas Aeruginosa from the Lung. Proc. Natl. Acad. Sci. USA 1997, 94, 12088–12093. [Google Scholar] [CrossRef] [Green Version]
  21. Jolly, A.L.; Takawira, D.; Oke, O.O.; Whiteside, S.A.; Chang, S.W.; Wen, E.R.; Quach, K.; Evans, D.J.; Fleiszig, S.M.J. Pseudomonas Aeruginosa-Induced Bleb-Niche Formation in Epithelial Cells Is Independent of Actinomyosin Contraction and Enhanced by Loss of Cystic Fibrosis Transmembrane-Conductance Regulator Osmoregulatory Function. mBio 2015, 6, e02533. [Google Scholar] [CrossRef] [Green Version]
  22. Del Barrio-Tofiño, E.; Zamorano, L.; Cortes-Lara, S.; López-Causapé, C.; Sánchez-Diener, I.; Cabot, G.; Bou, G.; Martínez-Martínez, L.; Oliver, A.; GEMARA-SEIMC/REIPI Pseudomonas study Group. Spanish Nationwide Survey on Pseudomonas Aeruginosa Antimicrobial Resistance Mechanisms and Epidemiology. J. Antimicrob. Chemother. 2019, 74, 1825–1835. [Google Scholar] [CrossRef]
  23. Recio, R.; Mancheño, M.; Viedma, E.; Villa, J.; Orellana, M.Á.; Lora-Tamayo, J.; Chaves, F. Predictors of Mortality in Bloodstream Infections Caused by Pseudomonas Aeruginosa and Impact of Antimicrobial Resistance and Bacterial Virulence. Antimicrob. Agents Chemother. 2020, 64, e01759-19. [Google Scholar] [CrossRef]
  24. Magiorakos, A.-P.; Srinivasan, A.; Carey, R.B.; Carmeli, Y.; Falagas, M.E.; Giske, C.G.; Harbarth, S.; Hindler, J.F.; Kahlmeter, G.; Olsson-Liljequist, B.; et al. Multidrug-Resistant, Extensively Drug-Resistant and Pandrug-Resistant Bacteria: An International Expert Proposal for Interim Standard Definitions for Acquired Resistance. Clin. Microbiol. Infect. Off. Publ. Eur. Soc. Clin. Microbiol. Infect. Dis. 2012, 18, 268–281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. CDC. Antibiotic Resistance Threats in the United States, 2019; U.S. Department of Health and Human Services, CDC: Atlanta, GA, USA, 2019. [CrossRef] [Green Version]
  26. Diggle, S.P.; Whiteley, M. Microbe Profile: Pseudomonas Aeruginosa: Opportunistic Pathogen and Lab Rat: This Article Is Part of the Microbe Profiles Collection. Microbiology 2020, 166, 30–33. [Google Scholar] [CrossRef] [PubMed]
  27. Gonçalves, T.; Vasconcelos, U. Colour Me Blue: The History and the Biotechnological Potential of Pyocyanin. Molecules 2021, 26, 927. [Google Scholar] [CrossRef]
  28. Michel-Briand, Y.; Baysse, C. The Pyocins of Pseudomonas Aeruginosa. Biochimie 2002, 84, 499–510. [Google Scholar] [CrossRef] [PubMed]
  29. Noto, M.J.; Burns, W.J.; Beavers, W.N.; Skaar, E.P. Mechanisms of Pyocyanin Toxicity and Genetic Determinants of Resistance in Staphylococcus Aureus. J. Bacteriol. 2017, 199, e00221-17. [Google Scholar] [CrossRef] [Green Version]
  30. Emmerich, R.; Löw, O. Bakteriolytische Enzyme als Ursache der erworbenen Immunität und die Heilung von Infectionskrankheiten durch dieselben. Z. Für Hyg. Infect. 1899, 31, 1–65. [Google Scholar] [CrossRef]
  31. Holloway, B.W. Genetic Recombination in Pseudomonas Aeruginosa. J. Gen. Microbiol. 1955, 13, 572–581. [Google Scholar] [CrossRef] [Green Version]
  32. Holloway, B.W.; Morgan, A.F. Genome Organization in Pseudomonas. Annu. Rev. Microbiol. 1986, 40, 79–105. [Google Scholar] [CrossRef]
  33. Holloway, B.W.; Krishnapillai, V.; Morgan, A.F. Chromosomal Genetics of Pseudomonas. Microbiol. Rev. 1979, 43, 73–102. [Google Scholar] [CrossRef]
  34. Holloway, B.W.; Römling, U.; Tümmler, B. Genomic Mapping of Pseudomonas Aeruginosa PAO. Microbiol. Read. Engl. 1994, 140 Pt 11, 2907–2929. [Google Scholar] [CrossRef] [Green Version]
  35. Römling, U.; Grothues, D.; Bautsch, W.; Tümmler, B. A Physical Genome Map of Pseudomonas Aeruginosa PAO. EMBO J. 1989, 8, 4081–4089. [Google Scholar] [CrossRef] [PubMed]
  36. Stover, C.K.; Pham, X.Q.; Erwin, A.L.; Mizoguchi, S.D.; Warrener, P.; Hickey, M.J.; Brinkman, F.S.; Hufnagle, W.O.; Kowalik, D.J.; Lagrou, M.; et al. Complete Genome Sequence of Pseudomonas Aeruginosa PAO1, an Opportunistic Pathogen. Nature 2000, 406, 959–964. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Oliveira, W.K.; Ferrarini, M.; Morello, L.G.; Faoro, H. Resistome Analysis of Bloodstream Infection Bacterial Genomes Reveals a Specific Set of Proteins Involved in Antibiotic Resistance and Drug Efflux. NAR Genom. Bioinforma 2020, 2, lqaa055. [Google Scholar] [CrossRef] [PubMed]
  38. Piddock, L.J.V. Multidrug-Resistance Efflux Pumps ? Not Just for Resistance. Nat. Rev. Microbiol. 2006, 4, 629–636. [Google Scholar] [CrossRef] [PubMed]
  39. Sun, J.; Deng, Z.; Yan, A. Bacterial Multidrug Efflux Pumps: Mechanisms, Physiology and Pharmacological Exploitations. Biochem. Biophys. Res. Commun. 2014, 453, 254–267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Hassan, K.A.; Liu, Q.; Elbourne, L.D.H.; Ahmad, I.; Sharples, D.; Naidu, V.; Chan, C.L.; Li, L.; Harborne, S.P.D.; Pokhrel, A.; et al. Pacing across the Membrane: The Novel PACE Family of Efflux Pumps Is Widespread in Gram-Negative Pathogens. Res. Microbiol. 2018, 169, 450–454. [Google Scholar] [CrossRef] [PubMed]
  41. Levy, S.B. Active Efflux Mechanisms for Antimicrobial Resistance. Antimicrob. Agents Chemother. 1992, 36, 695–703. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Nikaido, H. Prevention of Drug Access to Bacterial Targets: Permeability Barriers and Active Efflux. Science 1994, 264, 382–388. [Google Scholar] [CrossRef] [Green Version]
  43. Sanz-García, F.; Gil-Gil, T.; Laborda, P.; Ochoa-Sánchez, L.E.; Martínez, J.L.; Hernando-Amado, S. Coming from the Wild: Multidrug Resistant Opportunistic Pathogens Presenting a Primary, Not Human-Linked, Environmental Habitat. Int. J. Mol. Sci. 2021, 22, 8080. [Google Scholar] [CrossRef]
  44. Saier, M.H., Jr.; Paulsen, I.T.; Marek, K.S.; Pao, S.S.; Ronald, A.S.; Nikaido, H. Evolutionary Origins of Multidrug and Drug-Specific Efflux Pumps in Bacteria. FASEB J. 1998, 12, 265–274. [Google Scholar] [CrossRef] [Green Version]
  45. Shahini Shams Abadi, M.; Gholipour, A.; Hadi, N. The Highly Conserved Domain of RND Multidrug Efflux Pumps in Pathogenic Gram-Negative Bacteria. Cell. Mol. Biol. Noisy 2018, 64, 79–83. [Google Scholar] [CrossRef]
  46. McMurry, L.; Petrucci, R.E.; Levy, S.B. Active Efflux of Tetracycline Encoded by Four Genetically Different Tetracycline Resistance Determinants in Escherichia Coli. Proc. Natl. Acad. Sci. USA 1980, 77, 3974–3977. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Kodama, Y.; Inouye, S. Correlation of Susceptibility and Resistance of Twenty-Five Bacterial Strains by Analysis of MIC Database of Cephalosporins and Oxacephalosporins. J. Antibiot. 1997, 50, 246–255. [Google Scholar] [CrossRef]
  48. Yuan, L.; Zhai, Y.-J.; Wu, H.; Sun, H.-R.; He, Z.-P.; Wang, Y.-B.; Pan, Y.-S.; Kuang, N.-N.; Hu, G.-Z. Identification and Prevalence of RND Family Multidrug Efflux Pump OqxAB Genes in Enterococci Isolates from Swine Manure in China. J. Med. Microbiol. 2018, 67, 733–739. [Google Scholar] [CrossRef]
  49. Saier, M.H., Jr.; Tam, R.; Reizer, A.; Reizer, J. Two Novel Families of Bacterial Membrane Proteins Concerned with Nodulation, Cell Division and Transport. Mol. Microbiol. 1994, 11, 841–847. [Google Scholar] [CrossRef] [PubMed]
  50. Saier, M.H. Phylogenetic Approaches to the Identification and Characterization of Protein Families and Superfamilies. Microb. Comp. Genomics 1996, 1, 129–150. [Google Scholar] [CrossRef]
  51. Johnson, J.M.; Church, G.M. Alignment and Structure Prediction of Divergent Protein Families: Periplasmic and Outer Membrane Proteins of Bacterial Efflux Pumps11Edited by G. von Heijne. J. Mol. Biol. 1999, 287, 695–715. [Google Scholar] [CrossRef] [Green Version]
  52. Ma, D.; Cook, D.N.; Hearst, J.E.; Nikaido, H. Efflux Pumps and Drug Resistance in Gram-Negative Bacteria. Trends Microbiol. 1994, 2, 489–493. [Google Scholar] [CrossRef]
  53. Jamshidi, S.; Sutton, J.M.; Rahman, K.M. Mapping the Dynamic Functions and Structural Features of AcrB Efflux Pump Transporter Using Accelerated Molecular Dynamics Simulations. Sci. Rep. 2018, 8, 10470. [Google Scholar] [CrossRef]
  54. Wang, Z.; Fan, G.; Hryc, C.F.; Blaza, J.N.; Serysheva, I.I.; Schmid, M.F.; Chiu, W.; Luisi, B.F.; Du, D. An Allosteric Transport Mechanism for the AcrAB-TolC Multidrug Efflux Pump. eLife 2017, 6, e24905. [Google Scholar] [CrossRef]
  55. Yamaguchi, A.; Nakashima, R.; Sakurai, K. Structural Basis of RND-Type Multidrug Exporters. Front. Microbiol. 2015, 6, 327. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Zomot, E.; Yardeni, E.H.; Vargiu, A.V.; Tam, H.-K.; Malloci, G.; Ramaswamy, V.K.; Perach, M.; Ruggerone, P.; Pos, K.M.; Bibi, E. A New Critical Conformational Determinant of Multidrug Efflux by an MFS Transporter. J. Mol. Biol. 2018, 430, 1368–1385. [Google Scholar] [CrossRef] [PubMed]
  57. He, X.; Szewczyk, P.; Karyakin, A.; Evin, M.; Hong, W.-X.; Zhang, Q.; Chang, G. Structure of a Cation-Bound Multidrug and Toxic Compound Extrusion Transporter. Nature 2010, 467, 991–994. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Bolla, J.R.; Howes, A.C.; Fiorentino, F.; Robinson, C.V. Assembly and Regulation of the Chlorhexidine-Specific Efflux Pump AceI. Proc. Natl. Acad. Sci. USA 2020, 117, 17011–17018. [Google Scholar] [CrossRef] [PubMed]
  59. Al-Hamad, A.; Upton, M.; Burnie, J. Molecular Cloning and Characterization of SmrA, a Novel ABC Multidrug Efflux Pump from Stenotrophomonas Maltophilia. J. Antimicrob. Chemother. 2009, 64, 731–734. [Google Scholar] [CrossRef] [Green Version]
  60. Shcherbakov, A.A.; Hisao, G.; Mandala, V.S.; Thomas, N.E.; Soltani, M.; Salter, E.A.; Davis, J.H.; Henzler-Wildman, K.A.; Hong, M. Structure and Dynamics of the Drug-Bound Bacterial Transporter EmrE in Lipid Bilayers. Nat. Commun. 2021, 12, 172. [Google Scholar] [CrossRef]
  61. Marshall, N.J.; Piddock, L.J. Antibacterial Efflux Systems. Microbiol. Madr. Spain 1997, 13, 285–300. [Google Scholar]
  62. Sulavik, M.C.; Houseweart, C.; Cramer, C.; Jiwani, N.; Murgolo, N.; Greene, J.; DiDomenico, B.; Shaw, K.J.; Miller, G.H.; Hare, R.; et al. Antibiotic Susceptibility Profiles of Escherichia Coli Strains Lacking Multidrug Efflux Pump Genes. Antimicrob. Agents Chemother. 2001, 45, 1126–1136. [Google Scholar] [CrossRef] [Green Version]
  63. McDaniel, C.; Su, S.; Panmanee, W.; Lau, G.W.; Browne, T.; Cox, K.; Paul, A.T.; Ko, S.-H.B.; Mortensen, J.E.; Lam, J.S.; et al. A Putative ABC Transporter Permease Is Necessary for Resistance to Acidified Nitrite and EDTA in Pseudomonas Aeruginosa under Aerobic and Anaerobic Planktonic and Biofilm Conditions. Front. Microbiol. 2016, 7, 291. [Google Scholar] [CrossRef]
  64. Chen, L.; Duan, K. A PhoPQ-Regulated ABC Transporter System Exports Tetracycline in Pseudomonas Aeruginosa. Antimicrob. Agents Chemother. 2016, 60, 3016–3024. [Google Scholar] [CrossRef]
  65. Yero, D.; Díaz-Lobo, M.; Costenaro, L.; Conchillo-Solé, O.; Mayo, A.; Ferrer-Navarro, M.; Vilaseca, M.; Gibert, I.; Daura, X. The Pseudomonas Aeruginosa Substrate-Binding Protein Ttg2D Functions as a General Glycerophospholipid Transporter across the Periplasm. Commun. Biol. 2021, 4, 448. [Google Scholar] [CrossRef]
  66. Zhang, L.; Mah, T.-F. Involvement of a Novel Efflux System in Biofilm-Specific Resistance to Antibiotics. J. Bacteriol. 2008, 190, 4447–4452. [Google Scholar] [CrossRef] [Green Version]
  67. Poudyal, B.; Sauer, K. The ABC of Biofilm Drug Tolerance: The MerR-Like Regulator BrlR Is an Activator of ABC Transport Systems, with PA1874-77 Contributing to the Tolerance of Pseudomonas Aeruginosa Biofilms to Tobramycin. Antimicrob. Agents Chemother. 2018, 62, e01981-17. [Google Scholar] [CrossRef]
  68. Hall, C.W.; Zhang, L.; Mah, T.-F. PA3225 Is a Transcriptional Repressor of Antibiotic Resistance Mechanisms in Pseudomonas Aeruginosa. Antimicrob. Agents Chemother. 2017, 61, e02114-16. [Google Scholar] [CrossRef] [Green Version]
  69. Dulyayangkul, P.; Satapoomin, N.; Avison, M.B.; Charoenlap, N.; Vattanaviboon, P.; Mongkolsuk, S. Over-Expression of Hypochlorite Inducible Major Facilitator Superfamily (MFS) Pumps Reduces Antimicrobial Drug Susceptibility by Increasing the Production of MexXY Mediated by ArmZ in Pseudomonas Aeruginosa. Front. Microbiol. 2021, 11, 592153. [Google Scholar] [CrossRef]
  70. Bock, L.J.; Ferguson, P.M.; Clarke, M.; Pumpitakkul, V.; Wand, M.E.; Fady, P.-E.; Allison, L.; Fleck, R.A.; Shepherd, M.J.; Mason, A.J.; et al. Pseudomonas Aeruginosa Adapts to Octenidine via a Combination of Efflux and Membrane Remodelling. Commun. Biol. 2021, 4, 1058. [Google Scholar] [CrossRef]
  71. Bissonnette, L.; Champetier, S.; Buisson, J.P.; Roy, P.H. Characterization of the Nonenzymatic Chloramphenicol Resistance (CmlA) Gene of the In4 Integron of Tn1696: Similarity of the Product to Transmembrane Transport Proteins. J. Bacteriol. 1991, 173, 4493–4502. [Google Scholar] [CrossRef] [Green Version]
  72. Stokes, H.W.; Hall, R.M. Sequence Analysis of the Inducible Chloramphenicol Resistance Determinant in the Tn1696 Integron Suggests Regulation by Translational Attenuation. Plasmid 1991, 26, 10–19. [Google Scholar] [CrossRef]
  73. Li, X.-Z.; Poole, K.; Nikaido, H. Contributions of MexAB-OprM and an EmrE Homolog to Intrinsic Resistance of Pseudomonas Aeruginosa to Aminoglycosides and Dyes. Antimicrob. Agents Chemother. 2003, 47, 27–33. [Google Scholar] [CrossRef] [Green Version]
  74. Ninio, S.; Rotem, D.; Schuldiner, S. Functional Analysis of Novel Multidrug Transporters from Human Pathogens. J. Biol. Chem. 2001, 276, 48250–48256. [Google Scholar] [CrossRef]
  75. Duan, K.; Dammel, C.; Stein, J.; Rabin, H.; Surette, M.G. Modulation of Pseudomonas Aeruginosa Gene Expression by Host Microflora through Interspecies Communication. Mol. Microbiol. 2003, 50, 1477–1491. [Google Scholar] [CrossRef] [Green Version]
  76. He, G.-X.; Kuroda, T.; Mima, T.; Morita, Y.; Mizushima, T.; Tsuchiya, T. An H(+)-Coupled Multidrug Efflux Pump, PmpM, a Member of the MATE Family of Transporters, from Pseudomonas Aeruginosa. J. Bacteriol. 2004, 186, 262–265. [Google Scholar] [CrossRef] [Green Version]
  77. Zhao, J.; Hellwig, N.; Djahanschiri, B.; Khera, R.; Morgner, N.; Ebersberger, I.; Wang, J.; Michel, H. Assembly and Functional Role of PACE Transporter PA2880 from Pseudomonas Aeruginosa. Microbiol. Spectr. 2022, 10, e0145321. [Google Scholar] [CrossRef]
  78. Poole, K.; Krebes, K.; McNally, C.; Neshat, S. Multiple Antibiotic Resistance in Pseudomonas Aeruginosa: Evidence for Involvement of an Efflux Operon. J. Bacteriol. 1993, 175, 7363–7372. [Google Scholar] [CrossRef] [Green Version]
  79. Masuda, N.; Sakagawa, E.; Ohya, S.; Gotoh, N.; Tsujimoto, H.; Nishino, T. Substrate Specificities of MexAB-OprM, MexCD-OprJ, and MexXY-OprM Efflux Pumps in Pseudomonas Aeruginosa. Antimicrob. Agents Chemother. 2000, 44, 3322–3327. [Google Scholar] [CrossRef] [Green Version]
  80. Bialvaei, A.Z.; Rahbar, M.; Hamidi-Farahani, R.; Asgari, A.; Esmailkhani, A.; Mardani Dashti, Y.; Soleiman-Meigooni, S. Expression of RND Efflux Pumps Mediated Antibiotic Resistance in Pseudomonas Aeruginosa Clinical Strains. Microb. Pathog. 2021, 153, 104789. [Google Scholar] [CrossRef]
  81. Pesingi, P.V.; Singh, B.R.; Pesingi, P.K.; Bhardwaj, M.; Singh, S.V.; Kumawat, M.; Sinha, D.K.; Gandham, R.K. MexAB-OprM Efflux Pump of Pseudomonas Aeruginosa Offers Resistance to Carvacrol: A Herbal Antimicrobial Agent. Front. Microbiol. 2019, 10, 2664. [Google Scholar] [CrossRef]
  82. Poole, K.; Gotoh, N.; Tsujimoto, H.; Zhao, Q.; Wada, A.; Yamasaki, T.; Neshat, S.; Yamagishi, J.; Li, X.Z.; Nishino, T. Overexpression of the MexC-MexD-OprJ Efflux Operon in NfxB-Type Multidrug-Resistant Strains of Pseudomonas Aeruginosa. Mol. Microbiol. 1996, 21, 713–724. [Google Scholar] [CrossRef]
  83. Fraud, S.; Campigotto, A.J.; Chen, Z.; Poole, K. MexCD-OprJ Multidrug Efflux System of Pseudomonas Aeruginosa: Involvement in Chlorhexidine Resistance and Induction by Membrane-Damaging Agents Dependent upon the AlgU Stress Response Sigma Factor. Antimicrob. Agents Chemother. 2008, 52, 4478–4482. [Google Scholar] [CrossRef] [Green Version]
  84. Dong, N.; Zeng, Y.; Wang, Y.; Liu, C.; Lu, J.; Cai, C.; Liu, X.; Chen, Y.; Wu, Y.; Fang, Y.; et al. Distribution and Spread of the Mobilised RND Efflux Pump Gene Cluster TmexCD-ToprJ in Clinical Gram-Negative Bacteria: A Molecular Epidemiological Study. Lancet Microbe 2022, 3, e846–e856. [Google Scholar] [CrossRef]
  85. Lv, L.; Wan, M.; Wang, C.; Gao, X.; Yang, Q.; Partridge, S.R.; Wang, Y.; Zong, Z.; Doi, Y.; Shen, J.; et al. Emergence of a Plasmid-Encoded Resistance-Nodulation-Division Efflux Pump Conferring Resistance to Multiple Drugs, Including Tigecycline, in Klebsiella Pneumoniae. mBio 2020, 11, e02930-19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Köhler, T.; Michéa-Hamzehpour, M.; Henze, U.; Gotoh, N.; Curty, L.K.; Pechère, J.C. Characterization of MexE-MexF-OprN, a Positively Regulated Multidrug Efflux System of Pseudomonas Aeruginosa. Mol. Microbiol. 1997, 23, 345–354. [Google Scholar] [CrossRef] [PubMed]
  87. Aendekerk, S.; Ghysels, B.; Cornelis, P.; Baysse, C. Characterization of a New Efflux Pump, MexGHI-OpmD, from Pseudomonas Aeruginosa That Confers Resistance to Vanadium. Microbiol. Read. Engl. 2002, 148, 2371–2381. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Sakhtah, H.; Koyama, L.; Zhang, Y.; Morales, D.K.; Fields, B.L.; Price-Whelan, A.; Hogan, D.A.; Shepard, K.; Dietrich, L.E.P. The Pseudomonas Aeruginosa Efflux Pump MexGHI-OpmD Transports a Natural Phenazine That Controls Gene Expression and Biofilm Development. Proc. Natl. Acad. Sci. USA 2016, 113, E3538–E3547. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Sekiya, H.; Mima, T.; Morita, Y.; Kuroda, T.; Mizushima, T.; Tsuchiya, T. Functional Cloning and Characterization of a Multidrug Efflux Pump, MexHI-OpmD, from a Pseudomonas Aeruginosa Mutant. Antimicrob. Agents Chemother. 2003, 47, 2990–2992. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Chuanchuen, R.; Narasaki, C.T.; Schweizer, H.P. The MexJK Efflux Pump of Pseudomonas Aeruginosa Requires OprM for Antibiotic Efflux but Not for Efflux of Triclosan. J. Bacteriol. 2002, 184, 5036–5044. [Google Scholar] [CrossRef] [Green Version]
  91. Mima, T.; Sekiya, H.; Mizushima, T.; Kuroda, T.; Tsuchiya, T. Gene Cloning and Properties of the RND-Type Multidrug Efflux Pumps MexPQ-OpmE and MexMN-OprM from Pseudomonas Aeruginosa. Microbiol. Immunol. 2005, 49, 999–1002. [Google Scholar] [CrossRef]
  92. Ranjitkar, S.; Jones, A.K.; Mostafavi, M.; Zwirko, Z.; Iartchouk, O.; Barnes, S.W.; Walker, J.R.; Willis, T.W.; Lee, P.S.; Dean, C.R. Target (MexB)- and Efflux-Based Mechanisms Decreasing the Effectiveness of the Efflux Pump Inhibitor D13-9001 in Pseudomonas Aeruginosa PAO1: Uncovering a New Role for MexMN-OprM in Efflux of β-Lactams and a Novel Regulatory Circuit (MmnRS) Controlling MexMN Expression. Antimicrob. Agents Chemother. 2019, 63, e01718-18. [Google Scholar] [CrossRef] [Green Version]
  93. Li, Y.; Mima, T.; Komori, Y.; Morita, Y.; Kuroda, T.; Mizushima, T.; Tsuchiya, T. A New Member of the Tripartite Multidrug Efflux Pumps, MexVW–OprM, in Pseudomonas Aeruginosa. J. Antimicrob. Chemother. 2003, 52, 572–575. [Google Scholar] [CrossRef] [Green Version]
  94. Mine, T.; Morita, Y.; Kataoka, A.; Mizushima, T.; Tsuchiya, T. Expression in Escherichia Coli of a New Multidrug Efflux Pump, MexXY, from Pseudomonas Aeruginosa. Antimicrob. Agents Chemother. 1999, 43, 415–417. [Google Scholar] [CrossRef]
  95. Singh, M.; Sykes, E.M.E.; Li, Y.; Kumar, A. MexXY RND Pump of Pseudomonas Aeruginosa PA7 Effluxes Bi-Anionic β-Lactams Carbenicillin and Sulbenicillin When It Partners with the Outer Membrane Factor OprA but Not with OprM. Microbiol. Read. Engl. 2020, 166, 1095–1106. [Google Scholar] [CrossRef] [PubMed]
  96. Masuda, N.; Sakagawa, E.; Ohya, S.; Gotoh, N.; Tsujimoto, H.; Nishino, T. Contribution of the MexX-MexY-OprM Efflux System to Intrinsic Resistance in Pseudomonas Aeruginosa. Antimicrob. Agents Chemother. 2000, 44, 2242–2246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Mima, T.; Kohira, N.; Li, Y.; Sekiya, H.; Ogawa, W.; Kuroda, T.; Tsuchiya, T. Gene Cloning and Characteristics of the RND-Type Multidrug Efflux Pump MuxABC-OpmB Possessing Two RND Components in Pseudomonas Aeruginosa. Microbiol. Read. Engl. 2009, 155, 3509–3517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Mima, T.; Joshi, S.; Gomez-Escalada, M.; Schweizer, H.P. Identification and Characterization of TriABC-OpmH, a Triclosan Efflux Pump of Pseudomonas Aeruginosa Requiring Two Membrane Fusion Proteins. J. Bacteriol. 2007, 189, 7600–7609. [Google Scholar] [CrossRef] [Green Version]
  99. Winsor, G.L.; Griffiths, E.J.; Lo, R.; Dhillon, B.K.; Shay, J.A.; Brinkman, F.S.L. Enhanced Annotations and Features for Comparing Thousands of Pseudomonas Genomes in the Pseudomonas Genome Database. Nucleic Acids Res. 2016, 44, D646–D653. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Sanz-García, F.; Hernando-Amado, S.; López-Causapé, C.; Oliver, A.; Martínez, J.L. Low Ciprofloxacin Concentrations Select Multidrug-Resistant Mutants Overproducing Efflux Pumps in Clinical Isolates of Pseudomonas Aeruginosa. Microbiol. Spectr. 2022, 10, e0072322. [Google Scholar] [CrossRef]
  101. Langevin, A.M.; Dunlop, M.J. Stress Introduction Rate Alters the Benefit of AcrAB-TolC Efflux Pumps. J. Bacteriol. 2018, 200, e00525-17. [Google Scholar] [CrossRef] [Green Version]
  102. Adamiak, J.W.; Jhawar, V.; Bonifay, V.; Chandler, C.E.; Leus, I.V.; Ernst, R.K.; Schweizer, H.P.; Zgurskaya, H.I. Loss of RND-Type Multidrug Efflux Pumps Triggers Iron Starvation and Lipid A Modifications in Pseudomonas Aeruginosa. Antimicrob. Agents Chemother. 2021, 65, e0059221. [Google Scholar] [CrossRef]
  103. Nag, A.; Mehra, S. Involvement of the SCO3366 Efflux Pump from S. Coelicolor in Rifampicin Resistance and Its Regulation by a TetR Regulator. Appl. Microbiol. Biotechnol. 2022, 106, 2175–2190. [Google Scholar] [CrossRef]
  104. Yao, X.; Tao, F.; Zhang, K.; Tang, H.; Xu, P. Multiple Roles for Two Efflux Pumps in the Polycyclic Aromatic Hydrocarbon-Degrading Pseudomonas Putida Strain B6-2 (DSM 28064). Appl. Environ. Microbiol. 2017, 83, e01882-17. [Google Scholar] [CrossRef]
  105. Zwama, M.; Yamaguchi, A.; Nishino, K. Phylogenetic and Functional Characterisation of the Haemophilus Influenzae Multidrug Efflux Pump AcrB. Commun. Biol. 2019, 2, 340. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Mehla, J.; Malloci, G.; Mansbach, R.; López, C.A.; Tsivkovski, R.; Haynes, K.; Leus, I.V.; Grindstaff, S.B.; Cascella, R.H.; D’Cunha, N.; et al. Predictive Rules of Efflux Inhibition and Avoidance in Pseudomonas Aeruginosa. mBio 2021, 12, e02785-20. [Google Scholar] [CrossRef] [PubMed]
  107. Elkins, C.A.; Nikaido, H. Substrate Specificity of the RND-Type Multidrug Efflux Pumps AcrB and AcrD of Escherichia Coli Is Determined Predominantly by Two Large Periplasmic Loops. J. Bacteriol. 2002, 184, 6490–6498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Nishino, K.; Yamaguchi, A. Analysis of a Complete Library of Putative Drug Transporter Genes in Escherichia Coli. J. Bacteriol. 2001, 183, 5803–5812. [Google Scholar] [CrossRef] [Green Version]
  109. Laudy, A.E. Non-Antibiotics, Efflux Pumps and Drug Resistance of Gram-Negative Rods. Pol. J. Microbiol. 2018, 67, 129–135. [Google Scholar] [CrossRef] [Green Version]
  110. Ramos, J.L.; Duque, E.; Gallegos, M.-T.; Godoy, P.; Ramos-Gonzalez, M.I.; Rojas, A.; Teran, W.; Segura, A. Mechanisms of Solvent Tolerance in Gram-Negative Bacteria. Annu. Rev. Microbiol. 2002, 56, 743–768. [Google Scholar] [CrossRef]
  111. Garratt, I.; Aranega-Bou, P.; Sutton, J.M.; Moore, G.; Wand, M.E. Long-Term Exposure to Octenidine in a Simulated Sink Trap Environment Results in Selection of Pseudomonas Aeruginosa, Citrobacter, and Enterobacter Isolates with Mutations in Efflux Pump Regulators. Appl. Environ. Microbiol. 2021, 87, e00210-21. [Google Scholar] [CrossRef]
  112. Lennen, R.M.; Politz, M.G.; Kruziki, M.A.; Pfleger, B.F. Identification of Transport Proteins Involved in Free Fatty Acid Efflux in Escherichia Coli. J. Bacteriol. 2013, 195, 135–144. [Google Scholar] [CrossRef] [Green Version]
  113. Sabrin, A.; Gioe, B.W.; Gupta, A.; Grove, A. An EmrB Multidrug Efflux Pump in Burkholderia Thailandensis with Unexpected Roles in Antibiotic Resistance. J. Biol. Chem. 2019, 294, 1891–1903. [Google Scholar] [CrossRef] [Green Version]
  114. Ramos, J.L.; Martínez-Bueno, M.; Molina-Henares, A.J.; Terán, W.; Watanabe, K.; Zhang, X.; Gallegos, M.T.; Brennan, R.; Tobes, R. The TetR Family of Transcriptional Repressors. Microbiol. Mol. Biol. Rev. MMBR 2005, 69, 326–356. [Google Scholar] [CrossRef]
  115. Liu, Q.; Hassan, K.A.; Ashwood, H.E.; Gamage, H.K.A.H.; Li, L.; Mabbutt, B.C.; Paulsen, I.T. Regulation of the AceI Multidrug Efflux Pump Gene in Acinetobacter Baumannii. J. Antimicrob. Chemother. 2018, 73, 1492–1500. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Stubenrauch, C.J.; Bamert, R.S.; Wang, J.; Lithgow, T. A Noncanonical Chaperone Interacts with Drug Efflux Pumps during Their Assembly into Bacterial Outer Membranes. PLoS Biol. 2022, 20, e3001523. [Google Scholar] [CrossRef] [PubMed]
  117. Ma, D.; Cook, D.N.; Alberti, M.; Pon, N.G.; Nikaido, H.; Hearst, J.E. Molecular Cloning and Characterization of AcrA and AcrE Genes of Escherichia Coli. J. Bacteriol. 1993, 175, 6299–6313. [Google Scholar] [CrossRef] [Green Version]
  118. Schmalstieg, A.M.; Srivastava, S.; Belkaya, S.; Deshpande, D.; Meek, C.; Leff, R.; van Oers, N.S.C.; Gumbo, T. The Antibiotic Resistance Arrow of Time: Efflux Pump Induction Is a General First Step in the Evolution of Mycobacterial Drug Resistance. Antimicrob. Agents Chemother. 2012, 56, 4806–4815. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Murray, J.L.; Kwon, T.; Marcotte, E.M.; Whiteley, M. Intrinsic Antimicrobial Resistance Determinants in the Superbug Pseudomonas Aeruginosa. mBio 2015, 6, e01603-01615. [Google Scholar] [CrossRef] [Green Version]
  120. Chevalier, S.; Bouffartigues, E.; Bodilis, J.; Maillot, O.; Lesouhaitier, O.; Feuilloley, M.G.J.; Orange, N.; Dufour, A.; Cornelis, P. Structure, Function and Regulation of Pseudomonas Aeruginosa Porins. FEMS Microbiol. Rev. 2017, 41, 698–722. [Google Scholar] [CrossRef] [Green Version]
  121. Valot, B.; Guyeux, C.; Rolland, J.Y.; Mazouzi, K.; Bertrand, X.; Hocquet, D. What It Takes to Be a Pseudomonas Aeruginosa? The Core Genome of the Opportunistic Pathogen Updated. PLoS ONE 2015, 10, e0126468. [Google Scholar] [CrossRef]
  122. Venter, H.; Mowla, R.; Ohene-Agyei, T.; Ma, S. RND-Type Drug Efflux Pumps from Gram-Negative Bacteria: Molecular Mechanism and Inhibition. Front. Microbiol. 2015, 6, 377. [Google Scholar] [CrossRef]
  123. Daury, L.; Orange, F.; Taveau, J.-C.; Verchère, A.; Monlezun, L.; Gounou, C.; Marreddy, R.K.R.; Picard, M.; Broutin, I.; Pos, K.M.; et al. Tripartite Assembly of RND Multidrug Efflux Pumps. Nat. Commun. 2016, 7, 10731. [Google Scholar] [CrossRef] [Green Version]
  124. Du, D.; Wang-Kan, X.; Neuberger, A.; van Veen, H.W.; Pos, K.M.; Piddock, L.J.V.; Luisi, B.F. Multidrug Efflux Pumps: Structure, Function and Regulation. Nat. Rev. Microbiol. 2018, 16, 523–539. [Google Scholar] [CrossRef]
  125. López-Causapé, C.; Sommer, L.M.; Cabot, G.; Rubio, R.; Ocampo-Sosa, A.A.; Johansen, H.K.; Figuerola, J.; Cantón, R.; Kidd, T.J.; Molin, S.; et al. Evolution of the Pseudomonas Aeruginosa Mutational Resistome in an International Cystic Fibrosis Clone. Sci. Rep. 2017, 7, 5555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Colclough, A.L.; Alav, I.; Whittle, E.E.; Pugh, H.L.; Darby, E.M.; Legood, S.W.; McNeil, H.E.; Blair, J.M. RND Efflux Pumps in Gram-Negative Bacteria; Regulation, Structure and Role in Antibiotic Resistance. Future Microbiol. 2020, 15, 143–157. [Google Scholar] [CrossRef] [PubMed]
  127. Cabot, G.; Ocampo-Sosa, A.A.; Tubau, F.; Macia, M.D.; Rodríguez, C.; Moya, B.; Zamorano, L.; Suárez, C.; Peña, C.; Martínez-Martínez, L.; et al. Overexpression of AmpC and Efflux Pumps in Pseudomonas Aeruginosa Isolates from Bloodstream Infections: Prevalence and Impact on Resistance in a Spanish Multicenter Study. Antimicrob. Agents Chemother. 2011, 55, 1906–1911. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Jones, A.K.; Caughlan, R.E.; Woods, A.L.; Uehara, K.; Xie, L.; Barnes, S.W.; Walker, J.R.; Thompson, K.V.; Ranjitkar, S.; Lee, P.S.; et al. Mutations Reducing In Vitro Susceptibility to Novel LpxC Inhibitors in Pseudomonas Aeruginosa and Interplay of Efflux and Nonefflux Mechanisms. Antimicrob. Agents Chemother. 2019, 64, e01490-19. [Google Scholar] [CrossRef]
  129. Krause, K.M.; Haglund, C.M.; Hebner, C.; Serio, A.W.; Lee, G.; Nieto, V.; Cohen, F.; Kane, T.R.; Machajewski, T.D.; Hildebrandt, D.; et al. Potent LpxC Inhibitors with In Vitro Activity against Multidrug-Resistant Pseudomonas Aeruginosa. Antimicrob. Agents Chemother. 2019, 63, e00977-19. [Google Scholar] [CrossRef]
  130. Ito, A.; Sato, T.; Ota, M.; Takemura, M.; Nishikawa, T.; Toba, S.; Kohira, N.; Miyagawa, S.; Ishibashi, N.; Matsumoto, S.; et al. In Vitro Antibacterial Properties of Cefiderocol, a Novel Siderophore Cephalosporin, against Gram-Negative Bacteria. Antimicrob. Agents Chemother. 2018, 62, e01454-17. [Google Scholar] [CrossRef] [Green Version]
  131. Amieva, R.; Gil-Gil, T.; Martínez, J.L.; Alcalde-Rico, M. The MexJK Multidrug Efflux Pump Is Not Involved in Acquired or Intrinsic Antibiotic Resistance in Pseudomonas Aeruginosa, but Modulates the Bacterial Quorum Sensing Response. Int. J. Mol. Sci. 2022, 23, 7492. [Google Scholar] [CrossRef]
  132. López, C.A.; Travers, T.; Pos, K.M.; Zgurskaya, H.I.; Gnanakaran, S. Dynamics of Intact MexAB-OprM Efflux Pump: Focusing on the MexA-OprM Interface. Sci. Rep. 2017, 7, 16521. [Google Scholar] [CrossRef] [Green Version]
  133. Tsutsumi, K.; Yonehara, R.; Ishizaka-Ikeda, E.; Miyazaki, N.; Maeda, S.; Iwasaki, K.; Nakagawa, A.; Yamashita, E. Structures of the Wild-Type MexAB-OprM Tripartite Pump Reveal Its Complex Formation and Drug Efflux Mechanism. Nat. Commun. 2019, 10, 1520. [Google Scholar] [CrossRef] [Green Version]
  134. Poole, K.; Tetro, K.; Zhao, Q.; Neshat, S.; Heinrichs, D.E.; Bianco, N. Expression of the Multidrug Resistance Operon MexA-MexB-OprM in Pseudomonas Aeruginosa: MexR Encodes a Regulator of Operon Expression. Antimicrob. Agents Chemother. 1996, 40, 2021–2028. [Google Scholar] [CrossRef]
  135. Cao, L.; Srikumar, R.; Poole, K. MexAB-OprM Hyperexpression in NalC-Type Multidrug-Resistant Pseudomonas Aeruginosa: Identification and Characterization of the NalC Gene Encoding a Repressor of PA3720-PA3719. Mol. Microbiol. 2004, 53, 1423–1436. [Google Scholar] [CrossRef] [PubMed]
  136. Morita, Y.; Cao, L.; Gould, V.C.; Avison, M.B.; Poole, K. NalD Encodes a Second Repressor of the MexAB-OprM Multidrug Efflux Operon of Pseudomonas Aeruginosa. J. Bacteriol. 2006, 188, 8649–8654. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Tafti, F.A.; Eslami, G.; Zandi, H.; Barzegar, K. Mutations in Nalc Gene of Mex AB-OprM Efflux Pump in Carbapenem Resistant Pseudomonas Aeruginosa Isolated from Burn Wounds in Yazd, Iran. Iran. J. Microbiol. 2020, 12, 32–36. [Google Scholar] [PubMed]
  138. Ziha-Zarifi, I.; Llanes, C.; Köhler, T.; Pechere, J.-C.; Plesiat, P. In Vivo Emergence of Multidrug-Resistant Mutants of Pseudomonas Aeruginosa Overexpressing the Active Efflux System MexA-MexB-OprM. Antimicrob. Agents Chemother. 1999, 43, 287–291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Horna, G.; López, M.; Guerra, H.; Saénz, Y.; Ruiz, J. Interplay between MexAB-OprM and MexEF-OprN in Clinical Isolates of Pseudomonas Aeruginosa. Sci. Rep. 2018, 8, 16463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Choudhury, D.; Ghose, A.; Ghosh, A.; Dhar Chanda, D.; Das Talukdar, A.; Dutta Choudhury, M.; Paul, D.; Maurya, A.P.; Chakravarty, A.; Chakravorty, A.; et al. Premature Termination of MexR Leads to Overexpression of MexAB-OprM Efflux Pump in Pseudomonas Aeruginosa in a Tertiary Referral Hospital in India. PLoS ONE 2016, 11, e0149156. [Google Scholar] [CrossRef]
  141. Boutoille, D.; Corvec, S.; Caroff, N.; Giraudeau, C.; Espaze, E.; Caillon, J.; Plésiat, P.; Reynaud, A. Detection of an IS21 Insertion Sequence in the MexR Gene of Pseudomonas Aeruginosa Increasing Beta-Lactam Resistance. FEMS Microbiol. Lett. 2004, 230, 143–146. [Google Scholar] [CrossRef]
  142. Ma, Z.; Xu, C.; Zhang, X.; Wang, D.; Pan, X.; Liu, H.; Zhu, G.; Bai, F.; Cheng, Z.; Wu, W.; et al. A MexR Mutation Which Confers Aztreonam Resistance to Pseudomonas Aeruginosa. Front. Microbiol. 2021, 12, 659808. [Google Scholar] [CrossRef]
  143. Aguilar-Rodea, P.; Zúñiga, G.; Cerritos, R.; Rodríguez-Espino, B.A.; Gomez-Ramirez, U.; Nolasco-Romero, C.G.; López-Marceliano, B.; Rodea, G.E.; Mendoza-Elizalde, S.; Reyes-López, A.; et al. Nucleotide Substitutions in the MexR, NalC and NalD Regulator Genes of the MexAB-OprM Efflux Pump Are Maintained in Pseudomonas Aeruginosa Genetic Lineages. PLoS ONE 2022, 17, e0266742. [Google Scholar] [CrossRef]
  144. Suresh, M.; Nithya, N.; Jayasree, P.R.; Vimal, K.P.; Manish Kumar, P.R. Mutational Analyses of Regulatory Genes, MexR, NalC, NalD and MexZ of MexAB-OprM and MexXY Operons, in Efflux Pump Hyperexpressing Multidrug-Resistant Clinical Isolates of Pseudomonas Aeruginosa. World J. Microbiol. Biotechnol. 2018, 34, 83. [Google Scholar] [CrossRef]
  145. Beig, M.; Taheri, M.; Arabestani, M.R. Expression of MexAB-OprM Efflux Pump and OprD Porin in Carbapenemase Producing Pseudomonas Aeruginosa Clinical Isolates. Gene Rep. 2020, 20, 100744. [Google Scholar] [CrossRef]
  146. Glen, K.A.; Lamont, I.L. β-Lactam Resistance in Pseudomonas Aeruginosa: Current Status, Future Prospects. Pathogens 2021, 10, 1638. [Google Scholar] [CrossRef] [PubMed]
  147. Lodge, J.M.; Minchin, S.D.; Piddock, L.J.; Busby, J.W. Cloning, Sequencing and Analysis of the Structural Gene and Regulatory Region of the Pseudomonas Aeruginosa Chromosomal AmpC Beta-Lactamase. Biochem. J. 1990, 272, 627–631. [Google Scholar] [CrossRef] [PubMed]
  148. Grosjean, M.; Tazrout, S.; Bour, M.; Triponey, P.; Muller, C.; Jeannot, K.; Plésiat, P. Reassessment of the Cooperativity between Efflux System MexAB-OprM and Cephalosporinase AmpC in the Resistance of Pseudomonas Aeruginosa to β-Lactams. J. Antimicrob. Chemother. 2021, 76, 536–539. [Google Scholar] [CrossRef] [PubMed]
  149. Roy, P.H.; Tetu, S.G.; Larouche, A.; Elbourne, L.; Tremblay, S.; Ren, Q.; Dodson, R.; Harkins, D.; Shay, R.; Watkins, K.; et al. Complete Genome Sequence of the Multiresistant Taxonomic Outlier Pseudomonas Aeruginosa PA7. PLoS ONE 2010, 5, e8842. [Google Scholar] [CrossRef]
  150. Morita, Y.; Tomida, J.; Kawamura, Y. Primary Mechanisms Mediating Aminoglycoside Resistance in the Multidrug-Resistant Pseudomonas Aeruginosa Clinical Isolate PA7. Microbiol. Read. Engl. 2012, 158, 1071–1083. [Google Scholar] [CrossRef]
  151. Issa, K.H.B.; Phan, G.; Broutin, I. Functional Mechanism of the Efflux Pumps Transcription Regulators From Pseudomonas Aeruginosa Based on 3D Structures. Front. Mol. Biosci. 2018, 5, 57. [Google Scholar] [CrossRef] [Green Version]
  152. Muller, C.; Plésiat, P.; Jeannot, K. A Two-Component Regulatory System Interconnects Resistance to Polymyxins, Aminoglycosides, Fluoroquinolones, and β-Lactams in Pseudomonas Aeruginosa. Antimicrob. Agents Chemother. 2011, 55, 1211–1221. [Google Scholar] [CrossRef] [Green Version]
  153. Hay, T.; Fraud, S.; Lau, C.H.-F.; Gilmour, C.; Poole, K. Antibiotic Inducibility of the MexXY Multidrug Efflux Operon of Pseudomonas Aeruginosa: Involvement of the MexZ Anti-Repressor ArmZ. PLoS ONE 2013, 8, e56858. [Google Scholar] [CrossRef]
  154. Kawalek, A.; Modrzejewska, M.; Zieniuk, B.; Bartosik, A.A.; Jagura-Burdzy, G. Interaction of ArmZ with the DNA-Binding Domain of MexZ Induces Expression of MexXY Multidrug Efflux Pump Genes and Antimicrobial Resistance in Pseudomonas Aeruginosa. Antimicrob. Agents Chemother. 2019, 63, e01199-19. [Google Scholar] [CrossRef]
  155. Guénard, S.; Muller, C.; Monlezun, L.; Benas, P.; Broutin, I.; Jeannot, K.; Plésiat, P. Multiple Mutations Lead to MexXY-OprM-Dependent Aminoglycoside Resistance in Clinical Strains of Pseudomonas Aeruginosa. Antimicrob. Agents Chemother. 2014, 58, 221–228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Seupt, A.; Schniederjans, M.; Tomasch, J.; Häussler, S. Expression of the MexXY Aminoglycoside Efflux Pump and Presence of an Aminoglycoside-Modifying Enzyme in Clinical Pseudomonas Aeruginosa Isolates Are Highly Correlated. Antimicrob. Agents Chemother. 2020, 65, e01166-20. [Google Scholar] [CrossRef] [PubMed]
  157. López-Causapé, C.; Rubio, R.; Cabot, G.; Oliver, A. Evolution of the Pseudomonas Aeruginosa Aminoglycoside Mutational Resistome In Vitro and in the Cystic Fibrosis Setting. Antimicrob. Agents Chemother. 2018, 62, e02583-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Singh, M.; Yau, Y.C.W.; Wang, S.; Waters, V.; Kumar, A. MexXY Efflux Pump Overexpression and Aminoglycoside Resistance in Cystic Fibrosis Isolates of Pseudomonas Aeruginosa from Chronic Infections. Can. J. Microbiol. 2017, 63, 929–938. [Google Scholar] [CrossRef] [Green Version]
  159. Westbrock-Wadman, S.; Sherman, D.R.; Hickey, M.J.; Coulter, S.N.; Zhu, Y.Q.; Warrener, P.; Nguyen, L.Y.; Shawar, R.M.; Folger, K.R.; Stover, C.K. Characterization of a Pseudomonas Aeruginosa Efflux Pump Contributing to Aminoglycoside Impermeability. Antimicrob. Agents Chemother. 1999, 43, 2975–2983. [Google Scholar] [CrossRef] [Green Version]
  160. Vettoretti, L.; Plésiat, P.; Muller, C.; El Garch, F.; Phan, G.; Attrée, I.; Ducruix, A.; Llanes, C. Efflux Unbalance in Pseudomonas Aeruginosa Isolates from Cystic Fibrosis Patients. Antimicrob. Agents Chemother. 2009, 53, 1987–1997. [Google Scholar] [CrossRef] [Green Version]
  161. Shigemura, K.; Osawa, K.; Kato, A.; Tokimatsu, I.; Arakawa, S.; Shirakawa, T.; Fujisawa, M. Association of Overexpression of Efflux Pump Genes with Antibiotic Resistance in Pseudomonas Aeruginosa Strains Clinically Isolated from Urinary Tract Infection Patients. J. Antibiot. (Tokyo) 2015, 68, 568–572. [Google Scholar] [CrossRef]
  162. Jeannot, K.; Elsen, S.; Köhler, T.; Attree, I.; van Delden, C.; Plésiat, P. Resistance and Virulence of Pseudomonas Aeruginosa Clinical Strains Overproducing the MexCD-OprJ Efflux Pump. Antimicrob. Agents Chemother. 2008, 52, 2455–2462. [Google Scholar] [CrossRef] [Green Version]
  163. Gomis-Font, M.A.; Pitart, C.; Del Barrio-Tofiño, E.; Zboromyrska, Y.; Cortes-Lara, S.; Mulet, X.; Marco, F.; Vila, J.; López-Causapé, C.; Oliver, A. Emergence of Resistance to Novel Cephalosporin-β-Lactamase Inhibitor Combinations through the Modification of the Pseudomonas Aeruginosa MexCD-OprJ Efflux Pump. Antimicrob. Agents Chemother. 2021, 65, e0008921. [Google Scholar] [CrossRef]
  164. Morita, Y.; Komori, Y.; Mima, T.; Kuroda, T.; Mizushima, T.; Tsuchiya, T. Construction of a Series of Mutants Lacking All of the Four Major Mex Operons for Multidrug Efflux Pumps or Possessing Each One of the Operons from Pseudomonas Aeruginosa PAO1: MexCD-OprJ Is an Inducible Pump. FEMS Microbiol. Lett. 2001, 202, 139–143. [Google Scholar] [CrossRef]
  165. Srikumar, R.; Kon, T.; Gotoh, N.; Poole, K. Expression of Pseudomonas Aeruginosa Multidrug Efflux Pumps MexA-MexB-OprM and MexC-MexD-OprJ in a Multidrug-Sensitive Escherichia Coli Strain. Antimicrob. Agents Chemother. 1998, 42, 65–71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Masuda, N.; Gotoh, N.; Ohya, S.; Nishino, T. Quantitative Correlation between Susceptibility and OprJ Production in NfxB Mutants of Pseudomonas Aeruginosa. Antimicrob. Agents Chemother. 1996, 40, 909–913. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Mao, W.; Warren, M.S.; Black, D.S.; Satou, T.; Murata, T.; Nishino, T.; Gotoh, N.; Lomovskaya, O. On the Mechanism of Substrate Specificity by Resistance Nodulation Division (RND)-Type Multidrug Resistance Pumps: The Large Periplasmic Loops of MexD from Pseudomonas Aeruginosa Are Involved in Substrate Recognition. Mol. Microbiol. 2002, 46, 889–901. [Google Scholar] [CrossRef] [PubMed]
  168. Gotoh, N.; Tsujimoto, H.; Tsuda, M.; Okamoto, K.; Nomura, A.; Wada, T.; Nakahashi, M.; Nishino, T. Characterization of the MexC-MexD-OprJ Multidrug Efflux System in DeltamexA-MexB-OprM Mutants of Pseudomonas Aeruginosa. Antimicrob. Agents Chemother. 1998, 42, 1938–1943. [Google Scholar] [CrossRef] [Green Version]
  169. Mulet, X.; Moyá, B.; Juan, C.; Macià, M.D.; Pérez, J.L.; Blázquez, J.; Oliver, A. Antagonistic Interactions of Pseudomonas Aeruginosa Antibiotic Resistance Mechanisms in Planktonic but Not Biofilm Growth▿. Antimicrob. Agents Chemother. 2011, 55, 4560–4568. [Google Scholar] [CrossRef] [Green Version]
  170. Martínez-Ramos, I.; Mulet, X.; Moyá, B.; Barbier, M.; Oliver, A.; Albertí, S. Overexpression of MexCD-OprJ Reduces Pseudomonas Aeruginosa Virulence by Increasing Its Susceptibility to Complement-Mediated Killing. Antimicrob. Agents Chemother. 2014, 58, 2426–2429. [Google Scholar] [CrossRef] [Green Version]
  171. Cazares, A.; Moore, M.P.; Hall, J.P.J.; Wright, L.L.; Grimes, M.; Emond-Rhéault, J.-G.; Pongchaikul, P.; Santanirand, P.; Levesque, R.C.; Fothergill, J.L.; et al. A Megaplasmid Family Driving Dissemination of Multidrug Resistance in Pseudomonas. Nat. Commun. 2020, 11, 1370. [Google Scholar] [CrossRef] [Green Version]
  172. Köhler, T.; Epp, S.F.; Curty, L.K.; Pechère, J.-C. Characterization of MexT, the Regulator of the MexE-MexF-OprN Multidrug Efflux System of Pseudomonas Aeruginosa. J. Bacteriol. 1999, 181, 6300–6305. [Google Scholar] [CrossRef] [Green Version]
  173. Sobel, M.L.; Neshat, S.; Poole, K. Mutations in PA2491 (MexS) Promote MexT-Dependent MexEF-OprN Expression and Multidrug Resistance in a Clinical Strain of Pseudomonas Aeruginosa. J. Bacteriol. 2005, 187, 1246–1253. [Google Scholar] [CrossRef] [Green Version]
  174. Maseda, H.; Saito, K.; Nakajima, A.; Nakae, T. Variation of the MexT Gene, a Regulator of the MexEF-OprN Efflux Pump Expression in Wild-Type Strains of Pseudomonas Aeruginosa. FEMS Microbiol. Lett. 2000, 192, 107–112. [Google Scholar] [CrossRef]
  175. Fetar, H.; Gilmour, C.; Klinoski, R.; Daigle, D.M.; Dean, C.R.; Poole, K. MexEF-OprN Multidrug Efflux Operon of Pseudomonas Aeruginosa: Regulation by the MexT Activator in Response to Nitrosative Stress and Chloramphenicol. Antimicrob. Agents Chemother. 2011, 55, 508–514. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Morita, Y.; Tomida, J.; Kawamura, Y. Efflux-Mediated Fluoroquinolone Resistance in the Multidrug-Resistant Pseudomonas Aeruginosa Clinical Isolate PA7: Identification of a Novel MexS Variant Involved in Upregulation of the MexEF-OprN Multidrug Efflux Operon. Front. Microbiol. 2015, 6, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Richardot, C.; Juarez, P.; Jeannot, K.; Patry, I.; Plésiat, P.; Llanes, C. Amino Acid Substitutions Account for Most MexS Alterations in Clinical NfxC Mutants of Pseudomonas Aeruginosa. Antimicrob. Agents Chemother. 2016, 60, 2302–2310. [Google Scholar] [CrossRef] [Green Version]
  178. Iftikhar, A.; Asif, A.; Manzoor, A.; Azeem, M.; Sarwar, G.; Rashid, N.; Qaisar, U. Mutation in PvcABCD Operon of Pseudomonas Aeruginosa Modulates MexEF-OprN Efflux System and Hence Resistance to Chloramphenicol and Ciprofloxacin. Microb. Pathog. 2020, 149, 104491. [Google Scholar] [CrossRef] [PubMed]
  179. Balasubramanian, D.; Schneper, L.; Merighi, M.; Smith, R.; Narasimhan, G.; Lory, S.; Mathee, K. The Regulatory Repertoire of Pseudomonas Aeruginosa AmpC SS-Lactamase Regulator AmpR Includes Virulence Genes. PLoS ONE 2012, 7, e34067. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Westfall, L.W.; Carty, N.L.; Layland, N.; Kuan, P.; Colmer-Hamood, J.A.; Hamood, A.N. MvaT Mutation Modifies the Expression of the Pseudomonas Aeruginosa Multidrug Efflux Operon MexEF-OprN. FEMS Microbiol. Lett. 2006, 255, 247–254. [Google Scholar] [CrossRef] [Green Version]
  181. Ochs, M.M.; McCusker, M.P.; Bains, M.; Hancock, R.E.W. Negative Regulation of the Pseudomonas Aeruginosa Outer Membrane Porin OprD Selective for Imipenem and Basic Amino Acids. Antimicrob. Agents Chemother. 1999, 43, 1085–1090. [Google Scholar] [CrossRef] [Green Version]
  182. Sherrard, L.J.; Wee, B.A.; Duplancic, C.; Ramsay, K.A.; Dave, K.A.; Ballard, E.; Wainwright, C.E.; Grimwood, K.; Sidjabat, H.E.; Whiley, D.M.; et al. Emergence and Impact of OprD Mutations in Pseudomonas Aeruginosa Strains in Cystic Fibrosis. J. Cyst. Fibros. Off. J. Eur. Cyst. Fibros. Soc. 2022, 21, e35–e43. [Google Scholar] [CrossRef]
  183. Mlynarcik, P.; Kolar, M. Molecular Mechanisms of Polymyxin Resistance and Detection of Mcr Genes. Biomed. Pap. Med. Fac. Univ. Palacky Olomouc Czechoslov. 2019, 163, 28–38. [Google Scholar] [CrossRef] [Green Version]
  184. Yin, R.; Cheng, J.; Wang, J.; Li, P.; Lin, J. Treatment of Pseudomonas Aeruginosa Infectious Biofilms: Challenges and Strategies. Front. Microbiol. 2022, 13, 955286. [Google Scholar] [CrossRef]
  185. Nikaido, H.; Pagès, J.-M. Broad-Specificity Efflux Pumps and Their Role in Multidrug Resistance of Gram-Negative Bacteria. FEMS Microbiol. Rev. 2012, 36, 340–363. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Reza, A.; Sutton, J.M.; Rahman, K.M. Effectiveness of Efflux Pump Inhibitors as Biofilm Disruptors and Resistance Breakers in Gram-Negative (ESKAPEE) Bacteria. Antibiotics 2019, 8, 229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Alav, I.; Sutton, J.M.; Rahman, K.M. Role of Bacterial Efflux Pumps in Biofilm Formation. J. Antimicrob. Chemother. 2018, 73, 2003–2020. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Dawan, J.; Li, Y.; Lu, F.; He, X.; Ahn, J. Role of Efflux Pump-Mediated Antibiotic Resistance in Quorum Sensing-Regulated Biofilm Formation by Salmonella Typhimurium. Pathogens 2022, 11, 147. [Google Scholar] [CrossRef] [PubMed]
  189. Kvist, M.; Hancock, V.; Klemm, P. Inactivation of Efflux Pumps Abolishes Bacterial Biofilm Formation. Appl. Environ. Microbiol. 2008, 74, 7376–7382. [Google Scholar] [CrossRef] [Green Version]
  190. Davies, D.G.; Parsek, M.R.; Pearson, J.P.; Iglewski, B.H.; Costerton, J.W.; Greenberg, E.P. The Involvement of Cell-to-Cell Signals in the Development of a Bacterial Biofilm. Science 1998, 280, 295–298. [Google Scholar] [CrossRef] [Green Version]
  191. Pearson, J.P.; Van Delden, C.; Iglewski, B.H. Active Efflux and Diffusion Are Involved in Transport of Pseudomonas Aeruginosa Cell-to-Cell Signals. J. Bacteriol. 1999, 181, 1203–1210. [Google Scholar] [CrossRef] [Green Version]
  192. Sánchez, P.; Linares, J.F.; Ruiz-Díez, B.; Campanario, E.; Navas, A.; Baquero, F.; Martínez, J.L. Fitness of in Vitro Selected Pseudomonas Aeruginosa NalB and NfxB Multidrug Resistant Mutants. J. Antimicrob. Chemother. 2002, 50, 657–664. [Google Scholar] [CrossRef] [Green Version]
  193. Lamarche, M.G.; Déziel, E. MexEF-OprN Efflux Pump Exports the Pseudomonas Quinolone Signal (PQS) Precursor HHQ (4-Hydroxy-2-Heptylquinoline). PLoS ONE 2011, 6, e24310. [Google Scholar] [CrossRef]
  194. De Kievit, T.R.; Parkins, M.D.; Gillis, R.J.; Srikumar, R.; Ceri, H.; Poole, K.; Iglewski, B.H.; Storey, D.G. Multidrug Efflux Pumps: Expression Patterns and Contribution to Antibiotic Resistance in Pseudomonas Aeruginosa Biofilms. Antimicrob. Agents Chemother. 2001, 45, 1761–1770. [Google Scholar] [CrossRef]
  195. Waite, R.D.; Papakonstantinopoulou, A.; Littler, E.; Curtis, M.A. Transcriptome Analysis of Pseudomonas Aeruginosa Growth: Comparison of Gene Expression in Planktonic Cultures and Developing and Mature Biofilms. J. Bacteriol. 2005, 187, 6571–6576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Harrington, N.E.; Littler, J.L.; Harrison, F. Transcriptome Analysis of Pseudomonas Aeruginosa Biofilm Infection in an Ex Vivo Pig Model of the Cystic Fibrosis Lung. Appl. Environ. Microbiol. 2022, 88, e0178921. [Google Scholar] [CrossRef] [PubMed]
  197. Kaatz, G.W. Inhibition of Bacterial Efflux Pumps: A New Strategy to Combat Increasing Antimicrobial Agent Resistance. Expert Opin. Emerg. Drugs 2002, 7, 223–233. [Google Scholar] [CrossRef] [PubMed]
  198. Aínsa, J.A.; Blokpoel, M.C.J.; Otal, I.; Young, D.B.; De Smet, K.A.L.; Martín, C. Molecular Cloning and Characterization of Tap, a Putative Multidrug Efflux Pump Present in Mycobacterium Fortuitum and Mycobacterium Tuberculosis. J. Bacteriol. 1998, 180, 5836–5843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Hulen, C.; Racine, P.-J.; Feuilloley, M.; Elomri, A.; Lomri, N.-E. Effects of Verapamil and Two Bisbenzylisoquinolines, Curine and Guattegaumerine Extracted from Isolona Hexaloba, on the Inhibition of ABC Transporters from Pseudomonas Aeruginosa. Antibiotics 2022, 11, 700. [Google Scholar] [CrossRef] [PubMed]
  200. Lomovskaya, O.; Warren, M.S.; Lee, A.; Galazzo, J.; Fronko, R.; Lee, M.; Blais, J.; Cho, D.; Chamberland, S.; Renau, T.; et al. Identification and Characterization of Inhibitors of Multidrug Resistance Efflux Pumps in Pseudomonas Aeruginosa: Novel Agents for Combination Therapy. Antimicrob. Agents Chemother. 2001, 45, 105–116. [Google Scholar] [CrossRef] [Green Version]
  201. Piddock, L.J.; Williams, K.J.; Ricci, V. Accumulation of Rifampicin by Mycobacterium Aurum, Mycobacterium Smegmatis and Mycobacterium Tuberculosis. J. Antimicrob. Chemother. 2000, 45, 159–165. [Google Scholar] [CrossRef] [Green Version]
  202. Silva, P.E.; Bigi, F.; Santangelo, M.P.; Romano, M.I.; Martín, C.; Cataldi, A.; Aínsa, J.A. Characterization of P55, a Multidrug Efflux Pump in Mycobacterium Bovis and Mycobacterium Tuberculosis. Antimicrob. Agents Chemother. 2001, 45, 800–804. [Google Scholar] [CrossRef] [Green Version]
  203. Coban, A.Y.; Ekinci, B.; Durupinar, B. A Multidrug Efflux Pump Inhibitor Reduces Fluoroquinolone Resistance in Pseudomonas Aeruginosa Isolates. Chemotherapy 2004, 50, 22–26. [Google Scholar] [CrossRef]
  204. Pagès, J.-M.; Amaral, L. Mechanisms of Drug Efflux and Strategies to Combat Them: Challenging the Efflux Pump of Gram-Negative Bacteria. Biochim. Biophys. Acta 2009, 1794, 826–833. [Google Scholar] [CrossRef]
  205. Hirakata, Y.; Kondo, A.; Hoshino, K.; Yano, H.; Arai, K.; Hirotani, A.; Kunishima, H.; Yamamoto, N.; Hatta, M.; Kitagawa, M.; et al. Efflux Pump Inhibitors Reduce the Invasiveness of Pseudomonas Aeruginosa. Int. J. Antimicrob. Agents 2009, 34, 343–346. [Google Scholar] [CrossRef] [PubMed]
  206. Kriengkauykiat, J.; Porter, E.; Lomovskaya, O.; Wong-Beringer, A. Use of an Efflux Pump Inhibitor to Determine the Prevalence of Efflux Pump-Mediated Fluoroquinolone Resistance and Multidrug Resistance in Pseudomonas Aeruginosa. Antimicrob. Agents Chemother. 2005, 49, 565–570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Mitchell, C.J.; Stone, T.A.; Deber, C.M. Peptide-Based Efflux Pump Inhibitors of the Small Multidrug Resistance Protein from Pseudomonas Aeruginosa. Antimicrob. Agents Chemother. 2019, 63, e00730-19. [Google Scholar] [CrossRef] [Green Version]
  208. Stavri, M.; Piddock, L.J.V.; Gibbons, S. Bacterial Efflux Pump Inhibitors from Natural Sources. J. Antimicrob. Chemother. 2007, 59, 1247–1260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Figueroa-Romero, A.; Pons-Duran, C.; Gonzalez, R. Drugs for Intermittent Preventive Treatment of Malaria in Pregnancy: Current Knowledge and Way Forward. Trop. Med. Infect. Dis. 2022, 7, 152. [Google Scholar] [CrossRef] [PubMed]
  210. Vidal-Aroca, F.; Meng, A.; Minz, T.; Page, M.G.P.; Dreier, J. Use of Resazurin to Detect Mefloquine as an Efflux-Pump Inhibitor in Pseudomonas Aeruginosa and Escherichia Coli. J. Microbiol. Methods 2009, 79, 232–237. [Google Scholar] [CrossRef] [PubMed]
  211. Piddock, L.J.V.; Garvey, M.I.; Rahman, M.M.; Gibbons, S. Natural and Synthetic Compounds Such as Trimethoprim Behave as Inhibitors of Efflux in Gram-Negative Bacteria. J. Antimicrob. Chemother. 2010, 65, 1215–1223. [Google Scholar] [CrossRef] [Green Version]
  212. Adamson, D.H.; Krikstopaityte, V.; Coote, P.J. Enhanced Efficacy of Putative Efflux Pump Inhibitor/Antibiotic Combination Treatments versus MDR Strains of Pseudomonas Aeruginosa in a Galleria Mellonella in Vivo Infection Model. J. Antimicrob. Chemother. 2015, 70, 2271–2278. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Laborda, P.; Alcalde-Rico, M.; Chini, A.; Martínez, J.L.; Hernando-Amado, S. Discovery of Inhibitors of Pseudomonas aeruginosa Virulence through the Search for Natural-like Compounds with a Dual Role as Inducers and Substrates of Efflux Pumps. Environ. Microbiol. 2021, 23, 7396–7411. [Google Scholar] [CrossRef]
  214. Tambat, R.; Mahey, N.; Chandal, N.; Verma, D.K.; Jangra, M.; Thakur, K.G.; Nandanwar, H. A Microbe-Derived Efflux Pump Inhibitor of the Resistance-Nodulation-Cell Division Protein Restores Antibiotic Susceptibility in Escherichia Coli and Pseudomonas Aeruginosa. ACS Infect. Dis. 2022, 8, 255–270. [Google Scholar] [CrossRef]
  215. Yamasaki, S.; Koga, N.; Zwama, M.; Sakurai, K.; Nakashima, R.; Yamaguchi, A.; Nishino, K. Spatial Characteristics of the Efflux Pump MexB Determine Inhibitor Binding. Antimicrob. Agents Chemother. 2022, 66, e0067222. [Google Scholar] [CrossRef] [PubMed]
  216. Liu, Y.; Yang, L.; Molin, S. Synergistic Activities of an Efflux Pump Inhibitor and Iron Chelators against Pseudomonas Aeruginosa Growth and Biofilm Formation. Antimicrob. Agents Chemother. 2010, 54, 3960–3963. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Rampioni, G.; Pillai, C.R.; Longo, F.; Bondì, R.; Baldelli, V.; Messina, M.; Imperi, F.; Visca, P.; Leoni, L. Effect of Efflux Pump Inhibition on Pseudomonas Aeruginosa Transcriptome and Virulence. Sci. Rep. 2017, 7, 11392. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Multidrug efflux pumps in P. aeruginosa. (a) Schematic structures of the four main efflux pumps involved in antibiotic resistance in P. aeruginosa, showing the resistance-nodulation-cell division transporters (MexB, MexY, MexD and MexF) on the inner membrane; the periplasmic membrane fusion proteins (MexA, MexX, MexC and MexE) on the periplasm; and the channel-forming outer membrane factors (OprM, OprJ and OprN) on the outer membrane. Protein representations based on the Protein Databank (PDB) structures: 2V50, 4DK1 and 3D5K. (b) Organization of the Mex operons and upstream regulatory genes in the P. aeruginosa genome. Regulatory genes are represented in white, and the Mex coding sequences follow the same color patterns of the protein structures shown in (a). Created with BioRender.com.
Figure 1. Multidrug efflux pumps in P. aeruginosa. (a) Schematic structures of the four main efflux pumps involved in antibiotic resistance in P. aeruginosa, showing the resistance-nodulation-cell division transporters (MexB, MexY, MexD and MexF) on the inner membrane; the periplasmic membrane fusion proteins (MexA, MexX, MexC and MexE) on the periplasm; and the channel-forming outer membrane factors (OprM, OprJ and OprN) on the outer membrane. Protein representations based on the Protein Databank (PDB) structures: 2V50, 4DK1 and 3D5K. (b) Organization of the Mex operons and upstream regulatory genes in the P. aeruginosa genome. Regulatory genes are represented in white, and the Mex coding sequences follow the same color patterns of the protein structures shown in (a). Created with BioRender.com.
Ijms 23 15779 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lorusso, A.B.; Carrara, J.A.; Barroso, C.D.N.; Tuon, F.F.; Faoro, H. Role of Efflux Pumps on Antimicrobial Resistance in Pseudomonas aeruginosa. Int. J. Mol. Sci. 2022, 23, 15779. https://doi.org/10.3390/ijms232415779

AMA Style

Lorusso AB, Carrara JA, Barroso CDN, Tuon FF, Faoro H. Role of Efflux Pumps on Antimicrobial Resistance in Pseudomonas aeruginosa. International Journal of Molecular Sciences. 2022; 23(24):15779. https://doi.org/10.3390/ijms232415779

Chicago/Turabian Style

Lorusso, Andre Bittencourt, João Antônio Carrara, Carolina Deuttner Neumann Barroso, Felipe Francisco Tuon, and Helisson Faoro. 2022. "Role of Efflux Pumps on Antimicrobial Resistance in Pseudomonas aeruginosa" International Journal of Molecular Sciences 23, no. 24: 15779. https://doi.org/10.3390/ijms232415779

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