Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2010 Aug 20.
Published in final edited form as: Drugs. 2009 Aug 20;69(12):1555–1623. doi: 10.2165/11317030-000000000-00000

Efflux-Mediated Drug Resistance in Bacteria: an Update

Xian-Zhi Li 1, Hiroshi Nikaido 2
PMCID: PMC2847397  NIHMSID: NIHMS183678  PMID: 19678712

Abstract

Drug efflux pumps play a key role in drug resistance and also serve other functions in bacteria. There has been a growing list of multidrug and drug-specific efflux pumps characterized from bacteria of human, animal, plant and environmental origins. These pumps are mostly encoded on the chromosome although they can also be plasmid-encoded. A previous article (Li X-Z and Nikaido H, Drugs, 2004; 64[2]: 159–204) had provided a comprehensive review regarding efflux-mediated drug resistance in bacteria. In the past five years, significant progress has been achieved in further understanding of drug resistance-related efflux transporters and this review focuses on the latest studies in this field since 2003. This has been demonstrated in multiple aspects that include but are not limited to: further molecular and biochemical characterization of the known drug efflux pumps and identification of novel drug efflux pumps; structural elucidation of the transport mechanisms of drug transporters; regulatory mechanisms of drug efflux pumps; determining the role of the drug efflux pumps in other functions such as stress responses, virulence and cell communication; and development of efflux pump inhibitors. Overall, the multifaceted implications of drug efflux transporters warrant novel strategies to combat multidrug resistance in bacteria.


Antibacterial resistance continues to be a global public health concern, threatens the effectiveness of antibacterial therapy, and also challenges the efforts for developing novel antibacterials. A variety of bacterial pathogens isolated globally have now become multidrug resistant (MDR, here also used for “multidrug resistance”). Although antibacterial resistance occurs by numerous mechanisms, including enzymatic inactivation or modification of drugs, drug target alteration or protection, and lack of pro-drug activation, that due to the increased active efflux of the drugs is a major concern especially because a single species of multidrug efflux pump can produce a simultaneous resistance to a number of drugs, an MDR phenotype.[14] Efflux also acts synergistically with other resistance mechanisms to provide elevated level of resistance of clinical significance.[1]

A comprehensive review on bacterial drug efflux was prepared by us previously.[1] Since then, a very large amount of literature has been published in this field. In this article, we cover the advances in bacterial efflux systems since 2003 with emphasis on the clinical relevance of the drug exporters in various bacteria. In order to keep the size of this article within an accessible range, earlier literature cited in the previous review[1] was usually omitted. Recently, several reviews that focus on various aspects of the drug efflux transporters or MDR have been published.[310]

1. Drug Resistance in Bacteria: Emerging Features

Drug resistance in bacteria continues to escalate globally with the emergence of novel resistance patterns. This renders antibacterial therapy less effective and may lead us back to the “pre-antibiotic era”, thus the need for innovative approaches to tackle antibacterial resistance now.[11] We begin with the emerging features of antibacterial-resistant bacteria, in which drug efflux plays an important role. First, MDR is a phenotype increasingly associated with many pathogens. This also includes the extensive drug resistance (XDR) in Mycobacterium tuberculosis[12] and pan-resistance in especially Gram-negative bacteria[13] MDR can be caused by simultaneous presence of multiple individual resistance mechanisms, each of which can be either plasmid- or chromosome-mediated.[10] In a typical example, an R plasmid, which is often transferable or conjugative, causes MDR because it contains multiple resistance genes on a single molecule of DNA. Furthermore, the so-called resistance island, often on a chromosome, may contain a cluster of multiple resistance genes.[1417] Resistance genes are often also co-present with mobile genetic elements, e.g., transposons and integrons, and in this manner they move as a block between molecules of DNA, for example among different R plasmids and between plasmid and chromosome. In addition, MDR can be mediated or enhanced by the inaccessibility of drugs to their cellular targets as a result of the outer membrane (OM) impermeability and active drug efflux, which are often encoded by chromosomal genes.

Second, novel mechanisms of resistance with emerging resistance determinants have been reported. β-Lactamases continue to evolve with CTX-M type extended-spectrum β-lactamases, AmpC, and carbapenamases as major threats for β-lactam therapy.[18, 19] A variant of aminoglycoside-modifying enzyme, AAC(6′)-Ib-cr, can also modify fluoroquinolones and thus yield fluoroquinolone resistance.[20] Plasmid-mediated fluoroquinolone resistance due to the target protection (by qnr genes encoding proteins with a pentapeptide repeat) or efflux mechanism (by qep genes) is increasingly observed.[2123] Most worrisome are the pathogenic bacteria with a combination of resistance genes associated with mobile genetic elements. For instance, Gram-negative bacteria such as Acinetobacter and Pseudomonas as well as Enterobacteriaceae often possess this type of mechanisms in addition to drug efflux systems.[1, 6, 18, 24] A Salmonella Waycross isolate, obtained from a hospitalized elderly, possesses plasmid-borne class 1 integron harbouring resistance genes blaIMP-4 (encoding for a class B metallo-β-lactamase), aacA4 (an aminoglycoside-modifying enzyme), catB (chloramphenicol-acetyltransferase), qnrB4 (a pentapeptide repeat protein) and qacG (an efflux pump).[25]

Third, the resistant pathogens are not only isolated from hospitals and communities but also increasingly derived from other sources, e.g., animals, food products and environments.[15, 2628] Resistant bacteria and genetic determinants of resistance can transfer between animals and humans. This may be exemplified by the evolution and spread of methicillin-resistant Staphylococcus aureus (MRSA).[29, 30]

2. Drug Efflux Pumps in Bacteria: Structures and Mechanisms

Bacterial drug efflux pumps have been categorized into five families, i.e., the ATP-binding cassette (ABC) superfamily,[5] the major facilitator superfamily (MFS),[31] the multidrug and toxic compound extrusion (MATE) family,[32] the small multidrug resistance (SMR) family (a subgroup of the drug/metabolite transporter superfamily[33]), and the resistance-nodulation-division (RND) superfamily.[1, 8, 34, 35] In particular, drug exporters belonging to RND family play a key role in clinically relevant resistance in Gram-negative bacteria. A major achievement in the field has been the structural and biochemical elucidation of drug efflux pumps and these will be highlighted below.

2.1 Drug Efflux Transporters

2.1.1 RND Transporters

RND efflux systems, which function as proton/drug antiporters, are particularly widespread among Gram-negative bacteria and catalyze the active efflux of a wide variety of antibacterial substrates including many antibiotics and chemotherapeutic agents. Homologues exist even in higher animals, including the Niemann-Pick C1 Like 1 protein, shown to be involved in cholesterol absorption from intestine.[36] The extensive studies in recent years with the archetypal bacterial RND pumps AcrAB-TolC of Escherichia coli and MexAB-OprM of P. aeruginosa have revealed both the structure and mechanisms of RND pumps in the efflux of a very wide range of agents. RND transporters (e.g., AcrB and MexB) have large periplasmic domains and form tripartite complexes with the periplasmic adaptor proteins or membrane fusion proteins (MFPs) (AcrA and MexA) and OM channels (TolC and OprM).[8, 35, 37] The latter will be discussed below in two separate sections.

AcrB transporter is a homotrimer, and characteristically contains a large periplasmic domain that equals the transmembrane domain in size.[1, 38] When AcrB protein was co-crystallized with drugs, dyes, or deoxycholate, the ligands were found not only on the wall of central cavity of the transmembrane domain but also at the side of a large external cleft in the periplasmic domain.[1, 39, 40] These drugs may be on their correct pathway for extrusion,[3941] but lipophilic drugs may bind nonspecifically to hydrophobic spots on the protein surface. However, AcrB that was accidentally crystallized (together with a small protein YajC) contained ampicillin from the growth medium at the periplasmic site.[42]

Site-directed mutagenesis and/or structural studies have identified the key residues in the transmembrane domains, residues Asp407, Asp408, Lys940 and Thr978 in AcrB.[1, 38, 39, 4351] The recent titration study with dicyclohexylcarbodiimide showed the modification of Asp408, which has a pKa of 7.4.[52] These residues probably function as the proton relay network, ultimately resulting in drug extrusion.[45] In the periplasmic domain, a phenylalanine-rich binding site around Phe178 and Phe615 is revealed and the Phe610Ala point mutation has a significant impact on transport activity.[50] The replacement of AcrB residues 615 to 628 with the homologous MexB sequence (AcrB-615–628MexB) and more specifically the Gly616Asn substitution in AcrB have both resulted in the reduction of macrolide resistance of AcrB.[51] A single Val610Phe substitution in YhiV (MdtF), an AcrB homologue, altered spectrum of MDR by retaining or increasing the resistance to fluoroquinolones, linezolid, novobiocin and tetracyclines, while decreasing resistance to azithromycin and telithromycin, suggesting the involvement of the region around the residue in determining substrate recognition.[47] In vitro reconstitution of AcrD exporter showed that aminoglycosides are captured from both periplasm and cytoplasm,[53] consistent with our early prediction that β-lactams such as dianionic agents carbenicillin and ceftriaxone are captured from the periplasm.[54]

New crystallographic studies have now revealed the asymmetric trimer structure of AcrB where each AcrB protomer in the trimeric assembly goes through a cycle of conformational changes during drug export (Fig. 1).[5557] This asymmetric structure suggests the possible route of substrate binding and extrusion as well as the presence of an open pathway between the substrate binding pocket and the periplasm. It is especially important that Murakami and coworkers found that one protomer bound minocycline or doxorubicin in a hydrophobic binding site of the periplasmic domain, containing Phe610 mentioned above[55], which is separate from the external cleft. The three-step functionally rotating mechanism of transport describes each of the three protomers in one of the three functional states (i.e., access [loose], binding [tight], and extrusion [open]) and predicts that the drug bound in the periplasmic domain is extruded through the conformational change initiated by the protonation of one of the residues in the aforementioned network within the transmembrane domain.[55] (Recently an asymmetric structure of MexB trimer was elucidated, with the binding protomer containing the detergent molecule dodecylmaltoside within its periplasmic binding pocket[58]). The functionally rotating mechanism has also been further supported by recent biochemical studies with disulfide cross-linking as well as by the behaviour of covalently linked AcrB protomers.[5961] In particular, the new approach of Takatsuka and Nikaido[61] via the use of covalently linked trimer expressed from a constructed giant gene is a powerful tool for studying the transport mechanisms of drug pumps.[62] The “linked trimer” AcrB was not only expressed well but also functional in providing resistance to antibacterials. Intriguingly, the inactivation of only one of the three protomeric units in the linked trimer by either mutations in the “proton relay network” in the transmembrane domain or disulfide cross-linking of the external cleft in the periplasmic domain resulted in the total activity loss of the entire trimeric complex, thus providing a strong biochemical evidence for the functionally rotating mechanism of RND pump action.[61]

Fig. 1.

Fig. 1

Crystal structures of multidrug efflux transporters exemplified by AcrAB-TolC[55, 107, 119] and EmrD[70] of E. coli and Sav1866 of S. aureus.[56] Instead of AcrA, we show the recent complete structure of its homologue MexA.[108] See text and the relevant references for details of the structural properties of these transporters. The figures were drawn by Pymol (http://www.pymol.org) by using the coordinate files 2DRD (AcrB), 2V4D (MexA), 2VDE (TolC, an open form), 2GFP (EmrD), and 2HYD (Sav1866), obtained from the Protein Data Bank. The models were colored in rainbow colors (N-terminus blue, C-terminus red) and the approximate positions of membrane bilayers are indicated by horizontal lines. Note that for oligomeric proteins, the rainbow color was selected from the N-terminus of one protomer all the way to the C-terminus of another protomer. Proteins were positioned so that the external portion is up in the figure. The bound minocycline in the AcrB structure is shown in red rods (highlighted by a green arrow).

Steady-state fluorescence polarization was used to assess the interactions between fluorescent ligands and purified AcrB transporter;[63] however, again there is no guarantee that the ligands are binding to the relevant site within AcrB. Recently, the kinetic constants of AcrB were successfully determined in intact cells by using cephalosporins as the substrates.[64] Among the compounds tested, nitrocefin, with its two aromatic substituents, appeared to have the highest affinity to AcrB, but with only a low value of kcat (about 10/s). In contrast, compounds that were useful clinically, such as cefamandole, cephalothin, and cephaloridine apparently had lower affinities to AcrB but showed much higher values of kcat. Most remarkably, positive cooperativity was evident with the efflux of these compounds. Finally, cefazolin, with two hydrophilic substituents (a tetrazole and a thiadiazole), showed little evidence of efflux by AcrB.[64]

In an approach similar to the study of QacR and BmrR,[1] the repressor TtgR from P. putida was crystallized with antibiotics and plant antimicrobials, and revealed a large binding pocket with capacity for multiple binding interactions.[65] The repressor AcrR was crystallized;[66] however it is not known if it binds any small inducer molecules.

2.1.2 MFS Transporters

This family is known to represent the largest group of secondary active transporters[31] with well characterized multidrug pumps, including Bmr and Blt of Bacillus subtilis, MdfA of E. coli, LmrP of Lactobacillus lactis, NorA and QacA of S. aureus.[1] These transporters are antiporters that are thought to function as monomers. However, in Gram-negative bacteria, MFS efflux systems can function as components of tripartite systems together with the additional MFPs and OM channels (e.g., EmrAB-TolC and EmrKY-TolC of E. coli).[1] These systems enable the transporter to efficiently export the substrates across the double membranes of Gram-negative bacteria. This is unlike several single-component MFS transporters, which can export drugs only into periplasm.[1] However, even the transporters of the latter type can increase the resistance when the drugs exported into the periplasm are further taken up by the tripartite RND pumps, as shown by the pioneering study from the Lomovskaya group.[67]

To date, crystal structures are available for several MFS transporters such as the lactose/H+ permease LacY,[68] the glycerol-3-phosphate transporter GlpT[69] and the multidrug transporter EmrD,[70] which are all from E. coli. The common folding pattern consisting of two transmembrane domains that surround a substrate translocation pore may be shared by most MFS members.[71] However, LacY and GlpT permeases transport a relatively narrow range of structurally related substrates,[68, 69] while EmrD accommodates a range of hydrophobic agents including benzalkonium, carbonyl cyanide m-chlorophenylhydrazone, and sodium dodecylsulfate.[1, 72] The EmrD structure demonstrates an interior with mostly hydrophobic residues and also displays two long loops extended into the inner leaflet side of the cell membrane. The loop region can serve to recognize and bind substrates directly from the lipid bilayer (Fig. 1).[70] More recently, a low-resolution structure for MdfA has become available.[73] With the particular examples of MdfA and LmrP, a recent review discusses the physiological significance, multisubstrate specificities and the structural mechanisms of the MFS multidrug transporters.[74] LmrP functions as a facilitated diffusion catalyst in the absence of proton-motive force.[75] Site-directed mutagenesis studies with QacA produced interesting results, for example, a tryptophan-to-alanine change in the outside surface was compensated by changes in the inside loops of the protein.[76] An in-vitro study revealed that EmrAB of E. coli forms a dimer in contrast to the trimeric RND AcrB.[77]

2.1.3 MATE Transporters

This family is represented by NorM of Vibro parahaemolyticus[1, 32] and confers resistance to multiple cationic toxic agents (including fluoroquinolones) as H+- or Na+-antiporters. However, the substrate profiles are generally narrower than those of the RND transporters. Although there are only about 20 MATE transporters characterized to date,[32] the bacterial genome sequences contain many more examples, and intriguingly, the MATE proteins are present in all kingdoms of life.[78] For instance, two MATE genes were identified on human chromosome 17, named hMATE1 and hMATE.[79] When expressed in HEK293 cells, hMATE1 mediated H+-coupled electroneutral exchange of organic cations.[79, 80] A phylogenetic analysis has classified the mammalian MATE proteinsinto three subfamilies.[78, 81] To date, no crystal structures are available for any MATE transporters.

As discussed later, the majority of the bacterial MATE pumps have been identified by expression in a heterologous, antimicrobial-hypersusceptible E. coli. Thus, the functional significance of these pumps in the native hosts is usually unclear. The regulation of MATE pumps will be discussed below in section 9 with the staphylococcal MepR-regulated MepA pump.

2.1.4 SMR Transporters

This family of transporters is represented by EmrE of E. coli, which functions as a homodimer of a small four-transmembrane protein.[1, 33] The SMR family contains >250 annotated members, and is now grouped into three subclasses: the small multidrug pumps, the paired SMR proteins, and suppressors of groEL mutant proteins.[82] The SMR proteins may be encoded on the chromosomes or on plasmids and may be associated with integrons. The substrate specificity is not limited to the disinfectants and can extend to clinically relevant antibacterials such as aminoglycosides.[82, 83] Although EmrE exports its substrates only into the periplasm, it can cause significant resistance as the substrate is then taken up by constitutive tripartite RND pumps, such as AcrAB-TolC,[64, 84, 85] along the lines formulated earlier.[67]

EmrE appears to function as a dimer. The orientation of the two protomers within the dimer has been a subject of controversy. Biochemical studies indicated that the two protomers are inserted into the membrane in a parallel orientation.[86] In contrast, electron[87] and x-ray crystallography[88] suggested an antiparallel orientation, which is also favoured by another study.[89] In this connection, the paired SMR proteins[90] may be relevant. EbrAB of B. subtitilis is a heterodimer composed of two polypeptides, EbrA and EbrB, which are both required for activity. Importantly, EbrAB displays an anti-parallel membrane topology.[91] Thus one can argue for the presence of a conserved architecture for all SMR family members as antiparallel dimers.[88] However, it is difficult to imagine that EmrE is inserted into the membrane in two opposite orientations at equal probability, and it is difficult to exclude completely the possibility that the antiparallel dimer is an artefact of dissociation-association process during sample preparation. An especially strong result for the parallel orientation is that two emrE genes, linked together with very short (down to two amino acid residues) linker sequences, can function in the efflux.[92] Recent reviews present the opposing views on the structure of SMR dimers.[84, 93] A study has defined a minimum activity motif of G90LxLIxxGV98 within the fourth transmembrane segment in mediating the SMR protein dimerization.[94]

2.1.5 ABC Transporters

The multidrug transporters of ABC family are conserved from bacteria to humans and export a wide array of substrates, driven byATP hydrolysis. The structure of the S. aureus Sav1866 multidrug exporter has provided insight into ABC transporter-mediated multidrug efflux (Fig. 1).[95] Sav1866 is also a structural homologue of the human MDR P-glycoprotein. The outward-facing conformation of Sav1866 is triggered by ATP binding and reflects the ATP-bound state, with the two nucleotide-binding domains in close contact and the two transmembrane domains forming a central cavity that is presumably the drug translocation pathway (for a review of ABC transporters, see reference[96]). The latter is shielded from the inner leaflet of the lipid bilayer and from the cytoplasm, but exposed to the external medium. The inward-facing conformation is promoted by dissociation of the hydrolysis products ADP and phosphate and shows the substrate-binding site accessible from the cell interior.[95, 97] Similar outward/inward-facing conformations are also shared by the RND transporter AcrB.[98] Interestingly, an alignment of an extended region of the ABC transporter LmrA of L. lactis (a homologue of Sav1866) with a portion of the RND transporter MexB of P. aeruginosa reveals significant similarity.[99]

The newly reanalyzed structure of MsbA (an E. coli lipid flippase) further supports that the inward and outward openings are mediated by two different sets of transmembrane helix interactions and that large ranges of motion may be required for substrate transport.[100] Functional expression of Sav1866 in L. lactis deficient in LmrA and LmrCD transporters has shown that Sav1866 accommodates multiple toxic agents.[101] A truncated LmrA protein lacking the ATP-binding domain mediates a proton-ethidium symport reaction as a secondary-active multidrug uptake system without ATP.[102] In proteoliposome reconstitution studies, LmrA catalyzes Hoechst 33342 transport independent of auxiliary proteins in an ATP-dependent fashion and a transmembrane proton gradient-dependent fashion.[103] These results suggest that the transmembrane ligand transport and the utilization of energy source are sometimes not linked so tightly.

2.2 Membrane Fusion Proteins

The tripartite tansporter complexes also contain MFPs and OM channel proteins, as mentioned above. MFPs such as AcrA, EmrA, and MacA function as adaptor proteins in systems containing RND (AcrAB-TolC), MFS (EmrAB-TolC) and ABC (MacAB-TolC) pumps.[1] The bacterial genome sequences have shown the diversity of the MFPs with identification of many homology-defined clusters (e.g., AcrA/MexA, TriAB, MexH, MacA, EmrA/EmrK, CusB, VexL, and YknX clusters; the last one occurring in Bacillus[104]). The earlier crystal structures of AcrA and MexA showed elongated molecules with three linearly arranged domains: β-barrel, lipoyl and α-helical hairpin domains, but were missing the large domain containing both N- and C-terminus.[105107]. However, recent reanalysis of previous MexA data resulted in the successful modeling of the hitherto missing domain,[108] as shown in Fig. 1. This domain is essential in the assembly and function of the tripartite complex.[109] Crystal structure of MacA of MacAB-TolC indicates a domain orientation of MacA different from that of AcrA with a hexameric MacA observed.[110] Acidic pH induces oligomerization and conformational change of AcrA.[111] A conformational flexibility is evident in the α-helical hairpin domain and may be important in coupling between the MFP conformations and OM channel opening.[107] Molecular dynamics simulation of MexA has been published.[112]

Both AcrA and MexA play a key role in the pump complex assembly.[113118] AcrA drives TolC to fit the transporter complex.[119] Chimeric analysis of AcrA function reveals the importance of its C-terminal domain in its interaction with the AcrB pump.[120] Mutations at both N- and C-terminus of MexA compromise the MexAB-OprM efflux activity, with the N-terminus involved in oligomerization of MexA and/or interaction with OprM and the C-terminus in interaction with the transporter MexB.[115, 118, 121] Construction of the chimeric, functional AcrA-MexB-TolC complex has suggested a certain degree of flexibility in accommodation.[122] In addition, there are paired MFPs as shown with TriAB of P. aeruginosa, which are both essential for TriABC-mediated triclosan resistance.[123]

Although MFPs are often viewed as a mere “glue” in the tripartite complex, they may play a more important role, in that they activate the function of the pump directly. This was first shown in the in vitro assay of an RND pump AcrD, which must interact with AcrA to extrude substrates.[53] A strong AcrA stimulation of the activity of the isolated AcrB pump[124] may have a similar explanation.

2.3 Outer Membrane Channel Proteins

These trimeric proteins, represented by TolC and OprM, functions as channel proteins in the multi-component transporters of various families.[125] TolC of E. coli works together with an exceptionally wide range of transporters, belonging to the RND, MFS, and ABC family,[1] and a tolC mutant was found to be defective in the excretion of endogenous porphyrins.[126] Additional crystal structure of TolC in its partially open state reveals that the opening of the end of the α-helical barrel is accompanied by the exposure of three shallow intraprotomer grooves in the TolC trimer and there is a contact point with the MFP AcrA.[119] The crystal structures of OprM and the Vibrio cholerae VceC are now also available.[127, 128] Like TolC, the OprM channel is trimeric andcomposed of an OM-spanning β-barrel and a periplasmicα-helical barrel, with an overall length of 135Å,[127] a structure consistent with early mutational analysis.[129] In a cross-linking study performed after OprM/OprJ/OprN reconstitution into liposome, either OprM or OprN formed a trimer; but OprJ unexpectedly was reported to form a tetramer.[130]

Interaction of the OM proteins with other efflux components have been supported by genetic and biochemical evidence, e.g., interaction between AcrA-TolC;[113, 131133] AcrA-AcrB-TolC;[113, 134] AcrB-TolC;[135] MexA-OprM[114, 118]and chimerics[117] With the availability of structures of all three components, it is now possible to propose a model of the assembled tripartite structure;[108, 136, 137] the most recent model containing only one MFP for a protomer of RND pump.[108] Interestingly a study of interaction between MexA and OprM with an innovative approach also suggests a 1:1 or 2:1 stoichiometry.[138] The substrates of the transporters further stabilize the efflux pump complex as demonstrated with AcrAB-TolC.[134] Moreover, the amino acid substitutions in the lower α-helical barrel of TolC enabled TolC to function with non-cognate MexAB and to confer MDR.[139] Similarly, while TolC can replace VceC to function with VceAB pump, VceC does not functionally interact with AcrAB. Nevertheless, VceC gain-of-function mutants with the mutations located at the periplasmic tip of VceC have enabled VceC to function with AcrAB.[140] Finally, the TolC homologue, HI1462 of Haemophilus influenzae differs from the E. coli TolC in that it is anion-selective and contains an arginine residue lining the tunnel entrance.[141]

3. Drug Efflux in Gram-Negative Bacteria

Drug efflux is a key mechanism of resistance in Gram-negative bacteria. The major clinically relevant efflux systems belong to the RND efflux systems that are typically composed of a cytoplasmic membrane pump, an MFP and an OM channel protein as described above. Over the past several years, while those previously-studied drug efflux pumps including RND systems have been further characterized, novel efflux systems have also been identified in Gram-negative bacteria (Tables I and II).

Table 1.

Resistance-nodulation-division (RND) family multidrug efflux transporters in Gram-negative bacteria and mycobacteria reported or further characterized since 2003

Species Efflux system componenta
Regulator (Family) Substratesb References
MFP RND OMP
Acinetobacter baumannii AdeI AdeJ AdeK ? AO, BL, CM, EM, FQ, FU, LC, NO, PY, RF, SA, SDS, TC, TM [286, 296]
Acinetobacter genospecies 3 AdeD AdeF ? ? AM, CM, CP, CT, EB, EM, MP, RF, TC [287, 296]
AdeX AdeY AdeZ ? ? [288, 296]
Acinetobacter genospecies UT13 Unknown Unknown AdeC ? ? [733]
Aeromonas hydrophilia AheA AheB AheC AheR (TetR) BC, BL, EM, FU, LC, PR, TC, TM, TT [223]
Bacteroides fragilis BmeA1-16 BmeB1-16 BmeC1-16 Varied, e.g., BmeR5 (TetR) BL, CP, EB, MD, SDS [384, 389, 390]
Brucella suis BepD BepE BepC BepR (TetR) AM, BL, CAB, CV, DC, EB, EM, FQ, SDS, TC [349, 350]
BepF BepG BepC ? DC, NA [350]
Burkholderia cenocepacia CeoA CeoB OpcM CeoR (LysR) CM, FQ, TMP [325]
? ORF2 ? ? EB, FQ, SM, TPP [324]
Burkholderia glumae ToxG ToxH ToxI ToxR (LysR) TF [610]
Burkholderia pseudomallei BpeA BpeB OprB BpeR (TetR) AC, AG, EM [326, 327]
BpeE BpeF OprC BpeT (LysR) CM, TM [328]
Campylobacter jejuni CmeD CmeE CmeF ? AO, AP, CAB, EB, PM, SDS, TR [371, 373]
Chromohalobacter spp. ? ? HrdC ? BL, CM, EB, OS, TC [734]
Enterobacter aerogenes EefA EefB EefC ? CM, CP, EM, TC [182, 735]
Enterobacter cloacae AcrA AcrB TolC AcrR (TetR), RamA (MarA) AC, AG, BL, CL, CM, CP, CV, DC, EM, LC, LZ, SDS, SXT, TC, TG [736]
Erwinia amylovora AcrA AcrB ? AcrR (TetR) BB, CV, EB, MB, PH, SDS [596]
Escherichia coli OqxA (plasmidborne) OqxB (plasmidborne) ? ORF68 (plasmidborne) CM, CP, EB, OQ [154156]
Haemophilus influenzae AcrA AcrB ? AcrR (TetR) AP [321, 737]
Helicobacter pylori CznB CznA CznC ? MS [382]
HefB HefC HefA ? CM, CR, CX, EB, GM, ML, NO, TC [378, 380, 738]
Klebsiella pneumoniae AcrA AcrB AcrR AcrR (TetR), RamA (MarA) AC, CM, EB, EM, NA, NF, NO, TC, TG [187, 190, 739]
EefA EefB EefC ? Inorganic acid [191]
Morganella morganii AcrA AcrB ? AcrR (TetR) AC, CM, EB, EM, NA, NO, SDS, TC, TG, TM [269]
Mycobacterium tuberculosis ? MmpL7 ? ? INH [518]
Neisseria gonorrhoeae FarA FarB MtrE FarR (MarR) FA [740]
Pasteurella multocida ? ? PM0527 ? AO, CT, CV, EB, EM, LC, NO, RF, SDS, TM [741]
? ? PM1980 ? CT, RF, VC. [741]
Proteus mirabilis AcrA AcrB TolC ? AC, AP, CM, CP, MI, SAM, SDS, TC, TG, TM [267]
Pseudomonas aeruginosa MexA MexB OprM AmrR, MexR (MarR), NalC (TetR), NalD (TetR) AC, AO, AG, BL, MC, FQ, NO, OS, SF, TC, TG, TR [1, 228, 623, 624, 626]
MexM MexN OprM ? CM, TP [264]
MexP MexQ OpmM ? MC, FQ [264]
MexV MexW OprM ? AC, CM, EB, EM, FQ, TC [262]
TriAB TriC OpmH ? TR [123]
Pseudomonas fluorescens EmhA EmhB EmhC ? Polycyclic aromatic hydrocarbons, CM, NA [275, 276]
Pseudomonas stutzeri TbtA TbtB TbtM ? CM, NA, OS, SF, TT [278]
Pseudomonas syringae MexA MexB OprM PmeR (TetR) AC, AG, AO, BB, BC, CM, CV, DA, EB, EM, FQ, FU, NA, NT, RG, TC, TM, TPP [684]
PseB PseC PseA GacS/GacA AC, EM, TC [590]
Ralstonia solanacearum AcrA AcrB ? AcrR (TetR) AC, AP, BB, EB [595]
Salmonella Typhimurium MdtA MdtBC ? ? DC, NO, SDS [172]
MsdA MsdB MsdC, TolC ? DC, NO, SDS [172]
Serratia marcescens SdeA SdeB HasF SdeR (MarA) CM, EB, FQ, OS, SDS [198, 199]
SdeC SdeDE ? ? NO [198, 742]
SdeX SdeY ? ? AC, BAC, EM, NF, RG, TC [197]
Serratia spp. ZrpAD ZrpB ZrpC? PigZ (TetR) ? [206]
Stenotrophomonas maltophilia SmeG SmeH ? Smlt3169 (TetR) ? [306]
SmeI SmeJK ? ? AG, CP, TC [306]
SmeM SmeN ? ? ? [306]
SmeO SmeP ? Smlt3926 (TetR) ? [306]
SmeV SmeW SmeX Smlt1827 (LysR) ? [306]
SmeY SmeZ ? Smlt2199-2130 AG [306]
Vibrio cholerae VexA VexB ? VexR (TetR) MDR [207209]
VexC/BreA VexD/BreB ? VexR/BreR (TetR) BS [207209]
VexE VexF TolC ? BC, BS, DC, EB, EM, NF, NO, SDS, TC, TM [209]
Vibrio parahaemolyticus VmeA (VP1091) VmeB (VP1092) VpoC (VP0425) ? AC, BL, CP, CV, DC, EB, EM, NF, RG, SDS, TC, TM [743]
a

MFP=membrane fusion protein; OMP=outer membrane protein; RND=resistance-nodulation-division.

b

AC=acriflavine; AG=aminoglycosides; AM=amikacin; AO=acridine orange; AP=ampicillin; BC=benzalkonium chloride; BL=β-lactams; BS=bile salts; CAB=cetyltrimethylammonium bromide; CL= cholate; CM=chloramphenicol; CP=ciprofloxacin; CR=clarithromycin; CT=ceftazidime; CV=crystal violet; CX=cefotaxime; DA=daunorubicin, DC=deoxycholate; EB=ethidium bromide; EM=erythromycin; FA=fatty acids; FQ=fluoroquinolones; FU=fusidic acid; GM=gentamicin; INH=isoniazid; LC=lincosamides; LZ=linezolid; MB=methylene blue; MC=macrolides; MD=metronidazole; MI= minocycline; MP=meropenem; MS=metal salts; NA=nalidixic acid; NF=norfloxacin; NO=novobiocin; NT=nitrofurantoin OQ=olaquindox; OS=organic solvents; PH=Phloretin; PM=polymyxin B; PR=pristinamycin; PY=pyronine; RF=rifampin; RG=rhodamine 6G; SAM= ampicillin-sulbactam; SDS=sodium dodecyl sulphate; SA=safranin; SF=sulfonamides; SM=streptomycin; SXT=trimethoprim sulfamethoxazole; TC=tetracyclines; TF=toxoflavin; TG=tigecycline; TM=trimethoprim; TP=thiamphenicol; TPP=tetraphenylphosphonium; TR=triclosan; TT=tributyltin; VC=vancomycin; ?=the efflux components or regulators remain unknown or no genes linked to the transporter structural gene(s) are identified.

Table 2.

MFS, MATE, SMR and ABC multidrug efflux transporters in bacteria reported or further characterized since 2003

Transporter family/Organism Efflux pump Regulator (Family) Substratesa References
Major Facilitator Superfamily (MFS)
Acinetobacter baumannii SmvA (A1S_2057) ? EM, MV, QAC [299]
Bacillus subtilis Bmr3 ? FQ, PU [403, 744]
LmrB LmrA (TetR) DR, FQ, LC, PU [404, 405]
MdtP MdtR (MarR) AT, FU, NO, SM [408]
Bordetella bronchiseptica CmlB1 ? CM [745]
Clostridium difficile Cme ? EB, EM, SA [399]
Clostridium saccharolyticum Tet(40) ? TC [402]
Enterobacter aerogenes QepA (plasmidborne) ? FQ [184]
Enterococcus faecium EfmA ? DP, FQ, TPP [470]
Escherichia coli Mef(B) (plasmidborne) ? ML [533]
QepA, QepA2 (plasmidborne) ? FQ [23, 160, 161]
Helicobacter pylori Hp1181 ? Unknown [746]
Klebsiella pneumoniae KmrA (Ec) ? AC, DP, EB, HO, MV, TPP [193]
Listeria monocytogenes Lde ? AC, BC, EB, FQ [421, 424, 425]
MdrL LadR Unknown [422]
MdrM MarR Unknown [423]
MdrT TetR Unknown [423]
Mycobacterium smegmatis LfrA LfrR (TetR) AC, EB, FQ [514, 528]
Salmonella Typhimurium EmrAB ? DC, NA, NO [172]
MdfA ? CM, DR, NF, TC [172]
SmvA-OmpW ? MV [747]
Serratia marcescens SmfY ? AC, BC, DP, EB, NF [202]
Staphylococcus aureus MdeA ? BC, DQ, EB, FU, HO, MU, NO, QAC, TPP, VM [428, 748]
NorB MgrA (MarR), NorG (GntR) CT, EB, FQ [432434, 436, 749]
NorC MgrA (MarR) FQ [432]
SdrM (Ec) ? AC, EB, NF [431]
Tet38 MgrA (MarR) TC [433, 436]
Staphylococcus lentus FexA (plasmidborne) ? CM, FP [438]
Stenotrophomonas maltophilia Smlt0032 ? MC [306]
Smlt1528-1529-1530 ? Unknown [306]
Streptococcus agalactiae MefB, MefG ? MC [750]
Tet42 TetR TC [751]
Streptococcus suis SmrA ? FQ [752]
Vibro cholerae VceCAB VceR (TetR) CCCP, DC, NA, PA, PC [1, 210]
Xanthomonas albilineans AlbF ? AB [753]
Multidrug and Toxic Compound Extrusion (MATE) Family
Acinetobacter baumannii AbeM ? AC, AG, DN, DR, FQ, HO, RG [298]
Brucella melitensis NorMI ? AC, BB, FQ, GM, TPP [348]
Clostridium difficile CdeA ? AC, EB [398]
Erwinia amylovora NorM ? AP, BB, EB, CV, FQ, KM, MB, PH [754]
Haemophilus influenzae HmrM ? AC, BB, DC, DN, DP, DR, EB, HO, TPP [322]
Neisseria gonorrhoeae NorM ? CC [339]
Neisseria meningitidis NorM ? CC [339]
Pseudomonas aeruginosa PmpM ? AC, BC, EB, TPP [263]
Ralstonia solanacearum DinF ? AC, AP, BB, EB, TPP [595]
Salmonella Typhimurium MdtK ? AC, DR, NF [172]
Staphylococcus aureus MepA MepR (MarR) CT, EB, FQ, MDB, TG [429, 430, 666]
Vibrio cholerae NorM ? EB, FQ [217]
VcmB, VcmD, VcmH, VcmN ? AG, EB, FQ, HO [215]
VcrM ? AC, DP, EB, HO, RG, TPP [213]
Vibrio parahaemolyticus VmrA ? AC, DP, EB, TPP [755]
Small Multidrug Resistance (SMR) Family
Acinetobacter baumannii Smr (A1S_0710) ? DC, SDS [299]
Escherichia coli MdtJI DC, SDS, SP [153]
Serratia marcescens SsmE ? AC, EB, NF [203]
Staphylococcus aureus SepA ? AC, BC, CH [435]
ATP-Binding Cassette (ABC) Superfamily
Bacillus subtilis YtsCD YtsA BA [756]
YvcC (BmrA) ? AA, DR, HO [409]
Bifidobacterium breve AbcAB ? NI, PM [507]
Enterococcus faecalis EfrAB ? AC, DA, DP, DR, FQ, TC, TPP [463]
Enterococcus faecium MsrC ? MC, QP [467, 468]
Escherichia coli YojI Lrp MJ [150, 151]
Lactococcus lactis LmrCD LmrR (PadR) CL, DN, EB, HO, RG [451454]
Mycobacterium bovis BCG Bcg0231 ? AP, CM, SM, VC [757]
Mycobacterium tuberculosis Rv0194 ? AP, EM, NO, VC [757]
Rv1258C (Tap) ? FQ, RF, TC [1, 523]
Rv2686c- ? FQ [521]
Rv2687c-
Rv2688c
Neisseria gonorrhoeae MacAB ? ML [338]
Oenococus oeni OmrA ? ML, SL [758, 759]
Salmonella Typhimurium MacAB ? EM [172]
Serratia marcescens SmdAB ? DP, HO, NF, TC [204]
Staphylococcus aureus AbcA MgrA (MarR), NorG (GntR) BL [434, 760]
Sav1866 ? EB, HO, TPP [101]
Stenotrophomonas maltophilia Smlt1537-1538-1539 ? MC [306]
Smlt2642-2643 ? MC [306]
Streptococcus pneumoniae PatA, PatB ? FQ [483486]
SP2073/SP2075 ? AC, EB, FQ, NO [487]
Spr0812/Spr0813 BA [488]
Vibrio chlolera VcaM ? DN, DP, DR, FQ, HO, TC [214]
a

AA=7-aminoactinomycin D; AB=albicidin; AP=ampicillin; AT=actinomycin D; BA=bacitracin; BB=berberine; BC=benzalkonium chloride; CC=cationic compounds; CCCP=carbonyl cyanide m-chlorphenylhrazone; CH= chlorhexidine; CL= cholate; CM=chloramphenicol; CT=cetrimide; DA=daunorubicin, DN=Daunomycin; DP=4′,6-diamidino-2-phenylindole; DQ=dequalinium chloride; DR=doxorubicin; EB=ethidium bromide; EM=erythromycin; FP=florphenicol; FQ=fluoroquinolones; FU=fusidic acid; GM=gentamicin; HO=Hoechst 33342; KM=kanamycin; LC=lincosamides; MB=methylene blue; MC=macrolides; MDB=monovalent and divalent biocides; MJ=microcin J25; ML=metal salts; MU=mupirocin; MV=methyl viologen; NF=norfloxacin; NI=nisin; NO=novobiocin; PA=phenymercuric acetate; PC=pentachorophenol; PH=Phloretin; PM=polymyxin B; PU=puromycin; QAC=quaternary ammonium compounds; QP=quinupristin; RF=rifampin; SA=safranin; SL=Sodium laureate; SM=streptomycin; SP=spermidine; TC=tetracyclines; TPP=tetraphenylphosphonium; VC=vancomycin; VM=virginiamycin; VR=vancoresmycin; ?=the regulators remain unknown or no regulator genes linked to the transporter structural gene(s) are identified.

3.1 Gammaproteobacteria (Enterobacteriales, Vibrio, Aeromonas, Pseudomonas, Acinetobacter, Stenotrophomonas and Haemophilus)

3.1.1 Enterobacteriales

This large order (for the taxonomy followed in this article, see reference[142]) contains many genera that are important in human health.

Escherichia coli

Drug efflux systems in E. coli have been used as models for genetic and biochemical studies.[1] Of clinical relevance, efflux (presumably due to AcrAB) was shown to contribute to cefuroxime[143] and fluoroquinolone[144] resistance of strains from patients. The postantibiotic effect of multiple antibacterials was prolonged in an acrAB mutant.[145] Tigecycline was found to be a substrate for AcrAB and AcrEF pumps,[146, 147] in spite of its activity against tetracycline-resistant strains carrying plasmid-borne specific efflux genes, tet(B), tet(C), or tet(K). Moreover, tigecyline interacts with TetR repressor and induces the expression of Tet(B).[146] Mureidomycin A and C and simocyclinone D8 (an angucyclinone antibiotic) are also substrates for AcrAB-TolC.[148, 149] YojI, an ABC exporter, functions with TolC and mediates resistance to microcin J25[150] and its expression is modulated by the leucine-responsive regulatory protein (Lrp).[151] Preincubation of FloR pump-producing florfenicol-resistant strains with anti-FloR antibody, in the presence of lysozyme and ethylenediaminotetraacetic acid, increased the intracellular accumulation of florfenicol.[152] A paired SMR pump, MdtJI, exports spermidine.[153]

Novel plasmid-encoded drug pumps have also been identified in E. coli. OqxAB, a RND pump, mediates resistance to olaquindox (a growth promoter in pigs) and several other agents (Table I).[154156] Its function is dependent on TolC.[154] A plasmid-encoded fluoroquinolone resistance protein, QepA, is a 14-transmembrane-segment MFS transporter and causes a decreased accumulation of norfloxacin.[23] The qepA-harbouring isolates were further identified in clinical strains in Japan[157] as well as in isolates derived from companion and food-producing animals in China (with co-presence of Qnr, AAC(6′)-Ib-cr and the aminoglycoside resistance 16S rRNA methylase RmtB). [158, 159] Isolates harbouring large mobilizable plasmids, encoding QepA/QepA2 and CTX-M-15 β-lactamase, were recently recovered from clinical isolates from France and Canada.[160, 161]

Salmonella enterica spp

Infections associated with either nontyphoid or typhoid Salmonella are of global health concern and are complicated by the increasing prevalence of acquired MDR. Salmonellae possess multiple drug efflux systems including the AcrAB-TolC system.[1, 15, 162] An MDR isolate of S. Typhimurium derived from a patient treated with ciprofloxacin was an AcrAB overproducer.[163] Among 388 Salmonella of 35 serovars from animal and human origins, ca. 10% of the isolates were resistant to cyclohexane, a phenotype usually associated with AcrAB overexpression.[164] Clonal expansion among human and poultry isolates of quinolone-resistant S. Virchow probably emerged from a parental clone overproducing AcrAB.[165] Laboratory-selected and naturally occurring fluoroquinolone-resistant S. Typhimurium strains showed increased expressions of acrA, acrB, acrE, acrF, emrB, emrD, and mdlB as well as, to a lesser extent, of mdtB, mdtC, and emrA.[165] A complementary result is that ciprofloxacin-resistant S. Typhimurium mutants are difficult to select in the absence of AcrB and TolC.[166] In S. Typhimurium DT204 overexpression of acrAB plays a dominant role in fluoroquinolone resistance, and selection of fluoroquinolone resistant mutant in an acrB background resulted in the isolation of strains overexpressing acrEF through insertion of IS1 or IS10 elements.[167] AcrD and MdtABC pumps are also involved in metal resistance.[168] As described above, an S. Waycross isolate possesses a plasmid containing class 1 integron and MDR genes including the efflux pump gene qacG.[25] Efflux is a major mechanism for the adaptive resistance to erythromycin, benzalkonium chloride and triclosan in Salmonella spp.[169]

The TolC component is required for AcrAB to function.[1] Strains with tolC inactivation exhibited hypersusceptiblity to several antibacterials.[15, 170] Interestingly, TolC is required for the colonization of MDR S. Typhimurium in chick although AcrAB is not.[171] This may be partly because TolC is the OM component of many other efflux pumps, including the Salmonella-specific RND pump MsdAB.[172] However, another study showed that both tolC an acrB mutants colonized poorly and did not persist in the avian gut.[173] This is perhaps due to the impact of the AcrAB-TolC disruption on reduced expression of certain pathogenesis genes.[174]

There are also drug-specific pumps such as tetracycline-specific Tet pumps and phenicol-specific FloR.[1, 15] Other multidrug pumps, such as EmrAB, MdfA and MdtK, were also identified in S. Typhimurium (Table II).[172].

Enterobacter spp

Enterobacter aerogenes, a member of Enterobacteriales like E. coli, has emerged as an important nosocomial pathogen. Early studies revealed that the MDR clinical strains had a drastic porin reduction, altered O-polysaccharide, and active efflux of chloramphenicol.[175] AcrAB-TolC, a major RND pump from E. aerogenes,[1] mediated resistance to erythromycin and clarithromycin but not to telithromycin.[176] Chloramphenicol- or imipenem-selected resistant mutants displayed elevated AcrAB expression that was also associated with resistance to quinolones and tetracyclines.[177, 178] Tigecycline resistance was due to RamA-mediated overexpression of AcrAB.[179] Elevated MarA expression was triggered by certain antibiotics and phenolic compounds as well as by RamA activator[180] and was observed in imipenem-resistant isolates.[177] AcrAB-TolC was inhibited by chloroquinoline derivatives.[181] EefABC encodes another cryptic RND pump whose expression from multicopy plasmids conferred MDR.[182] Chloramphenicol-resistant mutants isolated in the laboratory showed detectable production of EefABC and showed resistance to erythromycin and ticarcillin, but not to fluoroquinolones, ketolides and detergents.[183] Novel resistance plasmids were also isolated with the co-presence of qepA, qnrS, rmtB and blaLAP-1 (for a Class A β-lactamase).[184]

An ertapenem-resistant isolate of Enterobacter cloacae exhibited reduction of ompD and ompF transcripts, and the inhibitory levels of multiple antibacterials for this isolate decreased in the presence of the efflux pump inhibitor (EPI) Phe-Arg-β-naphthylamide, suggesting the involvement of efflux mechanism.[185] Enterobacter gergoviae isolates from cosmetic formulations containing parabens showed high methylparaben inhibitory concentrations; the expression of a Phe-Arg-β-naphthylamide-sensitiveparaben efflux mechanism was responsible for the observed resistance, although there was no cross-resistance to other antibacterials.[186]

Klebsiella spp

Overexpression of AcrAB homologues, in some cases through AcrR mutations or RamA overexpression, have been observed in multidrug- or fluoroquinolone-resistant clinical isolates of Klebsiella pneumoniae and Klebsiella oxytoca.[187] Efflux also plays a key role in β-lactam resistance in clinical isolates.[188] The multidrug efflux was inhibited by alkoxyquinoline derivatives.[189] Decreased susceptibility to tigecycline in K. pneumoniae is also a result of RamA-activated AcrAB overexpression.[190] An RND pump, EefABC (see above), is involved in gastrointestinal colonization by K. pneumoniae and confers a tolerance response to inorganic and organic acids (Table I). EefA inactivation did not alter the susceptibility to bile salts, other detergents and antibiotics.[191] KmrA, an MFS transporter, confers resistance to multiple toxic agents when expressed from a low copy number plasmid (Table II).[192, 193] Over the past decade, qnr-containing plasmids have become widespread among fluoroquinolone-resistant bacteria including Klebsiella.[21, 22] This resistance mechanism interplays positively with the efflux pumps in producing clinically relevant resistance.[194] The K. oxytoca TolC protein lacks six residues around the region of the residues 280–290 of E. coli TolC that forms part of the loop exposed to the external side of the OM and the absence of these residues is involved in resistance to colicins.[195]

Serratia spp

Serratia marcescens is naturally resistant to multiple antibacterials.[196] Three RND pumps, SdeAB-HasF, SdeCDE and SdeXY, have been identified to date (Table I).[1, 197201] HasF is a TolC homologue.[200] Intriguingly, SdeCDE requires the paired RND pump components, SdeDE,[198] similar to MdtBC of E. coli. When expressed from plasmids, three additional pumps, SmdAB (a heterodimeric ABC-type), SmfY (an MFS-type) and SsmE (an SMR-type), confer, to a multidrug-susceptible E. coli, resistance to several structurally unrelated antibacterials (Table II).[202204] TetA(41) efflux protein and TetR(41) repressor were observed in an environmental strain of S. marcescens.[205] A TetR/AcrR family repressor PizR, which was identified because of its effect on the production of secondary metabolites (prodigiosin and carbapenem), is a specific repressor of a four-component RND efflux pump, ZrpADBC.[206] The overproduction of this pump may cause the removal of intracellular metabolites, resulting in lowered transcription of genes involved in secondary metabolism.

3.1.2 Vibrio spp

In contrast to the groups so far discussed, which belong to Enterobacteriales, Vibrio spp. belong to another order, Vibrionales. There are six operons for putative RND-type efflux transporters in the chromosome of V. cholerae O1, the causative pathogen for cholera. Two bile-regulated RND systems, VexAB and VexCD, are involved in bile resistance.[207] VexAB causes resistance to multiple antibacterials including bile acids, whereas deletion of VexCD, also known as BreAB, does not cause any change in the sensitivity to antibacterials, including bile salts. However, the simultaneous absence of VexB and VexD dramatically lowers the resistance to bile salts, and only to bile salts.[207] Further analysis regarding vexAB and breAB expression established that vexAB was induced in the presence of bile, novobiocin or sodium dodecylsulphate, whereas induction of breAB was specific to bile. BreR is a direct repressor of the breAB promoter and is able to autoregulate its own expression. Expression of breR and breAB is induced by the bile salts, which appear to abolish the complex formation between the repressor BreR and breAB and breR promoters.[208] In another study, all of the six RND operons were cloned from V. cholerae non-O1 and expressed in efflux-deficient E. coli; VexAB produced resistance to dyes and some resistance to deoxycholate, whereas VexEF conferred resistance to antibiotics but not to bile salts. Ethidium efflux activity via VexEF-TolC requires Na+ (Table I).[209]

The OM component, VceC coded in the vceCAB operon, is required for the function of the MFS-type VceAB pump.[210] The native VceC does not function in replacing TolC of the E. coli AcrAB-TolC, but the gain of function mutant of VceC with the amino acid substitutions located at the periplasmic tip has been isolated.[140] VceR repressor regulates vceCAB expression by alternating between mutually exclusive conformations[211] but positively regulates its own synthesis.[212]

Several MATE-type pumps and an ABC-type pump have also been characterized from Vibrio spp. (Table II).[213216] Mutational analysis of NorM identified functionally important residues that are mostly located in periplasmic loops.[217] Efflux plays a major role in quinolone resistance in Vibrio fluvialis.[218]

3.1.3 Aeromonas spp

The ubiquitous waterborne species (e.g., Aeromonas hydrophila and Aeromonas veronii) belong to yet another order, Aeromonales in Gammaproteobacteria. They cause intestinal infections in normal adults or children, as well as extraintestinal infections in immunocompromised hosts. There is an increasing resistance trend in this group.[219] Moreover, Aeromonas strains may serve as reservoirs for dissemination and transfer of resistance among humans, animals, plants and natural soil and water, because mobile resistance gene cassesstes are often found.[220]

Drug efflux contributes to resistance in Aeromonas.[219] The genomes of A. hydrophila and A. salmonicida show an abundance of transporters comparable to those of pseudomonads and vibrios, with the presence of putative drug efflux systems in A. hydrophilia including 10 RND transporters such as AheB,[221] and in A. salmonicida including RND exporters (AcrAB, MexF, and MexW), MFS pumps (MdtH and EmrD), MATE (NorM), SMR (EmrE) and ABC (there are several MacAB homologues) (for details see Antibiotic Resistance Gene Database: http://ardb.cbcb.umd.edu/cgi/ssquery.cgi?db=O&gn=a&sp=382245).[220] Up-regulation of RND genes occurs in the presence of erythromycin.[222] An RND system, AheABC, is responsible for intrinsic resistance, and extrudes 13 antibacterial substrates out of the 63 agents tested (Table I).[223]

The tetA(E) gene encoding a tetracycline efflux protein was observed in Aeromonas two decades ago,[224] and its prevalence, often associated with large plasmids, was recently reconfirmed in Aeromonas spp. derived from fish farms.[225] Quinolone-resistant A. salmonicida strains isolated from diseased fish not only carried mutations in the target genes but also indicated an important contribution of efflux.[226]

3.1.4 Pseudomonas spp. and Acinetobacter spp

The large order Pseudomonadales contain mostly soil bacteria, which often show MDR phenotype because their OM has an exceptionally low permeability owing to the mostly closed porin channels.[227] This makes the tripartite efflux systems, which work in synergy with the OM barrier,[1] exceedingly efficient. Pan-resistance in Gram-negative bacteria often occurs among this group.[13]

Pseudomonas aeruginosa and its relatives

Multidrug efflux systems in P. aeruginosa, particularly the RND-type Mex pumps, have been extensively investigated since their discovery in early 1990s.[1, 6, 228] Studies with clinical isolates including epidemic clones support the established role of the drug efflux pumps in MDR.[229234] Thus, efflux mechanisms are considered as a key factor in optimizing the treatment of P. aeruginosa infections.[235237] Meanwhile, approaches for detection of overexpressed Mex efflux systems are in development.[238]

Overexpression of MexAB-OprM and MexXY-OprM occurred, respectively, in 11% and 35% of 120 bacteraemic isolates from France, suggesting enhanced expression of the efflux systems without causing the loss of ability to cause severe bloodstream infections.[239] Isolates can also simultaneously overproduce multiple drug pumps and broaden the resistance profiles.[240] Carbapenem resistance was due to non-enzymatic mechanisms, active efflux and the OprD deficiency, in a large number of isolates from Bulgaria.[241] Efflux-type resistant mutants with broad cross-resistance were selected in vitro by ertrapenem, a carbapenem which contains a side-chain with an aminobenzoate moiety and is used increasingly against the community-acquired pathogens (although not active against P. aeruginosa).[242] Fosfomycin, with no lipophilic surface, is predictably a poor substrate for most Mex systems.[243] Tolerance of P. aeruginosa to tea tree oil is also associated with the OM barrier and efflux pumps.[244]

MexXY-OprM is necessary for adaptive resistance to aminoglycosides[245] and is overproduced in amikacin-resistant or MDR isolates including those producing PER-1 β-lactamase.[229, 246250] It also plays a key role in resistance to a fourth generation cephalosporin, cefepime.[251] A Phe1018Leu change in MexY, which increased MDR, was identified in isolates from cystic fibrosis patients, suggesting the need of MexXY in the hostile environment of cystic fibrosis lung.[252] Transposon mutagenesis showed that aminoglycoside resistance can be generated by the inactivation of the repressor mexZ (causing increased MexXY-OprM-mediated efflux), but also by the inactivation of other genes galU, nuoG and rplY, which respectively, may have produced unstable OM, reduced drug influx, and alteration of the target of aminoglycosides.[253] Gene disruption study indicates that two OM proteins, OpmG and OpmI, also function in MexXY-mediated aminoglycoside efflux.[254]

MDR pumps also impair the in vivo efficacy of fluoroquinolones or aminoglycosides in therapy of P. aeruginosa infections.[255257] One study using an animal model concluded that MexAB-OprM had insignificant impact on drug efficacy,[258] but it utilized meropenem and cefepime, weak substrates for this system. The combination of levofloxacin and imipenem prevented the emergence of high-level resistance in strains already lacking susceptibility to one or both drugs due to the Mex pump overexpression and OprD deficiency.[259]

MexJK requires OprM for erythromycin efflux, while a TolC homologue, OpmH, functions with MexJK for triclosan efflux, suggesting the preference among multiple OM proteins by an RND transporter and/or an MFP.[260, 261] Novel RND pumps (MexMN-OprM, MexPQ-OpmE, MexVW-OprM and TriABC-OpmH) and an MATE-type PmpM pump have been characterized from P. aeruginosa (Tables I and II) and TriABC-OpmH requires two MFP components, TriAB, for its function.[123, 262264] An increased expression of a probable ATP-binding component of an ABC transporter was observed in a ciprofloxacin-resistant strain.[265]

Several new antibacterials such as ceftobiprole (a fifth generation cephalosporin), doripenem and tigecycline are substrates for RND pumps.[266270] A hydrophobic indole derivative that inhibits P. aeruginosa growth by targeting MreB (a prokaryotic actin homologue) is, predictably, a substrate for MexAB-OprM.[271] As predicted from the synergy between the pumps and the OM barrier, an OM-permeabilizing polycationic compound 48/80[272] increased the susceptibility of P. aeruginosa to the hydrophobic biocide triclosan.[273]

Drug efflux systems have also been characterized from other Pseudomonas species. Pseudomonas putida DOT-T1E withstands solvents predominantly because it removes solvents from within the membrane interior by using three RND systems, TtgABC, TtgDEF, and TtgGHI.[1] Among them TtgABC plays a major role in the intrinsic antibiotic resistance.[274] The RND system EmhABC from Pseudomonas fluorescens accommodates nontoxic, highly hydrophobic polycyclic aromatic hydrocarbons and antibacterials.[275] Mutational analysis of EmhB suggested that the central cavity and periplasmic domains play an important role in the efflux function.[276] High-level benzalkonium chloride resistance in P. fluorescens is also attributable to efflux.[277] Pseudomonas stutzeri contains TbtABM pump associated with tributyltin resistance.[278] A putative ABC transporter PltHIJKN is required for the export of pyoluteorin, an amphiphilic antibiotic with a resorcinol linked to dichlorinated pyrrole via a ketone bridge, in Pseudomonas sp. M18, and can also confer resistance to pyoluteorin when expressed in E. coli.[279]

Acinetobacter spp

Clinically relevant Acinetobacter spp. are often related to Acinetobacter baumannii-Acinetobacter calcoaceticus complex.[280] Particularly A. baumannii and relatives have emerged as common nosocomial pathogens worldwide with high-levels of MDR increasingly observed.[281, 282] This is not so surprising, as Acinetobacter, as a member of Pseudomonadales, produces an OM of exceptionally low permeability.[283] Its OM appears to lack a trimeric porin found in Enterobacteriales, and to contain an OmpA/OprF homologue as the major porin.[284] With such an OM, efflux becomes extremely efficient in building resistance.[285] Additionally, MDR strains of Acinetobacter often contain individual genetic determinants that mediate resistance to β-lactams, aminoglycosides and fluoroquinolones.[18, 280]

Indeed RND efflux systems have been reported in Acinetobacter spp. (Table I). AdeIJK, identified in susceptible and resistant A. baumannii,[16] is likely responsible only for intrinsic resistance.[286] AdeDE (with an unidentified OM component) and AdeXYZ were reported in Acinetobacter genospecies 3.[287, 288] The comparative genomics of MDR in A. baumannii shows that adeABC genes are only present in the MDR isolate analyzed, but not in a susceptible strain. Intriguingly, the MDR isolate carries an 86 kb genomic resistance island containing an integrase gene and 45 individual resistance genes, including several uncharacterized RND systems.[16] The significance of the resistance island in MDR is further demonstrated with additional genome analysis.[17] AdeABC is regulated by the AdeRS two-component regulatory system[289] and also contributes to resistance to netilmicin and tigecycline.[290292] A 25-fold increase in adeB expression was observed in a tigecycline-resistant mutant.[292] Sequence analysis of an 850-bp fragment internal to adeB revealed many sequence types, suggesting the possibility of sequence-based adeB typing.[293] Efflux pump overproducers have been observed with resistant Acinetobacter including those from hospital outbreaks.[294] Tigecycline-non-susceptible A. baumannii that caused bloodstream infection was partly attributable to an efflux mechanism.[295] Distribution of AdeABC and AdeIJK was associated with the presence of class 1 integron.[296] An AdeT-associated putative RND pump involved in aminoglycoside resistance was also revealed via a comparison of the membrane subproteomes.[297] The pumps belonging to the MFS, SMR and MATE were also reported from A. baumannii (Table II).[298, 299] Transposon mutagenesis of A. baylyi identified a few genes encoding efflux proteins, AcrB or OprM homologues, responsible for intrinsic resistance.[300]

Drug-specific pumps such as tetracycline pumps Tet(A) or Tet(B) have also been found in MDR Acinetobacter[301, 302] which also contain AdeABC. Tet(A) may coexist with Tet(M) that is for ribosomal protection.[302, 303] Given that Tet(A) confers resistance to tetracycline butnot to minocycline, minocycline may be one of limited drugs to which some MDR Acinetobacter may still be susceptible.[296, 302]

3.1.5 Stenotrophomonas maltophilia

This species, found in various environments, used to be considered as a relative of Pseudomonas, but is now known to belong to a separate order, Xanthomonadales. It increasingly causes human infections that are difficult to treat, particularly due to the MDR phenotypes attributed to efflux mechanisms.[1, 304] Early studies indicate growth temperature-dependent variation of cell envelope lipids and proteins as well as antibiotic susceptibilities.[305] S. maltophilia contains a homologue of P. aeruginosa OprF, whereas no homologue of the trimeric, open porin of Enterobacteriales can be found. Thus OM permeability is probably low, and synergy with the tripartite drug efflux is expected to be efficient. In addition the S. maltophilia genome reveals a number of drug resistance determinants as well as potentially mobile genetic regions. Indeed 8 putative RND-type efflux systems are present and include the previously identified SmeABC and SmeDEF[1] as well as the new SmeGH, SmeIJK, SmeMN, SmeOP, SmeVWX, and SmeYZ (Table I).[306] Some of these putative RND pumps do not have an identified OM component, but proteins such as SmeC[307] might also be used by these other systems. Additional ABC-type and MFS-type transporters are shown in Table II.

SmeABC and SmeDEF contribute to MDR in clinical isolates.[308] The biocide triclosan can also select SmeDEF-overproducing mutants.[309] Mutations in the SmeT repressor can result in SmeDEF overproduction. Yet the drug resistance pattern is not completely reproducible among SmeDEF overproducers, and the contribution of so far unidentified drug efflux systems is suspected.[310, 311] The EPI Phe-Arg-β-naphthylamide does not affect the SmeDEF efflux activity.[312] Highly effective efflux mechanism is also suggested to preserve topoisomerase targets in S. maltophilia challenged by ciprofloxacin.[313]

3.1.6 Haemophilus influenzae

This organism belongs to the order Pasteurellales. The majority of H. influenzae strains were found to have a macrolide efflux mechanism.[314] Telithromycin and azithromycinefflux was further demonstrated in various clinical strains.[315] In fact, efflux is required for ribosomal protein mutations to produce high-level macrolide resistance.[316] Ciprofloxacin-nonsusceptible respiratory isolates were more hypermutable than the susceptible group and the hypermutability appeared to result in the stepwise accumulation of resistance mechanisms, including target modifications, loss of a porin protein, and increased efflux.[317]

The previously described AcrAB-TolC pump of H. influenzae [1, 318]also accommodates the peptide deformylase inhibitor LBM415, a lipophilic compound, which selected AcrR mutants.[319, 320] Overproduction of AcrAB (due to AcrR mutations) and alterations in penicillin-binding protein 3 were observed in β-lactamase-negative, high-level ampicillin-resistant H. influenzae of diverse geographical sources.[321] Single cysteine mutations were constructed in AcrB in positions identified as important for substrate recognition in order to investigate the accessibility of the cysteine to the hydrophilic thiol-reactive fluorophore fluorescein-5-maleimide and the results suggest that substrates induce conformational changes in AcrB.[49] Finally, a Na+-dependent NorM homologue, HmrM, conferred MDR when expressed in E. coli (Table II).[322]

3.2 Betaproteobacteria (Burkholderia, Neisseria and Brucella)

Class Betaproteobacteria contains many bacterial species that interact intimately with plants or animals. Possibly the patterns of their efflux activity reflect this mode of living.

3.2.1 Burkholderia spp

This species include several pathogenic members for diseases in humans and animals such as Burkholderia cepacia complex, Burkholderia mallei and Burkholderia pseudomallei. OM of these organisms contains, as the major porin, a trimeric protein (Omp38) that is a distant relative of E. coli OmpF. Nevertheless, permeation through this channel appears to be slower, by one or two orders of magnitude, than that through OmpF.[323] With the low permeability OM, efflux is expected to be efficient in creating resistance, and indeed genomes of several Burkholderia species contain many drug efflux pump genes (http://www.membranetransport.org). For instance, there are, respectively, 14, 9, and 12 RND-type transporters in the genomes of Burkholderia cenocepacia, B. mallei, and B. pseudomallei.[324] In B. cenocepacia expression of the four RND pumps was detectable and one of these pumps was inducible by chloramphenicol. When overexpressed in E. coli, an RND gene (orf2), whose expression was not detectable in B. cenocepacia, conferred resistance to several antibiotics and to ethidium bromide.[324] The ceoR repressor gene was identified upstream of the previously characterized RND operon, ceoAB-opcM.[325] In B. pseudomallei, two RND pumps, BpeAB-OprB (and its BmeR repressor) and BpeEF-OprC (and its BpeT repressor), were characterized (Table I)[326328] in addition to AmrAB-OprA.[1] BpeEF-OprC, homologous to CeoAB-OpcM of B. cenocepacia, conferred resistance to chloramphenicol and trimethoprim.[1] In both ceoR-ceoAB-opcM and bpeT-bpeEC-oprC gene complexes, there is an additional gene, llpE (encoding a lipase-like hydrolase protein) between the repressor gene and the RND structural genes.[325, 328, 329] The llpE gene is conserved in the isolates of B. cepacia complex and may benefit the bacterial survival in the cystic fibrosis lung.[329] A recent study confirmed widespread expression of 7 RND pumps in clinical B. pseudomallei strains.[330]

3.2.2 Neisseria spp

Neisseria porins, which are trimeric and show high permeability, have been studied intensively over the years. One of their outstanding characteristics is the anion selectivity that forms a contrast to the cation selectivity of enterobacterial porins;[331] this may be important in making Neisseria spp. much more susceptible to anionic penicillins. The major multidrug pump, MtrCDE,[1] also contributes to resistance to cationic antibacterial peptides in Neisseria meningitidis[332] as well as modulates the in vivo fitness of Neisseria gonorrhoeae.[333] It has been known that penicillin resistance in gonococci requires simultaneous overexpression of MtrCDE and a mutation in the PIB porin; in this remarkable example of OM/pump synergy, the porin mutation lowers the influx of penicillin thereby increasing the effectiveness of the pump to produce resistance.[334, 335] A clinical isolate with reduced ceftriaxone susceptibility and MDR also has a porin mutation and overexpression of MtrCDE.[336] High occurrence of simultaneous mutations in target enzymes and MtrRCDE was observed in quinolone-resistant N. gonorrhoeae.[337] Inactivation of the ABC transporter MacAB in clinical isolates only slightly decreased resistance to azithromycin and erythromycin but macAB overexpression enhanced the macrolide resistance of gonococci defective in MtrCDE pump.[338]

Mutations in either the gonococcal or meningococcal norM gene (for MATE pump) resulted in increased susceptibility to antibacterial cationic compounds.[339] Resistance to tetracycline and doxycycline (not minocycline) in N. meningitidis of various sources was associated with tet(B) drug-specific efflux gene.[340, 341] As well, the presence of a possible constitutive efflux pump for tetracycline resistance, which might be inhibited by reserpine, was also suggested in N. gonorrhoeae.[342] A gene for a TolC-like protein of N. meningitidis is cotranscribed with the gene for HlyD protein and is required for extracellular production of the repeats-in-toxin toxin FrpC. However, this TolC cannot functionally replace the OM protein MtrE of MtrCDE for antibacterial resistance.[343]

3.2.3 Brucella spp

These Gram-negative coccobacilli are members of the order Rhizobiales, and are related to Rhizobia and Agrobacterium. The genus contains several different species, each with slightly different specificity for host animals. They cause brucellosis, a zoonotic disease which may be transmitted to humans.[344] Brucella spp. contain trimeric porins homologous to E. coli porins, with roughly comparable permeability,[345] and with such a permeable OM the contribution of efflux to drug resistance is predicted to be not so extreme. Nevertheless, the genomes of Brucella abortus, Brucella canis, Brucella melitensis, Brucella ovis and Brucella suis show the presence of two dozens of the putative drug efflux transporters belonging to MFS, RND, SMR, and ABC (http://www.membranetransport.org).[346, 347] An MATE-family pump was also identified in B. melitensis, and its expression in a drug-hypersusceptible E. coli strain produces MDR (Table II), although the disruption of this gene in B. metitensis did not alter the susceptibility to ciprofloxacin.[348] BepC, a TolC homologue identified in B. suis,[349] is functionally involved in two RND systems, BepDE and BepFG, which interplay in providing MDR (Table I).[350] BepC with 25% identity to E. coli TolC was surprisingly able to complement TolC deficiency in E. coli in restoring MDR but not in haemolysin secretion.[349] Efflux also contributes to resistance to erythromycin and fluoroquinolones.[351, 352]

3.3 Epsilonproteobacteria (Campylobacter and Helicobacter)

3.3.1 Campylobacter spp

Campylobacter spp. are major foodborne pathogens and show increasing resistance to antibacterials. The major OM protein is a trimeric porin,[353] which seems to produce high permeability channels comparable to E. coli porins.[354]. Yet multidrug efflux pumps are reported to play a major role in drug resistance.[355, 356] Campylobacter jejuni contains at least 14 putative drug efflux pumps, including 3 RND (CmeB, CmeD, and Cj1373), 4 MFS, 4 SMR and 1 ABC transporters.[357] Two functionally characterized RND systems, CmeABC and CmeDEF, contribute to intrinsic resistance. Overproduction of CmeABC has been demonstrated in isolates that are resistant to macrolides, fluoroquinolones and tetracyclines.[355, 358365] Frequent variations in cmeB gene sequence have been observed.[366]

Macrolides and fluoroquinolones are the drugs of choice for therapy of Campylobacter infections. Chickens, which are fed tylosin-containing feed and infected with C. jejuni or Campylobacter coli, yielded resistant mutants that had the contribution from CmeABC.[367] Low-level macrolide resistance was also found due to CmeABC and other uncharacterized efflux pump(s) (but not to CmeDEF), and was minimized by the EPI Phe-Arg-β-naphthylamide.[368, 369] Enrofloxacin treatment of chickens infected with susceptible Campylobacter promoted the emergence of CmeABC-associated fluoroquinolone-resistant mutants.[355] A recent study shows antisense-mediated gene silencing by cemA-specific peptide nucleic acid for inhibition of CmeABC pump.[370]

Contribution of CmeDEF to intrinsic resistance is likely secondary compared with that of CmeABC.[371] Nevertheless, the cmeB/cmeF double mutants in C. jejuni showed further decrease in MDR than single mutants, and moreover, the double mutations impaired cell viability.[371] Disruption of cmeB did not affect the expression levelsof cmeF and vice versa.[372] There is evidence that a non-CmeB or -CmeF efflux pump or reduced uptake is involved in conferring MDR.[373]

Campylobacter isolates exhibit either high- or low-level erythromycin resistance phenotype. Cross-resistance to erythromycin, clarithromycin and the ketolide telithromycin was observed in the high-level resistant isolates due to mutations in the 23S rRNA. Low-level erythromycin resistance was, in contrast, mediated by Phe-Arg-β-naphthylamide-sensitive, macrolides/ketolide-selective efflux mechanism which remains unidentified.[368] A synergy between CmeABC and the ribosomal modifications was observed in macrolide resistance.[374] Recently, a theoretical model that explains such synergy has been proposed.[375]

3.3.2 Helicobacter pylori

This Gram-negative, microaerophilic bacterium inhabits various areas of the stomach and duodenum and is linked to the development of ulcers. Its OM contains a porin producing a large channel,[376] but it is unknown what fraction of this channel is open. H. pylori is intrinsically resistant to multiple antibacterials such as glycopeptides, polymyxins, nalidixic acid, trimethoprim, sulfonamides, nystatin,amphotericin B, and cycloheximide.[377] Two early studies suggested the absence of functional efflux mechanisms for intrinsic resistance in H. pylori,[378, 379] despite the presence of putative RND efflux systems (HefABC, HefDEF, and HefGHI).[378] However, a recent study revealed that inactivation of HefA renders a chloramphenicol-selected MDR mutant more susceptible to multiple antibacterials (Table I).[380] Re-examination of HefC, HefF and HefI mutants found that HefC is, in fact, involved in MDR (Table I).[380] Another study[381] identified 26 putative transporters belongingto the RND, MFS and ABC families as well as only oneputative MATE transporter that is involved in ethidium efflux. Characterization of the four TolC homologues revealed the involvement of one in resistance to ethidium bromide and another in resistance to novobiocin and sodium deoxycholate. Inactivation of the two TolC homologues increased susceptibility to metronidazole.[381] A TolC-like protein constitutes the OM component of the RND-type CznABC metal efflux pump that provides resistance to cadmium, zinc and nickel salts and is also essential for gastric colonization.[382]

3.4 Bacteroides spp

Bacteroides is a genus of Gram-negative anaerobes that is phylogenetically very far away from the phylum Proteobacteria discussed so far. Bacteroides constitutes substantial portion of the mammalian gastrointestinal flora which are bile-resistant. Analysis of Bacteriodes proteomes suggests a capacity to use a wide range of dietary polysaccharides.[383] Bacteriodes fragilis is considered both the most frequent clinical isolate and the most virulent Bacteroides species.[384] The B. fragilis cell envelope undergoes major changes in protein expression and ultrastructure in response to stressors such as bile and antibacterial agents. The latter may also act as signals for attachment and colonization.[383] Several proteins were reported as porins in B. fragilis OM, one study showing that the porins are very inefficient, by a factor of 10 or even more, compared with the E. coli porins.[385] Bacteroides are intrinsically resistant to a variety of structurally unrelated antibiotics including certain β-lactams and aminoglycosides.[386] Acquired resistance to erythromycin and tetracycline has been observed and prompted concerns that Bacteroides species may become a reservoir for resistance in other, more highly pathogenic bacterial strains.[387]

Drug efflux plays a key role in resistance in Bacteroides[1, 384, 388] and, for example, efflux of fluoroquinolonesby NorA/Bmr of B. fragilis and BexA of Bacteriodes thetaiotaomicron was described earlier.[1] On the basis of homology, 16 putative RND efflux pumps in B. fragilis, named bmeABC116, were identified.[389] Disruption of bmeB15 led to increased susceptibility to a range of antibacterials (Table I).[389] BmeABC5 conferred metronidazole resistance in a clinical isolate, which contained a mutation in the promoter region of bmeC5 (coding for the OM component), preventing the binding of the repressor BmeR5.[390] Expression of all bmeB genes except bmeB9 was detectable.[391] Construction of multiple deletion mutants demonstrated that seven BmeB pumpsare functional and have overlappingsubstrate profiles, and at least four confer intrinsic resistance in an additive manner.[391] MDR strains of Bacteriodes have been isolated clinically or selected in the laboratory with resistance attributable to elevated efflux activities of RND systems.[391] In other studies overexpression of various RND pumps was demonstrated in clinical isolates showing increased resistance levels to several drugs.[392, 393] A Bacteroides conjugative transposon, CTnGERM1, contains genes that are also observed in Gram-positive bacteria, such as a gene for Mef(A), a macrolide efflux pump.[394]

4. Drug Efflux in Gram-Positive Bacteria

The drug efflux pumps in Gram-positive bacteria are usually non-RND pumps and often the singleton protein pumps belonging to the MFS, MATE, SMR or ABC. Those pumps reported or further characterized since 2003 are listed in Table II. The significance of the drug pumps in individual bacteria is discussed below.

4.1 Members of Phylum Firmicutes (Clostridium, Bacillus, Listeria, Staphylococcus, Lactococcus, Lactobacillus, Enterococcus and Streptococcus)

Most of Gram-positive bacteria described below belong to the large phylum Firmicutes.

4.1.1 Clostridium spp

Clostridium is a genus of Gram-positive, obligate anaerobes that include at least four important pathogens in humans i.e., Clostridium botulinum, Clostridium difficile, Clostridium perfringens, and Clostridium tetani. In particular, C. difficile is a significant cause of pseudomembranous colitis as it can overgrow other bacteria and disrupt indigenous intestinal microflora during antimicrobial therapy. Clindamycin, third-generation cephalosporins, penicillins, and fluoroquinolones are considered to have the greatest risk factors for producing C. difficle infections.[395] This would also suggest intrinsic or acquired drug resistance in Clostridium, which is a factor promoting C. difficile outbreak in hospitals.[396] Indeed, the genome of a virulent and MDR C. difficile strain shows a large proportion of the genome with mobile genetic elements putatively responsible for the acquisition of genes involved in resistance and virulence.[397] The cdeA gene from a clinical C. difficile isolate encodes a Na+-coupled MATE efflux pump. When expressed on plasmid, this pump conferred MDR upon C. perfringens and E. coli (Table II). There was an elevated cdeA expression in C. difficile in the presence of ethidium bromide (although not ciprofloxacin).[398] Another efflux pump, Cme, a MefA/MefE homologue from C. difficile, was able to confer resistance in Enterococcus faecalis (Table II).[399] Fluoroquinolones show limited activities against anaerobic bacteria and the efflux appears to be a mechanism for the resistance in Clostridium hathewayi.[400] However, another study revealed that high-level fluoroquinolone resistance in toxin-A-negative, toxin-B-positive C. difficile isolates was associated with a novel mutation in the target gene gyrB and that efflux inhibitors had little impact on the resistance.[396] Tet efflux proteins such as Tet(P) and Tet(40) are also distributed in Clostridium.[401, 402]

4.1.2 Bacillus spp

Several multidrug pumps including Bmr, Blt and Bmr3 have been described earlier in B. substilis.[1, 403] A Bmr3-overproducing mutant selected by puromcyin exhibited the increased stability of bmr3 transcripts and MDR.[403] LmrB, a fourth multidrug efflux pump, was identified from spontaneous mutants of B. subtilis by puromycin and lincomycin selection.[404] The lmrB efflux gene and the lmrA repressor gene form an operon.[405] Mutations in two regions immediately downstream of the -10 lmrAB promoterregion increased lmr transcription in lincomycin-resistant mutants.[406] LmrA autogenously represses the transcription of lmrAB through binding to the lmrAB promoter region. Interestingly, LmrA also represses the expression of another gene, yxaG that encodes an iron-containing quercetin 2,3-dioxygenese. However, the latter apparently is not involved in MDR, although it forms an operon with yxaH that encodes a putative membrane protein and may function as a drug exporter.[405] Tet(L) tetracycline efflux protein from B. subtilis has been characterized as a dimer.[407]

An MDR operon mdtRP (encoding the MdtR repressor and the MFS pump MdtP) is involved in resistance to fusidic acid and other agents.[408] YvcC (BmrA), a functional ABC transporter in B. subtilis, is homologous to mammalian P-glycoprotein and to LmrA of L. lactis. This transporter was constitutively expressed in B. subtilis, and its deletion decreased ethidium efflux. Inverted membrane vesicles prepared from overexpression of YvcC in E. coli exhibited high transport activities for Hoechst 33342 (a lipophilic fluorescent bisbenzimide agent), doxorubicin, and 7-aminoactinomycin D.[409]

Fluoroquinolones such as ciprofloxacin are drugs recommended for the treatmentof anthrax. Studies have been carried out to identify the steps necessary to obtain high-level resistance to fluoroquinolones in Bacillus anthracis and to characterize the underlying mechanisms. Although GyrA and/or ParC mutations were the major mechanisms, efflux was also observed in the mutants obtained at certain steps.[410412] The identity of the efflux pump(s) remains unknown.

4.1.3 Listeria monocytogenes

The Gram-positive bacterium L. monocytogenes is a ubiquitous,intracellular pathogen implicated as the causative organism in various outbreaks of the foodborne disease, listeriosis. It is estimated that 20–30% of foodborne listeriosis infections in high-risk individuals may be fatal.[413] Although L. monocytogenes is usually susceptible to most antibacterials,[414, 415] strains resistant to some agents have been isolated recently.[416418]

Drug efflux determinants including floR, tet(A) and tet(K) have been also observed in L. monocytogenes.[417, 418] Isolates resistant to heavy metal (cadmium and arsenic) salts and benzalkonium chloride were also obtained[419, 420] and further resistance to ethidium bromide was likely associated with an efflux pump.[420] The multidrug transporter MdrL is partially responsible to adaptation of L. monocytogenes to benzalkonium chloride[1, 421] and can also be repressed by LadR, a PadR-related transcriptional regulator.[422] As described in section 8, this and two other multidrug transporters in L. monocytogenes were recently found to be involved in controlling the innate host immune response.[423] Another MFS drug pump, Lde, is a homologue of PmrA from Streptococcus pneumoniae and is involved in resistance to fluoroquinolones and toxic compounds.[424] Intriguingly, Lde as a bacterial efflux pump cooperates with a eukaryotic MRP-like efflux transporter to reduce the activity of ciprofloxacin, a substrate of both pumps, in J774 macrophages infected with L. monocytogenes.[425] The in vivo-induced virulence factor Hpt mediates uptake of fosfomycin in L. monocytogenes (which is resistant to fosfomycin in vitro), making an antibacterial in-vitro/in-vivo paradox, i.e., the bacteria are resistant in vitro but are susceptible to the drug in vivo.[426]

4.1.4 Staphylococcus spp

Efflux is an important resistance mechanism in S. aureus.[1, 427] In addition to the previously described chromosomal NorA pump and plasmid-encoded MsrA and QacA/B pumps,[1] additional chromosomally-encoded pumps have been characterized from S. aureus and these include the MFS-type NorB, NorC, MdeA, SdrM and Tet(38), the MATE-type MepA, the SMR-type SepA, and the ABC-type AbcA and Sav1866, which (except Tet[38]) are all multidrug pumps with the substrates shown in Table II.[428434] Intriguingly, the sepA gene is located immediately downstream of sdrM, and SepA is the only known SMR pump encoded on the chromosome in S. aureus.[435] Additionally, NorB can facilitate bacterial survival when overexpressed in a staphylococcal abscess and may contribute to the relative resistance of abscesses to antibacterial therapy, thus linking bacterial fitness and resistance in vivo.[436] Among the plasmid-coded genes, msr(A) not only encodes a macrolide efflux pump but also is required for expression of mph(C) that encodes a phosphotransferase for inactivating some macrolide antibiotics.[437] A plasmid-encoded phenicol efflux pump FexA was identified in Staphylococcus lentus.[438]

In a recent study, ca. 50% of the 232 bloodstream isolates of S. aureus were considered as strains exhibiting efflux of at least two structurally unrelated substrates. Frequencies of overexpressed efflux genes were mepA (4%), mdeA (11%), norA (23%), norB (25%) and norC (17%), and ca. 20% of the strains overexpressed two or more efflux genes.[439] The prevalence of msrA/msrB efflux genes was significantly higher in the invasive MRSA spa-type t067 than in the other MRSA spa-types in a national survey in Spain.[440] Exposure to several substrates significantly increased norA expression.[441] Low concentrations of several biocides and dyes also selected the mutants overexpressing mepA, mdeA, norA and norC, with mepA overexpression predominating. Overexpression was frequently associated with promoter-region or regulatory protein mutations.[442] Loss of NorA pump leading to susceptibility to fluoroquinolones was observed in laboratory-generated vancomycin intermediate resistant S. aureus strains.[443] The mef(A) efflux gene was detected in Staphylococcus sciuri resistant to macrolides, lincosamides, streptogramins, and linezolid.[444]

The SMR-type QacD pump, encoded by plasmids and conferring a low-level antiseptic resistance, has been found in both methicillin-susceptible S. aureus (MSSA) and MRSA, and the smr gene cassettes were classified into three types.[445, 446] High-level antiseptic resistance genes qacA and qacB were more frequent in MRSA isolates than in MSSA isolates.[445] QacA, but not QacB or QacC, confers in vitro resistance to thrombin-induced platelet microbial protein 1 (tPMP-1), a cationic antibacterial polypeptide, apparently by a mechanism that does not involve efflux.[447] The presence of Tet(M) ribosomal protection or Tet(K) efflux proteins has no discernible effect on the tigecycline activity for either MRSA or MSSA strains.[448] Novel agents may be sought to bypass the efflux mechanism. A novel des-fluoro[6] quinolone, DX-619, generates resistant S. aureus mutants only at a very low frequency; the mechanism of resistance in these mutant strains is unlikely to be the conventional ones.[449] Co-presence of qacA/B, qacG, qacH, qacJ and/or smr efflux genes were confirmed in Staphylococcus haemolyticus human isolates.[450]

4.1.5 Lactococcus lactis and Lactobacillus spp

In silico analysis of the genome of non-pathogenic, Gram-positive L. lactis suggests the presence of 40 putative drug transporters including the previously described LmrA (ABC transporter) and LmrP (MFS).[1, 451] An additional heterodimeric ABC transporter, named LmrCD (YdaG/YdbA), has been reported as a major determinant of both intrinsic and acquired MDR in L. lactis. Up-regulation of lmrCD in resistant strains was observed, while deletion of lmrCD led to the hypersusceptibility to toxic compounds including bile salts (Table II), but not to common antibiotics.[451, 452] Cholate-induced wild-type cells, which actively extrude cholate, differ from LmrCD-deficient, cholate-selected resistant cells, whose resistance seems to involve multiple responses.[453] A local transcriptional repressor of lmrCD, LmrR (YdaF), which belongs to the PadR family, interacts with drugs to cause lmrCD up-regulation.[454] Mdt(A), originally described in L. lactis, is a plasmid-encoded drug pump and confers resistance to macrolides, lincosamides, streptogramins and tetracycline.[1] A mutated mdt(A) gene containing inactivating mutations was identified in susceptible Lactococcus garvieae strains.[455]

Lactobacillus species are a major group of lactic acid bacteria and play an important role in promoting intestinal and vaginal health. An ABC multidrug exporter HorA, a homologue of LmrA of L. lactis, is involved in hop resistance in Lactobacillus brevis.[456] Additional unidentified proton-dependent pump also contributes to hop resistance.[457] Several Lactobacillus species derived from broiler chickens displayed tetracycline resistance due to the presence of the efflux genes tet(K), tet(L) or tet(Z) and the ribosomal protection genes tet(M) or tet(W).[458, 459] A tetracycline-resistant Lactobacillus sakei showed the coexistence of two different tetracycline resistance mechanisms, plasmid-carried efflux gene tet(L) and chromosomally-located transposon-associated tet(M).[460] Bile-mediated aminoglycoside sensitivity in Lactobacillus species likely results from increased membrane permeability.[461] Heterologous expression of BetL, a betaine uptake system of L. monocytogenes, enhances the stress tolerance of Lactobacillus salivarius.[462]

4.1.6 Enterococcus spp

This species is resistant to numerous antibacterials with efflux as a key mechanism of resistance as described previously.[1] Additional efflux systems have been found. EfrAB, an ABC exporter in E. faecalis, conferred MDR upon a drug-hypersensitive E. coli (Table II) and this efflux activity was inhibited by reserpine, verapamil, and o-vanadate inhibitors of ABC pumps.[463] Lsa pump is involved in intrinsic resistance to lincosamides and streptogramins in E. faecalis[1] and the lsa-like genes of clinical isolates susceptible to lincosamides and dalfopristin carried premature termination mutations.[464] However, acquired intermediate-level gentamicin resistance in E. faecalis was not associated with clear indications of an active efflux.[465] Sparfloxacin- or norfloxacin-selected resistant E. faecalis mutants contained a non-EmeA (see ref[1]), NorA-like pump.[466] Analysis of the foodborne Enterococcus faecium and E. faecalis indicated the presence of a number of the resistance determinants such as tet(L), tet(M) and tet(K) (for tetracycline resistance) and ermA,B,C, mefA,E, msrA/B and ereA,B (for erythromycin resistance). All E. faecium strains contained the msrC gene that encodes an erythromycin exporter,[467, 468] in spite of an early study that led to a different conclusion.[469] An MFS pump, EfmA of E. faecium was characterized (Table II).[470]

4.1.7 Streptococcus pneumoniae and relatives

This human pathogen causes many types of pneumococcal infection and is a common cause of bacterial meningitis. Efflux-mediated drug resistance is common in this species.[1] An in vivo exposure to ciprofloxacin resulted in predominately efflux-mediated resistant mutants, suggesting that efflux plays a central role in emergence of fluoroquinolone resistance.[471] Azithromycin selected for efflux-type low-level resistance to macrolides.[472]

The MFS-type MefA, MefE and PmrA exporters are involved in macrolide or fluoroquinolone resistance.[1] In macrolide-resistant isolates from various geographical areas, efflux mediated by MefA-and/or MefE were predominant.[473476] The presence of a tet(O)-mef(A)chimeric element indicated the genetic linkage between macrolide and tetracycline resistance.[477] Non-PmrA efflux pumps were also associated with fluoroquinolone resistance in S. pneumoniae.[478, 479] Consistently, fluoroquinolone-susceptible isolates did not exhibit efflux.[480] The activity of the ketolide telithromycin can also be reduced by MefA in Streptococcus pyogenes[481] although the ketolides possess enough activity against efflux-positive isolates.[482]

An MDR mutant obtained after exposure of capsulated wild-type strain of S. pneumoniae to ciprofloxacin constitutively overexpressed 22 genes including patA and patB that encode a heterodimeric ABC transporter (Table II). Expression of patAB was induced by ciprofloxacin in both wild-type and resistant strains.[483] Quinolones and distamycin also strongly induced patAB expression in fluoroquinolone-sensitive strains. A second group of quinolone-induced transporter genes are SP1587 and SP0287, which are homologues of, respectively, oxalate/formate antiporters and xanthine or uracil permeases belonging to the MFS.[484] Interestingly, the EPI reserpine selected MDR mutants that overexpressed PatA and PatB, despite the fact that only patA was involved in reserpine resistance.[485] Exposure to subinhibitory ciprofloxacin resulted in patAB-mediated efflux regardless of the expression of pmrA.[486] Another heterodimeric ABC-type pump, SP2073-SP2075, was identified from a PmrA-deficient S. pneumoniae strain to mediate intrinsic resistance to toxic agents and certain quinolones (Table II). Inactivation of other putative MFS, MATE, and ABC-type drug exporters did not alter the drug susceptibility. [487] An ABC transporter, Spr0812/0813, was required for intrinsic resistance to bacitracin, but an overexpression of a mutant Spr0813 permease lacking the two C-terminal helices resulted surprisingly in reduced susceptibility to vancoresmycin, an antibiotic of tetrameric acid (2,4-pyrrolidine-dione) class.[488]

Efflux pumps encoded by mef(A) and mef(E) genes are among the most common mechanisms of resistance to macrolides (M phenotype) in streptococci. These genes may be located on the chromosomes (e.g., chromosomal chimeric tet(O)-mef(A))[489, 490] but are more often associated with transferable elements such as the mef(E)-containing macrolide efflux genetic assembly (MEGA) element or mef(A)-containing transposons.[490492] The mef(A) elements of Streptococcus pyogenes are likely prophage-associated.[493] Conjugative transfer of mef(E) from viridans streptococci to S. pyogenes was demonstrated. In all cases of conjugal transfer of mef(E), the gene was carried on MEGA.[494] Another study suggested that mef(A) and mef(E) genes were also observed in ca. 9% of erythromycin-resistant isolates of Streptococcus agalactiae and transformation was considered the main mechanism for resistance gene acquisition.[495] The mef(A) gene has also been found in Gram-negative bacteria[496] and can be transferred to E. faecalis and E. coli recipients.[497]

A mel (msr(D)) gene that encodes an ABC transporter is cotranscribed with the mef(E). Both mel and mef(E) were inducible by macrolides and were required for macrolide resistance.[498, 499] The mef(E)-MEGA element was inserted into a Tn916-like genetic element to form a new composite element, Tn2009 containing both mef(E) and tet(M).[500] Tn2009 can further absorb erm(B) to form another new composite, Tn2010.[501] A novel mef gene variant, mef(I), was identified in Streptococcus pseudopneumoniae,[502]; mef(I), an adjacent new msr variant, and catQ chloramphenicol resistance gene form a composite structure, 5216IQ complex.[503]

4.2 Bifidobacterium spp. (a Member of Phylum Actinobacteria)

The Gram-positive, anaerobic bifidobacteria belong to the phylum Actinobacteria that contain also Corynebacterium and Mycobacterium. Bifidobacterium is an important natural inhabitant of the human intestinal microflora. Like other constituents of this microflora, Bifidobacterium has evolved to tolerate inhibitory factorsin the intestinal niche, such as bile salts and antibacterial peptides.[504, 505] Drug efflux is probably a major mechanism for such tolerance or resistance. A protein with 8 transmembrane segments, BbmR of Bifidobacterium breve exhibits characteristics reminiscentof MDR proteins and confers resistance to macrolides azithromycin, clarithromycin and dirithromycin.[506] Two genes (abcA and abcB) from B. breve encoding a putative ABC efflux transporter were coexpressed in the heterologous host L. lactis and conferred resistance to nisin and polymyxin B (Table II).[507] The ctr gene of Bifidobacterium longum encodes a cholate efflux exporter and confers resistance to cholate, chloramphenicol, and erythromycin in the heterologous host E. coli. Ctr belongs to the sodium/bile acid family of transporters, which had not been reported previously to cause antibiotic resistance.[504] A recent study identified bile salt-affected, envelope-associated proteins including ABC trasnporters of B. longum.[508]

5. Drug Efflux in Mycobacteria

The significance of drug efflux in mycobacteria has been discussed previously,[1] and is also the subject of other reviews.[509512] Indeed, mycobacteria such as M. tuberculosis and Mycobacterium smegmatis contain at least two or three dozens of putative drug efflux transporters.[513, 514] Several of them have been shown to be involved in resistance to aminoglycosides, chloramphenicol, fluoroquinolones, isoniazid, linezolid, rifampicin, tetracycline and other toxic compounds.[1, 510, 512, 514, 515] Enhanced killing of intracellular multidrug-resistant M. tuberculosis by efflux pump inhibitors (EPIs) was recently demonstrated.[516] Nevertheless, it is not entirely clear how these pumps create significant levels of resistance if they pump out drugs only into the “periplasmic space,” inside the highly impermeable cell wall.

The M. tuberculosis genome contains 13 putative RND-type transporters, designated MmpL (mycobacterial membrane proteins, large).[513] However, the inactivation of 11 out of 13 of the mmpL genes including mmpL7 did not alter the drug susceptibility.[517] The mmpL4, mmpL7, mmpL8, and mmpL11 genes are instead involved in the virulence in mice.[517] Nevertheless, when expressed in M. smegmatis, MmpL7 confers a high-level resistance to isoniazid due to efflux and this resistance level decreases in the presence of the EPIs.[518] MmpL7 also catalyzes the export of phthiocerol dimycocerosate in M. tuberculosis and an MmpL7-deficient mutant is attenuated for growth in the lungs.[517] MmpL8 exports a sulfatide precursor, 2,3-diacyl-α,α′-trehalose-2′-sulfate.[519]

There are at least 26 putative ABC drug exporters in M. tuberculosis.[520] The Rv2686c-2687c–2688c operon encodes an ABC exporter and its expression in M. smegmatis mediates resistance to fluoroquinolones, which is reduced in the presence of EPIs.[521] There are also 16 putative MFS drug efflux proteins proteins.[522] A Tap-like pump (Rv1258c) was overexpressed in an MDR clinical isolate of M. tuberculosis.[523] In M. bovis BCG the Mb2361c protein, homologous to the MFS-type Rv2333c pump from M. tuberculosis,[522] is involved in intrinsic resistance to spectinomycin and tetracycline.[524]

In an important development, the whiB7 gene, which is a primary regulatory gene and coordinates resistance to drugs, was characterized in M. tuberculosis. The whiB7 expression was induced by erythromycin, tetracycline, and streptomycin as well as by fatty acids and whiB7 deletion mutants were hypersusceptible to clarithromycin, erythromycin, lincomycin, spectinomycin and streptomycin. Induction of whiB7 was correlated with the expression of genes associated with resistance, including Rv1258c (tap for a drug efflux pump), Rv1473 (encoding a putative macrolide transporter) and Rv1988 (erm for ribosomal methyltransferases).[525]

When expressed in Mycobacterium bovis BCG, the M. tuberculosis iniA gene confers resistance to isoniazid and ethambutol, two first-line antituberculosis agents. These two agents also induce the expression of iniA and the linked iniB/iniC in M. tuberculosis, but iniA deletion results in increased susceptibility only to isoniazid. IniA may function as a MDR pump component, although the type of the pump remains unknown.[526] Alkyl diphenyl ethers that are high affinity InhA inhibitors with activity against drug-resistant M. tuberculosis mutants were unable to up-regulate a putative drug efflux pump.[527]

The M. smegmatis genome contains many genes encoding putative drug efflux pumps. The expression of the lfrA gene (encoding the first-identified mycobacterial efflux pump[1]) and the homologues of M. tuberculosis Rv1145, Rv1146, Rv1877, Rv2846c (efpA) and Rv3065 (mmr and emrE) was detectable.[514] Null mutants each carrying a deletion of lfrA, efpA or Rv1877 homologue produced increased susceptibility to various agents, indicating the role of these genes in intrinsic resistance.[514] The repressor LfrR for the LfrA pump was also identified and characterized.[514, 528]

6. Contribution of Efflux Pumps to Resistance in Bacteria of Animal and Environmental Origins

There is an increasing concern on drug resistance in bacteria of animal and environmental origin, which may serve as a reservoir of resistance genes and/or resistant strains in human infections.[2628, 529, 530] Efflux-mediated resistance has often been observed in animal pathogens. Fluoroquinolone resistance in canine P. aeruginosa isolates or in avian pathogenic E. coli isolates involves the efflux pump overexpression (e.g., AcrAB).[531, 532] A macrolide efflux gene, mef(B), which clusters with sulphonamide resistance gene sul3 on a plasmid, was recently reported from porcine E. coli.[533] MexXY overexpression occurred in P. aeruginosa from dairy cows with Pseudomonas mastitis.[534] AcrAB overexpression was found in MDR Salmonella isolated from diseased swine.[535] Efflux contributes to erythromycin and fluoroquinolone resistance in poultry and pig isolates of C. coli.[536] FloR pump mediates florfenicol resistance in pathogenic E. coli isolates of calves[537] and a plasmid-encoded, yet-unidentified chloramphenicol efflux pump was detected in E. coli isolated from poultry carcass.[538] Efflux activity in fluoroquinolone and tetracycline resistant Salmonella and E. coli of poultry origin contributes to reduced susceptibility to household antibacterial cleaning agents.[539]. A number of coagulase-negative S. epidermidis isolates obtained from milk, heifers and dairy cows carried MsrA efflux-based resistance to erythromycin.[540] MDR in E. coli of both avian and human sources was usually associated with tetracycline efflux genes.[541]

The plasmids harbouring qepA, qnr and aac(6′)-Ib-cr fluoroquinolone efflux/resistance genes were found highly prevalent among ceftiofur-resistant Enterobacteriaceae isolates from companion and food-producing animals.[158, 159] Plasmids containing oqxA (for an RND-type multidrug pump, see Table I) were prevalent in E. coli isolates derived from pigs.[155] Animal pathogens also carry plasmid-borne efflux genes such as floR (for florfenicol resistance) in bovine Pasteurella multocida[542] and tet(L) (for tetracycline resistance) in bovine Mannheimia and Pasteurella [543] and swine Actinobacillus pleuropneumoniae.[544] Inactivation of TolC or its homologue FtlC led to multidrug susceptibility in the zoonotic pathogen Francisella tularensis.[545] We note that overexpression of MarA-like regulator and AcrAB pump can cause MDR of Yersinia pestis, a possible agent of bioterrorism.[546]

Drug efflux pumps are also evident in environmental isolates. An erythromycin resistance mosaic plasmid, isolated from a sewage treatment plant, harbours resistance determinants, mel (for an ABC-type efflux transporter) and mph (for a macrolide-2′-phosphotransferase) as well as an integron-containing transposon element.[547] MexAB-OprM contributes MDR in P. aeruginosa isolated from farm environments and retail products.[548] The presence of the intI1 (class 1 integrase), qacE (multidrug efflux), and qacEΔ1 (attenuated qacE) genes was significantly higher for the isolates pre-exposed to quaternary ammonium-polluted environments.[549] Unidentified efflux mechanism contributes to phenicol resistance in MDR Chryseobacterium isolates from fish and aquatic habitats.[550] Efflux contribution to resistance in Aeromonas spp. from aquatic sources has been described above.[224226] Tetracycline efflux genes (and other resistance genes) were detected in uncultured soil bacteria which can also be a reservoir of resistance genes.[551] A novel tetracycline-specific Tet41 pump was identified in an environmental strain of S. marcescens.[205] The genome analysis reveals the presence of a number of putative drug efflux pumps of ABC, MFS, RND and SMR types in Chromobacterium violaceum, a Gram-negative bacterium commonly found in aquatic habitats of tropical and subtropical regions.[552]

7. Role of Efflux Pumps in Biofilm Resistance

Bacteria growing in biofilms are more resistant or tolerant to antibacterials than their planktonic counterparts.[553] This is in large part attributable to a strategy with the occurrence of persister cells that shut down the targets to protect cells from killing by antibacterials.[554] It seems possible that efflux pumps also contribute to resistance in biofilms. However, ciprofloxacin resistance in biofilms did not correlate with expression of AcrAB or MarA in E. coli or of MexAB-OprM in P. aeruginosa[555, 556] There was also no up-regulation of MexAB-OprM and MexCD-OprJ in biofilms of P. aeruginosa[557] and overproduction of these two systems did not affect the biofilm formation.[558] Nevertheless efflux pumps may affect drug-specific resistance in biofilms such that resistance to ofloxacin is dependent on the expressionof MexAB-OprM pump at a low ofloxacin concentration range[556] and the MexCD-OprJpump acts as a biofilm-specific mechanism for azithromycin resistance.[559] MexAB-OprM and pmr-mediated lipopolysaccharide modification are also linked to tolerance to colistin in P. aeruginosa biofilms.[560]

Zhang and Mah[561] recently reported the identification of a PA1874–1877-encoded efflux system in P. aeruginosa that is important for biofilm-specific resistance to tobramycin, gentamicin, and ciprofloxacin. Similarly, in an uropathogenic E. coli RapA regulatory protein appears to increase the transcription of a putative MDRpump gene yhcQ and evidence suggests this protein also contributes to the biofilm-specific penicillin G resistance.[562] The biocide triclosan up-regulates the expression of acrAB pump genes and marA pump activator gene in Salmonella biofilm cells.[563] Bile salt-induced B. fragilis cells with elevated RND pump expression increase the possibility for biofilm formation, with increased resistance possibly due to efflux.[564] It was noted that a polymicrobial-biofilm-associated MDR S. aureus isolates carried an MDR gene cluster including macrolide efflux gene msrA.[565]

8. Role of Drug Efflux Pumps beyond Drug Resistance

MDR efflux pumps can handle a wide range of structurally unrelated substrates including those compounds produced by higher organisms, such as bile salts, fatty acids and hormones.[1, 566571] Thus the pumps are likely to affect the interaction of bacteria with the host animals and plants. Indeed, the drug efflux pumps can respond to a range of stimuli including stress signals, and they influence the colonization, pathogenesis or virulence, cell communications, biofilm formation and other fitness responses.[572574] These functions may well be the physiological functions of at least some of the drug pumps, and may ensure the persistence of drug efflux transporters in evolution.[575] The physiological function of the best studied RND transporter AcrB is clearly to protect E. coli cells from the bile salts and fatty acids that are abundant in the intestinal tract, their normal habitat.[1, 10, 566] Finally, the conventional measurement of the minimal inhibitory concentrations alone may not indicate the extraordinary capacity of MDR transporter;[576] this was shown clearly by the observation that cephaloridine is pumped out strongly by AcrB although the susceptibility of E. coli to this drug is scarcely affected by the deletion of this pump.[64]

8.1 Bacterial Stress Responses

Bacteria possess a complex regulatory network to ensure a coordinated and effective response to various types of stress.[577] The role of MDR pumps in stress responses was demonstrated early by the stress-induced AcrAB expression in response to fatty acids, ethanol, high salt concentration, etc.[566] This will be discussed below under the Regulation.

P. aeruginosa RND pumps provide a good example for their stress response functions.[573] In response to the ribosome-targeting antibacterials,[578] expression of MexXY is elevated by the aberrant polypeptides and their oxidatively modified counterparts, and this may lead in turn to the removal of such polypeptides.[579, 580] Subsequent to the action of membrane-damaging agents, MexCD-OprJ likely exports the released membrane constituents.[573, 580] MexEF-OprN also may protect the cell by responding to hydrogen peroxide (H2O2) and nitric oxide.[573] Consistent with such ideas, the repressor MtrR of the MtrCDE pump controls 69 genes including rpoH, which encodes the general stress response sigma factor RpoH. RpoH-regulated genes also modulate levels of gonococcal susceptibility to H2O2.[581] Iron starvation in E. coli led to increased expression of the RND gene mdtF and a decrease in acrD.[582] Thus, MDR pumps may often function for toxic waste disposal rather than only for drug resistance.[583]

8.2 Colonization and Virulence

It has been reported that MDR transporters such as RND pumps contribute to the bacterial colonization in the host. For example, TolC (and its homologue) mutants of S. Typhimurium and V. cholerae are deficient in intestinal colonization.[171, 173, 174, 567] Administration of the EPI Phe-Arg-β-naphthylamide also decreased the colonization of C. jejuni.[570] Nishino et al.[172] determined the virulence role of 9 drug transporters of Salmonella, and concluded that MdtABC, MdsABC (a salmonella-specific RND complex) and MacAB were required for virulence and acrAB and acrEF null mutants had impaired ability in causing the mortality of mouse by oral route of infection. A strain deleted for all 9 pump genes did not cause mortality in mice. However, such results may not mean that the pump activity is directly connected with “virulence”, given that the major function of AcrB and its relatives in enteric bacteria is the protection of bacteria against bile salts. In this connection, it is regrettable that often little attention has been paid to the presence of bile salts, and oral challenge has been routinely used without further consideration.

However, there are data that cannot accommodate such trivial explanations. AcrB mutants of S. Typhimurium failed to invade macrophages in vitro AcrB mutants of S. Typhimurium failed to invade macrophages in vitro.[173] MexAB-OprM deletion mutant of P. aeruginosa was greatly reduced in its ability to infect cultured cells.[584] BesABC of Borrelia burgdorferi (a causative agent of Lyme borreliosis) is involved in virulence.[585] A functional MtrCDE system enhances gonococcal genital tract infection in female mice and MtrCDE-deficient gonococci are more rapidly cleared from mice secreting gonadal hormones.[568] BepFG-defective mutant of B. suis is attenuated in virulence.[350] CznABC metal pump of H. pylori is required for urease modulation and gastric colonization.[382] Exposure of B. fragilis cells to bile salts increases, in addition to efflux, bacterial co-aggregation and adhesion to intestinal epithelial cells.[564]

Some of the results above may be explained by the function of MDR pumps in exporting virulence factors. P. aeruginosa MexAB-OprM system exports virulence determinants[1] and contributes to the success of an epidemic clone.[586] A cystic fibrosis epidemic strain of P. aeruginosa overproduces both MexAB-OprM and MexXY-OprM and displays enhanced virulence.[587] However, overexpression of MexCD-OprJ and MexEF-OprN impairs the type III secretion system that delivers toxins to the cytoplasm of the host cells, and this is due to the lack of expression of ExsA, a master regulator of the type III secretion system.[588] MexCD-OprJ up-regulation also impairs bacterial growth and has a strain-specific, variable impact on rhamnolipid, elastase, phospholipase C, and pyocyanin production.[589] PseABC of P. syringae is involved in secretion of lipopeptide phytotoxins.[590] BpeAB-OprB of B. pseudomallei is neededfor optimal production of quorum-sensing-controlled virulencefactors such as siderophore and phospholipase C and for biofilmformation, and the bpeAB mutant is attenuated in their invasiveness and cytotoxicity.[327] Nevertheless, resistance involving pump overexpression may also result in biological cost and affect the fitness. SmeDEF overexpression in S. maltophilia leads to virulence reduction.[591] Reduced fitness has been observed with quinolone-resistant strains of E. coli and P. aeruginosa.[592, 593] Subsequently, fitness-compensatory mutations may beacquired for bacterial survival.[593]

MacAB-TolC of E. coli functions in the secretion of a peptide toxin, the heat-stable enterotoxin II, which is produced by enterotoxigenic E. coli.[594] AcrA and DinF (an MATE pump) of Ralstonia solanacearum contribute to bacterial wilt virulence,[595] and the phytoalexin-inducible AcrAB pump contributes to virulence in the fire blight pathogen, Erwinia amylovora, possibly by excluding these plant toxins.[596]

TolC homologues, as key pump components, are also required for virulence of a large number of bacteria, such as Brucella suis (BepC),[349] Francisella tularensis (causing tularemia)[545] and Salmonella,[172] as well as plant pathogens Erwinia chrysanthemi[597] and Xylella fastidiosa.[598] The TolC-like TdeA protein is required for leukotoxin export in Aggregatibacter actinomycetemcomitans, an oral commensal.[599]

8.3 Quorum Sensing

Bacteria use quorum sensing systems to control gene expression in response to cell density and environmental factors. The process involves the production and detection of extracellular signalling molecules called autoinducers.[600] Among those, N-acyl homoserine lactones have been studied most intensively.[601] There are numerous reports on the role of RND pumps in quorum sensing. However, many of the conclusions demand more careful analysis. This is because in the usual batch culture system, some cells that were turned on early (producers) will be secreting the autoinducer, while the rest of population (receivers) will be responding to this signalling molecule. Thus even when a given transporter exports the signal, it will have opposite effects on these two types of cells, and the outcome in the whole, mixed population is impossible to predict.

Early literature on the role of RND pumps on quorum sensing through N-acyl homoserine lactone was analyzed previously (section 2.2.1 of reference[1]). Moreover, N-acyl homoserine lactones should easily diffuse across any membrane as a lipophilic, uncharged molecule, and it is difficult to imagine that they need to be pumped out actively by an RND pump (although extremely hydrophobic members may be pumped out from within the membrane interior, to avoid self-poisoning of the producer cells). We thus concluded that there is no evidence that the secretion of N-acyl homoserine lactones by producer cells requires RND pumps, and the overproduction of RND pumps are likely to hinder the entry of autoinducers into receiver cells.[1] In fact, overproduction of MexEF-OprN system decreases, rather than increases, the production of N-acyl homoserine lactone autoinducers by the whole population.[1] Nevertheless, several reviews have emphasized the “role” of RND pumps in quorum sensing without careful analysis, and there are studies that “confirm” this purported “physiological” role of the pumps, and this myth of autoinducer export by RND pumps still continues. In an extension of an earlier study, it was shown that deletion mutants in the mexHI-opmD system are drastically reduced in the production of autoinducers.[602] However, as pointed out earlier,[1] such mutants overproduce other RND systems and the data may simply mean that the exclusion of autoinducers through efflux results in the failure to convert the receiver cells (the majority of the population) into producer cells, rather than the authors’ interpretation that MexHI-OpmD is essential in the export of autoinducers. Possibly a similar interpretation applies to the observation that deletion of BpeAB-OprB pump hinders the production of autoinducers in batch cultures of B. pseudomallei.[327, 603] In another study, AcrAB deletion mutant of E. coli was found to reach a 10% higher final density in the stationary phase, in comparison with the wild type. Furthermore, the culture supernatant of an AcrAB overproducer decreased the final density of the wild type culture. It was concluded that AcrAB pumped out N-acyl homoserine lactone signal.[604] In our opinion, the data are far too insufficient for such a conclusion.

An autoinducer enhances MexAB-OprMexpression but this activity is repressedby MexT (PA2492), the activator of MexEF-OprN pump.[605, 606] Macrolides modulate the quorum-sensing system of P. aeruginosa and the cell density-dependent expression of MexAB-OprM is repressed by a subinhibitory concentration of azithromycin.[607] The phenazine pyocyanin is a physiological signalling factor for the up-regulation of several quorum sensing-controlled genes including those encoding MexHI-OpmD pump.[608] Overexpression of the quorum sensing regulator SdiA in E. coli is linked to the increased levels of AcrAB pump.[1] The exogenous autoinducers N-acyl homoserine lactones modulate expression of four quorum sensing regulatory luxR genes and four bme RND pump genes and biofilm formation in B. fragilis.[609] In the plant pathogenic Burkholderia glumae, quorum sensing is involved in pathogenicity by the regulation of biosynthesis and export of a very hydrophilic phytotoxin toxoflavin (in which a pyrimidine is fused to a triazine) by ToxGHI RND pump.[610]

8.4. Other Cell Physiology

Drug efflux pumps also influence additional cell physiology other than those described above. AcrEF-deficient E. coli cells are defective in chromosome condensation and segregation and thus AcrEF plays a role in the maintenance of cell division, as the name of RND implies.[611] In L. monocytogenes, the MFS drug transporters such as MdrL and the newly-identified MdrM and MdrT control the magnitude of a host cytosolic surveillance pathway, leading to the production of several cytokines, a result linking bacterial MDR to host immunity.[423] Moreover, the genes encoded for MdrL, MdrM and MdrT are, respectively, linked to the regulatory genes encoding LadR, TetR and MarR,[423] suggesting the importance for controlling the efflux pump expression.

9. Regulation of Drug Efflux Pump Expression

9.1 Multiple-Level Genetic Regulation: Involvement of Local and Global Regulators/Modulators

Both the presence of numerous multidrug efflux systems and the overlapping functions of the MDR transporters require a well-regulated expression of these efflux systems, which can be subject to multiple levels of regulation. Indeed, involvement of a variety of local and global transcriptional regulators and other modulators underlines the complexity anddiversity of the mechanisms in regulation of drug effluxpumps, as previously described with the regulation of AcrAB-TolC efflux system of E. coli.[1] In particular, most regulators (e.g., AcrR of E. coli) of the efflux pumps fall into the TetR family of transcriptional repressors (Table I and II) (see reference[612] for a review). Crystallographic studies reveal AcrR (also CmeR of C. jejuni) as a dimeric two-domain molecule with an entirely helical architecture.[66, 613] The two-component systems EvgAS, PhoPQ and BaeSR also affect the expression of the E. coli exporters. The expression of emrKY, yhiUV, acrAB, mdfA and tolC is increased by the constitutive EvgS. PhoPQ further affects tolC expression as part of an interaction between EvgAS and PhoPQ.[1, 614] BaeSR induces expression of AcrD and MdtABC pumps in E. coli and S. Typhimurium.[168, 615] Indole, copper, or zinc (all in millimolar concentrations) induces these transporters, presumably by interacting with BaeSR.[168, 615] The repressor AcrS of AcrEF pump also represses AcrAB.[616] MdtEF (YhiUV) pump-mediated MDR is activated by AraC-XylS family regulators GadX[617] and YdeO[618] but repressed by the global regulator CRP,[619] which is involved in catabolite repression. Deletion of the E. coli hns gene, coding for the histone-like nucleoid-structuring protein H-NS, derepresses acrEF and mdtEF genes;[620] however, we are not aware of conditions that lower the expression level of H-NS.

P. aeruginosa contains 10 RND systems which have been characterized (Fig. 2).[1, 6] The complex regulation of these Mex pumps shown in Fig. 2 is used here for demonstrating the multiple levels of the efflux pump regulation. MexAB-OprM overexpression is typically associated with mutations in the linked mexR repressor gene as demonstrated in various mutants (earlier called nalB).[1, 621] The structure of MexR suggests effector-induced conformational changes for inhibiting DNA binding.[1] A recent study revealed that MexR is a redox-sensing regulator that senses peroxide stress to increase MexAB-OprM expression and drug resistance.[622]

Fig. 2.

Fig. 2

Regulation of the RND superfamily Mex transporters of P. aeruginosa. The efflux systems are shown in the light blue blocks with the respective transcriptional units presented in the solid grey lines. All regulators are shown in the green boxes, and their functions as repressor or activator are indicated, respectively, in the green or orange dotted arrows. The inverse relationship between MexAB-OprM expression and MexCD-OprJ/MexEF-OprN expression is marked by a double arrowed line. Interaction of the regulators (MexR or MexT) with the modulators (ArmR or MexT) is denoted by the double arrowed dotted lines. See text and relevant references for details of the regulation. MDA=membrane-damaging agents; QS AIs= quorum-sensing autoinducers.

In spite of the early identification of MexR, the MexAB-OprM regulation is more complicated, and also involves additional regulators/modulators such as NalC, NalD and AmrR (Fig. 2). NalC, encoded by PA3721, is a repressor of the TetR/AcrR family and negatively regulates the expression of the PA3720-PA3719 operon, located downstream of nalC. Thus, PA3720-PA3719 is overexpressed in nalC mutants. PA3719 encodes a protein modulator of only 53 amino acid residues, named AmrR, which functions as an anti-repressor that interacts with MexR and modulates MexR repressor activity.[623] The nalC mutants show modestly elevated expression of mexAB-oprM. The removal of AmrR decreases MexAB-OprM expression to wild-type levels and compromises MDR. These mutants also produce markedly elevated levels of MexR protein.[624] The crystal structure of MexR in complex with AmrR reveals the way the repressor activity is modulated.[625] Mutations in nalD (PA3574), which encodes a TetR familyrepressor,[612] are also responsiblefor mexAB-oprM overerexpression in some clinical isolates.[233, 240, 626] NalD binds to a second promoter upstreamof mexAB-oprM, directly repressing the effluxgene expression.[627]

An inverse relationship in expression between MexAB-OprM and other efflux pumps MexCD-OprJ or MexEF-OprN has been observed.[589, 628] This mechanism is also likely the cause of β-lactam hypersusceptibility in nfxC-type MexEF-OprN-overproducing mutants.[605] However, details of this regulatory mechanism(s) remain unknown. Increased mexAB-oprM expression is induced by N-butyryl homoserine lactone and this is repressed by MexT, a positive regulator of mexEF-oprN.[605] Mutations in mexS (PA2491 encoding a probable oxidoreductase) promote MexT-dependent mexEF-oprN expression and MDR in a clinical strain.[629] Inactivation of mexS resulted in up-regulation of the genes for efflux pumps (including MexCD-OprJ and MexEF-OprN), alginate synthesis and nitrate reduction as well as down-regulation of the genes for DNA replication, ribosome synthesis, virulence factor and lipopolysaccharide synthesis.[630] Macrolides such as azithromycin also reduce the expression of MexAB-OprM, possibly via the impact on the quorum sensing system.[607]

Expression of MexCD-OprJ is regulated by NfxB repressor.[1] This pump complex is induced by disinfectants and dyes but not by common antibiotics.[631] This induction was shown recently to involve an AlgU-dependent pathway.[573, 580] Membrane damaging-agents such as biocides, cationic antibacterial peptides, detergents and solvents disrupt the OM and/or cytoplasmic membrane by releasing the membrane lipid constituents, which signal the cytoplasmic-membrane-associated Muc proteins (homologues of Rse proteins of E. coli). The latter then activates AlgU (a homologue of E. coli Sigma E), a sigma factor that positively regulates mexCD-oprJ expression. Thus the physiological function of MexCD-OprJ may be the export of constituents from damaged membranes.[573] Moreover, the inactivation of the DNA oxidative repair system led to increased mutation frequency that also yielded nfxB mutations with MexCD-OprJ overproduction.[632] Overexpression of MexCD-OprJ in nfxB mutants decreases MexAB-OprM and MexXY expression.[589]

MexZ is a transcriptional repressor of the mexXY efflux operon, and purified MexZ shows specific binding with the mexZ-mexX intergenic site.[633] The mexXY operon is inducible by antibacterials targeting the ribosome.[578] The aberrant polypeptides produced and their oxidatively modified products (generated via reactive oxygen) interact with PA5471 in regulating MexXY.[573] Though the precise activity of PA5471 remains to be determined, disruption in gene PA5471 compromises the drug-inducible mexXY expression, and PA5471 itself is induced by the same ribosome-targeting agents that induce mexXY expression. PA5471 thus appears to modulate MexZ activity in affecting mexXY expression (Fig. 2).[579] The mexJK operon is constitutively expressed in mutants with defects in the upstream mexL gene. The MexL repressor regulates the expression of both mexL and mexJK.[634]

MvaT, a global regulator in P. aeruginosa, has been proposed as an H-NS-like protein involved in biofilm, quorum sensing and virulence.[635638] Deletion of mvaT resulted in increased resistance to chloramphenicol and norfloxacin but higher susceptibility to imipenem, and this was associated with increased expression of mexEF-oprN.[639]

The operons for other P. aeruginosa RND efflux systems such as MexHI-OpmD, MexMN-OprM, MexPQ-OpmE, MexVW-OprM and TriABC-OpmH are not associated with putative regulatory genes. How these pumps are regulated remains unknown. However, intriguingly, these efflux systems apparently have narrow or drug-specific substrate profiles, or their function as drug pumps had to be measured through the heterologous overexpression from plasmids (Table I).

Multiple mechanisms are also involved in regulation of MDR transporters in other Gram-negative bacteria. In most cases, local regulators encoded by the genes linked to the efflux genes are identified as shown in Tables I and II. The three RND efflux operons ttgABC, ttgDEF and ttgGHI of P. putida have, respectively, the adjacent repressor genes ttgR, ttgT and ttgV.[1, 640] TtgV and TtgT repressors bind with different affinities to the promoters of the RND efflux operons, and show a new model of regulation in cross-regulating TtgDEF and TtgGHI.[640, 641] In C. jejuni, CmeR functions as a transcriptional repressor for CmeABC by binding specifically to the inverted repeat sequences in the cmeABC promoter.[613, 642] In an enrofloxacin-selected MDR C. jejuni, a point mutation in the binding site of CmeR was responsible for the overproduction of CmeABC.[643]

Multiple regulatory pathways are involved in the high-level MDR in Salmonella.[644] In a highly invasive and MDR zoonotic pathogen S. Choleraesuis, gene for AcrR was inactivated by a stop codon insertion, resulting in the AcrAB overexpression for ciprofloxacin resistance.[645] Elevated expression of the MarA global activator was observed with increased levels of RND pumps, AcrB, AcrD, and AcrF, in posthterapy MDR S. Typhimurium.[646] Inactivation of marA impaired inducible MDR in S. Choleraesuis and the EPI Phe-Arg-β-naphthylamide reduced the MDR phenotype.[647] A MarA homologue, the global regulator Rma (RamA), which is not present in E. coli, is often overproduced in Salmonella spp. including MDR S. Enteritidis, S. Hadar, S. Paratyphi B and S. Typhimurium,[644, 648651] increasing the expression of AcrAB, AcrEF and MdtABC.[650, 651] Overexpression of AcrAB was also demonstrated in S. Typhimurium with the prolonged treatment with commercial disinfectants, although the isolates also exhibited reduced invasiveness.[652] The promoter region of macAB genes in S. Typhimurium harbours a binding site for the response regulator PhoP, which represses macAB transcription. PhoPQ is a major regulator of Salmonella virulence, thus indicating an inverse connection between a virulence determinant and a drug efflux system.[172]

MtrCDE of N. gonorrhoeae is repressed by MtrR repressor and activated by MtrA.[1, 653, 654] MtrR also negatively regulates FarR, a repressor involved in the regulation of FarAB, suggesting a coordinating mechanism for MtrCDE and FarAB expression.[655] In N. meningitidis, however, MtrCDE expression is regulated neither by MtrR nor MtrA. Instead, the MtrCDE-overproducing clinical isolates contain a unique insertion element, called Correia sequence, in the mtrCDE promoter region. A post-transcriptional regulation of the mtrCDE transcript by cleavage in the inverted repeat of the Correia element was also identified.[656] Expression of mtrCDE in gonococci is also inducible by membrane-acting hydrophobic antibacterial agents in a manner dependent on another envelope protein, MtrF. The mtrF expression is repressed not only by MtrR, but also by another repressor, MpeR, in an additive manner.[657, 658]

The regulation of MDR transporters in Gram-positive bacteria is exemplified by the staphylococcal pumps as presented in Fig. 3. Expression of NorA, NorB, NorC and AbcA pumps is affected by multiple regulators including MgrA (also called NorR or Rat) and NorG.[434, 659661] MgrA (multiple gene regulator) of the MarR family is a global regulator that controls autolysis, virulence, biofilm formation, and efflux pump expression.[659, 660, 662, 663] Overexpression of MgrA may either increase or repress norA expression based on the genetic background including, for example, presence or absence of the promoter region mutations of norA, i.e., flqB mutations that alone cause norA overexpression.[660, 662, 664] MgrA also augments expression of NorC, Tet38 and AbcA pumps (Table II and Fig. 3).[434] The function of MgrA on norA expression appears to require other regulators such as global regulators SarA[665] and Agr (accessory gene regulator).[660] The two-component regulatory system ArlR-ArsS, initially found to modulate autolytic activity in S. aureus, also affects norA expression.[1] Substrate exposure can also augment norA expression, likely via yet unidentified mediators.[441] NorG, a member of the GntR-like transcriptional regulator family, binds specifically to the promoters of the pump genes. MgrA is an indirect repressor for norB and a direct activator for abcA; NorG in contrast has an opposite effect on these pump genes.[434, 662] The regulation of the MATE pump MepA involves the repressor MepR, which is a substrate-responsive regulatory protein repressing both mepR and mepA expression.[442, 666] Single and multiple in vitro exposures to low concentrations of biocides and dyes generated S. aureus mutants overexpressing mepA and other pumps. In addition to regulatory protein mutations alterations in promoter regions were also found.[442] MepR binds the mepA operator as a dimer of dimers, but binds the mepR operator as a single dimer.[667] Regulation of the efflux pumps MdeA, SdrM and SepA remains unknown (Fig. 3).

Fig. 3.

Fig. 3

Regulation of multidrug or drug-specific efflux transporters of S. aureus. The efflux transporters are shown in colour blocks. All regulators are presented in the green boxes, and their functions as repressor or activator are indicated, respectively, by the green or orange dotted arrows. Unknown regulators are marked with a question mark (?) with the dotted grey lines linked to the relevant transporters. See text and relevant references for details of the regulation.

9.2 Phenotypic Induction of Drug Efflux Pump Expression

The expression of drug pumps is often subjected to induction by small molecules (e.g., antibiotics, biocides, bile salts and salicylate), including substrates of the pumps. The examples of such compounds or inducers are compiled in Table III. There are multiple mechanisms for the induction. A typical mechanism relies on the interaction of the particular inducers and the regulator proteins, as exemplified by the binding of multiple toxic agents with the BmrR or QacR repressors, which, respectively, impact on the expression of Bmr pump in B. subtilis[668] or QacA/QacB pumps of S. aureus.[669, 670] In the induction of P. aeruginosa MexCD-OprJ (and also MexXY),[578, 631] membrane-damaging agents act through AlgU-dependent pathway as described above.[573] A recent study examined the transcriptome response of P. aeruginosa to pentachlorophenol, a common environmental contaminant. Exposure to pentachlorophenol resulted in strong up-regulation of both MexAB-OprM and MexJK pumps and of the regulatory genes PA3720-PA3719 and PA3721, but the molecular mechanisms were not investigated.[671] The CzcRS two-component regulatory system is involved in heavy metal and carbapenem resistance in P. aeruginosa, and this resistance can be induced by zinc released from latex urinary catheters into urine.[672]

Table 3.

Compounds that induce expression of bacterial drug efflux pumps

Inducer Species Efflux pump References
Aminoglycosides P. aeruginosa MexXY [578, 579]
Benzoate B. fragilis, E. coli, K. pneumoniae Mar-associated pumps [673, 761]
Bile salts B. fragilis, Campylobacter spp., E. coli, Salmonella spp., V. cholerae AcrAB, Bme, CmeABC, VexAB, VexCD [566, 569, 762764]
Chloramphenicol B. cenocepacia, P. aeruginosa, P. putida RND pumps such as MexXY, TtgABC [274, 324, 578, 579]
Cytotoxic agents (ethidium bromide, rhodamine 6G and tetraphenylphosphonium chloride) and disinfectants (benzalkonium chloride and chlorhexidine) B. subtilits, E. coli, P. aeruginosa, S. aureus Blt, Bmr, MexCD-OprJ, MepA, NorA, QacA, QacB [1, 441, 442, 580, 631, 765]
Diazepam E. coli, K. pneumoniae Mar-associated pumps [761]
Ethanol E. coli AcrAB [566]
Fatty acids E. coli AcrAB [762]
Fluoroquinolones Salmonella spp., S. pneumoniae AcrAB, PatAB [483, 677]
Indole E. coli, Salmonella spp. AcrAB, AcrD, AcrEF, CusB, EmrK, MdtA, MdtE, MdtH [685, 764, 766]
Macrolides P. aeruginosa MexXY [578]
Phenolic acids (salicylic acid, t-cinnamic acid and benzoic acid) Erwinia chrysanthemi AcrAB, EmrAB [767]
Phytoalexins: naringenin and phloretin Erwinia amylovora AcrAB [596]
Salicylate B. cenocepacia, B. fragilis; E. coli, C. jejuni, C. coli, K. pneumoniae, M. tuberculosis, Salmonella spp., V. cholerae AcrAB, CeoAB-OpcM, CmeABC, VceCAB, Mar-regulated pumps; MgrA/SarRA-regulated pumps, and unidentified pumps [1, 210, 325, 673, 681, 761, 768770]
Salt (NaCl) A. baumannii, Chromohalobacter spp., E. coli HrdC associated pump(s), AcrAB and other RND pumps [566, 734, 771]
Tetracyclines E. coli, P. aeruginosa, P. putida Tet pumps, MexXY, TtgABC [274, 578, 579]

The drug pump BmeB expression in B. fragilis is induced by analgesics/antiseptics, detergents and disinfectants, but the mechanism is unknown.[673] Increased expression of RND pump genes, together with other changes in cell morphology, was seen upon exposure of B. fragilis to bile salts.[564] B. fragilis isolates from stool expressed more RND pumps than blood isolates, and withstood the bile salt stress better.[674] Bile acids are also implicated in the regulation of several V. cholerae RND pump genes.[209]

In enteric bacteria, some antibiotics induce efflux pump production through MarA, but the mechanism remains obscure except the salicylate binding (and inactivation) of MarR.[1] Recently, transketolase, an enzyme in the pentose phosphate pathway, was found to bind MarR specifically.[675] Since stresses often up-regulates the pentose phosphate pathway, this may mean a signalling pathway for the stress-induced overproduction of AcrAB-TolC through the MarA overproduction caused by the inactivation of MarR. SoxR, which is the repressor of another global regulator SoxS, is well-known to become inactivated by reactive oxygen radicals.[1] Rob, another global transcription activator, is larger than MarA and SoxS, and its activity is modulated directly by the binding of some AcrAB substrates.[1]

Microarray analysis revealed that exposure of S. Typhimurium to nalidixic acid at a subinhibitory concentration resulted in overexpression of 226 genes including efflux pump genes (e.g., acrA, emrA and tolC).[676] The overproduction of AcrAB-TolC and other proteins also occurred after the exposure of S. Typhimurium to ciprofloxacin.[677] Triclosan induced the expression of acrAB and marA in the biofilm cells of S. Typhimurium.[563] Nitric oxide decreased activity of fluoroquinolones via its activation of soxRS and marRAB regulons in S. Typhimurium.[678] RamR repressor controls RamA expression (and therefore AcrAB expression) in Salmonella, and most MDR clinical isolates had mutations in the ramA gene.[679]

Salicylate continues to demonstrate its impact on multiple gene expression including efflux pump genes (Table III). In S. aureus, salicylate induction down-regulates a multidrug pump repressor gene (mgrA) and sarR, which represses a gene (sarA) important for intrinsic resistance, likely representing a unique mechanism that allows S. aureus to resist antibacterial stress and toxicity. SarA also globally affects the expression of many virulence genes.[680, 681]

9.3 Growth-Dependent Expression of Drug Efflux Pumps

The expression of drug exporter genes can vary based on the phases of growth. The expression of mexAB-oprM from P. aeruginosa is increased in the stationary phase and enhanced by quorum-sensing autoinducers,[682] whereas the expression of the P. syringae mexAB-oprM and S. maltophilia smeDEF operons is maximal in early exponential phase.[683, 684] In E. coli, the expression of acrAB, emrAB, emrD, emrE, emrKY, mdfA, and ydgFE is relatively stable, but mdtEF expression is the highest at the late stationary phase. The latter effect is mediated by the stationary-phase sigma factor rpoS.[685] In a chemostat culture, acrAB expression in E. coli is affected by the growth rate. This regulation does not require RpoS.[686] Expression of both the B. subtilis efflux gene mdtP and its repressor gene mdtR decreases during the stationary phase.[408]

10. Efflux Pump Inhibitors

Most clinically used antibacterials were discovered between 1941 and 1968. Over the past four decades, there were only a few novel classes of antibacterials developed, i.e., the oxazolidinone linezolid, lipopeptide daptomycin, and a ketolide connected to a polar aromatic residue, platensimycin.[687, 688] Thus, development of novel antibacterial drugs has been challenged by the rapid emergence of bacterial resistance, especially MDR, as well as by the unwillingness of pharmaceutical companies.[11] Given the clinical significance of drug efflux pumps in pathogenic bacteria, exploration of EPIs has been under way and also a subject of reviews.[689699] Biochemical and structural elucidation of key efflux pumps of both prokaryotic and euokaryotic origins have also facilitated the mechanism-based design of EPIs.[700, 701]

Archetypal efflux pumps such as the E. coli AcrAB-TolC, the P. aeruginosa MexAB-OprM, and the S. aureus NorA have been used to screen and characterize the potential EPIs. EPIs of various sources have been investigated, including those derived from natural sources.[697, 702] With respect to MexAB-OprM-specific EPIs, compounds of synthetic pyridopyrimidine series have become available and these include a potential preclinical candidate quaternary ammonium analogue D13-9001, which potentiated the activity of levofloxacin and aztreonam against P. aeruginosa.[703707] Examples of various EPIs are shown in Table IV.

Table 4.

Bacterial efflux pump inhibitors

Inhibitors Efflux pump(s) targeted Antibacterials with activity enhanceda References
Gram-negative bacteria
Arylpiperazines: 1-(1-naphthylmethyl)-piperazine and others. RND pumps of A. baumannii, Citrobacter freundii, E. aerogenes, E. coli, K. pneumoniae, V. cholerae FQ, MA, TC [713, 719, 772774]
Carbonyl cyanide m-chlorophenylhydrazone (CCCP) Secondary transporters such as RND, MFS and MATE pumps MA [1]
Dipeptide amides (synthetic): Phe-Arg-β-naphthylamide (MC-207,110), MC-02,595, and MC-04,124 RND-type pumps of Gram-negative bacteria including Mex pumps of P. aeruginosa and AcrAB-TolC of E. coli FQ, MA, plant antimicrobials [1, 230, 356, 715, 718, 719, 773]
EA-371alpha and EA-371delta of Streptomyces MexAB-OprM of P. aeruginosa LF [1]
Extracts of Berberis aetnensis E. coli, P. aeruginosa, S. aureus pumps not reported CP [775]
Extracts of Commiphora molmol, Centella asiatica, Daucus carota, Citrus aurantium and Glycyrrhiza glabra AcrAB-TolC of E. coli CM, NA, TC [697]
Phenothiazines E. coli pumps MA [711]
Pyridopyrimidine series MexAB-OprM of P. aeruginosa AZ, FQ, [703707]
Quinoline derivatives AcrAB-TolC of E. aerogenes, E. coli, K. pneumoniae MA [181, 189, 693, 776]
Tetracycline analogues Tet pumps TC [1]
Thanatin Pumps of E. aerogenes and K. pneumoniae CM, NF, TC [777]
Gram-positive bacteria
3-aryl piperidines MepA and NorA of S. aureus FQ [778]
Baicalein (trihydroxy flavone) Tet(K) and unidentified pump(s)/mechanisms of S. aureus AP, CA, OX, TC [779]
Berberine NorA of S. aureus FQ [1]
Catechin gallates: epicatechin gallate and epigallocatechin gallate NorA and Tet(K) of S. aureus NF, TC [722, 780, 781]
Diterpenes: abietane (carnosic acid and carnosol), isopimarane and geranylgeranyl diterpenes Msr(A) and Tet(K) of S. aureus EM, TC [782, 783]
Diterpenes: ferruginol, pisiferol, 5-epipisiferol, formosanoxide, trans-communic acid, torulosal, the sesquiterpene oplopanonyl acetate and the germacrane 4 beta-hydroxygermacra-1(10)-5-diene NorA of S. aureus OX [784]
Extracts of Mezoneuron benthamianum and Securinega virosa S. aureus pumps EM, FQ, TC [785]
Extracts of Mirabilis jalapa Linn.: N-trans-feruloyl 4′-O-methyldopamine and synthetic N-trans-3,4-O-dimethylcaffeoyl tryptamine NorA of S. aureus NF [786]
Extracts of Punica granatum NorA of S. aureus AP, CM, GM, OX, TC [787]
Flavones: 5′-methoxyhydnocarpin NorA of S. aureus FQ [1, 718]
Flavonolignans: flavonoid tricin and silybin NorA of S. aureus BB, FQ [1, 788]
Fluoroquinolone derivatives MepA and NorA of S. aureus FQ [789]
GG918 (synthetic) Unidenitifed pump(s) of S. aureus (but known to target P-glycoprotein) FQ [790]
Grapefruit oil (coumarin, abergamottin epoxide and coumarin epoxide derivatives) S. aureus pump(s) not reported EB, NF [791]
Indoles: 2-aryl-5-nitro-1H-indoles NorA of S. aureus BB [792]
Indoles: 5-nitro-2-phenyl-1H-indole (INF55) and others NorA of S. aureus FQ [1, 717]
Kaempferol glycoside from Herissantia tiubae NorA of S. aureus EB, FQ [793]
Methoxylated flavones/isoflavones: chrysosplenol-D, chrysoplenetin, genistein, orobol and biochanin A, pterocarpan NorA of S. aureus BB, FQ [794796]
Oligosaccharides murucoidins and stoloniferin NorA of S. aureus NF [797]
Penta-substituted pyridine: 2,6-dimethyl-4-phenyl-pyridine-3,5-dicarboxylic acid diethyl ester MsrA of S. aureus FQ [709]
Phenothiazines: chlorpromazine and thioridazine Unknown but may be associated with NorA, Erm(A) and Erm(B) of S. aureus MA [697, 711, 798801]
Piperine: 1-piperoyl-piperidine and analogs NorA of S. aureus CP, EB, FQ [714, 802]
Piperidine alkaloids: julifloridine, juliflorine and juliprosine NorA of S. aureus FQ [702]
Polyacylated neohesperidosides NorA of S. aureus BB, FQ, RH [803]
Polyacylated oligosaccharides: orizabins NorA of S. aureus NF [804]
Porphyrin pheophorbide a NorA of S. aureus BB, FQ [788]
Pyrrolo [1,2-a] quinoxaline derivatives (omeprazole analogues; synthetic) NorA of S. aureus NF [805]
Reserpine (alkaloid) Bmr of B. subtilis, EfrAB of E. faecalis, NorA and Tet(K) of S. aureus, PmrA and PatAB of S. pneuomoniae FQ, TC [1, 463]
Resin glycosides of Ipomoea murucoides (murucoidins, pescaprein and stoloniferin) NorA of S. aureus FQ [797]
Spinosan A (arylbenzofuran aldehyde) NorA of S. aureus BB [796]
Stilbene (phenolic metabolite) NorA of S. aureus BB, EM, TC [708]
Totarol (phenolic diterpene) NorA of S. aureus EB, FQ [806]
Mycobacteria
CCCP M. tuberculosis SM [807]
Chlorpromazine M. avium EB, EM [808]
Dipeptide amide (synthetic): Phe-Arg-β-naphthylamide M. tuberculosis FQ [515]
Isoflavonoid (biochanin A), flavone (luteolin(and stilbene(resveratrol) M. smegmatis EB [809]
Reserpine M. smegmatis, M. tuberculosis EB, FQ [515, 809]
Thioridazine M. avium, M. tuberculosis EB, EM [732, 801, 808]
Verapamil M. avium, M. smegmatis, M. tuberculosis EB, EM, SM [807, 809]
a

AP=Ampicillin; AZ=aztreonam; BB=berberine; CA=cefmetazole; CM=chloramphenicol; CP=ciprofloxacin; FQ=fluoroquinolones; GM=gentamicin; LF=levofloxacin; MA=multiple antibacterials; NA=nalidixic acid; NF=norfloxacin; OX=oxacillin; RH=rhein; SM=streptomycin; TC=tetracyclines.

Some EPIs may be the substrates for the pumps they inhibit.[708] Some inhibitors may also target the MDR transporters of fungal and mammalian cells (e.g., the jatrophane diterpenoids[709, 710] and phenothiazines[711, 712]). The modes of action of some EPIs may not be limited to the inhibition of efflux pumps (in this case the term “resistance modulators” is more appropriate).[702] The EPI 1-(1-naphthylmethyl)-piperazine displays a paradoxical effect on A. baumannii isolate, where it unexpectedly decreased the susceptibility to tigecycline whereas the susceptibility to other tetracyclines was increased as expected.[713]

The combinational use of an EPI with antibacterial agents should potentiate the activity of antibacterials, and it would also reduce the frequency of emergence of resistant mutants.[1, 714, 715] For example, the presence of the EPI Phe-Arg-β-naphthylamide resulted in a up to 2,000-fold reduction in the minimum inhibitory concentrations of antibacterials known to be substrates of the Campylobacter CmeABC pump, and the frequency of emergence of erythromycin-resistant mutants in C. jejuni was reduced more than 1,000 fold.[715]

Some fluoroquinolone dimers remain active against NorA-overproducing S. aureus and do not inhibit ethidium efflux catalyzed by NorA.[716] Also, a hybrid between an EPI and a weak antibacterial, berberine, displayed an elevated antibacterial activity.[717] Several EPIs have also been usedin antibacterial photodynamic inactivation in combination with cationic phenothiazinium salts and light to enhancethe antibacterial activity.[718] Both 1-(1-naphthylmethyl)-piperazine and Phe-Arg-β-naphthylamide inhibit the production of the virulence factors cholera toxin and the toxin-coregulated pilus in V. cholerae.[719] Alkoxyquinoline derivatives (e.g., 2,8-dimethyl-4-(2′-pyrrolidinoethyl)-oxyquinoline) were able to inhibit antibacterial extrusion in E. aerogenes,[189] suggesting quinoline derivatives as promising efflux inhibitors for this species.[693] Among the clinical isolates of E. aerogenes, there was a noticeable increase in those containing an efflux mechanism susceptible to Phe-Arg-β-naphthylamide between 1995 and 2003.[720]

Understanding how EPIs block the transport of antibacterials is critical for designing and optimizing EPIs. Reserpine action is affected by the residues Phe143, Val286 and Ph306 of the Bmr pump, and these residues are also involved in determination of the substrate specificity.[721] NorA seems to have both high- and low-affinity binding sites to the phenolic metabolites catechin gallates, which paradoxically stimulated efflux at a lower concentration.[722] The differential impact of Phe-Arg-β-naphthylamide on the potentiation of carbencillin and levofloxacin/erythromycin, respectively, against MexAB-OprM-mediated resistance also may suggest the complexity of substrate recognition site.[8]

Finally, the susceptibility to efflux pump substrates in the presence and absence of an EPI has been used as a crude screen for the presence of efflux-based resistance mechanisms.[1, 439] The accuracy of reserpine, an EPI universally used for Gram-positive bacteria, in predicting pump gene overexpression was recently reassessed. The reserpine screen failed to identify many strains that overexpress one or more staphylococcal MDR pump genes, suggesting a need for development of an improved method.[723]

11. Conclusions

Bacteria have evolved sophisticated mechanisms of resistance including efficient drug efflux pumps that accommodate a wide range of substrates, both antibacterials and non-antibacterials. Efflux-mediated resistance can be clinically relevant and render antibacterial therapy ineffective. It also provides baseline resistance that helps the emergence of further resistance mechanisms such as drug inactivation or drug target modification. Thus it may be necessary, for the optimization of the pharmacokinetics and pharmacodynamics of antibacterial therapy, to take into account the activity (and its possible inhibition) of drug efflux pumps.[256, 724] Even the eukaryotic drug efflux pumps have been implicated in pharmacokinetics of antibacterials such that transport of fluoroquinolones is also mediated by mammalian transporters, which impact on the drug disposition or secretion.[725, 726]

The control of bacterial drug efflux pumps is a complex process with involvement of an intricate regulatory network that allows bacteria to sense and respond to a wide range of stress signals including, but not limited to, the presence of antibacterials. Such capability may require many genes, and thus usually a larger genome size.[727] In any case, the selection of efflux-pump overproducing strains depends on bacterial exposure to antibacterials, and limiting such exposure, including minimizing of antibacterial use, would limit the emergence of efflux-mediated drug resistance.[728, 729]

Structural and genetic studies have allowed the better understanding of the transport mechanisms of the efflux pumps, and these include the identification of amino acid residues or regions for rational design of drugs that may be able to evade efflux. Such agents or EPIs would be able to overcome the efflux-mediated resistance. In this regard, the activities of tigecycline against various pathogens are at least partly attributable to being an inferior substrate for specific Tet transporters. Interestingly, among the wide ranges of antibacterial substrates for bacterial MDR transporters, antibacterial peptides tend to be rather poor substrates, such that AcrAB, MexAB and NorA pumps do not confer resistance to several human antibacterial peptides,[730] although cases of pump-mediated resistance to such peptides are known.[332, 731] Significant efforts have been made to develop EPIs. EPIs are even considered in combating the XDR in M. tuberculosis[732]. However, it appears that none of the bacterial EPIs tested have ever entered into a clinical trial phase, although this may be the result of various factors, such as the high cost of running clinical trials for an EPI and then again for EPI-antibacterial combination.

The presence of MDR pumps in bacteria is certainly not just for drug resistance. However, understanding the physiological roles of the MDR pumps may continue to be rather difficult as the functions of these pumps are often involved in a complex, and overlapping network of reactions in the bacterial cell. In any case, in antibacterial therapy clearly we have an urgent need to overcome the negative effects caused by the MDR pumps. We hope that exciting new discoveries on these pumps will continue to arrive.

Acknowledgments

Research in the laboratory of H.N. has been supported by the U. S. Public Health Service (AI-09644). The views in this article do not necessarily reflect those of the X.-Z.L’s affiliation, Health Canada. Neither author has any disclosable interest relevant to the content of this article.

References

  • 1.Li X-Z, Nikaido H. Efflux-mediated drug resistance in bacteria. Drugs. 2004;64(2):159–204. doi: 10.2165/00003495-200464020-00004. [DOI] [PubMed] [Google Scholar]
  • 2.Poole K. Efflux-mediated antimicrobial resistance. J Antimicrob Chemother. 2005;56(1):20–51. doi: 10.1093/jac/dki171. [DOI] [PubMed] [Google Scholar]
  • 3.Alekshun MN, Levy SB. Molecular mechanisms of antibacterial multidrug resistance. Cell. 2007;128(6):1037–50. doi: 10.1016/j.cell.2007.03.004. [DOI] [PubMed] [Google Scholar]
  • 4.Higgins CF. Multiple molecular mechanisms for multidrug resistance transporters. Nature. 2007;446(7137):749–757. doi: 10.1038/nature05630. [DOI] [PubMed] [Google Scholar]
  • 5.Lubelski J, Konings WN, Driessen AJ. Distribution and physiology of ABC-type transporters contributing to multidrug resistance in bacteria. Microbiol Mol Biol Rev. 2007;71(3):463–76. doi: 10.1128/MMBR.00001-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Poole K. Efflux pumps as antimicrobial resistance mechanisms. Ann Med. 2007;39(3):162–76. doi: 10.1080/07853890701195262. [DOI] [PubMed] [Google Scholar]
  • 7.Lomovskaya O, Zgurskaya HI, Bostian KA, et al. Multidrug efflux pumps: structure, mechanism, and inhibition. In: Wax RG, Kim Lewis K, et al., editors. Bacterial resistance to antimicrobials. Boca Raton, Florida: CRC Press; 2008. pp. 45–70. [Google Scholar]
  • 8.Nikaido H, Takatsuka Y. Mechanisms of RND multidrug efflux pumps. Biochim Biophys Acta . 2009;1794(5):769–81. doi: 10.1016/j.bbapap.2008.10.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Pages JM, James CE, Winterhalter M. The porin and the permeating antibiotic: a selective diffusion barrier in Gram-negative bacteria. Nat Rev Microbiol. 2008;6(12):893–903. doi: 10.1038/nrmicro1994. [DOI] [PubMed] [Google Scholar]
  • 10.Nikaido H. Multidrug resistance in bacteria. Ann Rev Biochem. 2009;78:119–46. doi: 10.1146/annurev.biochem.78.082907.145923. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.The Royal Society London. Innovative mechanism tacking antibacterial resistance. London: 2008. [Last accessed on March 20, 2009]. http://royalsociety.org/document.asp?tip=0&id=7888. [Google Scholar]
  • 12.Jassal M, Bishai WR. Extensively drug-resistant tuberculosis. Lancet Infect Dis. 2009;9(1):19–30. doi: 10.1016/S1473-3099(08)70260-3. [DOI] [PubMed] [Google Scholar]
  • 13.Livermore DM. Minimising antibiotic resistance. Lancet Infect Dis. 2005;5(7):450–9. doi: 10.1016/S1473-3099(05)70166-3. [DOI] [PubMed] [Google Scholar]
  • 14.Mulvey MR, Boyd DA, Olson AB, et al. The genetics of Salmonella genomic island 1. Microbes Infect. 2006;8(7):1915–22. doi: 10.1016/j.micinf.2005.12.028. [DOI] [PubMed] [Google Scholar]
  • 15.Li X-Z. Antimicrobial resistance in Salmonella: features and mechanisms. In: Giordano LS, Moretti MA, editors. Salmonella infections: new research. Hauppauge, New York: Nova Science Publishers; 2008. pp. 1–43. [Google Scholar]
  • 16.Fournier PE, Vallenet D, Barbe V, et al. Comparative genomics of multidrug resistance in Acinetobacter baumannii. PLoS Genet. 2006;2(1):e7. doi: 10.1371/journal.pgen.0020007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Adams MD, Goglin K, Molyneaux N, et al. Comparative genome sequence analysis of multidrug-resistant Acinetobacter baumannii. J Bacteriol. 2008;190(24):8053–64. doi: 10.1128/JB.00834-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Livermore DM, Woodford N. The β-lactamase threat in Enterobacteriaceae, Pseudomonas and Acinetobacter. Trends Microbiol. 2006;14(9):413–20. doi: 10.1016/j.tim.2006.07.008. [DOI] [PubMed] [Google Scholar]
  • 19.Jacoby GA. AmpC β-lactamases. Clin Microbiol Rev. 2009;22(1):161–82. doi: 10.1128/CMR.00036-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Robicsek A, Strahilevitz J, Jacoby GA, et al. Fluoroquinolone-modifying enzyme: a new adaptation of a common aminoglycoside acetyltransferase. Nat Med. 2006;12(1):83–8. doi: 10.1038/nm1347. [DOI] [PubMed] [Google Scholar]
  • 21.Li X-Z. Quinolone resistance in bacteria: emphasis on plasmid-mediated mechanisms. Int J Antimicrob Agents. 2005;25(6):453–463. doi: 10.1016/j.ijantimicag.2005.04.002. [DOI] [PubMed] [Google Scholar]
  • 22.Robicsek A, Jacoby GA, Hooper DC. The worldwide emergence of plasmid-mediated quinolone resistance. Lancet Infect Dis. 2006;6(10):629–40. doi: 10.1016/S1473-3099(06)70599-0. [DOI] [PubMed] [Google Scholar]
  • 23.Yamane K, Wachino J, Suzuki S, et al. New plasmid-mediated fluoroquinolone efflux pump, QepA, found in an Escherichia coli clinical isolate. Antimicrob Agents Chemother. 2007;51(9):3354–60. doi: 10.1128/AAC.00339-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Piddock LJ. Clinically relevant chromosomally encoded multidrug resistance efflux pumps in bacteria. Clin Microbiol Rev. 2006;19(2):382–402. doi: 10.1128/CMR.19.2.382-402.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Nordmann P, Poirel L, Mak JK, et al. Multidrug-resistant Salmonella strains expressing emerging antibiotic resistance determinants. Clin Infect Dis. 2008;46(2):324–5. doi: 10.1086/524898. [DOI] [PubMed] [Google Scholar]
  • 26.Li X-Z, Mehrotra M, Ghimire S, et al. β-Lactam resistance and β-lactamases in bacteria of animal origin. Vet Microbiol. 2007;121(3–4):197–214. doi: 10.1016/j.vetmic.2007.01.015. [DOI] [PubMed] [Google Scholar]
  • 27.Aarestrup FM, Wegener HC, Collignon P. Resistance in bacteria of the food chain: epidemiology and control strategies. Expert Rev Anti Infect Ther. 2008;6(5):733–50. doi: 10.1586/14787210.6.5.733. [DOI] [PubMed] [Google Scholar]
  • 28.Weese SJ. Antimicrobial resistance in companion animals. Anim Health Res Rev. 2008;9(2):169–76. doi: 10.1017/S1466252308001485. [DOI] [PubMed] [Google Scholar]
  • 29.de Lencastre H, Oliveira D, Tomasz A. Antibiotic resistant Staphylococcus aureus: a paradigm of adaptive power. Curr Opin Microbiol. 2007;10(5):428–35. doi: 10.1016/j.mib.2007.08.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Wulf M, Voss A. MRSA in livestock animals-an epidemic waiting to happen? Clin Microbiol Infect. 2008;14(6):519–21. doi: 10.1111/j.1469-0691.2008.01970.x. [DOI] [PubMed] [Google Scholar]
  • 31.Pao SS, Paulsen IT, Saier MH., Jr Major facilitator superfamily. Microbiol Mol Biol Rev. 1998;62(1):1–34. doi: 10.1128/mmbr.62.1.1-34.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Kuroda T, Tsuchiya T. Multidrug efflux transporters in the MATE family. Biochim Biophys Acta. 2009;1794(5):763–8. doi: 10.1016/j.bbapap.2008.11.012. [DOI] [PubMed] [Google Scholar]
  • 33.Jack DL, Yang NM, Saier MH., Jr The drug/metabolite transporter superfamily. Eur J Biochem. 2001;268(13):3620–39. doi: 10.1046/j.1432-1327.2001.02265.x. [DOI] [PubMed] [Google Scholar]
  • 34.Tseng TT, Gratwick KS, Kollman J, et al. The RND permease superfamily: an ancient, ubiquitous and diverse family that includes human disease and development proteins. J Mol Microbiol Biotechnol. 1999;1(1):107–25. [PubMed] [Google Scholar]
  • 35.Seeger MA, Diederichs K, Eicher T, et al. The AcrB efflux pump: conformational cycling and peristalsis lead to multidrug resistance. Curr Drug Targets. 2008;9(9):729–49. doi: 10.2174/138945008785747789. [DOI] [PubMed] [Google Scholar]
  • 36.Altmann SW, Davis HR, Jr, Zhu LJ, et al. Niemann-Pick C1 Like 1 protein is critical for intestinal cholesterol absorption. Science. 2004;303(5661):1201–4. doi: 10.1126/science.1093131. [DOI] [PubMed] [Google Scholar]
  • 37.Murakami S. Multidrug efflux transporter, AcrB-the pumping mechanism. Curr Opin Struct Biol. 2008;18(4):459–65. doi: 10.1016/j.sbi.2008.06.007. [DOI] [PubMed] [Google Scholar]
  • 38.Murakami S, Yamaguchi A. Multidrug-exporting secondary transporters. Curr Opin Struct Biol. 2003;13(4):443–52. doi: 10.1016/s0959-440x(03)00109-x. [DOI] [PubMed] [Google Scholar]
  • 39.Yu EW, Aires JR, McDermott G, et al. A periplasmic drug-binding site of the AcrB multidrug efflux pump: a crystallographic and site-directed mutagenesis study. J Bacteriol. 2005;187(19):6804–15. doi: 10.1128/JB.187.19.6804-6815.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Drew D, Klepsch MM, Newstead S, et al. The structure of the efflux pump AcrB in complex with bile acid. Mol Membr Biol. 2008;25(8):677–82. doi: 10.1080/09687680802552257. [DOI] [PubMed] [Google Scholar]
  • 41.Yu EW, Aires JR, Nikaido H. AcrB multidrug efflux pump of Escherichia coli: composite substrate-binding cavity of exceptional flexibility generates its extremely wide substrate specificity. J Bacteriol. 2003;185(19):5657–64. doi: 10.1128/JB.185.19.5657-5664.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Tornroth-Horsefield S, Gourdon P, et al. Crystal structure of AcrB in complex with a single transmembrane subunit reveals another twist. Structure. 2007;15(12):1663–73. doi: 10.1016/j.str.2007.09.023. [DOI] [PubMed] [Google Scholar]
  • 43.Murakami S, Tamura N, Saito A, et al. Extramembrane central pore of multidrug exporter AcrB in Escherichia coli plays an important role in drug transport. J Biol Chem. 2004;279(5):3743–8. doi: 10.1074/jbc.M308893200. [DOI] [PubMed] [Google Scholar]
  • 44.Middlemiss JK, Poole K. Differential impact of MexB mutations on substrate selectivity of the MexAB-OprM multidrug efflux pump of Pseudomonas aeruginosa. J Bacteriol. 2004;186(5):1258–69. doi: 10.1128/JB.186.5.1258-1269.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Takatsuka Y, Nikaido H. Threonine-978 in the transmembrane segment of the multidrug efflux pump AcrB of Escherichia coli is crucial for drug transport as a probable component of the proton relay network. J Bacteriol. 2006;188(20):7284–9. doi: 10.1128/JB.00683-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Su CC, Li M, Gu R, et al. Conformation of the AcrB multidrug efflux pump in mutants of the putative proton relay pathway. J Bacteriol. 2006;188(20):7290–6. doi: 10.1128/JB.00684-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Bohnert JA, Schuster S, Fahnrich E, et al. Altered spectrum of multidrug resistance associated with a single point mutation in the Escherichia coli RND-type MDR efflux pump YhiV (MdtF) J Antimicrob Chemother. 2007;59(6):1216–22. doi: 10.1093/jac/dkl426. [DOI] [PubMed] [Google Scholar]
  • 48.Das D, Xu QS, Lee JY, et al. Crystal structure of the multidrug efflux transporter AcrB at 3.1A resolution reveals the N-terminal region with conserved amino acids. J Struct Biol. 2007;158(3):494–502. doi: 10.1016/j.jsb.2006.12.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Dastidar V, Mao W, Lomovskaya O, et al. Drug-induced conformational changes in multidrug efflux transporter AcrB from Haemophilus influenzae. J Bacteriol. 2007;189(15):5550–8. doi: 10.1128/JB.00471-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Bohnert JA, Schuster S, Seeger MA, et al. Site-directed mutagenesis reveals putative substrate binding residues in the Escherichia coli RND efflux pump AcrB. J Bacteriol. 2008;190(24):8225–9. doi: 10.1128/JB.00912-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Wehmeier C, Schuster S, Fahnrich E, et al. Site-directed mutagenesis reveals amino acid residues in the Escherichia coli RND efflux pump AcrB that confer macrolide resistance. Antimicrob Agents Chemother. 2009;53(1):329–30. doi: 10.1128/AAC.00921-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Seeger MA, von Ballmoos C, Verrey F, et al. Crucial role of Asp408 in the proton translocation pathway of multidrug transporter AcrB: evidence from site-directed mutagenesis and carbodiimide labeling. Biochemistry. 2009 doi: 10.1021/bi900446j. e-published May 28, 2009. [DOI] [PubMed] [Google Scholar]
  • 53.Aires JR, Nikaido H. Aminoglycosides are captured from both periplasm and cytoplasm by the AcrD multidrug efflux transporter of Escherichia coli. J Bacteriol. 2005;187(6):1923–9. doi: 10.1128/JB.187.6.1923-1929.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Li X-Z, Ma D, Livermore DM, et al. Role of efflux pump(s) in intrinsic resistance of Pseudomonas aeruginosa: active efflux as a contributing factor to β-lactam resistance. Antimicrob Agents Chemother. 1994;38(8):1742–52. doi: 10.1128/aac.38.8.1742. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Murakami S, Nakashima R, Yamashita E, et al. Crystal structures of a multidrug transporter reveal a functionally rotating mechanism. Nature. 2006;443(7108):173–9. doi: 10.1038/nature05076. [DOI] [PubMed] [Google Scholar]
  • 56.Seeger MA, Schiefner A, Eicher T, et al. Structural asymmetry of AcrB trimer suggests a peristaltic pump mechanism. Science. 2006;313(5791):1295–8. doi: 10.1126/science.1131542. [DOI] [PubMed] [Google Scholar]
  • 57.Sennhauser G, Amstutz P, Briand C, et al. Drug export pathway of multidrug exporter AcrB revealed by DARPin inhibitors. PLoS Biol. 2007;5(1):e7. doi: 10.1371/journal.pbio.0050007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Sennhauser G, Bukowska MA, Briand C, et al. Crystal structure of the multidrug exporter MexB from Pseudomonas aeruginosa. J Mol Biol. 2009;389(1):134–45. doi: 10.1016/j.jmb.2009.04.001. [DOI] [PubMed] [Google Scholar]
  • 59.Takatsuka Y, Nikaido H. Site-directed disulfide cross-linking shows that cleft flexibility in the periplasmic domain is needed for the multidrug efflux pump AcrB of Escherichia coli. J Bacteriol. 2007;189(23):8677–84. doi: 10.1128/JB.01127-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Seeger MA, von Ballmoos C, Eicher T, et al. Engineered disulfide bonds support the functional rotation mechanism of multidrug efflux pump AcrB. Nat Struct Mol Biol. 2008;15(2):199–205. doi: 10.1038/nsmb.1379. [DOI] [PubMed] [Google Scholar]
  • 61.Takatsuka Y, Nikaido H. Covalently linked trimer of the AcrB multidrug efflux pump provides support for the functional rotating mechanism. J Bacteriol. 2009;191(6):1729–37. doi: 10.1128/JB.01441-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Zgurskaya HI. Covalently linked AcrB giant offers a new powerful tool for mechanistic analysis of multidrug efflux in bacteria. J Bacteriol. 2009;191(6):1727–8. doi: 10.1128/JB.01718-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Su CC, Yu EW. Ligand-transporter interaction in the AcrB multidrug efflux pump determined by fluorescence polarization assay. FEBS Lett. 2007;581(25):4972–6. doi: 10.1016/j.febslet.2007.09.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Nagano K, Nikaido H. Kinetic behavior of the major multidrug efflux pump AcrB of Escherichia coli. Proc Natl Acad Sci USA. 2009;106(14):5854–8. doi: 10.1073/pnas.0901695106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Alguel Y, Meng C, Teran W, et al. Crystal structures of multidrug binding protein TtgR in complex with antibiotics and plant antimicrobials. J Mol Biol. 2007;369(3):829–40. doi: 10.1016/j.jmb.2007.03.062. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Li M, Gu R, Su CC, et al. Crystal structure of the transcriptional regulator AcrR from Escherichia coli. J Mol Biol. 2007;374(3):591–603. doi: 10.1016/j.jmb.2007.09.064. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Lee A, Mao W, Warren MS, et al. Interplay between efflux pumps may provide either additive or multiplicative effects on drug resistance. J Bacteriol. 2000;182(11):3142–50. doi: 10.1128/jb.182.11.3142-3150.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Abramson J, Smirnova I, Kasho V, et al. Structure and mechanism of the lactose permease of Escherichia coli. Science. 2003;301(5633):610–5. doi: 10.1126/science.1088196. [DOI] [PubMed] [Google Scholar]
  • 69.Huang Y, Lemieux MJ, Song J, et al. Structure and mechanism of the glycerol-3-phosphate transporter from Escherichia coli. Science. 2003;301(5633):616–20. doi: 10.1126/science.1087619. [DOI] [PubMed] [Google Scholar]
  • 70.Yin Y, He X, Szewczyk P, et al. Structure of the multidrug transporter EmrD from Escherichia coli. Science. 2006;312(5774):741–4. doi: 10.1126/science.1125629. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Law CJ, Maloney PC, Wang DN. Ins and outs of major facilitator superfamily antiporters. Annu Rev Microbiol . 2008;62:289–305. doi: 10.1146/annurev.micro.61.080706.093329. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Nishino K, Yamaguchi A. Analysis of a complete library of putative drug transporter genes in Escherichia coli. J Bacteriol. 2001;183(20):5803–12. doi: 10.1128/JB.183.20.5803-5812.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Sigal N, Lewinson O, Wolf SG, et al. E. coli multidrug transporter MdfA is a monomer. Biochemistry. 2007;46(17):5200–8. doi: 10.1021/bi602405w. [DOI] [PubMed] [Google Scholar]
  • 74.Fluman N, Bibi E. Bacterial multidrug transport through the lens of the major facilitator superfamily. Biochim Biophys Acta. 2009;1794(5):738–47. doi: 10.1016/j.bbapap.2008.11.020. [DOI] [PubMed] [Google Scholar]
  • 75.Mazurkiewicz P, Poelarends GJ, Driessen AJ, et al. Facilitated drug influx by an energy-uncoupled secondary multidrug transporter. J Biol Chem. 2004;279(1):103–8. doi: 10.1074/jbc.M306579200. [DOI] [PubMed] [Google Scholar]
  • 76.Hassan KA, Souhani T, Skurray RA, et al. Analysis of tryptophan residues in the staphylococcal multidrug transporter QacA reveals long-distance functional associations of residues on opposite sides of the membrane. J Bacteriol. 2008;190(7):2441–9. doi: 10.1128/JB.01864-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Tanabe M, Szakonyi G, Brown KA, et al. The multidrug resistance efflux complex, EmrAB from Escherichia coli forms a dimer in vitro. Biochem Biophys Res Commun. 2009;380(2):338–42. doi: 10.1016/j.bbrc.2009.01.081. [DOI] [PubMed] [Google Scholar]
  • 78.Omote H, Hiasa M, Matsumoto T, et al. The MATE proteins as fundamental transporters of metabolic and xenobiotic organic cations. Trends Pharmacol Sci. 2006;27(11):587–93. doi: 10.1016/j.tips.2006.09.001. [DOI] [PubMed] [Google Scholar]
  • 79.Otsuka M, Matsumoto T, Morimoto R, et al. A human transporter protein that mediates the final excretion step for toxic organic cations. Proc Natl Acad Sci USA. 2005;102(50):17923–8. doi: 10.1073/pnas.0506483102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Matsumoto T, Kanamoto T, Otsuka M, et al. Role of glutamate residues in substrate recognition by human MATE1 polyspecific H+/organic cation exporter. Am J Physiol Cell Physiol. 2008;294(4):C1074–8. doi: 10.1152/ajpcell.00504.2007. [DOI] [PubMed] [Google Scholar]
  • 81.Hiasa M, Matsumoto T, Komatsu T, et al. Functional characterization of testis-specific rodent multidrug and toxic compound extrusion 2, a class III MATE-type polyspecific H+/organic cation exporter. Am J Physiol Cell Physiol. 2007;293(5):C1437–44. doi: 10.1152/ajpcell.00280.2007. [DOI] [PubMed] [Google Scholar]
  • 82.Bay DC, Rommens KL, Turner RJ. Small multidrug resistance proteins: a multidrug transporter family that continues to grow. Biochim Biophys Acta. 2008;1778(9):1814–38. doi: 10.1016/j.bbamem.2007.08.015. [DOI] [PubMed] [Google Scholar]
  • 83.Li X-Z, Poole K, Nikaido H. Contributions of MexAB-OprM and an EmrE homolog to intrinsic resistance of Pseudomonas aeruginosa to aminoglycosides and dyes. Antimicrob Agents Chemother. 2003;47(1):27–33. doi: 10.1128/AAC.47.1.27-33.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Schuldiner S. EmrE, a model for studying evolution and mechanism of ion-coupled transporters. Biochim Biophys Acta. 2009;1794(5):748–62. doi: 10.1016/j.bbapap.2008.12.018. [DOI] [PubMed] [Google Scholar]
  • 85.Tal N, Schuldiner S. A coordinated network of transporters with overlapping specificities provides a robust survival strategy. Proc Natl Acad Sci USA. 2009 doi: 10.1073/pnas.0902400106. e-published May 18, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Schuldiner S. When biochemistry meets structural biology: the cautionary tale of EmrE. Trends Biochem Sci. 2007;32(6):252–8. doi: 10.1016/j.tibs.2007.04.002. [DOI] [PubMed] [Google Scholar]
  • 87.Fleishman SJ, Harrington SE, Enosh A, et al. Quasi-symmetry in the cryo-EM structure of EmrE provides the key to modeling its transmembrane domain. J Mol Biol. 2006;364(1):54–67. doi: 10.1016/j.jmb.2006.08.072. [DOI] [PubMed] [Google Scholar]
  • 88.Chen YJ, Pornillos O, Lieu S, et al. X-ray structure of EmrE supports dual topology model. Proc Natl Acad Sci USA. 2007;104(48):18999–9004. doi: 10.1073/pnas.0709387104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Rapp M, Seppala S, Granseth E, et al. Emulating membrane protein evolution by rational design. Science. 2007;315(5816):1282–4. doi: 10.1126/science.1135406. [DOI] [PubMed] [Google Scholar]
  • 90.Kikukawa T, Nara T, Araiso T, et al. Two-component bacterial multidrug transporter, EbrAB: mutations making each component solely functional. Biochim Biophys Acta. 2006;1758(5):673–9. doi: 10.1016/j.bbamem.2006.04.004. [DOI] [PubMed] [Google Scholar]
  • 91.Kikukawa T, Miyauchi S, Araiso T, et al. Anti-parallel membrane topology of two components of EbrAB, a multidrug transporter. Biochem Biophys Res Commun. 2007;358(4):1071–5. doi: 10.1016/j.bbrc.2007.05.032. [DOI] [PubMed] [Google Scholar]
  • 92.Steiner-Mordoch S, Soskine M, Solomon D, et al. Parallel topology of genetically fused EmrE homodimers. EMBO J. 2008;27(1):17–26. doi: 10.1038/sj.emboj.7601951. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Korkhov VM, Tate CG. An emerging consensus for the structure of EmrE. Acta Crystallogr D Biol Crystallogr. 2009;65(2):186–92. doi: 10.1107/S0907444908036640. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Poulsen BE, Rath A, Deber CM. The assembly motif of a bacterial small multidrug resistance protein. J Biol Chem. 2009;284(15):9870–5. doi: 10.1074/jbc.M900182200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Dawson RJP, Locher KP. Structure of a bacterial multidrug ABC transporter. Nature. 2006;443(7108):180–5. doi: 10.1038/nature05155. [DOI] [PubMed] [Google Scholar]
  • 96.Davidson AL, Chen J. ATP-binding cassette transporters in bacteria. Annu Rev Biochem. 2004;73:241–68. doi: 10.1146/annurev.biochem.73.011303.073626. [DOI] [PubMed] [Google Scholar]
  • 97.Hollenstein K, Dawson RJ, Locher KP. Structure and mechanism of ABC transporter proteins. Curr Opin Struct Biol. 2007;17(4):412–8. doi: 10.1016/j.sbi.2007.07.003. [DOI] [PubMed] [Google Scholar]
  • 98.Schuldiner S. Structural biology: the ins and outs of drug transport. Nature. 2006;443(7108):156–7. doi: 10.1038/443156b. [DOI] [PubMed] [Google Scholar]
  • 99.Kim SH, Chang AB, Saier MH., Jr Sequence similarity between multidrug resistance efflux pumps of the ABC and RND superfamilies. Microbiology. 2004;150(Pt 8):2493–5. doi: 10.1099/mic.0.27312-0. [DOI] [PubMed] [Google Scholar]
  • 100.Ward A, Reyes CL, Yu J, et al. Flexibility in the ABC transporter MsbA: alternating access with a twist. Proc Natl Acad Sci USA. 2007;104(48):19005–10. doi: 10.1073/pnas.0709388104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Velamakanni S, Yao Y, Gutmann DA, et al. Multidrug transport by the ABC transporter Sav1866 from Staphylococcus aureus. Biochemistry. 2008;47(35):9300–8. doi: 10.1021/bi8006737. [DOI] [PubMed] [Google Scholar]
  • 102.Venter H, Shilling RA, Velamakanni S, et al. An ABC transporter with a secondary-active multidrug translocator domain. Nature. 2003;426(6968):866–70. doi: 10.1038/nature02173. [DOI] [PubMed] [Google Scholar]
  • 103.Venter H, Velamakanni S, Balakrishnan L, van Veen HW. On the energy-dependence of Hoechst 33342 transport by the ABC transporter LmrA. Biochem Pharmacol. 2008;75(4):866–74. doi: 10.1016/j.bcp.2007.10.022. [DOI] [PubMed] [Google Scholar]
  • 104.Zgurskaya HI, Yamada Y, Tikhonova EB, et al. Structural and functional diversity of bacterial membrane fusion proteins. Biochim Biophys Acta. 2009;1794(5):794–807. doi: 10.1016/j.bbapap.2008.10.010. [DOI] [PubMed] [Google Scholar]
  • 105.Akama H, Matsuura T, Kashiwagi S, et al. Crystal structure of the membrane fusion protein, MexA, of the multidrug transporter in Pseudomonas aeruginosa. J Biol Chem. 2004;279(25):25939–42. doi: 10.1074/jbc.C400164200. [DOI] [PubMed] [Google Scholar]
  • 106.Higgins MK, Bokma E, Koronakis E, et al. Structure of the periplasmic component of a bacterial drug efflux pump. Proc Natl Acad Sci USA. 2004;101(27):9994–9. doi: 10.1073/pnas.0400375101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Mikolosko J, Bobyk K, Zgurskaya HI, et al. Conformational flexibility in the multidrug efflux system protein AcrA. Structure. 2006;14(3):577–87. doi: 10.1016/j.str.2005.11.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Symmons MF, Bokma E, Koronakis E, et al. The assembled structure of a complete tripartite bacterial multidrug efflux pump. Proc Natl Acad Sci USA. 2009;106(17):7173–8. doi: 10.1073/pnas.0900693106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Ge Q, Yamada Y, Zgurskaya H. The C-terminal domain of AcrA is essential for the assembly and function of the multidrug efflux pump AcrAB-TolC. J Bacteriol. 2009 doi: 10.1128/JB.00204-09. e-published May 1, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Yum S, Xu Y, Piao S, et al. Crystal structure of the periplasmic component of a tripartite macrolide-specific efflux pump. J Mol Biol. 2009;387(5):1286–97. doi: 10.1016/j.jmb.2009.02.048. [DOI] [PubMed] [Google Scholar]
  • 111.Ip H, Stratton K, Zgurskaya H, et al. pH-induced conformational changes of AcrA, the membrane fusion protein of Escherichia coli multidrug efflux system. J Biol Chem. 2003;278(50):50474–82. doi: 10.1074/jbc.M305152200. [DOI] [PubMed] [Google Scholar]
  • 112.Vaccaro L, Koronakis V, Sansom MS. Flexibility in a drug transport accessory protein: molecular dynamics simulations of MexA. Biophys J. 2006;91(2):558–64. doi: 10.1529/biophysj.105.080010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Touze T, Eswaran J, Bokma E, et al. Interactions underlying assembly of the Escherichia coli AcrAB-TolC multidrug efflux system. Mol Microbiol. 2004;53(2):697–706. doi: 10.1111/j.1365-2958.2004.04158.x. [DOI] [PubMed] [Google Scholar]
  • 114.Mokhonov VV, Mokhonova EI, Akama H, et al. Role of the membrane fusion protein in the assembly of resistance-nodulation-cell division multidrug efflux pump in Pseudomonas aeruginosa. Biochem Biophys Res Commun. 2004;322(2):483–9. doi: 10.1016/j.bbrc.2004.07.140. [DOI] [PubMed] [Google Scholar]
  • 115.Nehme D, Li X-Z, Elliot R, et al. Assembly of the MexAB-OprM multidrug efflux system of Pseudomonas aeruginosa: identification and characterization of mutations in mexA compromising MexA multimerization and interaction with MexB. J Bacteriol. 2004;186(10):2973–83. doi: 10.1128/JB.186.10.2973-2983.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Eda S, Maseda H, Yoshihara E, et al. Assignment of the outer-membrane-subunit-selective domain of the membrane fusion protein in the tripartite xenobiotic efflux pump of Pseudomonas aeruginosa. FEMS Microbiol Lett. 2006;254(1):101–7. doi: 10.1111/j.1574-6968.2005.00010.x. [DOI] [PubMed] [Google Scholar]
  • 117.Stegmeier JF, Polleichtner G, Brandes N, et al. Importance of the adaptor (membrane fusion) protein hairpin domain for the functionality of multidrug efflux pumps. Biochemistry. 2006;45(34):10303–12. doi: 10.1021/bi060320g. [DOI] [PubMed] [Google Scholar]
  • 118.Nehme D, Poole K. Assembly of the MexAB-OprM multidrug pump of Pseudomonas aeruginosa: component interactions defined by the study of pump mutant suppressors. J Bacteriol. 2007;189(17):6118–27. doi: 10.1128/JB.00718-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Bavro VN, Pietras Z, Furnham N, et al. Assembly and channel opening in a bacterial drug efflux machine. Mol Cell. 2008;30(1):114–21. doi: 10.1016/j.molcel.2008.02.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Elkins CA, Nikaido H. Chimeric analysis of AcrA function reveals the importance of its C-terminal domain in its interaction with the AcrB multidrug efflux pump. J Bacteriol. 2003;185(18):5349–56. doi: 10.1128/JB.185.18.5349-5356.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Nehme D, Poole K. Interaction of the MexA and MexB components of the MexAB-OprM multidrug efflux system of Pseudomonas aeruginosa: identification of MexA extragenic suppressors of a T578I mutation in MexB. Antimicrob Agents Chemother. 2005;49(10):4375–8. doi: 10.1128/AAC.49.10.4375-4378.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Krishnamoorthy G, Tikhonova EB, Zgurskaya HI. Fitting periplasmic membrane fusion proteins to inner membrane transporters: mutations that enable Escherichia coli AcrA to function with Pseudomonas aeruginosa MexB. J Bacteriol. 2008;190(2):691–8. doi: 10.1128/JB.01276-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Mima T, Joshi S, Gomez-Escalada M, et al. Identification and characterization of TriABC-OpmH, a triclosan efflux pump of Pseudomonas aeruginosa requiring two membrane fusion proteins. J Bacteriol. 2007;189(21):7600–9. doi: 10.1128/JB.00850-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Zgurskaya HI, Nikaido H. Bypassing the periplasm: reconstitution of the AcrAB multidrug efflux pump of Escherichia coli. Proc Natl Acad Sci USA. 1999;96(13):7190–5. doi: 10.1073/pnas.96.13.7190. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Nikaido H. Molecular basis of bacterial outer membrane permeability revisited. Microbiol Mol Biol Rev. 2003;67(4):593–656. doi: 10.1128/MMBR.67.4.593-656.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Tatsumi R, Wachi M. TolC-dependent exclusion of porphyrins in Escherichia coli. J Bacteriol. 2008;190(18):6228–33. doi: 10.1128/JB.00595-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Akama H, Kanemaki M, Yoshimura M, et al. Crystal structure of the drug discharge outer membrane protein, OprM, of Pseudomonas aeruginosa: dual modes of membrane anchoring and occluded cavity end. J Biol Chem. 2004;279(51):52816–9. doi: 10.1074/jbc.C400445200. [DOI] [PubMed] [Google Scholar]
  • 128.Federici L, Du D, Walas F, et al. The crystal structure of the outer membrane protein VceC from the bacterial pathogen Vibrio cholerae at 1.8 Å resolution. J Biol Chem. 2005;280(15):15307–14. doi: 10.1074/jbc.M500401200. [DOI] [PubMed] [Google Scholar]
  • 129.Li X-Z, Poole K. Mutational analysis of the OprM outer membrane component of the MexA-MexB-OprM multidrug efflux system of Pseudomonas aeruginosa. J Bacteriol. 2001;183(1):12–27. doi: 10.1128/JB.183.1.12-27.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Yoshihara E, Eda S. Diversity in the oligomeric channel structure of the multidrug efflux pumps in Pseudomonas aeruginosa. Microbiol Immunol. 2007;51(1):47–52. doi: 10.1111/j.1348-0421.2007.tb03889.x. [DOI] [PubMed] [Google Scholar]
  • 131.Gerken H, Misra R. Genetic evidence for functional interactions between TolC and AcrA proteins of a major antibiotic efflux pump of Escherichia coli. Mol Microbiol. 2004;54(3):620–31. doi: 10.1111/j.1365-2958.2004.04301.x. [DOI] [PubMed] [Google Scholar]
  • 132.Husain F, Humbard M, Misra R. Interaction between the TolC and AcrA proteins of a multidrug efflux system of Escherichia coli. J Bacteriol. 2004;186(24):8533–6. doi: 10.1128/JB.186.24.8533-8536.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Lobedanz S, Bokma E, Symmons MF, et al. A periplasmic coiled-coil interface underlying TolC recruitment and the assembly of bacterial drug efflux pumps. Proc Natl Acad Sci USA. 2007;104(11):4612–7. doi: 10.1073/pnas.0610160104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Tikhonova EB, Zgurskaya HI. AcrA, AcrB, and TolC of Escherichia coli form a stable intermembrane multidrug efflux complex. J Biol Chem. 2004;279(31):32116–24. doi: 10.1074/jbc.M402230200. [DOI] [PubMed] [Google Scholar]
  • 135.Tamura N, Murakami S, Oyama Y, et al. Direct interaction of multidrug efflux transporter AcrB and outer membrane channel TolC detected via site-directed disulfide cross-linking. Biochemistry. 2005;44(33):11115–21. doi: 10.1021/bi050452u. [DOI] [PubMed] [Google Scholar]
  • 136.Eswaran J, Koronakis E, Higgins MK, et al. Three’s company: component structures bring a closer view of tripartite drug efflux pumps. Curr Opin Struct Biol. 2004;14(6):741–7. doi: 10.1016/j.sbi.2004.10.003. [DOI] [PubMed] [Google Scholar]
  • 137.Misra R, Bavro VN. Assembly and transport mechanism of tripartite drug efflux systems. Biochim Biophys Acta. 2009;1794(5):817–25. doi: 10.1016/j.bbapap.2009.02.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Reffay M, Gambin Y, Benabdelhak H, et al. Tracking membrane protein association in model membranes. PLoS ONE. 2009;4(4):e5035. doi: 10.1371/journal.pone.0005035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Bokma E, Koronakis E, Lobedanz S, et al. Directed evolution of a bacterial efflux pump: adaptation of the E. coli TolC exit duct to the Pseudomonas MexAB translocase. FEBS Lett. 2006;580(22):5339–43. doi: 10.1016/j.febslet.2006.09.005. [DOI] [PubMed] [Google Scholar]
  • 140.Vediyappan G, Borisova T, Fralick JA. Isolation and characterization of VceC gain-of-function mutants that can function with the AcrAB multiple-drug-resistant efflux pump of Escherichia coli. J Bacteriol. 2006;188(11):3757–62. doi: 10.1128/JB.00038-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Polleichtner G, Andersen C. The channel-tunnel HI1462 of Haemophilus influenzae reveals differences to Escherichia coli TolC. Microbiology. 2006;152(Pt 6):1639–47. doi: 10.1099/mic.0.28805-0. [DOI] [PubMed] [Google Scholar]
  • 142.Garrity GM. Appendix 2. Taxonomic outline of the archaea and bacteria. 2. New York: Springer; 2005. Bergey’s Manual of Systematic Bacteriology. [Google Scholar]
  • 143.Kallman O, Fendukly F, Karlsson I, et al. Contribution of efflux to cefuroxime resistance in clinical isolates of Escherichia coli. Scand J Infect Dis. 2003;35(8):464–70. doi: 10.1080/00365540310014639. [DOI] [PubMed] [Google Scholar]
  • 144.Lautenbach E, Metlay JP, Weiner MG, et al. Gastrointestinal tract colonization with fluoroquinolone-resistant Escherichia coli in hospitalized patients: changes over time in risk factors for resistance. Infect Control Hosp Epidemiol. 2009;30(1):18–24. doi: 10.1086/592703. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Stubbings W, Bostock J, Ingham E, et al. Deletion of the multiple-drug efflux pump AcrAB in Escherichia coli prolongs the postantibiotic effect. Antimicrob Agents Chemother. 2005;49(3):1206–8. doi: 10.1128/AAC.49.3.1206-1208.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Hirata T, Saito A, Nishino K, et al. Effects of efflux transporter genes on susceptibility of Escherichia coli to tigecycline (GAR-936) Antimicrob Agents Chemother. 2004;48(6):2179–84. doi: 10.1128/AAC.48.6.2179-2184.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Keeney D, Ruzin A, McAleese F, et al. MarA-mediated overexpression of the AcrAB efflux pump results in decreased susceptibility to tigecycline in Escherichia coli. J Antimicrob Chemother. 2008;61(1):46–53. doi: 10.1093/jac/dkm397. [DOI] [PubMed] [Google Scholar]
  • 148.Gotoh N, Murata T, Ozaki T, et al. Intrinsic resistance of Escherichia coli to mureidomycin A and C due to expression of the multidrug efflux system AcrAB-TolC: comparison with the efflux systems of mureidomycin-susceptible Pseudomonas aeruginosa. J Infect Chemother. 2003;9(1):101–3. doi: 10.1007/s10156-002-0205-2. [DOI] [PubMed] [Google Scholar]
  • 149.Oppegard LM, Hamann BL, Streck KR, et al. In vivo and in vitro patterns of the activity of simocyclinone D8, an angucyclinone antibiotic from Streptomyces antibioticus. Antimicrob Agents Chemother. 2009;53(5):2110–9. doi: 10.1128/AAC.01440-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Delgado MA, Vincent PA, Farias RN, et al. YojI of Escherichia coli functions as a microcin J25 efflux pump. J Bacteriol. 2005;187(10):3465–70. doi: 10.1128/JB.187.10.3465-3470.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Socias SB, Vincent PA, Salomon RA. The leucine-responsive regulatory protein, Lrp, modulates microcin J25 intrinsic resistance in Escherichia coli by regulating expression of the YojI microcin exporter. J Bacteriol. 2009;191(4):1343–8. doi: 10.1128/JB.01074-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Wu B, Xia C, Du X, et al. Influence of anti-FloR antibody on florfenicol accumulation in florfenicol-resistant Escherichia coli and enzyme-linked immunosorbent assay for detection of florfenicol-resistant E. coli isolates. J Clin Microbiol. 2006;44(2):378–82. doi: 10.1128/JCM.44.2.378-382.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Higashi K, Ishigure H, Demizu R, et al. Identification of a spermidine excretion protein complex (MdtJI) in Escherichia coli. J Bacteriol. 2008;190(3):872–8. doi: 10.1128/JB.01505-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Hansen LH, Johannesen E, Burmolle M, et al. Plasmid-encoded multidrug efflux pump conferring resistance to olaquindox in Escherichia coli. Antimicrob Agents Chemother. 2004;48(9):3332–7. doi: 10.1128/AAC.48.9.3332-3337.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Hansen LH, Sorensen SJ, Jorgensen HS, et al. The prevalence of the OqxAB multidrug efflux pump amongst olaquindox-resistant Escherichia coli in pigs. Microb Drug Resist. 2005;11(4):378–82. doi: 10.1089/mdr.2005.11.378. [DOI] [PubMed] [Google Scholar]
  • 156.Hansen LH, Jensen LB, Sorensen HI, et al. Substrate specificity of the OqxAB multidrug resistance pump in Escherichia coli and selected enteric bacteria. J Antimicrob Chemother. 2007;60(1):145–7. doi: 10.1093/jac/dkm167. [DOI] [PubMed] [Google Scholar]
  • 157.Yamane K, Wachino J, Suzuki S, et al. Plasmid-mediated qepA gene among Escherichia coli clinical isolates from Japan. Antimicrob Agents Chemother. 2008;52(4):1564–6. doi: 10.1128/AAC.01137-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Ma J, Zeng Z, Chen Z, et al. High prevalence of plasmid-mediated quinolone resistance determinants qnr, aac(6′)-Ib-cr and qepA among ceftiofur-resistant enterobacteriaceae isolates from companion and food-producing animals. Antimicrob Agents Chemother. 2008;53(2):519–24. doi: 10.1128/AAC.00886-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Liu JH, Deng YT, Zeng ZL, et al. Coprevalence of plasmid-mediated quinolone resistance determinants QepA, Qnr, and AAC(6′)-Ib-cr among 16S rRNA methylase RmtB-producing Escherichia coli isolates from pigs. Antimicrob Agents Chemother. 2008;52(8):2992–3. doi: 10.1128/AAC.01686-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Cattoir V, Poirel L, Nordmann P. Plasmid-mediated quinolone resistance pump QepA2 in an Escherichia coli isolate from France. Antimicrob Agents Chemother. 2008;52(10):3801–4. doi: 10.1128/AAC.00638-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Baudry PJ, Nichol K, DeCorby M, et al. Mechanisms of resistance and mobility among multidrug-resistant CTX-M-producing Escherichia coli from Canadian intensive care units: the 1st report of QepA in North America. Diagn Microbiol Infect Dis. 2009;63(3):319–26. doi: 10.1016/j.diagmicrobio.2008.12.001. [DOI] [PubMed] [Google Scholar]
  • 162.Quinn T, O’Mahony R, Baird AW, et al. Multi-drug resistance in Salmonella enterica: efflux mechanisms and their relationships with the development of chromosomal resistance gene clusters. Curr Drug Targets. 2006;7(7):849–60. doi: 10.2174/138945006777709548. [DOI] [PubMed] [Google Scholar]
  • 163.Piddock LJ, White DG, Gensberg K, et al. Evidence for an efflux pump mediating multiple antibiotic resistance in Salmonella enterica serovar Typhimurium. Antimicrob Agents Chemother. 2000;44(11):3118–21. doi: 10.1128/aac.44.11.3118-3121.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Randall LP, Cooles SW, Sayers AR, et al. Association between cyclohexane resistance in Salmonella of different serovars and increased resistance to multiple antibiotics, disinfectants and dyes. J Med Microbiol. 2001;50(10):919–24. doi: 10.1099/0022-1317-50-10-919. [DOI] [PubMed] [Google Scholar]
  • 165.Chen S, Cui S, McDermott PF, et al. Contribution of target gene mutations and efflux to decreased susceptibility of Salmonella enterica serovar Typhimurium to fluoroquinolones and other antimicrobials. Antimicrob Agents Chemother. 2007;51(2):535–42. doi: 10.1128/AAC.00600-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Ricci V, Tzakas P, Buckley A, et al. Ciprofloxacin-resistant Salmonella enterica serovar Typhimurium strains are difficult to select in the absence of AcrB and TolC. Antimicrob Agents Chemother. 2006;50(1):38–42. doi: 10.1128/AAC.50.1.38-42.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Olliver A, Valle M, Chaslus-Dancla E, et al. Overexpression of the multidrug efflux operon acrEF by insertional activation with IS1 or IS10 elements in Salmonella enterica serovar typhimurium DT204 acrB mutants selected with fluoroquinolones. Antimicrob Agents Chemother. 2005;49(1):289–301. doi: 10.1128/AAC.49.1.289-301.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Nishino K, Nikaido E, Yamaguchi A. Regulation of multidrug efflux systems involved in multidrug and metal resistance of Salmonella enterica serovar Typhimurium. J Bacteriol. 2007;189(24):9066–9075. doi: 10.1128/JB.01045-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Braoudaki M, Hilton AC. Mechanisms of resistance in Salmonella enterica adapted to erythromycin, benzalkonium chloride and triclosan. Int J Antimicrob Agents. 2005;25(1):31–7. doi: 10.1016/j.ijantimicag.2004.07.016. [DOI] [PubMed] [Google Scholar]
  • 170.Murata T, Tseng W, Guina T, et al. PhoPQ-mediated regulation produces a more robust permeability barrier in the outer membrane of Salmonella enterica serovar Typhimurium. J Bacteriol. 2007;189(20):7213–22. doi: 10.1128/JB.00973-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Baucheron S, Mouline C, Praud K, et al. TolC but not AcrB is essential for multidrug-resistant Salmonella enterica serotype Typhimurium colonization of chicks. J Antimicrob Chemother. 2005;55(5):707–12. doi: 10.1093/jac/dki091. [DOI] [PubMed] [Google Scholar]
  • 172.Nishino K, Latifi T, Groisman EA. Virulence and drug resistance roles of multidrug efflux systems of Salmonella enterica serovar Typhimurium. Mol Microbiol. 2006;59(1):126–41. doi: 10.1111/j.1365-2958.2005.04940.x. [DOI] [PubMed] [Google Scholar]
  • 173.Buckley AM, Webber MA, Cooles S, et al. The AcrAB-TolC efflux system of serovar Typhimurium plays a role in pathogenesis. Cell Microbiol. 2006;8(5):847–56. doi: 10.1111/j.1462-5822.2005.00671.x. [DOI] [PubMed] [Google Scholar]
  • 174.Webber MA, Bailey AM, Blair JM, et al. The global consequence of disruption of the AcrAB-TolC efflux pump in Salmonella enterica includes reduced expression of SPI-1 and other attributes required to infect the host. J Bacteriol . 2009 doi: 10.1128/JB.00363-09. e-published May 1, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Gayet S, Chollet R, Molle G, et al. Modification of outer membrane protein profile and evidence suggesting an active drug pump in Enterobacter aerogenes clinical strains. Antimicrob Agents Chemother. 2003;47(5):1555–9. doi: 10.1128/AAC.47.5.1555-1559.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Chollet R, Chevalier J, Bryskier A, et al. The AcrAB-TolC pump is involved in macrolide resistance but not in telithromycin efflux in Enterobacter aerogenes and Escherichia coli. Antimicrob Agents Chemother. 2004;48(9):3621–4. doi: 10.1128/AAC.48.9.3621-3624.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Bornet C, Chollet R, Mallea M, et al. Imipenem and expression of multidrug efflux pump in Enterobacter aerogenes. Biochem Biophys Res Commun. 2003;301(4):985–90. doi: 10.1016/s0006-291x(03)00074-3. [DOI] [PubMed] [Google Scholar]
  • 178.Ghisalberti D, Masi M, Pages JM, et al. Chloramphenicol and expression of multidrug efflux pump in Enterobacter aerogenes. Biochem Biophys Res Commun. 2005;328(4):1113–8. doi: 10.1016/j.bbrc.2005.01.069. [DOI] [PubMed] [Google Scholar]
  • 179.Keeney D, Ruzin A, Bradford PA. RamA, a transcriptional regulator, and AcrAB, an RND-type efflux pump, are associated with decreased susceptibility to tigecycline in Enterobacter cloacae. Microb Drug Resist. 2007;13(1):1–6. doi: 10.1089/mdr.2006.9990. [DOI] [PubMed] [Google Scholar]
  • 180.Chollet R, Chevalier J, Bollet C, et al. RamA is an alternate activator of the multidrug resistance cascade in Enterobacter aerogenes. Antimicrob Agents Chemother. 2004;48(7):2518–23. doi: 10.1128/AAC.48.7.2518-2523.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Ghisalberti D, Mahamoud A, Chevalier J, et al. Chloroquinolines block antibiotic efflux pumps in antibiotic-resistant Enterobacter aerogenes isolates. Int J Antimicrob Agents. 2006;27(6):565–9. doi: 10.1016/j.ijantimicag.2006.03.010. [DOI] [PubMed] [Google Scholar]
  • 182.Masi M, Pages JM, Villard C, et al. The eefABC multidrug efflux pump operon is repressed by H-NS in Enterobacter aerogenes. J Bacteriol. 2005;187(11):3894–7. doi: 10.1128/JB.187.11.3894-3897.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Masi M, Pages J-M, Pradel E. Production of the cryptic EefABC efflux pump in Enterobacter aerogenes chloramphenicol-resistant mutants. J Antimicrob Chemother. 2006;57(6):1223–6. doi: 10.1093/jac/dkl139. [DOI] [PubMed] [Google Scholar]
  • 184.Park YJ, Yu JK, Kim SI, et al. Accumulation of plasmid-mediated fluoroquinolone resistance genes, qepA and qnrS1, in enterobacter aerogenes co-producing RmtB and class A β-lactamase LAP-1. Ann Clin Lab Sci. 2009;39(1):55–9. [PubMed] [Google Scholar]
  • 185.Szabo D, Silveira F, Hujer AM, et al. Outer membrane protein changes and efflux pump expression together may confer resistance to ertapenem in Enterobacter cloacae. Antimicrob Agents Chemother. 2006;50(8):2833–5. doi: 10.1128/AAC.01591-05. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Davin-Regli A, Chollet R, Bredin J, et al. Enterobacter gergoviae and the prevalence of efflux in parabens resistance. J Antimicrob Chemother. 2006;57(4):757–60. doi: 10.1093/jac/dkl023. [DOI] [PubMed] [Google Scholar]
  • 187.Schneiders T, Amyes SG, Levy SB. Role of AcrR and RamA in fluoroquinolone resistance in clinical Klebsiella pneumoniae isolates from Singapore. Antimicrob Agents Chemother. 2003;47(9):2831–7. doi: 10.1128/AAC.47.9.2831-2837.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Pages JM, Lavigne JP, Leflon-Guibout V, et al. Efflux pump, the masked side of β-lactam resistance in Klebsiella pneumoniae clinical isolates. PLoS ONE. 2009;4(3):e4817. doi: 10.1371/journal.pone.0004817. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189.Chevalier J, Bredin J, Mahamoud A, et al. Inhibitors of antibiotic efflux in resistant Enterobacter aerogenes and Klebsiella pneumoniae strains. Antimicrob Agents Chemother. 2004;48(3):1043–6. doi: 10.1128/AAC.48.3.1043-1046.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Ruzin A, Visalli MA, Keeney D, et al. Influence of transcriptional activator RamA on expression of multidrug efflux pump AcrAB and tigecycline susceptibility in Klebsiella pneumoniae. Antimicrob Agents Chemother. 2005;49(3):1017–22. doi: 10.1128/AAC.49.3.1017-1022.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Coudeyras S, Nakusi L, Charbonnel N, et al. A tripartite efflux pump involved in gastrointestinal colonization by Klebsiella pneumoniae confers a tolerance response to inorganic acid. Infect Immun. 2008;76(10):4633–41. doi: 10.1128/IAI.00356-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Ogawa W, Li DW, Yu P, et al. Multidrug resistance in Klebsiella pneumoniae MGH78578 and cloning of genes responsible for the resistance. Biol Pharm Bull. 2005;28(8):1505–8. doi: 10.1248/bpb.28.1505. [DOI] [PubMed] [Google Scholar]
  • 193.Ogawa W, Koterasawa M, Kuroda T, et al. KmrA multidrug efflux pump from Klebsiella pneumoniae. Biol Pharm Bull. 2006;29(3):550–3. doi: 10.1248/bpb.29.550. [DOI] [PubMed] [Google Scholar]
  • 194.Rodriguez-Martinez JM, Pichardo C, Garcia I, et al. Activity of ciprofloxacin and levofloxacin in experimental pneumonia caused by Klebsiella pneumoniae deficient in porins, expressing active efflux and producing QnrA1. Clin Microbiol Infect. 2008;14(7):691–7. doi: 10.1111/j.1469-0691.2008.02020.x. [DOI] [PubMed] [Google Scholar]
  • 195.Fenosa A, Fuste E, Ruiz L, et al. Role of TolC in Klebsiella oxytoca resistance to antibiotics. J Antimicrob Chemother. 2009;63(4):668–74. doi: 10.1093/jac/dkp027. [DOI] [PubMed] [Google Scholar]
  • 196.Stock I, Grueger T, Wiedemann B. Natural antibiotic susceptibility of strains of Serratia marcescens and the S.liquefaciens complex: S. liquefaciens sensu stricto, S. proteamaculans and S. grimesii. Int J Antimicrob Agents. 2003;22(1):35–47. doi: 10.1016/s0924-8579(02)00163-2. [DOI] [PubMed] [Google Scholar]
  • 197.Chen J, Kuroda T, Huda MN, et al. An RND-type multidrug efflux pump SdeXY from Serratia marcescens. J Antimicrob Chemother. 2003;52(2):176–9. doi: 10.1093/jac/dkg308. [DOI] [PubMed] [Google Scholar]
  • 198.Kumar A, Worobec EA. Cloning, sequencing, and characterization of the SdeAB multidrug efflux pump of Serratia marcescens. Antimicrob Agents Chemother. 2005;49(4):1495–1. doi: 10.1128/AAC.49.4.1495-1501.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Begic S, Worobec EA. The role of the Serratia marcescens SdeAB multidrug efflux pump and TolC homologue in fluoroquinolone resistance studied via gene-knockout mutagenesis. Microbiology. 2008;154(Pt 2):454–61. doi: 10.1099/mic.0.2007/012427-0. [DOI] [PubMed] [Google Scholar]
  • 200.Kumar A, Worobec EA. HasF, a TolC-homolog of Serratia marcescens, is involved in energy-dependent efflux. Can J Microbiol. 2005;51(6):497–500. doi: 10.1139/w05-029. [DOI] [PubMed] [Google Scholar]
  • 201.Begic S, Worobec EA. Fluoroquinolone resistance of Serratia marcescens: sucrose, salicylate, temperature, and pH induction of phenotypic resistance. Can J Microbiol. 2007;53(11):1239–45. doi: 10.1139/W07-097. [DOI] [PubMed] [Google Scholar]
  • 202.Shahcheraghi F, Minato Y, Chen J, et al. Molecular cloning and characterization of a multidrug efflux pump, SmfY, from Serratia marcescens. Biol Pharm Bull. 2007;30(4):798–800. doi: 10.1248/bpb.30.798. [DOI] [PubMed] [Google Scholar]
  • 203.Minato Y, Shahcheraghi F, Ogawa W, et al. Functional gene cloning and characterization of the SsmE multidrug efflux pump from Serratia marcescens. Biol Pharm Bull. 2008;31(3):516–9. doi: 10.1248/bpb.31.516. [DOI] [PubMed] [Google Scholar]
  • 204.Matsuo T, Chen J, Minato Y, et al. SmdAB, a heterodimeric ABC-Type multidrug efflux pump, in Serratia marcescens. J Bacteriol. 2008;190(2):648–54. doi: 10.1128/JB.01513-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205.Thompson SA, Maani EV, Lindell AH, et al. Novel tetracycline resistance determinant isolated from an environmental strain of Serratia marcescens. Appl Environ Microbiol. 2007;73(7):2199–206. doi: 10.1128/AEM.02511-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Gristwood T, Fineran PC, Everson L, et al. PigZ, a TetR/AcrR family repressor, modulates secondary metabolism via the expression of a putative four-component resistance-nodulation-cell-division efflux pump, ZrpADBC, in Serratia sp. ATCC 39006. Mol Microbiol. 2008;69(2):418–35. doi: 10.1111/j.1365-2958.2008.06291.x. [DOI] [PubMed] [Google Scholar]
  • 207.Bina JE, Provenzano D, Wang C, et al. Characterization of the Vibrio cholerae vexAB and vexCD efflux systems. Arch Microbiol. 2006;186(3):171–81. doi: 10.1007/s00203-006-0133-5. [DOI] [PubMed] [Google Scholar]
  • 208.Cerda FA, Ringelberg CS, Taylor RK. The bile response repressor, BreR, regulates expression of the Vibrio cholerae breAB efflux system operon. J Bacteriol. 2008;190(22):7441–52. doi: 10.1128/JB.00584-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Rahman MM, Matsuo T, Ogawa W, et al. Molecular cloning and characterization of all RND-type efflux transporters in Vibrio cholerae non-O1. Microbiol Immunol. 2007;51(11):1061–70. doi: 10.1111/j.1348-0421.2007.tb04001.x. [DOI] [PubMed] [Google Scholar]
  • 210.Woolley RC, Vediyappan G, Anderson M, et al. Characterization of the Vibrio cholerae vceCAB multiple-drug resistance efflux operon in Escherichia coli. J Bacteriol. 2005;187(15):5500–3. doi: 10.1128/JB.187.15.5500-5503.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Borges-Walmsley MI, Du D, McKeegan KS, et al. VceR regulates the vceCAB drug efflux pump operon of Vibrio cholerae by alternating between mutually exclusive conformations that bind either drugs or promoter DNA. J Mol Biol. 2005;349(2):387–400. doi: 10.1016/j.jmb.2005.03.045. [DOI] [PubMed] [Google Scholar]
  • 212.Alatoom AA, Aburto R, Hamood AN, et al. VceR negatively regulates the vceCAB MDR efflux operon and positively regulates its own synthesis in Vibrio cholerae 569B. Can J Microbiol. 2007;53(7):888–900. doi: 10.1139/W07-054. [DOI] [PubMed] [Google Scholar]
  • 213.Huda MN, Chen J, Morita Y, et al. Gene cloning and characterization of VcrM, a Na+-coupled multidrug efflux pump, from Vibrio cholerae non-O1. Microbiol Immunol. 2003;47(6):419–27. doi: 10.1111/j.1348-0421.2003.tb03379.x. [DOI] [PubMed] [Google Scholar]
  • 214.Huda N, Lee EW, Chen J, et al. Molecular cloning and characterization of an ABC multidrug efflux pump, VcaM, in non-O1 Vibrio cholerae. Antimicrob Agents Chemother. 2003;47(8):2413–7. doi: 10.1128/AAC.47.8.2413-2417.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Begum A, Rahman MM, Ogawa W, et al. Gene cloning and characterization of four MATE family multidrug efflux pumps from Vibrio cholerae non-O1. Microbiol Immunol. 2005;49(11):949–57. doi: 10.1111/j.1348-0421.2005.tb03690.x. [DOI] [PubMed] [Google Scholar]
  • 216.Gupta AK, Chauhan DS, Srivastava K, et al. Estimation of efflux mediated multi-drug resistance and its correlation with expression levels of two major efflux pumps in mycobacteria. J Commun Dis. 2006;38(3):246–54. [PubMed] [Google Scholar]
  • 217.Singh AK, Haldar R, Mandal D, et al. Analysis of the topology of Vibrio cholerae NorM and identification of amino acid residues involved in norfloxacin resistance. Antimicrob Agents Chemother. 2006;50(11):3717–23. doi: 10.1128/AAC.00460-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Srinivasan VB, Virk RK, Kaundal A, et al. Mechanism of drug resistance in clonally related clinical isolates of Vibrio fluvialis isolated in Kolkata, India. Antimicrob Agents Chemother. 2006;50(7):2428–32. doi: 10.1128/AAC.01561-05. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219.Balotescu C, Israil A, Radu R, et al. Aspects of constitutive and acquired antibioresistance in Aeromonas hydrophila strains isolated from water sources. Roum Arch Microbiol Immunol. 2003;62(3–4):179–89. [PubMed] [Google Scholar]
  • 220.Reith ME, Singh RK, Curtis B, et al. The genome of Aeromonas salmonicida. subspsalmonicida A449: insights into the evolution of a fish pathogen. BMC Genomics. 2008;9:427. doi: 10.1186/1471-2164-9-427. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221.Seshadri R, Joseph SW, Chopra AK, et al. Genome sequence of Aeromonas hydrophila ATCC 7966T: jack of all trades. J Bacteriol. 2006;188(23):8272–82. doi: 10.1128/JB.00621-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Rangrez AY, Kulkarni G, Dhotre D, Patole MS, Shouche YS. Prevalence of RND type multidrug efflux pump in the genus Aeromonas. Icfai J Biotech. 2008;2(1):72–80. [Google Scholar]
  • 223.Hernould M, Gagne S, Fournier M, et al. Role of the AheABC efflux pump in Aeromonas hydrophila intrinsic multidrug resistance. Antimicrob Agents Chemother. 2008;52(4):1559–63. doi: 10.1128/AAC.01052-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224.Marshall B, Morrissey S, Flynn P, et al. A new tetracycline-resistance determinant, class E, isolated from Enterobacteriaceae. Gene. 1986;50(1–3):111–7. doi: 10.1016/0378-1119(86)90315-x. [DOI] [PubMed] [Google Scholar]
  • 225.Agerso Y, Bruun MS, Dalsgaard I, et al. The tetracycline resistance gene tet(E) is frequently occurring and present on large horizontally transferable plasmids in Aeromonas spp. from fish farms. Aquaculture. 2007;266(1–4):47–52. [Google Scholar]
  • 226.Giraud E, Blanc G, Bouju-Albert A, et al. Donnay-Moreno C. Mechanisms of quinolone resistance and clonal relationship among Aeromonas salmonicida strains isolated from reared fish with furunculosis. J Med Microbiol. 2004;53(Pt 9):895–901. doi: 10.1099/jmm.0.45579-0. [DOI] [PubMed] [Google Scholar]
  • 227.Sugawara E, Nestorovich EM, Bezrukov SM, et al. Pseudomonas aeruginosa porin OprF exists in two different conformations. J Biol Chem. 2006;281(24):16220–9. doi: 10.1074/jbc.M600680200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228.Li X-Z, Nikaido H, Poole K. Role of MexA-MexB-OprM in antibiotic efflux in Pseudomonas aeruginosa. Antimicrob Agents Chemother. 1995;39(9):1948–53. doi: 10.1128/aac.39.9.1948. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Deplano A, Denis O, Poirel L, et al. Molecular characterization of an epidemic clone of panantibiotic-resistant Pseudomonas aeruginosa. J Clin Microbiol. 2005;43(3):1198–1204. doi: 10.1128/JCM.43.3.1198-1204.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230.Kriengkauykiat J, Porter E, Lomovskaya O, et al. Use of an efflux pump inhibitor to determine the prevalence of efflux pump-mediated fluoroquinolone resistance and multidrug resistance in Pseudomonas aeruginosa. Antimicrob Agents Chemother. 2005;49(2):565–70. doi: 10.1128/AAC.49.2.565-570.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 231.Pournaras S, Maniati M, Spanakis N, et al. Spread of efflux pump-overexpressing, non-metallo-β-lactamase-producing, meropenem-resistant but ceftazidime-susceptible Pseudomonas aeruginosa in a region with blaVIM endemicity. J Antimicrob Chemother. 2005;56(4):761–4. doi: 10.1093/jac/dki296. [DOI] [PubMed] [Google Scholar]
  • 232.Dumas JL, van Delden C, Perron K, et al. Analysis of antibiotic resistance gene expression in Pseudomonas aeruginosa by quantitative real-time-PCR. FEMS Microbiol Lett. 2006;254(2):217–25. doi: 10.1111/j.1574-6968.2005.00008.x. [DOI] [PubMed] [Google Scholar]
  • 233.Quale J, Bratu S, Gupta J, et al. Interplay of efflux system, ampC, and oprD expression in carbapenem resistance of Pseudomonas aeruginosa clinical isolates. Antimicrob Agents Chemother. 2006;50(5):1633–41. doi: 10.1128/AAC.50.5.1633-1641.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 234.Driscoll JA, Brody SL, Kollef MH. The epidemiology, pathogenesis and treatment of Pseudomonas aeruginosa infections. Drugs. 2007;67(3):351–68. doi: 10.2165/00003495-200767030-00003. [DOI] [PubMed] [Google Scholar]
  • 235.Burgess DS. Use of pharmacokinetics and pharmacodynamics to optimize antimicrobial treatment of Pseudomonas aeruginosa infections. Clin Infect Dis. 2005;40(Suppl 2):S99–104. doi: 10.1086/426189. [DOI] [PubMed] [Google Scholar]
  • 236.Rossolini GM, Mantengoli E. Treatment and control of severe infections caused by multiresistant Pseudomonas aeruginosa. Clin Microbiol Infect. 2005;11(Suppl 4):17–32. doi: 10.1111/j.1469-0691.2005.01161.x. [DOI] [PubMed] [Google Scholar]
  • 237.Boutoille D, Jacqueline C, Le Mabecque V, et al. In vivo impact of the MexAB-OprM efflux system on β-lactam efficacy in an experimental model of Pseudomonas aeruginosa infection. Int J Antimicrob Agents. 2009;33(5):417–20. doi: 10.1016/j.ijantimicag.2008.10.029. [DOI] [PubMed] [Google Scholar]
  • 238.Mesaros N, Glupczynski Y, Avrain L, et al. A combined phenotypic and genotypic method for the detection of Mex efflux pumps in Pseudomonas aeruginosa. J Antimicrob Chemother. 2007;59(3):378–86. doi: 10.1093/jac/dkl504. [DOI] [PubMed] [Google Scholar]
  • 239.Hocquet D, Berthelot P, Roussel-Delvallez M, et al. Pseudomonas aeruginosa may accumulate drug resistance mechanisms without losing its ability to cause bloodstream infections. Antimicrob Agents Chemother. 2007;51(10):3531–6. doi: 10.1128/AAC.00503-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 240.Llanes C, Hocquet D, Vogne C, et al. Clinical strains of Pseudomonas aeruginosa overproducing MexAB-OprM and MexXY efflux pumps simultaneously. Antimicrob Agents Chemother. 2004;48(5):1797–802. doi: 10.1128/AAC.48.5.1797-1802.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 241.Strateva T, Ouzounova-Raykova V, Markova B, et al. Problematic clinical isolates of Pseudomonas aeruginosa from the university hospitals in Sofia, Bulgaria: current status of antimicrobial resistance and prevailing resistance mechanisms. J Med Microbiol. 2007;56(Pt 7):956–63. doi: 10.1099/jmm.0.46986-0. [DOI] [PubMed] [Google Scholar]
  • 242.Livermore DM, Mushtaq S, Warner M. Selectivity of ertapenem for Pseudomonas aeruginosa mutants cross-resistant to other carbapenems. J Antimicrob Chemother. 2005;55(3):306–11. doi: 10.1093/jac/dki009. [DOI] [PubMed] [Google Scholar]
  • 243.Mikuniya T, Kato Y, Kariyama R, et al. Synergistic effect of fosfomycin and fluoroquinolones against Pseudomonas aeruginosa growing in a biofilm. Acta Med Okayama. 2005;59(5):209–16. doi: 10.18926/AMO/31977. [DOI] [PubMed] [Google Scholar]
  • 244.Longbottom CJ, Carson CF, Hammer KA, et al. Tolerance of Pseudomonas aeruginosa to Melaleuca alternifolia (tea tree) oil is associated with the outer membrane and energy-dependent cellular processes. J Antimicrob Chemother. 2004;54(2):386–92. doi: 10.1093/jac/dkh359. [DOI] [PubMed] [Google Scholar]
  • 245.Hocquet D, Vogne C, El Garch F, et al. MexXY-OprM efflux pump is necessary for a adaptive resistance of Pseudomonas aeruginosa to aminoglycosides. Antimicrob Agents Chemother. 2003;47(4):1371–5. doi: 10.1128/AAC.47.4.1371-1375.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 246.Sobel ML, McKay GA, Poole K. Contribution of the MexXY multidrug transporter to aminoglycoside resistance in Pseudomonas aeruginosa clinical isolates. Antimicrob Agents Chemother. 2003;47(10):3202–7. doi: 10.1128/AAC.47.10.3202-3207.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 247.Islam S, Jalal S, Wretlind B. Expression of the MexXY efflux pump in amikacin-resistant isolates of Pseudomonas aeruginosa. Clin Microbiol Infect. 2004;10(10):877–83. doi: 10.1111/j.1469-0691.2004.00991.x. [DOI] [PubMed] [Google Scholar]
  • 248.Vogne C, Aires JR, Bailly C, et al. Role of the multidrug efflux system MexXY in the emergence of moderate resistance to aminoglycosides among Pseudomonas aeruginosa isolates from patients with cystic fibrosis. Antimicrob Agents Chemother. 2004;48(5):1676–80. doi: 10.1128/AAC.48.5.1676-1680.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 249.Wolter DJ, Smith-Moland E, Goering RV, et al. Multidrug resistance associated with mexXY expression in clinical isolates of Pseudomonas aeruginosa from a Texas hospital. Diagn Microbiol Infect Dis. 2004;50(1):43–50. doi: 10.1016/j.diagmicrobio.2004.05.004. [DOI] [PubMed] [Google Scholar]
  • 250.Llanes C, Neuwirth C, El Garch F, et al. Genetic analysis of a multiresistant strain of Pseudomonas aeruginosa producing PER-1 β-lactamase. Clin Microbiol Infect. 2006;12(3):270–8. doi: 10.1111/j.1469-0691.2005.01333.x. [DOI] [PubMed] [Google Scholar]
  • 251.Hocquet D, Nordmann P, El Garch F, et al. Involvement of the MexXY-OprM efflux system in emergence of cefepime resistance in clinical strains of Pseudomonas aeruginosa. Antimicrob Agents Chemother. 2006;50(4):1347–51. doi: 10.1128/AAC.50.4.1347-1351.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 252.Vettoretti L, Plesiat P, Muller C, et al. Efflux unbalance in cystic fibrosis isolates of Pseudomonas aeruginosa. Antimicrob Agents Chemother . 2009;53(5):1987–97. doi: 10.1128/AAC.01024-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 253.El’Garch F, Jeannot K, Hocquet D, et al. Cumulative effects of several nonenzymatic mechanisms on the resistance of Pseudomonas aeruginosa to aminoglycosides. Antimicrob Agents Chemother. 2007;51(3):1016–21. doi: 10.1128/AAC.00704-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 254.Jo JT, Brinkman FS, Hancock RE. Aminoglycoside efflux in Pseudomonas aeruginosa: involvement of novel outer membrane proteins. Antimicrob Agents Chemother. 2003;47(3):1101–11. doi: 10.1128/AAC.47.3.1101-1111.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 255.Dupont P, Hocquet D, Jeannot K, et al. Bacteriostatic and bactericidal activities of eight fluoroquinolones against MexAB-OprM-overproducing clinical strains of Pseudomonas aeruginosa. J Antimicrob Chemother. 2005;55(4):518–22. doi: 10.1093/jac/dki030. [DOI] [PubMed] [Google Scholar]
  • 256.Griffith DC, Corcoran E, Lofland D, et al. Pharmacodynamics of levofloxacin against Pseudomonas aeruginosa with reduced susceptibility due to different efflux pumps: do elevated MICs always predict reduced in vivo efficacy? Antimicrob Agents Chemother. 2006;50(5):1628–32. doi: 10.1128/AAC.50.5.1628-1632.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 257.Martha B, Croisier D, Durand D, et al. In-vivo impact of the MexXY efflux system on aminoglycoside efficacy in an experimental model of Pseudomonas aeruginosa pneumonia treated with tobramycin. Clin Microbiol Infect. 2006;12(5):426–32. doi: 10.1111/j.1469-0691.2006.01371.x. [DOI] [PubMed] [Google Scholar]
  • 258.Ong CT, Tessier PR, Li C, et al. Comparative in vivo efficacy of meropenem, imipenem, and cefepime against Pseudomonas aeruginosa expressing MexA-MexB-OprM efflux pumps. Diagn Microbiol Infect Dis. 2007;57(2):153–61. doi: 10.1016/j.diagmicrobio.2006.06.014. [DOI] [PubMed] [Google Scholar]
  • 259.Lister PD, Wolter DJ, Wickman PA, et al. Levofloxacin/imipenem prevents the emergence of high-level resistance among Pseudomonas aeruginosa strains already lacking susceptibility to one or both drugs. J Antimicrob Chemother. 2006;57(5):999–1003. doi: 10.1093/jac/dkl063. [DOI] [PubMed] [Google Scholar]
  • 260.Chuanchuen R, Karkhoff-Schweizer RR, Schweizer HP. High-level triclosan resistance in Pseudomonas aeruginosa is solely a result of efflux. Am J Infect Control. 2003;31(2):124–7. doi: 10.1067/mic.2003.11. [DOI] [PubMed] [Google Scholar]
  • 261.Chuanchuen R, Murata T, Gotoh N, et al. Substrate-dependent utilization of OprM or OpmH by the Pseudomonas aeruginosa MexJK efflux pump. Antimicrob Agents Chemother. 2005;49(5):2133–6. doi: 10.1128/AAC.49.5.2133-2136.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 262.Li Y, Mima T, Komori Y, et al. A new member of the tripartite multidrug efflux pumps, MexVW-OprM, in Pseudomonas aeruginosa. J Antimicrob Chemother. 2003;52(4):572–5. doi: 10.1093/jac/dkg390. [DOI] [PubMed] [Google Scholar]
  • 263.He GX, Kuroda T, Mima T, et al. An H+-coupled multidrug efflux pump, PmpM, a member of the MATE family of transporters, from Pseudomonas aeruginosa. J Bacteriol. 2004;186(1):262–5. doi: 10.1128/JB.186.1.262-265.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 264.Mima T, Sekiya H, Mizushima T, et al. Gene cloning and properties of the RND-type multidrug efflux pumps MexPQ-OpmE and MexMN-OprM from Pseudomonas aeruginosa. Microbiol Immunol. 2005;49(11):999–1002. doi: 10.1111/j.1348-0421.2005.tb03696.x. [DOI] [PubMed] [Google Scholar]
  • 265.Zhou J, Hao D, Wang X, et al. An important role of a “probable ATP-binding component of ABC transporter” during the process of Pseudomonas aeruginosa resistance to fluoroquinolone. Proteomics. 2006;6(8):2495–503. doi: 10.1002/pmic.200501354. [DOI] [PubMed] [Google Scholar]
  • 266.Dean CR, Visalli MA, Projan SJ, et al. Efflux-mediated resistance to tigecycline (GAR-936) in Pseudomonas aeruginosa PAO1. Antimicrob Agents Chemother. 2003;47(3):972–8. doi: 10.1128/AAC.47.3.972-978.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 267.Visalli MA, Murphy E, Projan SJ, et al. AcrAB multidrug efflux pump is associated with reduced levels of susceptibility to tigecycline (GAR-936) in Proteus mirabilis. Antimicrob Agents Chemother. 2003;47(2):665–9. doi: 10.1128/AAC.47.2.665-669.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 268.Mushtaq S, Ge Y, Livermore DM. Doripenem versus Pseudomonas aeruginosa in vitro: activity against characterized isolates, mutants, and transconjugants and resistance selection potential. Antimicrob Agents Chemother. 2004;48(8):3086–92. doi: 10.1128/AAC.48.8.3086-3092.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 269.Ruzin A, Keeney D, Bradford PA. AcrAB efflux pump plays a role in decreased susceptibility to tigecycline in Morganella morganii. Antimicrob Agents Chemother. 2005;49(2):791–3. doi: 10.1128/AAC.49.2.791-793.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 270.Baum EZ, Crespo-Carbone SM, Morrow B, et al. Effect of MexXY overexpression on ceftobiprole susceptibility in Pseudomonas aeruginosa. Antimicrob Agents Chemother. 2009 doi: 10.1128/AAC.00018-09. e-published May 11, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 271.Robertson GT, Doyle TB, Du Q, et al. A novel indole compound that inhibits Pseudomonas aeruginosa growth by targeting MreB is a substrate for MexAB-OprM. J Bacteriol. 2007;189(19):6870–81. doi: 10.1128/JB.00805-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 272.Vaara M. Agents that increase the permeability of the outer membrane. Microbiol Rev. 1992;56(3):395–411. doi: 10.1128/mr.56.3.395-411.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 273.Ellison ML, Roberts AL, Champlin FR. Susceptibility of compound 48/80-sensitized Pseudomonas aeruginosa to the hydrophobic biocide triclosan. FEMS Microbiol Lett. 2007;269(2):295–300. doi: 10.1111/j.1574-6968.2007.00640.x. [DOI] [PubMed] [Google Scholar]
  • 274.Teran W, Felipe A, Segura A, et al. Antibiotic-dependent induction of Pseudomonas putida DOT-T1E TtgABC efflux pump is mediated by the drug binding repressor TtgR. Antimicrob Agents Chemother. 2003;47(10):3067–72. doi: 10.1128/AAC.47.10.3067-3072.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 275.Hearn EM, Dennis JJ, Gray MR, et al. Identification and characterization of the emhABC efflux system for polycyclic aromatic hydrocarbons in Pseudomonas fluorescens cLP6a. J Bacteriol. 2003;185(21):6233–40. doi: 10.1128/JB.185.21.6233-6240.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 276.Hearn EM, Gray MR, Foght JM. Mutations in the central cavity and periplasmic domain affect efflux activity of the resistance-nodulation-division pump EmhB from Pseudomonas fluorescens cLP6a. J Bacteriol. 2006;188(1):115–23. doi: 10.1128/JB.188.1.115-123.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 277.Nagai K, Murata T, Ohta S, et al. Two different mechanisms are involved in the extremely high-level benzalkonium chloride resistance of a Pseudomonas fluorescens strain. Microbiol Immunol. 2003;47(10):709–15. doi: 10.1111/j.1348-0421.2003.tb03440.x. [DOI] [PubMed] [Google Scholar]
  • 278.Jude F, Arpin C, Brachet-Castang C, et al. TbtABM, a multidrug efflux pump associated with tributyltin resistance in Pseudomonas stutzeri. FEMS Microbiol Lett. 2004;232(1):7–14. doi: 10.1016/S0378-1097(04)00012-6. [DOI] [PubMed] [Google Scholar]
  • 279.Huang X, Yan A, Zhang X, et al. Identification and characterization of a putative ABC transporter PltHIJKN required for pyoluteorin production in Pseudomonas sp. M18. Gene. 2006;376(1):68–78. doi: 10.1016/j.gene.2006.02.009. [DOI] [PubMed] [Google Scholar]
  • 280.Bergogne-Berezin E, Towner KJ. Acinetobacter spp. as nosocomial pathogens: microbiological, clinical, and epidemiological features. Clin Microbiol Rev. 1996;9(2):148–65. doi: 10.1128/cmr.9.2.148. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 281.Peleg AY, Paterson DL. Multidrug-resistant Acinetobacter: a threat to the antibiotic era. Intern Med J. 2006;36(8):479–82. doi: 10.1111/j.1445-5994.2006.01130.x. [DOI] [PubMed] [Google Scholar]
  • 282.Gilad J, Carmeli Y. Treatment options for multidrug-resistant Acinetobacter species. Drugs. 2008;68(2):165–89. doi: 10.2165/00003495-200868020-00003. [DOI] [PubMed] [Google Scholar]
  • 283.Sato K, Nakae T. Outer membrane permeability of Acinetobacter calcoaceticus and its implication in antibiotic resistance. J Antimicrob Chemother. 1991;28(1):35–45. doi: 10.1093/jac/28.1.35. [DOI] [PubMed] [Google Scholar]
  • 284.Yun SH, Choi CW, Park SH, et al. Proteomic analysis of outer membrane proteins from Acinetobacter baumannii DU202 in tetracycline stress condition. J Microbiol. 2008;46(6):720–7. doi: 10.1007/s12275-008-0202-3. [DOI] [PubMed] [Google Scholar]
  • 285.Vila J, Marti S, Sanchez-Cespedes J. Porins, efflux pumps and multidrug resistance in Acinetobacter baumannii. J Antimicrob Chemother. 2007;59(6):1210–5. doi: 10.1093/jac/dkl509. [DOI] [PubMed] [Google Scholar]
  • 286.Damier-Piolle L, Magnet S, Bremont S, et al. AdeIJK, a resistance-nodulation-cell division pump effluxing multiple antibiotics in Acinetobacter baumannii. Antimicrob Agents Chemother. 2008;52(2):557–62. doi: 10.1128/AAC.00732-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 287.Chau SL, Chu YW, Houang ET. Novel resistance-nodulation-cell division efflux system AdeDE in Acinetobacter genomic DNA group 3. Antimicrob Agents Chemother. 2004;48(10):4054–5. doi: 10.1128/AAC.48.10.4054-4055.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 288.Chu YW, Chau SL, Houang ET. Presence of active efflux systems AdeABC, AdeDE and AdeXYZ in different Acinetobacter genomic DNA groups. J Med Microbiol. 2006;55(Pt 4):477–8. doi: 10.1099/jmm.0.46433-0. [DOI] [PubMed] [Google Scholar]
  • 289.Marchand I, Damier-Piolle L, Courvalin P, et al. Expression of the RND-type efflux pump AdeABC in Acinetobacter baumannii is regulated by the AdeRS two-component system. Antimicrob Agents Chemother. 2004;48(9):3298–304. doi: 10.1128/AAC.48.9.3298-3304.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 290.Nemec A, Maixnerova M, van der Reijden TJ, et al. Relationship between the AdeABC efflux system gene content, netilmicin susceptibility and multidrug resistance in a genotypically diverse collection of Acinetobacter baumannii strains. J Antimicrob Chemother. 2007;60(3):483–9. doi: 10.1093/jac/dkm231. [DOI] [PubMed] [Google Scholar]
  • 291.Ruzin A, Keeney D, Bradford PA. AdeABC multidrug efflux pump is associated with decreased susceptibility to tigecycline in Acinetobacter calcoaceticus-Acinetobacter baumannii complex. J Antimicrob Chemother. 2007;59(5):1001–4. doi: 10.1093/jac/dkm058. [DOI] [PubMed] [Google Scholar]
  • 292.Peleg AY, Adams J, Paterson DL. Tigecycline efflux as a mechanism for nonsusceptibility in Acinetobacter baumannii. Antimicrob Agents Chemother. 2007;51(6):2065–9. doi: 10.1128/AAC.01198-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 293.Huys G, Cnockaert M, Nemec A, et al. Sequence-based typing of adeB as a potential tool to identify intraspecific groups among clinical strains of multidrug-resistant Acinetobacter baumannii. J Clin Microbiol. 2005;43(10):5327–31. doi: 10.1128/JCM.43.10.5327-5331.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 294.Higgins PG, Wisplinghoff H, Stefanik D, et al. Selection of topoisomerase mutations and overexpression of adeB mRNA transcripts during an outbreak of Acinetobacter baumannii. J Antimicrob Chemother. 2004;54(4):821–3. doi: 10.1093/jac/dkh427. [DOI] [PubMed] [Google Scholar]
  • 295.Peleg AY, Potoski BA, Rea R, et al. Acinetobacter baumannii bloodstream infection while receiving tigecycline: a cautionary report. J Antimicrob Chemother. 2007;59(1):128–31. doi: 10.1093/jac/dkl441. [DOI] [PubMed] [Google Scholar]
  • 296.Lin L, Ling BD, Li X-Z. Distribution of the multidrug efflux pump genes, adeABC, adeDE and adeIJK, and class 1 integron genes in multiple-antimicrobial-resistant clinical isolates of Acinetobacter baumannii-Acinetobacter calcoaceticus complex. Int J Antimicrob Agents. 2009;33(1):27–32. doi: 10.1016/j.ijantimicag.2008.06.027. [DOI] [PubMed] [Google Scholar]
  • 297.Siroy A, Cosette P, Seyer D, et al. Global comparison of the membrane subproteomes between a multidrug-resistant Acinetobacter baumannii strain and a reference strain. J Proteome Res. 2006;5(12):3385–98. doi: 10.1021/pr060372s. [DOI] [PubMed] [Google Scholar]
  • 298.Su XZ, Chen J, Mizushima T, Kuroda T, et al. AbeM, an H+-coupled Acinetobacter baumannii multidrug efflux pump belonging to the MATE family of transporters. Antimicrob Agents Chemother. 2005;49(10):4362–4. doi: 10.1128/AAC.49.10.4362-4364.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 299.Gebreyes W, Srinivasan V, Rajamohan G, et al. Novel secondary active transporters conferring antimicrobial resistance in Acinetobacter baumannii with broad substrate specificity. C1-1048. Abstracts of 48th ICAAC/IDSA 46th Annual Meeting; Washington DC. 2008. [Google Scholar]
  • 300.Gomez MJ, Neyfakh AA. Identification of genes involved in intrinsic antibiotic resistance of Acinetobacter baylyi. Antimicrob Agents Chemother. 2006;50(11):3562–7. doi: 10.1128/AAC.00579-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 301.Guardabassi L, Dijkshoorn L, Collard JM, et al. Distribution and in-vitro transfer of tetracycline resistance determinants in clinical and aquatic Acinetobacter strains. J Med Microbiol. 2000;49(10):929–36. doi: 10.1099/0022-1317-49-10-929. [DOI] [PubMed] [Google Scholar]
  • 302.Huys G, Cnockaert M, Vaneechoutte M, et al. Distribution of tetracycline resistance genes in genotypically related and unrelated multiresistant Acinetobacter baumannii strains from different European hospitals. Res Microbiol. 2005;156(3):348–55. doi: 10.1016/j.resmic.2004.10.008. [DOI] [PubMed] [Google Scholar]
  • 303.Ribera A, Ruiz J, Vila J. Presence of the Tet M determinant in a clinical isolate of Acinetobacter baumannii. Antimicrob Agents Chemother. 2003;47(7):2310–2. doi: 10.1128/AAC.47.7.2310-2312.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 304.Nicodemo AC, Paez JI. Antimicrobial therapy for Stenotrophomonas maltophilia infections. Eur J Clin Microbiol Infect Dis. 2007;26(4):229–37. doi: 10.1007/s10096-007-0279-3. [DOI] [PubMed] [Google Scholar]
  • 305.Rahmati-Bahram A, Magee JT, Jackson SK. Temperature-dependent aminoglycoside resistance in Stenotrophomonas (Xanthomonas) maltophilia; alterations in protein and lipopolysaccharide with growth temperature. J Antimicrob Chemother. 1996;37(4):665–76. doi: 10.1093/jac/37.4.665. [DOI] [PubMed] [Google Scholar]
  • 306.Crossman LC, Gould VC, Dow JM, et al. The complete genome, comparative and functional analysis of Stenotrophomonas maltophilia reveals an organism heavily shielded by drug resistance determinants. Genome Biol. 2008;9(4):R74. doi: 10.1186/gb-2008-9-4-r74. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 307.Li X-Z, Zhang L, Poole K. SmeC, an outer membrane multidrug efflux protein of Stenotrophomonas maltophilia. Antimicrob Agents Chemother. 2002;46(2):333–43. doi: 10.1128/AAC.46.2.333-343.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 308.Chang LL, Chen HF, Chang CY, et al. Contribution of integrons, and SmeABC and SmeDEF efflux pumps to multidrug resistance in clinical isolates of Stenotrophomonas maltophilia. J Antimicrob Chemother. 2004;53(3):518–21. doi: 10.1093/jac/dkh094. [DOI] [PubMed] [Google Scholar]
  • 309.Sanchez P, Moreno E, Martinez JL. The biocide triclosan selects Stenotrophomonas maltophilia mutants that overproduce the SmeDEF multidrug efflux pump. Antimicrob Agents Chemother. 2005;49(2):781–2. doi: 10.1128/AAC.49.2.781-782.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 310.Gould VC, Okazaki A, Howe RA, et al. Analysis of sequence variation among smeDEF multi drug efflux pump genes and flanking DNA from defined 16S rRNA subgroups of clinical Stenotrophomonas maltophilia isolates. J Antimicrob Chemother. 2004;54(2):348–53. doi: 10.1093/jac/dkh367. [DOI] [PubMed] [Google Scholar]
  • 311.Gould VC, Avison MB. SmeDEF-mediated antimicrobial drug resistance in Stenotrophomonas maltophilia clinical isolates having defined phylogenetic relationships. J Antimicrob Chemother. 2006;57(6):1070–6. doi: 10.1093/jac/dkl106. [DOI] [PubMed] [Google Scholar]
  • 312.Sanchez P, Le U, Martinez JL. The efflux pump inhibitor Phe-Arg-β-naphthylamide does not abolish the activity of the Stenotrophomonas maltophilia SmeDEF multidrug efflux pump. J Antimicrob Chemother. 2003;51(4):1042–5. doi: 10.1093/jac/dkg181. [DOI] [PubMed] [Google Scholar]
  • 313.Valdezate S, Vindel A, Saez-Nieto JA, et al. Preservation of topoisomerase genetic sequences during in vivo and in vitro development of high-level resistance to ciprofloxacin in isogenic Stenotrophomonas maltophilia strains. J Antimicrob Chemother. 2005;56(1):220–3. doi: 10.1093/jac/dki182. [DOI] [PubMed] [Google Scholar]
  • 314.Peric M, Bozdogan B, Jacobs MR, et al. Effects of an efflux mechanism and ribosomal mutations on macrolide susceptibility of Haemophilus influenzae clinical isolates. Antimicrob Agents Chemother. 2003;47(3):1017–22. doi: 10.1128/AAC.47.3.1017-1022.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 315.Bogdanovich T, Bozdogan B, Appelbaum PC. Effect of efflux on telithromycin and macrolide susceptibility in Haemophilus influenzae. Antimicrob Agents Chemother. 2006;50(3):893–8. doi: 10.1128/AAC.50.3.893-898.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 316.Peric M, Bozdogan B, Galderisi C, et al. Inability of L22 ribosomal protein alteration to increase macrolide MICs in the absence of efflux mechanism in Haemophilus influenzae HMC-S. J Antimicrob Chemother. 2004;54(2):393–400. doi: 10.1093/jac/dkh364. [DOI] [PubMed] [Google Scholar]
  • 317.Perez-Vazquez M, Roman F, Garcia-Cobos S, et al. Fluoroquinolone resistance in Haemophilus influenzae is associated with hypermutability. Antimicrob Agents Chemother. 2007;51(4):1566–9. doi: 10.1128/AAC.01437-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 318.Trepod CM, Mott JE. Identification of the Haemophilus influenzae tolC gene by susceptibility profiles of insertionally inactivated efflux pump mutants. Antimicrob Agents Chemother. 2004;48(4):1416–8. doi: 10.1128/AAC.48.4.1416-1418.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 319.Dean CR, Narayan S, Daigle DM, et al. Role of the AcrAB-TolC efflux pump in determining susceptibility of Haemophilus influenzae to the novel peptide deformylase inhibitor LBM415. Antimicrob Agents Chemother. 2005;49(8):3129–35. doi: 10.1128/AAC.49.8.3129-3135.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 320.Bogdanovich T, Smith KA, Clark C, et al. Activity of LBM415 compared to those of 11 other agents against Haemophilus species. Antimicrob Agents Chemother. 2006;50(7):2323–9. doi: 10.1128/AAC.00106-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 321.Kaczmarek FS, Gootz TD, Dib-Hajj F, et al. Genetic and molecular characterization of β-lactamase-negative ampicillin-resistant Haemophilus influenzae with unusually high resistance to ampicillin. Antimicrob Agents Chemother. 2004;48(5):1630–9. doi: 10.1128/AAC.48.5.1630-1639.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 322.Xu XJ, Su XZ, Morita Y, et al. Molecular cloning and characterization of the HmrM multidrug efflux pump from Haemophilus influenzae Rd. Microbiol Immunol. 2003;47(12):937–43. doi: 10.1111/j.1348-0421.2003.tb03467.x. [DOI] [PubMed] [Google Scholar]
  • 323.Siritapetawee J, Prinz H, Krittanai C, et al. Expression and refolding of Omp38 from Burkholderia pseudomallei and Burkholderia thailandensis, and its function as a diffusion porin. Biochem J. 2004;384(Pt 3):609–17. doi: 10.1042/BJ20041102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 324.Guglierame P, Pasca MR, De Rossi E, et al. Efflux pump genes of the resistance-nodulation-division family in Burkholderia cenocepacia genome. BMC Microbiol. 2006;6:66. doi: 10.1186/1471-2180-6-66. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 325.Nair BM, Cheung KJ, Jr, Griffith A, et al. Salicylate induces an antibiotic efflux pump in Burkholderia cepacia complex genomovar III (B. cenocepacia) J Clin Invest. 2004;113(3):464–73. doi: 10.1172/JCI19710. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 326.Chan YY, Tan TM, Ong YM, et al. BpeAB-OprB, a multidrug efflux pump in Burkholderia pseudomallei. Antimicrob Agents Chemother. 2004;48(4):1128–35. doi: 10.1128/AAC.48.4.1128-1135.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 327.Chan YY, Chua KL. The Burkholderia pseudomallei BpeAB-OprB efflux pump: expression and impact on quorum sensing and virulence. J Bacteriol. 2005;187(14):4707–19. doi: 10.1128/JB.187.14.4707-4719.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 328.Kumar A, Chua KL, Schweizer HP. Method for regulated expression of single-copy efflux pump genes in a surrogate Pseudomonas aeruginosa strain: identification of the BpeEF-OprC chloramphenicol and trimethoprim efflux pump of Burkholderia pseudomallei 1026b. Antimicrob Agents Chemother. 2006;50(10):3460–3. doi: 10.1128/AAC.00440-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 329.Nair BM, Joachimiak LA, Chattopadhyay S, Montano I, Burns JL. Conservation of a novel protein associated with an antibiotic efflux operon in Burkholderia cenocepacia. FEMS Microbiol Lett. 2005;245(2):337. doi: 10.1016/j.femsle.2005.03.027. [DOI] [PubMed] [Google Scholar]
  • 330.Kumar A, Mayo M, Trunck LA, et al. Expression of resistance-nodulation-cell-division efflux pumps in commonly used Burkholderia pseudomallei strains and clinical isolates from northern Australia. Trans R Soc Trop Med Hyg. 2008;102(Suppl 1):S145–51. doi: 10.1016/S0035-9203(08)70032-4. [DOI] [PubMed] [Google Scholar]
  • 331.Young JD, Blake M, Mauro A, et al. Properties of the major outer membrane protein from Neisseria gonorrhoeae incorporated into model lipid membranes. Proc Natl Acad Sci USA. 1983;80(12):3831–5. doi: 10.1073/pnas.80.12.3831. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 332.Tzeng YL, Ambrose KD, Zughaier S, et al. Cationic antimicrobial peptide resistance in Neisseria meningitidis. J Bacteriol. 2005;187(15):5387–96. doi: 10.1128/JB.187.15.5387-5396.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 333.Warner DM, Folster JP, Shafer WM, et al. Regulation of the MtrC-MtrD-MtrE efflux-pump system modulates the in vivo fitness of Neisseria gonorrhoeae. J Infect Dis. 2007;196(12):1804–12. doi: 10.1086/522964. [DOI] [PubMed] [Google Scholar]
  • 334.Olesky M, Zhao S, Rosenberg RL, et al. Porin-mediated antibiotic resistance in Neisseria gonorrhoeae: ion, solute, and antibiotic permeation through PIB proteins with penB mutations. J Bacteriol. 2006;188(7):2300–8. doi: 10.1128/JB.188.7.2300-2308.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 335.Shafer WM, Folster JP. Towards an understanding of chromosomally mediated penicillin resistance in Neisseria gonorrhoeae: evidence for a porin-efflux pump collaboration. J Bacteriol. 2006;188(7):2297–9. doi: 10.1128/JB.188.7.2297-2299.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 336.Tanaka M, Nakayama H, Huruya K, et al. Analysis of mutations within multiple genes associated with resistance in a clinical isolate of Neisseria gonorrhoeae with reduced ceftriaxone susceptibility that shows a multidrug-resistant phenotype. Int J Antimicrob Agents. 2006;27(1):20–6. doi: 10.1016/j.ijantimicag.2005.08.021. [DOI] [PubMed] [Google Scholar]
  • 337.Dewi BE, Akira S, Hayashi H, et al. High occurrence of simultaneous mutations in target enzymes and MtrRCDE efflux system in quinolone-resistant Neisseria gonorrhoeae. Sex Transm Dis. 2004;31(6):353–9. doi: 10.1097/00007435-200406000-00007. [DOI] [PubMed] [Google Scholar]
  • 338.Rouquette-Loughlin CE, Balthazar JT, Shafer WM. Characterization of the MacA-MacB efflux system in Neisseria gonorrhoeae. J Antimicrob Chemother. 2005;56(5):856–60. doi: 10.1093/jac/dki333. [DOI] [PubMed] [Google Scholar]
  • 339.Rouquette-Loughlin C, Dunham SA, Kuhn M, et al. The NorM efflux pump of Neisseria gonorrhoeae and Neisseria meningitidis recognizes antimicrobial cationic compounds. J Bacteriol. 2003;185(3):1101–6. doi: 10.1128/JB.185.3.1101-1106.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 340.Crawford SA, Fiebelkorn KR, Patterson JE, et al. International clone of Neisseria meningitidis serogroup A with tetracycline resistance due to tet(B) Antimicrob Agents Chemother. 2005;49(3):1198–200. doi: 10.1128/AAC.49.3.1198-1200.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 341.Jorgensen JH, Crawford SA, Fiebelkorn KR. Susceptibility of Neisseria meningitidis to 16 antimicrobial agents and characterization of resistance mechanisms affecting some agents. J Clin Microbiol. 2005;43(7):3162–71. doi: 10.1128/JCM.43.7.3162-3171.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 342.Ruiz J, Ribera A, Jurado A, et al. Evidence for a reserpine-affected mechanism of resistance to tetracycline in Neisseria gonorrhoeae. APMIS. 2005;113(10):670–4. doi: 10.1111/j.1600-0463.2005.apm_303.x. [DOI] [PubMed] [Google Scholar]
  • 343.Kamal N, Rouquette-Loughlin C, Shafer WM. The TolC-like protein of Neisseria meningitidis is required for extracellular production of the repeats-in-toxin toxin FrpC but not for resistance to antimicrobials recognized by the Mtr efflux pump system. Infect Immun. 2007;75(12):6008–12. doi: 10.1128/IAI.01995-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 344.Pappas G, Papadimitriou P, Christou L, et al. Future trends in human brucellosis treatment. Expert Opin Investig Drugs. 2006;15(10):1141–9. doi: 10.1517/13543784.15.10.1141. [DOI] [PubMed] [Google Scholar]
  • 345.Douglas JT, Rosenberg EY, Nikaido H, et al. Porins of Brucella species. Infect Immun. 1984;44(1):16–21. doi: 10.1128/iai.44.1.16-21.1984. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 346.DelVecchio VG, Kapatral V, Elzer P, et al. The genome of Brucella melitensis. Vet Microbiol. 2002;90(1–4):587–92. doi: 10.1016/s0378-1135(02)00238-9. [DOI] [PubMed] [Google Scholar]
  • 347.Paulsen IT, Seshadri R, Nelson KE, et al. The Brucella suis genome reveals fundamental similarities between animal and plant pathogens and symbionts. Proc Natl Acad Sci USA. 2002;99(20):13148–53. doi: 10.1073/pnas.192319099. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 348.Braibant M, Guilloteau L, Zygmunt MS. Functional characterization of Brucella melitensis NorMI, an efflux pump belonging to the multidrug and toxic compound extrusion family. Antimicrob Agents Chemother. 2002;46(9):3050–3. doi: 10.1128/AAC.46.9.3050-3053.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 349.Posadas DM, Martin FA, Sabio y Garcia JV, et al. The TolC homologue of Brucella suis is involved in resistance to antimicrobial compounds and virulence. Infect Immun. 2007;75(1):379–89. doi: 10.1128/IAI.01349-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 350.Martin FA, Posadas DM, Carrica MC, et al. Interplay between two RND systems mediating antimicrobial resistance in Brucella suis. J Bacteriol. 2009;191(8):2530–40. doi: 10.1128/JB.01198-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 351.Halling SM, Jensen AE. Intrinsic and selected resistance to antibiotics binding the ribosome: analyses of Brucella 23S rrn, L4, L22, EF-Tu1, EF-Tu2, efflux and phylogenetic implications. BMC Microbiol. 2006;6:84. doi: 10.1186/1471-2180-6-84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 352.Ravanel N, Gestin B, Maurin M. In vitro selection of fluoroquinolone resistance in Brucella melitensis. Int J Antimicrob Agents. 2009 doi: 10.1016/j.ijantimicag.2009.01.002. Epub ahead of print. [DOI] [PubMed] [Google Scholar]
  • 353.Labesse G, Garnotel E, Bonnel S, et al. MOMP, a divergent porin from Campylobacter: cloning and primary structural characterization. Biochem Biophys Res Commun. 2001;280(1):380–7. doi: 10.1006/bbrc.2000.4129. [DOI] [PubMed] [Google Scholar]
  • 354.Page WJ, Huyer G, Huyer M, et al. Characterization of the porins of Campylobacter jejuni and Campylobacter coli and implications for antibiotic susceptibility. Antimicrob Agents Chemother. 1989;33(3):297–303. doi: 10.1128/aac.33.3.297. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 355.Luo N, Sahin O, Lin J, et al. In vivo selection of Campylobacter isolates with high levels of fluoroquinolone resistance associated with gyrA mutations and the function of the CmeABC efflux pump. Antimicrob Agents Chemother. 2003;47(1):390–4. doi: 10.1128/AAC.47.1.390-394.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 356.Mamelli L, Amoros JP, Pages JM, et al. A phenylalanine-arginine β-naphthylamide sensitive multidrug efflux pump involved in intrinsic and acquired resistance of Campylobacter to macrolides. Int J Antimicrob Agents. 2003;22(3):237–41. doi: 10.1016/s0924-8579(03)00199-7. [DOI] [PubMed] [Google Scholar]
  • 357.Zhang Q, Plummer P. Mechanisms of antibiotic resistance in Campylobacter. In: Nachamkin I, Szymanski C, Blaser M, editors. Campylobacter. Washibgton DC: ASM Press; 2008. pp. 263–76. [Google Scholar]
  • 358.Corcoran D, Quinn T, Cotter L, et al. Characterization of a cmeABC operon in a quinolone-resistant Campylobacter coli isolate of Irish origin. Microb Drug Resist. 2005;11(4):303–8. doi: 10.1089/mdr.2005.11.303. [DOI] [PubMed] [Google Scholar]
  • 359.Ge B, McDermott PF, White DG, et al. Role of efflux pumps and topoisomerase mutations in fluoroquinolone resistance in Campylobacter jejuni and Campylobacter coli. Antimicrob Agents Chemother. 2005;49(8):3347–54. doi: 10.1128/AAC.49.8.3347-3354.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 360.Olah PA, Doetkott C, Fakhr MK, et al. Prevalence of the Campylobacter multi-drug efflux pump (CmeABC) in Campylobacter spp. isolated from freshly processed turkeys. Food Microbiol. 2006;23(5):453–60. doi: 10.1016/j.fm.2005.06.004. [DOI] [PubMed] [Google Scholar]
  • 361.Yan M, Sahin O, Lin J, et al. Role of the CmeABC efflux pump in the emergence of fluoroquinolone-resistant Campylobacter under selection pressure. J Antimicrob Chemother. 2006;58(6):1154–9. doi: 10.1093/jac/dkl412. [DOI] [PubMed] [Google Scholar]
  • 362.Gibreel A, Wetsch NM, Taylor DE. Contribution of the CmeABC efflux pump to macrolide and tetracycline resistance in Campylobacter jejuni. Antimicrob Agents Chemother. 2007;51(9):3212–6. doi: 10.1128/AAC.01592-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 363.Caldwell DB, Wang Y, Lin J. Development, stability, and molecular mechanisms of macrolide resistance in Campylobacter jejuni. Antimicrob Agents Chemother. 2008;52(11):3947–54. doi: 10.1128/AAC.00450-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 364.Hanninen ML, Hannula M. Spontaneous mutation frequency and emergence of ciprofloxacin resistance in Campylobacter jejuni and Campylobacter coli. J Antimicrob Chemother. 2007;60(6):1251–7. doi: 10.1093/jac/dkm345. [DOI] [PubMed] [Google Scholar]
  • 365.Piddock LJ, Griggs D, Johnson MM, et al. Persistence of Campylobacter species, strain types, antibiotic resistance and mechanisms of tetracycline resistance in poultry flocks treated with chlortetracycline. J Antimicrob Chemother. 2008;62(2):303–15. doi: 10.1093/jac/dkn190. [DOI] [PubMed] [Google Scholar]
  • 366.Cagliero C, Cloix L, Cloeckaert A, et al. High genetic variation in the multidrug transporter cmeB gene in Campylobacter jejuni and Campylobacter coli. J Antimicrob Chemother. 2006;58(1):168–72. doi: 10.1093/jac/dkl212. [DOI] [PubMed] [Google Scholar]
  • 367.Lin J, Yan M, Sahin O, et al. Effect of macrolide usage on emergence of erythromycin-resistant Campylobacter isolates in chickens. Antimicrob Agents Chemother. 2007;51(5):1678–86. doi: 10.1128/AAC.01411-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 368.Mamelli L, Prouzet-Mauleon V, Pages J-M, et al. Molecular basis of macrolide resistance in Campylobacter: role of efflux pumps and target mutations. J Antimicrob Chemother. 2005;56(3):491–7. doi: 10.1093/jac/dki253. [DOI] [PubMed] [Google Scholar]
  • 369.Mamelli L, Demoulin E, Prouzet-Mauleon V, et al. Prevalence of efflux activity in low-level macrolide-resistant Campylobacter species. J Antimicrob Chemother. 2007;59(2):327–8. doi: 10.1093/jac/dkl476. [DOI] [PubMed] [Google Scholar]
  • 370.Jeon B, Zhang Q. Sensitization of Campylobacter jejuni to fluoroquinolone and macrolide antibiotics by antisense inhibition of the CmeABC multidrug efflux transporter. J Antimicrob Chemother. 2009;63(5):946–8. doi: 10.1093/jac/dkp067. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 371.Akiba M, Lin J, Barton YW, et al. Interaction of CmeABC and CmeDEF in conferring antimicrobial resistance and maintaining cell viability in Campylobacter jejuni. J Antimicrob Chemother. 2006;57(1):52–60. doi: 10.1093/jac/dki419. [DOI] [PubMed] [Google Scholar]
  • 372.Pumbwe L, Randall LP, Woodward MJ, et al. Expression of the efflux pump genes cmeB, cmeF and the porin gene porA in multiple-antibiotic-resistant Campylobacter jejuni. J Antimicrob Chemother. 2004;54(2):341–7. doi: 10.1093/jac/dkh331. [DOI] [PubMed] [Google Scholar]
  • 373.Pumbwe L, Randall LP, Woodward MJ, et al. Evidence for multiple-antibiotic resistance in Campylobacter jejuni not mediated by CmeB or CmeF. Antimicrob Agents Chemother. 2005;49(4):1289–93. doi: 10.1128/AAC.49.4.1289-1293.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 374.Cagliero C, Mouline C, Cloeckaert A, et al. Synergy between efflux pump CmeABC and modifications in ribosomal proteins L4 and L22 in conferring macrolide resistance in Campylobacter jejuni and Campylobacter coli. Antimicrob Agents Chemother. 2006;50(11):3893–6. doi: 10.1128/AAC.00616-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 375.Fange D, Nilsson K, Tenson T, et al. Drug efflux pump deficiency and drug target resistance masking in growing bacteria. Proc Natl Acad Sci USA. 2009;106(20):8215–20. doi: 10.1073/pnas.0811514106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 376.Doig P, Exner MM, Hancock RE, et al. Isolation and characterization of a conserved porin protein from Helicobacter pylori. J Bacteriol. 1995;177(19):5447–52. doi: 10.1128/jb.177.19.5447-5452.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 377.Megraud F, Lehours P. Helicobacter pylori detection and antimicrobial susceptibility testing. Clin Microbiol Rev. 2007;20(2):280–322. doi: 10.1128/CMR.00033-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 378.Bina JE, Alm RA, Uria-Nickelsen M, et al. Helicobacter pylori uptake and efflux: basis for intrinsic susceptibility to antibiotics in vitro. Antimicrob Agents Chemother. 2000;44(2):248–54. doi: 10.1128/aac.44.2.248-254.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 379.DeLoney CR, Schiller NL. Characterization of an in vitro-selected amoxicillin-resistant strain of Helicobacter pylori. Antimicrob Agents Chemother. 2000;44(12):3368–73. doi: 10.1128/aac.44.12.3368-3373.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 380.Liu ZQ, Zheng PY, Yang PC. Efflux pump gene hefA of Helicobacter pylori plays an important role in multidrug resistance. World J Gastroenterol. 2008;14(33):5217–22. doi: 10.3748/wjg.14.5217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 381.van Amsterdam K, Bart A, van der Ende A. A Helicobacter pylori TolC efflux pump confers resistance to metronidazole. Antimicrob Agents Chemother. 2005;49(4):1477–82. doi: 10.1128/AAC.49.4.1477-1482.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 382.Stahler FN, Odenbreit S, Haas R, et al. The novel Helicobacter pylori CznABC metal efflux pump is required for cadmium, zinc, and nickel resistance, urease modulation, and gastric colonization. Infect Immun. 2006;74(7):3845–52. doi: 10.1128/IAI.02025-05. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 383.Pumbwe L, Skilbeck CA, Wexler HM. The Bacteroides fragilis cell envelope: quarterback, linebacker, coach-or all three? Anaerobe. 2006;12(5–6):211–20. doi: 10.1016/j.anaerobe.2006.09.004. [DOI] [PubMed] [Google Scholar]
  • 384.Wexler HM. Bacteroides: the good, the bad, and the nitty-gritty. Clin Microbiol Rev. 2007;20(4):593–621. doi: 10.1128/CMR.00008-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 385.Kanazawa K, Kobayashi Y, Nakano M, et al. Identification of three porins in the outer membrane of Bacteroides fragilis. FEMS Microbiol Lett. 1995;127(3):181–6. [Google Scholar]
  • 386.Hecht DW. Prevalence of antibiotic resistance in anaerobic bacteria: worrisome developments. Clin Infect Dis. 2004;39(1):92–7. doi: 10.1086/421558. [DOI] [PubMed] [Google Scholar]
  • 387.Salyers A, Shoemaker NB. Reservoirs of antibiotic resistance genes. Anim Biotechnol. 2006;17(2):137–46. doi: 10.1080/10495390600957076. [DOI] [PubMed] [Google Scholar]
  • 388.Ricci V, Peterson ML, Rotschafer JC, et al. Role of topoisomerase mutations and efflux in fluoroquinolone resistance of Bacteroides fragilis clinical isolates and laboratory mutants. Antimicrob Agents Chemother. 2004;48(4):1344–6. doi: 10.1128/AAC.48.4.1344-1346.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 389.Ueda O, Wexler HM, Hirai K, et al. Sixteen homologs of the mex-type multidrug resistance efflux pump in Bacteroides fragilis. Antimicrob Agents Chemother. 2005;49(7):2807–15. doi: 10.1128/AAC.49.7.2807-2815.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 390.Pumbwe L, Chang A, Smith RL, et al. BmeRABC5 is a multidrug efflux system that can confer metronidazole resistance in Bacteroides fragilis. Microb Drug Resist. 2007;13(2):96–101. doi: 10.1089/mdr.2007.719. [DOI] [PubMed] [Google Scholar]
  • 391.Pumbwe L, Glass D, Wexler HM. Efflux pump overexpression in multiple-antibiotic-resistant mutants of Bacteroides fragilis. Antimicrob Agents Chemother. 2006;50(9):3150–3. doi: 10.1128/AAC.00141-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 392.Pumbwe L, Chang A, Smith RL, et al. Clinical significance of overexpression of multiple RND-family efflux pumps in Bacteroides fragilis isolates. J Antimicrob Chemother. 2006;58(3):543–8. doi: 10.1093/jac/dkl278. [DOI] [PubMed] [Google Scholar]
  • 393.Pumbwe L, Wareham DW, Aduse-Opoku J, et al. Genetic analysis of mechanisms of multidrug resistance in a clinical isolate of Bacteroides fragilis. Clin Microbiol Infect. 2007;13(2):183–9. doi: 10.1111/j.1469-0691.2006.01620.x. [DOI] [PubMed] [Google Scholar]
  • 394.Wang Y, Wang GR, Shelby A, et al. A newly discovered Bacteroides conjugative transposon, CTnGERM1, contains genes also found in Gram-positive bacteria. Appl Environ Microbiol. 2003;69(8):4595–603. doi: 10.1128/AEM.69.8.4595-4603.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 395.Owens RC, Jr, Donskey CJ, Gaynes RP, et al. Antimicrobial-associated risk factors for Clostridium difficile infection. Clin Infect Dis. 2008;46(Suppl 1):S19–31. doi: 10.1086/521859. [DOI] [PubMed] [Google Scholar]
  • 396.Drudy D, Quinn T, O’Mahony R, et al. High-level resistance to moxifloxacin and gatifloxacin associated with a novel mutation in gyrB in toxin-A-negative, toxin-B-positive Clostridium difficile. J Antimicrob Chemother. 2006;58(6):1264–7. doi: 10.1093/jac/dkl398. [DOI] [PubMed] [Google Scholar]
  • 397.Sebaihia M, Wren BW, Mullany P, et al. The multidrug-resistant human pathogen Clostridium difficile has a highly mobile, mosaic genome. Nat Genet. 2006;38(7):779–86. doi: 10.1038/ng1830. [DOI] [PubMed] [Google Scholar]
  • 398.Dridi L, Tankovic J, Petit JC. CdeA of Clostridium difficile, a new multidrug efflux transporter of the MATE family. Microb Drug Resist. 2004;10(3):191–6. doi: 10.1089/mdr.2004.10.191. [DOI] [PubMed] [Google Scholar]
  • 399.Lebel S, Bouttier S, Lambert T. The cme gene of Clostridium difficile confers multidrug resistance in Enterococcus faecalis. FEMS Microbiol Lett. 2004;238(1):93–100. doi: 10.1016/j.femsle.2004.07.022. [DOI] [PubMed] [Google Scholar]
  • 400.Rafii F, Park M, Wynne R. Evidence for active drug efflux in fluoroquinolone resistance in Clostridium hathewayi. Chemotherapy. 2005;51(5):256–62. doi: 10.1159/000087253. [DOI] [PubMed] [Google Scholar]
  • 401.Bannam TL, Johanesen PA, Salvado CL, et al. The Clostridium perfringens TetA(P) efflux protein contains a functional variant of the Motif A region found in major facilitator superfamily transport proteins. Microbiology. 2004;150(Pt 1):127–34. doi: 10.1099/mic.0.26614-0. [DOI] [PubMed] [Google Scholar]
  • 402.Kazimierczak KA, Rincon MT, Patterson AJ, et al. A new tetracycline efflux gene, tet(40), is located in tandem with tet(O/32/O) in a human gut firmicute bacterium and in metagenomic library clones. Antimicrob Agents Chemother. 2008;52(11):4001–9. doi: 10.1128/AAC.00308-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 403.Ohki R, Tateno K. Increased stability of bmr3 mRNA results in a multidrug-resistant phenotype in Bacillus subtilis. J Bacteriol. 2004;186(21):7450–5. doi: 10.1128/JB.186.21.7450-7455.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 404.Murata M, Ohno S, Kumano M, et al. Multidrug resistant phenotype of Bacillus subtilis spontaneous mutants isolated in the presence of puromycin and lincomycin. Can J Microbiol. 2003;49(2):71–7. doi: 10.1139/w03-014. [DOI] [PubMed] [Google Scholar]
  • 405.Yoshida K, Ohki YH, Murata M, et al. Bacillus subtilis LmrA is a repressor of the lmrAB and yxaGH operons: identification of its binding site and functional analysis of lmrB and yxaGH. J Bacteriol. 2004;186(17):5640–8. doi: 10.1128/JB.186.17.5640-5648.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 406.Kumano M, Fujita M, Nakamura K, et al. Lincomycin resistance mutations in two regions immediately downstream of the -10 region of lmr promoter cause overexpression of a putative multidrug efflux pump in Bacillus subtilis mutants. Antimicrob Agents Chemother. 2003;47(1):432–5. doi: 10.1128/AAC.47.1.432-435.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 407.Safferling M, Griffith H, Jin J, et al. TetL tetracycline efflux protein from Bacillus subtilis is a dimer in the membrane and in detergent solution. Biochemistry. 2003;42(47):13969–76. doi: 10.1021/bi035173q. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 408.Kim J-Y, Inaoka T, Hirooka K, et al. Identification and characterization of a novel multidrug resistance operon mdtRP (yusOP) of Bacillus subtilis. J Bacteriol. 2009;191(10):3273–81. doi: 10.1128/JB.00151-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 409.Steinfels E, Orelle C, Fantino JR, et al. Characterization of YvcC (BmrA), a multidrug ABC transporter constitutively expressed in Bacillus subtilis. Biochemistry. 2004;43(23):7491–502. doi: 10.1021/bi0362018. [DOI] [PubMed] [Google Scholar]
  • 410.Price LB, Vogler A, Pearson T, et al. In vitro selection and characterization of Bacillus anthracis mutants with high-level resistance to ciprofloxacin. Antimicrob Agents Chemother. 2003;47(7):2362–5. doi: 10.1128/AAC.47.7.2362-2365.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 411.Grohs P, Podglajen I, Gutmann L. Activities of different fluoroquinolones against Bacillus anthracis mutants selected in vitro and harboring topoisomerase mutations. Antimicrob Agents Chemother. 2004;48(8):3024–7. doi: 10.1128/AAC.48.8.3024-3027.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 412.Bast DJ, Athamna A, Duncan CL, et al. Type II topoisomerase mutations in Bacillus anthracis associated with high-level fluoroquinolone resistance. J Antimicrob Chemother. 2004;54(1):90–4. doi: 10.1093/jac/dkh294. [DOI] [PubMed] [Google Scholar]
  • 413.Ramaswamy V, Cresence VM, Rejitha JS, et al. Listeria - review of epidemiology and pathogenesis. J Microbiol Immunol Infect. 2007;40(1):4–13. [PubMed] [Google Scholar]
  • 414.Marco F, Almela M, Nolla-Salas J, et al. In vitro activities of 22 antimicrobial agents against Listeria monocytogenes strains isolated in Barcelona, Spain. The Collaborative Study Group of Listeriosis of Barcelona. Diagn Microbiol Infect Dis. 2000;38(4):259–61. doi: 10.1016/s0732-8893(00)00208-x. [DOI] [PubMed] [Google Scholar]
  • 415.Hof H. Listeriosis: therapeutic options. FEMS Immunol Med Microbiol. 2003;35(3):203–5. doi: 10.1016/S0928-8244(02)00466-2. [DOI] [PubMed] [Google Scholar]
  • 416.Lyon SA, Berrang ME, Fedorka-Cray PJ, et al. Antimicrobial resistance of Listeria monocytogenes isolated from a poultry further processing plant. Foodborne Pathog Dis. 2008;5(3):253–9. doi: 10.1089/fpd.2007.0070. [DOI] [PubMed] [Google Scholar]
  • 417.Li Q, Sherwood JS, Logue CM. Antimicrobial resistance of Listeria spp. recovered from processed bison. Lett Appl Microbiol. 2007;44(1):86–91. doi: 10.1111/j.1472-765X.2006.02027.x. [DOI] [PubMed] [Google Scholar]
  • 418.Srinivasan V, Nam HM, Nguyen LT, Tamilselvam B, Murinda SE, Oliver SP. Prevalence of antimicrobial resistance genes in Listeria monocytogenes isolated from dairy farms. Foodborne Pathog Dis. 2005;2(3):201–11. doi: 10.1089/fpd.2005.2.201. [DOI] [PubMed] [Google Scholar]
  • 419.Mullapudi S, Siletzky RM, Kathariou S. Heavy-metal and benzalkonium chloride resistance of Listeria monocytogenes isolates from the environment of turkey-processing plants. Appl Environ Microbiol. 2008;74(5):1464–8. doi: 10.1128/AEM.02426-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 420.Soumet C, Ragimbeau C, Maris P. Screening of benzalkonium chloride resistance in Listeria monocytogenes strains isolated during cold smoked fish production. Lett Appl Microbiol. 2005;41(3):291–6. doi: 10.1111/j.1472-765X.2005.01763.x. [DOI] [PubMed] [Google Scholar]
  • 421.Romanova NA, Wolffs PF, Brovko LY, et al. Role of efflux pumps in adaptation and resistance of Listeria monocytogenes to benzalkonium chloride. Appl Environ Microbiol. 2006;72(5):3498–503. doi: 10.1128/AEM.72.5.3498-3503.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 422.Huillet E, Velge P, Vallaeys T, et al. LadR, a new PadR-related transcriptional regulator from Listeria monocytogenes, negatively regulates the expression of the multidrug efflux pump MdrL. FEMS Microbiol Lett. 2006;254(1):87–94. doi: 10.1111/j.1574-6968.2005.00014.x. [DOI] [PubMed] [Google Scholar]
  • 423.Crimmins GT, Herskovits AA, Rehder K, et al. Listeria monocytogenes multidrug resistance transporters activate a cytosolic surveillance pathway of innate immunity. Proc Natl Acad Sci USA. 2008;105(29):10191–6. doi: 10.1073/pnas.0804170105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 424.Godreuil S, Galimand M, Gerbaud G, et al. Efflux pump Lde is associated with fluoroquinolone resistance in Listeria monocytogenes. Antimicrob Agents Chemother. 2003;47(2):704–8. doi: 10.1128/AAC.47.2.704-708.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 425.Lismond A, Tulkens PM, Mingeot-Leclercq MP, et al. Cooperation between prokaryotic (Lde) and eukaryotic (MRP) efflux transporters in J774 macrophages infected with Listeria monocytogenes: studies with ciprofloxacin and moxifloxacin. Antimicrob Agents Chemother. 2008;52(9):3040–6. doi: 10.1128/AAC.00105-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 426.Scortti M, Lacharme-Lora L, Wagner M, et al. Coexpression of virulence and fosfomycin susceptibility in Listeria: molecular basis of an antimicrobial in vitro-in vivo paradox. Nat Med. 2006;12(5):515–7. doi: 10.1038/nm1396. [DOI] [PubMed] [Google Scholar]
  • 427.Hassan KA, Skurray RA, Brown MH. Active export proteins mediating drug resistance in staphylococci. J Mol Microbiol Biotechnol. 2007;12(3–4):180–96. doi: 10.1159/000099640. [DOI] [PubMed] [Google Scholar]
  • 428.Huang J, O’Toole PW, Shen W, et al. Novel chromosomally encoded multidrug efflux transporter MdeA in Staphylococcus aureus. Antimicrob Agents Chemother. 2004;48(3):909–17. doi: 10.1128/AAC.48.3.909-917.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 429.Kaatz GW, McAleese F, Seo SM. Multidrug resistance in Staphylococcus aureus due to overexpression of a novel multidrug and toxin extrusion (MATE) transport protein. Antimicrob Agents Chemother. 2005;49(5):1857–64. doi: 10.1128/AAC.49.5.1857-1864.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 430.McAleese F, Petersen P, Ruzin A, et al. A novel MATE family efflux pump contributes to the reduced susceptibility of laboratory-derived Staphylococcus aureus mutants to tigecycline. Antimicrob Agents Chemother. 2005;49(5):1865–71. doi: 10.1128/AAC.49.5.1865-1871.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 431.Yamada Y, Hideka K, Shiota S, et al. Gene cloning and characterization of SdrM, a chromosomally-encoded multidrug efflux pump, from Staphylococcus aureus. Biol Pharm Bull. 2006;29(3):554–6. doi: 10.1248/bpb.29.554. [DOI] [PubMed] [Google Scholar]
  • 432.Truong-Bolduc QC, Strahilevitz J, Hooper DC. NorC, a new efflux pump regulated by MgrA of Staphylococcus aureus. Antimicrob Agents Chemother. 2006;50(3):1104–7. doi: 10.1128/AAC.50.3.1104-1107.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 433.Truong-Bolduc QC, Dunman PM, Strahilevitz J, et al. MgrA is a multiple regulator of two new efflux pumps in Staphylococcus aureus. J Bacteriol. 2005;187(7):2395–405. doi: 10.1128/JB.187.7.2395-2405.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 434.Truong-Bolduc QC, Hooper DC. The transcriptional regulators NorG and MgrA modulate resistance to both quinolones and β-lactams in Staphylococcus aureus. J Bacteriol. 2007;189(8):2996–3005. doi: 10.1128/JB.01819-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 435.Narui K, Noguchi N, Wakasugi K, et al. Cloning and characterization of a novel chromosomal drug efflux gene in Staphylococcus aureus. Biol Pharm Bull. 2002;25(12):1533–6. doi: 10.1248/bpb.25.1533. [DOI] [PubMed] [Google Scholar]
  • 436.Ding Y, Onodera Y, Lee JC, et al. NorB, an efflux pump in Staphylococcus aureus MW2, contributes to bacterial fitness in abscesses. J Bacteriol. 2008;190(21):7123–9. doi: 10.1128/JB.00655-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 437.Matsuoka M, Inoue M, Endo Y, et al. Characteristic expression of three genes, msr(A), mph(C) and erm(Y), that confer resistance to macrolide antibiotics on Staphylococcus aureus. FEMS Microbiol Lett. 2003;220(2):287–93. doi: 10.1016/S0378-1097(03)00134-4. [DOI] [PubMed] [Google Scholar]
  • 438.Kehrenberg C, Schwarz S. fexA, a novel Staphylococcus lentus gene encoding resistance to florfenicol and chloramphenicol. Antimicrob Agents Chemother. 2004;48(2):615–8. doi: 10.1128/AAC.48.2.615-618.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 439.DeMarco CE, Cushing LA, Frempong-Manso E, et al. Efflux-related resistance to norfloxacin, dyes, and biocides in bloodstream isolates of Staphylococcus aureus. Antimicrob Agents Chemother. 2007;51(9):3235–9. doi: 10.1128/AAC.00430-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 440.Perez-Vazquez M, Vindel A, Marcos C, et al. Spread of invasive Spanish Staphylococcus aureus spa-type t067 associated with a high prevalence of the aminoglycoside-modifying enzyme gene ant(4′)-Ia and the efflux pump genes msrA/msrB. J Antimicrob Chemother. 2009;63(1):21–31. doi: 10.1093/jac/dkn430. [DOI] [PubMed] [Google Scholar]
  • 441.Kaatz GW, Seo SM. Effect of substrate exposure and other growth condition manipulations on norA expression. J Antimicrob Chemother. 2004;54(2):364–9. doi: 10.1093/jac/dkh341. [DOI] [PubMed] [Google Scholar]
  • 442.Huet AA, Raygada JL, Mendiratta K, et al. Multidrug efflux pump overexpression in Staphylococcus aureus after single and multiple in vitro exposures to biocides and dyes. Microbiology. 2008;154(Pt 10):3144–53. doi: 10.1099/mic.0.2008/021188-0. [DOI] [PubMed] [Google Scholar]
  • 443.Bhateja P, Purnapatre K, Dube S, et al. Characterisation of laboratory-generated vancomycin intermediate resistant Staphylococcus aureus strains. Int J Antimicrob Agents. 2006;27(3):201–11. doi: 10.1016/j.ijantimicag.2005.10.008. [DOI] [PubMed] [Google Scholar]
  • 444.Stepanovic S, Martel A, Dakic I, et al. Resistance to macrolides, lincosamides, streptogramins, and linezolid among members of the Staphylococcus sciuri group. Microb Drug Resist. 2006;12(2):115–20. doi: 10.1089/mdr.2006.12.115. [DOI] [PubMed] [Google Scholar]
  • 445.Alam MM, Kobayashi N, Uehara N, Watanabe N. Analysis on distribution and genomic diversity of high-level antiseptic resistance genes qacA and qacB in human clinical isolates of Staphylococcus aureus. Microb Drug Resist. 2003;9(2):109–21. doi: 10.1089/107662903765826697. [DOI] [PubMed] [Google Scholar]
  • 446.Alam MM, Ishino M, Kobayashi N. Analysis of genomic diversity and evolution of the low-level antiseptic resistance gene smr in Staphylococcus aureus. Microb Drug Resist. 2003;9(Suppl 1):S1–7. doi: 10.1089/107662903322541838. [DOI] [PubMed] [Google Scholar]
  • 447.Bayer AS, Kupferwasser LI, Brown MH, et al. Low-level resistance of Staphylococcus aureus to thrombin-induced platelet microbicidal protein 1 in vitro associated with qacA gene carriage is independent of multidrug efflux pump activity. Antimicrob Agents Chemother. 2006;50(7):2448–54. doi: 10.1128/AAC.00028-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 448.Jones CH, Tuckman M, Howe AY, et al. Diagnostic PCR analysis of the occurrence of methicillin and tetracycline resistance genes among Staphylococcus aureus isolates from phase 3 clinical trials of tigecycline for complicated skin and skin structure infections. Antimicrob Agents Chemother. 2006;50(2):505–10. doi: 10.1128/AAC.50.2.505-510.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 449.Strahilevitz J, Truong-Bolduc QC, Hooper DC. DX-619, a novel des-fluoro(6) quinolone manifesting low frequency of selection of resistant Staphylococcus aureus mutants: quinolone resistance beyond modification of type II topoisomerases. Antimicrob Agents Chemother. 2005;49(12):5051–7. doi: 10.1128/AAC.49.12.5051-5057.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 450.Correa JE, De Paulis A, Predari S, et al. First report of qacG, qacH and qacJ genes in Staphylococcus haemolyticus human clinical isolates. J Antimicrob Chemother. 2008;62(5):956–60. doi: 10.1093/jac/dkn327. [DOI] [PubMed] [Google Scholar]
  • 451.Lubelski J, de Jong A, van Merkerk R, et al. LmrCD is a major multidrug resistance transporter in Lactococcus lactis. Mol Microbiol. 2006;61(3):771–81. doi: 10.1111/j.1365-2958.2006.05267.x. [DOI] [PubMed] [Google Scholar]
  • 452.Lubelski J, Mazurkiewicz P, van Merkerk R, et al. ydaG and ydbA of Lactococcus lactis encode a heterodimeric ATP-binding cassette-type multidrug transporter. J Biol Chem. 2004;279(33):34449–55. doi: 10.1074/jbc.M404072200. [DOI] [PubMed] [Google Scholar]
  • 453.Zaidi AH, Bakkes PJ, Lubelski J, et al. The ABC-type multidrug resistance transporter LmrCD is responsible for an extrusion-based mechanism of bile acid resistance in Lactococcus lactis. J Bacteriol. 2008;190(22):7357–66. doi: 10.1128/JB.00485-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 454.Agustiandari H, Lubelski J, van den Berg van Saparoea HB, et al. LmrR is a transcriptional repressor of expression of the multidrug ABC transporter LmrCD in Lactococcus lactis. J Bacteriol. 2008;190(2):759–63. doi: 10.1128/JB.01151-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 455.Walther C, Rossano A, Thomann A, et al. Antibiotic resistance in Lactococcus species from bovine milk: presence of a mutated multidrug transporter mdt(A) gene in susceptible Lactococcus garvieae strains. Vet Microbiol. 2008;131(3–4):348–57. doi: 10.1016/j.vetmic.2008.03.008. [DOI] [PubMed] [Google Scholar]
  • 456.Sakamoto K, Margolles A, van Veen HW, et al. Hop resistance in the beer spoilage bacterium Lactobacillus brevis is mediated by the ATP-binding cassette multidrug transporter HorA. J Bacteriol. 2001;183(18):5371–5. doi: 10.1128/JB.183.18.5371-5375.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 457.Suzuki K, Sami M, Kadokura H, et al. Biochemical characterization of horA-independent hop resistance mechanism in Lactobacillus brevis. Int J Food Microbiol. 2002;76(3):223–30. doi: 10.1016/s0168-1605(02)00016-8. [DOI] [PubMed] [Google Scholar]
  • 458.Cauwerts K, Pasmans F, Devriese LA, et al. Cloacal Lactobacillus isolates from broilers often display resistance toward tetracycline antibiotics. Microb Drug Resist. 2006;12(4):284–8. doi: 10.1089/mdr.2006.12.284. [DOI] [PubMed] [Google Scholar]
  • 459.Roberts MC. Update on acquired tetracycline resistance genes. FEMS Microbiol Lett. 2005;245(2):195–203. doi: 10.1016/j.femsle.2005.02.034. [DOI] [PubMed] [Google Scholar]
  • 460.Ammor MS, Gueimonde M, Danielsen M, et al. Two different tetracycline resistance mechanisms, plasmid-carried tet(L) and chromosomally located transposon-associated tet(M), coexist in Lactobacillus sakei Rits 9. Appl Environ Microbiol. 2008;74(5):1394–401. doi: 10.1128/AEM.01463-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 461.Elkins CA, Mullis LB. Bile-mediated aminoglycoside sensitivity in Lactobacillus species likely results from increased membrane permeability attributable to cholic acid. Appl Environ Microbiol. 2004;70(12):7200–9. doi: 10.1128/AEM.70.12.7200-7209.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 462.Sheehan VM, Sleator RD, Fitzgerald GF, et al. Heterologous expression of BetL, a betaine uptake system, enhances the stress tolerance of Lactobacillus salivarius UCC118. Appl Environ Microbiol. 2006;72(3):2170–7. doi: 10.1128/AEM.72.3.2170-2177.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 463.Lee EW, Huda MN, Kuroda T, et al. EfrAB, an ABC multidrug efflux pump in Enterococcus faecalis. Antimicrob Agents Chemother. 2003;47(12):3733–8. doi: 10.1128/AAC.47.12.3733-3738.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 464.Dina J, Malbruny B, Leclercq R. Nonsense mutations in the lsa-like gene in Enterococcus faecalis isolates susceptible to lincosamides and streptogramins A. Antimicrob Agents Chemother. 2003;47(7):2307–9. doi: 10.1128/AAC.47.7.2307-2309.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 465.Aslangul E, Massias L, Meulemans A, et al. Acquired gentamicin resistance by permeability impairment in Enterococcus faecalis. Antimicrob Agents Chemother. 2006;50(11):3615–21. doi: 10.1128/AAC.00390-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 466.Oyamada Y, Ito H, Inoue M, et al. Topoisomerase mutations and efflux are associated with fluoroquinolone resistance in Enterococcus faecalis. J Med Microbiol. 2006;55(Pt 10):1395–401. doi: 10.1099/jmm.0.46636-0. [DOI] [PubMed] [Google Scholar]
  • 467.Singh KV, Malathum K, Murray BE. Disruption of an Enterococcus faecium species-specific gene, a homologue of acquired macrolide resistance genes of staphylococci, is associated with an increase in macrolide susceptibility. Antimicrob Agents Chemother. 2001;45(1):263–6. doi: 10.1128/AAC.45.1.263-266.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 468.Reynolds E, Cove JH. Enhanced resistance to erythromycin is conferred by the enterococcal msrC determinant in Staphylococcus aureus. J Antimicrob Chemother. 2005;55(2):260–4. doi: 10.1093/jac/dkh541. [DOI] [PubMed] [Google Scholar]
  • 469.Werner G, Hildebrandt B, Witte W. The newly described msrC gene is not equally distributed among all isolates of Enterococcus faecium. Antimicrob Agents Chemother. 2001;45(12):3672–3. doi: 10.1128/AAC.45.12.3672-3673.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 470.Nishioka T, Ogawa W, Kuroda T, et al. Gene cloning and characterization of EfmA, a multidrug efflux pump, from Enterococcus faecium. Biol Pharm Bull. 2009;32(3):483–8. doi: 10.1248/bpb.32.483. [DOI] [PubMed] [Google Scholar]
  • 471.Jumbe NL, Louie A, Miller MH, et al. Quinolone efflux pumps play a central role in emergence of fluoroquinolone resistance in Streptococcus pneumoniae. Antimicrob Agents Chemother. 2006;50(1):310–7. doi: 10.1128/AAC.50.1.310-317.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 472.Sevillano D, Aguilar L, Alou L, et al. Effects of antimicrobials on the competitive growth of Streptococcus pneumoniae: a pharmacodynamic in vitro model approach to selection of resistant populations. J Antimicrob Chemother. 2006;58(4):794–801. doi: 10.1093/jac/dkl307. [DOI] [PubMed] [Google Scholar]
  • 473.Felmingham D, Canton R, Jenkins SG. Regional trends in β-lactam, macrolide, fluoroquinolone and telithromycin resistance among Streptococcus pneumoniae isolates 2001–2004. J Infect. 2007;55(2):111–8. doi: 10.1016/j.jinf.2007.04.006. [DOI] [PubMed] [Google Scholar]
  • 474.Wierzbowski AK, Swedlo D, Boyd D, et al. Molecular epidemiology and prevalence of macrolide efflux genes mef(A) and mef(E) in Streptococcus pneumoniae obtained in Canada from 1997 to 2002. Antimicrob Agents Chemother. 2005;49(3):1257–61. doi: 10.1128/AAC.49.3.1257-1261.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 475.Song JH, Chang HH, Suh JY, et al. Macrolide resistance and genotypic characterization of Streptococcus pneumoniae in Asian countries: a study of the Asian Network for Surveillance of Resistant Pathogens (ANSORP) J Antimicrob Chemother. 2004;53(3):457–63. doi: 10.1093/jac/dkh118. [DOI] [PubMed] [Google Scholar]
  • 476.Farrell DJ, Morrissey I, Bakker S, et al. Molecular epidemiology of multiresistant Streptococcus pneumoniae with both erm(B)- and mef(A)-mediated macrolide resistance. J Clin Microbiol. 2004;42(2):764–8. doi: 10.1128/JCM.42.2.764-768.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 477.Bacciaglia A, Brenciani A, Varaldo PE, et al. SmaI typeability and tetracycline susceptibility and resistance in Streptococcus pyogenes isolates with efflux-mediated erythromycin resistance. Antimicrob Agents Chemother. 2007;51(8):3042–3. doi: 10.1128/AAC.00249-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 478.Brenwald NP, Appelbaum P, Davies T, et al. Evidence for efflux pumps, other than PmrA, associated with fluoroquinolone resistance in Streptococcus pneumoniae. Clin Microbiol Infect. 2003;9(2):140–3. doi: 10.1046/j.1469-0691.2003.00482.x. [DOI] [PubMed] [Google Scholar]
  • 479.Martinez-Garriga B, Vinuesa T, Hernandez-Borrell J, et al. The contribution of efflux pumps to quinolone resistance in Streptococcus pneumoniae clinical isolates. Int J Med Microbiol. 2007;297(3):187–95. doi: 10.1016/j.ijmm.2007.01.004. [DOI] [PubMed] [Google Scholar]
  • 480.Schurek KN, Adam HJ, Siemens CG, et al. Are fluoroquinolone-susceptible isolates of Streptococcus pneumoniae really susceptible? A comparison of resistance mechanisms in Canadian isolates from 1997 and 2003. J Antimicrob Chemother. 2005;56(4):769–72. doi: 10.1093/jac/dki315. [DOI] [PubMed] [Google Scholar]
  • 481.Canton R, Mazzariol A, Morosini M-I, et al. Telithromycin activity is reduced by efflux in Streptococcus pyogenes. J Antimicrob Chemother. 2005;55(4):489–95. doi: 10.1093/jac/dki033. [DOI] [PubMed] [Google Scholar]
  • 482.Hisanaga T, Hoban DJ, Zhanel GG. Mechanisms of resistance to telithromycin in Streptococcus pneumoniae. J Antimicrob Chemother. 2005;56(3):447–50. doi: 10.1093/jac/dki249. [DOI] [PubMed] [Google Scholar]
  • 483.Marrer E, Satoh AT, Johnson MM, et al. Global transcriptome analysis of the responses of a fluoroquinolone-resistant Streptococcus pneumoniae mutant and its parent to ciprofloxacin. Antimicrob Agents Chemother. 2006;50(1):269–78. doi: 10.1128/AAC.50.1.269-278.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 484.Marrer E, Schad K, Satoh AT, et al. Involvement of the putative ATP-dependent efflux proteins PatA and PatB in fluoroquinolone resistance of a multidrug-resistant mutant of Streptococcus pneumoniae. Antimicrob Agents Chemother. 2006;50(2):685–93. doi: 10.1128/AAC.50.2.685-693.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 485.Garvey MI, Piddock LJ. The efflux pump inhibitor reserpine selects multidrug-resistant Streptococcus pneumoniae strains that overexpress the ABC transporters PatA and PatB. Antimicrob Agents Chemother. 2008;52(5):1677–85. doi: 10.1128/AAC.01644-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 486.Avrain L, Garvey M, Mesaros N, et al. Selection of quinolone resistance in Streptococcus pneumoniae exposed in vitro to subinhibitory drug concentrations. J Antimicrob Chemother. 2007;60(5):965–72. doi: 10.1093/jac/dkm292. [DOI] [PubMed] [Google Scholar]
  • 487.Robertson GT, Doyle TB, Lynch AS. Use of an efflux-deficient Streptococcus pneumoniae strain panel to identify ABC-class multidrug transporters involved in intrinsic resistance to antimicrobial agents. Antimicrob Agents Chemother. 2005;49(11):4781–3. doi: 10.1128/AAC.49.11.4781-4783.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 488.Becker P, Hakenbeck R, Henrich B. An ABC transporter of Streptococcus pneumoniae involved in susceptibility to vancoresmycin and bacitracin. Antimicrob Agents Chemother. 2009;53(5):2034–41. doi: 10.1128/AAC.01485-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 489.Brenciani A, Ojo KK, Monachetti A, et al. Distribution and molecular analysis of mef(A)-containing elements in tetracycline-susceptible and -resistant Streptococcus pyogenes clinical isolates with efflux-mediated erythromycin resistance. J Antimicrob Chemother. 2004;54(6):991–8. doi: 10.1093/jac/dkh481. [DOI] [PubMed] [Google Scholar]
  • 490.D’Ercole S, Petrelli D, Prenna M, et al. Distribution of mef(A)-containing genetic elements in erythromycin-resistant isolates of Streptococcus pyogenes from Italy. Clin Microbiol Infect. 2005;11(11):927–30. doi: 10.1111/j.1469-0691.2005.01250.x. [DOI] [PubMed] [Google Scholar]
  • 491.Santagati M, Iannelli F, Cascone C, et al. The novel conjugative transposon tn1207.3 carries the macrolide efflux gene mef(A) in Streptococcus pyogenes. Microb Drug Resist. 2003;9(3):243–7. doi: 10.1089/107662903322286445. [DOI] [PubMed] [Google Scholar]
  • 492.Figueiredo TA, Aguiar SI, Melo-Cristino J, et al. DNA methylase activity as a marker for the presence of a family of phage-like elements conferring efflux-mediated macrolide resistance in streptococci. Antimicrob Agents Chemother. 2006;50(11):3689–94. doi: 10.1128/AAC.00782-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 493.Giovanetti E, Brenciani A, Vecchi M, et al. Prophage association of mef(A) elements encoding efflux-mediated erythromycin resistance in Streptococcus pyogenes. J Antimicrob Chemother. 2005;55(4):445–51. doi: 10.1093/jac/dki049. [DOI] [PubMed] [Google Scholar]
  • 494.Jonsson M, Swedberg G. Macrolide resistance can be transferred by conjugation from viridans streptococci to Streptococcus pyogenes. Int J Antimicrob Agents. 2006;28(2):101–3. doi: 10.1016/j.ijantimicag.2006.02.023. [DOI] [PubMed] [Google Scholar]
  • 495.Marimon JM, Valiente A, Ercibengoa M, et al. Erythromycin resistance and genetic elements carrying macrolide efflux genes in Streptococcus agalactiae. Antimicrob Agents Chemother. 2005;49(12):5069–74. doi: 10.1128/AAC.49.12.5069-5074.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 496.Cousin SL, Jr, Whittington WL, Roberts MC. Acquired macrolide resistance genes and the 1 bp deletion in the mtrR promoter in Neisseria gonorrhoeae. J Antimicrob Chemother. 2003;51(1):131–3. doi: 10.1093/jac/dkg040. [DOI] [PubMed] [Google Scholar]
  • 497.Ojo KK, Ulep C, Van Kirk N, et al. The mef(A) gene predominates among seven macrolide resistance genes identified in Gram-negative strains representing 13 genera, isolated from healthy Portuguese children. Antimicrob Agents Chemother. 2004;48(9):3451–6. doi: 10.1128/AAC.48.9.3451-3456.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 498.Daly MM, Doktor S, Flamm R, et al. Characterization and prevalence of MefA, MefE, and the associated msr(D) gene in Streptococcus pneumoniae clinical isolates. J Clin Microbiol. 2004;42(8):3570–4. doi: 10.1128/JCM.42.8.3570-3574.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 499.Ambrose KD, Nisbet R, Stephens DS. Macrolide efflux in Streptococcus pneumoniae is mediated by a dual efflux pump (mel and mef) and is erythromycin inducible. Antimicrob Agents Chemother. 2005;49(10):4203–9. doi: 10.1128/AAC.49.10.4203-4209.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 500.Del Grosso M, Scotto d’Abusco A, Iannelli F, et al. Tn2009, a Tn916-like element containing mef(E) in Streptococcus pneumoniae. Antimicrob Agents Chemother. 2004;48(6):2037–42. doi: 10.1128/AAC.48.6.2037-2042.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 501.Del Grosso M, Camilli R, Iannelli F, et al. The mef(E)-carrying genetic element (MEGA) of Streptococcus pneumoniae: insertion sites and association with other genetic elements. Antimicrob Agents Chemother. 2006;50(10):3361–6. doi: 10.1128/AAC.00277-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 502.Cochetti I, Vecchi M, Mingoia M, et al. Molecular characterization of pneumococci with efflux-mediated erythromycin resistance and identification of a novel mef gene subclass, mef(I) Antimicrob Agents Chemother. 2005;49(12):4999–5006. doi: 10.1128/AAC.49.12.4999-5006.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 503.Mingoia M, Vecchi M, Cochetti I, et al. Composite structure of Streptococcus pneumoniae containing the erythromycin efflux resistance gene mefI and the chloramphenicol resistance gene catQ. Antimicrob Agents Chemother. 2007;51(11):3983–7. doi: 10.1128/AAC.00790-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 504.Price CE, Reid SJ, Driessen AJ, et al. The Bifidobacterium longum NCIMB 702259T ctr gene codes for a novel cholate transporter. Appl Environ Microbiol. 2006;72(1):923–6. doi: 10.1128/AEM.72.1.923-926.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 505.De Dea Lindner J, Canchaya C, Zhang Z, et al. Exploiting Bifidobacterium genomes: the molecular basis of stress response. Int J Food Microbiol. 2007;120(1–2):13–24. doi: 10.1016/j.ijfoodmicro.2007.06.016. [DOI] [PubMed] [Google Scholar]
  • 506.Margolles A, Moreno JA, van Sinderen D, et al. Macrolide resistance mediated by a Bifidobacterium breve membrane protein. Antimicrob Agents Chemother. 2005;49(10):4379–81. doi: 10.1128/AAC.49.10.4379-4381.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 507.Margolles A, Florez AB, Moreno JA, et al. Two membrane proteins from Bifidobacterium breve UCC2003 constitute an ABC-type multidrug transporter. Microbiology. 2006;152(Pt 12):3497–505. doi: 10.1099/mic.0.29097-0. [DOI] [PubMed] [Google Scholar]
  • 508.Ruiz L, Coute Y, Sanchez B, de Los Reyes-Gavilan CG, et al. The cell-envelope proteome of Bifidobacterium longum in an in vitro bile environment. Microbiology. 2009;155(Pt 3):957–67. doi: 10.1099/mic.0.024273-0. [DOI] [PubMed] [Google Scholar]
  • 509.Viveiros M, Leandro C, Amaral L. Mycobacterial efflux pumps and chemotherapeutic implications. Int J Antimicrob Agents. 2003;22(3):274–8. doi: 10.1016/s0924-8579(03)00208-5. [DOI] [PubMed] [Google Scholar]
  • 510.De Rossi E, Ainsa JA, Riccardi G. Role of mycobacterial efflux transporters in drug resistance: an unresolved question. FEMS Microbiol Rev. 2006;30(1):36–52. doi: 10.1111/j.1574-6976.2005.00002.x. [DOI] [PubMed] [Google Scholar]
  • 511.Nguyen L, Thompson CJ. Foundations of antibiotic resistance in bacterial physiology: the mycobacterial paradigm. Trends Microbiol. 2006;14(7):304–12. doi: 10.1016/j.tim.2006.05.005. [DOI] [PubMed] [Google Scholar]
  • 512.Louw GE, Warren RM, Gey van Pittius NC, et al. A balancing act: efflux/influx in mycobacterial drug resistance. Antimicrob Agents Chemother. 2009 doi: 10.1128/AAC.01577-08. E-published May 18, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 513.Cole ST, Brosch R, Parkhill J, Garnier T, Churcher C, Harris D, et al. Deciphering the biology of Mycobacterium tuberculosis from the complete genome sequence. Nature. 1998;393(6685):537–44. doi: 10.1038/31159. [DOI] [PubMed] [Google Scholar]
  • 514.Li X-Z, Zhang L, Nikaido H. Efflux pump-mediated intrinsic drug resistance in Mycobacterium smegmatis. Antimicrob Agents Chemother. 2004;48(7):2415–23. doi: 10.1128/AAC.48.7.2415-2423.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 515.Escribano I, Rodriguez JC, Llorca B, et al. Importance of the efflux pump systems in the resistance of Mycobacterium tuberculosis to fluoroquinolones and linezolid. Chemotherapy. 2007;53(6):397–401. doi: 10.1159/000109769. [DOI] [PubMed] [Google Scholar]
  • 516.Amaral L, Martins M, Viveiros M. Enhanced killing of intracellular multidrug-resistant Mycobacterium tuberculosis by compounds that affect the activity of efflux pumps. J Antimicrob Chemother. 2007;59(6):1237–46. doi: 10.1093/jac/dkl500. [DOI] [PubMed] [Google Scholar]
  • 517.Domenech P, Reed MB, Barry CE., 3rd Contribution of the Mycobacterium tuberculosis MmpL protein family to virulence and drug resistance. Infect Immun. 2005;73(6):3492–501. doi: 10.1128/IAI.73.6.3492-3501.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 518.Pasca MR, Guglierame P, De Rossi E, et al. mmpL7 gene of Mycobacterium tuberculosis is responsible for isoniazid efflux in Mycobacterium smegmatis. Antimicrob Agents Chemother. 2005;49(11):4775–7. doi: 10.1128/AAC.49.11.4775-4777.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 519.Domenech P, Reed MB, Dowd CS, et al. The role of MmpL8 in sulfatide biogenesis and virulence of Mycobacterium tuberculosis. J Biol Chem. 2004;279(20):21257–65. doi: 10.1074/jbc.M400324200. [DOI] [PubMed] [Google Scholar]
  • 520.Braibant M, Gilot P, Content J. The ATP binding cassette (ABC) transport systems of Mycobacterium tuberculosis. FEMS Microbiol Rev. 2000;24(4):449–67. doi: 10.1111/j.1574-6976.2000.tb00550.x. [DOI] [PubMed] [Google Scholar]
  • 521.Pasca MR, Guglierame P, Arcesi F, et al. Rv2686c-Rv2687c-Rv2688c, an ABC fluoroquinolone efflux pump in Mycobacterium tuberculosis. Antimicrob Agents Chemother. 2004;48(8):3175–8. doi: 10.1128/AAC.48.8.3175-3178.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 522.De Rossi E, Arrigo P, Bellinzoni M, et al. The multidrug transporters belonging to major facilitator superfamily in Mycobacterium tuberculosis. Mol Med. 2002;8(11):714–24. [PMC free article] [PubMed] [Google Scholar]
  • 523.Siddiqi N, Das R, Pathak N, et al. Mycobacterium tuberculosis isolate with a distinct genomic identity overexpresses a Tap-like efflux pump. Infection. 2004;32(2):109–11. doi: 10.1007/s15010-004-3097-x. [DOI] [PubMed] [Google Scholar]
  • 524.Ramon-Garcia S, Martin C, De Rossi E, et al. Contribution of the Rv2333c efflux pump (the Stp protein) from Mycobacterium tuberculosis to intrinsic antibiotic resistance in Mycobacterium bovis BCG. J Antimicrob Chemother. 2007;59(3):544–7. doi: 10.1093/jac/dkl510. [DOI] [PubMed] [Google Scholar]
  • 525.Morris RP, Nguyen L, Gatfield J, et al. Ancestral antibiotic resistance in Mycobacterium tuberculosis. Proc Natl Acad Sci USA. 2005;102(34):12200–5. doi: 10.1073/pnas.0505446102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 526.Colangeli R, Helb D, Sridharan S, et al. The Mycobacterium tuberculosis iniA gene is essential for activity of an efflux pump that confers drug tolerance to both isoniazid and ethambutol. Mol Microbiol. 2005;55(6):1829–40. doi: 10.1111/j.1365-2958.2005.04510.x. [DOI] [PubMed] [Google Scholar]
  • 527.Sullivan TJ, Truglio JJ, Boyne ME, et al. High affinity InhA inhibitors with activity against drug-resistant strains of Mycobacterium tuberculosis. ACS Chem Biol. 2006;1(1):43–53. doi: 10.1021/cb0500042. [DOI] [PubMed] [Google Scholar]
  • 528.Buroni S, Manina G, Guglierame P, et al. LfrR is a repressor that regulates expression of the efflux pump LfrA in Mycobacterium smegmatis. Antimicrob Agents Chemother. 2006;50(12):4044–52. doi: 10.1128/AAC.00656-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 529.Aarestrup F. Antimicrobial resistance in bacteria of animal origin. Washington DC: ASM Press; 2006. [DOI] [PubMed] [Google Scholar]
  • 530.Prescott JF. Antimicrobial use in food and companion animals. Anim Health Res Rev. 2008;9(2):127–33. doi: 10.1017/S1466252308001473. [DOI] [PubMed] [Google Scholar]
  • 531.Teresa Tejedor M, Martin JL, Navia M, et al. Mechanisms of fluoroquinolone resistance in Pseudomonas aeruginosa isolates from canine infections. Vet Microbiol. 2003;94(4):295–301. doi: 10.1016/s0378-1135(03)00129-9. [DOI] [PubMed] [Google Scholar]
  • 532.Zhao S, Maurer JJ, Hubert S, et al. Antimicrobial susceptibility and molecular characterization of avian pathogenic Escherichia coli isolates. Vet Microbiol. 2005;107(3–4):215. doi: 10.1016/j.vetmic.2005.01.021. [DOI] [PubMed] [Google Scholar]
  • 533.Liu J, Keelan P, Bennett PM, et al. Characterization of a novel macrolide efflux gene, mef(B), found linked to sul3 in porcine Escherichia coli. J Antimicrob Chemother. 2009;63(3):423–6. doi: 10.1093/jac/dkn523. [DOI] [PubMed] [Google Scholar]
  • 534.Chuanchuen R, Wannaprasat W, Ajariyakhajorn K, et al. Role of the MexXY multidrug efflux pump in moderate aminoglycoside resistance in Pseudomonas aeruginosa isolates from Pseudomonas mastitis. Microbiol Immunol. 2008;52(8):392–8. doi: 10.1111/j.1348-0421.2008.00051.x. [DOI] [PubMed] [Google Scholar]
  • 535.White DG, Zhao S, McDermott PF, et al. Characterization of integron mediated antimicrobial resistance in Salmonella isolated from diseased swine. Can J Vet Res. 2003;67(1):39–47. [PMC free article] [PubMed] [Google Scholar]
  • 536.Payot S, Avrain L, Magras C, et al. Relative contribution of target gene mutation and efflux to fluoroquinolone and erythromycin resistance, in French poultry and pig isolates of Campylobacter coli. Int J Antimicrob Agents. 2004;23(5):468–72. doi: 10.1016/j.ijantimicag.2003.12.008. [DOI] [PubMed] [Google Scholar]
  • 537.Du X, Xia C, Shen J, et al. Characterization of florfenicol resistance among calf pathogenic Escherichia coli. FEMS Microbiol Lett. 2004;236(2):183–9. doi: 10.1016/j.femsle.2004.05.013. [DOI] [PubMed] [Google Scholar]
  • 538.Moreira MA, Oliveira JA, Teixeira LM, et al. Detection of a chloramphenicol efflux system in Escherichia coli isolated from poultry carcass. Vet Microbiol. 2005;109(1–2):75–81. doi: 10.1016/j.vetmic.2005.04.012. [DOI] [PubMed] [Google Scholar]
  • 539.Thorrold CA, Letsoalo ME, Duse AG, et al. Efflux pump activity in fluoroquinolone and tetracycline resistant Salmonella and E. coli implicated in reduced susceptibility to household antimicrobial cleaning agents. Int J Food Microbiol. 2007;113(3):315–20. doi: 10.1016/j.ijfoodmicro.2006.08.008. [DOI] [PubMed] [Google Scholar]
  • 540.Sawant AA, Gillespie BE, Oliver SP. Antimicrobial susceptibility of coagulase-negative Staphylococcus species isolated from bovine milk. Vet Microbiol. 2009;134(1–2):73–81. doi: 10.1016/j.vetmic.2008.09.006. [DOI] [PubMed] [Google Scholar]
  • 541.Miles TD, McLaughlin W, Brown PD. Antimicrobial resistance of Escherichia coli isolates from broiler chickens and humans. BMC Vet Res. 2006;2:7. doi: 10.1186/1746-6148-2-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 542.Kehrenberg C, Schwarz S. Plasmid-borne florfenicol resistance in Pasteurella multocida. J Antimicrob Chemother. 2005;55(5):773–5. doi: 10.1093/jac/dki102. [DOI] [PubMed] [Google Scholar]
  • 543.Kehrenberg C, Catry B, Haesebrouck F, et al. tet(L)-mediated tetracycline resistance in bovine Mannheimia and Pasteurella isolates. J Antimicrob Chemother. 2005;56(2):403–6. doi: 10.1093/jac/dki210. [DOI] [PubMed] [Google Scholar]
  • 544.Blanco M, Gutierrez-Martin CB, Rodriguez-Ferri EF, et al. Distribution of tetracycline resistance genes in Actinobacillus pleuropneumoniae isolates from Spain. Antimicrob Agents Chemother. 2006;50(2):702–8. doi: 10.1128/AAC.50.2.702-708.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 545.Gil H, Platz GJ, Forestal CA, et al. Deletion of TolC orthologs in Francisella tularensis identifies roles in multidrug resistance and virulence. Proc Natl Acad Sci USA. 2006;103(34):12897–902. doi: 10.1073/pnas.0602582103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 546.Udani RA, Levy SB. MarA-like regulator of multidrug resistance in Yersinia pestis. Antimicrob Agents Chemother. 2006;50(9):2971–5. doi: 10.1128/AAC.00015-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 547.Schluter A, Szczepanowski R, Kurz N, et al. Erythromycin resistance-conferring plasmid pRSB105, isolated from a sewage treatment plant, harbors a new macrolide resistance determinant, an integron-containing Tn402-like element, and a large region of unknown function. Appl Environ Microbiol. 2007;73(6):1952–60. doi: 10.1128/AEM.02159-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 548.Kim SH, Wei CI. Antibiotic resistance and Caco-2 cell invasion of Pseudomonas aeruginosa isolates from farm environments and retail products. Int J Food Microbiol. 2007;115(3):356–63. doi: 10.1016/j.ijfoodmicro.2006.12.033. [DOI] [PubMed] [Google Scholar]
  • 549.Gaze WH, Abdouslam N, Hawkey PM, et al. Incidence of class 1 integrons in a quaternary ammonium compound-polluted environment. Antimicrob Agents Chemother. 2005;49(5):1802–7. doi: 10.1128/AAC.49.5.1802-1807.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 550.Michel C, Matte-Tailliez O, Kerouault B, et al. Resistance pattern and assessment of phenicol agents’ minimum inhibitory concentration in multiple drug resistant Chryseobacterium isolates from fish and aquatic habitats. J Appl Microbiol. 2005;99(2):323–32. doi: 10.1111/j.1365-2672.2005.02592.x. [DOI] [PubMed] [Google Scholar]
  • 551.Riesenfeld CS, Goodman RM, Handelsman J. Uncultured soil bacteria are a reservoir of new antibiotic resistance genes. Environ Microbiol. 2004;6(9):981–9. doi: 10.1111/j.1462-2920.2004.00664.x. [DOI] [PubMed] [Google Scholar]
  • 552.Fantinatti-Garboggini F, Almeida R, Portillo Vdo A, et al. Drug resistance in Chromobacterium violaceum. Genet Mol Res. 2004;3(1):134–47. [PubMed] [Google Scholar]
  • 553.Mah TF, Pitts B, Pellock B, et al. A genetic basis for Pseudomonas aeruginosa biofilm antibiotic resistance. Nature. 2003;426(6964):306–10. doi: 10.1038/nature02122. [DOI] [PubMed] [Google Scholar]
  • 554.Lewis K. Persister cells, dormancy and infectious disease. Nat Rev Microbiol. 2007;5(1):48–56. doi: 10.1038/nrmicro1557. [DOI] [PubMed] [Google Scholar]
  • 555.Maira-Litran T, Allison DG, Gilbert P. An evaluation of the potential of the multiple antibiotic resistance operon (mar) and the multidrug efflux pump acrAB to moderate resistance towards ciprofloxacin in Escherichia coli biofilms. J Antimicrob Chemother. 2000;45(6):789–95. doi: 10.1093/jac/45.6.789. [DOI] [PubMed] [Google Scholar]
  • 556.Brooun A, Liu S, Lewis K. A dose-response study of antibiotic resistance in Pseudomonas aeruginosa biofilms. Antimicrob Agents Chemother. 2000;44(3):640–6. doi: 10.1128/aac.44.3.640-646.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 557.De Kievit TR, Parkins MD, Gillis RJ, et al. Multidrug efflux pumps: expression patterns and contribution to antibiotic resistance in Pseudomonas aeruginosa biofilms. Antimicrob Agents Chemother. 2001;45(6):1761–70. doi: 10.1128/AAC.45.6.1761-1770.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 558.Sanchez P, Linares JF, Ruiz-Diez B, et al. Fitness of in vitro selected Pseudomonas aeruginosa nalB and nfxB multidrug resistant mutants. J Antimicrob Chemother. 2002;50(5):657–64. doi: 10.1093/jac/dkf185. [DOI] [PubMed] [Google Scholar]
  • 559.Gillis RJ, White KG, Choi K-H, et al. Molecular basis of azithromycin-resistant Pseudomonas aeruginosa biofilms. Antimicrob Agents Chemother. 2005;49(9):3858–67. doi: 10.1128/AAC.49.9.3858-3867.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 560.Pamp SJ, Gjermansen M, Johansen HK, et al. Tolerance to the antimicrobial peptide colistin in Pseudomonas aeruginosa biofilms is linked to metabolically active cells, and depends on the pmr and mexAB-oprM genes. Mol Microbiol. 2008;68(1):223–40. doi: 10.1111/j.1365-2958.2008.06152.x. [DOI] [PubMed] [Google Scholar]
  • 561.Zhang L, Mah TF. Involvement of a novel efflux system in biofilm-specific resistance to antibiotics. J Bacteriol. 2008;190(13):4447–52. doi: 10.1128/JB.01655-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 562.Lynch SV, Dixon L, Benoit MR, et al. Role of the rapA gene in controlling antibiotic resistance of Escherichia coli biofilms. Antimicrob Agents Chemother. 2007;51(10):3650–8. doi: 10.1128/AAC.00601-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 563.Tabak M, Scher K, Hartog E, et al. Effect of triclosan on Salmonella Typhimurium at different growth stages and in biofilms. FEMS Microbiol Lett. 2007;267(2):200–6. doi: 10.1111/j.1574-6968.2006.00547.x. [DOI] [PubMed] [Google Scholar]
  • 564.Pumbwe L, Skilbeck CA, Nakano V, et al. Bile salts enhance bacterial co-aggregation, bacterial-intestinal epithelial cell adhesion, biofilm formation and antimicrobial resistance of Bacteroides fragilis. Microb Pathog. 2007;43(2–3):78–87. doi: 10.1016/j.micpath.2007.04.002. [DOI] [PubMed] [Google Scholar]
  • 565.Weigel LM, Donlan RM, Shin DH, et al. High-level vancomycin-resistant Staphylococcus aureus isolates associated with a polymicrobial biofilm. Antimicrob Agents Chemother. 2007;51(1):231–8. doi: 10.1128/AAC.00576-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 566.Ma D, Cook DN, Alberti M, et al. Genes acrA and acrB encode a stress-induced efflux system of Escherichia coli. Mol Microbiol. 1995;16(1):45–55. doi: 10.1111/j.1365-2958.1995.tb02390.x. [DOI] [PubMed] [Google Scholar]
  • 567.Bina JE, Mekalanos JJ. Vibrio cholerae tolC is required for bile resistance and colonization. Infect Immun. 2001;69(7):4681–5. doi: 10.1128/IAI.69.7.4681-4685.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 568.Jerse AE, Sharma ND, Simms AN, et al. A gonococcal efflux pump system enhances bacterial survival in a female mouse model of genital tract infection. Infect Immun. 2003;71(10):5576–82. doi: 10.1128/IAI.71.10.5576-5582.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 569.Lin J, Cagliero C, Guo B, Barton YW, Maurel MC, Payot S, et al. Bile salts modulate expression of the CmeABC multidrug efflux pump in Campylobacter jejuni. J Bacteriol. 2005;187(21):7417–24. doi: 10.1128/JB.187.21.7417-7424.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 570.Lin J, Martinez AL. Effect of efflux pump inhibitors on bile resistance and in vivo colonization of Campylobacter jejuni. J Antimicrob Chemother. 2006;58(5):966–72. doi: 10.1093/jac/dkl374. [DOI] [PubMed] [Google Scholar]
  • 571.Elkins CA, Mullis LB. Mammalian steroid hormones are substrates for the major RND- and MFS-type tripartite multidrug efflux pumps of Escherichia coli. J Bacteriol. 2006;188(3):1191–5. doi: 10.1128/JB.188.3.1191-1195.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 572.Piddock LJ. Multidrug-resistance efflux pumps - not just for resistance. Nat Rev Microbiol. 2006;4(8):629–36. doi: 10.1038/nrmicro1464. [DOI] [PubMed] [Google Scholar]
  • 573.Poole K. Bacterial multidrug efflux pumps serve other functions. Microbe. 2008;3(4):179–185. [Google Scholar]
  • 574.Nishino K, Nikaido E, Yamaguchi A. Regulation and physiological function of multidrug efflux pumps in Escherichia coli and Salmonella. Biochim Biophys Acta. 2009 doi: 10.1016/j.bbapap.2009.02.002. Epub ahead of print. [DOI] [PubMed] [Google Scholar]
  • 575.Krulwich TA, Lewinson O, Padan E, et al. Do physiological roles foster persistence of drug/multidrug-efflux transporters? A case study. Nat Rev Microbiol. 2005;3(7):566–72. doi: 10.1038/nrmicro1181. [DOI] [PubMed] [Google Scholar]
  • 576.Elkins CA, Mullis LB. Substrate competition studies using whole-cell accumulation assays with the major tripartite multidrug efflux pumps of Escherichia coli. Antimicrob Agents Chemother. 2007;51(3):923–9. doi: 10.1128/AAC.01048-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 577.Giuliodori AM, Gualerzi CO, Soto S, et al. Review on bacterial stress topics. Ann NY Acad Sci. 2007;1113:95–104. doi: 10.1196/annals.1391.008. [DOI] [PubMed] [Google Scholar]
  • 578.Jeannot K, Sobel ML, El Garch F, et al. Plesiat P. Induction of the MexXY efflux pump in Pseudomonas aeruginosa is dependent on drug-ribosome interaction. J Bacteriol. 2005;187(15):5341–6. doi: 10.1128/JB.187.15.5341-5346.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 579.Morita Y, Sobel ML, Poole K. Antibiotic inducibility of the MexXY multidrug efflux system of Pseudomonas aeruginosa: involvement of the antibiotic-inducible PA5471 gene product. J Bacteriol. 2006;188(5):1847–55. doi: 10.1128/JB.188.5.1847-1855.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 580.Fraud S, Campigotto AJ, Chen Z, et al. MexCD-OprJ multidrug efflux system of Pseudomonas aeruginosa: involvement in chlorhexidine resistance and induction by membrane-damaging agents dependent upon the AlgU stress response sigma factor. Antimicrob Agents Chemother. 2008;52(12):4478–82. doi: 10.1128/AAC.01072-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 581.Folster JP, Johnson PJ, Jackson L, et al. MtrR modulates rpoH expression and levels of antimicrobial resistance in Neisseria gonorrhoeae. J Bacteriol. 2009;191(1):287–97. doi: 10.1128/JB.01165-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 582.Bleuel C, Grosse C, Taudte N, et al. TolC is involved in enterobactin efflux across the outer membrane of Escherichia coli. J Bacteriol. 2005;187(19):6701–7. doi: 10.1128/JB.187.19.6701-6707.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 583.Helling RB, Janes BK, Kimball H, Tran T, Bundesmann M, Check P, et al. Toxic waste disposal in Escherichia coli. J Bacteriol. 2002;184(13):3699–703. doi: 10.1128/JB.184.13.3699-3703.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 584.Hirakata Y, Srikumar R, Poole K, et al. Multidrug efflux systems play an important role in the invasiveness of Pseudomonas aeruginosa. J Exp Med. 2002;196(1):109–18. doi: 10.1084/jem.20020005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 585.Bunikis I, Denker K, Ostberg Y, et al. An RND-type efflux system in Borrelia burgdorferi is involved in virulence and resistance to antimicrobial compounds. PLoS Pathog. 2008;4(2):e1000009. doi: 10.1371/journal.ppat.1000009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 586.Hocquet D, Bertrand X, Kohler T, et al. Genetic and phenotypic variations of a resistant Pseudomonas aeruginosa epidemic clone. Antimicrob Agents Chemother. 2003;47(6):1887–94. doi: 10.1128/AAC.47.6.1887-1894.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 587.Salunkhe P, Smart CH, Morgan JA, et al. A cystic fibrosis epidemic strain of Pseudomonas aeruginosa displays enhanced virulence and antimicrobial resistance. J Bacteriol. 2005;187(14):4908–20. doi: 10.1128/JB.187.14.4908-4920.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 588.Linares JF, Lopez JA, Camafeita E, et al. Overexpression of the multidrug efflux pumps MexCD-OprJ and MexEF-OprN is associated with a reduction of type III secretion in Pseudomonas aeruginosa. J Bacteriol. 2005;187(4):1384–91. doi: 10.1128/JB.187.4.1384-1391.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 589.Jeannot K, Elsen S, Kohler T, et al. Resistance and virulence of Pseudomonas aeruginosa clinical strains overproducing the MexCD-OprJ efflux pump. Antimicrob Agents Chemother. 2008;52(7):2455–62. doi: 10.1128/AAC.01107-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 590.Kang H, Gross DC. Characterization of a resistance-nodulation-cell division transporter system associated with the syr-syp genomic island of Pseudomonas syringae pv. syringae. Appl Environ Microbiol. 2005;71(9):5056–65. doi: 10.1128/AEM.71.9.5056-5065.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 591.Alonso A, Morales G, Escalante R, et al. Overexpression of the multidrug efflux pump SmeDEF impairs Stenotrophomonas maltophilia physiology. J Antimicrob Chemother. 2004;53(3):432–4. doi: 10.1093/jac/dkh074. [DOI] [PubMed] [Google Scholar]
  • 592.Kugelberg E, Lofmark S, Wretlind B, et al. Reduction of the fitness burden of quinolone resistance in Pseudomonas aeruginosa. J Antimicrob Chemother. 2005;55(1):22–30. doi: 10.1093/jac/dkh505. [DOI] [PubMed] [Google Scholar]
  • 593.Komp Lindgren P, Marcusson LL, et al. Biological cost of single and multiple norfloxacin resistance mutations in Escherichia coli implicated in urinary tract infections. Antimicrob Agents Chemother. 2005;49(6):2343–51. doi: 10.1128/AAC.49.6.2343-2351.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 594.Yamanaka H, Kobayashi H, Takahashi E, et al. MacAB is involved in the secretion of Escherichia coli heat-stable enterotoxin II. J Bacteriol. 2008;190(23):7693–8. doi: 10.1128/JB.00853-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 595.Brown DG, Swanson JK, Allen C. Two host-induced Ralstonia solanacearum genes, acrA and dinF, encode multidrug efflux pumps and contribute to bacterial wilt virulence. Appl Environ Microbiol. 2007;73(9):2777–86. doi: 10.1128/AEM.00984-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 596.Burse A, Weingart H, Ullrich MS. The phytoalexin-inducible multidrug efflux pump AcrAB contributes to virulence in the fire blight pathogen, Erwinia amylovora. Mol Plant Microbe Interact. 2004;17(1):43–54. doi: 10.1094/MPMI.2004.17.1.43. [DOI] [PubMed] [Google Scholar]
  • 597.Barabote RD, Johnson OL, Zetina E, et al. Erwinia chrysanthemi tolC is involved in resistance to antimicrobial plant chemicals and is essential for phytopathogenesis. J Bacteriol. 2003;185(19):5772–8. doi: 10.1128/JB.185.19.5772-5778.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 598.Reddy JD, Reddy SL, Hopkins DL, et al. TolC is required for pathogenicity of Xylella fastidiosa in Vitis vinifera grapevines. Mol Plant Microbe Interact. 2007;20(4):403–10. doi: 10.1094/MPMI-20-4-0403. [DOI] [PubMed] [Google Scholar]
  • 599.Crosby JA, Kachlany SC. TdeA, a TolC-like protein required for toxin and drug export in Aggregatibacter (Actinobacillus) actinomycetemcomitans. Gene. 2007;388(1–2):83–92. doi: 10.1016/j.gene.2006.10.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 600.Camilli A, Bassler BL. Bacterial small-molecule signaling pathways. Science. 2006;311(5764):1113–6. doi: 10.1126/science.1121357. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 601.Parsek MR, Greenberg EP. Acyl-homoserine lactone quorum sensing in Gram-negative bacteria: a signaling mechanism involved in associations with higher organisms. Proc Natl Acad Sci USA. 2000;97(16):8789–93. doi: 10.1073/pnas.97.16.8789. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 602.Aendekerk S, Diggle SP, Song Z, et al. The MexGHI-OpmD multidrug efflux pump controls growth, antibiotic susceptibility and virulence in Pseudomonas aeruginosa via 4-quinolone-dependent cell-to-cell communication. Microbiology. 2005;151(4):1113–25. doi: 10.1099/mic.0.27631-0. [DOI] [PubMed] [Google Scholar]
  • 603.Chan YY, Bian HS, Tan TM, et al. Control of quorum sensing by a Burkholderia pseudomallei multidrug efflux pump. J Bacteriol. 2007;189(11):4320–4. doi: 10.1128/JB.00003-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 604.Yang S, Lopez CR, Zechiedrich EL. Quorum sensing and multidrug transporters in Escherichia coli. Proc Natl Acad Sci USA. 2006;103(7):2386–91. doi: 10.1073/pnas.0502890102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 605.Maseda H, Sawada I, Saito K, et al. Enhancement of the mexAB-oprM efflux pump expression by a quorum-sensing autoinducer and its cancellation by a regulator, MexT, of the mexEF-oprN efflux pump operon in Pseudomonas aeruginosa. Antimicrob Agents Chemother. 2004;48(4):1320–8. doi: 10.1128/AAC.48.4.1320-1328.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 606.Sawada I, Maseda H, Nakae T, et al. A quorum-sensing autoinducer enhances the mexAB-oprM efflux-pump expression without the MexR-mediated regulation in Pseudomonas aeruginosa. Microbiol Immunol. 2004;48(5):435–9. doi: 10.1111/j.1348-0421.2004.tb03533.x. [DOI] [PubMed] [Google Scholar]
  • 607.Sugimura M, Maseda H, Hanaki H, et al. Macrolide antibiotic-mediated downregulation of MexAB-OprM efflux pump expression in Pseudomonas aeruginosa. Antimicrob Agents Chemother. 2008;52(11):4141–4. doi: 10.1128/AAC.00511-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 608.Dietrich LE, Price-Whelan A, Petersen A, et al. The phenazine pyocyanin is a terminal signalling factor in the quorum sensing network of Pseudomonas aeruginosa. Mol Microbiol. 2006;61(5):1308–21. doi: 10.1111/j.1365-2958.2006.05306.x. [DOI] [PubMed] [Google Scholar]
  • 609.Pumbwe L, Skilbeck CA, Wexler HM. Presence of quorum-sensing systems associated with multidrug resistance and biofilm formation in Bacteroides fragilis. Microb Ecol. 2008;56(3):412–9. doi: 10.1007/s00248-007-9358-3. [DOI] [PubMed] [Google Scholar]
  • 610.Kim J, Kim JG, Kang Y, et al. Quorum sensing and the LysR-type transcriptional activator ToxR regulate toxoflavin biosynthesis and transport in Burkholderia glumae. Mol Microbiol. 2004;54(4):921–34. doi: 10.1111/j.1365-2958.2004.04338.x. [DOI] [PubMed] [Google Scholar]
  • 611.Lau SY, Zgurskaya HI. Cell division defects in Escherichia coli deficient in the multidrug efflux transporter AcrEF-TolC. J Bacteriol. 2005;187(22):7815–25. doi: 10.1128/JB.187.22.7815-7825.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 612.Ramos JL, Martinez-Bueno M, Molina-Henares AJ, et al. The TetR family of transcriptional repressors. Microbiol Mol Biol Rev. 2005;69(2):326–56. doi: 10.1128/MMBR.69.2.326-356.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 613.Gu R, Su CC, Shi F, et al. Crystal structure of the transcriptional regulator CmeR from Campylobacter jejuni. J Mol Biol. 2007;372(3):583–93. doi: 10.1016/j.jmb.2007.06.072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 614.Eguchi Y, Oshima T, Mori H, et al. Transcriptional regulation of drug efflux genes by EvgAS, a two-component system in Escherichia coli. Microbiology. 2003;149(Pt 10):2819–28. doi: 10.1099/mic.0.26460-0. [DOI] [PubMed] [Google Scholar]
  • 615.Nishino K, Honda T, Yamaguchi A. Genome-wide analyses of Escherichia coli gene expression responsive to the BaeSR two-component regulatory system. J Bacteriol. 2005;187(5):1763–72. doi: 10.1128/JB.187.5.1763-1772.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 616.Hirakawa H, Takumi-Kobayashi A, Theisen U, et al. AcrS/EnvR represses expression of the acrAB multidrug efflux genes in Escherichia coli. J Bacteriol. 2008;190(18):6276–9. doi: 10.1128/JB.00190-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 617.Nishino K, Senda Y, Yamaguchi A. The AraC-family regulator GadX enhances multidrug resistance in Escherichia coli by activating expression of mdtEF multidrug efflux genes. J Infect Chemother. 2008;14(1):23–9. doi: 10.1007/s10156-007-0575-y. [DOI] [PubMed] [Google Scholar]
  • 618.Nishino K, Senda Y, Hayashi-Nishino M, et al. Role of the AraC-XylS family regulator YdeO in multi-drug resistance of Escherichia coli. J Antibiot (Tokyo) 2009;62(5):251–7. doi: 10.1038/ja.2009.23. [DOI] [PubMed] [Google Scholar]
  • 619.Nishino K, Senda Y, Yamaguchi A. CRP regulator modulates multidrug resistance of Escherichia coli by repressing the mdtEF multidrug efflux genes. J Antibiot (Tokyo) 2008;61(3):120–7. doi: 10.1038/ja.2008.120. [DOI] [PubMed] [Google Scholar]
  • 620.Nishino K, Yamaguchi A. Role of histone-like protein H-NS in multidrug resistance of Escherichia coli. J Bacteriol. 2004;186(5):1423–9. doi: 10.1128/JB.186.5.1423-1429.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 621.Boutoille D, Corvec S, Caroff N, et al. Detection of an IS21 insertion sequence in the mexR gene of Pseudomonas aeruginosa increasing β-lactam resistance. FEMS Microbiol Lett. 2004;230(1):143–6. doi: 10.1016/S0378-1097(03)00882-6. [DOI] [PubMed] [Google Scholar]
  • 622.Chen H, Hu J, Chen PR, et al. The Pseudomonas aeruginosa multidrug efflux regulator MexR uses an oxidation-sensing mechanism. Proc Natl Acad Sci USA. 2008;105(36):13586–91. doi: 10.1073/pnas.0803391105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 623.Daigle DM, Cao L, Fraud S, et al. Protein modulator of multidrug efflux gene expression in Pseudomonas aeruginosa. J Bacteriol. 2007;189(15):5441–51. doi: 10.1128/JB.00543-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 624.Cao L, Srikumar R, Poole K. MexAB-OprM hyperexpression in NalC-type multidrug-resistant Pseudomonas aeruginosa: identification and characterization of the nalC gene encoding a repressor of PA3720-PA3719. Mol Microbiol. 2004;53(5):1423–36. doi: 10.1111/j.1365-2958.2004.04210.x. [DOI] [PubMed] [Google Scholar]
  • 625.Wilke MS, Heller M, Creagh AL, et al. The crystal structure of MexR from Pseudomonas aeruginosa in complex with its antirepressor ArmR. Proc Natl Acad Sci USA. 2008;105(39):14832–7. doi: 10.1073/pnas.0805489105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 626.Sobel ML, Hocquet D, Cao L, et al. Mutations in PA3574 (nalD) lead to increased MexAB-OprM expression and multidrug resistance in laboratory and clinical isolates of Pseudomonas aeruginosa. Antimicrob Agents Chemother. 2005;49(5):1782–6. doi: 10.1128/AAC.49.5.1782-1786.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 627.Morita Y, Cao L, Gould G, et al. nalD encodes a second repressor of the mexAB-oprM multidrug efflux operon of Pseudomonas aeruginosa. J Bacteriol. 2006;188(24):8649–54. doi: 10.1128/JB.01342-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 628.Li X-Z, Barre N, Poole K. Influence of the MexA-MexB-OprM multidrug efflux system on expression of the MexC-MexD-OprJ and MexE-MexF-OprN multidrug efflux systems in Pseudomonas aeruginosa. J Antimicrob Chemother. 2000;46(6):885–93. doi: 10.1093/jac/46.6.885. [DOI] [PubMed] [Google Scholar]
  • 629.Sobel ML, Neshat S, Poole K. Mutations in PA2491 (mexS) promote MexT-dependent mexEF-oprN expression and multidrug resistance in a clinical strain of Pseudomonas aeruginosa. J Bacteriol. 2005;187(4):1246–53. doi: 10.1128/JB.187.4.1246-1253.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 630.Yoo J, Byeon J, Yoo J, et al. Role of PA2491 Gene in multi-drug resistant Pseudomonas aeruginosa. Abstracts of 48th ICAAC/IDSA 46th Annual Meeting; Washington DC. 2008. pp. C1–1055. [Google Scholar]
  • 631.Morita Y, Murata T, Mima T, et al. Induction of mexCD-oprJ operon for a multidrug efflux pump by disinfectants in wild-type Pseudomonas aeruginosa PAO1. J Antimicrob Chemother. 2003;51(4):991–4. doi: 10.1093/jac/dkg173. [DOI] [PubMed] [Google Scholar]
  • 632.Mandsberg LF, Ciofu O, Kirkby N, et al. Antibiotic resistance in Pseudomonas aeruginosa strains with increased mutation frequency due to inactivation of the DNA oxidative repair system. Antimicrob Agents Chemother. 2009;53(6):2483–91. doi: 10.1128/AAC.00428-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 633.Matsuo Y, Eda S, Gotoh N, et al. MexZ-mediated regulation of mexXY multidrug efflux pump expression in Pseudomonas aeruginosa by binding on the mexZ-mexX intergenic DNA. FEMS Microbiol Lett. 2004;238(1):23–8. doi: 10.1016/j.femsle.2004.07.010. [DOI] [PubMed] [Google Scholar]
  • 634.Chuanchuen R, Gaynor JB, Karkhoff-Schweizer R, et al. Molecular characterization of MexL, the transcriptional repressor of the mexJK multidrug efflux operon in Pseudomonas aeruginosa. Antimicrob Agents Chemother. 2005;49(5):1844–51. doi: 10.1128/AAC.49.5.1844-1851.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 635.Rosenthal RS, Rodwell VW. Purification and characterization of the heteromeric transcriptional activator MvaT of the Pseudomonas mevalonii mvaAB operon. Protein Sci. 1998;7(1):178–84. doi: 10.1002/pro.5560070118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 636.Diggle SP, Winzer K, Lazdunski A, et al. Advancing the quorum in Pseudomonas aeruginosa: MvaT and the regulation of N-acylhomoserine lactone production and virulence gene expression. J Bacteriol. 2002;184(10):2576–86. doi: 10.1128/JB.184.10.2576-2586.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 637.Tendeng C, Soutourina OA, Danchin A, et al. MvaT proteins in Pseudomonas spp.: a novel class of H-NS-like proteins. Microbiology. 2003;149(Pt 11):3047–50. doi: 10.1099/mic.0.C0125-0. [DOI] [PubMed] [Google Scholar]
  • 638.Vallet-Gely I, Donovan KE, Fang R, et al. Repression of phase-variable cup gene expression by H-NS-like proteins in Pseudomonas aeruginosa. Proc Natl Acad Sci USA. 2005;102(31):11082–7. doi: 10.1073/pnas.0502663102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 639.Westfall LW, Carty NL, Layland N, et al. mvaT mutation modifies the expression of the Pseudomonas aeruginosa multidrug efflux operon mexEF-oprN. FEMS Microbiol Lett. 2006;255(2):247–54. doi: 10.1111/j.1574-6968.2005.00075.x. [DOI] [PubMed] [Google Scholar]
  • 640.Teran W, Felipe A, Fillet S, et al. Complexity in efflux pump control: cross-regulation by the paralogues TtgV and TtgT. Mol Microbiol. 2007;66(6):1416–28. doi: 10.1111/j.1365-2958.2007.06004.x. [DOI] [PubMed] [Google Scholar]
  • 641.Fillet S, Velez M, Lu D, et al. TtgV represses two different promoters by recognizing different sequences. J Bacteriol. 2009;191(6):1901–9. doi: 10.1128/JB.01504-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 642.Lin J, Akiba M, Sahin O, et al. CmeR functions as a transcriptional repressor for the multidrug efflux pump CmeABC in Campylobacter jejuni. Antimicrob Agents Chemother. 2005;49(3):1067–75. doi: 10.1128/AAC.49.3.1067-1075.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 643.Cagliero C, Maurel MC, Cloeckaert A, et al. Regulation of the expression of the CmeABC efflux pump in Campylobacter jejuni: identification of a point mutation abolishing the binding of the CmeR repressor in an in vitro-selected multidrug-resistant mutant. FEMS Microbiol Lett. 2007;267(1):89–94. doi: 10.1111/j.1574-6968.2006.00558.x. [DOI] [PubMed] [Google Scholar]
  • 644.O’Regan E, Quinn T, Pages JM, et al. Multiple regulatory pathways associated with high-level ciprofloxacin and multi-drug resistance in Salmonella enterica serovar Enteritidis -involvement of ramA and other global regulators. Antimicrob Agents Chemother. 2009;53(3):1080–7. doi: 10.1128/AAC.01005-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 645.Chiu CH, Tang P, Chu C, et al. The genome sequence of Salmonella enterica serovar Choleraesuis, a highly invasive and resistant zoonotic pathogen. Nucleic Acids Res. 2005;33(5):1690–8. doi: 10.1093/nar/gki297. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 646.Eaves DJ, Ricci V, Piddock LJ. Expression of acrB, acrF, acrD, marA, and soxS in Salmonella enterica serovar Typhimurium: role in multiple antibiotic resistance. Antimicrob Agents Chemother. 2004;48(4):1145–50. doi: 10.1128/AAC.48.4.1145-1150.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 647.Tibbetts RJ, Lin TL, Wu CC. Insertional mutation of marA vitiates inducible multiple antimicrobial resistance in Salmonella enterica subsp. enterica serovar Choleraesuis. Vet Microbiol. 2005;109(3–4):267–74. doi: 10.1016/j.vetmic.2005.05.016. [DOI] [PubMed] [Google Scholar]
  • 648.Yassien MA, Ewis HE, Lu CD, et al. Molecular cloning and characterization of the Salmonella enterica serovar Paratyphi B rma gene, which confers multiple drug resistance in Escherichia coli. Antimicrob Agents Chemother. 2002;46(2):360–6. doi: 10.1128/AAC.46.2.360-366.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 649.van der Straaten T, Janssen R, Mevius DJ, et al. Salmonella gene rma (ramA) and multiple-drug-resistant Salmonella enterica serovar Tphimurium. Antimicrob Agents Chemother. 2004;48(6):2292–4. doi: 10.1128/AAC.48.6.2292-2294.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 650.Feuerriegel S, Heisig P. Role of global regulator Rma for multidrug efflux-mediated fluoroquinolone resistance in Salmonella. Microb Drug Resist. 2008;14(4):259–63. doi: 10.1089/mdr.2008.0846. [DOI] [PubMed] [Google Scholar]
  • 651.Zheng J, Cui S, Meng J. Effect of transcriptional activators RamA and SoxS on expression of multidrug efflux pumps AcrAB and AcrEF in fluoroquinolone-resistant Salmonella Typhimurium. J Antimicrob Chemother. 2009;63(1):95–102. doi: 10.1093/jac/dkn448. [DOI] [PubMed] [Google Scholar]
  • 652.Karatzas KAG, Webber MA, Jorgensen F, et al. Prolonged treatment of Salmonella enterica serovar Typhimurium with commercial disinfectants selects for multiple antibiotic resistance, increased efflux and reduced invasiveness. J Antimicrob Chemother. 2007;60(5):947–55. doi: 10.1093/jac/dkm314. [DOI] [PubMed] [Google Scholar]
  • 653.Rouquette C, Harmon JB, Shafer WM. Induction of the mtrCDE-encoded efflux pump system of Neisseria gonorrhoeae requires MtrA, an AraC-like protein. Mol Microbiol. 1999;33(3):651–8. doi: 10.1046/j.1365-2958.1999.01517.x. [DOI] [PubMed] [Google Scholar]
  • 654.Hoffmann KM, Williams D, Shafer WM, et al. Characterization of the multiple transferable resistance repressor, MtrR, from Neisseria gonorrhoeae. J Bacteriol. 2005;187(14):5008–12. doi: 10.1128/JB.187.14.5008-5012.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 655.Lee EH, Rouquette-Loughlin C, Folster JP, et al. FarR regulates the farAB-encoded efflux pump of Neisseria gonorrhoeae via an MtrR regulatory mechanism. J Bacteriol. 2003;185(24):7145–52. doi: 10.1128/JB.185.24.7145-7152.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 656.Rouquette-Loughlin CE, Balthazar JT, Hill SA, et al. Modulation of the mtrCDE-encoded efflux pump gene complex of Neisseria meningitidis due to a Correia element insertion sequence. Mol Microbiol. 2004;54(3):731–41. doi: 10.1111/j.1365-2958.2004.04299.x. [DOI] [PubMed] [Google Scholar]
  • 657.Veal WL, Shafer WM. Identification of a cell envelope protein (MtrF) involved in hydrophobic antimicrobial resistance in Neisseria gonorrhoeae. J Antimicrob Chemother. 2003;51(1):27–37. doi: 10.1093/jac/dkg031. [DOI] [PubMed] [Google Scholar]
  • 658.Folster JP, Shafer WM. Regulation of mtrF expression in Neisseria gonorrhoeae and its role in high-level antimicrobial resistance. J Bacteriol. 2005;187(11):3713–20. doi: 10.1128/JB.187.11.3713-3720.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 659.Luong TT, Newell SW, Lee CY. Mgr, a novel global regulator in Staphylococcus aureus. J Bacteriol. 2003;185(13):3703–10. doi: 10.1128/JB.185.13.3703-3710.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 660.Truong-Bolduc QC, Zhang X, Hooper DC. Characterization of NorR protein, a multifunctional regulator of norA expression in Staphylococcus aureus. J Bacteriol. 2003;185(10):3127–38. doi: 10.1128/JB.185.10.3127-3138.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 661.Ingavale SS, Van Wamel W, Cheung AL. Characterization of RAT, an autolysis regulator in Staphylococcus aureus. Mol Microbiol. 2003;48(6):1451–66. doi: 10.1046/j.1365-2958.2003.03503.x. [DOI] [PubMed] [Google Scholar]
  • 662.Kaatz GW, Thyagarajan RV, Seo SM. Effect of promoter region mutations and mgrA overexpression on transcription of norA, which encodes a Staphylococcus aureus multidrug efflux transporter. Antimicrob Agents Chemother. 2005;49(1):161–9. doi: 10.1128/AAC.49.1.161-169.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 663.Trotonda MP, Tamber S, Memmi G, et al. MgrA represses biofilm formation in Staphylococcus aureus. Infect Immun. 2008;76(12):5645–54. doi: 10.1128/IAI.00735-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 664.Fournier B, Truong-Bolduc QC, Zhang X, et al. A mutation in the 5′ untranslated region increases stability of norA mRNA, encoding a multidrug resistance transporter of Staphylococcus aureus. J Bacteriol. 2001;183(7):2367–71. doi: 10.1128/JB.183.7.2367-2371.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 665.Cheung AL, Zhang G. Global regulation of virulence determinants in Staphylococcus aureus by the SarA protein family. Front Biosci. 2002;7:d1825–42. doi: 10.2741/A882. [DOI] [PubMed] [Google Scholar]
  • 666.Kaatz GW, DeMarco CE, Seo SM. MepR, a repressor of the Staphylococcus aureus MATE family multidrug efflux pump MepA, is a substrate-responsive regulatory protein. Antimicrob Agents Chemother. 2006;50(4):1276–81. doi: 10.1128/AAC.50.4.1276-1281.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 667.Kumaraswami M, Schuman JT, Seo SM, et al. Structural and biochemical characterization of MepR, a multidrug binding transcription regulator of the Staphylococcus aureus multidrug efflux pump MepA. Nucleic Acids Res. 2009;37(4):1211–24. doi: 10.1093/nar/gkn1046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 668.Heldwein EE, Brennan RG. Crystal structure of the transcription activator BmrR bound to DNA and a drug. Nature. 2001;409(6818):378–82. doi: 10.1038/35053138. [DOI] [PubMed] [Google Scholar]
  • 669.Schumacher MA, Miller MC, Brennan RG. Structural mechanism of the simultaneous binding of two drugs to a multidrug-binding protein. EMBO J. 2004;23(15):2923–30. doi: 10.1038/sj.emboj.7600288. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 670.Murray DS, Schumacher MA, Brennan RG. Crystal structures of QacR-diamidine complexes reveal additional multidrug-binding modes and a novel mechanism of drug charge neutralization. J Biol Chem. 2004;279(14):14365–71. doi: 10.1074/jbc.M313870200. [DOI] [PubMed] [Google Scholar]
  • 671.Muller JF, Stevens AM, Craig J, et al. Transcriptome analysis reveals that multidrug efflux genes are upregulated to protect Pseudomonas aeruginosa from pentachlorophenol stress. Appl Environ Microbiol. 2007;73(14):4550–8. doi: 10.1128/AEM.00169-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 672.Perron K, Caille O, Rossier C, et al. CzcR-CzcS, a two-component system involved in heavy metal and carbapenem resistance in Pseudomonas aeruginosa. J Biol Chem. 2004;279(10):8761–8. doi: 10.1074/jbc.M312080200. [DOI] [PubMed] [Google Scholar]
  • 673.Pumbwe L, Skilbeck CA, Wexler HM. Induction of multiple antibiotic resistance in Bacteroides fragilis by benzene and benzene-derived active compounds of commonly used analgesics, antiseptics and cleaning agents. J Antimicrob Chemother. 2007;60(6):1288–97. doi: 10.1093/jac/dkm363. [DOI] [PubMed] [Google Scholar]
  • 674.Pumbwe L, Skilbeck CA, Wexler HM. Impact of anatomic site on growth, efflux-pump expression, cell structure, and stress responsiveness of Bacteroides fragilis. Curr Microbiol. 2007;55(4):362–5. doi: 10.1007/s00284-007-0278-8. [DOI] [PubMed] [Google Scholar]
  • 675.Domain F, Bina XR, Levy SB. Transketolase A, an enzyme in central metabolism, derepresses the marRAB multiple antibiotic resistance operon of Escherichia coli by interaction with MarR. Mol Microbiol. 2007;66(2):383–94. doi: 10.1111/j.1365-2958.2007.05928.x. [DOI] [PubMed] [Google Scholar]
  • 676.Dowd SE, Killinger-Mann K, Blanton J, et al. Positive adaptive state: microarray evaluation of gene expression in Salmonella enterica Typhimurium exposed to nalidixic acid. Foodborne Pathog Dis. 2007;4(2):187–200. doi: 10.1089/fpd.2006.0075. [DOI] [PubMed] [Google Scholar]
  • 677.Coldham NG, Randall LP, Piddock LJ, et al. Effect of fluoroquinolone exposure on the proteome of Salmonella enterica serovar Typhimurium. J Antimicrob Chemother. 2006;58(6):1145–53. doi: 10.1093/jac/dkl413. [DOI] [PubMed] [Google Scholar]
  • 678.Coban AY, Durupinar B. The effect of nitric oxide combined with fluoroquinolones against Salmonella enterica serovar Typhimurium in vitro. Mem Inst Oswaldo Cruz. 2003;98(3):419–23. doi: 10.1590/s0074-02762003000300023. [DOI] [PubMed] [Google Scholar]
  • 679.Abouzeed YM, Baucheron S, Cloeckaert A. ramR mutations involved in efflux-mediated multidrug resistance in Salmonella enterica serovar Typhimurium. Antimicrob Agents Chemother. 2008;52(7):2428–34. doi: 10.1128/AAC.00084-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 680.Riordan JT, O’Leary JO, Gustafson JE. Contributions of sigB and sarA to distinct multiple antimicrobial resistance mechanisms of Staphylococcus aureus. Int J Antimicrob Agents. 2006;28(1):54–61. doi: 10.1016/j.ijantimicag.2006.01.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 681.Riordan JT, Muthaiyan A, Van Voorhies W, et al. Response of Staphylococcus aureus to salicylate challenge. J Bacteriol. 2007;189(1):220–7. doi: 10.1128/JB.01149-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 682.Evans K, Poole K. The MexA-MexB-OprM multidrug efflux system of Pseudomonas aeruginosa is growth-phase regulated. FEMS Microbiol Lett. 1999;173(1):35–9. doi: 10.1111/j.1574-6968.1999.tb13481.x. [DOI] [PubMed] [Google Scholar]
  • 683.Alonso A, Martinez JL. Cloning and characterization of SmeDEF, a novel multidrug efflux pump from Stenotrophomonas maltophilia. Antimicrob Agents Chemother. 2000;44(11):3079–86. doi: 10.1128/aac.44.11.3079-3086.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 684.Stoitsova SO, Braun Y, Ullrich MS, et al. Characterization of the RND-type multidrug efflux pump MexAB-OprM of the plant pathogen Pseudomonas syringae. Appl Environ Microbiol. 2008;74(11):3387–93. doi: 10.1128/AEM.02866-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 685.Kobayashi A, Hirakawa H, Hirata T, et al. Growth phase-dependent expression of drug exporters in Escherichia coli and its contribution to drug tolerance. J Bacteriol. 2006;188(16):5693–703. doi: 10.1128/JB.00217-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 686.Rand JD, Danby SG, Greenway DL, et al. Increased expression of the multidrug efflux genes acrAB occurs during slow growth of Escherichia coli. FEMS Microbiol Lett. 2002;207(1):91–5. doi: 10.1111/j.1574-6968.2002.tb11034.x. [DOI] [PubMed] [Google Scholar]
  • 687.Miller HI. Are we being outdone by bacteria? Novel antibioitcs and more cautious use of drugs needed to quelch drug-resistant bugs. Genetic Engineering News. 2006;26(10):6–8. [Google Scholar]
  • 688.Wang J, Soisson SM, Young K, et al. Platensimycin is a selective FabF inhibitor with potent antibiotic properties. Nature. 2006;441(7091):358–61. doi: 10.1038/nature04784. [DOI] [PubMed] [Google Scholar]
  • 689.Kaatz GW. Bacterial efflux pump inhibition. Curr Opin Investig Drugs. 2005;6(2):191–8. [PubMed] [Google Scholar]
  • 690.Marquez B. Bacterial efflux systems and efflux pumps inhibitors. Biochimie. 2005;87(12):1137–47. doi: 10.1016/j.biochi.2005.04.012. [DOI] [PubMed] [Google Scholar]
  • 691.Pages JM, Masi M, Barbe J. Inhibitors of efflux pumps in Gram-negative bacteria. Trends Mol Med. 2005;11(8):382–9. doi: 10.1016/j.molmed.2005.06.006. [DOI] [PubMed] [Google Scholar]
  • 692.Lynch AS. Efflux systems in bacterial pathogens: an opportunity for therapeutic intervention? An industry view. Biochem Pharmacol. 2006;71(7):949–56. doi: 10.1016/j.bcp.2005.10.021. [DOI] [PubMed] [Google Scholar]
  • 693.Mahamoud A, Chevalier J, Davin-Regli A, et al. Quinoline derivatives as promising inhibitors of antibiotic efflux pump in multidrug resistant Enterobacter aerogenes isolates. Curr Drug Targets. 2006;7(7):843–7. doi: 10.2174/138945006777709557. [DOI] [PubMed] [Google Scholar]
  • 694.Lomovskaya O, Bostian KA. Practical applications and feasibility of efflux pump inhibitors in the clinic- a vision for applied use. Biochem Pharmacol. 2006;71(7):910–8. doi: 10.1016/j.bcp.2005.12.008. [DOI] [PubMed] [Google Scholar]
  • 695.Lomovskaya O, Zgurskaya HI, Totrov M, et al. Waltzing transporters and ‘the dance macabre’ between humans and bacteria. Nat Rev Drug Discov. 2007;6(1):56–65. doi: 10.1038/nrd2200. [DOI] [PubMed] [Google Scholar]
  • 696.Mahamoud A, Chevalier J, Alibert-Franco S, et al. Antibiotic efflux pumps in Gram-negative bacteria: the inhibitor response strategy. J Antimicrob Chemother. 2007;59(6):1223–9. doi: 10.1093/jac/dkl493. [DOI] [PubMed] [Google Scholar]
  • 697.Stavri M, Piddock LJV, Gibbons S. Bacterial efflux pump inhibitors from natural sources. J Antimicrob Chemother. 2007;59(6):1247–60. doi: 10.1093/jac/dkl460. [DOI] [PubMed] [Google Scholar]
  • 698.Gibbons S. Phytochemicals for bacterial resistance--strengths, weaknesses and opportunities. Planta Med. 2008;74(6):594–602. doi: 10.1055/s-2008-1074518. [DOI] [PubMed] [Google Scholar]
  • 699.Martins M, Dastidar SG, Fanning S, et al. Potential role of non-antibiotics (helper compounds) in the treatment of multidrug-resistant Gram-negative infections: mechanisms for their direct and indirect activities. Int J Antimicrob Agents. 2008;31(3):198–208. doi: 10.1016/j.ijantimicag.2007.10.025. [DOI] [PubMed] [Google Scholar]
  • 700.McKeegan KS, Borges-Walmsley MI, Walmsley AR. Structural understanding of efflux-mediated drug resistance: potential routes to efflux inhibition. Curr Opin Pharmacol. 2004;4(5):479–86. doi: 10.1016/j.coph.2004.07.002. [DOI] [PubMed] [Google Scholar]
  • 701.McDevitt CA, Callaghan R. How can we best use structural information on P-glycoprotein to design inhibitors? Pharmacol Ther. 2007;113(2):429–41. doi: 10.1016/j.pharmthera.2006.10.003. [DOI] [PubMed] [Google Scholar]
  • 702.Gibbons S. Plants as a source of bacterial resistance modulators and anti-infective agents. Phytochem Rev. 2005;4(1):63–78. [Google Scholar]
  • 703.Nakayama K, Kawato H, Watanabe J, et al. MexAB-OprM specific efflux pump inhibitors in Pseudomonas aeruginosaPart 3: Optimization of potency in the pyridopyrimidine series through the application of a pharmacophore model. Bioorg Med Chem Lett. 2004;14(2):475–9. doi: 10.1016/j.bmcl.2003.10.060. [DOI] [PubMed] [Google Scholar]
  • 704.Nakayama K, Kuru N, Ohtsuka M, et al. MexAB-OprM specific efflux pump inhibitors in Pseudomonas aeruginosa. Part 4: Addressing the problem of poor stability due to photoisomerization of an acrylic acid moiety. Bioorg Med Chem Lett. 2004;14(10):2493–7. doi: 10.1016/j.bmcl.2004.03.007. [DOI] [PubMed] [Google Scholar]
  • 705.Yoshida K, Nakayama K, Kuru N, et al. MexAB-OprM specific efflux pump inhibitors in Pseudomonas aeruginosa. Part 5: Carbon-substituted analogues at the C-2 position. Bioorg Med Chem. 2006;14(6):1993–2004. doi: 10.1016/j.bmc.2005.10.043. [DOI] [PubMed] [Google Scholar]
  • 706.Yoshida K, Nakayama K, Yokomizo Y, et al. MexAB-OprM specific efflux pump inhibitors in Pseudomonas aeruginosa. Part 6: exploration of aromatic substituents. Bioorg Med Chem. 2006;14(24):8506–18. doi: 10.1016/j.bmc.2006.08.037. [DOI] [PubMed] [Google Scholar]
  • 707.Yoshida K, Nakayama K, Ohtsuka M, et al. MexAB-OprM specific efflux pump inhibitors in Pseudomonas aeruginosa. Part 7: highly soluble and in vivo active quaternary ammonium analogue D13–9001, a potential preclinical candidate. Bioorg Med Chem. 2007;15(22):7087–97. doi: 10.1016/j.bmc.2007.07.039. [DOI] [PubMed] [Google Scholar]
  • 708.Belofsky G, Percivill D, Lewis K, et al. Phenolic metabolites of Dalea versicolor that enhance antibiotic activity against model pathogenic bacteria. J Nat Prod. 2004;67(3):481–4. doi: 10.1021/np030409c. [DOI] [PubMed] [Google Scholar]
  • 709.Marquez B, Neuville L, Moreau NJ, et al. Multidrug resistance reversal agent from Jatropha elliptica. Phytochemistry. 2005;66(15):1804–11. doi: 10.1016/j.phytochem.2005.06.008. [DOI] [PubMed] [Google Scholar]
  • 710.Hohmann J, Redei D, Forgo P, et al. Jatrophane diterpenoids from Euphorbia mongolica as modulators of the multidrug resistance of L5128 mouse lymphoma cells. J Nat Prod. 2003;66(7):976–9. doi: 10.1021/np030036f. [DOI] [PubMed] [Google Scholar]
  • 711.Gracio MA, Gracio AJ, Viveiros M, et al. Since phenothiazines alter antibiotic susceptibility of microorganisms by inhibiting efflux pumps, are these agents useful for evaluating similar pumps in phenothiazine-sensitive parasites? Int J Antimicrob Agents. 2003;22(3):347–51. doi: 10.1016/s0924-8579(03)00204-8. [DOI] [PubMed] [Google Scholar]
  • 712.Kolaczkowski M, Michalak K, Motohashi N. Phenothiazines as potent modulators of yeast multidrug resistance. Int J Antimicrob Agents. 2003;22(3):279–83. doi: 10.1016/s0924-8579(03)00214-0. [DOI] [PubMed] [Google Scholar]
  • 713.Bean DC, Wareham DW. Paradoxical effect of 1-(1-naphthylmethyl)-piperazine on resistance to tetracyclines in multidrug-resistant Acinetobacter baumannii. J Antimicrob Chemother. 2009;63(2):349–52. doi: 10.1093/jac/dkn493. [DOI] [PubMed] [Google Scholar]
  • 714.Khan IA, Mirza ZM, Kumar A, et al. Piperine, a phytochemical potentiator of ciprofloxacin against Staphylococcus aureus. Antimicrob Agents Chemother. 2006;50(2):810–2. doi: 10.1128/AAC.50.2.810-812.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 715.Martinez A, Lin J. Effect of an efflux pump inhibitor on the function of the multidrug efflux pump CmeABC and antimicrobial resistance in Campylobacter. Foodborne Pathog Dis. 2006;3(4):393–402. doi: 10.1089/fpd.2006.3.393. [DOI] [PubMed] [Google Scholar]
  • 716.Kerns RJ, Rybak MJ, Kaatz GW, et al. Piperazinyl-linked fluoroquinolone dimers possessing potent antibacterial activity against drug-resistant strains of Staphylococcus aureus. Bioorg Med Chem Lett. 2003;13(10):1745–9. doi: 10.1016/s0960-894x(03)00208-7. [DOI] [PubMed] [Google Scholar]
  • 717.Ball AR, Casadei G, Samosorn S, et al. Conjugating berberine to a multidrug efflux pump inhibitor creates an effective antimicrobial. ACS Chem Biol. 2006;1(9):594–600. doi: 10.1021/cb600238x. [DOI] [PubMed] [Google Scholar]
  • 718.Tegos GP, Masago K, Aziz F, Higginbotham A, et al. Inhibitors of bacterial multidrug efflux pumps potentiate antimicrobial photoinactivation. Antimicrob Agents Chemother. 2008;52(9):3202–9. doi: 10.1128/AAC.00006-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 719.Bina XR, Philippart JA, Bina JE. Effect of the efflux inhibitors 1-(1-naphthylmethyl)-piperazine and phenyl-arginine-β-naphthylamide on antimicrobial susceptibility and virulence factor production in Vibrio cholerae. J Antimicrob Chemother. 2009;63(1):103–8. doi: 10.1093/jac/dkn466. [DOI] [PubMed] [Google Scholar]
  • 720.Chevalier J, Mulfinger C, Garnotel E, et al. Identification and evolution of drug efflux pump in clinical Enterobacter aerogenes strains isolated in 1995 and 2003. PLoS ONE. 2008;3(9):e3203. doi: 10.1371/journal.pone.0003203. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 721.Klyachko KA, Schuldiner S, Neyfakh AA. Mutations affecting substrate specificity of the Bacillus subtilis multidrug transporter Bmr. J Bacteriol. 1997;179(7):2189–93. doi: 10.1128/jb.179.7.2189-2193.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 722.Gibbons S, Moser E, Kaatz GW. Catechin gallates inhibit multidrug resistance (MDR) in Staphylococcus aureus. Planta Med. 2004;70(12):1240–2. doi: 10.1055/s-2004-835860. [DOI] [PubMed] [Google Scholar]
  • 723.Frempong-Manso E, Raygada JL, Demarco CE, et al. Inability of a reserpine-based screen to identify strains overexpressing efflux pump genes in clinical isolates of Staphylococcus aureus. Int J Antimicrob Agents. 2009;33(4):360–3. doi: 10.1016/j.ijantimicag.2008.10.016. [DOI] [PubMed] [Google Scholar]
  • 724.Zhanel GG, Johanson C, Laing N, et al. Pharmacodynamic activity of telithromycin at simulated clinically achievable free-drug concentrations in serum and epithelial lining fluid against efflux (mefE)-producing macrolide-resistant Streptococcus pneumoniae for which telithromycin MICs vary. Antimicrob Agents Chemother. 2005;49(5):1943–8. doi: 10.1128/AAC.49.5.1943-1948.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 725.Michot JM, Seral C, Van Bambeke F, et al. Influence of efflux transporters on the accumulation and efflux of four quinolones (ciprofloxacin, levofloxacin, garenoxacin, and moxifloxacin) in J774 macrophages. Antimicrob Agents Chemother. 2005;49(6):2429–37. doi: 10.1128/AAC.49.6.2429-2437.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 726.Alvarez AI, Perez M, Prieto JG, et al. Fluoroquinolone efflux mediated by ABC transporters. J Pharm Sci. 2008;97(9):3483–93. doi: 10.1002/jps.21233. [DOI] [PubMed] [Google Scholar]
  • 727.Projan SJ. (Genome) size matters. Antimicrob Agents Chemother. 2007;51(4):1133–4. doi: 10.1128/AAC.01370-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 728.Bronzwaer SL, Cars O, Buchholz U, et al. A European study on the relationship between antimicrobial use and antimicrobial resistance. Emerg Infect Dis. 2002;8(3):278–82. doi: 10.3201/eid0803.010192. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 729.Goossens H. Antibiotic consumption and link to resistance. Clin Microbiol Infect. 2009;15(Suppl 3):12–5. doi: 10.1111/j.1469-0691.2009.02725.x. [DOI] [PubMed] [Google Scholar]
  • 730.Rieg S, Huth A, Kalbacher H, Kern WV. Resistance against antimicrobial peptides is independent of Escherichia coli AcrAB, Pseudomonas aeruginosa MexAB and Staphylococcus aureus NorA efflux pumps. Int J Antimicrob Agents. 2009;33(2):174–6. doi: 10.1016/j.ijantimicag.2008.07.032. [DOI] [PubMed] [Google Scholar]
  • 731.Brissette CA, Lukehart SA. Mechanisms of decreased susceptibility to β-defensins by Treponema denticola. Infect Immun. 2007;75(5):2307–15. doi: 10.1128/IAI.01718-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 732.Amaral L, Martins M, Viveiros M, et al. Promising therapy of XDR-TB/MDR-TB with thioridazine an inhibitor of bacterial efflux pumps. Curr Drug Targets. 2008;9(9):816–9. doi: 10.2174/138945008785747798. [DOI] [PubMed] [Google Scholar]
  • 733.Espinal PA, Marti S, Sanchez-Cespedes J, et al. First detection of adeC component of the efflux pump AdeABC in an Acinetobacter genospecies 13TU. Abstracts of 48th ICAAC/IDSA 46th Annual Meeting; Washington DC. 2008. pp. C1–1049. [Google Scholar]
  • 734.Tokunaga H, Mitsuo K, Ichinose S, et al. Salt-inducible multidrug efflux pump protein in the moderately halophilic bacterium Chromohalobacter sp. Appl Environ Microbiol. 2004;70(8):4424–31. doi: 10.1128/AEM.70.8.4424-4431.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 735.Masi M, Saint N, Molle G, et al. The Enterobacter aerogenes outer membrane efflux proteins TolC and EefC have different channel properties. Biochim Biophys Acta. 2007;1768(10):2559–67. doi: 10.1016/j.bbamem.2007.06.008. [DOI] [PubMed] [Google Scholar]
  • 736.Perez A, Canle D, Latasa C, et al. Cloning, nucleotide sequencing, and analysis of the AcrAB-TolC efflux pump of Enterobacter cloacae and determination of its involvement in antibiotic resistance in a clinical Isolate. Antimicrob Agents Chemother. 2007;51(9):3247–53. doi: 10.1128/AAC.00072-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 737.Cerquetti M, Giufre M, Cardines R, et al. First characterization of heterogeneous resistance to imipenem in invasive nontypeable Haemophilus influenzae isolates. Antimicrob Agents Chemother. 2007;51(9):3155–61. doi: 10.1128/AAC.00335-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 738.Kutschke A, de Jonge BL. Compound efflux in Helicobacter pylori. Antimicrob Agents Chemother. 2005;49(7):3009–10. doi: 10.1128/AAC.49.7.3009-3010.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 739.Mazzariol A, Zuliani J, Cornaglia G, et al. AcrAB efflux system: expression and contribution to fluoroquinolone resistance in Klebsiella spp. Antimicrob Agents Chemother. 2002;46(12):3984–6. doi: 10.1128/AAC.46.12.3984-3986.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 740.Lee EH, Hill SA, Napier R, et al. Integration host factor is required for FarR repression of the farAB-encoded efflux pump of Neisseria gonorrhoeae. Mol Microbiol. 2006;60(6):1381–400. doi: 10.1111/j.1365-2958.2006.05185.x. [DOI] [PubMed] [Google Scholar]
  • 741.Hatfaludi T, Al-Hasani K, Dunstone M, et al. Characterization of TolC efflux pump proteins from Pasteurella multocida. Antimicrob Agents Chemother. 2008;52(11):4166–71. doi: 10.1128/AAC.00245-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 742.Begic S, Worobec EA. Characterization of the Serratia marcescens SdeCDE multidrug efflux pump studied via gene knockout mutagenesis. Can J Microbiol. 2008;54(5):411–6. doi: 10.1139/w08-019. [DOI] [PubMed] [Google Scholar]
  • 743.Matsuo T, Hayashi K, Morita Y, et al. VmeAB, an RND-type multidrug efflux transporter in Vibrio parahaemolyticus. Microbiology. 2007;153(Pt 12):4129–37. doi: 10.1099/mic.0.2007/009597-0. [DOI] [PubMed] [Google Scholar]
  • 744.Ohki R, Murata M. bmr3, a third multidrug transporter gene of Bacillus subtilis. J Bacteriol. 1997;179(4):1423–7. doi: 10.1128/jb.179.4.1423-1427.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 745.Kadlec K, Kehrenberg C, Schwarz S. Efflux-mediated resistance to florfenicol and/or chloramphenicol in Bordetella bronchiseptica: identification of a novel chloramphenicol exporter. J Antimicrob Chemother. 2007;59(2):191–6. doi: 10.1093/jac/dkl498. [DOI] [PubMed] [Google Scholar]
  • 746.Morrison S, Ward A, Hoyle CJ, et al. Cloning, expression, purification and properties of a putative multidrug resistance efflux protein from Helicobacter pylori. Int J Antimicrob Agents. 2003;22(3):242–9. doi: 10.1016/s0924-8579(03)00222-x. [DOI] [PubMed] [Google Scholar]
  • 747.Gil F, Ipinza F, Fuentes J, et al. The ompW (porin) gene mediates methyl viologen (paraquat) efflux in Salmonella enterica serovar Typhimurium. Res Microbiol. 2007;158(6):529–36. doi: 10.1016/j.resmic.2007.05.004. [DOI] [PubMed] [Google Scholar]
  • 748.Yamada Y, Shiota S, Mizushima T, et al. Functional gene cloning and characterization of MdeA, a multidrug efflux pump from Staphylococcus aureus. Biol Pharm Bull. 2006;29(4):801–4. doi: 10.1248/bpb.29.801. [DOI] [PubMed] [Google Scholar]
  • 749.Overton TW, Justino MC, Li Y, et al. Widespread distribution in pathogenic bacteria of di-iron proteins that repair oxidative and nitrosative damage to iron-sulfur centers. J Bacteriol. 2008;190(6):2004–13. doi: 10.1128/JB.01733-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 750.Cai Y, Kong F, Gilbert GL. Three new macrolide efflux (mef) gene variants in Streptococcus agalactiae. J Clin Microbiol. 2007;45(8):2754–5. doi: 10.1128/JCM.00579-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 751.Brown MG, Mitchell EH, Balkwill DL. Tet 42, a novel tetracycline resistance determinant isolated from deep terrestrial subsurface bacteria. Antimicrob Agents Chemother. 2008;52(12):4518–21. doi: 10.1128/AAC.00640-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 752.Escudero JA, San Millan A, Hidalgo L, et al. Identification and characterisation of SmrA, a novel fluoroquinolone efflux pump in Streptococcus suis. C1-1945. Abstracts of 48th ICAAC/IDSA 46th Annual Meeting; Washington DC. 2008. [Google Scholar]
  • 753.Bostock JM, Huang G, Hashimi SM, et al. A DHA14 drug efflux gene from Xanthomonas albilineans confers high-level albicidin antibiotic resistance in Escherichia coli. J Appl Microbiol. 2006;101(1):151–60. doi: 10.1111/j.1365-2672.2006.02899.x. [DOI] [PubMed] [Google Scholar]
  • 754.Burse A, Weingart H, Ullrich MS. NorM, an Erwinia amylovora multidrug efflux pump involved in in vitro competition with other epiphytic bacteria. Appl Environ Microbiol. 2004;70(2):693–703. doi: 10.1128/AEM.70.2.693-703.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 755.Chen J, Morita Y, Huda MN, et al. VmrA, a member of a novel class of Na+-coupled multidrug efflux pumps from Vibrio parahaemolyticus. J Bacteriol. 2002;184(2):572–6. doi: 10.1128/JB.184.2.572-576.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 756.Bernard R, Joseph P, Guiseppi A, et al. YtsCD and YwoA, two independent systems that confer bacitracin resistance to Bacillus subtilis. FEMS Microbiol Lett. 2003;228(1):93–7. doi: 10.1016/S0378-1097(03)00738-9. [DOI] [PubMed] [Google Scholar]
  • 757.Danilchanka O, Mailaender C, Niederweis M. Identification of a novel multidrug efflux pump of Mycobacterium tuberculosis. Antimicrob Agents Chemother. 2008;52(7):2503–11. doi: 10.1128/AAC.00298-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 758.Bourdineaud JP, Nehme B, Tesse S, et al. A bacterial gene homologous to ABC transporters protect Oenococcus oeni from ethanol and other stress factors in wine. Int J Food Microbiol. 2004;92(1):1–14. doi: 10.1016/S0168-1605(03)00162-4. [DOI] [PubMed] [Google Scholar]
  • 759.Achard-Joris M, van den Berg van Saparoea HB, Driessen AJ, et al. Heterologously expressed bacterial and human multidrug resistance proteins confer cadmium resistance to Escherichia coli. Biochemistry. 2005;44(15):5916–22. doi: 10.1021/bi047700r. [DOI] [PubMed] [Google Scholar]
  • 760.Schrader-Fischer G, Berger-Bachi B. The AbcA transporter of Staphylococcus aureus affects cell autolysis. Antimicrob Agents Chemother. 2001;45(2):407–12. doi: 10.1128/AAC.45.2.407-412.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 761.Tavio MM, Vila J, Perilli M, et al. Enhanced active efflux, repression of porin synthesis and development of Mar phenotype by diazepam in two enterobacteria strains. J Med Microbiol. 2004;53(Pt 11):1119–22. doi: 10.1099/jmm.0.45613-0. [DOI] [PubMed] [Google Scholar]
  • 762.Rosenberg EY, Bertenthal D, Nilles ML, et al. Bile salts and fatty acids induce the expression of Escherichia coli AcrAB multidrug efflux pump through their interaction with Rob regulatory protein. Mol Microbiol. 2003;48(6):1609–19. doi: 10.1046/j.1365-2958.2003.03531.x. [DOI] [PubMed] [Google Scholar]
  • 763.Prouty AM, Brodsky IE, Falkow S, et al. Bile-salt-mediated induction of antimicrobial and bile resistance in Salmonella typhimurium. Microbiology. 2004;150(Pt 4):775–83. doi: 10.1099/mic.0.26769-0. [DOI] [PubMed] [Google Scholar]
  • 764.Nikaido E, Yamaguchi A, Nishino K. AcrAB multidrug efflux pump regulation in Salmonella enterica serovar Typhimurium by RamA in response to environmental signals. J Biol Chem. 2008;283(35):24245–53. doi: 10.1074/jbc.M804544200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 765.Langsrud S, Sundheim G, Holck AL. Cross-resistance to antibiotics of Escherichia coli adapted to benzalkonium chloride or exposed to stress-inducers. J Appl Microbiol. 2004;96(1):201–8. doi: 10.1046/j.1365-2672.2003.02140.x. [DOI] [PubMed] [Google Scholar]
  • 766.Hirakawa H, Inazumi Y, Masaki T, et al. Indole induces the expression of multidrug exporter genes in Escherichia coli. Mol Microbiol. 2005;55(4):1113–26. doi: 10.1111/j.1365-2958.2004.04449.x. [DOI] [PubMed] [Google Scholar]
  • 767.Ravirala RS, Barabote RD, Wheeler DM, et al. Efflux pump gene expression in Erwinia chrysanthemi is induced by exposure to phenolic acids. Mol Plant Microbe Interact. 2007;20(3):313–20. doi: 10.1094/MPMI-20-3-0313. [DOI] [PubMed] [Google Scholar]
  • 768.Denkin S, Byrne S, Jie C, et al. Gene expression profiling analysis of Mycobacterium tuberculosis genes in response to salicylate. Arch Microbiol. 2005;184(3):152–7. doi: 10.1007/s00203-005-0037-9. [DOI] [PubMed] [Google Scholar]
  • 769.Escribano I, Rodriguez JC, Pertegas V, Cebrian L, et al. Relation between induction of the mar operon and cyclohexane tolerance and reduction in fluoroquinolone susceptibility in Salmonella spp. J Infect Chemother. 2006;12(4):177–80. doi: 10.1007/s10156-006-0456-9. [DOI] [PubMed] [Google Scholar]
  • 770.Hannula M, Hanninen ML. Effect of putative efflux pump inhibitors and inducers on the antimicrobial susceptibility of Campylobacter jejuni and Campylobacter coli. J Med Microbiol. 2008;57(Pt 7):851–5. doi: 10.1099/jmm.0.47823-0. [DOI] [PubMed] [Google Scholar]
  • 771.Hood MI, Skaar EP. Sodium chloride exposure induces expression of antibiotic resistance in Acinetobacter baumannii. C1-3726. Abstracts of 48th ICAAC/IDSA 46th Annual Meeting; Washington DC. 2008. [Google Scholar]
  • 772.Bohnert JA, Kern WV. Selected arylpiperazines are capable of reversing multidrug resistance in Escherichia coli overexpressing RND efflux pumps. Antimicrob Agents Chemother. 2005;49(2):849–52. doi: 10.1128/AAC.49.2.849-852.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 773.Schumacher A, Steinke P, Bohnert JA, et al. Effect of 1-(1-naphthylmethyl)-piperazine, a novel putative efflux pump inhibitor, on antimicrobial drug susceptibility in clinical isolates of Enterobacteriaceae other than Escherichia coli. J Antimicrob Chemother. 2006;57(2):344–8. doi: 10.1093/jac/dki446. [DOI] [PubMed] [Google Scholar]
  • 774.Pannek S, Higgins PG, Steinke P, et al. Multidrug efflux inhibition in Acinetobacter baumannii: comparison between 1-(1-naphthylmethyl)-piperazine and phenyl-arginine-β-naphthylamide. J Antimicrob Chemother. 2006;57(5):970–4. doi: 10.1093/jac/dkl081. [DOI] [PubMed] [Google Scholar]
  • 775.Musumeci R, Speciale A, Costanzo R, et al. Berberis aetnensis C. Presl. extracts: antimicrobial properties and interaction with ciprofloxacin. Int J Antimicrob Agents. 2003;22(1):48–53. doi: 10.1016/s0924-8579(03)00085-2. [DOI] [PubMed] [Google Scholar]
  • 776.Mallea M, Mahamoud A, Chevalier J, et al. Alkylaminoquinolines inhibit the bacterial antibiotic efflux pump in multidrug-resistant clinical isolates. Biochem J. 2003;376(Pt 3):801–5. doi: 10.1042/BJ20030963. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 777.Pages JM, Dimarcq JL, Quenin S, et al. Thanatin activity on multidrug resistant clinical isolates of Enterobacter aerogenes and Klebsiella pneumoniae. Int J Antimicrob Agents. 2003;22(3):265–9. doi: 10.1016/s0924-8579(03)00201-2. [DOI] [PubMed] [Google Scholar]
  • 778.German N, Kaatz GW, Kerns RJ. Synthesis and evaluation of PSSRI-based inhibitors of Staphylococcus aureus multidrug efflux pumps. Bioorg Med Chem Lett. 2008;18(4):1368–73. doi: 10.1016/j.bmcl.2008.01.014. [DOI] [PubMed] [Google Scholar]
  • 779.Fujita M, Shiota S, Kuroda T, et al. Remarkable synergies between baicalein and tetracycline, and baicalein and β-lactams against methicillin-resistant Staphylococcus aureus. Microbiol Immunol. 2005;49(4):391–6. doi: 10.1111/j.1348-0421.2005.tb03732.x. [DOI] [PubMed] [Google Scholar]
  • 780.Hamilton-Miller JM, Shah S. Activity of the tea component epicatechin gallate and analogues against methicillin-resistant Staphylococcus aureus. J Antimicrob Chemother. 2000;46(5):852–3. doi: 10.1093/jac/46.5.852. [DOI] [PubMed] [Google Scholar]
  • 781.Roccaro SA, Blanco AR, Giuliano F, et al. Epigallocatechin-gallate enhances the activity of tetracycline in staphylococci by inhibiting its efflux from bacterial cells. Antimicrob Agents Chemother. 2004;48(6):1968–73. doi: 10.1128/AAC.48.6.1968-1973.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 782.Gibbons S, Oluwatuyi M, Veitch NC, et al. Bacterial resistance modifying agents from Lycopus europaeus. Phytochemistry. 2003;62(1):83–7. doi: 10.1016/s0031-9422(02)00446-6. [DOI] [PubMed] [Google Scholar]
  • 783.Oluwatuyi M, Kaatz GW, Gibbons S. Antibacterial and resistance modifying activity of Rosmarinus officinalis. Phytochemistry. 2004;65(24):3249–54. doi: 10.1016/j.phytochem.2004.10.009. [DOI] [PubMed] [Google Scholar]
  • 784.Smith EC, Williamson EM, Wareham N, et al. Antibacterials and modulators of bacterial resistance from the immature cones of Chamaecyparis lawsoniana. Phytochemistry. 2007;68(2):210–7. doi: 10.1016/j.phytochem.2006.10.001. [DOI] [PubMed] [Google Scholar]
  • 785.Dickson RA, Houghton PJ, Hylands PJ, et al. Antimicrobial, resistance-modifying effects, antioxidant and free radical scavenging activities of Mezoneuron benthamianum Baill., Securinega virosa Roxb. & Wlld. and Microglossa pyrifolia Lam. Phytother Res. 2006;20(1):41–5. doi: 10.1002/ptr.1799. [DOI] [PubMed] [Google Scholar]
  • 786.Michalet S, Cartier G, David B, et al. N-caffeoylphenalkylamide derivatives as bacterial efflux pump inhibitors. Bioorg Med Chem Lett. 2007;17(6):1755–8. doi: 10.1016/j.bmcl.2006.12.059. [DOI] [PubMed] [Google Scholar]
  • 787.Braga LC, Leite AA, Xavier KG, et al. Synergic interaction between pomegranate extract and antibiotics against Staphylococcus aureus. Can J Microbiol. 2005;51(7):541–7. doi: 10.1139/w05-022. [DOI] [PubMed] [Google Scholar]
  • 788.Stermitz FR, Beeson TD, Mueller PJ, et al. Staphylococcus aureus MDR efflux pump inhibitors from a Berberis and a Mahonia (sensu strictu) species. Biochem Syst Ecol. 2001;29(8):793–8. doi: 10.1016/s0305-1978(01)00025-4. [DOI] [PubMed] [Google Scholar]
  • 789.German N, Wei P, Kaatz GW, et al. Synthesis and evaluation of fluoroquinolone derivatives as substrate-based inhibitors of bacterial efflux pumps. Eur J Med Chem. 2008;43(11):2453–63. doi: 10.1016/j.ejmech.2008.01.042. [DOI] [PubMed] [Google Scholar]
  • 790.Gibbons S, Oluwatuyi M, Kaatz GW. A novel inhibitor of multidrug efflux pumps in Staphylococcus aureus. J Antimicrob Chemother. 2003;51(1):13–7. doi: 10.1093/jac/dkg044. [DOI] [PubMed] [Google Scholar]
  • 791.Abulrob AN, Suller MT, Gumbleton M, et al. Identification and biological evaluation of grapefruit oil components as potential novel efflux pump modulators in methicillin-resistant Staphylococcus aureus bacterial strains. Phytochemistry. 2004;65(22):3021–7. doi: 10.1016/j.phytochem.2004.08.044. [DOI] [PubMed] [Google Scholar]
  • 792.Samosorn S, Bremner JB, Ball A, et al. Synthesis of functionalized 2-aryl-5-nitro-1H-indoles and their activity as bacterial NorA efflux pump inhibitors. Bioorg Med Chem. 2006;14(3):857–65. doi: 10.1016/j.bmc.2005.09.019. [DOI] [PubMed] [Google Scholar]
  • 793.Falcao-Silva VS, Silva DA, Souza MD, et al. Modulation of drug resistance in Staphylococcus aureus by a kaempferol glycoside from Herissantia tiubae (Malvaceae) Phytother Res. 2009 doi: 10.1002/ptr.2695. e-published February 17, 2009. [DOI] [PubMed] [Google Scholar]
  • 794.Stermitz FR, Scriven LN, Tegos G, et al. Two flavonols from Artemisa annua which potentiate the activity of berberine and norfloxacin against a resistant strain of Staphylococcus aureus. Planta Med. 2002;68(12):1140–1. doi: 10.1055/s-2002-36347. [DOI] [PubMed] [Google Scholar]
  • 795.Morel C, Stermitz FR, Tegos G, et al. Isoflavones as potentiators of antibacterial activity. J Agric Food Chem. 2003;51(19):5677–9. doi: 10.1021/jf0302714. [DOI] [PubMed] [Google Scholar]
  • 796.Belofsky G, Carreno R, Lewis K, et al. Metabolites of the “smoke tree”, Dalea spinosa, potentiate antibiotic activity against multidrug-resistant Staphylococcus aureus. J Nat Prod. 2006;69(2):261–4. doi: 10.1021/np058057s. [DOI] [PubMed] [Google Scholar]
  • 797.Cherigo L, Pereda-Miranda R, Fragoso-Serrano M, et al. Inhibitors of bacterial multidrug efflux pumps from the resin glycosides of Ipomoea murucoides. J Nat Prod. 2008;71(6):1037–45. doi: 10.1021/np800148w. [DOI] [PubMed] [Google Scholar]
  • 798.Kaatz GW, Moudgal VV, Seo SM, et al. Phenothiazines and thioxanthenes inhibit multidrug efflux pump activity in Staphylococcus aureus. Antimicrob Agents Chemother. 2003;47(2):719–26. doi: 10.1128/AAC.47.2.719-726.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 799.Kristiansen MM, Leandro C, Ordway D, et al. Phenothiazines alter resistance of methicillin-resistant strains of Staphylococcus aureus (MRSA) to oxacillin in vitro. Int J Antimicrob Agents. 2003;22(3):250–3. doi: 10.1016/s0924-8579(03)00200-0. [DOI] [PubMed] [Google Scholar]
  • 800.Kristiansen MM, Leandro C, Ordway D, et al. Thioridazine reduces resistance of methicillin-resistant Saphylococcus aureus by inhibiting a reserpine-sensitive efflux pump. In Vivo. 2006;20(3):361–6. [PubMed] [Google Scholar]
  • 801.Kristiansen JE, Hendricks O, Delvin T, et al. Reversal of resistance in microorganisms by help of non-antibiotics. J Antimicrob Chemother. 2007;59(6):1271–9. doi: 10.1093/jac/dkm071. [DOI] [PubMed] [Google Scholar]
  • 802.Sangwan PL, Koul JL, Koul S, et al. Piperine analogs as potent Staphylococcus aureus NorA efflux pump inhibitors. Bioorg Med Chem. 2008;16(22):9847–57. doi: 10.1016/j.bmc.2008.09.042. [DOI] [PubMed] [Google Scholar]
  • 803.Stermitz FR, Cashman KK, Halligan KM, et al. Polyacylated neohesperidosides from Geranium caespitosum: bacterial multidrug resistance pump inhibitors. Bioorg Med Chem Lett. 2003;13(11):1915–8. doi: 10.1016/s0960-894x(03)00316-0. [DOI] [PubMed] [Google Scholar]
  • 804.Pereda-Miranda R, Kaatz GW, 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(3):406–9. doi: 10.1021/np050227d. [DOI] [PubMed] [Google Scholar]
  • 805.Vidaillac C, Guillon J, Arpin C, et al. Synthesis of omeprazole analogues and evaluation of these as potential inhibitors of the multidrug efflux pump NorA of Staphylococcus aureus. Antimicrob Agents Chemother. 2007;51(3):831–8. doi: 10.1128/AAC.01306-05. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 806.Smith EC, Kaatz GW, Seo SM, et al. The phenolic diterpene totarol inhibits multidrug efflux pump activity in Staphylococcus aureus. Antimicrob Agents Chemother. 2007;51(12):4480–3. doi: 10.1128/AAC.00216-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 807.Spies FS, da Silva PE, Ribeiro MO, et al. Identification of mutations related to streptomycin resistance in clinical isolates of Mycobacterium tuberculosis and possible involvement of efflux mechanism. Antimicrob Agents Chemother. 2008;52(8):2947–9. doi: 10.1128/AAC.01570-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 808.Rodrigues L, Wagner D, Viveiros M, et al. Thioridazine and chlorpromazine inhibition of ethidium bromide efflux in Mycobacterium avium and Mycobacterium smegmatis. J Antimicrob Chemother. 2008;61(5):1076–82. doi: 10.1093/jac/dkn070. [DOI] [PubMed] [Google Scholar]
  • 809.Lechner D, Gibbons S, Bucar F. Plant phenolic compounds as ethidium bromide efflux inhibitors in Mycobacterium smegmatis. J Antimicrob Chemother. 2008;62(2):345–8. doi: 10.1093/jac/dkn178. [DOI] [PubMed] [Google Scholar]

RESOURCES