Next Article in Journal
Behavioral and Psychiatric Symptoms in Patients with Severe Traumatic Brain Injury: A Comprehensive Overview
Next Article in Special Issue
Terminalia chebula-Assisted Silver Nanoparticles: Biological Potential, Synthesis, Characterization, and Ecotoxicity
Previous Article in Journal
Sinapic Acid Attenuate Liver Injury by Modulating Antioxidant Activity and Inflammatory Cytokines in Thioacetamide-Induced Liver Cirrhosis in Rats
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Inhibition of Multidrug Efflux Pumps Belonging to the Major Facilitator Superfamily in Bacterial Pathogens

1
Department of Biology, Eastern New Mexico University, Station 33, Portales, NM 88130, USA
2
ICAR-Central Institute of Fisheries Education (CIFE), Mumbai 400061, India
*
Author to whom correspondence should be addressed.
Biomedicines 2023, 11(5), 1448; https://doi.org/10.3390/biomedicines11051448
Submission received: 14 April 2023 / Revised: 7 May 2023 / Accepted: 10 May 2023 / Published: 15 May 2023
(This article belongs to the Special Issue Research of Molecules to Fight Antimicrobial Resistance)

Abstract

:
Bacterial pathogens resistant to multiple structurally distinct antimicrobial agents are causative agents of infectious disease, and they thus constitute a serious concern for public health. Of the various bacterial mechanisms for antimicrobial resistance, active efflux is a well-known system that extrudes clinically relevant antimicrobial agents, rendering specific pathogens recalcitrant to the growth-inhibitory effects of multiple drugs. In particular, multidrug efflux pump members of the major facilitator superfamily constitute central resistance systems in bacterial pathogens. This review article addresses the recent efforts to modulate these antimicrobial efflux transporters from a molecular perspective. Such investigations can potentially restore the clinical efficacy of infectious disease chemotherapy.

Graphical Abstract

1. Bacterial Pathogens

Bacterial pathogens of concern within the context of the worldwide emergence and spread of antimicrobial resistance (AMR) are classified under the group ESKAPEE, comprising Enterococcus faecium, Staphylococcus aureus, Acinetobacter baumannii, Klebsiella pneumoniae, Pseudomonas aeruginosa, Enterobacter species, and Escherichia coli [1]. These bacterial species represent some of the extremely drug-resistant strains that threaten the relevance of existing antimicrobial therapy due to their current ability to resist all available antimicrobials [2]. ESKAPEE pathogens are the major cause of healthcare-associated infections (nosocomial), and it is estimated that more than 80% of global deaths are due to ESKAPEE pathogens [3].
The Gram-positive enteric bacterium Enterococcus faecium is the causative agent of neonatal meningitis, endocarditis, and bacteremia and is a common agent of nosocomial infections, second only to staphylococci [4]. The bacterium is intrinsically resistant to many antibiotics, such as the β-lactams and the cephalosporins, due to the overproduction of low-affinity penicillin-binding proteins (PBPs), aminoglycosides, and trimethoprim-sulfamethoxazole. In contrast, the bacterium has acquired resistance to several others, such as vancomycin, linezolid, tigecycline, and daptomycin, through point mutations and the acquisition of resistance plasmids [5,6]. Hospital environments, including medical devices, offer ideal niches for colonization by vancomycin-resistant enterococci (VRE), making these common sources of potentially fatal nosocomial infections in susceptible populations [6,7]. Enterococci are among the high-priority pathogens owing to the rapid increase in infections attributed to drug-resistant isolates [8].
Staphylococcus aureus is the most common human and animal pathogen responsible for skin, soft tissue, wound infections, life-threatening pneumonia, endocarditis, and infections associated with indwelling devices in hospital settings [9,10]. This bacterium is of great importance due to its resistance against numerous antimicrobial agents, and S. aureus has been involved in more than 100,000 deaths attributable to methicillin-resistant S. aureus (MRSA AMR in 2019 [3]. Although MRSA microorganisms were initially identified as healthcare-associated (HA-MRSA) strains, subsequently, distinct lineages of MRSA emerged from livestock and the community, termed livestock-associated (LA-MRSA) and community-associated (CA-MRSA) [11]. Due to the high genetic and phenotypic variation and their ability to adapt to various environmental conditions, staphylococci have become resistant to the most currently used antimicrobials. MRSA strains are typically resistant to β-lactams and cephalosporins, making it necessary to use alternate antibiotics such as glycopeptides (vancomycin, teicoplanin), tigecycline, daptomycin, and linezolid [12,13]. With the emergence of clonal types of MRSA resistant to each of these antimicrobials, the treatment options against this pathogen are becoming critically constricted to a minority of individual antibiotics or combination therapies involving more than one antibiotic [14].
The Gram-negative pathogens Klebsiella pneumoniae, Acinetobacter baumannii, Pseudomonas aeruginosa, and Enterobacter spp. (Enterobacter cloacae complex, ECC) are high-priority pathogens owing to their plethora of mechanisms to resist multiple antibiotics, colonize hospital environments and devices, and cause serious nosocomial infections [1]. K. pneumoniae ranked third in global deaths associated with AMR in 2019 [3]. Resistance to all β-lactam antibiotics, including cephalosporins, carbapenems, aminoglycosides, and fluoroquinolones, has left clinicians with few options to treat infections by these pathogens [4,15,16]. The carbapenem-resistant Enterobacterales (CRE) group, producing KPC, VIM, IMP, and NDM carbapenemases, is often associated with severe infections with a fatality rate as high as 40% [17]. Similarly, P. aeruginosa is a serious pathogen causing various infections, from wound infections to fatal bacteremia, pneumonia, and lung infections in patients with cystic fibrosis and in immunocompromised individuals. This bacterium possesses diverse drug-resistance mechanisms, both intrinsic and acquired, and produces antibiotic degradative enzymes such as AmpC lactamases, extended spectrum-β-lactamases (ESBLs), and carbapenemases, and has numerous efflux pumps, all of which synergistically make this bacterium extremely drug-resistant [18,19]. Its ability to form strong biofilms with a robust quorum-sensing system has enabled this bacterium to resist disinfectants and antimicrobials and successfully persist in the host environment to establish chronic infections [2,18,20]. Certain lineages of P. aeruginosa, such as ST235 and ST175, have emerged as main agents of serious nosocomial infections [21,22].

2. Antimicrobial Resistance Mechanisms

Antimicrobial agents are purportedly meant to eliminate microorganisms, particularly those deleterious to the health of plants, animals, and humans. Besides these, antimicrobials are also useful in preserving foods from spoilage. Antimicrobial agents could be of chemical or natural origin. The antibiotic penicillin is recognized as the first microbe-derived antimicrobial compound employed clinically to treat infectious diseases. The three-decade period following the discovery of penicillin is termed the “golden period of antimicrobial therapy”, as numerous antibiotics were discovered and inducted into clinical use [23]. However, along with the discovery of antimicrobials, bacteria that survived and grew in antibiotics were also discovered, suggesting that such bacteria can be present in the environment or possibly evolve during antibiotic exposure. Penicillin-resistant staphylococci were isolated from clinical samples in the early 1940s, and with the widespread use of antibiotics in the following two decades, the emergence and spread of antibiotic-resistant strains increased dramatically in clinical settings [24]. The development of resistance followed the discovery of new antimicrobials, and this phenomenon is often aided by the development of cross-resistance to closely related antibiotics or antibiotics belonging to the same group.
The bacterial mechanisms of antibiotic resistance are diverse and are broadly classified into (i) enzymatic inactivation of antibiotics, (ii) modification of drug target, (iii) deceased drug permeability, and (iv) active efflux of antibiotics [25]. The enzymatic inactivation is achieved by hydrolysis of antibiotics by bacterial enzymes such as the β-lactamases that degrade β-lactam antibiotics such as the penicillins, cephalosporins, and carbapenems, macrolide esterases, and the fosfomycin-inactivating enzymes. The second mechanism of enzymatic inactivation of antibiotics involves the structural modification of drugs by group transfer activities from enzymes such as acetyltransferases, phosphotransferases, glycosyltransferases, and nucleotidyltransferases [26,27]. The group transfer mechanism results in the irreversible modification of the antibiotic structure, which can no longer bind to its target. The antibiotics susceptible to this resistance mechanism include chloramphenicol, aminoglycosides, macrolides, and fosfomycin [23].
Alternatively, the antibiotic targets may be modified to circumvent binding by the antibiotics. This resistance may be achieved by mutations in the genes encoding the proteins which act as antibiotic targets, such as the penicillin-binding proteins (PBPs), the modification of which results in the inability of β-lactams to bind to PBPs resulting in resistance development [28]. Mutations in QRDRs (quinolone resistance determinant regions) result in the modification of ribosomal targets leading to the development of resistance to quinolone–fluoroquinolone antibiotics [29].
The outer membrane porins often regulate the entry of antimicrobials into the bacterial cell, the structural modification of which can result in reduced permeability of drugs through the cellular membrane. Such porin modifications result from mutations in porin-encoding genes under antibiotic pressure, and bacteria expressing defective porin proteins resist the entry of antibiotics such as the aminoglycosides, chloramphenicol, tetracyclines, β-lactams, and fluoroquinolones [30].
The efflux-mediated antimicrobial resistance mechanism involves the transmembrane proteins, which transport structurally diverse substrates across the membrane and primarily function to extrude toxic metabolites, Kreb’s cycle intermediates such as salts, sugars, vitamins, fatty acids, and amino acids, among others [9,31]. Based on the source of energy that drives this efflux process, the transporter proteins are broadly grouped into primary active transporters, which use ATP, and secondary active transporters, which utilize the concentration gradient established across the membrane by the primary active transport and respiration [32,33].

3. Antimicrobial Transporter Superfamilies

Many of the transport systems used by microorganisms can fall into one of a variety of protein superfamilies, the vast majority of which have been incorporated into a massive transporter classification database (TCDB) [34]. These and other biological transport systems have been systematically and taxonomically organized and are readily accessible in the TCDB, a continually updated resource [35]. The transporter superfamilies are established according to similarities in evolutionary origin, sequence, protein structure, and energization modes [36,37].
Solute transporters that involve antimicrobial agents and bacterial pathogens can be grouped into a handful of superfamilies. The drug/metabolite transporter (DMT) superfamily consists of many protein families, such as the small multidrug resistance (SMR) family [38]. The multidrug/oligosaccharidyl-lipid/polysaccharide (MOP) superfamily harbors [39] several distantly related protein families, one of which is the multidrug and toxic compound extrusion transporters (MATE) family consisting of numerous cation-based multidrug antiporters [40]. The resistance–nodulation–cell division (RND) superfamily contains many solute transporters, many of which are found in bacterial pathogens conferring resistance to multiple antimicrobial agents [41]. The proteobacterial antimicrobial compound efflux (PACE) family encompasses members that are multidrug efflux pump systems in bacteria [42]. One of the largest groups is the major facilitator superfamily (MFS), consisting of secondary active transporters and passive facilitators [43]. This review focuses primarily on the multidrug efflux pumps of the MFS as they are extensively studied, ubiquitous across all living taxa, and involved in conferring drug resistance in clinical pathogens, making them good targets for modulation [9,44]. The extensively studied bacterial antimicrobial transporter superfamilies are illustrated in Figure 1.

3.1. Major Facilitator Superfamily (MFS)

The MFS solute transporters consist of passive and secondary active transporters such as uniporters, symporters, and antiporters [45]. The protein members of the MFS typically have between 300 and 600 amino acids with 12 or 14 transmembrane domains composed of α-helices [46]. For instance, the LmrS multidrug efflux pump from S. aureus is predicted to harbor 14 membrane-spanning helices, with the N- and C-termini facing the cytoplasmic side of the cell membrane, Figure 2.
The MFS transporters share similarities in amino acid sequences [47], including several functional, highly conserved sequence motifs [48] and protein structures. The MFS transporters have diverse, structurally distinct membrane transport substrates, such as sugars, amino acids, metabolic intermediates, nucleic acids, ions, and antimicrobial agents [47].

MFS Antimicrobial Efflux Pump Structure and Function

The first transporter of the MFS to be characterized physiologically and at the molecular level was the E. coli tetracycline efflux pump, TetA, discovered in the laboratory of Levy [49,50]. Shortly afterward, related tetracycline transporters were found in Gram-negative and -positive microorganisms, collectively called TetA, with sub-classes including A-D and others [51]. Interestingly, it was discovered that these so-called single-substrate TetA transporters were homologous to bacterial multidrug efflux pumps, such as NorA [52], MdfA [53], and QacA [54], which are among the now most intensively studied antimicrobial transporters of the MFS.
High-resolution protein structures have been elucidated for several multidrug efflux pumps of the MFS. Multidrug efflux pumps of the MFS that have been crystalized and structures determined include EmrD [55], YajR [56,57], MdfA [53,58], and SotB [59], all from E. coli, and more recently, NorC from S. aureus [60] and LmrP from Lactococcus lactis [61]. These MFS drug transport proteins generally have two global domains, the N-terminus and the C-terminus domains, each harboring six-helical bundles [62] related by a so-called twofold pseudosymmetrical axis that runs perpendicular to the membrane bilayer [63]. This structural motif, called the MFS fold, appears in many multidrug efflux pump proteins and symporters of the superfamily [64,65]. Furthermore, these antimicrobial transporters alternately expose their substrate binding sites on either side of the membrane to mediate drug and ion translocation across the membrane during the transport cycle, a catalytic mechanism known as the alternating-access model [66,67].
Thus far, the MFS antimicrobial and multidrug efflux pumps share a highly conserved signature sequence called motif C [68] and the antiporter motif [69,70] in the fifth transmembrane helix. The antiporter motif C has the consensus amino acid sequence of “G (X)8 G (X)3 G P (X)2 G G” [71] consisting largely of a highly hydrophobic region with a proline manifested as a so-called GP dipeptide, where X is any amino acid [69,70]. The residues of the antiporter motif appear to play major mechanistic roles during drug and multidrug transport [48]. The first evidence to demonstrate the functional importance of the antiporter motif was reported on TetA(C), in which the highly conserved residue Gly-147, part of the GP dipeptide, was necessary for tetracycline resistance [70]. Since then, other transporters have been reported to have a variety of functional roles attributable to residues of motif C, such as mediating the direction of substrate transport [72], change in protein conformation during transport [73], forming an ion or substrate leak barrier [74,75,76], protein stabilization [77], drug binding [78], forming an accessible central cavity for binding substrates [79], and constituting the interface boundary between the two helical bundles [80] of the MFS antimicrobial efflux pumps. More recently, it was reported that motif C’s residues form a flexible hinge structure that undergoes conformational changes during transport and a regulator switch mechanism for that hinge’s conformation change [81]. As new studies are needed to fully understand the mechanisms played by the antiporter motif C structure, it has become an important target for resistance modulation [44], especially in multidrug-resistant bacterial pathogens [82,83].
Another widely studied conserved sequence motif is referred to as motif A with the consensus sequence “G (X)3 D R/K X G R R”, which is present in the intracellular loop between helices two and three of the vast majority of transporters of the MFS [36,47,84]. These conserved residues are found in symporters of the MFS, reviewed elsewhere [48]. The laboratories of Yamaguchi and Levy were the first to independently report evidence for a functional role of motif A residues in an antimicrobial efflux pump of the MFS [85,86]. They showed that the dipeptide Ser-Asp in the loop conferred drug transport and formed a gate in the tetracycline pumps TetA(B) and TetA(C) [51,87]. Analyses of related MFS multidrug efflux pumps from a structure–function approach in many laboratories have provided physiological evidence for residues of motif A. For example, the structures formed by the motif A consensus sequence were shown to confer the substrate pathway through the transporters [88,89], structural stability mediated by salt bridges [56], conformation change regulation [88,89], an interface boundary between the N- and C-terminal bundles [56], an electrochemical ion-gradient potential sensor [90,91], and a switch that influences binding site orientation [91]. These studies further implicate a critical solute transport role in virtually all protein members of the MFS. Continued analyses of the residues are anticipated to shed new insight into the molecular mechanism of drug transport across the membrane and the generation of mutants with altered specificities for transporting desirable substrates. Thus, the residues of motif A and the structure and functions conferred by them continue to be a focus in mechanistic studies of solute transport in proteins of the MFS [23,92,93].

4. Efflux Pump Inhibition

Multidrug efflux pumps of the MFS are known to confound treatment efforts toward clinical bacterial infections [23]. Thus, understanding the modulation of antimicrobial agent efflux by MFS drug pumps in potentially pathogenic microorganisms is of considerable interest to restore the clinical efficacy of treatment of infections caused by multidrug-resistant pathogens [44,94]. Naturally occurring plant-derived agents as resistance modulators in such cases represent a promising avenue [31]. Our laboratory discovered that a plant-derived extract from Allium sativum and its bioactive agent, allyl sulfide, reduced resistance to substrates and drug transport activity of the MFS multidrug efflux pump EmrD-3 from Vibrio cholerae [95]. In the same study, we demonstrated that A. sativum extract showed a synergistic reduction in the antimicrobial susceptibility in host cells harboring EmrD-3, pointing to this resistance mechanism as a suitable target for the modulation of resistance in severe cholera infections [96]. In another study, we found extracts from the common food spice Cuminum cyminum directly inhibited the transport activity of the LmrS multidrug efflux pump from S. aureus. We showed that the compound cumin aldehyde restored the susceptibility levels of antimicrobial agents that are known substrates of the LmrS multidrug efflux pump [97]. Therefore, the MFS transporter LmrS is another desirable target for developing putative modulatory agents against infection from MRSA [1,9,94].
The NorA multidrug efflux pump system of S. aureus has been a good target for transport inhibition studies reviewed elsewhere [98]. In S. aureus, plant-derived chalcones showed synergistic activity with antimicrobial agents ciprofloxacin, norfloxacin, and ethidium bromide involving the MFS proteins NorA and MepA [99]. Molecular docking analysis of these chalcones showed close interactions with NorA residues Ser-337, Met-338, Gly-339, and Asn-340, implicating these residues as good molecular targets for transport studies and resistance modulation [99]. The affected residues directly involve ciprofloxacin and norfloxacin binding and transport across the membrane through NorA. Recently, it was reported that the saponin secondary metabolite hecogenin acetate, while demonstrating antibacterial effects on isolates of S. aureus, nevertheless failed to show an inhibitory effect on drug efflux activities in NorA and MepA [100].
Interestingly, evaluation of the 1,8-naphthyridine sulfonamides showed that they could inhibit both β-lactamases and the QacC and QacA/B efflux pumps in S. aureus [101]. Thus, these agents and their derivatives show promise in addressing infection by pathogens by targeting more than one resistance mechanism, as in the case of resistances conferred by enzymatic inactivation and active drug efflux. More recently, berberine, a naturally occurring plant compound and known efflux pump inhibitor of Mdr1p, an MFS multidrug efflux pump from Candida albicans [102], reduced the resistance to multiple antimicrobial agents conferred by MdfA of E. coli [103]. Using a combination of molecular simulation dynamics and physiological analysis of transport by MdfA across the membrane, berberine was reported to affect the formation of salt bridges and alter the hydrophobic interactions of MdfA with water in the membrane during the transport cycle to inhibit transport [103]. Another study showed that resistance of C. albicans to the antifungal agent fluconazole conferred by the MFS pump Mdr1 could be blocked by plant extracts from Acalypha communis and Solanum atriplicifolium. Extracts from Argentinian native plants reverse fluconazole resistance in Candida species [104]. Although the biochemical nature of these plant extracts was not identified in the study, the inhibitory effects on Mdr1 are of interest, as these extracts are considered non-toxic and potent in their activities. Table 1 shows the potential efflux pump inhibitors that hold tremendous promise in restoring the antimicrobial susceptibility of important bacterial pathogens.

5. Antimicrobial Resistance and Biofilms

Dynamics between Biofilm Formation and Antibiotic Resistance

Bacteria form biofilms in most scenarios where the dispersal and planktonic stages are considered intermediate or transitional. Biofilms are reported from the sea, rivers, food processing surfaces, medical implants, and the International Space Station [141]. Biofilms are medically important since almost 65–80% of human chronic infections are attributed to pathogenic biofilms [142]. The biofilm has become a menace in the food processing industry due to its persistence on food contact surfaces [143]. Biofilms are a survival strategy for bacteria to escape environmental stress, including predation by the bacteriophage [144]. Environmental stress triggers the transition of free-swimming planktonic forms to sessile forms that attach to a biotic or abiotic surface [145].
There are various aspects in the relationship between antibiotic resistance and biofilm formation. One aspect is that the biofilm acts as the antibiotic resistance gene pool, thus facilitating the emergence of antibiotic resistance bacteria. Another is that the biofilms enhance the survivability of drug-resistant bacteria in a harsh environment, thus helping them to sustain in various environments for a long time. A study conducted in the Yangtze Estuary found that antibiotic-resistant genes (ARG) were high in biofilms, followed by sediment and water. The biofilm is an evident sink for ARG [146]. Hence, biofilm acts as a reservoir of antibiotic-resistant genes, thus facilitating antibiotic resistance in the bacterial biofilm community. Ratajczak et al. [147] found a positive correlation between MDR and biofilm formation in P. aeruginosa. Another aspect is that biofilms are a survival strategy for bacteria with low levels of antimicrobial resistance.
A statistically negative correlation was observed between biofilm formation and MDR in Acinetobacter baumanii. The MDR and XDR isolates formed weak biofilms compared to non-MDR isolates producing robust biofilms [148]. Exposure to sub-inhibitory and sub-lethal concentrations of different antimicrobials triggers biofilm formation in different bacteria. A study conducted with P. aeruginosa found that exposure to aminoglycosides, particularly tobramycin, had the most effect on biofilm formation. Neither polymyxin B, a peptide antibiotic cationic like the aminoglycosides, nor carbenicillin or chloramphenicol had any effect on biofilm formation. Thus, P. aeruginosa forms biofilm as a specific response to the aminoglycoside antibiotics [149]. A recent study shed a different light on understanding antimicrobial resistance in biofilms. The zone of inhibition formed during tobramycin disc diffusion resulted from the transition of P. aeruginosa from planktonic to biofilm growth mode [150]. The bacterial biofilm confers protection from antibiotics through various means. Antimicrobial resistance or tolerance can be grouped into extracellular, cellular, and nuclear components, i.e., the biofilm matrix, the physiological state of bacteria, and genetic determinants. Resistant microorganisms can grow in the presence of a bactericidal or bacteriostatic antimicrobial agent at a concentration normally inhibitory to growth measured as minimum inhibitory concentration (MIC).
In contrast, tolerance to an antimicrobial agent is the ability of a microorganism to survive but neither grow nor die in the presence of a bactericidal antimicrobial agent measured as minimum bactericidal concentration (MBC) [151]. The reduced susceptibility of biofilm to antibiotics could be due to complex interactions, and hence a clear demarcation as antibiotic resistance or tolerance might not be feasible in all the settings. Hence many authors use the term recalcitrance to denote the reduced susceptibility of the biofilm community to the antibiotics tested [152,153,154]. The mechanisms involved in antibiotic recalcitrance in a biofilm are depicted in Figure 3.

6. Role of the Biofilm Matrix in Antibiotic Recalcitrance

The biofilm matrix comprises components such as the extracellular polymeric substance (EPS), extracellular DNA (eDNA), proteins, and lipids inside which the bacterial cells are embedded. In most biofilms, the bacterial cells account for less than 10% of the biofilm’s dry mass [155]. The biofilm composition differs between species and even within the species. The MRSA biofilms are more proteinaceous compared to the polysaccharide-rich MSSA biofilms [156].

7. MFS Transporters and Biofilms

Efflux pumps play important roles in the formation of biofilms, as well as in the antibiotic resistance of biofilm bacteria. The overexpression of tetracycline resistance efflux pump TetA(C) in Escherichia coli biofilm contributes to forming mature biofilms, stress tolerance, and antimicrobial resistance [157]. The inactivation of efflux pumps abolishes biofilm formation, suggesting that efflux pumps as essential for biofilm formation and its persistence [158]. Pinostrobin, a plant-derived flavonoid compound, has anti-efflux and anti-biofilm activities. This compound supposedly interacts more efficiently with MFS efflux pumps of Gram-positive bacteria, reducing the MIC of ciprofloxacin by 128 times in MRSA [159]. In Salmonella enterica serovar Typhimurium, efflux pump knockout mutants lacking EmrAB or MdfA were found to be deficient in biofilm formation due to the mutants’ inability to produce curli, an essential component of biofilm matrix [160]. Boeravinone B, a known NorA multidrug efflux pump inhibitor, also inhibited biofilm formation by S. aureus [134].
Since biofilm formation is strongly associated with the quorum-sensing mechanism of bacteria, efflux pumps promote biofilm formation by reducing the impact of antibacterials and extrusion of quorum-sensing signaling molecules [161,162]. Efflux pump genes are overexpressed in biofilms, and this corresponds with the overexpression of quorum-sensing genes, suggesting a strong relationship between these two processes [163,164,165]. Biologically derived polyamines such as cadaverine, putrescine, spermidine, and spermine are important in bacteria for oxidative stress tolerance, biofilm formation, and persistence. These molecules are also substrates for efflux pumps, such as the AmvA protein of Acinetobacter baumannii, suggesting that the biofilm formation and efflux pump activities can be interrelated [166].
The efflux-pump-mediated biofilm formation is a complex process involving an interplay between multiple pathways, and the net effect of these interactions might vary in different bacterial species. For example, the inactivation of LmrB in Streptococcus mutans resulted in increased EPS secretion and biofilm formation while upregulating other efflux genes’ expression [167]. In many Gram-negative and -positive bacteria, efflux pumps contribute positively to biofilm formation [168]. Efflux pump inhibitors (EPI), in many instances, not only enhance the susceptibility of bacteria to antimicrobials but also reduce or inhibit biofilm formation [169]. Efflux pumps purportedly contribute to biofilm formation at various stages, such as (i) the initiation of biofilm formation in which quorum-sensing molecules play important roles, and the overexpressed efflux pumps might participate in extruding quorum-sensing molecules to bring about a desired effect; (ii) biofilm maturation, during which toxic metabolites need to be expelled from the cellular environment; and (iii) biofilm persistence, in which the biofilm bacteria are protected from an antimicrobial and noxious substance in the surrounding environment [168,170].
The quorum-sensing system also regulates the expression of virulence genes, suggesting that antibiotic resistance, biofilm formation, virulence, and persistence of pathogenic bacteria stay interlinked and provide opportunities to identify the means of interfering with these processes with the ultimate goal of pathogen control [171]. Apart from clinical implications, efflux pump inhibition and the consequent quorum-sensing inhibition could positively impact the shelf life of highly perishable products such as fish. A recent report suggests that black pepper essential oil (BPEO) and its bioactive compounds, limonene (LIM) and β-caryophyllene (CAR), could inhibit efflux pumps and quorum sensing in the fish spoilage bacterium Pseudomonas psychrophila and reduce its spoilage potential [172]. Compounds such as thioridazine and chlorpromazine significantly reduced the gene expression of efflux pumps such as norB, norC, abcA, and mepA and impaired their ability to form biofilms [173]. Copper nanoparticles (CuNPs) have also been shown to inhibit efflux pumps and biofilm formation in S. aureus and P. aeruginosa [174].
Similarly, toluidine blue O (TBO)-mediated photodynamic therapy (PDT) resulted in decreased expression of the norA, norB, sepA, mepA, and mdeA efflux pump genes and impaired biofilm formation by S. aureus strains [175]. Menadione (vitamin K3) has been recently shown to have EPI activity on norA involving two pathways: direct interaction with the NorA protein and indirectly affecting the expression of the norA gene [135]. Nilotinib, a tyrosine kinase inhibitor, significantly reduced S. aureus biofilm formation when combined with ciprofloxacin, suggesting that this compound interacted with the NorA efflux pump leading to diminished activity [176]. The antifungal ketoconazole is an inhibitor of the NorA efflux pump and biofilm formation in S. aureus [177]. MFS efflux pumps are important for biofilm formation in E. coli as evidenced by deficient biofilm formation by the mutant E. coli K12 strain lacking emrD, emrE, emrK, acrD, acrE, and mdtE efflux pump genes [178]. In Shigella flexneri, the efflux pump EmrKY contributes to intracellular survival in macrophages, and its loss results in reduced biofilm formation and increased susceptibility to DNA-damaging substances [179].

8. Future Directions

One new field of study involves the modulation of bacterial antimicrobial efflux pump alteration in expression by nanoparticles [180,181]. We anticipate that this area shows further promise toward generating new treatment strategies against potentially untreatable infections caused by multidrug-resistant pathogens [182]. Surprisingly, although many studies show inhibition of multidrug resistance in transporters of the MFS [9,23,92,93,183], few, if any, of these modulatory agents have reached clinical trials, suggesting that a focus on this area of infectious disease investigation is lacking. The reasons for this apparent disparity are unclear.
One promising avenue can be found in the continued analysis of the MFS antimicrobial transporters’ conserved amino acid signature sequences, especially those transporters expressed in pathogenic microorganisms. Conserved amino acids of the MFS solute transporters represent functionally important aspects that drive multidrug efflux. Studying these drug efflux pumps’ structure–functional natures can reveal critical physiological systems conferring antimicrobial resistance. These molecular mechanisms of antimicrobial transport across the membrane are good targets for developing efflux pump inhibitors [184].

Author Contributions

Conceptualization, S.K. and M.F.V.; writing—original draft preparation, S.K., M.L., J.S., D.B. and M.F.V.; writing—review and editing, S.K., M.L. and M.F.V. All authors have read and agreed to the published version of the manuscript.

Funding

The investigations reviewed in this publication and reported from our research laboratories were supported in part by internal research grants (IRG) from ENMU and grants from the National Institute of General Medical Sciences (P20GM103451), and the National Institutes of Health, the US Department of Education, HSI-STEM (P031C110114).

Acknowledgments

The authors thank Kenwyn Cradock (ENMU) for helpful commentaries.

Conflicts of Interest

The authors of this publication declare no conflict of interest.

References

  1. Stephen, J.; Salam, F.; Lekshmi, M.; Kumar, S.H.; Varela, M.F. The Major Facilitator Superfamily and Antimicrobial Resistance Efflux Pumps of the ESKAPEE Pathogen Staphylococcus aureus. Antibiotics 2023, 12, 343. [Google Scholar] [CrossRef] [PubMed]
  2. Panda, S.K.; Buroni, S.; Swain, S.S.; Bonacorsi, A.; da Fonseca Amorim, E.A.; Kulshrestha, M.; da Silva, L.C.N.; Tiwari, V. Recent Advances to Combat ESKAPE Pathogens with Special Reference to Essential Oils. Front. Microbiol. 2022, 13, 1029098. [Google Scholar] [CrossRef] [PubMed]
  3. Murray, C.J.L.; Ikuta, K.S.; Sharara, F.; Swetschinski, L.; Aguilar, G.R.; Gray, A.; Han, C.; Bisignano, C.; Rao, P.; Wool, E.; et al. Global Burden of Bacterial Antimicrobial Resistance in 2019: A Systematic Analysis. Lancet 2022, 399, 629–655. [Google Scholar] [CrossRef]
  4. Mancuso, G.; Midiri, A.; Gerace, E.; Biondo, C. Bacterial Antibiotic Resistance: The Most Critical Pathogens. Pathogens 2021, 10, 1310. [Google Scholar] [CrossRef] [PubMed]
  5. Hollenbeck, B.L.; Rice, L.B. Intrinsic and Acquired Resistance Mechanisms in Enterococcus. Virulence 2012, 3, 421. [Google Scholar] [CrossRef] [PubMed]
  6. Miller, W.R.; Munita, J.M.; Arias, C.A. Mechanisms of Antibiotic Resistance in Enterococci. Expert. Rev. Anti Infect. Ther. 2014, 12, 1221–1236. [Google Scholar] [CrossRef]
  7. Ayobami, O.; Willrich, N.; Reuss, A.; Eckmanns, T.; Markwart, R. The Ongoing Challenge of Vancomycin-Resistant Enterococcus faecium and Enterococcus faecalis in Europe: An Epidemiological Analysis of Bloodstream Infections. Emerg. Microbes Infect. 2020, 9, 1180–1193. [Google Scholar] [CrossRef]
  8. Tacconelli, E.; Carrara, E.; Savoldi, A.; Harbarth, S.; Mendelson, M.; Monnet, D.L.; Pulcini, C.; Kahlmeter, G.; Kluytmans, J.; Carmeli, Y. Discovery, Research, and Development of New Antibiotics: The Who Priority List of Antibiotic-Resistant Bacteria and Tuberculosis. Lancet Infect. Dis. 2018, 18, 318–327. [Google Scholar] [CrossRef]
  9. Andersen, J.L.; He, G.-X.; Kakarla, P.; Kc, R.; Kumar, S.; Lakra, W.S.; Mukherjee, M.M.; Ranaweera, I.; Shrestha, U.; Tran, T.; et al. Multidrug Efflux Pumps from Enterobacteriaceae, Vibrio cholerae and Staphylococcus aureus Bacterial Food Pathogens. Int. J. Environ. Res. Public Health 2015, 12, 1487–1547. [Google Scholar] [CrossRef]
  10. Lekshmi, M.; Stephen, J.; Ojha, M.; Kumar, S.; Varela, M. Staphylococcus aureus Antimicrobial Efflux Pumps and Their Inhibitors: Recent Developments. AIMS Med. Sci. 2022, 9, 367–393. [Google Scholar] [CrossRef]
  11. Hassoun, A.; Linden, P.K.; Friedman, B. Incidence, Prevalence, and Management of MRSA Bacteremia across Patient Populations—A Review of Recent Developments in MRSA Management and Treatment. Crit. Care 2017, 21, 211. [Google Scholar] [CrossRef] [PubMed]
  12. Vestergaard, M.; Frees, D.; Ingmer, H. Antibiotic Resistance and the MRSA Problem. Microbiol. Spectr. 2019, 7, 2. [Google Scholar] [CrossRef] [PubMed]
  13. Nandhini, P.; Kumar, P.; Mickymaray, S.; Alothaim, A.S.; Somasundaram, J.; Rajan, M. Recent Developments in Methicillin-Resistant Staphylococcus aureus (MRSA) Treatment: A Review. Antibiotics 2022, 11, 606. [Google Scholar] [CrossRef] [PubMed]
  14. Shoaib, M.; Aqib, A.I.; Muzammil, I.; Majeed, N.; Bhutta, Z.A.; Kulyar, M.F.-A.; Fatima, M.; Zaheer, C.-N.F.; Muneer, A.; Murtaza, M.; et al. MRSA Compendium of Epidemiology, Transmission, Pathophysiology, Treatment, and Prevention within One Health Framework. Front. Microbiol. 2023, 13, 1067284. [Google Scholar] [CrossRef]
  15. Pendleton, J.N.; Gorman, S.P.; Gilmore, B.F. Clinical Relevance of the ESKAPE Pathogens. Expert. Rev. Anti. Infect. Ther. 2013, 11, 297–308. [Google Scholar] [CrossRef]
  16. Tommasi, R.; Brown, D.G.; Walkup, G.K.; Manchester, J.I.; Miller, A.A. ESKAPEing the Labyrinth of Antibacterial Discovery. Nat. Rev. Drug Discov. 2015, 14, 529–542. [Google Scholar] [CrossRef]
  17. Tzouvelekis, L.S.; Markogiannakis, A.; Psichogiou, M.; Tassios, P.T.; Daikos, G.L. Carbapenemases in Klebsiella pneumoniae and Other Enterobacteriaceae: An Evolving Crisis of Global Dimensions. Clin. Microbiol. Rev. 2012, 25, 682–707. [Google Scholar] [CrossRef]
  18. Mulani, M.S.; Kamble, E.E.; Kumkar, S.N.; Tawre, M.S.; Pardesi, K.R. Emerging Strategies to Combat ESKAPE Pathogens in the Era of Antimicrobial Resistance: A Review. Front. Microbiol. 2019, 10, 539. [Google Scholar] [CrossRef]
  19. 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]
  20. Reynolds, D.; Kollef, M. The Epidemiology and Pathogenesis and Treatment of Pseudomonas aeruginosa Infections: An Update. Drugs 2021, 81, 2117–2131. [Google Scholar] [CrossRef]
  21. Cabot, G.; López-Causapé, C.; Ocampo-Sosa, A.A.; Sommer, L.M.; Domínguez, M.Á.; Zamorano, L.; Juan, C.; Tubau, F.; Rodríguez, C.; Moyà, B.; et al. Deciphering the Resistome of the Widespread Pseudomonas aeruginosa Sequence Type 175 International High-Risk Clone through Whole-Genome Sequencing. Antimicrob. Agents Chemother. 2016, 60, 7415–7423. [Google Scholar] [CrossRef] [PubMed]
  22. Treepong, P.; Kos, V.N.; Guyeux, C.; Blanc, D.S.; Bertrand, X.; Valot, B.; Hocquet, D. Global Emergence of the Widespread Pseudomonas aeruginosa ST235 Clone. Clin. Microbiol. Infect. 2018, 24, 258–266. [Google Scholar] [CrossRef] [PubMed]
  23. Varela, M.F.; Stephen, J.; Lekshmi, M.; Ojha, M.; Wenzel, N.; Sanford, L.M.; Hernandez, A.J.; Parvathi, A.; Kumar, S.H. Bacterial Resistance to Antimicrobial Agents. Antibiotics 2021, 10, 593. [Google Scholar] [CrossRef]
  24. Lobanovska, M.; Pilla, G. Penicillin’s Discovery and Antibiotic Resistance: Lessons for the Future? Yale J. Biol. Med. 2017, 90, 135–145. [Google Scholar]
  25. Blair, J.M.A.; Webber, M.A.; Baylay, A.J.; Ogbolu, D.O.; Piddock, L.J.V. Molecular mechanisms of antibiotic resistance. Nat. Rev. Microbiol. 2015, 13, 42–51. [Google Scholar] [CrossRef] [PubMed]
  26. Lupo, A.; Coyne, S.; Berendonk, T.U. Origin and Evolution of Antibiotic Resistance: The Common Mechanisms of Emergence and Spread in Water Bodies. Front. Microbiol. 2012, 3, 18. [Google Scholar] [CrossRef]
  27. Wright, G.D. Molecular mechanisms of antibiotic resistance. Chem. Commun. 2011, 47, 4055–4061. [Google Scholar] [CrossRef]
  28. Zapun, A.; Contreras-Martel, C.; Vernet, T. Penicillin-Binding Proteins and β-Lactam Resistance. FEMS Microbiol. Rev. 2008, 32, 361–385. [Google Scholar] [CrossRef]
  29. Jacoby, G.A. Mechanisms of Resistance to Quinolones. Clin. Infect. Dis. 2005, 41, S120–S126. [Google Scholar] [CrossRef]
  30. Delcour, A.H. Outer membrane permeability and antibiotic resistance. Biochim. Biophys. Acta 2009, 1794, 808–816. [Google Scholar] [CrossRef]
  31. Rao, M.; Padyana, S.; Dipin, K.; Kumar, S.; Nayak, B.; Varela, M.F. Antimicrobial Compounds of Plant Origin as Efflux Pump Inhibitors: New Avenues for Controlling Multidrug Resistant Pathogens. J. Antimicrob. Agents 2018, 4, 1–6. [Google Scholar] [CrossRef]
  32. Henderson, P.J. Studies of Translocation Catalysis. Biosci. Rep. 1991, 11, 477–538; discussion 534–538. [Google Scholar] [CrossRef] [PubMed]
  33. Saier, M.H. Vectorial Metabolism and the Evolution of Transport Systems. J. Bacteriol. 2000, 182, 5029–5035. [Google Scholar] [CrossRef] [PubMed]
  34. Saier, M.H.; Reddy, V.S.; Tsu, B.V.; Ahmed, M.S.; Li, C.; Moreno-Hagelsieb, G. The Transporter Classification Database (TCDB): Recent Advances. Nucleic Acids Res. 2016, 44, D372–D379. [Google Scholar] [CrossRef] [PubMed]
  35. Saier, M.H.; Reddy, V.S.; Moreno-Hagelsieb, G.; Hendargo, K.J.; Zhang, Y.; Iddamsetty, V.; Lam, K.J.K.; Tian, N.; Russum, S.; Wang, J.; et al. The Transporter Classification Database (TCDB): 2021 Update. Nucleic Acids Res. 2021, 49, D461–D467. [Google Scholar] [CrossRef]
  36. Griffith, J.; Sansom, C. The Transporter Factsbook; Elsevier Science: Amsterdam, The Netherlands, 1997; ISBN 978-0-08-054265-2. [Google Scholar]
  37. Broome-Smith, J.K.; Symposium, S.; Baumberg, S.; Stirling, C.J.; Ward, F.B. Transport of Molecules Across Microbial Membranes; Cambridge University Press: Cambridge, UK, 1999; ISBN 978-0-521-77270-9. [Google Scholar]
  38. Burata, O.E.; Yeh, T.J.; Macdonald, C.B.; Stockbridge, R.B. Still Rocking in the Structural Era: A Molecular Overview of the Small Multidrug Resistance (SMR) Transporter Family. J. Biol. Chem. 2022, 298, 102482. [Google Scholar] [CrossRef]
  39. Hvorup, R.N.; Winnen, B.; Chang, A.B.; Jiang, Y.; Zhou, X.F.; Saier, M.H., Jr. The Multidrug/Oligosaccharidyl-Lipid/Polysaccharide (Mop) Exporter Superfamily. Eur. J. Biochem. 2003, 270, 799–813. [Google Scholar] [CrossRef]
  40. Kuroda, T.; Tsuchiya, T. Multidrug Efflux Transporters in the MATE Family. Biochim. Biophys. Acta 2009, 1794, 763–768. [Google Scholar] [CrossRef]
  41. Nikaido, H. RND Transporters in the Living World. Res. Microbiol. 2018, 169, 363–371. [Google Scholar] [CrossRef]
  42. 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. 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]
  43. Saier, M.H., Jr.; Beatty, J.T.; Goffeau, A.; Harley, K.T.; Heijne, W.H.; Huang, S.C.; Jack, D.L.; Jahn, P.S.; Lew, K.; Liu, J.; et al. The Major Facilitator Superfamily. J. Mol. Microbiol. Biotechnol. 1999, 1, 257–279. [Google Scholar]
  44. Kumar, S.; He, G.; Kakarla, P.; Shrestha, U.; Ranjana, K.C.; Ranaweera, I.; Willmon, T.M.; Barr, S.R.; Hernandez, A.J.; Varela, M.F. Bacterial Multidrug Efflux Pumps of the Major Facilitator Superfamily as Targets for Modulation. Infect. Disord. Drug Targets 2016, 16, 28–43. [Google Scholar] [CrossRef] [PubMed]
  45. Pao, S.S.; Paulsen, I.T.; Saier, M.H. Major Facilitator Superfamily. Microbiol. Mol. Biol. Rev. 1998, 62, 1–34. [Google Scholar] [CrossRef] [PubMed]
  46. Saidijam, M.; Benedetti, G.; Ren, Q.; Xu, Z.; Hoyle, C.J.; Palmer, S.L.; Ward, A.; Bettaney, K.E.; Szakonyi, G.; Meuller, J.; et al. Microbial Drug Efflux Proteins of the Major Facilitator Superfamily. Curr. Drug Targets 2006, 7, 793–811. [Google Scholar] [CrossRef] [PubMed]
  47. Griffith, J.K.; Baker, M.E.; Rouch, D.A.; Page, M.G.; Skurray, R.A.; Paulsen, I.T.; Chater, K.F.; Baldwin, S.A.; Henderson, P.J. Membrane Transport Proteins: Implications of Sequence Comparisons. Curr. Opin. Cell. Biol. 1992, 4, 684–695. [Google Scholar] [CrossRef]
  48. Kumar, S.; Ranjana, K.; Sanford, L.M.; Hernandez, A.J.; Kakarla, P.; Varela, M.F. Structural and Functional Roles of Two Evolutionarily Conserved Amino Acid Sequence Motifs within Solute Transporters of the Major Facilitator Superfamily. Trends Cell. Mol. Biol. 2016, 11, 41–53. [Google Scholar]
  49. Levy, S.B.; McMurry, L. Plasmid-Determined Tetracycline Resistance Involves New Transport Systems for Tetracycline. Nature 1978, 276, 90–92. [Google Scholar] [CrossRef]
  50. McMurry, L.M.; Cullinane, J.C.; Petrucci, R.E., Jr.; Levy, S.B. Active Uptake of Tetracycline by Membrane Vesicles from Susceptible Escherichia coli. Antimicrob. Agents Chemother. 1981, 20, 307–313. [Google Scholar] [CrossRef]
  51. McMurry, L.; Petrucci, R.E., Jr.; 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]
  52. Yoshida, H.; Bogaki, M.; Nakamura, S.; Ubukata, K.; Konno, M. Nucleotide Sequence and Characterization of the Staphylococcus aureus norA Gene, Which Confers Resistance to Quinolones. J. Bacteriol. 1990, 172, 6942–6949. [Google Scholar] [CrossRef]
  53. Edgar, R.; Bibi, E. MdfA, an Escherichia coli Multidrug Resistance Protein with an Extraordinarily Broad Spectrum of Drug Recognition. J. Bacteriol. 1997, 179, 2274–2280. [Google Scholar] [CrossRef] [PubMed]
  54. Tennent, J.M.; Lyon, B.R.; Midgley, M.; Jones, I.G.; Purewal, A.S.; Skurray, R.A. Physical and Biochemical Characterization of the qacA Gene Encoding Antiseptic and Disinfectant Resistance in Staphylococcus aureus. J. Gen. Microbiol. 1989, 135, 1–10. [Google Scholar] [CrossRef] [PubMed]
  55. Yin, Y.; He, X.; Szewczyk, P.; Nguyen, T.; Chang, G. Structure of the Multidrug Transporter EmrD from Escherichia coli. Science 2006, 312, 741–744. [Google Scholar] [CrossRef] [PubMed]
  56. Jiang, D.; Zhao, Y.; Wang, X.; Fan, J.; Heng, J.; Liu, X.; Feng, W.; Kang, X.; Huang, B.; Liu, J.; et al. Structure of the YajR Transporter Suggests a Transport Mechanism Based on the Conserved Motif A. Proc. Natl. Acad. Sci. USA 2013, 110, 14664–14669. [Google Scholar] [CrossRef] [PubMed]
  57. Jiang, D.; Zhao, Y.; Fan, J.; Liu, X.; Wu, Y.; Feng, W.; Zhang, X.C. Atomic Resolution Structure of the E. coli YajR Transporter YAM Domain. Biochem. Biophys. Res. Commun. 2014, 450, 929–935. [Google Scholar] [CrossRef] [PubMed]
  58. Heng, J.; Zhao, Y.; Liu, M.; Liu, Y.; Fan, J.; Wang, X.; Zhang, X.C. Substrate-Bound Structure of the E. coli Multidrug Resistance Transporter MdfA. Cell. Res. 2015, 25, 1060–1073. [Google Scholar] [CrossRef]
  59. Xiao, Q.; Sun, B.; Zhou, Y.; Wang, C.; Guo, L.; He, J.; Deng, D. Visualizing the Nonlinear Changes of a Drug-Proton Antiporter from Inward-Open to Occluded State. Biochem. Biophys. Res. Commun. 2021, 534, 272–278. [Google Scholar] [CrossRef]
  60. Kumar, S.; Mahendran, I.; Athreya, A.; Ranjan, R.; Penmatsa, A. Isolation and Structural Characterization of a Zn(2+)-Bound Single-Domain Antibody against NorC, a Putative Multidrug Efflux Transporter in Bacteria. J. Biol. Chem. 2020, 295, 55–68. [Google Scholar] [CrossRef]
  61. Debruycker, V.; Hutchin, A.; Masureel, M.; Ficici, E.; Martens, C.; Legrand, P.; Stein, R.A.; McHaourab, H.S.; Faraldo-Gómez, J.D.; Remaut, H.; et al. An Embedded Lipid in the Multidrug Transporter LmrP Suggests a Mechanism for Polyspecificity. Nat. Struct. Mol. Biol. 2020, 27, 829–835. [Google Scholar] [CrossRef]
  62. Maloney, P.C. Bacterial Transporters. Curr. Opin. Cell. Biol. 1994, 6, 571–582. [Google Scholar] [CrossRef]
  63. Huang, Y.; Lemieux, M.J.; Song, J.; Auer, M.; Wang, D.-N. Structure and Mechanism of the Glycerol-3-Phosphate Transporter from Escherichia coli. Science 2003, 301, 616–620. [Google Scholar] [CrossRef]
  64. Ranaweera, I.; Shrestha, U.; Ranjana, K.C.; Kakarla, P.; Willmon, T.M.; Hernandez, A.J.; Mukherjee, M.M.; Barr, S.R.; Varela, M.F. Structural Comparison of Bacterial Multidrug Efflux Pumps of the Major Facilitator Superfamily. Trends Cell. Mol. Biol. 2015, 10, 131–140. [Google Scholar] [PubMed]
  65. Fowler, P.W.; Orwick-Rydmark, M.; Radestock, S.; Solcan, N.; Dijkman, P.M.; Lyons, J.A.; Kwok, J.; Caffrey, M.; Watts, A.; Forrest, L.R.; et al. Gating Topology of the Proton-Coupled Oligopeptide Symporters. Structure 2015, 23, 290–301. [Google Scholar] [CrossRef]
  66. Tanford, C. Mechanism of Free Energy Coupling in Active Transport. Annu. Rev. Biochem. 1983, 52, 379–409. [Google Scholar] [CrossRef]
  67. Jencks, W.P. From Chemistry to Biochemistry to Catalysis to Movement. Annu. Rev. Biochem. 1997, 66, 1–18. [Google Scholar] [CrossRef]
  68. Paulsen, I.T.; Skurray, R.A. Topology, Structure and Evolution of Two Families of Proteins Involved in Antibiotic and Antiseptic Resistance in Eukaryotes and Prokaryotes—An Analysis. Gene 1993, 124, 1–11. [Google Scholar] [CrossRef]
  69. Varela, M.F.; Griffith, J.K. Nucleotide and deduced protein sequences of the class D tetracycline resistance determinant: Relationship to other antimicrobial transport proteins. Antimicrob. Agents Chemother. 1993, 37, 1253–1258. [Google Scholar] [CrossRef]
  70. Varela, M.F.; Sansom, C.E.; Griffith, J.K. Mutational Analysis and Molecular Modelling of an Amino acid Sequence Motif Conserved in Antiporters but not Symporters in a Transporter Superfamily. Mol. Membr. Biol. 1995, 12, 313–319. [Google Scholar] [CrossRef]
  71. Rouch, D.A.; Cram, D.S.; DiBerardino, D.; Littlejohn, T.G.; Skurray, R.A. Efflux-Mediated Antiseptic Resistance Gene qacA from Staphylococcus aureus: Common Ancestry with Tetracycline- and Sugar-Transport Proteins. Mol. Microbiol. 1990, 4, 2051–2062. [Google Scholar] [CrossRef]
  72. Maiden, M.C.; Davis, E.O.; Baldwin, S.A.; Moore, D.C.; Henderson, P.J. Mammalian and Bacterial Sugar Transport Proteins Are Homologous. Nature 1987, 325, 641–643. [Google Scholar] [CrossRef]
  73. Ginn, S.L.; Brown, M.H.; Skurray, R.A. The TetA (K) Tetracycline/H+ Antiporter from Staphylococcus aureus: Mutagenesis and Functional Analysis of Motif C. J. Bacteriol. 2000, 182, 1492–1498. [Google Scholar] [CrossRef]
  74. Konishi, S.; Iwaki, S.; Kimura-Someya, T.; Yamaguchi, A. Cysteine-Scanning Mutagenesis around Transmembrane Segment VI of Tn10-Encoded Metal-Tetracycline/H(+) Antiporter. FEBS Lett. 1999, 461, 315–318. [Google Scholar] [CrossRef] [PubMed]
  75. Jin, J.; Krulwich, T.A. Site-Directed Mutagenesis Studies of Selected Motif and Charged Residues and of Cysteines of the Multifunctional Tetracycline Efflux Protein Tet(L). J. Bacteriol. 2002, 184, 1796–1800. [Google Scholar] [CrossRef]
  76. De Jesus, M.; Jin, J.; Guffanti, A.A.; Krulwich, T.A. Importance of the GP Dipeptide of the Antiporter Motif and Other Membrane-Embedded Proline and Glycine Residues in Tetracycline Efflux Protein Tet(L). Biochemistry 2005, 44, 12896–12904. [Google Scholar] [CrossRef]
  77. Saraceni-Richards, C.A.; Levy, S.B. Evidence for Interactions between Helices 5 and 8 and a Role for the Interdomain Loop in Tetracycline Resistance Mediated by Hybrid Tet Proteins. J. Biol. Chem. 2000, 275, 6101–6106. [Google Scholar] [CrossRef] [PubMed]
  78. Hassan, K.A.; Galea, M.; Wu, J.; Mitchell, B.A.; Skurray, R.A.; Brown, M.H. Functional Effects of Intramembranous Proline Substitutions in the Staphylococcal Multidrug Transporter QacA. FEMS Microbiol. Lett. 2006, 263, 76–85. [Google Scholar] [CrossRef] [PubMed]
  79. Pasrija, R.; Banerjee, D.; Prasad, R. Structure and Function Analysis of CaMdr1p, a Major Facilitator Superfamily Antifungal Efflux Transporter Protein of Candida albicans: Identification of Amino Acid Residues Critical for Drug/H+ Transport. Eukaryot. Cell. 2007, 6, 443–453. [Google Scholar] [CrossRef]
  80. Yaffe, D.; Radestock, S.; Shuster, Y.; Forrest, L.R.; Schuldiner, S. Identification of Molecular Hinge Points Mediating Alternating Access in the Vesicular Monoamine Transporter VMAT2. Proc. Natl. Acad. Sci. USA 2013, 110, E1332–E1341. [Google Scholar] [CrossRef]
  81. Luo, J.; Parsons, S.M. Conformational Propensities of Peptides Mimicking Transmembrane Helix 5 and Motif C in Wild-Type and Mutant Vesicular Acetylcholine Transporters. ACS Chem. Neurosci. 2010, 1, 381–390. [Google Scholar] [CrossRef]
  82. Lekshmi, M.; Ammini, P.; Adjei, J.; Sanford, L.M.; Shrestha, U.; Kumar, S.; Varela, M.F. Modulation of Antimicrobial Efflux Pumps of the Major Facilitator Superfamily in Staphylococcus aureus. AIMS Microbiol. 2018, 4, 1–18. [Google Scholar] [CrossRef]
  83. Lekshmi, M.; Parvathi, A.; Kumar, S.; Varela, M.F. Efflux Pump-Mediated Quorum Sensing: New Avenues for Modulation of Antimicrobial Resistance and Bacterial Virulence. In Biotechnological Applications of Quorum Sensing Inhibitors; Kalia, V.C., Ed.; Springer: Singapore, 2018; pp. 127–142. ISBN 978-981-10-9026-4. [Google Scholar]
  84. Henderson, P.J.F. Proton-Linked Sugar Transport Systems in Bacteria. J. Bioenerg. Biomembr. 1990, 22, 525–569. [Google Scholar] [CrossRef]
  85. Tamura, N.; Konishi, S.; Yamaguchi, A. Mechanisms of Drug/H+ Antiport: Complete Cysteine-Scanning Mutagenesis and the Protein Engineering Approach. Curr. Opin. Chem. Biol. 2003, 7, 570–579. [Google Scholar] [CrossRef]
  86. Nelson, M.L.; Levy, S.B. The History of the Tetracyclines. Ann. N. Y. Acad. Sci. 2011, 1241, 17–32. [Google Scholar] [CrossRef]
  87. Yamaguchi, A.; Ono, N.; Akasaka, T.; Noumi, T.; Sawai, T. Metal-Tetracycline/H+ Antiporter of Escherichia coli Encoded by a Transposon, Tn10. The Role of the Conserved Dipeptide, Ser65-Asp66, in Tetracycline Transport. J. Biol. Chem. 1990, 265, 15525–15530. [Google Scholar] [CrossRef]
  88. Yamaguchi, A.; Akasaka, T.; Kimura, T.; Sakai, T.; Adachi, Y.; Sawai, T. Role of the Conserved Quartets of Residues Located in the N- and C-Terminal Halves of the Transposon Tn10-Encoded Metal-Tetracycline/H+ Antiporter of Escherichia coli. Biochemistry 1993, 32, 5698–5704. [Google Scholar] [CrossRef]
  89. Kimura, T.; Shiina, Y.; Sawai, T.; Yamaguchi, A. Cysteine-Scanning Mutagenesis around Transmembrane Segment III of Tn10-Encoded Metal-Tetracycline/H+ Antiporter. J. Biol. Chem. 1998, 273, 5243–5247. [Google Scholar] [CrossRef] [PubMed]
  90. Bolhuis, H.; Poelarends, G.; van Veen, H.W.; Poolman, B.; Driessen, A.J.; Konings, W.N. The Lactococcal lmrP Gene Encodes a Proton Motive Force-Dependent Drug Transporter. J. Biol. Chem. 1995, 270, 26092–26098. [Google Scholar] [CrossRef]
  91. Masureel, M.; Martens, C.; Stein, R.A.; Mishra, S.; Ruysschaert, J.-M.; Mchaourab, H.S.; Govaerts, C. Protonation Drives the Conformational Switch in the Multidrug Transporter LmrP. Nat. Chem. Biol. 2014, 10, 149–155. [Google Scholar] [CrossRef] [PubMed]
  92. Varela, M.F.; Kumar, S. Strategies for Discovery of New Molecular Targets for Anti-Infective Drugs. Curr. Opin. Pharmacol. 2019, 48, 57–68. [Google Scholar] [CrossRef] [PubMed]
  93. Stephen, J.; Lekshmi, M.; Ammini, P.; Kumar, S.H.; Varela, M.F. Membrane Efflux Pumps of Pathogenic Vibrio Species: Role in Antimicrobial Resistance and Virulence. Microorganisms 2022, 10, 382. [Google Scholar] [CrossRef]
  94. Mohanty, H.; Pachpute, S.; Yadav, R.P. Mechanism of Drug Resistance in Bacteria: Efflux Pump Modulation for Designing of New Antibiotic Enhancers. Folia Microbiol. (Praha) 2021, 66, 727–739. [Google Scholar] [CrossRef]
  95. Bruns, M.M.; Kakarla, P.; Floyd, J.T.; Mukherjee, M.M.; Ponce, R.C.; Garcia, J.A.; Ranaweera, I.; Sanford, L.M.; Hernandez, A.J.; Willmon, T.M.; et al. Modulation of the Multidrug Efflux Pump EmrD-3 from Vibrio cholerae by Allium sativum Extract and the Bioactive Agent Allyl Sulfide plus Synergistic Enhancement of Antimicrobial Susceptibility by A. sativum Extract. Arch. Microbiol. 2017, 199, 1103–1112. [Google Scholar] [CrossRef] [PubMed]
  96. Smith, K.P.; Kumar, S.; Varela, M.F. Identification, Cloning, and Functional Characterization of EmrD-3, a Putative Multidrug Efflux Pump of the Major Facilitator Superfamily from Vibrio cholerae O395. Arch. Microbiol. 2009, 191, 903–911. [Google Scholar] [CrossRef]
  97. Floyd, J.L.; Smith, K.P.; Kumar, S.H.; Floyd, J.T.; Varela, M.F. LmrS Is a Multidrug Efflux Pump of the Major Facilitator Superfamily from Staphylococcus aureus. Antimicrob. Agents Chemother. 2010, 54, 5406–5412. [Google Scholar] [CrossRef]
  98. Kumar, G.; Kiran Tudu, A. Tackling Multidrug-Resistant Staphylococcus aureus by Natural Products and Their Analogues Acting as NorA Efflux Pump Inhibitors. Bioorg. Med. Chem. 2023, 80, 117187. [Google Scholar] [CrossRef] [PubMed]
  99. Rocha, J.E.; de Freitas, T.S.; Xavier, J.C.; Pereira, R.L.S.; Pereira Junior, F.N.; Nogueira, C.E.S.; Marinho, M.M.; Bandeira, P.N.; Rodrigues, L.G.; Marinho, E.S.; et al. ADMET Study, Spectroscopic Characterization and Effect of Synthetic Nitro Chalcone in Combination with Norfloxacin, Ciprofloxacin, and Ethidium Bromide against Staphylococcus aureus Efflux Pumps. Fundam. Clin. Pharmacol. 2023, 37, 163–173. [Google Scholar] [CrossRef] [PubMed]
  100. Santos Araujo, N.J.; Pereira da Silva, A.R.; Costa, M.D.S.; Pereira Silva, C.A.; Sampaio de Freitas, T.; Oliveira de Sousa, E.; Barbosa Filho, J.M.; Lobo Soares de Matos, Y.M.; Melo Coutinho, H.D.; Andrade-Pinheiro, J.C. Evaluation of the Antibacterial Activity of Hecogenin Acetate and Its Inhibitory Potential of NorA and MepA Efflux Pumps from Staphylococcus aureus. Microb. Pathog. 2023, 174, 105925. [Google Scholar] [CrossRef]
  101. Oliveira-Tintino, C.D.d.M.; Tintino, S.R.; Justino de Araújo, A.C.; dos Santos Barbosa, C.R.; Ramos Freitas, P.; Araújo Neto, J.B.d.; Begnini, I.M.; Rebelo, R.A.; Silva, L.E.d.; Mireski, S.L.; et al. Efflux Pump (QacA, QacB, and QacC) and β-Lactamase Inhibitors? An Evaluation of 1,8-Naphthyridines against Staphylococcus aureus Strains. Molecules 2023, 28, 1819. [Google Scholar] [CrossRef]
  102. Tong, Y.; Zhang, J.; Sun, N.; Wang, X.M.; Wei, Q.; Zhang, Y.; Huang, R.; Pu, Y.; Dai, H.; Ren, B.; et al. Berberine Reverses Multidrug Resistance in Candida albicans by Hijacking the Drug Efflux Pump Mdr1p. Sci. Bull. Beijing 2021, 66, 1895–1905. [Google Scholar] [CrossRef]
  103. Li, Y.; Ge, X. Role of Berberine as a Potential Efflux Pump Inhibitor against MdfA from Escherichia coli: In Vitro and In Silico Studies. Microbiol. Spectr. 2023, 11, e03324-22. [Google Scholar] [CrossRef]
  104. Gil, F.; Laiolo, J.; Bayona-Pacheco, B.; Cannon, R.D.; Ferreira-Pereira, A.; Carpinella, M.C. Extracts from Argentinian Native Plants Reverse Fluconazole Resistance in Candida Species by Inhibiting the Efflux Transporters Mdr1 and Cdr1. BMC Complement. Med. Ther. 2022, 22, 264. [Google Scholar] [CrossRef]
  105. Khan, I.A.; Mirza, Z.M.; Kumar, A.; Verma, V.; Qazi, G.N. Piperine, a Phytochemical Potentiator of Ciprofloxacin against Staphylococcus aureus. Antimicrob. Agents Chemother. 2006, 50, 810–812. [Google Scholar] [CrossRef]
  106. Nargotra, A.; Sharma, S.; Koul, J.L.; Sangwan, P.L.; Khan, I.A.; Kumar, A.; Taneja, S.C.; Koul, S. Quantitative Structure Activity Relationship (QSAR) of Piperine Analogsfor Bacterial NorA Efflux Pump Inhibitors. Eur. J. Med. Chem. 2009, 44, 4128–4135. [Google Scholar] [CrossRef]
  107. Seukep, A.J.; Kuete, V.; Nahar, L.; Sarker, S.D.; Guo, M. Plant-Derived Secondary Metabolites as the Main Source of Efflux Pump Inhibitors and Methods for Identification. J. Pharm. Anal. 2020, 10, 277–290. [Google Scholar] [CrossRef]
  108. Bame, J.R.; Graf, T.N.; Junio, H.A.; Iii, R.O.B.; Jarmusch, S.A.; El-Elimat, T.; Iii, J.O.F.; Oberlies, N.H.; Cech, R.A.; Cech, N.B. Sarothrin from Alkanna orientalis Is an Antimicrobial Agent and Efflux Pump Inhibitor. Planta Med. 2013, 79, 327–329. [Google Scholar] [CrossRef] [PubMed]
  109. Wang, S.-Y.; Sun, Z.-L.; Liu, T.; Gibbons, S.; Zhang, W.-J.; Qing, M. Flavonoids from Sophora moorcroftiana and Their Synergistic Antibacterial Effects on MRSA. Phytother. Res. PTR 2014, 28, 1071–1076. [Google Scholar] [CrossRef]
  110. Stermitz, F.R.; Scriven, L.N.; Tegos, G.; Lewis, K. Two Flavonols from Artemisa annua which Potentiate the Activity of Berberine and Norfloxacin Against a Resistant Strain of Staphylococcus aureus. Planta Med. 2002, 68, 1140–1141. [Google Scholar] [CrossRef]
  111. Smith, E.C.J.; Williamson, E.M.; Wareham, N.; Kaatz, G.W.; Gibbons, S. Antibacterials and Modulators of Bacterial Resistance from the Immature Cones of Chamaecyparis lawsoniana. Phytochemistry 2007, 68, 210–217. [Google Scholar] [CrossRef] [PubMed]
  112. Smith, E.C.J.; Kaatz, G.W.; Seo, S.M.; Wareham, N.; Williamson, E.M.; Gibbons, S. The Phenolic Diterpene Totarol Inhibits Multidrug Efflux Pump Activity in Staphylococcus aureus. Antimicrob. Agents Chemother. 2007, 51, 4480–4483. [Google Scholar] [CrossRef]
  113. Morel, C.; Stermitz, F.R.; Tegos, G.; Lewis, K. Isoflavones As Potentiators of Antibacterial Activity. J. Agric. Food Chem. 2003, 51, 5677–5679. [Google Scholar] [CrossRef]
  114. Roy, S.K.; Kumari, N.; Pahwa, S.; Agrahari, U.C.; Bhutani, K.K.; Jachak, S.M.; Nandanwar, H. NorA Efflux Pump Inhibitory Activity of Coumarins from Mesua ferrea. Fitoterapia 2013, 90, 140–150. [Google Scholar] [CrossRef]
  115. Holler, J.G.; Christensen, S.B.; Slotved, H.-C.; Rasmussen, H.B.; Gúzman, A.; Olsen, C.-E.; Petersen, B.; Mølgaard, P. Novel Inhibitory Activity of the Staphylococcus aureus NorA Efflux Pump by a Kaempferol Rhamnoside Isolated from Persea lingue Nees. J. Antimicrob. Chemother. 2012, 67, 1138–1144. [Google Scholar] [CrossRef]
  116. Marquez, B.; Neuville, L.; Moreau, N.J.; Genet, J.-P.; dos Santos, A.F.; Caño de Andrade, M.C.; Goulart Sant’Ana, A.E. Multidrug Resistance Reversal Agent from Jatropha elliptica. Phytochemistry 2005, 66, 1804–1811. [Google Scholar] [CrossRef]
  117. Ponnusamy, K.; Ramasamy, M.; Savarimuthu, I.; Paulraj, M.G. Indirubin Potentiates Ciprofloxacin Activity in the NorA Efflux Pump of Staphylococcus aureus. Scand. J. Infect. Dis. 2010, 42, 500–505. [Google Scholar] [CrossRef]
  118. Holler, J.G.; Slotved, H.-C.; Mølgaard, P.; Olsen, C.E.; Christensen, S.B. Chalcone Inhibitors of the NorA Efflux Pump in Staphylococcus aureus Whole Cells and Enriched Everted Membrane Vesicles. Bioorg. Med. Chem. 2012, 20, 4514–4521. [Google Scholar] [CrossRef]
  119. Hellewell, L.; Bhakta, S. Chalcones, Stilbenes and Ketones Have Anti-Infective Properties via Inhibition of Bacterial Drug-Efflux and Consequential Synergism with Antimicrobial Agents. Access. Microbiol. 2020, 2, acmi000105. [Google Scholar] [CrossRef]
  120. Pereda-Miranda, R.; Kaatz, G.W.; Gibbons, S. Polyacylated Oligosaccharides from Medicinal Mexican Morning Glory Species as Antibacterials and Inhibitors of Multidrug Resistance in Staphylococcus aureus. J. Nat. Prod. 2006, 69, 406–409. [Google Scholar] [CrossRef]
  121. Sabatini, S.; Gosetto, F.; Iraci, N.; Barreca, M.L.; Massari, S.; Sancineto, L.; Manfroni, G.; Tabarrini, O.; Dimovska, M.; Kaatz, G.W.; et al. Re-Evolution of the 2-Phenylquinolines: Ligand-Based Design, Synthesis, and Biological Evaluation of a Potent New Class of Staphylococcus aureus NorA Efflux Pump Inhibitors to Combat Antimicrobial Resistance. J. Med. Chem. 2013, 56, 4975–4989. [Google Scholar] [CrossRef]
  122. Felicetti, T.; Mangiaterra, G.; Cannalire, R.; Cedraro, N.; Pietrella, D.; Astolfi, A.; Massari, S.; Tabarrini, O.; Manfroni, G.; Barreca, M.L.; et al. C-2 Phenyl Replacements to Obtain Potent Quinoline-Based Staphylococcus aureus NorA Inhibitors. J. Enzym. Inhib. Med. Chem. 2020, 35, 584–597. [Google Scholar] [CrossRef]
  123. Felicetti, T.; Cannalire, R.; Nizi, M.G.; Tabarrini, O.; Massari, S.; Barreca, M.L.; Manfroni, G.; Schindler, B.D.; Cecchetti, V.; Kaatz, G.W.; et al. Studies on 2-Phenylquinoline Staphylococcus aureus NorA Efflux Pump Inhibitors: New Insights on the C-6 Position. Eur. J. Med. Chem. 2018, 155, 428–433. [Google Scholar] [CrossRef]
  124. Gibbons, S.; Oluwatuyi, M.; Kaatz, G.W. A Novel Inhibitor of Multidrug Efflux Pumps in Staphylococcus aureus. J. Antimicrob. Chemother. 2003, 51, 13–17. [Google Scholar] [CrossRef] [PubMed]
  125. Chovanová, R.; Mezovská, J.; Vaverková, Š.; Mikulášová, M. The Inhibition the Tet(K) Efflux Pump of Tetracycline Resistant Staphylococcus epidermidis by Essential Oils from Three Salvia Species. Lett. Appl. Microbiol. 2015, 61, 58–62. [Google Scholar] [CrossRef]
  126. Stermitz, F.R.; Tawara-Matsuda, J.; Lorenz, P.; Mueller, P.; Zenewicz, L.; Lewis, K. 5′-Methoxyhydnocarpin-D and Pheophorbide A:  Berberis Species Components That Potentiate Berberine Growth Inhibition of Resistant Staphylococcus aureus. J. Nat. Prod. 2000, 63, 1146–1149. [Google Scholar] [CrossRef]
  127. Kakarla, P.; Floyd, J.; Mukherjee, M.; Devireddy, A.R.; Inupakutika, M.A.; Ranweera, I.; Kc, R.; Shrestha, U.; Cheeti, U.R.; Willmon, T.M.; et al. Inhibition of the Multidrug Efflux Pump LmrS from Staphylococcus aureus by Cumin Spice Cuminum cyminum. Arch. Microbiol. 2017, 199, 465–474. [Google Scholar] [CrossRef]
  128. Samosorn, S.; Tanwirat, B.; Muhamad, N.; Casadei, G.; Tomkiewicz, D.; Lewis, K.; Suksamrarn, A.; Prammananan, T.; Gornall, K.C.; Beck, J.L.; et al. Antibacterial Activity of Berberine-NorA Pump Inhibitor Hybrids with a Methylene Ether Linking Group. Bioorg. Med. Chem. 2009, 17, 3866–3872. [Google Scholar] [CrossRef]
  129. Waditzer, M.; Bucar, F. Flavonoids as Inhibitors of Bacterial Efflux Pumps. Molecules 2021, 26, 6904. [Google Scholar] [CrossRef]
  130. Shiu, W.K.P.; 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]
  131. Chérigo, L.; Pereda-Miranda, R.; Fragoso-Serrano, M.; Jacobo-Herrera, N.; Kaatz, G.W.; Gibbons, S. Inhibitors of Bacterial Multidrug Efflux Pumps from the Resin Glycosides of Ipomoea murucoides. J. Nat. Prod. 2008, 71, 1037–1045. [Google Scholar] [CrossRef] [PubMed]
  132. Gupta, V.K.; Tiwari, N.; Gupta, P.; Verma, S.; Pal, A.; Srivastava, S.K.; Darokar, M.P. A Clerodane Diterpene from Polyalthia longifolia as a Modifying Agent of the Resistance of Methicillin Resistant Staphylococcus aureus. Phytomedicine 2016, 23, 654–661. [Google Scholar] [CrossRef]
  133. Kalia, N.P.; Mahajan, P.; Mehra, R.; Nargotra, A.; Sharma, J.P.; Koul, S.; Khan, I.A. Capsaicin, a Novel Inhibitor of the NorA Efflux Pump, Reduces the Intracellular Invasion of Staphylococcus aureus. J. Antimicrob. Chemother. 2012, 67, 2401–2408. [Google Scholar] [CrossRef]
  134. Singh, S.; Kalia, N.P.; Joshi, P.; Kumar, A.; Sharma, P.R.; Kumar, A.; Bharate, S.B.; Khan, I.A. Boeravinone B, A Novel Dual Inhibitor of NorA Bacterial Efflux Pump of Staphylococcus aureus and Human P-Glycoprotein, Reduces the Biofilm Formation and Intracellular Invasion of Bacteria. Front. Microbiol. 2017, 8, 1868. [Google Scholar] [CrossRef]
  135. Tintino, S.R.; Oliveira-Tintino, C.D.; Campina, F.F.; Silva, R.L.; Costa Mdo, S.; Menezes, I.R.; Calixto-Junior, J.T.; Siqueira-Junior, J.P.; Coutinho, H.D.; Leal-Balbino, T.C.; et al. Evaluation of the Tannic Acid Inhibitory Effect against the NorA Efflux Pump of Staphylococcus aureus. Microb. Pathog. 2016, 97, 9–13. [Google Scholar] [CrossRef]
  136. Ojeda-Sana, A.M.; Repetto, V.; Moreno, S. Carnosic Acid Is an Efflux Pumps Modulator by Dissipation of the Membrane Potential in Enterococcus faecalis and Staphylococcus aureus. World J. Microbiol. Biotechnol. 2013, 29, 137–144. [Google Scholar] [CrossRef]
  137. Vázquez, N.M.; Fiorilli, G.; Cáceres Guido, P.A.; Moreno, S. Carnosic Acid Acts Synergistically with Gentamicin in Killing Methicillin-Resistant Staphylococcus aureus Clinical Isolates. Phytomedicine 2016, 23, 1337–1343. [Google Scholar] [CrossRef]
  138. Chan, B.C.L.; Han, X.Q.; Lui, S.L.; Wong, C.W.; Wang, T.B.Y.; Cheung, D.W.S.; Cheng, S.W.; Ip, M.; Han, S.Q.B.; Yang, X.-S.; et al. Combating against Methicillin-Resistant Staphylococcus aureus-Two Fatty Acids from Purslane (Portulaca Oleracea L.) Exhibit Synergistic Effects with Erythromycin. J. Pharm. Pharmacol. 2015, 67, 107–116. [Google Scholar] [CrossRef] [PubMed]
  139. Gibbons, S.; Moser, E.; Kaatz, G.W. Catechin Gallates Inhibit Multidrug Resistance (MDR) in Staphylococcus aureus. Planta Med. 2004, 70, 1240–1242. [Google Scholar] [CrossRef]
  140. Joshi, P.; Singh, S.; Wani, A.; Sharma, S.; Jain, S.K.; Singh, B.; Gupta, B.D.; Satti, N.K.; Koul, S.; Khan, I.A.; et al. Osthol and Curcumin as Inhibitors of Human Pgp and Multidrug Efflux Pumps of Staphylococcus aureus: Reversing the Resistance against Frontline Antibacterial Drugs. Med. Chem. Comm. 2014, 5, 1540–1547. [Google Scholar] [CrossRef]
  141. Sobisch, L.-Y.; Rogowski, K.M.; Fuchs, J.; Schmieder, W.; Vaishampayan, A.; Oles, P.; Novikova, N.; Grohmann, E. Biofilm Forming Antibiotic Resistant Gram-Positive Pathogens Isolated From Surfaces on the International Space Station. Front. Microbiol. 2019, 10, 543. [Google Scholar] [CrossRef]
  142. Davies, D. Understanding Biofilm Resistance to Antibacterial Agents. Nat. Rev. Drug Discov. 2003, 2, 114–122. [Google Scholar] [CrossRef]
  143. Bridier, A.; Sanchez-Vizuete, P.; Guilbaud, M.; Piard, J.-C.; Naitali, M.; Briandet, R. Biofilm-Associated Persistence of Food-Borne Pathogens. Food Microbiol. 2015, 45, 167–178. [Google Scholar] [CrossRef]
  144. Fernández, L.; Escobedo, S.; Gutiérrez, D.; Portilla, S.; Martínez, B.; García, P.; Rodríguez, A. Bacteriophages in the Dairy Environment: From Enemies to Allies. Antibiotics 2017, 6, 27. [Google Scholar] [CrossRef]
  145. O’Toole, G.; Kaplan, H.B.; Kolter, R. Biofilm Formation as Microbial Development. Annu. Rev. Microbiol. 2000, 54, 49–79. [Google Scholar] [CrossRef] [PubMed]
  146. Guo, X.-P.; Yang, Y.; Lu, D.-P.; Niu, Z.-S.; Feng, J.-N.; Chen, Y.-R.; Tou, F.-Y.; Garner, E.; Xu, J.; Liu, M.; et al. Biofilms as a Sink for Antibiotic Resistance Genes (ARGs) in the Yangtze Estuary. Water Res. 2018, 129, 277–286. [Google Scholar] [CrossRef]
  147. Ratajczak, M.; Kamińska, D.; Nowak-Malczewska, D.M.; Schneider, A.; Dlugaszewska, J. Relationship between Antibiotic Resistance, Biofilm Formation, Genes Coding Virulence Factors and Source of Origin of Pseudomonas aeruginosa Clinical Strains. Ann. Agric. Environ. Med. AAEM 2021, 28, 306–313. [Google Scholar] [CrossRef]
  148. Qi, L.; Li, H.; Zhang, C.; Liang, B.; Li, J.; Wang, L.; Du, X.; Liu, X.; Qiu, S.; Song, H. Relationship between Antibiotic Resistance, Biofilm Formation, and Biofilm-Specific Resistance in Acinetobacter baumannii. Front. Microbiol. 2016, 7, 483. [Google Scholar] [CrossRef]
  149. Hoffman, L.R.; D’Argenio, D.A.; MacCoss, M.J.; Zhang, Z.; Jones, R.A.; Miller, S.I. Aminoglycoside Antibiotics Induce Bacterial Biofilm Formation. Nature 2005, 436, 1171–1175. [Google Scholar] [CrossRef]
  150. Høiby, N.; Henneberg, K.-Å.; Wang, H.; Stavnsbjerg, C.; Bjarnsholt, T.; Ciofu, O.; Johansen, U.R.; Sams, T. Formation of Pseudomonas aeruginosa Inhibition Zone during Tobramycin Disk Diffusion Is Due to Transition from Planktonic to Biofilm Mode of Growth. Int. J. Antimicrob. Agents 2019, 53, 564–573. [Google Scholar] [CrossRef] [PubMed]
  151. 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]
  152. Pletzer, D.; Mansour, S.C.; Hancock, R.E.W. Synergy between Conventional Antibiotics and Anti-Biofilm Peptides in a Murine, Sub-Cutaneous Abscess Model Caused by Recalcitrant ESKAPE Pathogens. PLoS Pathog. 2018, 14, e1007084. [Google Scholar] [CrossRef] [PubMed]
  153. Dastgheyb, S.; Parvizi, J.; Shapiro, I.M.; Hickok, N.J.; Otto, M. Effect of Biofilms on Recalcitrance of Staphylococcal Joint Infection to Antibiotic Treatment. J. Infect. Dis. 2015, 211, 641–650. [Google Scholar] [CrossRef]
  154. Lebeaux, D.; Ghigo, J.-M.; Beloin, C. Biofilm-Related Infections: Bridging the Gap between Clinical Management and Fundamental Aspects of Recalcitrance toward Antibiotics. Microbiol. Mol. Biol. Rev. MMBR 2014, 78, 510–543. [Google Scholar] [CrossRef]
  155. Flemming, H.-C.; Wingender, J. The Biofilm Matrix. Nat. Rev. Microbiol. 2010, 8, 623–633. [Google Scholar] [CrossRef] [PubMed]
  156. McCarthy, H.; Rudkin, J.K.; Black, N.S.; Gallagher, L.; O’Neill, E.; O’Gara, J.P. Methicillin Resistance and the Biofilm Phenotype in Staphylococcus aureus. Front. Cell. Infect. Microbiol. 2015, 5, 1. [Google Scholar] [CrossRef] [PubMed]
  157. Ito, A.; Taniuchi, A.; May, T.; Kawata, K.; Okabe, S. Increased Antibiotic Resistance of Escherichia coli in Mature Biofilms. Appl. Environ. Microbiol. 2009, 75, 4093–4100. [Google Scholar] [CrossRef]
  158. Kvist, M.; Hancock, V.; Klemm, P. Inactivation of Efflux Pumps Abolishes Bacterial Biofilm Formation. Appl. Environ. Microbiol. 2008, 74, 7376–7382. [Google Scholar] [CrossRef] [PubMed]
  159. Christena, L.R.; Subramaniam, S.; Vidhyalakshmi, M.; Mahadevan, V.; Sivasubramanian, A.; Nagarajan, S. Dual Role of Pinostrobin-a Flavonoid Nutraceutical as an Efflux Pump Inhibitor and Antibiofilm Agent to Mitigate Food Borne Pathogens. RSC Adv. 2015, 5, 61881–61887. [Google Scholar] [CrossRef]
  160. Baugh, S.; Ekanayaka, A.S.; Piddock, L.J.V.; Webber, M.A. Loss of or Inhibition of All Multidrug Resistance Efflux Pumps of Salmonella enterica Serovar Typhimurium Results in Impaired Ability to Form a Biofilm. J. Antimicrob. Chemother. 2012, 67, 2409–2417. [Google Scholar] [CrossRef]
  161. Rezaie, P.; Pourhajibagher, M.; Chiniforush, N.; Hosseini, N.; Bahador, A. The Effect of Quorum-Sensing and Efflux Pumps Interactions in Pseudomonas aeruginosa Against Photooxidative Stress. J. Lasers Med. Sci. 2018, 9, 161–167. [Google Scholar] [CrossRef]
  162. Kaur, B.; Gupta, J.; Sharma, S.; Sharma, D.; Sharma, S. Focused Review on Dual Inhibition of Quorum Sensing and Efflux Pumps: A Potential Way to Combat Multi Drug Resistant Staphylococcus aureus Infections. Int. J. Biol. Macromol. 2021, 190, 33–43. [Google Scholar] [CrossRef] [PubMed]
  163. Rahmati, S.; Yang, S.; Davidson, A.L.; Zechiedrich, E.L. Control of the AcrAB Multidrug Efflux Pump by Quorum-Sensing Regulator SdiA. Mol. Microbiol. 2002, 43, 677–685. [Google Scholar] [CrossRef] [PubMed]
  164. Schembri, M.A.; Kjaergaard, K.; Klemm, P. Global Gene Expression in Escherichia coli Biofilms. Mol. Microbiol. 2003, 48, 253–267. [Google Scholar] [CrossRef] [PubMed]
  165. 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]
  166. Short, F.L.; Liu, Q.; Shah, B.; Clift, H.E.; Naidu, V.; Li, L.; Prity, F.T.; Mabbutt, B.C.; Hassan, K.A.; Paulsen, I.T. The Acinetobacter baumannii Disinfectant Resistance Protein, AmvA, Is a Spermidine and Spermine Efflux Pump. Commun. Biol. 2021, 4, 1114. [Google Scholar] [CrossRef] [PubMed]
  167. Liu, J.; Zhang, J.; Guo, L.; Zhao, W.; Hu, X.; Wei, X. Inactivation of a Putative Efflux Pump (LmrB) in Streptococcus Mutans Results in Altered Biofilm Structure and Increased Exopolysaccharide Synthesis: Implications for Biofilm Resistance. Biofouling 2017, 33, 481–493. [Google Scholar] [CrossRef] [PubMed]
  168. 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. [Google Scholar] [CrossRef]
  169. Lamut, A.; Peterlin Mašič, L.; Kikelj, D.; Tomašič, T. Efflux Pump Inhibitors of Clinically Relevant Multidrug Resistant Bacteria. Med. Res. Rev. 2019, 39, 2460–2504. [Google Scholar] [CrossRef]
  170. 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]
  171. Singh, S.; Datta, S.; Narayanan, K.B.; Rajnish, K.N. Bacterial Exo-Polysaccharides in Biofilms: Role in Antimicrobial Resistance and Treatments. J. Genet. Eng. Biotechnol. 2021, 19, 140. [Google Scholar] [CrossRef]
  172. Tomaś, N.; Myszka, K.; Wolko, Ł.; Nuc, K.; Szwengiel, A.; Grygier, A.; Majcher, M. Effect of Black Pepper Essential Oil on Quorum Sensing and Efflux Pump Systems in the Fish-Borne Spoiler Pseudomonas psychrophila KM02 Identified by RNA-Seq, RT-QPCR and Molecular Docking Analyses. Food Control 2021, 130, 108284. [Google Scholar] [CrossRef]
  173. Aggarwal, S.; Singh, D.V. Efflux Pumps and Biofilm Formation by Both Methicillin-Resistant and Methicillin-Sensitive Staphylococcus aureus Strains. Indian. J. Exp. Biol. IJEB 2020, 58, 527–538. [Google Scholar] [CrossRef]
  174. Christena, L.R.; Mangalagowri, V.; Pradheeba, P.; Ahmed, K.B.A.; Shalini, B.I.S.; Vidyalakshmi, M.; Anbazhagan, V.; Subramanian, N.S. Copper Nanoparticles as an Efflux Pump Inhibitor to Tackle Drug Resistant Bacteria. RSC Adv. 2015, 5, 12899–12909. [Google Scholar] [CrossRef]
  175. Yu, Y.; Zhao, Y.; He, Y.; Pang, J.; Yang, Z.; Zheng, M.; Yin, R. Inhibition of Efflux Pump Encoding Genes and Biofilm Formation by Sub-Lethal Photodynamic Therapy in Methicillin Susceptible and Resistant Staphylococcus aureus. Photodiagnosis Photodyn. Ther. 2022, 39, 102900. [Google Scholar] [CrossRef] [PubMed]
  176. Zimmermann, S.; Klinger-Strobel, M.; Bohnert, J.A.; Wendler, S.; Rödel, J.; Pletz, M.W.; Löffler, B.; Tuchscherr, L. Clinically Approved Drugs Inhibit the Staphylococcus aureus Multidrug NorA Efflux Pump and Reduce Biofilm Formation. Front. Microbiol. 2019, 10, 2762. [Google Scholar] [CrossRef]
  177. Abd El-Baky, R.M.; Sandle, T.; John, J.; Abuo-Rahma, G.E.-D.A.; Hetta, H.F. A Novel Mechanism of Action of Ketoconazole: Inhibition of the NorA Efflux Pump System and Biofilm Formation in Multidrug-Resistant Staphylococcus aureus. Infect. Drug Resist. 2019, 12, 1703–1718. [Google Scholar] [CrossRef] [PubMed]
  178. Matsumura, K.; Furukawa, S.; Ogihara, H.; Morinaga, Y. Roles of Multidrug Efflux Pumps on the Biofilm Formation of Escherichia coli K-12. Biocontrol Sci. 2011, 16, 69–72. [Google Scholar] [CrossRef] [PubMed]
  179. Pasqua, M.; Grossi, M.; Scinicariello, S.; Aussel, L.; Barras, F.; Colonna, B.; Prosseda, G. The MFS Efflux Pump EmrKY Contributes to the Survival of Shigella within Macrophages. Sci. Rep. 2019, 9, 2906. [Google Scholar] [CrossRef]
  180. Pelgrift, R.Y.; Friedman, A.J. Nanotechnology as a Therapeutic Tool to Combat Microbial Resistance. Adv. Drug Deliv. Rev. 2013, 65, 1803–1815. [Google Scholar] [CrossRef]
  181. Nejabatdoust, A.; Zamani, H.; Salehzadeh, A. Functionalization of ZnO Nanoparticles by Glutamic Acid and Conjugation with Thiosemicarbazide Alters Expression of Efflux Pump Genes in Multiple Drug-Resistant Staphylococcus aureus Strains. Microb. Drug Resist. 2019, 25, 966–974. [Google Scholar] [CrossRef]
  182. Seena, S.; Rai, A. Nanoengineering Approaches to Fight Multidrug-Resistant Bacteria. In Non-Traditional Approaches to Combat. Antimicrobial Drug Resistance; Wani, M.Y., Ahmad, A., Eds.; Springer Nature: Singapore, 2023; pp. 221–248. ISBN 978-981-19916-7-7. [Google Scholar]
  183. Lekshmi, M.; Ammini, P.; Kumar, S.; Varela, M.F. The Food Production Environment and the Development of Antimicrobial Resistance in Human Pathogens of Animal Origin. Microorganisms 2017, 5, 11. [Google Scholar] [CrossRef]
  184. Lowrence, R.C.; Subramaniapillai, S.G.; Ulaganathan, V.; Nagarajan, S. Tackling Drug Resistance with Efflux Pump Inhibitors: From Bacteria to Cancerous Cells. Crit. Rev. Microbiol. 2019, 45, 334–353. [Google Scholar] [CrossRef]
Figure 1. Bacterial antimicrobial transporter superfamilies. The ABC transporter superfamily consists of primary active transport systems that use ATP hydrolysis to drive antimicrobial efflux from bacterial cells. The antimicrobial efflux pumps belonging to the MFS, SMR, MATE, and PACE use ion/drug antiport mechanisms to extrude antimicrobial agents from the cytoplasm. The RND superfamily transporters are multi-component systems to efflux antimicrobial solutes to the external milieu of the bacterium.
Figure 1. Bacterial antimicrobial transporter superfamilies. The ABC transporter superfamily consists of primary active transport systems that use ATP hydrolysis to drive antimicrobial efflux from bacterial cells. The antimicrobial efflux pumps belonging to the MFS, SMR, MATE, and PACE use ion/drug antiport mechanisms to extrude antimicrobial agents from the cytoplasm. The RND superfamily transporters are multi-component systems to efflux antimicrobial solutes to the external milieu of the bacterium.
Biomedicines 11 01448 g001
Figure 2. Predicted structure of LmrS from S. aureus. The LmrS multidrug efflux pump of the MFS has 14 predicted transmembrane domains that are α-helical in nature.
Figure 2. Predicted structure of LmrS from S. aureus. The LmrS multidrug efflux pump of the MFS has 14 predicted transmembrane domains that are α-helical in nature.
Biomedicines 11 01448 g002
Figure 3. Mechanism of antibiotic recalcitrance in the biofilm matrix. The biofilm’s color gradient (yellow) depicts the availability of nutrients and oxygen from high to low and darker to lighter shades. The color differentiation in the bacterial cells depicts their physiological state, with the lighter ones in active phases and the darker green ones in less active stationary phases. Biofilm greatly reduced the diffusion of certain antibiotic molecules (blue triangles). 1. The exopolysaccharide. 2. eDNA. 3. Bacterial autolysis releases antibiotic-binding and degrading molecules into the matrix. 4. Persister cells. 5. Mixed species in the biofilm. 6. Horizontal gene transfer and increased frequency of mutation. 7. Quorum sensing. 8. Biofilm-specific antibiotic resistance genes. 9. Efflux pump.
Figure 3. Mechanism of antibiotic recalcitrance in the biofilm matrix. The biofilm’s color gradient (yellow) depicts the availability of nutrients and oxygen from high to low and darker to lighter shades. The color differentiation in the bacterial cells depicts their physiological state, with the lighter ones in active phases and the darker green ones in less active stationary phases. Biofilm greatly reduced the diffusion of certain antibiotic molecules (blue triangles). 1. The exopolysaccharide. 2. eDNA. 3. Bacterial autolysis releases antibiotic-binding and degrading molecules into the matrix. 4. Persister cells. 5. Mixed species in the biofilm. 6. Horizontal gene transfer and increased frequency of mutation. 7. Quorum sensing. 8. Biofilm-specific antibiotic resistance genes. 9. Efflux pump.
Biomedicines 11 01448 g003
Table 1. Natural compounds as efflux pump inhibitors (EPIs) against efflux pumps of the MFS family.
Table 1. Natural compounds as efflux pump inhibitors (EPIs) against efflux pumps of the MFS family.
InhibitorEfflux PumpReference
Plant-derived alkaloid compounds (reserpine, piperines, and piperine analogs)NorA, Bmr, MdeA, LmrA, PmrA[105,106,107]
Flavonoids (genistein, sarothrin)NorA[108,109]
Flavones (chrysosplenol-D and chrysoplenetin)NorA[110]
Diterpenes (ferruginol)NorA[111,112]
IsolflavonesNorA[113]
CoumarinsNorA[114]
Kaempferol rhamnosideNorA[115]
2,6-dimethyl-4-phenyl-pyridine-3,5-dicarboxylic acid diethyl esterNorA, MsrA[116]
IndirubinNorA[117]
Chalcones (4-phenoxy-4′-dimethylamino ethoxy chalcone) NorA[118,119]
Oligosaccharides (orizabin) NorA[120]
Derivatives of 2-phenylquinoline NorA[121,122,123]
Abietane diterpenesTet(K), Msr(A) [124]
Essential oilsTetK[125]
5′-Methoxyhydnocarpin-D and Pheophorbide ANorA[126]
Cumin seed oil, cumin aldehyde LmrS [127]
Plant-derived alkaloid compounds (berberine and palmatine)NorA, MdfA[103,128,129]
SarothrinNorA[108]
OlympicinNorA[130]
MurucoidinsNorA[131]
Clerodane diterpene 16α-
hydroxycleroda-3,13 (14)-Z-dien-15,16-olid 6
NorB, NorC[132]
Verapamil, capsaicin, boeravinone BNorA, QacA[133,134]
Cholecalciferol and alpha-tocopherolTetK, MsrA[101,135]
Carnosic acidMsrA, TetK, and NorA[136,137]
Linoleic and oleic acidsMsrA[138]
Epigallocatechin gallate, Epicatechin gallate TetK[139]
Osthtol NorA, MdeA, TetK, MsrA[140]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Varela, M.F.; Stephen, J.; Bharti, D.; Lekshmi, M.; Kumar, S. Inhibition of Multidrug Efflux Pumps Belonging to the Major Facilitator Superfamily in Bacterial Pathogens. Biomedicines 2023, 11, 1448. https://doi.org/10.3390/biomedicines11051448

AMA Style

Varela MF, Stephen J, Bharti D, Lekshmi M, Kumar S. Inhibition of Multidrug Efflux Pumps Belonging to the Major Facilitator Superfamily in Bacterial Pathogens. Biomedicines. 2023; 11(5):1448. https://doi.org/10.3390/biomedicines11051448

Chicago/Turabian Style

Varela, Manuel F., Jerusha Stephen, Deeksha Bharti, Manjusha Lekshmi, and Sanath Kumar. 2023. "Inhibition of Multidrug Efflux Pumps Belonging to the Major Facilitator Superfamily in Bacterial Pathogens" Biomedicines 11, no. 5: 1448. https://doi.org/10.3390/biomedicines11051448

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