Next Article in Journal
Effects of Curcumin and Tetracycline Gel on Experimental Induced Periodontitis as an Anti-Inflammatory, Osteogenesis Promoter and Enhanced Bone Density through Altered Iron Levels: Histopathological Study
Next Article in Special Issue
Halogenated Pyrrolopyrimidines with Low MIC on Staphylococcus aureus and Synergistic Effects with an Antimicrobial Peptide
Previous Article in Journal
Effect of β-Glucan Supplementation on Growth Performance and Intestinal Epithelium Functions in Weaned Pigs Challenged by Enterotoxigenic Escherichia coli
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Bacterial Multidrug Efflux Pumps at the Frontline of Antimicrobial Resistance: An Overview

Ministry of Agriculture Laboratory for Risk Assessment of Quality and Safety of Livestock and Poultry Products, Huazhong Agricultural University, Wuhan 430070, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Antibiotics 2022, 11(4), 520; https://doi.org/10.3390/antibiotics11040520
Submission received: 15 February 2022 / Revised: 1 April 2022 / Accepted: 8 April 2022 / Published: 13 April 2022

Abstract

:
Multidrug efflux pumps function at the frontline to protect bacteria against antimicrobials by decreasing the intracellular concentration of drugs. This protective barrier consists of a series of transporter proteins, which are located in the bacterial cell membrane and periplasm and remove diverse extraneous substrates, including antimicrobials, organic solvents, toxic heavy metals, etc., from bacterial cells. This review systematically and comprehensively summarizes the functions of multiple efflux pumps families and discusses their potential applications. The biological functions of efflux pumps including their promotion of multidrug resistance, biofilm formation, quorum sensing, and survival and pathogenicity of bacteria are elucidated. The potential applications of efflux pump-related genes/proteins for the detection of antibiotic residues and antimicrobial resistance are also analyzed. Last but not least, efflux pump inhibitors, especially those of plant origin, are discussed.

1. Introduction

Bacteria evolve mechanisms of defense against harmful external substances that threaten their survival. Transporters are located on the bacterial cell membrane and play important roles in decreasing the concentration of endogenous and extraneous substances and promoting the growth of bacteria. It should be noticed that some transporters are also involved in the biodegradation of environmental toxic compounds [1]. Therefore, bacterial efflux pumps not only work as a functional transporter but also defend from environmental stress to maintain bacteria survival. Although some efflux pumps have specific substrates, some transporters can pump out multiple different kinds of antimicrobials, inducing multidrug resistance [2]. Sometimes, the extrusion of organic solvents or other substrates leads to the overexpression of transporters, thus creating co-selection of antimicrobial resistance features [3]. The overexpression of efflux pumps would also impact bacterial pathogenicity, involving biofilms and quorum sensing (QS) [4,5]. The efflux pumps export not only antimicrobials but also virulence determinants, including adhesins, toxins, or other proteins that are important for colonization in host cells [6].
Many studies have described efflux pumps-mediated resistance mechanisms, and more and more new efflux transporters and related proteins have been discovered. In recent years, additional functions of transporters were investigated, which involved bacterial virulence and self-protection against environmental pollutants. Nevertheless, the exact mechanisms and active domains of efflux pump transporters have not been clearly elucidated. There are many influencing factors on the inner and outer bacterial membrane, stimulating pump activity and promoting structure alterations of the transporters in the fluid membrane environment.
This article briefly summarizes the functions of efflux pumps in common Gram-positive and Gram-negative bacteria. Applications of efflux pump-related regulator proteins for detecting antimicrobial resistance and antibiotic residues, as well as newly discovered plant-derived efflux pump inhibitors, are discussed.

2. Functions of Efflux Pumps

Efflux transporters are mainly grouped into the following superfamilies: ATP-Binding Cassette (ABC) superfamily, Multidrug and Toxic Compound Extrusion (MATE) superfamily, Major Facilitator Superfamily (MFS), Resistance Nodulation and Cell Division (RND) superfamily, and Small Multidrug Resistance (SMR) superfamily. In 2015, a novel transporter family was identified, known as the Proteobacterial Antimicrobial Compound Efflux (PACE) superfamily [7]. The discovery and identification of bacterial multidrug-resistant (MDR) efflux pumps is still ongoing [8]. One of the differences among efflux pumps is their source of energy. The ABC family members function as efflux pumps through the hydrolysis of ATP for energy supply [9], while other transporters including the MATE, MFS, RND, and SMR superfamilies utilize the proton-motive force provided by H+ or the electrochemical gradient of Na+ to supply energy and then extrude multiple compounds [10]. In addition, there are also differences in the composition of efflux transporters. The RND-type multidrug efflux proteins use a tripartite complex as the structural basis of the pump, including an outer-membrane canal protein (OMP), an inner-membrane transporter, and a membrane fusion protein (MFP) that connect the above two components to pump out drugs and other harmful compounds [11,12].

2.1. Efflux Pump-Mediated Antimicrobial Resistance

The resistance mechanisms to antimicrobials are based on changes in drug targets, structural modifications or degradation of drugs, decreased permeability of outer membrane proteins to prevent the drugs from entering the cells, and enhanced efflux transporters to reduce intracellular drug concentrations. It was generally believed that efflux pumps and outer membrane proteins did not have a synergistic effect on reducing intracellular drug concentrations. However, a recent study found an interaction between active efflux pumps and the permeability barrier of the outer membrane in Burkholderia thailandensis [13]. In fact, the overexpression of efflux pumps plays an essential role in the acquisition of antimicrobial resistance, even of multidrug resistance. Understanding the molecular structure of efflux pumps and their crucial drug binding sites is critical for the development of efflux pump inhibitors. The structures of different efflux pump families have been reviewed in detail and will be briefly discussed below [14,15]. In the following, the main focus is on the effects of efflux pumps on biofilm formation, bacterial virulence and invasion, and oxidative stress responses, in addition to the acquisition of antimicrobial resistance.

2.1.1. ABC Superfamily

In Gram-negative microorganisms, the most extensively studied ABC-type transporters is the tripartite complexes MacAB-TolC efflux pump, which actively extrudes substrates including macrolides and polypeptide virulence factors powered by the ATPase MacB and participates in the secretion of enterotoxin TⅡ in Escherichia coli [16,17,18]. Rough-core lipopolysaccharide (LPS) or similar glycolipids is also regarded as a physiological substrate of the MacAB-TolC pump [18].The inner membrane protein MacB works as a homodimer complex that contains an N-terminal nucleotide binding domain that binds ATP and a C-terminal cytoplasmic tail [18]. The membrane fusion protein MacA binds the lipopolysaccharide core specifically and is activated by ATPase [18]. The tripartite-complex structure composed of the outer membrane protein TolC, the inner membrane protein MacB, and the periplasmic protein MacA provides a structural site for substrate transport [17]. The latest research shows that the lack of the MacAB efflux pump in Serratia marcescens increases the susceptibility to aminoglycosides and polymyxins, decreases the swimming motility and the ability of biofilm formation, even causes the loss of the ability to cope with superoxide stress [19]. In addition, MacAB also confers resistance to several penicillin-type antibiotics and As(III) in Agrobacterium tumefaciens 5A [20].
In Gram-positive microorganisms, the ABC transporter consists of a single transmembrane protein, such as EfrAB in Enterococcus faecalis, LmrA in Lactococcus lactis [21], Msr in Streptococcus, and PatA/B in Streptococcus pneumoniae [22,23]. The MDR pump EfrAB is a heterodimeric transporter that extrudes gentamicin, streptomycin, and chloramphenicol [24]. The expression of EfrAB is highly induced by sub-inhibitory concentration(sub-MIC) of these antibiotics and down-regulated by sub-MIC of EDTA (3 mM) [24]. The LmrA protein functions as a homodimer, comprising a nucleotide-binding domain and six alpha-helix transmembrane domains, which primarily recognize and transport macrolides and lincosamides [25]. The Msr protein harbors two nucleotide-binding domains with no transmembrane domain, conferring resistance to macrolides in Streptococcus and cooperating with the Mef transport family [26,27]. Nearly the whole family of hydrophilic fluoroquinolones, including ciprofloxacin and norfloxacin, is the substrate of PatA/B efflux pumps [28].

2.1.2. MATE Superfamily

All members of the MATE family present 12 alpha-helical transmembrane regions powered by electrochemical ion gradients [29]. MATE transporters are classified into NorM, DinF (DNA damage-inducible protein F), and eukaryotic subfamilies according to amino-acid sequence similarity [30]. The substrates of MATE family transporters are diverse and have different chemical structures, but fluoroquinolones are the substrates recognized by almost all transporters [31].
Among the Gram-negative bacteria, the most studied pump is the NorM efflux pump in Neisseria gonorrhoeae and Vibrio cholera [32]. The NorM efflux pump exports substrates including antimicrobial cationic compounds (quaternary ammonium compounds) and antimicrobials such as ciprofloxacin and solithromycin in N. gonorrhoeae [33]. NorM has the ability to extrude intracellular reactive oxygen species, therefore reducing the damage of oxidative stress [34]. Mutations of the DinF transport system confer increased susceptibility to moxifloxacin, ciprofloxacin, and levofloxacin in pneumococci [35]. The competence-induced protein A is encoded by cinA, recA, and dinF, which form an operon that is induced by quinolones through the SOS response [35].

2.1.3. MFS Superfamily

The MFS family is the largest characterized family of transporters, especially in Gram-positive microorganisms. Most members of this superfamily are MDR transporters and are constituted by 12 or 14 transmembrane segments [29]. The pumps Lde in Listeria monocytogenes [36] and NorA in Staphylococcus aureus [37] specifically extrude hydrophilic fluoroquinolones (e.g., ciprofloxacin and norfloxacin), while Mef in S. pneumoniae pumps out macrolides [38]. It has been found that the transcriptional start of mef(E)/mel genes locates in the mef(E)/mel promoter, and the attenuation of transcription will regulate mef-mediated macrolide resistance [39]. In addition, Msr of ABC family and Mef of MFS family enhanced macrolides extrusion synergistically, which increased resistance to 14-, 15-member ring macrolides [40].
In addition to extruding drugs, MFS transporter proteins also play important roles in other biological pathways. For example, MdrT and MdrM contribute to promoting the host immune response by activating IFN-β production of the type I interferon response and to maintaining cell wall stability [41]. Tet38, a chromosomally encoded MFS efflux pump, has an impact on multiple steps of the host cell invasion process of S. aureus, including adhesion, internalization, and trafficking in epithelial cells, and the subsequent step of the S. aureus infection in epithelial cells including bacterial viability and trafficking in phagolysosomes [41]. The AbaQ MFS transporter is also involved in Acinetobacter baumannii motility and virulence, as the loss of the abaQ gene decreases bacterial motility and virulence [41]. Additionally, the inactivation of genes encoding the RND, MATE, SMR, and ABC efflux pumps also reduces bacterial motility and virulence in comparison with the parental strain [42].

2.1.4. RND Superfamily

RND efflux pumps are generally present in Gram-negative bacteria, which expel an extensive spectrum of antibiotics and toxic compounds They include AcrAB-TolC in E. coli, AdeABC in A. baumannii [43], CmeABC in Campylobacter jejuni [44], MexAB-OprM in Pseudomonas aeruginosa [45], MtrCDE in N. gonorrhea [46], OqxAB in Klebsiella pneumoniae and Salmonella enterica [47,48], SmeABC in Stenotrophomonas maltophilia, and TtgABC in Pseudomonas putida [49,50]. The Inner transporters such as AcrB, AdeB, CmeB, MexB, TtgB, SmeB, and MtrD are responsible for specific substrate binding and the transportation of different classes of drugs, playing vital roles in clinically relevant resistance [51]. For instance, mutations of the acrB gene would cause the failure of ciprofloxacin therapy [52]. The expression of the above-mentioned efflux pumps is also regulated by transcriptional regulatory proteins belonging to the TetR family, including AcrR [53], CmeR [54], NalC/NalD [55,56], TtgR [57], SmeT [58], and MtrR [59], as well as MexR belonging to MarR family [60]. Table 1 shows the efflux pumps regulated by the TetR family and summarizes the residues that have been proved to be important for the binding of the activating molecule by mutational analysis. Amino acid residues in efflux pumps may be important sites for substrates binding, and the substitution of amino acid residues may affect the affinity of the substrates. For example, Val 141, Phe 626, Tyr 330, and Phe 180 in the substrate-binding domain of the OqxB transporter protein are crucial for binding and transporting different substrates [61]. Additionally, mutations of important amino acid residues are associated with efflux pump-mediated multidrug resistance. The latest research shows that changing the glycine residue 311 to the acidic amino acid glutamic acid (G311E) in the MATE family protein DTX6 markedly increases Arabidopsis plants’ resistance to paraquat and diquat [62].
The multidrug resistance efflux pumps AcrAB-TolC, MexAB-OprM, CmeABC, MtrCDE of the RND family are involved in the survival and pathogenicity of bacteria [6]. Several studies have found that AcrAB-tolC efflux pump in S. Typhimurium and E. coli affect bacterial adhesion and invasion in host cells and colonization and persistence in animals [6]. Bacteria produce pore-forming toxins, degrading enzymes, and virulence factors of the secretory system to exert their pathogenicity and evade the attack from the host immune system. The outer membrane protein TolC can transport not only drugs but also hemolysin and colicin V [63,64].
Heavy metal ions are poisonous or inhibit the growth and survival of bacteria at specific concentrations. Due to the widespread application of heavy metal ions in antimicrobials such as antiseptic disinfectants, bacteria have to take measures to deal with the stress of heavy metals. In E. coli, the RND superfamily plays an important role in the resistance to antibiotics and heavy metals. There are seven known RND proteins; AcrB, AcrD, AcrF, MdtB, MdtC, and YhiV are multidrug efflux pumps of the hydrophobic and amphiphilic efflux RND (HAE-RND) protein family, while CusA is a member of the heavy-metal efflux RND (HME-RND) family that exports Cu(I) and Ag(I) [69]. CusA together with CusB membrane fusion protein and CusC channel protein constitute the complete tripartite CusCBA efflux complexes [70]. Long et al. predicted the mechanisms of metal ions export [70]. Firstly, Cu(I) is combined with the chaperone CusF and then is delivered to three-methionine metal-binding sites (M49, M64, and M66) at the long N-terminal tail of CusB; secondly, the metal ion is transferred to the three-methionine cluster (M573, M623, and M672) inside the periplasmic cleft of CusA and then is released into the central funnel of CusA; finally, the metal ion enters the CusC channel for its final extrusion [70]. Bacteria protect themselves from toxic components of organic pollutants by efflux pump biodegradation. TtgABC, a tripartite RND efflux pump, confers toluene tolerance in P. putida [71].
The stress caused by the abuse of antimicrobials increases the occurrence of functional mutations in the RND family, possibly enhancing its efflux function. The substrate specificity of efflux pumps correlates with differences of amino acid residues in the drug-binding pocket, and bacteria become less sensitive to antimicrobials after the substitution of these amino acid residues. Numerous examples of mutations in the RND efflux pump which have been identified from clinical and environmental isolates and laboratory-evolved strains affect antimicrobial resistance and are fully described in this article [72]. The acquisition of a functional mutation in an efflux pump may be an adaptive response of bacteria to antimicrobials and undoubtedly increases difficulties in the clinical treatment of bacterial infections.

2.1.5. SMR Superfamily

The SMR family members consist of short polypeptide (100–150 amino acids) and span the cytoplasmic membrane as four transmembrane α-helices [73]. Those proteins with short hydrophilic loops permit to solubilization a broad spectrum of drugs, such as disinfecting agents including quaternary ammonium compounds, toxic lipophilic compounds including DNA intercalating dyes, toxic metabolites including nicotine intermediates, and polyamine compounds like spermidine [73,74]. In Gram-negative bacteria, the AbeS in A. baumannii transports sterilant ethidium, acriflavine, and benzalkonium, which also significantly increase resistance to amikacin [75,76]. The clinical strains of K. pneumoniae possesses the activated pump KpnEF and resists to benzalkonium chloride, chlorhexidine, and some other antiseptics [77]. The EmrE protein, present in both E. coli and P. aeruginosa, recognizes and mediates the extrusion of toxic polyaromatic compounds [78,79]. Specifically, the Qac protein is associated with resistance to some antiseptics and antibiotics [80]. The qacA/B genes is more frequently observed in S. aureus and E. faecalis, while the qacE gene is widely spread in Enterobacteriaceae and Pseudomonas spp. [80].
Genes encoding SMR proteins often occur on multidrug resistance plasmids and mobile genetic elements called integrons, which increase the risk of horizontal transmission of resistance [74]. Overexpression of efflux pumps induced by the exposure to QAC facilitates the horizontal transfer of mobile genetic elements carrying FQ resistance determinants (qnr, aac(60)-Ib-cr, oqxAB, qepAB) in Class 1 integrons (QacED1) [81]. Co-transfer of disinfectant-resistant genes and antibiotics-resistant genes among different species seriously influences the bactericidal effects of disinfectants and antibiotics.

2.1.6. PACE Superfamily

The PACE transporter family has been identified in the last 5 years, and the Acel protein in A. baumannii is the first transporter protein found in the PACE family [82]. This transporter contributes to extruding biosynthetic biocides (e.g., chlorhexidine, acriflavine, proflavine, dequalinium, and benzalkonium) [7]. It has been found that genomes encoding the PACE transporter family are highly conserved among bacterial species, which suggests that PACE transporter-related genes are acquired perpendicularly and maintain their characteristic in host species [83].
AceI (Acinetobacter chlorhexidine efflux protein I) is similar to members of the SMR family in size and in the predicted secondary structure. AceI contains two tandem bacterial transmembrane pairs. BTP-domain proteins homologous to AceI have also been discovered in many pathogens such as Burkholderia, Enterobacter, Klebsiella, Pseudomonas, and Salmonella species [83,84,85]. The structure of the AceI protein is in equilibrium between a monomer and a dimer. Increases in chlorhexidine concentration and pH promote the formation of an acetylated dimer, and the binding of chlorhexidine to the transcriptional protein AceR increases aceI transcription, thus extruding chlorhexidine [82].
Table 2 and Table 3 summarize the major efflux pumps and their specific substrates in Gram-positive and Gram-negative bacteria, respectively.

2.2. Efflux Pumps Affect Biofilm Formation and Quorum Sensing (QS)

Biofilms, a microbial community attached to a surface, contribute to bacterial resistance and tolerance. Interestingly, the function of biofilms is directly or indirectly influenced by efflux pumps [94,95]. For instance, sub-inhibitory concentration of tigecycline directly restrain biofilm formation by downregulating the adeG efflux gene in A. baumannii [96]. QS is formed among cells, promotes the mutual communication of cells, and participates in bacterial activities through signal transduction. QS cooperation with biofilms improves the viability of bacteria by sensing changes in environmental signals. For example, the MexAB-OprM efflux pump extrudes acylated homoserine lactones with the contribution of QS, and overexpression of the Mex pump in P. aeruginosa results in the release of QS signals [6]. It has been found that QS among bacteria will be impeded if efflux pumps activity is hindered by inhibitors [97]. Similarly, the proper concentration of efflux pump inhibitors prevents biofilm formation, as observed for inhibitors of the NorA efflux pump in S. aureus [98].
The proteomic profiles of two A. baumannii strains grown in planktonic stationary phase or in mature solid–liquid biofilm were compared using proteomics, and the MacAB-TolC efflux pump was found to play an essential role in biofilm formation [99]. The efflux pump, which helps pathogen adapt to bad conditions occurring in mature biofilms, is involved in envelope stress responses that maintain membrane rigidity and confer resistance to high osmotic stress [99]. The MFS transporter proteins Pmt and AbaF are involved in A. baumannii biofilm formation. Pmt extrudes extracellular DNA and plays an essential role in forming the biofilm structure, while AbaF releases biofilm materials [41].
The RND-type MDR efflux system is well studied. It has been found that a number of RND transporters contribute to alterations of the membrane involved in bacterial functions, especially biofilm formation [100]. This was shown for A. baumannii efflux pump AdeB and its regulatory protein AdeRS. Deletion of adeAB genes or inhibition of the expression of these genes leads to the reduction or prevention of biofilm formation and of the QS system [101,102]. There is a positive correlation between biofilm formation and the mRNA levels of the efflux pump genes adeB, adeG, and adeI, which is altered by sub-MICs of colistin or polymyxin B [103]. The pump MexGHI of P. aeruginosa transport phenazine, which is required for biofilm morphogenesis [104]. The QS system has been shown to regulate the expression of RND efflux pumps. Conversely, the RND transporters can also influence QS via translocation of quorum signals [105]. It was shown that virulence and quorum sensing molecules of P. aeruginosa would be lost and reduced when the efflux gene mexI is mutated [106]. The ABC-type efflux system also plays a role in resistance to antifungal agents in fungi, particularly in Candida species. Similar to the RND pumps, it is involved in the secretion of quorum-sensing molecules and affects biofilms’ behavior [107].

3. Regulation of Efflux Pumps and Application of Efflux Pump-Related Genes/Proteins

3.1. Regulation Mechanisms of Efflux Pumps

In addition to exposure to antimicrobials or disinfectants, efflux pump expression is also regulated by a variety of regulatory systems and proteins. Currently, single regulatory proteins for the MDR efflux pump are mainly classified into four categories, i.e., AraC, MarR, MerR, and TetR [108,109,110]. These regulatory proteins contain DNA-binding domains and ligand-binding domains. For instance, the TetR family is a substrate-dependent transcriptional repressor and controls the expression of the efflux-related tet genes, which confer resistance to tetracyclines [111]. FepA is a novel type of MATE efflux pump, whose overexpression results in resistance to fluoroquinolones. It is regulated by FepR, a TetR-type repressor, which increases the MICs of norfloxacin and ciprofloxacin in L. monocytogenes [67]. The expression of the AceI protein, a transporter protein of the PACE family, is regulated by the LysR-type transcriptional regulator AceR [82]. There are extensively studies reporting that the transcriptional regulatory protein MepR, a MarR-type repressor, inactivates the MepA protein, whose substrates include biocides, fluoroquinolones, and tigecycline [92,93]. These regulators primarily respond to several types of signals including superoxide and peroxide (e.g., H2O2) [112], antimicrobials (e.g., antibiotics), toxic reagents (e.g., methylglyoxal), and biocides (e.g., triclosan) [113,114].
Another type of efflux pumps regulation is based on the two-component regulatory system, consisting of a cognate response regulator and a sensor histidine protein kinase, such as SoxRS, AdeRS, BaeRS, SmeRS, MacRS, ArlRS, EvgSA, BaeSR, CpxAR [115,116,117,118,119,120,121]. Each sensor detects a specific intracellular signal (e.g., antibiotics) and then induces histidine auto-phosphorylation, thus transferring the phosphate group to the proper aspartic acid and activating the response regulator [122]. For instance, it is known that AcrAB in E. coli is controlled by a series of different regulatory systems, including MarA, Rob, SoxS, AcrR, EnvR, MprA, PhoP, RpoE [123]. These regulatory systems are activated by external environmental signals, such as pH, the concentrations of antimicrobials, divalent metal ions, organic solvents, the growth phase, and oxidative stress [123].

3.2. Determination of Antimicrobial Resistance Based on Efflux Pump Gene Expression

Microorganisms harbor MDR efflux pumps resulting in inherent or acquired resistance to antimicrobial agents. The resistance-related genes encode the constituent proteins or regulatory proteins of efflux pumps, which function as the first line of defense against drugs, maintaining the survival of the bacteria [124]. Many efflux pump genes can be used for the rapid detection of antimicrobial resistance, which can be efficiently verified by PCR identification and MIC determination. The RND family of efflux pumps recognizes a large number of substrates, extruding the majority of drugs and increasing antimicrobial resistance. Examples are the Acr pump in E. coli, the Ade pump in A. baumannii, and the Mex pump in P. aeruginosa [76,125]. According to whether the efflux gene is expressed or not, we can determine antibiotic resistance and guide clinical therapy [126]. For instance, in the carbapenem-resistant A. baumannii isolate, the expression of the efflux genes adeB, adeG, and adeJ was increased by different folds [127]. Similarly, a stimulated expression of adeB and adeJ was also found in bacteria resistant to tetracyclines [128]. However, there is a slight relationship between substrates concentration and velocity of maximal transportation. For example, although cefaloridine can be strongly excreted by AcrB, it still possesses antibacterial activity in the presence of AcrB, which explains that the effective antibacterial dose of cefaloridine is much lower than the concentration required for the efflux [129].
In addition, some single-substrate efflux pumps are associated with high level of resistance and MICs. These single-substrate efflux proteins include the macrolide-specific efflux pumps MacAB in E. coli and Mef in S. pneumoniae [38], the hydrophilic fluoroquinolones efflux pump OqxAB in E. coli and S. enterica [130], the tetracycline-mediated efflux pump TetA/TetO in E. coli [131], etc.

3.3. Detection of Antibiotic Residues Based on Efflux Pump Proteins

Presently, efflux pump-related proteins are still under research for antibiotic residue detection, and only two regulatory proteins are used for the screening of antibiotics residues, including TetR and TtgR (belongs to TetR family). Receptor proteins involved in gene transcriptional regulation have been proven to be a powerful tool for detecting low antibiotic concentrations [132]. Hyerim et al. established a bioreporter method based on TetR and tetR promoter to detect doxycycline, using a green fluorescence protein gene as reporter gene, and 5nM doxycycline would induce high expression levels of green fluorescence protein [133]. Weber et al. developed an in vitro indirect enzyme-linked immunosorbent assay using TetR-tetO to accurately and rapidly detect tetracycline antibiotics, with the detection limits of doxycycline and tetracycline of 0.1 and 1.9 ng/mL, respectively [134]. Moreover, Espinosa-Urgel et al. established a novel microbial biosensor based on TtgR to detect drug concentrations around 22 µM with high fluorescence intensity [135]. It was demonstrated that green fluorescence protein-fused TtgR, a TtgABC efflux pump transcription regulator in P. putida, is most sensitive to ceftazidime, ciprofloxacin, and tetracyclines [135].

4. Efflux Pump Inhibitors (EPIs) in Antimicrobial Therapy

Efflux pumps act as a new target for antimicrobial combination therapy, enabling synthetic or plant-derived EPIs to assist antibiotics therapy against bacterial infections [136,137]. Multi-drug resistance is largely mediated by efflux pumps; hence, the development of efflux pump inhibitors is necessary to curb antimicrobial resistance. There are various ways to block or bypass the action of the efflux pumps, including decreasing the binding affinity of antibiotics to the transporter by modifying the drug’s chemical structure, increasing the permeability of the outer membrane to increase the intracellular drug concentrations, inhibiting or knocking out efflux pump-related genes, impairing the ATP energy supply, or designing substances that can compete with antimicrobials for the action site of efflux pumps to competitively inhibit the efflux activity [138].
A number of inhibitors have been discovered by computational analysis or artificial extraction from plants. Through high-throughput virtual screening of natural compound collections against NorM—a MATE transporter from N. gonorrhea—authors found that Terminalia chebula, a compound from an Indian medicinal plant, destroyed the binding of Na+ and ligands and turned the NorM transporter into a closed state [139]. Phenylalanyl–arginine β-naphthylamide, an inhibitor of AcrB in E. coli, has been shown to inhibit the extrusion of drugs via binding to the hydrophobic pocket of AcrB [140]. It can be combined with carolacton, a secondary metabolite in myxobacteria, for potential use in antimicrobial chemotherapy against AcrAB-TolC [141]. Tannic acid also acts as an inhibitor of S. aureus multidrug efflux pump, Tet, Msr, and others [142]. It has been found to significantly reduce the MIC of drugs such as tetracycline and erythromycin [142]. However, EPIs combined with antibiotics to assist MDR therapy in the clinic is an obviously potential challenge, depending on the intrinsic permeability properties of the bacterial outer membranes. Yang et al. demonstrated that the coupled use of tobramycin and EPI promoted the binding of EPI with tetracycline and acted on MDR P. aeruginosa [143]. They also explored the effect of tobramycin–EPI conjugates in relation to with fluroquinolone, rifampicin, and fosfomycin, showing that they effectively reduced the MIC80 of these drugs and exerted a strong joint effect [144]. Adamson et al. showed that the combination of EPI (trimethoprim and sertraline) and levofloxacin against the overexpression of MexAB-OprM pumps in P. aeruginosa produced advanced benefit compared with the monotherapy with levofloxacin [145]. Prasch et al. conducted similar research and demonstrated that when the above efflux pump inhibitors are co-administered with antibiotics, the antibiotic therapeutic dose can be reduced [146].
In addition to the EPIs mentioned above, plant-derived EPIs have also been researched (Table 4). Based on various extraction mechanisms, more than 20 different potential plant-derived EPIs have been reported [147,148]. In addition to good anti-inflammatory and antibacterial effects, compounds extracted from vegetal including berberine, Artesunate, and Curcumin inhibit the efflux pump activity of the Gram-negative bacilli E. coli and P. aeruginosa [149,150,151]. Vegetables (e.g., Momordica balsamina), seeds (e.g., milk thistle seeds), spices (e.g., pepper and cumin), essential oils derived from aromatic plants (e.g., trans-cinnamaldehyde and eugenol), etc., are excellent sources of EPIs [147,152,153,154,155,156]. Flavonoids, particularly flavonolignans, were found to have great application prospects in the fight against multidrug resistance by inhibiting bacterial ABC transporters and other efflux pumps [157].

5. Conclusions and Perspectives

This work reviewed functional studies of various efflux pump families whose substrates include heavy metals, disinfectants, preservatives, toxins, and virulence factors, in addition to conventional antibiotics. Apart from mediating multidrug resistance, efflux pumps also confer resistance to heavy metals and disinfectants, and even cross-resistance to them. This suggests that efflux pumps have much more complex actions than we initially thought. Efflux pumps play important roles in biological processes such as biofilm formation, quorum sensing, bacterial adhesion to host cells, and invasiveness. This further highlights the importance of developing efflux pump inhibitors. Many discovered efflux pump inhibitors are in clinical trials, though some have even been excluded due to their excessive toxicity. Therefore, there is an urgent need to find safe, green, and harmless efflux pump inhibitors. Plant-derived extracts are interesting candidates. Overall, the functional diversity of efflux pumps remains to be discovered, and the development and utilization of efflux pump inhibitors still require further exploration.
According to preceding research, proteins of the TetR family can be used for the detection of antibiotic residues involved in drug resistance. In addition to high-affinity transcriptional regulatory proteins reported above, there are other potential proteins that recognize a single substrate and have a specific drug binding domain. As shown in Table 3, regulatory and constitutive proteins have specific drug binding sites that can be used for the detection of antibiotics residues involved in the binding. Therefore, efflux pump-related proteins can also be used to detect antibiotic residues, especially transcriptional regulators, though only TetR family proteins are currently applied in practice. It is imperatively demanded that researchers discover more appropriate proteins with high binding specificity for antibiotics. Moreover, this article mentions the connections between the formation of biofilms and the expression of efflux pumps genes, which offers new perspectives to widen fundamental research. This article illustrated the functions of efflux genes and relevant transporters, a topic that requires constant exploration to provide guidance for clinical applications.

Author Contributions

Writing—original draft preparation, L.H. (Lulu Huang), H.G., C.W. and C.X.; writing—review and editing, L.H. (Lulu Huang), H.G., C.W.; supervision, M.D., L.H. (Lingli Huang), X.W., H.H. and G.C.; funding acquisition, G.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research and the APC was funded by National Natural Science Foundation of China [No. 32072921].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Abbreviations

ABCATP-binding cassette;
ATPadenosine triphosphate;
EPIefflux pump inhibitor;
HTHhelix–turn–helix;
IMFinner-membrane fusion protein;
MATEmultidrug and toxic microbial extrusion;
MDRmultidrug resistant;
MFPmembrane fusion protein;
MFSmajor facilitator super family;
Mgmagnesium;
MICminimum inhibitory concentration;
OMFouter membrane protein;
PACEProteobacterial Antimicrobial Compound Efflux;
QSQuorum sensing;
RNDresistance nodulation and cell division;
SMRsmall multidrug resistance;
SNPsingle-nucleotide polymorphism.

References

  1. Ganas, P.; Mihasan, M.; Igloi, G.L.; Brandsch, R. A two-component small multidrug resistance pump functions as a metabolic valve during nicotine catabolism by Arthrobacter nicotinovorans. Microbiology 2007, 153, 1546–1555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Hernando-Amado, S.; Blanco, P.; Alcalde-Rico, M.; Corona, F.; Reales-Calderon, J.A.; Sanchez, M.B.; Martinez, J.L. Multidrug efflux pumps as main players in intrinsic and acquired resistance to antimicrobials. Drug Resist. Updat. 2016, 28, 13–27. [Google Scholar] [CrossRef] [PubMed]
  3. Blair, J.M.; Piddock, L.J. How to Measure Export via Bacterial Multidrug Resistance Efflux Pumps. mBio 2016, 7, e00840-16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Kong, E.F.; Tsui, C.; Kucharikova, S.; Van Dijck, P.; Jabra-Rizk, M.A. Modulation of Staphylococcus aureus Response to Antimicrobials by the Candida albicans Quorum Sensing Molecule Farnesol. Antimicrob. Agents Chemother. 2017, 61, e01573-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Alcalde-Rico, M.; Hernando-Amado, S.; Blanco, P.; Martinez, J.L. Multidrug Efflux Pumps at the Crossroad between Antibiotic Resistance and Bacterial Virulence. Front. Microbiol. 2016, 7, 1483. [Google Scholar] [CrossRef] [Green Version]
  6. Piddock, L.J. Multidrug-resistance efflux pumps—Not just for resistance. Nat. Rev. Microbiol. 2006, 4, 629–636. [Google Scholar] [CrossRef]
  7. 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, 69, 450–454. [Google Scholar] [CrossRef]
  8. Li, L.; Tetu, S.G.; Paulsen, I.T.; Hassan, K.A. A Transcriptomic Approach to Identify Novel Drug Efflux Pumps in Bacteria. Methods Mol. Biol. 2018, 1700, 221–235. [Google Scholar] [CrossRef]
  9. Verchere, A.; Broutin, I.; Picard, M. Photo-induced proton gradients for the in vitro investigation of bacterial efflux pumps. Sci. Rep. 2012, 2, 306. [Google Scholar] [CrossRef] [Green Version]
  10. Kim, Y.C.; Hummer, G. Proton-pumping mechanism of cytochrome c oxidase: A kinetic master-equation approach. Biochim. Biophys. Acta 2012, 1817, 526–536. [Google Scholar] [CrossRef] [Green Version]
  11. Daury, L.; Orange, F.; Taveau, J.C.; Verchere, A.; Monlezun, L.; Gounou, C.; Marreddy, R.K.; Picard, M.; Broutin, I.; Pos, K.M.; et al. Tripartite assembly of RND multidrug efflux pumps. Nat. Commun. 2016, 7, 10731. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Neuberger, A.; Du, D.; Luisi, B.F. Structure and mechanism of bacterial tripartite efflux pumps. Res. Microbiol. 2018, 169, 401–413. [Google Scholar] [CrossRef] [PubMed]
  13. Krishnamoorthy, G.; Weeks, J.W.; Zhang, Z.; Chandler, C.E.; Xue, H.; Schweizer, H.P.; Ernst, R.K.; Zgurskaya, H.I. Efflux Pumps of Burkholderia thailandensis Control the Permeability Barrier of the Outer Membrane. Antimicrob. Agents Chemother. 2019, 63, e00956-19. [Google Scholar] [CrossRef] [PubMed]
  14. Kumar, S.; Lekshmi, M.; Parvathi, A.; Ojha, M.; Wenzel, N.; Varela, M.F. Functional and Structural Roles of the Major Facilitator Superfamily Bacterial Multidrug Efflux Pumps. Microorganisms 2020, 8, 266. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. 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]
  16. Jo, I.; Hong, S.; Lee, M.; Song, S.; Kim, J.S.; Mitra, A.K.; Hyun, J.; Lee, K.; Ha, N.C. Stoichiometry and mechanistic implications of the MacAB-TolC tripartite efflux pump. Biochem. Biophys. Res. Commun. 2017, 494, 668–673. [Google Scholar] [CrossRef]
  17. Fitzpatrick, A.W.P.; Llabrés, S.; Neuberger, A.; Blaza, J.N.; Bai, X.C.; Okada, U.; Murakami, S.; van Veen, H.W.; Zachariae, U.; Scheres, S.H.W.; et al. Structure of the MacAB-TolC ABC-type tripartite multidrug efflux pump. Nat. Microbiol. 2017, 2, 17070. [Google Scholar] [CrossRef] [Green Version]
  18. Lu, S.; Zgurskaya, H.I. MacA, a periplasmic membrane fusion protein of the macrolide transporter MacAB-TolC, binds lipopolysaccharide core specifically and with high affinity. J. Bacteriol. 2013, 195, 4865–4872. [Google Scholar] [CrossRef] [Green Version]
  19. Shirshikova, T.V.; Sierra-Bakhshi, C.G.; Kamaletdinova, L.K.; Matrosova, L.E.; Khabipova, N.N.; Evtugyn, V.G.; Khilyas, I.V.; Danilova, I.V.; Mardanova, A.M.; Sharipova, M.R.; et al. The ABC-Type Efflux Pump MacAB Is Involved in Protection of Serratia marcescens against Aminoglycoside Antibiotics, Polymyxins, and Oxidative Stress. mSphere 2021, 6, e00033-21. [Google Scholar] [CrossRef]
  20. Shi, K.; Cao, M.; Li, C.; Huang, J.; Zheng, S.; Wang, G. Efflux proteins MacAB confer resistance to arsenite and penicillin/macrolide-type antibiotics in Agrobacterium tumefaciens 5A. World J. Microbiol. Biotechnol. 2019, 35, 115. [Google Scholar] [CrossRef]
  21. Hellmich, U.A.; Monkemeyer, L.; Velamakanni, S.; van Veen, H.W.; Glaubitz, C. Effects of nucleotide binding to LmrA: A combined MAS-NMR and solution NMR study. Biochim. Biophys. Acta 2015, 1848, 3158–3165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Iannelli, F.; Santagati, M.; Santoro, F.; Oggioni, M.R.; Stefani, S.; Pozzi, G. Nucleotide sequence of conjugative prophage Phi1207.3 (formerly Tn1207.3) carrying the mef(A)/msr(D) genes for efflux resistance to macrolides in Streptococcus pyogenes. Front. Microbiol. 2014, 5, 687. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Garvey, M.I.; Baylay, A.J.; Wong, R.L.; Piddock, L.J. Overexpression of patA and patB, which encode ABC transporters, is associated with fluoroquinolone resistance in clinical isolates of Streptococcus pneumoniae. Antimicrob. Agents Chemother. 2011, 55, 190–196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Lavilla Lerma, L.; Benomar, N.; Valenzuela, A.S.; Casado Munoz Mdel, C.; Galvez, A.; Abriouel, H. Role of EfrAB efflux pump in biocide tolerance and antibiotic resistance of Enterococcus faecalis and Enterococcus faecium isolated from traditional fermented foods and the effect of EDTA as EfrAB inhibitor. Food Microbiol. 2014, 44, 249–257. [Google Scholar] [CrossRef] [PubMed]
  25. Hellmich, U.A.; Lyubenova, S.; Kaltenborn, E.; Doshi, R.; van Veen, H.W.; Prisner, T.F.; Glaubitz, C. Probing the ATP hydrolysis cycle of the ABC multidrug transporter LmrA by pulsed EPR spectroscopy. J. Am. Chem. Soc. 2012, 134, 5857–5862. [Google Scholar] [CrossRef] [PubMed]
  26. Zhang, Y.; Tatsuno, I.; Okada, R.; Hata, N.; Matsumoto, M.; Isaka, M.; Isobe, K.; Hasegawa, T. Predominant role of msr(D) over mef(A) in macrolide resistance in Streptococcus pyogenes. Microbiology 2016, 162, 46–52. [Google Scholar] [CrossRef] [PubMed]
  27. Tatsuno, I.; Isaka, M.; Masuno, K.; Hata, N.; Matsumoto, M.; Hasegawa, T. Functional Predominance of msr(D), Which Is More Effective as mef(A)-Associated Than mef(E)-Associated, Over mef(A)/mef(E) in Macrolide Resistance in Streptococcus pyogenes. Microb. Drug Resist. 2018, 24, 1089–1097. [Google Scholar] [CrossRef]
  28. Baylay, A.J.; Piddock, L.J. Clinically relevant fluoroquinolone resistance due to constitutive overexpression of the PatAB ABC transporter in Streptococcus pneumoniae is conferred by disruption of a transcriptional attenuator. J. Antimicrob. Chemother. 2015, 70, 670–679. [Google Scholar] [CrossRef] [Green Version]
  29. Du, D.; van Veen, H.W.; Murakami, S.; Pos, K.M.; Luisi, B.F. Structure, mechanism and cooperation of bacterial multidrug transporters. Curr. Opin. Struct. Biol. 2015, 33, 76–91. [Google Scholar] [CrossRef] [Green Version]
  30. Lu, M. Structures of multidrug and toxic compound extrusion transporters and their mechanistic implications. Channels 2016, 10, 88–100. [Google Scholar] [CrossRef] [Green Version]
  31. Kuroda, T.; Tsuchiya, T. Multidrug efflux transporters in the MATE family. Biochim. Biophys. Acta 2009, 1794, 763–768. [Google Scholar] [CrossRef] [PubMed]
  32. Kusakizako, T.; Miyauchi, H.; Ishitani, R.; Nureki, O. Structural biology of the multidrug and toxic compound extrusion superfamily transporters. Biochim. Biophys. Acta Biomembr. 2020, 1862, 183154. [Google Scholar] [CrossRef]
  33. Rouquette-Loughlin, C.E.; Dhulipala, V.; Reimche, J.L.; Raterman, E.; Begum, A.A.; Jerse, A.E.; Shafer, W.M. cis- and trans-Acting Factors Influence Expression of the norM-Encoded Efflux Pump of Neisseria gonorrhoeae and Levels of Gonococcal Susceptibility to Substrate Antimicrobials. Antimicrob. Agents Chemother. 2018, 62, e00821-18. [Google Scholar] [CrossRef] [Green Version]
  34. Guelfo, J.R.; Rodríguez-Rojas, A.; Matic, I.; Blázquez, J. A MATE-family efflux pump rescues the Escherichia coli 8-oxoguanine-repair-deficient mutator phenotype and protects against H(2)O(2) killing. PLoS Genet. 2010, 6, e1000931. [Google Scholar] [CrossRef] [Green Version]
  35. Tocci, N.; Iannelli, F.; Bidossi, A.; Ciusa, M.L.; Decorosi, F.; Viti, C.; Pozzi, G.; Ricci, S.; Oggioni, M.R. Functional analysis of pneumococcal drug efflux pumps associates the MATE DinF transporter with quinolone susceptibility. Antimicrob. Agents Chemother. 2013, 57, 248–253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Jiang, X.; Zhou, L.; Gao, D.; Wang, Y.; Wang, D.; Zhang, Z.; Chen, M.; Su, Y.; Li, L.; Yan, H.; et al. Expression of efflux pump gene lde in ciprofloxacin-resistant foodborne isolates of Listeria monocytogenes. Microbiol. Immunol. 2012, 56, 843–846. [Google Scholar] [CrossRef] [PubMed]
  37. Costa, S.S.; Viveiros, M.; Amaral, L.; Couto, I. Multidrug Efflux Pumps in Staphylococcus aureus: An Update. Open Microbiol. J. 2013, 7, 59–71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Bley, C.; van der Linden, M.; Reinert, R.R. mef(A) is the predominant macrolide resistance determinant in Streptococcus pneumoniae and Streptococcus pyogenes in Germany. Int. J. Antimicrob. Agents 2011, 37, 425–431. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Chancey, S.T.; Bai, X.H.; Kumar, N.; Drabek, E.F.; Daugherty, S.C.; Colon, T.; Ott, S.; Sengamalay, N.; Sadzewicz, L.; Tallon, L.J.; et al. Transcriptional Attenuation Controls Macrolide Inducible Efflux and Resistance in Streptococcus pneumoniae and in Other Gram-Positive Bacteria Containing mef/mel (msr(D)) Elements. PLoS ONE 2015, 10, e0116254. [Google Scholar] [CrossRef]
  40. Nunez-Samudio, V.; Chesneau, O. Functional interplay between the ATP binding cassette Msr(D) protein and the membrane facilitator superfamily Mef(E) transporter for macrolide resistance in Escherichia coli. Res. Microbiol. 2013, 164, 226–235. [Google Scholar] [CrossRef]
  41. Pasqua, M.; Bonaccorsi di Patti, M.C.; Fanelli, G.; Utsumi, R.; Eguchi, Y.; Trirocco, R.; Prosseda, G.; Grossi, M.; Colonna, B. Host-Bacterial Pathogen Communication: The Wily Role of the Multidrug Efflux Pumps of the MFS Family. Front. Mol. Biosci. 2021, 8, 723274. [Google Scholar] [CrossRef] [PubMed]
  42. Pérez-Varela, M.; Corral, J.; Aranda, J.; Barbé, J. Roles of Efflux Pumps from Different Superfamilies in the Surface-Associated Motility and Virulence of Acinetobacter baumannii ATCC 17978. Antimicrob. Agents Chemother. 2019, 63, e02190-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Leus, I.V.; Weeks, J.W.; Bonifay, V.; Smith, L.; Richardson, S.; Zgurskaya, H.I. Substrate specificities and efflux efficiencies of RND efflux pumps of Acinetobacter baumannii. J. Bacteriol. 2018, 200, e00049-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Perez-Boto, D.; Acebo, P.; Garcia-Pena, F.J.; Abad, J.C.; Echeita, M.A.; Amblar, M. Isolation of a point mutation associated with altered expression of the CmeABC efflux pump in a multidrug-resistant Campylobacter jejuni population of poultry origin. J. Glob. Antimicrob. Resist. 2015, 3, 115–122. [Google Scholar] [CrossRef]
  45. Dreier, J.; Ruggerone, P. Interaction of antibacterial compounds with RND efflux pumps in Pseudomonas aeruginosa. Front. Microbiol. 2015, 6, 660. [Google Scholar] [CrossRef] [Green Version]
  46. Castanheira, M.; Deshpande, L.M.; Jones, R.N.; Farrell, D.J. Evaluation of quinolone resistance-determining region mutations and efflux pump expression in Neisseria meningitidis resistant to fluoroquinolones. Diagn. Microbiol. Infect. Dis. 2012, 72, 263–266. [Google Scholar] [CrossRef]
  47. Yuan, J.; Xu, X.; Guo, Q.; Zhao, X.; Ye, X.; Guo, Y.; Wang, M. Prevalence of the oqxAB gene complex in Klebsiella pneumoniae and Escherichia coli clinical isolates. J. Antimicrob. Chemother. 2012, 67, 1655–1659. [Google Scholar] [CrossRef] [Green Version]
  48. Wong, M.H.; Chan, E.W.; Xie, L.; Li, R.; Chen, S. IncHI2 Plasmids Are the Key Vectors Responsible for oqxAB Transmission among Salmonella Species. Antimicrob. Agents Chemother. 2016, 60, 6911–6915. [Google Scholar] [CrossRef] [Green Version]
  49. Cho, H.H.; Sung, J.Y.; Kwon, K.C.; Koo, S.H. Expression of Sme efflux pumps and multilocus sequence typing in clinical isolates of Stenotrophomonas maltophilia. Ann. Lab. Med. 2012, 32, 38–43. [Google Scholar] [CrossRef]
  50. Basler, G.; Thompson, M.; Tullman-Ercek, D.; Keasling, J. A Pseudomonas putida efflux pump acts on short-chain alcohols. Biotechnol. Biofuels 2018, 11, 136. [Google Scholar] [CrossRef] [Green Version]
  51. Verma, P.; Maurya, P.; Tiwari, M.; Tiwari, V. In-silico interaction studies suggest RND efflux pump mediates polymyxin resistance in Acinetobacter baumannii. J. Biomol. Struct. Dyn. 2017, 37, 95–103. [Google Scholar] [CrossRef] [PubMed]
  52. Blair, J.M.; Bavro, V.N.; Ricci, V.; Modi, N.; Cacciotto, P.; Kleinekathfer, U.; Ruggerone, P.; Vargiu, A.V.; Baylay, A.J.; Smith, H.E.; et al. AcrB drug-binding pocket substitution confers clinically relevant resistance and altered substrate specificity. Proc. Natl. Acad. Sci. USA 2015, 112, 3511–3516. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Manjasetty, B.A.; Halavaty, A.S.; Luan, C.H.; Osipiuk, J.; Mulligan, R.; Kwon, K.; Anderson, W.F.; Joachimiak, A. Loop-to-helix transition in the structure of multidrug regulator AcrR at the entrance of the drug-binding cavity. J. Struct. Biol. 2016, 194, 18–28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Routh, M.D.; Su, C.C.; Zhang, Q.; Yu, E.W. Structures of AcrR and CmeR: Insight into the mechanisms of transcriptional repression and multi-drug recognition in the TetR family of regulators. Biochim. Biophys. Acta 2009, 1794, 844–851. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Ghosh, S.; Cremers, C.M.; Jakob, U.; Love, N.G. Chlorinated phenols control the expression of the multidrug resistance efflux pump MexAB-OprM in Pseudomonas aeruginosa by interacting with NalC. Mol. Microbiol. 2011, 79, 1547–1556. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Chen, W.; Wang, D.; Zhou, W.; Sang, H.; Liu, X.; Ge, Z.; Zhang, J.; Lan, L.; Yang, C.G.; Chen, H. Novobiocin binding to NalD induces the expression of the MexAB-OprM pump in Pseudomonas aeruginosa. Mol. Microbiol. 2016, 100, 749–758. [Google Scholar] [CrossRef] [Green Version]
  57. Fernandez-Escamilla, A.M.; Fernandez-Ballester, G.; Morel, B.; Casares-Atienza, S.; Ramos, J.L. Molecular Binding Mechanism of TtgR Repressor to Antibiotics and Antimicrobials. PLoS ONE 2015, 10, e0138469. [Google Scholar] [CrossRef]
  58. Hernandez, A.; Mate, M.J.; Sanchez-Diaz, P.C.; Romero, A.; Rojo, F.; Martinez, J.L. Structural and functional analysis of SmeT, the repressor of the Stenotrophomonas maltophilia multidrug efflux pump SmeDEF. J. Biol. Chem. 2009, 284, 14428–14438. [Google Scholar] [CrossRef] [Green Version]
  59. Johnson, P.J.; Shafer, W.M. The Transcriptional Repressor, MtrR, of the mtrCDE Efflux Pump Operon of Neisseria gonorrhoeae Can Also Serve as an Activator of “off Target” Gene (glnE) Expression. Antibiotics 2015, 4, 188–197. [Google Scholar] [CrossRef] [Green Version]
  60. Anandapadamanaban, M.; Pilstal, R.; Andresen, C.; Trewhella, J.; Moche, M.; Wallner, B.; Sunnerhagen, M. Mutation-Induced Population Shift in the MexR Conformational Ensemble Disengages DNA Binding: A Novel Mechanism for MarR Family Derepression. Structure 2016, 24, 1311–1321. [Google Scholar] [CrossRef] [Green Version]
  61. Xu, S.; Chen, G.; Liu, Z.; Xu, D.; Wu, Z.; Li, Z.; Hong, M. Site-Directed Mutagenesis Reveals Crucial Residues in Escherichia coli Resistance-Nodulation-Division Efflux Pump OqxB. Microb. Drug Resist. 2020, 26, 550–560. [Google Scholar] [CrossRef]
  62. Lv, Z.; Zhao, M.; Wang, W.; Wang, Q.; Huang, M.; Li, C.; Lian, Q.; Xia, J.; Qi, J.; Xiang, C.; et al. Changing Gly311 to an acidic amino acid in the MATE family protein DTX6 enhances Arabidopsis resistance to the dihydropyridine herbicides. Mol. Plant 2021, 14, 2115–2125. [Google Scholar] [CrossRef] [PubMed]
  63. Bhakdi, S.; Mackman, N.; Menestrina, G.; Gray, L.; Hugo, F.; Seeger, W.; Holland, I.B. The hemolysin of Escherichia coli. Eur. J. Epidemiol. 1988, 4, 135–143. [Google Scholar] [CrossRef] [PubMed]
  64. Gilson, L.; Mahanty, H.K.; Kolter, R. Genetic analysis of an MDR-like export system: The secretion of colicin V. EMBO J. 1990, 9, 3875–3884. [Google Scholar] [CrossRef] [PubMed]
  65. Wehmeier, C.; Schuster, S.; Fähnrich, E.; Kern, W.V.; Bohnert, J.A. Site-directed mutagenesis reveals amino acid residues in the Escherichia coli RND efflux pump AcrB that confer macrolide resistance. Antimicrob. Agents Chemother. 2009, 53, 329–330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Wright, D.J.; Tate, C.G. Isolation and characterisation of transport-defective substrate-binding mutants of the tetracycline antiporter TetA(B). Biochim. Biophys. Acta 2015, 1848, 2261–2270. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Guérin, F.; Galimand, M.; Tuambilangana, F.; Courvalin, P.; Cattoir, V. Overexpression of the novel MATE fluoroquinolone efflux pump FepA in Listeria monocytogenes is driven by inactivation of its local repressor FepR. PLoS ONE 2014, 9, e106340. [Google Scholar] [CrossRef] [Green Version]
  68. Singh, A.K.; Haldar, R.; Mandal, D.; Kundu, M. Analysis of the topology of Vibrio cholerae NorM and identification of amino acid residues involved in norfloxacin resistance. Antimicrob. Agents Chemother. 2006, 50, 3717–3723. [Google Scholar] [CrossRef] [Green Version]
  69. Delmar, J.A.; Su, C.C.; Yu, E.W. Bacterial multidrug efflux transporters. Annu. Rev. Biophys. 2014, 43, 93–117. [Google Scholar] [CrossRef] [Green Version]
  70. Long, F.; Su, C.C.; Lei, H.T.; Bolla, J.R.; Do, S.V.; Yu, E.W. Structure and mechanism of the tripartite CusCBA heavy-metal efflux complex. Philos. Trans. R. Soc. Lond. Ser. B Biol. Sci. 2012, 367, 1047–1058. [Google Scholar] [CrossRef]
  71. Blanco, P.; Hernando-Amado, S.; Reales-Calderon, J.A.; Corona, F.; Lira, F.; Alcalde-Rico, M.; Bernardini, A.; Sanchez, M.B.; Martinez, J.L. Bacterial Multidrug Efflux Pumps: Much More Than Antibiotic Resistance Determinants. Microorganisms 2016, 4, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Zwama, M.; Nishino, K. Ever-Adapting RND Efflux Pumps in Gram-Negative Multidrug-Resistant Pathogens: A Race against Time. Antibiotics 2021, 10, 774. [Google Scholar] [CrossRef] [PubMed]
  73. Bay, D.C.; Rommens, K.L.; Turner, R.J. Small multidrug resistance proteins: A multidrug transporter family that continues to grow. Biochim. Biophys. Acta 2008, 1778, 1814–1838. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Bay, D.C.; Turner, R.J. Diversity and evolution of the small multidrug resistance protein family. BMC Evol. Biol. 2009, 9, 140. [Google Scholar] [CrossRef] [Green Version]
  75. Lytvynenko, I.; Brill, S.; Oswald, C.; Pos, K.M. Molecular basis of polyspecificity of the Small Multidrug Resistance Efflux Pump AbeS from Acinetobacter baumannii. J. Mol. Biol. 2016, 428, 644–657. [Google Scholar] [CrossRef]
  76. Lin, M.F.; Lin, Y.Y.; Tu, C.C.; Lan, C.Y. Distribution of different efflux pump genes in clinical isolates of multidrug-resistant Acinetobacter baumannii and their correlation with antimicrobial resistance. J. Microbiol. Immunol. Infect. = Wei Mian Yu Gan Ran Za Zhi 2017, 50, 224–231. [Google Scholar] [CrossRef] [Green Version]
  77. Srinivasan, V.B.; Rajamohan, G. KpnEF, a new member of the Klebsiella pneumoniae cell envelope stress response regulon, is an SMR-type efflux pump involved in broad-spectrum antimicrobial resistance. Antimicrob. Agents Chemother. 2013, 57, 4449–4462. [Google Scholar] [CrossRef] [Green Version]
  78. Banigan, J.R.; Gayen, A.; Cho, M.K.; Traaseth, N.J. A structured loop modulates coupling between the substrate-binding and dimerization domains in the multidrug resistance transporter EmrE. J. Biol. Chem. 2015, 290, 805–814. [Google Scholar] [CrossRef] [Green Version]
  79. Padariya, M.; Kalathiya, U.; Baginski, M. Structural and dynamic insights on the EmrE protein with TPP(+) and related substrates through molecular dynamics simulations. Chem. Phys. Lipids 2017, 212, 1–11. [Google Scholar] [CrossRef]
  80. Jaglic, Z.; Cervinkova, D. Genetic basis of resistance to quaternary ammonium compounds—The qac genes and their role: A review. Vet. Med. 2012, 57, 275–281. [Google Scholar] [CrossRef] [Green Version]
  81. Buffet-Bataillon, S.; Tattevin, P.; Maillard, J.Y.; Bonnaure-Mallet, M.; Jolivet-Gougeon, A. Efflux pump induction by quaternary ammonium compounds and fluoroquinolone resistance in bacteria. Future Microbiol. 2016, 11, 81–92. [Google Scholar] [CrossRef] [PubMed]
  82. 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]
  83. Hassan, K.A.; Elbourne, L.D.; Li, L.; Gamage, H.K.; Liu, Q.; Jackson, S.M.; Sharples, D.; Kolsto, A.B.; Henderson, P.J.; Paulsen, I.T. An ace up their sleeve: A transcriptomic approach exposes the AceI efflux protein of Acinetobacter baumannii and reveals the drug efflux potential hidden in many microbial pathogens. Front. Microbiol. 2015, 6, 333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Coenye, T.; Van Acker, H.; Peeters, E.; Sass, A.; Buroni, S.; Riccardi, G.; Mahenthiralingam, E. Molecular mechanisms of chlorhexidine tolerance in Burkholderia cenocepacia biofilms. Antimicrob. Agents Chemother. 2011, 55, 1912–1919. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Nde, C.W.; Jang, H.J.; Toghrol, F.; Bentley, W.E. Global transcriptomic response of Pseudomonas aeruginosa to chlorhexidine diacetate. Environ. Sci. Technol. 2009, 43, 8406–8415. [Google Scholar] [CrossRef]
  86. Lin, H.T.; Bavro, V.N.; Barrera, N.P.; Frankish, H.M.; Velamakanni, S.; van Veen, H.W.; Robinson, C.V.; Borges-Walmsley, M.I.; Walmsley, A.R. MacB ABC transporter is a dimer whose ATPase activity and macrolide-binding capacity are regulated by the membrane fusion protein MacA. J. Biol. Chem. 2009, 284, 1145–1154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Fluman, N.; Adler, J.; Rotenberg, S.A.; Brown, M.H.; Bibi, E. Export of a single drug molecule in two transport cycles by a multidrug efflux pump. Nat. Commun. 2014, 5, 4615. [Google Scholar] [CrossRef] [Green Version]
  88. Fluman, N.; Bibi, E. Bacterial multidrug transport through the lens of the major facilitator superfamily. Biochim. Biophys. Acta 2009, 1794, 738–747. [Google Scholar] [CrossRef]
  89. Rincón, G.; Radice, M.; Giovanakis, M.; Di Conza, J.A.; Gutkind, G. First report of plasmid-mediated fluoroquinolone efflux pump QepA in Escherichia coli clinical isolate ST68, in South America. Diagn. Microbiol. Infect. Dis. 2014, 79, 70–72. [Google Scholar] [CrossRef]
  90. Atac, N.; Kurt-Azap, O.; Dolapci, I.; Yesilkaya, A.; Ergonul, O.; Gonen, M.; Can, F. The Role of AcrAB-TolC Efflux Pumps on Quinolone Resistance of E. coli ST131. Curr. Microbiol. 2018, 75, 1661–1666. [Google Scholar] [CrossRef] [Green Version]
  91. Saw, H.T.; Webber, M.A.; Mushtaq, S.; Woodford, N.; Piddock, L.J. Inactivation or inhibition of AcrAB-TolC increases resistance of carbapenemase-producing Enterobacteriaceae to carbapenems. J. Antimicrob. Chemother. 2016, 71, 1510–1519. [Google Scholar] [CrossRef] [PubMed]
  92. Dabul, A.N.G.; Avaca-Crusca, J.S.; Van Tyne, D.; Gilmore, M.S.; Camargo, I. Resistance in In Vitro Selected Tigecycline-Resistant Methicillin-Resistant Staphylococcus aureus Sequence Type 5 Is Driven by Mutations in mepR and mepA Genes. Microb. Drug Resist. 2018, 24, 519–526. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Banchs, C.; Poulos, S.; Nimjareansuk, W.S.; Joo, Y.E.; Faham, S. Substrate binding to the multidrug transporter MepA. Biochim. Biophys. Acta 2014, 1838, 2539–2546. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Hall, C.W.; Mah, T.F. Molecular mechanisms of biofilm-based antibiotic resistance and tolerance in pathogenic bacteria. FEMS Microbiol. Rev. 2017, 41, 276–301. [Google Scholar] [CrossRef]
  95. 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] [Green Version]
  96. Chen, H.L.; Cao, J.M.; Zhou, C.; Liu, H.Y.; Zhang, X.X.; Zhou, T.L. Biofilm Formation Restrained by Subinhibitory Concentrations of Tigecyclin in Acinetobacter baumannii Is Associated with Downregulation of Efflux Pumps. Chemotherapy 2017, 62, 128–133. [Google Scholar] [CrossRef]
  97. Gupta, D.; Singh, A.; Khan, A.U. Nanoparticles as Efflux Pump and Biofilm Inhibitor to Rejuvenate Bactericidal Effect of Conventional Antibiotics. Nanoscale Res. Lett. 2017, 12, 454. [Google Scholar] [CrossRef]
  98. Sabatini, S.; Piccioni, M.; Felicetti, T.; De Marco, S.; Manfroni, G.; Pagiotti, R.; Nocchetti, M.; Cecchetti, V.; Pietrella, D. Investigation on the effect of known potent S. aureus NorA efflux pump inhibitors on the staphylococcal biofilm formation. RSC Adv. 2017, 7, 37007–37014. [Google Scholar] [CrossRef] [Green Version]
  99. Robin, B.; Nicol, M.; Le, H.; Tahrioui, A.; Schaumann, A.; Vuillemenot, J.B.; Vergoz, D.; Lesouhaitier, O.; Jouenne, T.; Hardouin, J.; et al. MacAB-TolC Contributes to the Development of Acinetobacter baumannii Biofilm at the Solid-Liquid Interface. Front. Microbiol. 2021, 12, 785161. [Google Scholar] [CrossRef]
  100. Yoon, E.J.; Balloy, V.; Fiette, L.; Chignard, M.; Courvalin, P.; Grillot-Courvalin, C. Contribution of the Ade Resistance-Nodulation-Cell Division-Type Efflux Pumps to Fitness and Pathogenesis of Acinetobacter baumannii. mBio 2016, 7, e00697-16. [Google Scholar] [CrossRef] [Green Version]
  101. Dou, Y.; Song, F.; Guo, F.; Zhou, Z.D.; Zhu, C.L.; Xiang, J.; Huan, J.N. Acinetobacter baumannii quorum-sensing signalling molecule induces the expression of drug-resistance genes. Mol. Med. Rep. 2017, 15, 4061–4068. [Google Scholar] [CrossRef] [Green Version]
  102. Richmond, G.E.; Evans, L.P.; Anderson, M.J.; Wand, M.E.; Bonney, L.C.; Ivens, A.; Chua, K.L.; Webber, M.A.; Sutton, J.M.; Peterson, M.L.; et al. The Acinetobacter baumannii Two-Component System AdeRS Regulates Genes Required for Multidrug Efflux, Biofilm Formation, and Virulence in a Strain-Specific Manner. mBio 2016, 7, e00430-16. [Google Scholar] [CrossRef] [Green Version]
  103. Sato, Y.; Unno, Y.; Ubagai, T.; Ono, Y. Sub-minimum inhibitory concentrations of colistin and polymyxin B promote Acinetobacter baumannii biofilm formation. PLoS ONE 2018, 13, e0194556. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. 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] [Green Version]
  105. Liang, Z.B.; Chen, Y.M.; Chen, Y.; Cheng, Y.Y.; Zhang, L.H. RND efflux pump and its interrelationship with quorum sensing system. Yi Chuan = Hered. 2016, 38, 894–901. [Google Scholar] [CrossRef]
  106. Wolloscheck, D.; Krishnamoorthy, G.; Nguyen, J.; Zgurskaya, H.I. Kinetic Control of Quorum Sensing in Pseudomonas aeruginosa by Multidrug Efflux Pumps. ACS Infect. Dis. 2018, 4, 185–195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Cannon, R.D.; Holmes, A.R. Learning the ABC of oral fungal drug resistance. Mol. Oral. Microbiol. 2015, 30, 425–437. [Google Scholar] [CrossRef] [PubMed]
  108. Santiago, A.E.; Yan, M.B.; Tran, M.; Wright, N.; Luzader, D.H.; Kendall, M.M.; Ruiz-Perez, F.; Nataro, J.P. A large family of anti-activators accompanying XylS/AraC family regulatory proteins. Mol. Microbiol. 2016, 101, 314–332. [Google Scholar] [CrossRef]
  109. Deochand, D.K.; Grove, A. MarR family transcription factors: Dynamic variations on a common scaffold. Crit. Rev. Biochem. Mol. Biol. 2017, 52, 595–613. [Google Scholar] [CrossRef]
  110. Chang, C.C.; Lin, L.Y.; Zou, X.W.; Huang, C.C.; Chan, N.L. Structural basis of the mercury(II)-mediated conformational switching of the dual-function transcriptional regulator MerR. Nucleic Acids Res. 2015, 43, 7612–7623. [Google Scholar] [CrossRef] [Green Version]
  111. Deng, W.; Li, C.; Xie, J. The underling mechanism of bacterial TetR/AcrR family transcriptional repressors. Cell Signal 2013, 25, 1608–1613. [Google Scholar] [CrossRef] [PubMed]
  112. Hwang, S.; Zhang, Q.; Ryu, S.; Jeon, B. Transcriptional regulation of the CmeABC multidrug efflux pump and the KatA catalase by CosR in Campylobacter jejuni. J. Bacteriol. 2012, 194, 6883–6891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Juarez, P.; Jeannot, K.; Plesiat, P.; Llanes, C. Toxic Electrophiles Induce Expression of the Multidrug Efflux Pump MexEF-OprN in Pseudomonas aeruginosa through a Novel Transcriptional Regulator, CmrA. Antimicrob. Agents Chemother. 2017, 61, e00585-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Hernandez, A.; Ruiz, F.M.; Romero, A.; Martinez, J.L. The binding of triclosan to SmeT, the repressor of the multidrug efflux pump SmeDEF, induces antibiotic resistance in Stenotrophomonas maltophilia. PLoS Pathog. 2011, 7, e1002103. [Google Scholar] [CrossRef] [Green Version]
  115. Ruiz, C.; Levy, S.B. Regulation of acrAB expression by cellular metabolites in Escherichia coli. J. Antimicrob. Chemother. 2014, 69, 390–399. [Google Scholar] [CrossRef] [Green Version]
  116. Sun, J.R.; Chiang, Y.S.; Shang, H.S.; Perng, C.L.; Yang, Y.S.; Chiueh, T.S. Phenotype microarray analysis of the AdeRS two-component system in Acinetobacter baumannii. Eur. J. Clin. Microbiol. Infect. Dis. 2017, 36, 2343–2353. [Google Scholar] [CrossRef]
  117. Lin, M.F.; Lin, Y.Y.; Lan, C.Y. The Role of the Two-Component System BaeSR in Disposing Chemicals through Regulating Transporter Systems in Acinetobacter baumannii. PLoS ONE 2015, 10, e0132843. [Google Scholar] [CrossRef]
  118. Wu, C.J.; Huang, Y.W.; Lin, Y.T.; Ning, H.C.; Yang, T.C. Inactivation of SmeSyRy Two-Component Regulatory System Inversely Regulates the Expression of SmeYZ and SmeDEF Efflux Pumps in Stenotrophomonas maltophilia. PLoS ONE 2016, 11, e0160943. [Google Scholar] [CrossRef]
  119. Lin, Y.T.; Huang, Y.W.; Liou, R.S.; Chang, Y.C.; Yang, T.C. MacABCsm, an ABC-type tripartite efflux pump of Stenotrophomonas maltophilia involved in drug resistance, oxidative and envelope stress tolerances and biofilm formation. J. Antimicrob. Chemother. 2014, 69, 3221–3226. [Google Scholar] [CrossRef]
  120. Walker, J.N.; Crosby, H.A.; Spaulding, A.R.; Salgado-Pabon, W.; Malone, C.L.; Rosenthal, C.B.; Schlievert, P.M.; Boyd, J.M.; Horswill, A.R. The Staphylococcus aureus ArlRS two-component system is a novel regulator of agglutination and pathogenesis. PLoS Pathog. 2013, 9, e1003819. [Google Scholar] [CrossRef]
  121. Nishino, K.; Yamasaki, S.; Nakashima, R.; Zwama, M.; Hayashi-Nishino, M. Function and Inhibitory Mechanisms of Multidrug Efflux Pumps. Front. Microbiol. 2021, 12, 737288. [Google Scholar] [CrossRef] [PubMed]
  122. Nishino, K. Regulation of the Expression of Bacterial Multidrug Exporters by Two-Component Signal Transduction Systems. Methods Mol. Biol. 2018, 1700, 239–251. [Google Scholar] [CrossRef] [PubMed]
  123. Henderson, P.J.F.; Maher, C.; Elbourne, L.D.H.; Eijkelkamp, B.A.; Paulsen, I.T.; Hassan, K.A. Physiological Functions of Bacterial “Multidrug” Efflux Pumps. Chem. Rev. 2021, 121, 5417–5478. [Google Scholar] [CrossRef] [PubMed]
  124. Li, X.Z.; Plesiat, P.; Nikaido, H. The challenge of efflux-mediated antibiotic resistance in Gram-negative bacteria. Clin. Microbiol. Rev. 2015, 28, 337–418. [Google Scholar] [CrossRef] [Green Version]
  125. Goli, H.R.; Nahaei, M.R.; Rezaee, M.A.; Hasani, A.; Samadi Kafil, H.; Aghazadeh, M.; Sheikhalizadeh, V. Contribution of mexAB-oprM and mexXY (-oprA) efflux operons in antibiotic resistance of clinical Pseudomonas aeruginosa isolates in Tabriz, Iran. Infect. Genet. Evol. J. Mol. Epidemiol. Evol. Genet. Infect. Dis. 2016, 45, 75–82. [Google Scholar] [CrossRef]
  126. Murugan, N.; Malathi, J.; Therese, K.L.; Madhavan, H.N. Application of six multiplex PCR’s among 200 clinical isolates of Pseudomonas aeruginosa for the detection of 20 drug resistance encoding genes. Kaohsiung J. Med. Sci. 2018, 34, 79–88. [Google Scholar] [CrossRef]
  127. Zhang, Y.; Li, Z.; He, X.; Ding, F.; Wu, W.; Luo, Y.; Fan, B.; Cao, H. Overproduction of efflux pumps caused reduced susceptibility to carbapenem under consecutive imipenem-selected stress in Acinetobacter baumannii. Infect. Drug Resist. 2017, 11, 457–467. [Google Scholar] [CrossRef] [Green Version]
  128. Gerson, S.; Nowak, J.; Zander, E.; Ertel, J.; Wen, Y.; Krut, O.; Seifert, H.; Higgins, P.G. Diversity of mutations in regulatory genes of resistance-nodulation-cell division efflux pumps in association with tigecycline resistance in Acinetobacter baumannii. J. Antimicrob. Chemother. 2018, 73, 1501–1508. [Google Scholar] [CrossRef]
  129. Nagano, K.; Nikaido, H. Kinetic behavior of the major multidrug efflux pump AcrB of Escherichia coli. Proc. Natl. Acad. Sci. USA 2009, 106, 5854–5858. [Google Scholar] [CrossRef] [Green Version]
  130. Wang, J.; Guo, Z.W.; Zhi, C.P.; Yang, T.; Zhao, J.J.; Chen, X.J.; Zeng, L.; Lv, L.C.; Zeng, Z.L.; Liu, J.H. Impact of plasmid-borne oqxAB on the development of fluoroquinolone resistance and bacterial fitness in Escherichia coli. J. Antimicrob. Chemother. 2017, 72, 1293–1302. [Google Scholar] [CrossRef] [Green Version]
  131. Huang, J.J.; Hu, H.Y.; Wu, Y.H.; Wei, B.; Lu, Y. Effect of chlorination and ultraviolet disinfection on tetA-mediated tetracycline resistance of Escherichia coli. Chemosphere 2013, 90, 2247–2253. [Google Scholar] [CrossRef] [PubMed]
  132. Melamed, S.; Naftaly, S.; Belkin, S. Improved detection of antibiotic compounds by bacterial reporter strains achieved by manipulations of membrane permeability and efflux capacity. Appl. Microbiol. Biotechnol. 2014, 98, 2267–2277. [Google Scholar] [CrossRef]
  133. Hong, H.; Park, W. TetR repressor-based bioreporters for the detection of doxycycline using Escherichia coli and Acinetobacter oleivorans. Appl. Microbiol. Biotechnol. 2014, 98, 5039–5050. [Google Scholar] [CrossRef] [PubMed]
  134. Weber, C.C.; Link, N.; Fux, C.; Zisch, A.H.; Weber, W.; Fussenegger, M. Broad-spectrum protein biosensors for class-specific detection of antibiotics. Biotechnol. Bioeng. 2005, 89, 9–17. [Google Scholar] [CrossRef] [PubMed]
  135. Espinosa-Urgel, M.; Serrano, L.; Ramos, J.L.; Fernandez-Escamilla, A.M. Engineering Biological Approaches for Detection of Toxic Compounds: A New Microbial Biosensor Based on the Pseudomonas putida TtgR Repressor. Mol. Biotechnol. 2015, 57, 558–564. [Google Scholar] [CrossRef]
  136. Alibert, S.; Diarra, J.N.; Hernandez, J.; Stutzmann, A.; Fouad, M.; Boyer, G.; Pages, J.M. Multidrug efflux pumps and their role in antibiotic and antiseptic resistance: A pharmacodynamic perspective. Expert Opin. Drug Met. 2017, 13, 301–309. [Google Scholar] [CrossRef]
  137. Opperman, T.J.; Nguyen, S. Recent advances toward a molecular mechanism of efflux pump inhibition. Front. Microbiol. Front. Microbiol. 2015, 6, 421. [Google Scholar] [CrossRef] [Green Version]
  138. Jamshidi, S.; Sutton, J.M.; Rahman, K.M. An overview of bacterial efflux pumps and computational approaches to study efflux pump inhibitors. Future Med. Chem. 2016, 8, 195–210. [Google Scholar] [CrossRef]
  139. Kesherwani, M.; Michael Gromiha, M.; Fukui, K.; Velmurugan, D. Identification of novel natural inhibitor for NorM—A multidrug and toxic compound extrusion transporter—An insilico molecular modeling and simulation studies. J. Biomol. Struct. Dyn. 2017, 35, 58–77. [Google Scholar] [CrossRef]
  140. Kinana, A.D.; Vargiu, A.V.; May, T.; Nikaido, H. Aminoacyl beta-naphthylamides as substrates and modulators of AcrB multidrug efflux pump. Proc. Natl. Acad. Sci. USA 2016, 113, 1405–1410. [Google Scholar] [CrossRef] [Green Version]
  141. Donner, J.; Reck, M.; Bunk, B.; Jarek, M.; App, C.B.; Meier-Kolthoff, J.P.; Overmann, J.; Muller, R.; Kirschning, A.; Wagner-Dobler, I. The Biofilm Inhibitor Carolacton Enters Gram-Negative Cells: Studies Using a TolC-Deficient Strain of Escherichia coli. mSphere 2017, 2, e00375-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Tintino, S.R.; Morais-Tintino, C.D.; Campina, F.F.; Costa, M.D.; Menezes, I.R.A.; de Matos, Y.M.L.S.; Calixto, J.T.; Pereira, P.S.; Siqueira, J.P.; Leal-Balbino, T.C.; et al. Tannic acid affects the phenotype of Staphylococcus aureus resistant to tetracycline and erythromycin by inhibition of efflux pumps. Bioorganic Chem. 2017, 74, 197–200. [Google Scholar] [CrossRef] [PubMed]
  143. Yang, X.; Goswami, S.; Gorityala, B.K.; Dornalaon, R.; Lyu, Y.F.; Kumar, A.; Zhanel, G.G.; Schweizer, F. A Tobramycin Vector Enhances Synergy and Efficacy of Efflux Pump Inhibitors against Multidrug-Resistant Gram-Negative Bacteria. J. Med. Chem. 2017, 60, 3913–3932. [Google Scholar] [CrossRef] [PubMed]
  144. Yang, X.; Domalaon, R.; Lyu, Y.; Zhanel, G.G.; Schweizer, F. Tobramycin-Linked Efflux Pump Inhibitor Conjugates Synergize Fluoroquinolones, Rifampicin and Fosfomycin against Multidrug-Resistant Pseudomonas aeruginosa. J. Clin. Med. 2018, 7, 158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. 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] [Green Version]
  146. Prasch, S.; Bucar, F. Plant derived inhibitors of bacterial efflux pumps: An update. Phytochem. Rev. 2015, 14, 961–974. [Google Scholar] [CrossRef]
  147. Shiu, W.K.; Malkinson, J.P.; Rahman, M.M.; Curry, J.; Stapleton, P.; Gunaratnam, M.; Neidle, S.; Mushtaq, S.; Warner, M.; Livermore, D.M.; et al. A new plant-derived antibacterial is an inhibitor of efflux pumps in Staphylococcus aureus. Int. J. Antimicrob. Agents 2013, 42, 513–518. [Google Scholar] [CrossRef]
  148. Garvey, M.I.; Rahman, M.M.; Gibbons, S.; Piddock, L.J. Medicinal plant extracts with efflux inhibitory activity against Gram-negative bacteria. Int. J. Antimicrob. Agents 2011, 37, 145–151. [Google Scholar] [CrossRef] [Green Version]
  149. Laudadio, E.; Cedraro, N.; Mangiaterra, G.; Citterio, B.; Mobbili, G.; Minnelli, C.; Bizzaro, D.; Biavasco, F.; Galeazzi, R. Natural Alkaloid Berberine Activity against Pseudomonas aeruginosa MexXY-Mediated Aminoglycoside Resistance: In Silico and in Vitro Studies. J. Nat. Prod. 2019, 82, 1935–1944. [Google Scholar] [CrossRef]
  150. Negi, N.; Prakash, P.; Gupta, M.L.; Mohapatra, T.M. Possible Role of Curcumin as an Efflux Pump Inhibitor in Multi Drug Resistant Clinical Isolates of Pseudomonas aeruginosa. J. Clin. Diagn. Res. 2014, 8, DC04–DC07. [Google Scholar] [CrossRef]
  151. Li, B.; Yao, Q.; Pan, X.C.; Wang, N.; Zhang, R.; Li, J.; Ding, G.; Liu, X.; Wu, C.; Ran, D.; et al. Artesunate enhances the antibacterial effect of {beta}-lactam antibiotics against Escherichia coli by increasing antibiotic accumulation via inhibition of the multidrug efflux pump system AcrAB-TolC. J. Antimicrob. Chemother. 2011, 66, 769–777. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Muniz, D.F.; Dos Santos Barbosa, C.R.; de Menezes, I.R.A.; de Sousa, E.O.; Pereira, R.L.S.; Junior, J.T.C.; Pereira, P.S.; de Matos, Y.; da Costa, R.H.S.; de Morais Oliveira-Tintino, C.D.; et al. In vitro and in silico inhibitory effects of synthetic and natural eugenol derivatives against the NorA efflux pump in Staphylococcus Aureus. Food Chem. 2021, 337, 127776. [Google Scholar] [CrossRef] [PubMed]
  153. Solnier, J.; Martin, L.; Bhakta, S.; Bucar, F. Flavonoids as Novel Efflux Pump Inhibitors and Antimicrobials Against Both Environmental and Pathogenic Intracellular Mycobacterial Species. Molecules 2020, 25, 734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Tariq, A.; Sana, M.; Shaheen, A.; Ismat, F.; Mahboob, S.; Rauf, W.; Mirza, O.; Iqbal, M.; Rahman, M. Restraining the multidrug efflux transporter STY4874 of Salmonella Typhi by reserpine and plant extracts. Lett. Appl. Microbiol. 2019, 69, 161–167. [Google Scholar] [CrossRef] [PubMed]
  155. Li, J.; Liu, D.; Tian, X.; Koseki, S.; Chen, S.; Ye, X.; Ding, T. Novel antibacterial modalities against methicillin resistant Staphylococcus aureus derived from plants. Crit. Rev. Food Sci. Nutr. 2019, 59, S153–S161. [Google Scholar] [CrossRef]
  156. Karumathil, D.P.; Nair, M.S.; Gaffney, J.; Kollanoor-Johny, A.; Venkitanarayanan, K. Trans-Cinnamaldehyde and Eugenol Increase Acinetobacter baumannii Sensitivity to Beta-Lactam Antibiotics. Front. Microbiol. 2018, 9, 1011. [Google Scholar] [CrossRef]
  157. Chambers, C.S.; Viktorová, J.; Řehořová, K.; Biedermann, D.; Turková, L.; Macek, T.; Křen, V.; Valentová, K. Defying Multidrug Resistance! Modulation of Related Transporters by Flavonoids and Flavonolignans. J. Agric. Food Chem. 2020, 68, 1763–1779. [Google Scholar] [CrossRef]
Table 1. Efflux pumps regulated by the TetR family.
Table 1. Efflux pumps regulated by the TetR family.
MicroorganismsEfflux PumpRegulatorsCrucial Amino Acid ResiduesSubstratesReferences
C. jejuniCmeABCCmeR-Multidrug[54]
E. coliAcrB
TetA
AcrR
TetR
Gly-616
His-64, Thr-103, Arg-104, Pro-105
Macrolide
Tetracyclines and Mg2+ complex
[65,66]
L. monocytogenesFepAFepR-Fluoroquinolones[67]
N. gonorrhoeaeMtrCDEMtrR-Hydrophobic antibiotics[59]
V. choleraeNorM-Glu-124, Glu190, Asp-155, Gly-184, Gly-187, Lys-185, Pro-189, Cys-196, and Tyr-384Norfloxacin[68]
P. aeruginosaMexAB-OprMMexR-Novobiocin[55,56]
P. putidaTtgABCTtgRSer-77, Glu-78, Asn-110, His-114Tetracyclines, Chloramphenicol[57]
S. altophiliaSmeDEFSmeTHis-67, Ser-96, His-167Tetracyclines, Chloramphenicol,
Quinolones
[58]
Table 2. Efflux pumps present in Gram-negative bacteria.
Table 2. Efflux pumps present in Gram-negative bacteria.
Efflux Pump FamilyEfflux PumpRegulatorOrganismsSubstrates (Class)Resistance to Specific Antibiotics aReferences
ABCMacAB-TolCPhoPQE. coli,
N. gonorrhoeae,
S. maltophilia
MacrolidesEM[18,86]
MATENorM/N. gonorrhoeaeFluoroquinolones, EB, Rhodamine 6GNF, CP[29,33]
MFSEmrAB-TolCEmrRE. coliCotrimoxazole/[40]
MdfA, MdtM/Tigecycline, chloramphenicolDC, CM[86,87,88]
QepAQepRFluoroquinolonesFQ[89]
TetATetRTigecyclineTC[66]
RNDAcrAB-TolCAcrRE. coli,
K. pneumoniae,
S. enterica
β-lactams, FluoroquinolonesKF, CM, FQ, P[90,91]
AdeABCAadR, AadSA. baumanniiAminoglycosides, Erythromycin, Chloramphenicol, Fluoroquinolones, Tetracyclines, Trimethoprim, some β-lactams, Bile saltsAZI[43]
CmeABCCmeRC. jejuniβ-lactams, Tetracyclines, QuinolonesTC[44]
MexAB-OprMNalC/NalDP. aeruginosaQuinolonesCM, CP, TC, SM[45]
MtrCDEMtrR, MtrRN. gonorrhoeaeFluoroquinolonesCP, RF[46]
OqxABOqxRE. coli,
K. pneumoniae,
S. enterica
Chloramphenicol, FluoroquinolonesCM, NT, NF, CP, LEV[47,48]
SmeDEFSmeTS. maltophiliaAminoglycosides, Trimethoprim
Tetracyclines, Chloramphenicol
GM, CZ, IMP, MP, CAR, TC[49]
TtgABCTtgRP. putidaChloramphenicolCM, TC[50]
SMRAbeS/A. baumanniiEthidium, Acriflavine, BenzalkoniumEM, NO[72,73]
EmrE/E. coli,
P. aeruginosa
Quaternary ammonium compoundsQuaternary ammonium compounds[75,76]
KpnEF/K. pneumoniaeBenzalkonium chloride, ChlorhexidineCT, EM, RF, TC, SM[74]
a ACR, acriflavine; ADM, adriamycin; AG, aminoglycosides; AZI, azithromycin; CAR, carbenicillin; CLI, clindamycin; CM, chloramphenicol; CP, ciprofloxacin; CR, clarithromycin; CT, Colistin; CZ, cefoperazone; DA, dalfopristin; DAU, daunomycin; DC, doxycycline; DTM, distamycin; EB, ethidium bromide; EM, erythromycin; FQ, fluoroquinolones; FU, fusidic acid; IMP, imipenem; KF, cephalosporins; GM, gentamicin; LEV, levofloxacin; ML, macrolides; MP, meropenem; NF, norfloxacin; NO, novobiocin; NT, nitrofurantoin; OF, ofloxacin; P, penicillins; RF, rifampicin; ROX, roxithromycin; SM, streptomycin; TC, tetracycline; TM, trimethoprim. “/” mean no transcription regulators found.
Table 3. Efflux pumps present in Gram-positive bacteria.
Table 3. Efflux pumps present in Gram-positive bacteria.
Efflux Pump FamilyEfflux PumpRegulatorOrganismsSubstratesResistance to Specific Antibiotics aReference
ABCEfrAB/E. faecalisacriflavine, ethidium bromide, safranin O, DAPI, daunomycin, doxorubicin, novobiocin, arbekacin, doxycycline and norfloxacinGM, SM, CM[24]
LmrA/L. lactisMacrolides, Lincosamides, StreptograminsDAU, ADM[21,25]
Msr/StreptococcusMacrolidesML[22,26,27]
PatA/PatB/S. pneumoniaeeFluoroquinolonesFQ[23,28]
MATEFepAFepRL. monocytogenesFluoroquinolonesNF, CP[64]
MepAMepRS. aureusFluoroquinolones, Tigecycline, PentamidineDT[92,93]
MFSLde/L. monocytogenesFluoroquinolonesACR, EB[36]
NorA, NorB, NorCMgrA, NorG, ArlRSS. aureusFluoroquinolonesNF, CP[37]
Mef/S. pneumoniaeMacrolidesEM, AZI, ROX, CR[38,39,40]
SMRQacQacRS. aureus,
Enterococcus spp.,
E. faecalis
Quaternary ammonium compoundsQuaternary ammonium compounds[77]
a ACR, acriflavine; ADM, adriamycin; AZI, azithromycin; CP, ciprofloxacin; CR, clarithromycin; CM, chloramphenicol; DA, dalfopristin; DAU, daunomycin; DT, dirithromycin; EB, ethidium bromide; EM, erythromycin; GM, gentamicin; FQ, fluoroquinolones; ML, macrolides; NF, norfloxacin; ROX, roxithromycin; SM, streptomycin.
Table 4. Plant-derived EPIs.
Table 4. Plant-derived EPIs.
Bioactive CompoundsBacterial SpeciesPharmacological ActivityReferences
BerberineP. aeruginosaInhibited the multidrug efflux system MexXY-OprM[149]
ArtesunateE. coliInhibited the multidrug efflux pump system AcrAB-TolC[151]
CurcuminP. aeruginosaInhibited the expression of efflux pump[150]
plant-derived flavonoids such as skullcapflavone II and nobiletinMycobacterial SpeciesInhibited the activity of the efflux pump and decreased the rifampicin-resistance level[153]
Extracts of milk thistle seeds and reserpineSalmonella TyphiInhibited an efflux transporter STY4874[154]
Hypericum olympicum L. cf. uniflorum-derived natural productS. aureusInhibited NorA multidrug efflux pump activity[147]
diterpene isolated from Chamaecyparis lawsoniana: ferruginolMethicillin-resistant S. aureus (MRSA)Inhibited the TetK pump[155]
quinine isolated from Cinchona tree’s barkInhibited the activity of the efflux pump
piperine isolated from the Piperaceae family
harmaline isolated from Perganum harmala
4′,5′-O-dicaffeoylquinic acid isolated from wormwood (Artemisia absinthium)
triterpenoids from Momordica balsamina
carnosic acid from Rosmarinus officinalis
coumarins from Mesua ferrea
clerodane diterpene from Polyalthia longifoliaDownregulation of MFS and MATE family efflux genes such as norA, norB, norC, mdeA, and mepA
cumin spice (Cuminum cyminum)inhibited LmrS drug transport (a proton-driven multidrug efflux pump in MRSA)
trans-cinnamaldehyde and eugenolA. baumanniidownregulated the expression of efflux pump-related gene adeABC
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Huang, L.; Wu, C.; Gao, H.; Xu, C.; Dai, M.; Huang, L.; Hao, H.; Wang, X.; Cheng, G. Bacterial Multidrug Efflux Pumps at the Frontline of Antimicrobial Resistance: An Overview. Antibiotics 2022, 11, 520. https://doi.org/10.3390/antibiotics11040520

AMA Style

Huang L, Wu C, Gao H, Xu C, Dai M, Huang L, Hao H, Wang X, Cheng G. Bacterial Multidrug Efflux Pumps at the Frontline of Antimicrobial Resistance: An Overview. Antibiotics. 2022; 11(4):520. https://doi.org/10.3390/antibiotics11040520

Chicago/Turabian Style

Huang, Lulu, Cuirong Wu, Haijiao Gao, Chao Xu, Menghong Dai, Lingli Huang, Haihong Hao, Xu Wang, and Guyue Cheng. 2022. "Bacterial Multidrug Efflux Pumps at the Frontline of Antimicrobial Resistance: An Overview" Antibiotics 11, no. 4: 520. https://doi.org/10.3390/antibiotics11040520

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