commentary antibiotic efflux pumps · the drug efflux pumps in eucaryotic cells ( [7]; drug efflux...

14
COMMENTARY Antibiotic Efflux Pumps Franc ¸oise Van Bambeke,* Elisabetta Balzi‡§ and Paul M. Tulkens* *UNITE ´ DE PHARMACOLOGIE CELLULAIRE ET MOLE ´ CULAIRE,UNIVERSITE ´ CATHOLIQUE DE LOUVAIN,BRUSSELS; AND ‡UNITE ´ DE BIOCHIMIE PHYSIOLOGIQUE,UNIVERSITE ´ CATHOLIQUE DE LOUVAIN,LOUVAIN-LA-NEUVE,BELGIUM ABSTRACT. Active efflux from procaryotic as well as eucaryotic cells strongly modulates the activity of a large number of antibiotics. Effective antibiotic transport has now been observed for many classes of drug efflux pumps. Thus, within the group of primary active transporters, predominant in eucaryotes, six families belonging to the ATP-binding cassette superfamily, and including the P-glycoprotein in the MDR (Multi Drug Resistance) group and the MRP (Multidrug Resistance Protein), have been recognized as being responsible for antibiotic efflux. Within the class of secondary active transporters (antiports, symports, and uniports), ten families of antibiotic efflux pumps have been described, distributed in five superfamilies [SMR (Small Multidrug Resistance), MET (Multidrug Endosomal Transporter), MAR (Multi Antimicrobial Resistance), RND (Resistance Nodulation Division), and MFS (Major Facilitator Superfamily)]. Nowadays antibiotic efflux pumps are believed to contribute significantly to acquired bacterial resistance because of the very broad variety of substrates they recognize, their expression in important pathogens, and their cooperation with other mechanisms of resistance. Their presence also explains high-level intrinsic resistances found in specific organisms. Stable mutations in regulatory genes can produce phenotypes of irreversible multidrug resistance. In eucaryotes, antibiotic efflux pumps modulate the accumulation of antimicrobials in phagocytic cells and play major roles in their transepithelial transport. The existence of antibiotic efflux pumps, and their impact on therapy, must now be taken fully into account for the selection of novel antimicrobials. The design of specific, potent inhibitors appears to be an important goal for the improved control of infectious diseases in the near future. BIOCHEM PHARMACOL 60;4:457– 470, 2000. © 2000 Elsevier Science Inc. KEY WORDS. antibiotic; efflux pump; resistance; transport; chemotherapy; infection Biological membranes most likely appeared very early on during evolution to isolate hydrophilic microdomains from the surrounding medium, allowing catalyzed reactions to occur in an efficient manner. Biomembranes, accordingly, constitute efficient barriers towards hydrophilic molecules, most of which can penetrate cells only by specific inward transport systems or find their entry restricted to the endocytic pathway. A contrario, biomembranes are easily crossed by amphiphilic compounds, since these are able to diffuse through both the hydrophilic and the hydrophobic domains of the bilayer. Therefore, it is not surprising that mechanisms were devised, also very early, to protect cells from the disordered invasion by amphiphilic molecules, many of which are endowed with biological activities leading to potentially harmful effects. A major mechanism in this respect is constituted by active outward transport. Although efflux systems have been known for many years, their importance, both in terms of number and variety of substrates, has become clearly recognized only very re- cently. Drugs are often amphiphilic, whether by selection or by design, ensuring their wide tissue distribution and/or their penetration into membrane-protected compartments. Therefore, it comes as no surprise that many drugs should fall into this category of exogenous compounds for which efflux mechanisms, globally referred to as ‘drug efflux pumps,’ are numerous and fairly active [1]. Thus, more and more membrane-spanning proteins involved in the outward transport of a surprisingly large variety of drugs have been recognized and characterized over the last years in almost all cell types, from procaryotes and archebacteria through fungi and higher eucaryotes. This now raises the question of which drug is not transported. For eucaryotic cells, drug efflux pumps have been viewed by many authors as com- plementing the cytochrome P 450 — or other enzyme-based detoxification systems (e.g. Ref. 2) to achieve efficient protection against ‘chemical invaders’. Both systems, in- deed, show broad specificity and may even work in synergy (see, for example, the concept of a concerted barrier in enterocytes [3, 4]). Drug efflux, indeed, decreases the load on enzyme-mediated detoxification systems, thereby avoid- ing their saturation, while chemical modifications by the enzyme-based systems, which usually increase the am- phiphilicity of drugs, provide drug pumps with better substrates [2, 5]. Moreover, most drug efflux pumps have a broad substrate specificity and, therefore, may deal with a wide range of drugs of completely unrelated pharmacolog- ical classes. The present commentary will focus on antibi- ² Corresponding author: F. Van Bambeke, Pharm.D., UCL 73.70 avenue E. Mounier 73, B-1200 Bruxelles, Belgium. Tel. (32) 2-764-73-78; FAX (32) 2-764-73-73; E-mail: [email protected] § Present affiliation: DG Research, European Commission, Brussels, Belgium. Biochemical Pharmacology, Vol. 60, pp. 457– 470, 2000. ISSN 0006-2952/00/$–see front matter © 2000 Elsevier Science Inc. All rights reserved. PII S0006-2952(00)00291-4

Upload: others

Post on 13-Jul-2021

8 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: COMMENTARY Antibiotic Efflux Pumps · the drug efflux pumps in eucaryotic cells ( [7]; drug efflux transporters are classically energized by ATP). The second-ary active transporters,

COMMENTARY

Antibiotic Efflux PumpsFrancoise Van Bambeke,*† Elisabetta Balzi‡§ and Paul M. Tulkens*

*UNITE DE PHARMACOLOGIE CELLULAIRE ET MOLECULAIRE, UNIVERSITE CATHOLIQUE DE LOUVAIN, BRUSSELS; AND

‡UNITE DE BIOCHIMIE PHYSIOLOGIQUE, UNIVERSITE CATHOLIQUE DE LOUVAIN, LOUVAIN-LA-NEUVE, BELGIUM

ABSTRACT. Active efflux from procaryotic as well as eucaryotic cells strongly modulates the activity of a largenumber of antibiotics. Effective antibiotic transport has now been observed for many classes of drug efflux pumps.Thus, within the group of primary active transporters, predominant in eucaryotes, six families belonging to theATP-binding cassette superfamily, and including the P-glycoprotein in the MDR (Multi Drug Resistance) groupand the MRP (Multidrug Resistance Protein), have been recognized as being responsible for antibiotic efflux.Within the class of secondary active transporters (antiports, symports, and uniports), ten families of antibioticefflux pumps have been described, distributed in five superfamilies [SMR (Small Multidrug Resistance), MET(Multidrug Endosomal Transporter), MAR (Multi Antimicrobial Resistance), RND (Resistance NodulationDivision), and MFS (Major Facilitator Superfamily)]. Nowadays antibiotic efflux pumps are believed tocontribute significantly to acquired bacterial resistance because of the very broad variety of substrates theyrecognize, their expression in important pathogens, and their cooperation with other mechanisms of resistance.Their presence also explains high-level intrinsic resistances found in specific organisms. Stable mutations inregulatory genes can produce phenotypes of irreversible multidrug resistance. In eucaryotes, antibiotic effluxpumps modulate the accumulation of antimicrobials in phagocytic cells and play major roles in theirtransepithelial transport. The existence of antibiotic efflux pumps, and their impact on therapy, must now betaken fully into account for the selection of novel antimicrobials. The design of specific, potent inhibitorsappears to be an important goal for the improved control of infectious diseases in the near future. BIOCHEM

PHARMACOL 60;4:457–470, 2000. © 2000 Elsevier Science Inc.

KEY WORDS. antibiotic; efflux pump; resistance; transport; chemotherapy; infection

Biological membranes most likely appeared very early onduring evolution to isolate hydrophilic microdomains fromthe surrounding medium, allowing catalyzed reactions tooccur in an efficient manner. Biomembranes, accordingly,constitute efficient barriers towards hydrophilic molecules,most of which can penetrate cells only by specific inwardtransport systems or find their entry restricted to theendocytic pathway. A contrario, biomembranes are easilycrossed by amphiphilic compounds, since these are able todiffuse through both the hydrophilic and the hydrophobicdomains of the bilayer. Therefore, it is not surprising thatmechanisms were devised, also very early, to protect cellsfrom the disordered invasion by amphiphilic molecules,many of which are endowed with biological activitiesleading to potentially harmful effects. A major mechanismin this respect is constituted by active outward transport.Although efflux systems have been known for many years,their importance, both in terms of number and variety ofsubstrates, has become clearly recognized only very re-cently. Drugs are often amphiphilic, whether by selectionor by design, ensuring their wide tissue distribution and/or

their penetration into membrane-protected compartments.Therefore, it comes as no surprise that many drugs shouldfall into this category of exogenous compounds for whichefflux mechanisms, globally referred to as ‘drug effluxpumps,’ are numerous and fairly active [1]. Thus, more andmore membrane-spanning proteins involved in the outwardtransport of a surprisingly large variety of drugs have beenrecognized and characterized over the last years in almostall cell types, from procaryotes and archebacteria throughfungi and higher eucaryotes. This now raises the question ofwhich drug is not transported. For eucaryotic cells, drugefflux pumps have been viewed by many authors as com-plementing the cytochrome P450—or other enzyme-baseddetoxification systems (e.g. Ref. 2) to achieve efficientprotection against ‘chemical invaders’. Both systems, in-deed, show broad specificity and may even work in synergy(see, for example, the concept of a concerted barrier inenterocytes [3, 4]). Drug efflux, indeed, decreases the loadon enzyme-mediated detoxification systems, thereby avoid-ing their saturation, while chemical modifications by theenzyme-based systems, which usually increase the am-phiphilicity of drugs, provide drug pumps with bettersubstrates [2, 5]. Moreover, most drug efflux pumps have abroad substrate specificity and, therefore, may deal with awide range of drugs of completely unrelated pharmacolog-ical classes. The present commentary will focus on antibi-

† Corresponding author: F. Van Bambeke, Pharm.D., UCL 73.70 avenueE. Mounier 73, B-1200 Bruxelles, Belgium. Tel. (32) 2-764-73-78; FAX(32) 2-764-73-73; E-mail: [email protected]

§ Present affiliation: DG Research, European Commission, Brussels,Belgium.

Biochemical Pharmacology, Vol. 60, pp. 457–470, 2000. ISSN 0006-2952/00/$–see front matter© 2000 Elsevier Science Inc. All rights reserved. PII S0006-2952(00)00291-4

Page 2: COMMENTARY Antibiotic Efflux Pumps · the drug efflux pumps in eucaryotic cells ( [7]; drug efflux transporters are classically energized by ATP). The second-ary active transporters,

otics, since the role of drug efflux mechanisms, as a majorcause of bacterial resistance, has been recognized onlyrecently. It therefore needs to be stressed and examined indetail if one wishes to cope with this major challenge inanti-infective therapy. Quite interestingly, also, it nowappears that antibiotic transport mechanisms play impor-tant roles in eucaryotic cells by modulating the pharmaco-logical and toxicological profile of antibiotics. Thus, theidentification and characterization of the correspondinggenes and gene products, in bacteria as well as in eucaryoticcells, have not only an immediate, fundamental interest butalso open interesting perspectives for new and more ratio-nal developments in chemotherapy.

STRUCTURE AND PHYLOGENY

Transporters can be classified on the basis of three maincriteria, namely the energy source, the phylogenic relation-ship, and the substrate specificity. A 4-digit nomenclaturehas been proposed recently, constructed in analogy withenzyme nomenclature and in which the first group of digitsrefers to the mode of transport and energy source, thesecond and the third to the phylogeny (superfamilies andfamilies), and the fourth to the substrate ( [6, 7]; see also theweb site ,http://www.biology.ucsd.edu/;msaier/transport/titlepage.html.). The so-called primary active transport-ers use various forms of energy and constitute the bulk ofthe drug efflux pumps in eucaryotic cells ( [7]; drug effluxtransporters are classically energized by ATP). The second-ary active transporters, acting as symports and antiports(i.e. coupling the drug efflux to the downhill transport of anion along a concentration gradient), are predominant inbacteria [6]. Within each of these two main classes, phylo-genic studies have led to the recognition of superfamilies,families, and clusters, in correlation with their substratespecificity [6]. Yet, most drug efflux pumps confer a multi-drug resistance phenotype, corresponding to the largevariety of substrates they may recognize, including severalclasses of antibiotics as well as non-antibiotic drugs. Anti-biotic-specific efflux pumps appear to be restricted to thoseorganisms producing antibiotics and are often an integralpart of the corresponding biosynthetic pathway [8].

Table 1 lists the main drug efflux pumps acting onantibiotics described thus far, within the correspondingfamilies of drug transporters. We also mention in whichorganisms they are mainly found (common bacteria, thespecial category of antibiotic-producing organisms, com-mon fungi, and mammalian cells). The secondary activetransporters, being of a simpler structure, will be presentedfirst. In this group, pumps acting on antibiotics are found inthe so-called SMR* family, the MET family, the RND

family, the MFS, and the MAR family. SMR have beenfound thus far only in bacteria [9], whereas MET [10, 11]seem to be restricted to superior eucaryotes. Other familiesare widely distributed, but antibiotic efflux pumps have notbeen described in all organisms (viz. the RND [12]). TheSMR family can be divided into two groups of geneproducts; one of them is immediately related to drug efflux,whereas the other is not, which suggests that evolutionfrom the ancestor transporter towards a gene product witha drug efflux phenotype occurred only once [9, 13]. Highlyconserved motifs are involved in transport activities as wellas in binding of the substrates. MET members present ageneral organization similar to that of SMR members butare characterized by signal motifs rich in tyrosines at theirC-termini, which direct them to intracellular compart-ments of eucaryotic cells [11]. Members of the RNDsuperfamily share common topological features, namely 2large extracytoplasmic loops and 12 transmembrane seg-ments resulting from an internal duplication of a geneencoding a 6-transmembrane segment (see Ref. 12 forreview). All the known members of this superfamily havethe function of efflux transporter, but proteins conferringmultidrug resistance are grouped in two of the sevenfamilies recognized in this superfamily. Thus, the HAE1family is largely predominant and includes the well knowndrug efflux pumps of Gram-negative bacteria with verybroad substrate specificity. They probably all derive from asingle ancestor [13, 14]. In contrast, there is only onemember characterized so far in the HAE2 family. Yet,Mycobacterium tuberculosis possesses 10 genes encodingproteins of this family, which could explain the poorsensitivity of this organism to many common antibiotics, ifall of them were to correspond to antibiotic efflux pumps[12]. The situation is more complex for drug efflux pumpsbelonging to the MFS [15, 16]. Indeed, drug-specific andmultidrug efflux pumps appear to be randomly interspersedin families (see, for example, the DHA2 [also calledDHA14] family), indicating that narrowing and broadeningof specificity have occurred repeatedly during evolution[13, 14]. Yet, they are also thought to derive from acommon ancestor. Moreover, sequence analysis suggeststhat a simple duplication of a gene encoding a 6-transmem-brane segment protein led to the appearance of the 12-transmembrane segment family (DHA1 [also calledDHA12]). Then, the 14-transmembrane segment family(DHA2 [also called DHA14]) evolved from the insertion ofan increasingly hydrophobic central loop of the DHA12precursor into the membrane (the DHA12 family actuallydisplays a long intracytosolic peptide loop running betweenthe 6th and the 7th transmembrane segments, which mayhave provided the necessary scaffold for these two addi-tional membrane spanning segments seen in DHA14).

* Abbreviations: ABC, ATP Binding Cassette; CFTR, Cystic FibrosisTransmembrane conductance Regulator; CT, Conjugate Transporters;DHA, Drug:H1 Antiporters; HAE, Hydrophobic Amphiphilic Efflux;MAR, Multi Antimicrobial Resistance; MET, Multidrug EndosomalTransporters; MDR, Multi Drug Resistance; MFS, Major FacilitatorSuperfamily; (c)MOAT, (canalicular) Multispecific Organic Anion

Transporters; MRP, Multidrug Resistance Proteins; OAT, Organic AnionTransporters; OCT, Organic Cation Transporters; PgP, P-glycoprotein;PDR, Pleiotropic Drug Resistance; RND, Resistance Nodulation Division;SET, Sugar Efflux Transporters; and SMR, Small Multidrug Resistance.

458 F. Van Bambeke et al.

Page 3: COMMENTARY Antibiotic Efflux Pumps · the drug efflux pumps in eucaryotic cells ( [7]; drug efflux transporters are classically energized by ATP). The second-ary active transporters,

Some of the conserved motifs throughout the MFS havebeen shown to play a role in activity. The MAR familypresents no sequence homology with MFS, but a similartopological organization [17].

Within the group of primary active transporters, theABC drug efflux pumps have been classified extensively infamilies according to structural homology [18–22]. Morerecently, they have been distinguished phylogenetically onthe basis of their import or export activity [23]. Whereasimport pumps are present only in procaryotes, efflux pumpsare maintained in both procaryotes and eucaryotes, suggest-ing that selection of the transport directionality occurredbefore divergence between procaryotes and eucaryotes [23].ABC domains generally present a high degree of homology,whereas transmembrane domains differ between transport-ers, and might contribute to defining their substrate speci-ficity [18]. Considering efflux pumps only, two families, theMDR and the CT2, deserve special attention since theycorrespond to the most studied and pharmacologicallyimportant transporters. They also show functional inter-changeability between different types of organisms [24, 25].The MDR family, which includes the well-known P-glycoprotein in eucaryotes (PgP, a product of the MDR1gene as shown in Table 1), is responsible for the efflux of awide range of drugs besides antibiotics, including antican-cer agents. The CT2 family, comprising MRP in superioreucaryotes and Yor1 in Saccharomyces cerevisiae, also isinvolved in the efflux of many drugs, again includingantibiotics and anticancer agents [2, 21, 26–28]. Thisfamily is phylogenetically close to the CFTR family, achloride channel, which, however, does not transport drugs(its mutation causes cystic fibrosis). The PDR transportersshare several biochemical features with the human PgP[19–21] and constitute the major class of ABC drug effluxpumps in yeasts and fungi. Finally, proteins from theDrugE1 family are involved in drug-specific efflux in anti-biotic-producing organisms [13, 23].

ORGANIZATION AND FUNCTION

Figure 1 shows in a combined fashion the topologicalorganization of the main antibiotic-extruding pumps pre-sented in Table 1 together with a schematic view of theirmode of operation and the type of antibiotics transported.The mechanisms of transport and of substrate recognition,however, remain largely unknown in most instances, andmany of the current views are based on extrapolations fromdata obtained with transporters of physiological substrates(the assumption is that proteins deriving from a commonancestor have maintained sufficient similarity not only ofstructure but also of function throughout evolution). SMR,RND, and most MFS transporters (and probably also theMET and MATE transporters) use a proton gradient as thedriving force. A minimum of 12 transmembrane segmentsseems required for activity, so that SMR transporters areprobably organized in trimers [29, 30]. The putative mech-anism of drug transport, as established by site-directed

mutagenesis of a SMR transporter, could involve thefollowing steps: (i) exchange between the drug and a protonfixed on a charged residue; (ii) translocation of the drug bya series of conformational changes driving it through ahydrophobic pathway; and (iii) replacement of the drug bya proton in the external medium and return to the initialconformational state [14, 31]. The overall result of thetransport is therefore an exchange between the drug and aproton (antiport). The residue responsible for proton ex-change in SMR could be a conserved glutamate [32]. Thesame mechanism probably applies to MFS transport, but theproton exchanger may be a conserved arginine [29] (again,a conserved acidic residue may be involved in the recogni-tion of positively charged substrates [33]). This mode ofextrusion explains why both SMR and MFS transporterspreferentially extrude cationic molecules (see below). Lessis known about transport by RND members. Like SMR andMFS transporters, however, RND transporters possesshighly conserved charged residues in their transmembranesegments, which may play an important role in substratebinding or proton transport [34]. In Gram-negative bacte-ria, the 2 large extracellular loops of RND are thought tointeract with two other proteins [34], namely the ‘mem-brane fusion protein’ (connecting the inner membrane tothe outer membrane) and the ‘outer membrane protein’(crossing this outer membrane). This tripartite proteincomplex allows the bacteria to extrude the substrate di-rectly into the external medium, bypassing the periplasmand thereby amplifying the efficacy of the transporter. It isnow proposed to be a common feature for RND, MFS, andABC efflux pumps in Gram-negative bacteria, with the‘membrane fusion proteins’ differing between transportsystems, whereas the ‘outer membrane proteins’ are proba-bly common to all three transporters [35]. ABC transport-ers, which contain two ATP binding cassettes (ABCdomains), derive their energy from the hydrolysis of ATP.An additional characteristic of MRP transporters is thattheir activity is strictly dependent on the presence ofglutathione. Whether glutathione directly activates thetransporter, or is itself co-transported with the drug, or evenforms a conjugate to the drug, is not clear, but its weakionic character is thought to be critical (see Ref. 26 forreview). As for proton antiporters, a conformational changeof the ABC protein is necessary for drug extrusion andprobably is triggered by drug binding and ATP hydrolysis[36, 37].

The exact mechanism of drug transport is still contro-versial. Among the different models that have been pro-posed, the two most likely ones present efflux pumps asacting either like hydrophobic ‘vacuum cleaners’ or likeflippases. In the first model, the drug is thought to movefreely into the lipid phase of the membrane, then reachingthe protein and its central channel, from where it is activelyexpelled outwardly. In the second model, the drug is alsothought to reach the protein from within the membrane,but then would be flipped to the outer layer (as proposed forphosphatidylcholine flip-flop under the action of flippases).

Antibiotic Efflux Pumps 459

Page 4: COMMENTARY Antibiotic Efflux Pumps · the drug efflux pumps in eucaryotic cells ( [7]; drug efflux transporters are classically energized by ATP). The second-ary active transporters,

TA

BL

E1.

Cla

ssif

icat

ion*

and

illus

trat

ive

exam

ples

ofdr

ugef

flux

pum

psex

trud

ing

the

anti

biot

ics

used

incl

inic

alpr

acti

ce†

Mec

hani

smof

tran

spor

tSu

perf

amily

‡Fa

mily

‡B

acte

ria

Ant

ibio

tic

prod

ucer

sFu

ngi

Mam

mal

s

#2.(

1–77

):Se

cond

ary

acti

vetr

ansp

orte

r(s

ympo

rts,

anti

port

s,un

ipor

ts)

#2.7

.(1–

2)SM

RSm

all

Mul

tidr

ugR

esis

tanc

e

#2.7

.1.

Smal

lM

ulti

drug

Efflu

x

●Em

rEof

E.

coli

(ery

,sul

f,te

t)M

mr

ofM

.tu

berc

ulos

is(e

ry)

#2.7

4.(1

)M

ETM

ulti

drug

Endo

som

alT

rans

port

er

#2.7

4.1.

●M

TP

ofM

usm

uscu

lus

(ery

)

#2.6

.(1–

7)R

ND

Res

ista

nce

Nod

ulat

ion

Div

isio

n

#2.6

.2.H

AE1

Hyd

roph

obe

Am

phip

hile

Efflu

x

●A

crof

E.

coli

(bla

c,ch

l,er

y,fu

s,na

l,ri

f,te

t)●

Mex

ofP.

aeru

gino

sa(a

g,b

lac,

inhi

b,b

lac’

ase,

chl,

fq,f

us,1

4m

L,15

mL,

rif,

tet)

●M

trD

ofN

.go

norr

heae

(bla

c,ch

l,er

y,fu

s,ri

f,te

t)●

Am

rBof

B.

pseu

dom

alei

(ag,

ery)

#2.6

5.H

AE2

Hyd

roph

obe

Am

phip

hile

Efflu

x

●A

ctll3

ofS.

coel

icol

or(a

ctin

orho

din)

#2.1

.(1–

29)

MFS

Maj

orFa

cilit

ator

Supe

rfam

ily

#2.1

.2.D

HA

1D

rug:

H1

Ant

ipor

ters

-112

span

ners

(DH

A12

)

●T

etA

,B,E

ofE

.co

li(c

hl,n

al,t

et)

●T

etC

ofP.

aeru

gino

sa(t

et)

●T

etH

ofP.

mul

toci

da(t

et)

●C

mlA

ofP.

aeru

gino

sa(c

hl)

●B

crof

E.

coli

(sul

f)●

Nor

Aof

S.au

reus

(chl

,fq,

tet)

●B

ltof

B.

subt

ilis

(chl

,fq)

●C

aMD

R1

ofC

.al

bica

ns(a

zo,c

hl)

●Fl

r1of

S.ce

revi

siae

(azo

)

●V

MA

T1

ofR

attu

sno

rveg

icus

(mon

oam

ines

)

#2.1

.3.D

HA

2D

rug:

H1

Ant

ipor

ters

-214

span

ners

(DH

A14

)

●Em

rBof

E.

coli

(nal

)●

Mdf

Aof

E.

coli

(ag,

chl,

ery,

fq,r

if,te

t)●

LfrA

ofM

.sm

egm

atis

(fq)

●T

etK

ofS.

aure

us(t

et)

●Pt

rof

S.pr

istin

aesp

iralis

(pri

s,re

f)●

LmrA

ofS.

linco

lnen

sis(l

in)

●Pu

r8of

S.lip

man

ii(p

ur)

●R

ifPof

A.

med

iterr

anei

(rif)

●A

tr1

ofS.

cere

visia

e(a

zo)

#2.1

.19

OC

TO

rgan

icC

atio

nT

rans

port

ers

●O

ct1

ofR

.no

rveg

icus

(org

anic

cati

ons,

xeno

biot

ics)

[hum

anho

mol

ogs:

Oct

1–2]

●O

at1

ofR

.no

rveg

icus

(org

anic

anio

ns,b

lac)

[hum

anho

mol

ogs:

Oat

1–3]

460 F. Van Bambeke et al.

Page 5: COMMENTARY Antibiotic Efflux Pumps · the drug efflux pumps in eucaryotic cells ( [7]; drug efflux transporters are classically energized by ATP). The second-ary active transporters,

#2.1

.20.

SET

Suga

rEf

flux

Tra

nspo

rter

s

●Se

tAof

E.

coli

(su

gars

;ag)

#2.1

.21.

DH

A3

Dru

g:H

1

Ant

ipor

ters

-312

span

ners

●M

efA

ofS.

pyog

enes

(14m

l,15

ml,

ole)

●M

efE

ofS.

pneu

mon

iae

(14m

l,15

ml)

●C

mr

ofC

.gl

utam

icum

(chl

,ery

,pur

,tet

)●

Tet

Vof

M.

segm

atis

(tet

)●

Tap

ofM

.fo

rtui

tum

and

M.

tube

rcul

osis

(tet

)

#2.6

6.(1

)M

AR

Mul

tiA

ntim

icro

bial

Res

ista

nce

#2.6

6.1.

●N

orM

ofV

.pa

raha

emol

ytic

us(a

g,fq

)

#3.(

1–11

):Pr

imar

yac

tive

tran

spor

ters

#3.1

.(1–

70):

AB

CA

TP

Bin

ding

Cas

sett

e

#3.1

.35.

Dru

gE1

Dru

gEx

port

er-1

●M

srA

ofS.

epid

erm

idis

(ery

)●

Ole

Cof

S.an

tibio

ticus

(ole

)●

Srm

Bof

S.am

bofa

cien

s(m

l)●

Tir

cof

S.fr

adia

e(t

yl)

#3.1

.47.

Dru

gE2

Dru

gEx

port

er-2

●Lm

rAof

L.la

ctis

(dru

gs)

#3.1

.61.

MD

RM

ulti

drug

Res

ista

nce

●M

DR

1of

H.

sapi

ens§

(pho

spho

lipid

s;fq

,lm

,ml,

rif,

tet)

#3.1

.65

PDR

Plei

otro

pic

Dru

gR

esis

tanc

e

●Pd

r5of

S.ce

revi

siae

(azo

,chl

,ery

,lm

,tet

)●

Snq2

ofS.

cere

visia

e(a

zo)

●C

DR

1of

C.

albi

cans

(azo

,chl

)●

Atr

A,B

ofA

.ni

dula

ns(a

g,az

o)

#3.1

.67

CT

1C

onju

gate

Tra

nspo

rter

-1

●Y

cf1

ofS.

cere

visia

e(c

onju

gate

s)

#3.1

.68

CT

2C

onju

gate

Tra

nspo

rter

-2

●Y

or1

ofS.

cere

visia

e(e

ry,t

et)

●M

RP1

–6of

H.

sapi

ensi

(con

juga

tes,

phos

phol

ipid

s;fq

)

*Thi

scl

assi

ficat

ion,

base

don

that

prop

osed

inR

efs,

6an

d7,

cons

ider

sfir

stth

em

ode

oftr

ansp

ort

and

ener

gyso

urce

,and

wit

hin

each

mec

hani

smra

nks

the

tran

spor

ters

phyl

ogen

etic

ally

.The

first

digi

tin

each

num

ber

refe

rsto

the

type

oftr

ansp

orte

r,th

ese

cond

one,

toth

e(s

uper

)fam

ilyan

dth

eth

ird

one,

toth

e(s

ub)f

amily

.Int

erva

lsgi

ven

betw

een

brac

kets

corr

espo

ndto

the

num

bero

ffam

ilies

iden

tifie

dso

fari

na

supe

rfam

ily,o

rofs

ubfa

mili

esin

afa

mily

.Not

eth

atm

echa

nism

#1an

d#4

and

the

corr

espo

ndin

gtr

ansp

orte

rsar

eno

tsh

own

here

sinc

eth

eydo

not

corr

espo

ndto

drug

efflu

xpu

mps

.In

each

fam

ily,w

epr

esen

ton

lyth

ose

subf

amili

esw

here

anti

biot

ictr

ansp

orte

rsha

vebe

enev

iden

ced.

Thi

scl

assi

ficat

ion

isup

date

dre

gula

rly.

(see

http

://w

ww

-bio

log.

ucsd

.edu

/;m

saie

r/tr

ansp

ort/

titl

epag

e.ht

ml;

http

://w

ww

-bio

log.

ucsd

.edu

/;m

saie

r/tr

ansp

ort/

2_1.

htm

l;/2

.6.h

tml;

/22

7.ht

ml;

/22

66.h

tml;

/2_7

4.ht

ml;

/3_1

.htm

l).

† Illus

trat

ive

exam

ples

ofkn

own

antib

iotic

tran

spor

ters

;sub

stra

telis

tsar

eno

tex

haus

tive

but

corr

espo

ndto

the

pres

ent

stat

eof

know

ledg

e(o

nly

the

mai

ncl

inic

ally

rele

vant

antib

iotic

san

dan

tifun

gals

tran

spor

ted

are

show

n;fo

rso

me

tran

spor

ters

,we

give

the

phys

iolo

gica

lsub

stra

teif

know

n).A

bbre

viat

ions

:ag,

amin

ogly

cosid

es;a

zo,a

zole

s;b

lac;

b-la

ctam

s;in

hib

bla

c’as

e,in

hibi

tor

ofb

-lact

amas

e;ch

l,ch

lora

mph

enic

ol;e

ry,e

ryth

rom

ycin

;fq,

fluor

oqui

nolo

nes;

fus,

fusid

icac

id;l

m,l

inco

sam

ides

;ml,

mac

rolid

es(1

4ml,

mac

rolid

esw

itha

14-a

tom

mac

rocy

cle;

15m

l,m

acro

lides

with

a15

-ato

mm

acro

cycl

e);n

al,n

alid

ixic

acid

;ole

,ole

ando

myc

in;p

ris,p

ristin

amyc

in;p

ur,p

urom

ycin

;rif,

rifam

pici

n;su

lf,su

lfam

ides

;tet

,ter

acyc

lines

;and

tyl,

tylo

sine.

‡So

me

supe

rfam

ilies

and

fam

ilies

are

som

etim

esre

ferr

edto

asfa

mili

esan

dsu

bfam

ilies

whe

nth

enu

mbe

rof

mem

bers

inth

egr

oup

issm

all.

§M

DR

1is

also

refe

rred

toas

PgP.

iM

RP2

,MR

P3,M

RP4

,and

MR

P5ar

eal

sore

ferr

edto

ascM

OA

T,M

OA

T-D

,MO

AT

-B,a

ndM

OA

T-C

[116

].

Antibiotic Efflux Pumps 461

Page 6: COMMENTARY Antibiotic Efflux Pumps · the drug efflux pumps in eucaryotic cells ( [7]; drug efflux transporters are classically energized by ATP). The second-ary active transporters,

In contrast to the previous, more conventional models ofsimple pore-like channels oriented towards export, whichpick up their substrate from the cytosol and orient ittowards the outside of the cell thanks to a regulator valve,the two models presented assume that the drug is extractedprimarily from the inner phase of the membrane. Indeed,strong membrane anchoring is probably a common require-

ment for substrates of drug efflux pumps. Moreover, trans-port has been suggested to occur not only from the cytosolicface of the membrane of eucaryotes and Gram-positiveorganisms (see data with the MFS transporter QacA ofStaphylococcus aureus; [38]), but also from both the innerand the outer leaflets of the inner membrane of Gram-negative organisms, ensuring clearance from the cytosol as

FIG. 1. Topology, mechanism of action, and typical substrates of the main classes of antibiotic efflux pumps (constructed from data andschemes adapted from Refs. 14, 117, and 118). The MET and MAR families are not represented since information on those pumpsis still scanty (MET pumps possess 4 transmembrane segments, like SMR pumps; MAR pumps present the same topologicalorganization as MFS). Topology: the membrane is represented in grey, with the extracellular milieu at the top of each scheme, andproteins as a chain of circles, the solid ones corresponding to conserved motifs (identified by letters). In ABC transporters, the locationsof the ATP binding cassettes are indicated by two black circles. The 5 transmembrane segments at the N-terminal part of MRP (in thediscontinuous line square) are present in MRP1, MRP2 (also called cMOAT), and MRP3 [116]. Numerous other organizations ofABC and transmembrane domains have been proposed (for instance, a mirror image of that shown here for the Pgp, or ABCtransporters in which transmembrane segments and ATP binding domains are not fused [22]). Mechanism of action: SMR, RND, andMFS transporters are drug-H1 antiporters. H1 probably is exchanged from a conserved glutamate (E) of the ‘a’ conserved domain inSMR and from a conserved arginine (R) of the ‘b’ conserved domain in MFS. Recognition of cationic drugs may imply the sameconserved glutamate for SMR transporters and a conserved acid residue (glutamate or aspartate; E in the figure) in the ‘d’ domain forMFS transporters. No corresponding data are available so far for RND transporters. ABC transporters involved in drug efflux use ATPas an energy source. In all types of transporters, the drug seems to be extracted from the membrane rather than from the cytosol. Thetransporter then could act as a flippase (catalyzing the flip-flop of the drug from the inner to the outer face of the membrane) or as a‘vacuum cleaner’ (moving the drug from the membrane to the central domain of a channel closed to the cytosolic face of the membranebut open to its extracellular face, as already shown for MDR [also known as PgP]). MRP transporters also require the presence ofglutathione, which could be conjugated to the drug prior to or during extrusion. Antibiotics: classes of antibacterial agents for whichtransport has been described for at least one pump in each family are grouped according to their general physicochemical character (seeFig. 2).

462 F. Van Bambeke et al.

Page 7: COMMENTARY Antibiotic Efflux Pumps · the drug efflux pumps in eucaryotic cells ( [7]; drug efflux transporters are classically energized by ATP). The second-ary active transporters,

well as from the periplasm (see the model of extrusion byRND transporters in Refs. 39 and 40). Several lines ofexperimental evidence, which, however, are based mostoften on studies with non-antibiotic substrates, support thisassumption. First, the only common feature of all pumpsubstrates is a liposolubility that must be sufficient to ensurea proper penetration into the bilayer [41]. Second, substrateor inhibitor molecules can compete with lipophilic fluoro-phores for insertion into the membrane [41, 42]. Third,lipophilic fluorophores are themselves substrates of thepumps [43]. Fourth, the rates and kinetics of efflux are notdirectly related to the cytosolic drug content [41, 44]. Thestructural characteristics of MDR proteins rather favor thehypothesis of a vacuum cleaner. Indeed, the 12 transmem-brane domains are organized in such a way that they forma large aqueous pore, open to the extracellular medium butclosed towards the cytosol (somewhat like an empty, openbottle or beaker floating on the surface of water [44, 45]). Inaddition, these transmembrane segments are rich in aro-matic amino acids, which could help the hydrophobicsubstrates to travel into this channel [46]. Functionalstudies, in contrast, favor the flip-flop hypothesis, since thephysiological transporters of phospholipids and of glutathi-one conjugates, which belong to the MDR and CT2families, respectively, are known as flippases [47–49].

ANTIBIOTIC SUBSTRATE SPECIFICITY

A key characteristic of the antibiotic efflux pumps is thevariety of molecules they may transport, which actually canbe related directly to their well-known poor substratespecificity. Considering the pharmacochemical aspects first,it is clear that only very minimal common structuraldeterminants are necessary to obtain detectable transport.Nevertheless, for each class of transporters, investigatorshave tried to determine which substrate features are themost specific. Results available thus far are presented in asummarized fashion in Fig. 1. Through the use of simulta-neous multiple disruption of several transporter genes,however, it has become evident that the substrate specific-ity is very broad and overlapping across a large array ofdistinct transporters [28, 50, 51]. The substrate specificity ofthe MET and the MAR transporters, for instance, has notyet been established with sufficient details and, for theMAR transporters, appears dubious since the NorM trans-porter representative of this family is claimed to recognizeboth fluoroquinolones and aminoglycosides, two classes ofantibiotics with strikingly different physicochemical prop-erties [17]. Although these substrate specificities may ap-pear difficult to establish, a unifying hypothesis is that most,if not all, transporters recognize molecules with a polar,often slightly charged head associated with a hydrophobicdomain (see Fig. 2). Unfortunately, the importance oflipophilia is often difficult to ascertain in the absence ofpublished studies with homogenous series of drug deriva-tives. It, nevertheless, appears striking for the RND multi-drug efflux pumps when considering the data available on

the b-lactams, for which a transport ranking of cloxacil-lin . nafcillin . penicillin G . carbenicillin . penicillinN (the latter being almost not transported) has beendemonstrated clearly in close relationship with their corre-sponding octanol/water partition coefficients [52]. Gener-ally speaking, also, the physicochemical properties of theantibiotics transported by a given class of pumps correspondto those of the non-antibiotic drugs as well as of those ofthe putative physiological substrates. A major discrepancy,however, concerns chloramphenicol, which is extruded byMFS despite its neutral character [e.g. Ref. 53]. Site-directed mutagenesis studies, however, suggest that therecognition of this drug is probably mediated by interac-tions different from those observed for other antibiotics[33]. Also remarkable is the rarity of pumps for aminogly-cosides [54, 55], but this can probably be explained by thehigh hydrophilicity of these antibiotics, which preventstheir entry into cells by nonspecific diffusion (aminoglyco-side antibiotics largely mimic polyamines, an essentialsubstrate for many types of cells, and use their inwardtransport system for entering both bacteria [for activity] andspecific eucaryotic cells [causing toxicity; see Refs. 56 and57 for reviews]; aminoglycoside-producing organisms gen-erally tend to protect themselves not by efflux pumps, butby the production of aminoglycoside-inactivating enzymes[58]). Similarly, the absence of efflux pumps acting onglycopeptide antibiotics has been explained, at least inbacteria, by the fact that these bulky, largely hydrophilicdrugs act in the outer space of Gram-positive bacteria, andnever cross the bacterial membrane (glycopeptides areinactive against Gram-negative bacteria precisely becausethey cannot cross the outer membrane of these organisms[59]). But the situation has evolved so quickly for the otherclasses of antibiotics that we cannot exclude the possibilitythat efflux will eventually be demonstrated for glycopep-tides also if appropriate studies are undertaken. Beyondthese specific considerations, it must also be emphasizedthat a given antibiotic may be a substrate for different typesof pumps, so that (i) it may be expelled by differentorganisms for which no common transporter has beenidentified so far (giving the false impression that thetransporter is ubiquitous), and (ii) modulation of theactivity of a given transporter may be compensated for by amodulation in the opposite direction of another trans-porter, with, therefore, no or little change in the cellularaccumulation of the drug (giving the false impression thatthe drug is not transported). Moreover, a given pump mayextrude not only different antibiotics within the same classbut also different classes of antibiotics [28, 39, 40, 55, 60].Finally, a single cell may possess a vast and complex arsenalof efflux pumps allowing for the extrusion of a very broadspectrum of drugs (viz. Escherichia coli and S. cerevisiae, thecomplete genome sequencing of which has revealed theexistence of more than 250 putative transporters monopo-lizing 10% of the total genetic material [6, 7]).

Antibiotic Efflux Pumps 463

Page 8: COMMENTARY Antibiotic Efflux Pumps · the drug efflux pumps in eucaryotic cells ( [7]; drug efflux transporters are classically energized by ATP). The second-ary active transporters,

MICROBIOLOGICAL AND THERAPEUTICSIGNIFICANCE

Considered clinically important only for tetracyclines untila few years ago, antibiotic efflux pumps appear nowadays asa major component of microbial resistance to many classesof major antibiotics [39, 40, 60]. For at least three of them,namely the tetracyclines [61, 62], the macrolides [63], andthe fluoroquinolones [64] (which are interesting to analyzein this context since they are totally synthetic, amphiphiliccompounds with no known ‘natural’ counterpart), antibi-otic efflux appears sufficient per se to confer a medium orhigh level of resistance, defeating medically applicabletreatments of the corresponding infections with theseantibiotics. Typical examples include Streptococcus pyogenes[65] and to some extent S. pneumoniae [65, 66]. Antibioticefflux may also contribute to the decreased susceptibility ofS. aureus to fluoroquinolones [64, 67] and of Pseudomonas

aeruginosa to many classes of antibiotics [68]. Most insidi-ously, antibiotic efflux may be found in association withother mechanisms, such as antibiotic inactivation, to con-fer high-level resistance on bacteria. In some respects, thisphenomenon bears similarities with the cooperation ofdrug-extruding pumps and the cytochrome P450-based deg-radation pathways in enterocytes, which we presented inthe introduction of this review. A typical example is givenby the cooperation between the penicillin efflux pumps andthe b-lactamases, both of which may effectively decreasethe concentration of b-lactams in the periplasmic space ofGram-negative bacteria to the point where penicillin-binding proteins are no longer saturated. These bacteriathen display the surprising phenotype of high-level resis-tance without being high-level producers of b-lactamase[69–71]. The situation is more subtle for fluoroquinolones,but illustrates quite well the cooperation between two

FIG. 2. Structural formulae of the main antibiotics for which efflux has been demonstrated through at least one type of the transportersdescribed in Table 1 and Fig. 1. The molecules are presented so as to show their amphipathic character when appropriate (the zonesconsidered more lipophilic are surrounded by a solid line, and those considered more hydrophilic, by a dotted line). The charged groupsare systematically oriented upwards (fluoroquinolones are represented twice, since these can act as cations as well as anions, andaccordingly are transported by RND, MFS, MDR, and MRP pumps). The local pH, which may vary widely from one type of organismto another and from the precise location of the pumps, strongly influences the ionization of these groups and their role in recognitionby the transporters. The structures shown emphasize the common characters of each class of antibiotics, since structural variationswithin each class (denoted by the existence of variable substituents [R]) do not appear to alter their recognition properties markedly,systematically, or specifically.

464 F. Van Bambeke et al.

Page 9: COMMENTARY Antibiotic Efflux Pumps · the drug efflux pumps in eucaryotic cells ( [7]; drug efflux transporters are classically energized by ATP). The second-ary active transporters,

apparently unrelated mechanisms of resistance. Resistanceto these antibiotics may result from point mutations at thelevel of the drug targets (DNA gyrase/topoisomerase IV). Asingle mutation most usually gives rise to only low- ormedium-level resistance, and the bacteria may still beconsidered as sensitive in routine microbiological testing.The combination of two or more mutations, however, willconfer high levels of resistance [72, 73]. Because thesemutations are easily obtained in many bacterial species invitro, they were considered as being primarily responsiblefor the resistance seen in the clinic. It now appears,however, that many, if not most, of the organisms with thephenotype of low- to medium-level resistance to fluoro-quinolones harbor one or several efflux mechanisms [73–75]. Recent data also show that single point mutations andthe existence of fluoroquinolone efflux pumps producesynergistic effects [74, 76]. We speculate that efflux pumps,by decreasing the cellular concentration of fluoroquino-lones, may facilitate the selection of mutants with two ormore mutations, thereby increasing the risk of emergence ofhighly resistant organisms. A further striking demonstra-tion of the key role of antibiotic efflux pumps in bacterialresistance is given by the recent observation that thedisruption of one or several of their genes, or their directpharmacological inhibition, results in a major increase oftheir intrinsic susceptibility to the corresponding antibiot-ics. It may also decrease the frequency of appearance ofresistant mutants [51, 76]. Moving to so-called non-suscep-tible organisms, it now appears that, in many cases, thisphenotype is not due to an intrinsic lack of susceptibility(absence of target or impermeability) as was thought for along time, but is rather caused by the presence of effluxpumps that are constitutionally very active against the drugunder study [40, 54, 76–78]. It is also important tounderscore the role of stable mutations at the level of theregulatory genes controlling the expression of multidrugpumps [40]. An example of this multidrug regulation circuitis well described in yeast, where it has been shown thatexposure to a given single drug could lead to mutations inregulatory genes provoking the constitutive and simulta-neous overexpression of several multidrug efflux pumps (ofdifferent types). This results in the irreversible acquisitionof a phenotype of multidrug resistance (see Ref. 79 forreview), a situation commonly observed in pathogenicfungi resistant to multiple drugs, which indeed often over-express multidrug efflux pumps [80, 81]. Moving to bacte-ria, we now know that patients infected by P. aeruginosaand treated by a b-lactam alone (or in combination) maybecome colonized rapidly by strains with a mutation in theregulator of the genes encoding the MexA–MexB–OprMpump [82]. These strains are resistant not only to b-lactams,but also to fluoroquinolones, tetracyclines, chlorampheni-col, and trimethoprim. Each drug, presumably, can beexpected to also be the inducer of regulatory mechanismsresponding to cytotoxic insult by overexpression of drugefflux pumps. Therefore, it is likely that ecological pressurethrough an inappropriate use of antibiotics will sooner or

later lead to the selection of strains showing stable resis-tance to a wide range of unrelated drugs (this may imply notonly bacteria but also other organisms that have beenexposed incidentally to the same type of drug). A second,but so far unproven, risk is related to the apparent plasticityand promiscuity of the drug transporters with respect totheir potential substrates. This could lead to the selection oftransporters with increased efficacy. This self-adaptation ofbacteria has been well documented with b-lactamases andaminoglycoside-inactivating enzymes. Each introduction ofmolecules designed to resist the action of these enzymes hasbeen followed rapidly by the emergence of apparently newenzymes acting against these ‘new’ antibiotics. Yet, thegenetic analysis showed that the new enzymes sometimesdiffered from the old ones only by single amino acidsubstitutions [56]. Likewise, in the case of drug effluxpumps, it is known that single amino acid substitutions mayaffect substrate specificity drastically [6, 33].

The clinical impact of antibiotic efflux pumps on resis-tance in clinics remains difficult to establish, since we lacklarge-scale and international statistics comparing theirprevalence with that of the other resistance mechanisms.Yet, very recent surveys already published point to alarmingfigures of 40–90% of some bacterial pathogens (S. pneu-moniae, S. pyogenes, and P. aeruginosa) bearing effluxmechanisms for the major classes of clinically availableantibiotics ( [66, 78, 83–85]; see also abstracts 1211, 1212,1216–1218, 1220–1223, 1225, and 1228 of the 39thInterscience Conference on Antimicrobial Agents andChemotherapy [ICAAC], San Francisco, CA, 1999). Theabundance of research papers describing antibiotic effluxmechanisms contrasts, however, with the rarity of data fromclinical microbiology laboratories. This raises the questionof the adequacy of the routine procedures to detect thesestrains, which most often may be classified as moderatelyresistant and erroneously assigned to a conventional mech-anism of resistance in the absence of further detailedinvestigation. Multiresistant organisms will need to bescreened critically in this context.

Moving now to eucaryotes, it becomes very clear nowa-days that the transepithelial movement of several drugs,including antibiotics, implies transport systems. The con-certed action of different pumps located both on thebasolateral and the apical membranes of epithelial cells hasbeen proposed to account for the preferential transfer ofcertain antibiotics from the blood to the excretory pathway.This cooperation is best evidenced in the liver, where OATand cMOAT (also called MRP2, see Table 1) ensure theunidirectional transfer of drugs to the bile (see Ref. 86 forreview). It is also suspected to occur in the kidney proximaltubules [87], and secretory transepithelial transport ofantibiotics has been demonstrated in the intestine andairway epithelia [88, 89]. Practically speaking, the activityof pumps explains the poor intestinal resorption of severalantibiotics [88, 90, 91] and gives a rational explanation forthe behavior of the so-called orally available penicillins orcephalosporins. The recognition of the existence of antibi-

Antibiotic Efflux Pumps 465

Page 10: COMMENTARY Antibiotic Efflux Pumps · the drug efflux pumps in eucaryotic cells ( [7]; drug efflux transporters are classically energized by ATP). The second-ary active transporters,

otic transporters now may be put into use for more rationaldesign in the future [92]. Similarly, antibiotic transportmechanisms operating in the liver and kidney explain someof the elimination features of b-lactams and fluoroquino-lones [93, 94]. The recognition of the existence of antibi-otic-extruding pumps in macrophages, which act on b-lac-tams and fluoroquinolones [95, 96], and, for MDR-expres-sing cells, on macrolides, tetracyclines, lincosamides, andrifamycins as well [97, 98], has shed new light on the lackof, or potentially reduced activity of, these antibioticsagainst intracellular bacteria. Antibiotic efflux pumps will,indeed, reduce the amount of drug present in phagocytes toa point where it may no longer exceed the minimalinhibitory concentration at the site of infection. This wasclearly demonstrated in the case of Listeria monocytogenes,which causes cytosolic infections (addition of an inhibitorof the fluoroquinolone efflux pump markedly increases theactivity of these antibiotics [96]). The situation may bemore complex for infections affecting the vacuolar appara-tus, because efflux pumps in this case could promote theaccumulation of antibiotics from the cytosol into thesevacuoles, since their membranes partly derive from invagi-nations of the pericellular membrane [2, 99].

CHEMOTHERAPEUTIC PERSPECTIVES

The search for new anti-infective agents now must takeinto account the growing problem of resistance. In thiscontext, glycylcyclines and ketolides were designed and/orscreened specifically for action against strains displaying theantibiotic efflux pumps recognizing their parent compounds(tetracyclines and macrolides ([100, 101]; see also abstracts2133, 2137, and 2140 in the 39th ICAAC). The failuresencountered with antibiotics designed to specifically resistinactivating enzymes (b-lactamases, aminoglycoside-modi-fying enzymes, and so forth) show, however, that chemical‘improvements’ are likely to be overcome quickly by bac-teria. Obtaining specific and potent inhibitors of antibiotictransporters, therefore, appears today as an important ob-jective in anti-infective chemotherapy. Although basicknowledge is still scanty in many areas, several lines ofresearch can be followed safely using either ligand-based ortarget-based design approaches. Ligand-based approacheshave long been the main source of new chemotherapeuticagents and, therefore, could be followed usefully in thiscase. Yet, they may fall short of an accurate definition of astarting pharmacophore. Indeed, we have seen that there isactually little stringent structural requirement for a drug tobe transported. Target-based approaches, on their side, mayoffer more opportunities for defining a specific ligand,especially since we now have a growing capacity to under-take precise molecular modeling allowing one to defineligands. In this context, it is important to emphasize thateffective inhibitors do not necessarily have to be directedagainst the binding site of the natural substrate. Forexample, effective inhibitors acting on proteins of thepicornavirus capsid have been designed to bind to func-

tional groups situated out of the site of interaction of theprotein with its target [102]. Moreover, for members of theABC transporter superfamily, it also appears possible toinhibit the ATPase activity of the transporter rather thanits efflux capacity [44, 103]. Several groups of both aca-demic and industry-based researchers are heavily engagedin the search for pump inhibitors [104–107]. A majordifficulty, however, may arise from the fact that antibiotic-extruding pumps could be proteins with important physio-logical functions, the manipulation of which may causeunexpected toxicities. In this context, efforts directed tospecifically inhibit antibiotic-extruding pumps operatingonly in procaryotes may offer significantly greater chancesof effective therapeutic success ([108]; promising com-pounds have also been presented in this respect at the 39thICAAC [viz. abstracts 1264–1272]). An indirect therapeu-tic application of our present knowledge of the antibiotic-extruding pumps could be the use of antibiotic moleculesthemselves to inhibit the transport of other chemothera-peutic agents such as anticancer drugs [109–113]. Therationale of this approach is that we already possess effec-tive pharmacophores with low levels of toxicity that arebacked by long clinical experience [114]. Although thisapproach may seem attractive at first glance, it will, in ouropinion, quickly stumble on the problem of the overuse ofantibiotics, which is one of the major causes of worldwideantibacterial resistance. This non-antibiotic use of antibi-otics, therefore, may create an uncontrollable problem, notonly at the level of the patient but also at that of thecommunity, as clearly exemplified by the use of antibioticsas food additives [115]. We therefore suggest that there isnot only room but also a necessity to design truly effectiveand finely tuned specific inhibitors of antibiotic effluxtransporters, which will usefully complement our presentanti-infective armamentarium. This could be achieved byconcerted genomic and proteomic studies targeted towardsthe discovery of specific genes and gene products, withcomplete biochemical and functional characterization inpurified systems.

We thank Dr. O. Lomovskaya (Microcide Pharmaceuticals, Inc.,Mountain View, CA), Prof. M. H. Saier (Department of Biology,University of California at San Diego, La Jolla, CA), Prof. A.Goffeau (Unite de Biochimie Physiologique, Universite Catholique deLouvain, Louvain-La-Neuve, Belgium), Prof. Y. Glupczynski (Cli-niques Universitaires de l’Universite Catholique de Louvain a MontGodinne, Belgium), and Dr. M. P. Mingeot-Leclercq (Unite dePharmacologie Cellulaire et Moleculaire, Universite Catholique deLouvain, Bruxelles, Belgium) for critical reading of our manuscript andvaluable suggestions. F. V. B is Charge de Recherches of the BelgianFonds National de la Recherche Scientifique. This work was supportedby the Fonds Special de Recherches (FSR) of the Universite Catholiquede Louvain, the Belgian Fonds de la Recherche Scientifique Medicale(Grant 3.4516.94), the Belgian Fonds National de la RechercheScientifique, and the Belgian Service de la Politique Scientifique ‘Polesd’Attractions Interuniversitaires’.

466 F. Van Bambeke et al.

Page 11: COMMENTARY Antibiotic Efflux Pumps · the drug efflux pumps in eucaryotic cells ( [7]; drug efflux transporters are classically energized by ATP). The second-ary active transporters,

References

1. Kolaczkowski M and Goffeau A, Active efflux by multidrugtransporters as one of the strategies to evade chemotherapyand novel practical implications of yeast pleiotropic drugresistance. Pharmacol Ther 76: 219–242, 1997.

2. Ishikawa T, Li ZS, Lu YP and Rea PA, The GS-X pump inplant, yeast, and animal cells: Structure, function, and geneexpression. Biosci Rep 17: 189–207, 1997.

3. Wacher VJ, Silverman JA, Zhang Y and Benet LZ, Role ofP-glycoprotein and cytochrome P450 3A in limiting oralabsorption of peptides and peptidomimetics. J Pharm Sci 87:1322–1330, 1998.

4. Hall SD, Thummel KE, Watkins PB, Lown KS, Benet LZ,Paine MF and Wrighton SA, Molecular and physical mech-anisms of first-pass extraction. Drug Metab Dispos 27: 161–166, 1999.

5. Suzuki H and Sugiyama Y, Excretion of GSSG and gluta-thione conjugates mediated by MRP1 and cMOAT/MRP2.Semin Liver Dis 18: 359–376, 1998.

6. Paulsen IT, Sliwinski MK and Saier MH Jr, Microbialgenome analyses: Global comparisons of transport capabili-ties based on phylogenies, bioenergetics and substrate spec-ificities. J Mol Biol 277: 573–592, 1998.

7. Paulsen IT, Sliwinski MK, Nelissen B, Goffeau A and SaierMH Jr, Unified inventory of established and putative trans-porters encoded within the complete genome of Saccharo-myces cerevisiae. FEBS Lett 430: 116–125, 1998.

8. Mendez C and Salas JA, ABC transporters in antibiotic-producing actinomycetes. FEMS Microbiol Lett 158: 1–8,1998.

9. Paulsen IT, Skurray RA, Tam R, Saier MH, Turner RJ,Weiner JH and Grinius LL, The SMR family: A novel familyof multidrug efflux proteins involved with the efflux oflipophilic drugs. Mol Microbiol 19: 1167–1175, 1996.

10. Hogue DL, Ellison MJ, Young JD and Cass CE, Identifica-tion of a novel membrane transporter associated with intra-cellular membranes by phenotypic complementation in theyeast Saccharomyces cerevisiae. J Biol Chem 271: 9801–9808,1996.

11. Hogue DL, Kerby L and Ling V, A mammalian lysosomalmembrane protein confers multidrug resistance upon expres-sion in Saccharomyces cerevisiae. J Biol Chem 274: 12877–12882, 1999.

12. Tseng TT, Gratwick KS, Kolman J, Park D, Nies DH,Goffeau A and Saier MH, The RND permease superfamily:An ancient, ubiquitous and diverse family that includeshuman disease and development proteins. J Mol MicrobiolBiotechnol 1: 107–125, 1999.

13. Saier MH, Paulsen IT, Sliwinski MK, Pao SS, Skurray RAand Nikaido H, Evolutionary origins of multidrug anddrug-specific efflux pumps in bacteria. FASEB J 12: 265–274, 1998.

14. Paulsen IT, Brown MH and Skurray RA, Proton-dependentmultidrug efflux systems. Microbiol Rev 60: 575–608, 1996.

15. Marger MD and Saier MH, A major superfamily of mem-brane facilitators that catalyse uniport, symport and antiport.Trends Biochem Sci 18: 13–20, 1993.

16. Pao SS, Paulsen IT and Saier MH, Major facilitator super-family. Microbiol Mol Biol Rev 62: 1–34, 1998.

17. Morita Y, Kodama K, Shiota S, Mine T, Kataoka A,Mizushima T and Tsuchiya T, NorM, a putative multidrugefflux protein of Vibrio parahaemolyticus and its homolog inEscherichia coli. Antimicrob Agents Chemother 42: 1778–1782, 1998.

18. Higgins CF, ABC transporters: From microorganisms toman. Annu Rev Cell Biol 8: 67–113, 1992.

19. Decottignies A and Goffeau A, Complete inventory of theyeast ABC proteins. Nat Genet 15: 137–145, 1997.

20. Croop JM, Evolutionary relationships among ABC trans-porters. Methods Enzymol 292: 101–116, 1998.

21. Taglicht D and Michaelis S, Saccharomyces cerevisiae ABCproteins and their relevance to human health and disease.Methods Enzymol 292: 130–162, 1998.

22. van Veen H and Konings WN, The ABC family of multi-drug transporters in microorganisms. Biochim Biophys Acta1365: 31–36, 1998.

23. Saurin W, Hofnung M and Dassa E, Getting in or out: Earlysegregation between importers and exporters in the evolu-tion of ATP-binding cassette (ABC) transporters. J Mol Evol48: 22–41, 1999.

24. Ruetz S, Brault M, Kast C, Hemenway C, Heitman J, GrantCE, Cole SPC, Deeley RG and Gros P, Functional expres-sion of the multidrug resistance-associated protein in theyeast Saccharomyces cerevisiae. J Biol Chem 271: 4154–4160,1996.

25. van Veen HW, Callaghan R, Soceneantu L, Sardini A,Konings WN and Higgins CF, A bacterial antibiotic-resis-tance gene that complements the human multidrug-resis-tance P-glycoprotein gene. Nature 391: 291–295, 1998.

26. Cole SPC and Deeley RG, Multidrug resistance mediated bythe ATP-binding cassette transporter protein MRP. Bioes-says 20: 931–940, 1998.

27. Paul S, Breuninger LM, Tew KD, Shen H and Kruh GD,ATP-dependent uptake of natural product cytotoxic drugsby membrane vesicles establishes MRP as a broad specificitytransporter. Proc Natl Acad Sci USA 93: 6929–6934, 1996.

28. Kolaczkowski M, Kolaczkowska A, Luczynski J, Stanislaw Wand Goffeau A, In vivo characterization of the drug resis-tance profile of the major ABC transporters and othercomponents of the yeast pleiotropic drug resistance network.Microb Drug Resist 4: 143–158, 1998.

29. Paulsen IT and Skurray RA, Topology, structure and evolu-tion of two families of proteins involved in antibiotic andantiseptic resistance in eukaryotes and prokaryotes—ananalysis. Gene 124: 1–11, 1993.

30. Maloney PC, Bacterial and plant antiporters. J Exp Biol 196:439–442, 1994.

31. Mordoch SS, Granot D, Lebendiker M and Schuldiner S,Scanning cysteine accessibility of EmrE, an H1-coupledmultidrug transporter from Escherichia coli, reveals a hydro-phobic pathway for solutes. J Biol Chem 274: 19480–19486,1999.

32. Grinius LL and Goldberg EB, Bacterial multidrug resistanceis due to a single membrane protein which functions as adrug pump. J Biol Chem 269: 29998–30004, 1994.

33. Edgar R and Bibi E, A single membrane-embedded negativecharge is critical for recognizing positively charged drugs bythe Escherichia coli multidrug resistance protein MdfA.EMBO J 18: 822–832, 1999.

34. Guan L, Ehrmann M, Yoneyama H and Nakae T, Membranetopology of the xenobiotic-exporting subunit, MexB, of theMexA,B-OprM extrusion pump in Pseudomonas aeruginosa.J Biol Chem 274: 10517–10522, 1999.

35. Paulsen IT, Park JH, Choi PS and Saier MH, A family ofGram-negative bacterial outer membrane factors that func-tion in the export of proteins, carbohydrates, drugs andheavy metals from Gram-negative bacteria. FEMS MicrobiolLett 156: 1–8, 1997.

36. Sonveaux N, Shapiro AB, Goormaghtigh E, Ling V andRuysschaert JM, Secondary and tertiary structure changes ofreconstituted P-glycoprotein. A Fourier transform attenu-ated total reflection infrared spectroscopy analysis. J BiolChem 271: 24617–24624, 1996.

Antibiotic Efflux Pumps 467

Page 12: COMMENTARY Antibiotic Efflux Pumps · the drug efflux pumps in eucaryotic cells ( [7]; drug efflux transporters are classically energized by ATP). The second-ary active transporters,

37. Sharom FJ, The P-glycoprotein efflux pump: How does ittransport drugs? J Membr Biol 160: 161–175, 1997.

38. Mitchell BA, Paulsen IT, Brown MH and Skurray RA,Bioenergetics of the staphylococcal multidrug export proteinQacA. Identification of distinct binding sites for monova-lent and divalent cations. J Biol Chem 274: 3541–3548,1999.

39. Nikaido H, Multidrug efflux pumps of Gram-negative bac-teria. J Bacteriol 178: 5853–5859, 1996.

40. Nikaido H, Multiple antibiotic resistance and efflux. CurrOpin Microbiol 1: 516–523, 1998.

41. Higgins CF and Gottesman MM, Is the multidrug trans-porter a flippase? Trends Biochem Sci 17: 18–21, 1992.

42. Wadkins RM and Houghton PJ, The role of drug-lipidinteractions in the biological activity of modulators ofmulti-drug resistance. Biochim Biophys Acta 1153: 225–236,1993.

43. Bolhuis H, van Veen HW, Molenaar D, Poolman B, Dries-sen AJM and Konings WN, Multidrug resistance in Lacto-coccus lactis: Evidence for ATP-dependent drug extrusionfrom the inner leaflet of the cytoplasmic membrane. EMBOJ 15: 4239–4245, 1996.

44. Ashida H, Oonishi T and Uyesaka N, Kinetic analysis of themechanism of action of the multidrug transporter. J TheorBiol 195: 219–232, 1998.

45. Rosenberg MF, Callaghan R, Ford RC and Higgins CF,Structure of the multidrug resistance P-glycoprotein to 2.5nm resolution determined by electron microscopy and imageanalysis. J Biol Chem 272: 10685–10694, 1997.

46. Pawagi AB, Wang J, Silverman M, Reithmeier RA andDeber CM, Transmembrane aromatic amino acid distribu-tion in P-glycoprotein. A functional role in broad substratespecificity. J Mol Biol 235: 554–564, 1994.

47. Smith AJ, Timmermans-Hereijgers JL, Roelofsen B, WirtzKW, van Blitterswijk WJ, Smit JJ, Schinkel AH and Borst P,The human MDR3 P-glycoprotein promotes translocation ofphosphatidylcholine through the plasma membrane of fibro-blasts from transgenic mice. FEBS Lett 354: 263–266, 1994.

48. Zhou Y, Gottesman MM and Pastan I, Studies with humanMDR1-MDR2 chimeras demonstrate the functional ex-changeability of a major transmembrane segment of themultidrug transporter of phosphatidylcholine flippase. MolCell Biol 19: 1450–1459, 1999.

49. Sokal A, Pulaski L, Rychlik B, Fortuniak A and Bartosz G,Is the glutathione S-conjugate pump a flippase? Biochem MolBiol Int 44: 97–105, 1998.

50. Decottignies A, Grant AM, Nichols JW, de Wet H, McIn-tosh DB and Goffeau A, ATPase and multidrug transportactivities of the overexpressed yeast ABC protein Yor1p.J Biol Chem 273: 12612–12622, 1998.

51. Hsieh PC, Siegel SA, Rogers B, Davis D and Lewis K,Bacteria lacking a multidrug pump: A sensitive tool for drugdiscovery. Proc Natl Acad Sci USA 95: 6602–6606, 1998.

52. Nikaido H, Basina M, Nguyen V and Rosenberg EY, Mul-tidrug efflux pump AcrAB of Salmonella typhimurium excretesonly those b-lactam antibiotics containing lipophilic sidechains. J Bacteriol 180: 4686–4692, 1998.

53. Nilsen IW, Bakke I, Vader A, Olsvik O and El-Gewely MR,Isolation of cmr, a novel Escherichia coli chloramphenicolresistance gene encoding a putative efflux pump. J Bacteriol178: 3188–3193, 1996.

54. Aires JR, Kohler T, Nikaido H and Plesiat P, Involvement ofan active efflux system in the natural resistance of Pseudo-monas aeruginosa to aminoglycosides. Antimicrob Agents Che-mother 43: 2624–2628, 1999.

55. Moore RA, DeShazer D, Reckseidler S, Weissman A andWoods DE, Efflux-mediated aminoglycoside and macrolide

resistance in Burkholderia pseudomallei. Antimicrob AgentsChemother 43: 465–470, 1999.

56. Mingeot-Leclercq MP, Glupczynski Y and Tulkens PM,Aminoglycosides: Activity and resistance. Antimicrob AgentsChemother 43: 727–737, 1999.

57. Mingeot-Leclercq MP and Tulkens PM, Aminoglycosides:Nephrotoxicity. Antimicrob Agents Chemother 43: 1003–1012, 1999.

58. Cundliffe E, Self-protection mechanisms in antibiotic pro-ducers. Ciba Found Symp 171: 208–214, 1992.

59. Hancock REW and Farmer SW, Mechanism of uptake ofdeglucoteicoplanin amide derivatives across outer mem-branes of Escherichia coli and Pseudomonas aeruginosa. Anti-microb Agents Chemother 37: 453–456, 1993.

60. Nakae T, Yoshihara E and Yoneyama H, Multiantibioticresistance caused by active drug extrusion in hospital patho-gens. J Infect Chemother 3: 173–183, 1997.

61. Ball PR, Shales SW and Chopra I, Plasmid-mediated tetra-cycline resistance in Escherichia coli involves increased effluxof the antibiotic. Biochem Biophys Res Commun 93: 74–81,1980.

62. McMurry L, Petrucci RE Jr and Levy SB, Active efflux oftetracycline encoded by four genetically different tetracy-cline resistance determinants in Escherichia coli. Proc NatlAcad Sci USA 77: 3974–3977, 1980.

63. Clancy J, Petitpas J, Dib-Hajj F, Yuan W, Cronan M,Kamath AV, Bergeron J and Retsema JA, Molecular cloningand functional analysis of a novel macrolide-resistancedeterminant, mefA, from Streptococcus pyogenes. Mol Micro-biol 22: 867–879, 1996.

64. Ubukata K, Itoh-Yamashita N and Konno M, Cloning andexpression of the norA gene for fluoroquinolone resistance inStaphylococcus aureus. Antimicrob Agents Chemother 33:1535–1539, 1989.

65. Sutcliffe J, Tait-Kamradt A and Wondrack L, Streptococcuspneumoniae and Streptococcus pyogenes resistant to macrolidesbut sensitive to clindamycin: A common resistance patternmediated by an efflux system. Antimicrob Agents Chemother40: 1817–1824, 1996.

66. Brenwald NP, Gill MJ and Wise R, Prevalence of a putativeefflux mechanism among fluoroquinolone-resistant clinicalisolates of Streptococcus pneumoniae. Antimicrob Agents Che-mother 42: 2032–2035, 1998.

67. Kaatz GW, Seo SM and Ruble CA, Efflux-mediated fluoro-quinolone resistance in Staphylococcus aureus. AntimicrobAgents Chemother 37: 1086–1094, 1993.

68. Li X-Z, Nikaido H and Poole K, Role of MexA-MexB-OprMin antibiotic efflux in Pseudomonas aeruginosa. AntimicrobAgents Chemother 39: 1948–1953, 1995.

69. Masuda N, Gotoh N, Ishii C, Sakagawa E, Ohya S andNishino T, Interplay between chromosomal b-lactamase andthe MexAB-oprM efflux system in intrinsic resistance tob-lactams in Pseudomonas aeruginosa. Antimicrob AgentsChemother 43: 400–402, 1999.

70. Lakaye B, Dubus A, Lepage S, Groslambert S and Frere JM,When drug inactivation renders the target irrelevant toantibiotic resistance: A case story with b-lactams. MolMicrobiol 31: 89–101, 1999.

71. Nakae T, Nakajima A, Ono T, Saito K and Yoneyama H,Resistance to b-lactam antibiotics in Pseudomonas aeruginosadue to interplay between the MexAB-OprM efflux pumpand b-lactamase. Antimicrob Agents Chemother 43: 1301–1303, 1999.

72. Janoir C, Zeller V, Kitzis M-D, Moreau NJ and Gutmann L,High-level fluoroquinolone resistance in Streptococcus pneu-moniae requires mutations in parC and gyrA. AntimicrobAgents Chemother 40: 2760–2764, 1996.

73. Jalal S and Wretlind B, Mechanims of quinolone resistance

468 F. Van Bambeke et al.

Page 13: COMMENTARY Antibiotic Efflux Pumps · the drug efflux pumps in eucaryotic cells ( [7]; drug efflux transporters are classically energized by ATP). The second-ary active transporters,

in clinical strains of Pseudomonas aeruginosa. Microb DrugResist 4: 257–261, 1998.

74. Banerjee SK, Bhatt K, Rana S, Misra P and Chakraborti PK,Involvement of an efflux system in mediating high level offluoroquinolone resistance in Mycobacterium smegmatis. Bio-chem Biophys Res Commun 226: 362–368, 1996.

75. Tanaka M, Sakuma S, Takahashi K, Nagahuzi T, Saika T,Kobayashi I and Kumazawa J, Analysis of quinolone resis-tance mechanisms in Neisseria gonorrhoeae isolates in vitro.Sex Transm Infect 74: 59–62, 1998.

76. Lomovskaya O, Lee A, Hoshino K, Ishida H, Mistry A,Warren MS, Boyer E, Chamberland S and Lee VJ, Use of agenetic approach to evaluate the consequences of inhibitionof efflux pumps in Pseudomonas aeruginosa. Antimicrob AgentsChemother 43: 1340–1346, 1999.

77. Li X-Z, Ma D, Livermore DM and Nikaido H, Role of effluxpump(s) in intrinsic resistance of Pseudomonas aeruginosa:Active efflux as a contributing factor to b-lactam resistance.Antimicrob Agents Chemother 38: 1742–1752, 1994.

78. Nikaido H, Antibiotic resistance caused by Gram-negativemultidrug efflux pumps. Clin Infect Dis 27: S32–S41, 1998.

79. Balzi E and Goffeau A, Yeast multidrug resistance: The PDRnetwork. J Bioenerg Biomembr 27: 71–76, 1995.

80. Sanglard D, Kuchler K, Ischer F, Pagani J-L, Monod M andBille J, Mechanisms of resistance to azole antifungal agentsin Candida albicans isolates from AIDS patients involvespecific multidrug transporters. Antimicrob Agents Chemother39: 2378–2386, 1995.

81. White TC, Marr KA and Bowden R, Clinical, cellular andmolecular factors that contribute to antifungal drug resis-tance. Clin Microbiol Rev 11: 382–402, 1998.

82. Ziha-Zarifi I, Llanes C, Kohler T, Pechere J-C and Plesiat P,In vivo emergence of multidrug-resistant mutants of Pseudo-monas aeruginosa overexpressing the active efflux systemMexA-MexB-OprM. Antimicrob Agents Chemother 43: 287–291, 1999.

83. Limia A, Jimenez ML, Delgado T, Sanchez I, Lopez S andLopez-Brea M, Phenotypic characterization of erythromycinresistance in strains of the genus Streptococcus isolated fromclinical specimens. Rev Esp Quimiter 11: 216–220, 1998.

84. Orden B, Perez-Trallero E, Montes M and Martinez R,Erythromycin resistance of Streptococcus pyogenes in Madrid.Pediatr Infect Dis J 17: 470–473, 1998.

85. Shortridge VD, Doern GV, Brueggemann AB, Beyer JM andFlamm RK, Prevalence of macrolide resistance mechanismsin Streptococcus pneumoniae isolates from a multicenterantibiotic resistance surveillance study conducted in theUnited States in 1994–1995. Clin Infect Dis 29: 1186–1188,1999.

86. Stieger B and Meier PJ, Bile acid and xenobiotic transportersin liver. Curr Opin Cell Biol 10: 462–467, 1998.

87. Roch-Ramel F, Renal transport of organic anions. Curr OpinNephrol Hypertens 7: 517–524, 1998.

88. Saitoh H, Fujisaki H, Aungst BJ and Miyazaki K, Restrictedintestinal absorption of some b-lactam antibiotics by anenergy-dependent efflux system in rat intestine. Pharm Res14: 645–649, 1997.

89. Cavet ME, West M and Simmons NL, Transepithelialtransport of the fluoroquinolone ciprofloxacin by humanairway epithelial Calu-3 cells. Antimicrob Agents Chemother41: 2693–2698, 1997.

90. Dautrey S, Felice K, Petiet A, Lacour B, Carbon C andFarinotti R, Active intestinal elimination of ciprofloxacin inrats: Modulation by different substrates. J Pharmacol 127:1728–1734, 1999.

91. Saitoh H, Gerard C and Aungst BJ, The secretory intestinaltransport of some beta-lactam antibiotics and anionic com-

pounds: A mechanism contributing to poor oral absorption.J Pharmacol Exp Ther 278: 205–211, 1996.

92. Snyder NJ, Tabas LB, Berry DM, Duckworth DC, Spry DOand Dantzig AH, Structure-activity relationship of carba-cephalosporins and cephalosporins: Antibacterial activityand interaction with the intestinal proton-dependent dipep-tide transport carrier of Caco-2 cells. Antimicrob AgentsChemother 41: 1649–1657, 1997.

93. Race JE, Grassl SM, Williams WJ and Holtzman EJ, Molec-ular cloning and characterization of two novel human renalorganic anion transporters (hOAT1 and hOAT3). BiochemBiophys Res Commun 255: 508–514, 1999.

94. Sasabe H, Terasaki T, Tsuji A and Sugiyama Y, Carrier-mediated hepatic uptake of quinolone antibiotics in the rat.J Pharmacol Exp Ther 282: 162–171, 1997.

95. Cao CX, Silverstein SC, Neu HC and Steinberg TH, J774macrophages secrete antibiotics via organic anion transport-ers. J Infect Dis 165: 322–328, 1992.

96. Rudin DE, Gao PX, Cao CX, Neu HC and Silverstein SC,Gemfibrozil enhances the listericidal effects of fluoroquino-lone antibiotics in J774 macrophages. J Exp Med 176:1439–1447, 1992.

97. Nichterlein T, Kretschmar M, Siegsmund M and Hof H,Erythromycin is ineffective against Listeria monocytogenes inmultidrug resistant cells. J Chemother 7: 184–188, 1995.

98. Nichterlein T, Kretschmar M, Schadt A, Meyer A, Wild-feuer A, Laufen H and Hof H, Reduced intracellular activityof antibiotics against Listeria monocytogenes in multidrugresistant cells. Int J Antimicrob Agents 10: 119–125, 1998.

99. Oh YK and Straubinger RM, Cellular delivery of liposome-delivered anionic compounds modulated by a probenecid-sensitive anion transporter. Pharm Res 14: 1203–1209, 1997.

100. Testa RT, Petersen PJ, Jacobus NV, Sum PE, Lee VJ andTally FP, In vitro and in vivo antibacterial activities of theglycylcyclines, a new class of semisynthetic tetracyclines.Antimicrob Agents Chemother 37: 2270–2277, 1993.

101. Chu DT, Recent developments in macrolides and ketolides.Curr Opin Microbiol 2: 467–474, 1999.

102. McCarthy J, Hogle JM and Karplus M, Use of the multiplecopy simultaneous search (MCSS) method to design a newclass of picornavirus capsid binding drugs. Proteins 29:32–58, 1997.

103. Litman T, Zeuthen T, Skovsgaard T and Stein WD, Struc-ture-activity relationships of P-glycoprotein interactingdrugs: Kinetic characterization of their effects on ATPaseactivity. Biochim Biophys Acta 1361: 159–168, 1997.

104. Ford JM and Hait WN, Pharmacologic circumvention ofmultidrug resistance. Cytotechnology 12: 171–212, 1993.

105. Tmej C, Chiba P, Huber M, Richter E, Hitzler M, SchaperKJ and Ecker G, A combined Hansch/Free-Wilson approachas predictive tool in QSAR studies on propafenone-typemodulators of multidrug resistance. Arch Pharm (Weinheim)331: 233–240, 1998.

106. Dinh TQ, Smith CD, Du X and Armstrong RW, Design,synthesis, and evaluation of the multidrug resistance-revers-ing activity of D-glucose mimetics of hapalosin. J Med Chem41: 981–987, 1998.

107. Ecker G, Huber M, Schmid D and Chiba P, The importanceof a nitrogen atom in modulators of multidrug resistance.Mol Pharmacol 56: 791–796, 1999.

108. Markham PN, Westhaus E, Klyachko K, Johnson ME andNeyfakh AA, Multiple novel inhibitors of the NorA multi-drug transporter of Staphylococcus aureus. Antimicrob AgentsChemother 43: 2404–2408, 1999.

109. Gosland MP, Lum BL and Sikic BI, Reversal by cefoperazoneof resistance to etoposide, doxorubicin and vinblastine inmultidrug resistant human sarcoma cells. Cancer Res 49:6901–6905, 1989.

Antibiotic Efflux Pumps 469

Page 14: COMMENTARY Antibiotic Efflux Pumps · the drug efflux pumps in eucaryotic cells ( [7]; drug efflux transporters are classically energized by ATP). The second-ary active transporters,

110. Crosta L, Candiloro V, Meli M, Tolomeo M, Rausa L andDusonchet L, Lacidipine and josamycin: Two new multidrugresistance modulators. Anticancer Res 14: 2685–2690, 1994.

111. Fardel O, Lecureur V, Loyer P and Guillouzo A, Rifampicinenhances anti-cancer drug accumulation and activity inmultidrug-resistant cells. Biochem Pharmacol 49: 1255–1260,1995.

112. Phung-Ba V, Warnery A, Scherman D and Wils P, Interac-tion of pristinamycin IA with P-glycoprotein in human intes-tinal epithelial cells. Eur J Pharmacol 288: 187–192, 1995.

113. Takano M, Hasegawa R, Fukuda T, Yumoto R, Nagai J andMurakami T, Interaction with P-glycoprotein and transportof erythromycin, midazolam and ketoconazole in Caco-2cells. Eur J Pharmacol 358: 289–294, 1998.

114. Raderer M and Scheithauer W, Clinical trials of agents thatreverse multidrug resistance. Cancer 72: 3553–3563, 1993.

115. Wegener HC, Aarestrup FM, Jensen LB, Hammerum AMand Bager F, Use of antimicrobial growth promoters in foodanimals and Enterococcus faecium resistance to therapeuticantimicrobial drugs in Europe. Emerg Infect Dis 5: 329–335,1999.

116. Belinsky MG, Bain LJ, Testa JR and Kruh GD, Character-ization of MOAT-C and MOAT-D, new members of theMRP/cMOAT subfamily of transporter proteins. J NatlCancer Inst 90: 1735–1741, 1998.

117. Georges E, Tsuruo T and Ling V, Topology of P-glycoproteinas determined by epitope mapping of MRK-16 monoclonalantibodies. J Biol Chem 268: 1792–1798, 1993.

118. Bakos E, Hegedus T, Hollo Z, Welker E, Tusnady G, ZamanGJR and Sarkadi B, Membrane topology and glycosylation ofthe human multidrug resistance-associated protein. J BiolChem 271: 12322–12326, 1996.

470 F. Van Bambeke et al.