ArticlePDF Available

Abstract and Figures

Secondary multidrug transporters use ion concentration gradients to energize the removal from cells of various antibiotics. The Escherichia coli multidrug transporter MdfA exchanges a single proton with a single monovalent cationic drug molecule. This stoichiometry renders the efflux of divalent cationic drugs energetically unfavourable, as it requires exchange with at least two protons. Here we show that surprisingly, MdfA catalyses efflux of divalent cations, provided that they have a unique architecture: the two charged moieties must be separated by a long linker. These drugs are exchanged for two protons despite the apparent inability of MdfA to exchange two protons for a single drug molecule. Our results suggest that these drugs are transported in two consecutive transport cycles, where each cationic moiety is transported as if it were a separate substrate. We propose that secondary transport can adopt a processive-like mode of action, thus expanding the substrate spectrum of multidrug transporters.
Content may be subject to copyright.
ARTICLE
Received 20 May 2014 |Accepted 7 Jul 2014 |Published 8 Aug 2014
Export of a single drug molecule in two transport
cycles by a multidrug efflux pump
Nir Fluman1,w, Julia Adler1,w, Susan A. Rotenberg2, Melissa H. Brown3& Eitan Bibi1
Secondary multidrug transporters use ion concentration gradients to energize the removal
from cells of various antibiotics. The Escherichia coli multidrug transporter MdfA exchanges a
single proton with a single monovalent cationic drug molecule. This stoichiometry renders the
efflux of divalent cationic drugs energetically unfavourable, as it requires exchange with at
least two protons. Here we show that surprisingly, MdfA catalyses efflux of divalent cations,
provided that they have a unique architecture: the two charged moieties must be separated
by a long linker. These drugs are exchanged for two protons despite the apparent inability of
MdfA to exchange two protons for a single drug molecule. Our results suggest that these
drugs are transported in two consecutive transport cycles, where each cationic moiety is
transported as if it were a separate substrate. We propose that secondary transport can adopt
a processive-like mode of action, thus expanding the substrate spectrum of multidrug
transporters.
DOI: 10.1038/ncomms5615
1Department of Biological Chemistry, Weizmann Institute of Science, Rehovot 76100, Israel. 2Department of Chemistry and Biochemistry, Queens College—
City University of New York, Flushing, New York 11367, USA. 3School of Biological Sciences, Flinders University, Adelaide South Australia 5001, Australia.
wPresent address: Department of Molecular Genetics, Weizmann Institute of Science, Rehovot 76100, Israel. Correspondence and requests for materials
should be addressed to E.B. (email: e.bibi@weizmann.ac.il).
NATURE COMMUNICATIONS | 5:4615 | DOI: 10.1038/ncomms5615 | www.nature.com/naturecommunications 1
&2014 Macmillan Publishers Limited. All rights reserved.
Multidrug resistance (Mdr) transporters are ubiquitous
membrane proteins that provide cells with defense
against various cytotoxic compounds. They function as
multispecific efflux pumps, each able to expel a broad spectrum of
chemically dissimilar drugs out of the cell1. Bacteria typically
harbour numerous Mdr transporter-encoding genes in their
genomes, which function as secondary transporters driven by the
proton motive force2. Bacterial Mdr transporters exist in several
phylogenetically and mechanistically distinct protein families3.Of
those, Hþ/multidrug antiporters of the Major Facilitator
Superfamily (MFS) comprise the largest group1.
Different MFS Mdr transporters work with distinct proton
antiport stoichiometries. Although a stoichiometry of one proton
per drug was observed with MdfA from E. coli4,5, QacA from
Staphylococcus aureus and LmrP from Lactococcus lactis use
stoichiometries of up to 2 and 3, respectively6,7. Recent progress
has shed light on the mechanism of proton transport by MdfA, in
which a membrane-embedded acidic residue facilitates proton
binding and transport4. Mutagenesis experiments indicate that
increasing the proton stoichiometry of MdfA from 1 to 2 is
possible by introducing an additional acidic residue in a number
of membrane-embedded locations8. LmrP, which works with a
stoichiometry of up to three protons, contains three membrane-
embedded acidic residues; each functions as a module that
facilitates the transport of a single proton7.
As shown recently, the stoichiometry of one proton per
drug renders MdfA inactive in efflux of divalent cationic
(dicationic) drugs, which require import of two protons per
transport cycle8. The underlying reason is that such an antiport
reaction (export of drug2þin exchange for a single H þ) involves
net charge efflux that opposes the membrane potential (Dc,
interior negative). Indeed, MdfA mutants that are capable of
pumping out two protons per drug are able to confer resistance
against dicationic drugs8. Surprisingly, we show here that also
wild-type MdfA can actively pump certain dicationic drugs.
However, in this case the dicationic drugs must have a distinct
chemical architecture, in which the two cationic moeities are
separated by a long linker. Our studies suggest that each of
these drugs is transported in two successive transport cycles,
where the cationic moieties are treated as if they were separate
monocationic substrates. This unexpected mechanism expands
the Mdr spectrum of MdfA towards a new group of dicationic
drugs and indicates that secondary transporters can operate in a
processive-like manner.
Results
MdfA confers resistance against unique dicationic drugs.
During our characterization of the drug recognition spectrum of
MdfA, we were surprised to find that it confers resistance against
the toxic dicationic compound dequalinium (Dq). The chemical
structure of Dq reveals that it contains two identical cationic
pharmacophores separated by a long, ten-carbon chain linker
(Fig. 1a). To examine whether these characteristics of Dq (sym-
metry and spacer length) are relevant to the unexpected activity,
we challenged MdfA with several dicationic drugs (Fig. 1a). In
some of these compounds, both positive charges are relatively
close (40,6-diamidino-2-phenylindole (DAPI), diminazene (Dmn)
and methyl viologen), while in others the positive charges are
separated by long linkers (Dq, chlorhexidine (Chx) and penta-
midine (Pent)). For simplicity, we term the two groups of com-
pounds SDCs (short dicationic compounds) and LDCs (long
dicationic compounds), respectively. Strikingly, when challenged
against all these drugs, we observed that MdfA could only protect
E. coli against LDCs, demonstrating that symmetry alone is not
sufficient and that the linker length is important (Fig. 1b).
The resistance conferred by MdfA suggests that unlike SDCs,
LDCs are actively extruded by the transporter. To test this, we
characterized the ability of MdfA to prevent Dq accumulation
in vivo in whole cells. The fluorescence of Dq is quenched when it
enters cells (Fig. 2a) and can be de-quenched on lysis of cells that
accumulated Dq (Fig. 2b). Figure 2a,b show that expression of
MdfA from a multicopy plasmid indeed prevented intracellular
accumulation of Dq compared with cells harbouring an empty
vector or expressing the inactive mutant MdfA(R112M)4,9.In
addition, extrusion of Dq was completely inhibited when another
substrate of MdfA, chloramphenicol (CAM), was added as a
competitor (Fig. 2c), as shown previously for another MdfA
substrate10. These results suggest that MdfA mediates active
transport of Dq.
MdfA binds both LDCs and SDCs. How does MdfA distinguish
between the two groups of dicationic drugs? One possibility is
that MdfA cannot bind SDCs. To test this, we measured the
ability of SDCs and LDCs to compete with a radiolabelled sub-
strate, the monovalent [3H]-tetraphenylphosphonium (TPP), for
binding to MdfA. The results show that with the exception of
methyl viologen, all the tested dicationic compounds are able to
outcompete [3H]-TPP-binding in a concentration-dependent
manner (Fig. 3). The concentration dependency suggests that
MdfA has similar affinities to these compounds as those reported
for other known substrates10. Thus, the reason for the
discrimination between the two types of dicationic substrates
does not stem from differences in their ability to bind the
transporter.
MdfA catalyses LDC/proton antiport. To further evaluate the
indication that MdfA actively exports LDCs, we performed
transport experiments with everted membrane vesicles. As MdfA-
catalysed active efflux relies on drug/proton antiport, we mea-
sured substrate-dependent proton transport by MdfA in vesicles
from cells expressing MdfA or harbouring an empty vector. The
proton concentration gradient (DpH) across these membranes
was followed by measuring the fluorescence of a DpH-sensitive
indicator, 9-amino-6-chloro-2-methoxyacridine (ACMA). Ener-
gization of everted membrane vesicles by D,L-lactate oxidation
generated a DpH (acidic inside) as observed by quenching of
ACMA fluorescence (Fig. 4). The generated DpH was fully dis-
sipated by adding the proton conducting uncoupler carbonyl
cyanide m-chlorophenyl hydrazone (Fig. 4). Addition of sub-
strates from the LDCs group (Dq, Pent or Chx) or the control
monocationic substrate TPP partially dissipated DpH only in
membranes expressing MdfA (Fig. 4, compare left and right
panels). Thus, MdfA indeed catalyses proton/LDC exchange. In
contrast, addition of the SDC compound Dmn did not trigger
proton translocation. Notably, this assay is not quantitative
because calculating the amount of protons transported by MdfA
in the presence of different compounds might be complicated by
other factors such as the ability of the different substrates to
distribute across the membrane via MdfA-independent routes.
Consequently, as expected, the magnitude to which different
substrates affect DpH (as reflected by the fluorescence change)
did not show any correlation with the charges on the substrates.
MdfA-catalysed transport of LDCs is electroneutral.As
explained earlier, it is energetically unfavourable for any anti-
porter to export dicationic substrates by using a proton/substrate
stoichiometry of 1, as the resulting reaction would involve net
export of a positive charge against the membrane potential.
Therefore, as wild-type MdfA usually works with a stoichiometry
of 1, the transport of LDCs is enigmatic. To better understand
ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms5615
2NATURE COMMUNICATIONS | 5:4615 | DOI: 10.1038/ncomms5615 | www.nature.com/naturecommunications
&2014 Macmillan Publishers Limited. All rights reserved.
b
Dq fluorescence (a.u.)
0
100
200
300
Vec MdfA MdfA
(R112M)
0 min
14 min
a
0
50
100
0510
Dq fluorescence (%)
Time (min)
R112M
Vec
wt c
Vec
wt
0
50
100
0510
Time (min)
Dq fluorescence (%)
Figure 2 | MdfA-mediated Dq transport. (a) Characterization of Dq accumulation by E. coli UTL2 mdfAHkan overexpressing MdfA or its inactive mutant
MdfA(R112M), or cells harbouring an empty plasmid. (b) Quantitative Dq accumulation. Cells were suspended in buffer containing glucose and Dq (1mM),
and allowed to accumulate Dq for 0 or 14 min. Dq fluorescence in these cells was determined after the cells were harvested and lysed, relieving their
fluorescence quenching effect. Error bars indicate range of duplicate measurements. (c) Dq accumulation by control or MdfA-expressing cells in the
presence of 0.8 mM CAM. (a,c) The experiments were repeated at least three times and the results shown are representative.
a
Dq
Met-Viol
NH2
+N
N+
+NN+
Dmn
NN
DAPI
NH
Chx
NH Cl
N
H
N
H
N
H
NH
Cl
N
HN
HN
H
Pent
H2N
OO
1600 µg ml–1
TPP
200 µg ml–1
Dmn
400 µg ml–1
EtBr
4 µg ml–1
CAM
10 µg ml–1
DAPI
No drug
180 µg ml–1
Pent
400 µg ml–1
Met-Viol
1 µg ml–1
Chx
b
Vec
MdfA
Vec
MdfA
Vec
MdfA
50 µg ml–1
Dq
Vec
MdfA
NH+
2NH+
2
NH2
NH+
2
NH+
2
H2NH+
2N
NH2
NH+
2
NH2
H+
2N
NH2
NH+
2
NH2
N
H
Figure 1 | MdfA confers resistance only against a unique class of symmetric dicationic compounds (LDC). (a) Structures of the tested dicationic drugs.
Left: long dicationic compounds (LDCs): dequalinium (Dq), chlorhexidine (Chx) and pentamidine (Pent). Right: short dicationic compounds (SDCs):
40,6-diamidino-2-phenylindole (DAPI), diminazene (Dmn) and methyl viologen (Met-Viol). (b)GrowthofE. coli UTmdfAHkan expressing MdfA from a
multicopy plasmid (or harboring empty vector) on solid media supplemented with the indicated drugs. Tenfold serial dilutions of cells were plated from left to
right and grown overnight. The upper panels show controls of growth in the absence of drug or in the presence of previously known monocationic (TPP/
ethidium bromide (EtBr)) or neutral (CAM) MdfA substrates. The experiments were repeated at least three times and the results shown are representative.
NATURE COMMUNICATIONS | DOI: 10.1038/ncomms5615 ARTICLE
NATURE COMMUNICATIONS | 5:4615 | DOI: 10.1038/ncomms5615 | www.nature.com/naturecommunications 3
&2014 Macmillan Publishers Limited. All rights reserved.
this important phenomenon, we characterized the electrogenicity
of the LDCs transport reactions. We assumed that MdfA may use
a proton/dicationic substrate stoichiometry of Z2, resulting in
electroneutral or electrogenic transport, respectively, which
involves no export of a net positive charge(s).
As a first attempt to probe transport energetics in living cells,
we determined the effect of the external pH on MdfA-mediated
resistance to neutral, monocationic and LDC drugs. When the
external pH rises above the intracellular level of B7.6, the pH
gradient across the membrane (DpH) reverses its direction and
becomes acidic inside relative to outside11. Thus, electroneutral
transport reactions, which rely energetically only on DpH,
become unfavourable. In contrast, electrogenic transport
reactions are still supported by the electrical potential (Dc,
interior negative)5. Our results confirmed previous experiments
with MdfA5, which showed that resistance against the neutral
substrate CAM is driven by both components of the electro-
chemical potential (Dc and DpH), whereas resistance against the
monovalent cation ethidium bromide is solely dependent on the
DpH (Fig. 5a). Figure 5a also shows clearly that MdfA-mediated
resistance against LDCs (Dq, Chx and Pent), just like ethidium
bromide, is pH dependent, suggesting that the transport reaction
with these drugs is electroneutral. To further assess the proposed
electroneutral mode of antiport in the case of LDCs, we measured
Dq accumulation in living cells depleted of the DpH (by addition
of the ionophore nigericin and external pH 8). These conditions
make drug efflux possible only if it can be driven by Dc.
However, Fig. 5b shows that this was not the case and MdfA
could not catalyse antiport of Dq when Dc (interior negative) is
the sole driving force.
Taken together, the results indicate that the transport of the
dicationic drug Dq is electroneutral and depends solely on the
DpH. These results are intriguing, because Dq and each of the
other LDCs carry two positive charges, meaning that two protons
are exchanged for one drug molecule during the antiport.
However, as MdfA cannot translocate two protons simultaneously,
export of LDCs requires a novel mode of transport (see later).
LDC binding triggers release of a single proton from MdfA.
How does the long linker between the charges in LDCs (Fig. 1a)
allow transport with an increased proton/drug stoichiometry of 2?
Recent studies have indicated that the antiport mechanism of
MdfA relies on competition between substrate and proton on
binding to the transporter. Monocationic substrates were shown
to trigger the release of a single stoichiometric proton from MdfA,
in agreement with an antiport stoichiometry of 1 (ref. 4). Here,
one possible scenario is that the linker in LDCs enables the two
cationic moieties to trigger proton release from two distinct
proton-binding sites, if those exist in MdfA. We therefore tested
whether LDCs compete with only one or maybe two protons for
binding. To this end, we compared the amount of protons
released from MdfA on binding of the LDCs Dq and Pent with
the amount released by binding of the monocationic TPP. Proton
release was inferred from an increase in the solution’s acidity, as
0
20
40
60
80
100
120
25 µM
125 µM
Relative [3H]TPP-binding (%)
LDCsSDCs
No drug
DAPI
Dmn
Met-Viol
Chx
Dq
Pent
*
**
**
**
**
**
**
*
**
Figure 3 | MdfA binds both groups of dicationic drugs (LDCs and SDCs).
[3H]-TPP binding by wild-type MdfA and its inhibition by addition of
increasing concentrations (25 or 125 mM) of dicationic drugs. Error bars
indicate s.d. of triplicate measurements (*Po10 2;**Pr10 4compared
with the no drug sample, one tailed t-test). The experiments were
repeated at least three times and the results shown are representative.
Lactate CCCP
Substrate
Dq
MdfA
vesicles
Control
vesicles
Pent
TPP
Chx
01234
100
200
Time (min)
Dmn
300
Fluorescence (a.u.)
Figure 4 | Proton transport by control or MdfA-containing everted
membrane vesicles. The transmembrane DpH was followed by measuring
ACMA fluorescence. Lactate (added after 0.5 min, blue arrow) and
carbonyl cyanide m-chlorophenyl hydrazone (CCCP; after 3 min, black
arrow) were used to generate or abolish DpH, respectively. Substrates were
added at 2 min (orange arrow), in the following concentrations: TPP
(125 mM), Dq (25 mM), Pent (125 mM), Chx (25 mM) and Dmn (12.5 mM).
The experiments were repeated at least three times and the results shown
are representative.
ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms5615
4NATURE COMMUNICATIONS | 5:4615 | DOI: 10.1038/ncomms5615 | www.nature.com/naturecommunications
&2014 Macmillan Publishers Limited. All rights reserved.
measured by the fluorescent pH indicator fluorescein4. Notably,
Chx could not be tested in this assay because it generates high
artificial fluorescent signals (not shown). Figure 6 shows that
despite the differences in transport stoichiometry, the dicationic
drugs Dq and Pent, and the monocationic compound TPP, all
induced the release of a single stoichiometric proton from MdfA.
Thus, it appears that MdfA contains a single proton-binding site.
Dq binds to MdfA with 1:1 stoichiometry. Another purely
theoretical possibility is that LDCs bind to an MdfA dimer where
each MdfA molecule binds one of the separable cationic moieties,
thus transporting the substrate cooperatively. We do not favour
this possibility because MdfA and other MFS transporters are
known to function as monomers12,13. In addition, the several
available structures of MFS members show that the binding sites
are located deep within the proteins14–18 and therefore the spacer
separating the two cationic moieties of LDCs would not fit the
length required for binding to separate transporters in the
membrane. Indeed, we saw no evidence for cross-linking of MdfA
dimers on Dq binding, when attempting to cross-link with
formaldehyde, paraformaldehyde or disuccinimidylsuberate (data
not shown). To further examine the binding stoichiometry, we
characterized the interaction of purified MdfA protein with Dq.
Binding was measured fluorimetrically, by following Dq
fluorescence. The results show that on MdfA binding, the
fluorescence of Dq is quenched (Fig. 7a). This effect is not
observed when Dq is mixed with the inactive mutant
MdfA(R112M)4, or when TPP is added as a competitive
inhibitor (Fig. 7a). We next determined the affinity of Dq
binding to MdfA by two complementary approaches, in which
one component (either MdfA or Dq) is kept constant at a low
concentration, while the other component is titrated at increasing
concentrations. Figure 7b shows the effect of increasing [MdfA]
on Dq fluorescence and Fig. 7c shows the competitive inhibition
of TPP binding by MdfA, mediated by increasing [Dq]. The two
approaches yield comparable dissociation constants, with
K
d
B0.7 mM for MdfA at ambient temperature measured by Dq
fluorescence (Fig. 7b) and K
I
of B2mM determined from [3H]-
TPP binding-competition experiments at 4 °C (Fig. 7c). Notably,
both binding curves are hyperbolic, suggesting no cooperativity in
binding, consistent with a binding stoichiometry of 1.
To evaluate the results more directly, we conducted Dq-
mediated cross-linking of MdfA. Dq is a photoactivatable
compound that on ultraviolet irradiation is able to react
covalently with nearby residues19. The results show that Dq can
cross-link a functionally active20 split mutant of MdfA, in which
the 8 amino-terminal (N8) and the 4 carboxy-terminal (C4)
transmembrane helices were expressed as separate polypeptides
with C-terminal hexahistidine tags (Fig. 7d). This indicates that
Dq can interact with two separate regions of MdfA, at least
transiently. In contrast, Dq could not cross-link two molecules of
full-length MdfA (Fig. 7d), suggesting that interaction of Dq with
an MdfA dimer does not take place.
Export of dicationic substrates by other Mdr transporters. The
results of all the previous experiments support the notion that
MdfA translocates two protons in exchange for a single LDC
molecule, but probably not simultaneously. This led us to propose
another explanation for the increased stoichiometry during
transport of LDCs. We hypothesized that MdfA transports these
unique drug molecules in two successive transport cycles.
According to this scenario, MdfA handles each positively charged
group as if it were a separate substrate, which is transported
against a proton (see model in Fig. 8).
The suggestion that transport of a single substrate proceeds in
two consecutive cycles relies only on the substrate properties.
Therefore, we reasoned that this ability to transport LDCs might
not be unique to MdfA but perhaps shared by other MFS Mdr
transporters. To test the generality of this phenomenon, we
studied QacA, a different MFS-related Mdr transporter from
S. aureus. QacA transports both monocationic and dicationic
drugs21, and previous studies showed that the activity of QacA
against the dicationic drugs depends on an acidic residue at
position 323 (ref. 22). Neutralizing D323 by mutagenesis
interrupted the activity towards dications while retaining active
efflux of monocationic drugs6. We speculated that a mechanism
involving two consecutive transport cycles should enable the
D323 mutant of QacA to transport LDCs, but not SDCs. This was
tested by drug resistance assays using E. coli harbouring
heterologously expressed QacA or QacA(D323C). As expected,
the results show that wild-type QacA was able to confer resistance
against all tested dicationic drugs (LDCs and SDCs) (Fig. 9a).
pH 6.5
pH 7.0
pH 7.5
pH 8.0
Vec
MdfA
Vec
MdfA
Vec
MdfA
Vec
MdfA
No drug
4 µg ml–1 CAM
400 µg ml–1 EtBr
50 µg ml–1 Dq
200 µg ml–1 Pent
1 µg ml–1 Chx
0
50
100
0510
Time (min)
Dq fluorescence (%)
Vec
wt
ba
Figure 5 | pH dependence of MdfA-mediated Mdr. (a) pH-dependent drug resistance assay. The experiment was done as in Fig. 1b, with a pH-buffering
agent (Bis-Tris-Propane) added to the medium to generate the pH indicated to the right. (b) Accumulation of Dq in DpH-dissipated cells. Dissipation
of DpH was achieved by conducting the assay at pH 8 in the presence of 2 mM nigericin. The membranes were permeabilized to nigericin by an
EDTA-ethanol treatment, which did not affect the assay when the DpH was not dissipated. The experiments were repeated at least three times and the
results shown are representative.
NATURE COMMUNICATIONS | DOI: 10.1038/ncomms5615 ARTICLE
NATURE COMMUNICATIONS | 5:4615 | DOI: 10.1038/ncomms5615 | www.nature.com/naturecommunications 5
&2014 Macmillan Publishers Limited. All rights reserved.
Remarkably, however, although the QacA(D323C) mutant had
partially lost its ability to transport SDCs, it retained resistance
activity against LDCs (Dq and Chx) (Fig. 9a). These results
support our hypothesis that a monocationic Mdr transporter can
transport dicationic substrates that belong to the LDC group. In
support of this notion, previous studies have shown that QacB, a
close variant of QacA harbouring a neutral residue at position 323
and consequently lacked activity against most dications, was
found to retain activity towards several dicationic drugs. The
common denominator of these drugs is that their two cationic
moieties are separated by a long linker6,22,23, and meet our
criteria for LDCs.
Finally, as an independent complementary approach, we
characterized the activity of MdfA(V231E). Unlike wild-type
MdfA, this mutant is able to translocate two protons per
transport cycle8, rendering it active against both LDCs and
Fluorescence (a.u.)
500
540
580
480
520
560
500
540
580
440
480
520
TPP
TPP
2 µM
HCl
Dq
Dq
2 µM
HCl
2 µM
HCl
Pent
Pent
TPP Dq
Pent
2 µM
HCl
123456
Time (min)
0
Pent
TPP
Dq
No
MdfA
Figure 6 | Substrate-induced proton release. MdfA was solubilized and
purified to homogeneity and dialysed to remove buffering agents. The
protein was diluted to 2 mM and the solution acidity was monitored in a
time-dependent manner by measuring the pH-sensitive fluorescence of the
probe fluorescein. Before the assay, the protein was titrated by NaOH or
HCl to a fluorescent signal of B550 (pHB6.5). Notably, the absence of a
buffering agent generates spontaneous slow drifts in the acidity, creating a
background signal. The substrate-induced changes are distinguished
from the background by their much higher rate of changes. The arrows
denote additions of HCl (2 mM), TPP (150 mM), Pent (300 mM) or Dq
(50 mM). HCl was added in an MdfA-equimolar amount to reveal the signal
expected by a stoichiometric release of a single proton per MdfA molecule.
This signal is nearly identical to that of substrates (TPP/Dq/Pent),
suggesting that all substrates release a single proton from MdfA. Substrates
were added at saturating concentrations. As a control, substrates were
added twice to examine the saturability of binding and the related proton
release. In addition, substrates did not appreciably affect fluorescence when
added to a solution without MdfA (lower panel). The experiments were
repeated at least three times and the results shown are representative.
K
d
=0.7±0.1 µM
K
I
=2.2±0.2 µM
0
50
100
150
200
250
02040
60
70
80
90
100
0510
0
Dq
Dq + MdfA
Dq + MdfA + TPP
Dq + MdfA(R112M)
40
80
120
Dq fluorescence (%)
[3H]-TPP binding (counts)
[MdfA] (µM)
[Dq] (µM)
Dq fluorescence (%)
ab
cd
10
17
26
34
43
72
95
N8
C4
Dq
UV
+–+ +–+
–++ –++
MdfA
kDa
*
Figure 7 | Affinity and stoichiometry of Dq binding by purified MdfA.
(a) Quenching of 0.2 mM Dq fluorescence by 2 mM wild-type MdfA or
MdfA(R112M). Effect of adding 1 mM TPP to the mixture is also shown.
Error bars indicate s.d. of triplicate measurements (*Po10 4, two-tailed t-
test). (b) Quenching of 0.2 mM Dq fluorescence by increasing MdfA
concentrations. The binding of Dq to MdfA was analysed by nonlinear
regression fitting (line), yielding the indicated K
d
.(c) Inhibition of [3H]-TPP
binding to MdfA by increasing Dq concentrations. The results were
analysed by non-linear regression fitting (line), yielding the indicated K
I
.
Error bars indicate s.d. of triplicate measurements. (d) Dq-mediated
photoinduced cross-linking of MdfA. Membranes expressing functional
split-MdfA (C-terminally tagged N8-His
6
and C4-His
6
, left lanes) or wild-
type MdfA-His
6
(right lanes) were incubated in the absence or presence of
100 mM Dq and irradiated with ultraviolet. Proteins were separated by SDS–
PAGE and western blotting against the His tag. Cross-linked product of N8
and C4 is indicated by a star.
+
+
+
+
+
+
+
+
+
+
Periplasm
Cytoplasm
Figure 8 | Model for transport of divalent cations in two successive
transport cycles. The transporter first binds and transports only one
cationic moiety (sphere) of the substrate. Once this moiety is expelled, the
transporter binds and then transports the second half of the substrate. Each
transport cycle is depicted as an interconversion between inward- and
outward-facing conformations.
ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms5615
6NATURE COMMUNICATIONS | 5:4615 | DOI: 10.1038/ncomms5615 | www.nature.com/naturecommunications
&2014 Macmillan Publishers Limited. All rights reserved.
SDCs (as QacA) (Fig. 9b, upper panel). If MdfA(V231E) exports
LDCs in two successive cycles, in exchange for 2 protons per cycle
(a total of 4 protons), the antiport would be electrogenic. In
contrast, this mutant might catalyse electroneutral SDC/proton
antiport, as these compounds are exported in a single cycle in
exchange for two protons. To assess this prediction in vivo,we
tested the pH dependence of MdfA(V231E) resistance activity
against dications. The results show that this mutant could not
confer resistance against SDCs (DAPI and Dmn) at elevated pH,
suggesting electroneutral transport of these drugs (exchange of
two protons per one dicationic substrate) (Fig. 9b, lower panel).
In contrast to SDCs, MdfA(V231E) conferred resistance against
LDCs (Chx and Dq) even at pH 8 (Fig. 9b, lower panel),
suggesting that their transport is indeed electrogenic. When
considered in the context of our hypothesis that LDCs are
transported in two consecutive cycles, the indication that even in
the absence of DpH the mutant can energize Dq efflux suggests
that the transport is likely to be electrogenic and each of the two
positive moieties of Dq might be exchanged for two protons.
Together, the results suggest that MdfA(V231E) catalyses efflux of
LDCs with an overall higher proton exchange stoichiometry than
SDCs, in support of the hypothesis that LDCs are exported by
MFS secondary Mdr transporters in two consecutive cycles.
Discussion
Recently, important aspects of the mechanism underlying MdfA-
mediated antiport were elucidated4. A key principle in the
mechanism is that substrate binding triggers release of a single
stoichiometric proton from MdfA. This offers an explanation for
the proton/drug antiport stoichiometry of 1 that was observed
with MdfA5. Different MFS Mdr transporters work with distinct
proton antiport stoichiometries. For example, QacA from
S. aureus and LmrP from L. lactis use stoichiometries of up to
2 or 3 (refs 6,7). Compared with most Mdr transporters, the
proton/drug stoichiometry of 1 (as shown for MdfA) is relatively
low, rendering it generally inactive in efflux of dicationic drugs8.
Indeed, MdfA mutants that are capable of pumping two protons
per drug molecule acquire resistance also against dicationic
drugs8. Therefore, in the case of wild-type MdfA, the discovery of
efflux of certain dicationic drugs (LDCs) was surprising and
posed a mechanistic riddle. An important clue to solving this was
that MdfA seems to confer resistance only to LDCs, whose two
positive charges are separated by a long linker.
Our studies showed that LDCs are transported in a manner
similar to the monocationic drugs. Namely, they are exchanged
with protons, bind the transporter with a 1:1 stoichiometry and,
most importantly, their transport is electroneutral. This suggests
that they are probably exchanged for two protons. Nevertheless,
binding of LDCs induces the release of only a single proton from
MdfA, just like monocationic drugs4. If so, why is their proton/
drug antiport stoichiometry twice higher? The unique
architecture of these substrates, with two charges that are
separated by a long linker, offers a hypothetical solution. We
propose that MdfA exports LDCs in two successive transport
cycles, in which each charged moiety is transported as if it were a
separate substrate. In each cycle, one proton is transported, thus
providing a total of two protons imported during the process,
which ends after the full release of the LDC molecule from MdfA.
According to this scenario, the carbon chain that separates the
two charged domains of the LDCs must remain buried inside the
transporter, or in the lipid-transporter interface, just before the
second transport cycle (Fig. 8). Such a hydrophobic linker may
behave as a carbon chain of lipids and integrate into the lipid–
protein structure without interrupting the ability of MdfA to
undergo transport-related conformational changes. Through
additional studies with QacA and its D323C mutant and with
the MdfA mutant V231E, we obtained evidence in support of the
proposed mechanism and the notion that it might be universal in
secondary Mdr transporters. Briefly, the results show that
QacA(D323C), which is in principal a monocationic drug
pump, can confer resistance against LDCs. Furthermore, we
showed that an MdfA mutant that generally exchanges two
protons per dicationic drugs from the SDC group can transport
also LDCs; however, these compounds are exchanged for more
than two protons. This is consistent with the hypothesis that
LDCs are transported in two sequential cycles also by this MdfA
mutant. It seems that the simplest explanation for these
observations is that MFS Mdr transporters handle certain
dicationic drugs as if they were two separate monocations,
which are transported separately.
Transport of a substrate in more than a single cycle has never
been demonstrated for secondary or primary active solute
transporters. Transported substrates are usually small molecules
that do not justify such a mechanism, because they can be
transported in a single step. Transport of larger molecules, such as
proteins across the membrane is usually catalysed by machineries
that generate a continuous transmembrane channel24. The
pH 6.5 Vec
MdfA
100 µg ml–1 Dmn 4 µg ml–1 DAPI
No drug 0.5 µg ml–1 Chx 20 µg ml–1 Dq
MdfA(V231E)
10 µg ml–1 Dmn 1 µg ml–1 DAPI
No drug 0.5 µg ml–1 Chx 20 µg ml–1 Dq
pH 8 Vec
MdfA
MdfA(V231E)
100 µg ml–1 Dmn 10 µg ml–1 DAPI
No drug 2 µg ml–1 Chx 25 µg ml–1 Dq
Vec
QacA
QacA(D323C)
a
b
Figure 9 | Effect of acidic residue substitutions on resistance against dications by QacA and MdfA. The experiments were done as in
Figs 1b and 5. (a) Mdr by E. coli expressing QacA or its D323C mutant. (b) pH dependence of multidrug resistance by E. coli UTmdfAHkan expressing
MdfA or its V231E mutant. Resistance to Dmn and DAPI at pH 8 was tested at lower drug concentrations than pH 6.5 to make sure that even
low-level resistance is not conferred. The experiments were repeated at least three times and the results shown are representative.
NATURE COMMUNICATIONS | DOI: 10.1038/ncomms5615 ARTICLE
NATURE COMMUNICATIONS | 5:4615 | DOI: 10.1038/ncomms5615 | www.nature.com/naturecommunications 7
&2014 Macmillan Publishers Limited. All rights reserved.
continuous channel enables such large substrates to be transport-
ed in a step-wise, processive manner. In active transporters, the
substrate-conducting channel is always blocked on either side of
the membrane. The transport mechanism works such that when
the blockade is opened on one side, it closes on the other side of
the membrane, enabling the substrate to bind or dissociate from
either side of the membrane alternatively. This blockade should
hinder processive transport; however, our results suggest that
relatively narrow, hydrophobic linkers may still be threaded
through the blocked channel. This mechanism might be shared
by other active transporters that translocate large substrates with
long linkers. The unusual processive mechanism broadens the
substrate profile of Mdr transporters and may extend the
spectrum of substrates that are accessible for active transporters
in general.
Methods
Drug resistance assay.Antibacterial resistance was assayed as described25.
Briefly, E. coli UTmdfA::Kan or UTL2mdfA::Kan25 harbouring empty vector or
expressing MdfA were grown aerobically to A
600
of 1 and a series of 10-fold
dilutions was prepared. Four microlitres of the serial dilution were spotted on drug-
containing LB-agar plates and tested for growth after overnight incubation at 37 °C.
pH-dependent resistance was tested by adding 75 mM bis-Tris-propane buffer at
the indicated pH to the media as described5. Resistance conferred by QacA was
assayed similarly with the following modifications: (i) the E. coli strain used was
DH5a; (ii) the control vector was pBluescriptII, on which the plasmids encoding
QacA (pSK4322) and QacA(D323C) (pSK4432) are based; (iii) liquid cultures were
grown in 2YT media supplemented with 0.1 mg ml 1ampicillin. All drugs were
from Sigma Aldrich.
MdfA overexpression and purification.E. coli cells harbouring plasmid pUC18/
Para/mdfA- His6 were grown at 37 °C in LB medium supplemented with ampicillin
(200 mgml1). Overnight cultures were diluted to 0.07 A
600
units, grown to 1.0
A
600
units and induced with 0.2% arabinose for 1 h. A typical 10-l culture yielded
15 g (wet weight) of cells. Cell pellets were washed once in 150 ml of 50 mM KPi
(pH 7.3) supplemented with 5 mM MgSO
4
and collected by centrifugation (10 min,
10,000 g). Next, the cells were suspended in 150 ml of the same buffer containing
10 mgml1DNAse and 1 mM phenylmethylsulphonyl fluoride and passed three
times through a liquidizer (Emulsiflex-C5, Avestin) (10,000 psi) for disruption. Cell
debris was removed by centrifugation (30 min, 10,000 g) and the membranes were
collected by ultracentrifugation (1 h, 100,000 g). The membranes were homo-
genized in 27 ml of urea buffer (20mM Tris-HCl pH 8, 0.5 M NaCl, 5 M urea, 10%
glycerol, 28 mM b-mercaptoethanol and 1 mM phenylmethylsulp honyl fluoride),
incubated for 30 min at 4 °C and collected by ultracentrifugation (2.5 h, 100,000g).
The membranes were washed with 27 ml of membrane buffer (20 mM Tris–HCl
pH 8, 0.5 M NaCl, 10% glycerol and 3.5 mM b-mercaptoethanol). Finally, the
membranes were suspended by homogenization in 7 ml of the same buffer and
aliquots of 1 ml were snap-frozen in liquid nitrogen and stored at 80 °C.
For MdfA purification, thawed membranes were diluted fivefold in
solubilization buffer (20 mM Tris–HCl pH 8, 0.5 M NaCl, 10% glycerol, 0.1%
b-dodecyl maltopyranoside (DDM), 5 mM imidazole) and solubilized by gradual
addition of 10% DDM (Anatrace) to a final concentration of 1.1% and
homogenized. Insoluble material was discarded by ultracentrifugation (30 min,
100,000g) and the soluble fraction was mixed with solubilization buffer-
equilibrated Talon beads (Clontech) (typically 0.5 ml for 2 ml of thawed
membranes). Next, the mixture was agitated for 3 h at 4 °C and the suspension was
poured into a column. The column was then washed twice with ten-column
volumes of solubilization buffer. MdfA was eluted in elution buffer (20 mM
Tris–HCl pH 7.2, 0.12 M NaCl, 10% glycerol, 0.1% DDM, 100 mM imidazole). The
protein was collected after initial discard of 0.75 column volume. The protein was
then dialysed overnight against dialysis buffer (20 mM Tris-HCl pH 7.2, 0.12 M
NaCl, 10% glycerol, 0.01% DDM) at 4 °C. Protein concentration was determined by
measuring A
280
(1 mg ml 1B2.1 A
280
)13.
Measurement of [3H]-TPP binding.Radiolabelled TPP binding was carried out
by immobilizing purified MdfA-His
6
on Ni2þbeads, allowing it to bind [3H]-TPP
(American Radiolabeled Chemicals) and counting the MdfA-associated radio-
activity as described13. For assessment of competition, the indicated concentrations
of competitors were added during [3H]-TPP binding. To determine K
I
, the data
were fitted to the equation B¼B
max
B
max
([Dq]/([Dq þK
I
)) using nonlinear
regression, where Bis the amount of bound [3H]-TPP (in counts) and B
max
is the
maximal amount of B(without added Dq).
Dq transport in whole cells.Overnight cultures of E. coli UTL2mdfA::kan cells
harbouring empty vector (pT7-5) or plasmid pT7-5/mdfA-His
6
were diluted
50-fold and grown at 37 °C in LB broth supplemented with ampicillin
(100 mgml1) and kanamycin (30 mgml1)uptoanA
600
of B1 and transferred
to ice. Ice-cold cell suspensions (0.3 A
600
units) were pelleted and washed once with
1 ml of wash buffer (50 mM KPi pH 7, 0.2% glucose, 2mM MgSO
4
), centrifuged
and resuspended in 110 ml of the same buffer. From these cells, 100 ml were added
to a cuvette containing 1,900 ml of pre-warmed (37 °C) wash buffer supplemented
with 1 mM Dq (Sigma). The suspension was continuously stirred and warmed while
measuring Dq fluorescence quenching (excitation 330 nm, emission 370 nm).
For experiments with CAM, the wash buffers were supplemente d with 0.8 mM
CAM (Sigma).
For dissipation of DpH, the protocol was modified as follows: cell suspensions
were pelleted, resuspended and incubated for 10 min in modified wash buffer
(50 mM KPi pH 8, 1% EtOH, 2 mM EDTA, 2 mM nigericin (Fermentek)). Cells
were centrifuged and resuspended in 110 ml of the same buffer. From these cells,
100 ml were added to cuvette containing 1,900 ml of pre-warmed (37 °C) modified
wash buffer supplemented with 1 mM Dq and 0.2% glucose, and the fluorescence
was recorded.
Estimation of Dq fluorescence inside cells was done similarly with
modifications. Cells (A
600
of 1.5) were pelleted and resuspended in 100 ml wash
buffer. To this suspension, 1.5 ml of Dq-supplemented wash buffer was added and
the cells were allowed to accumulate Dq for 0 or 14 min at 37 °C. Cells were
centrifuged (1 min, 13,000 r.p.m., tabletop centrifuge, Eppendorf 5424)
immediately after incubation and the supernatant was discarded. Pellets were
resuspended in 100 ml water, mixed with 1.5 ml of 1% SDS and incubated at
ambient temperature for several minutes. Dq fluorescence in these lysates was
determined fluorimetrically.
Proton transport in everted membrane vesicles.Proton antiport was assayed as
described4.E. coli UT5600/mdfA::kan cells overexpressing MdfA (from plasmid
pT7-5/araP/mdfA-His6) or harbouring empty vector were grown to A
600
of B1.2,
induced for 1 h by arabinose and everted membrane vesicles were prepared as
described26. Membranes were frozen by liquid N
2
in 10% glycerol, 5 mM MgSO
5
,
50 mM KPi pH 7.3 in a concentration of 20 mg ml 1protein and stored at
80 °C. Membranes were thawed quickly at 46 °C for experiments. For each
measurement, 20 ml membranes were diluted into 2 ml of pre-warmed (30 °C)
proton transport solution (50 m mM K-gluconate, 50 mM KCl, 10mM MgSO4,
titrated to pH 6 and supplemented with 1 mM ACMA (Sigma-Aldrich)).
Membranes were equilibrated for 4 min before fluorescence measurement
(excitation and emission wavelengths of 409 and 474, respectively). The samples
were continuously stirred during measurement. At the indicated times,
D,L-lactate (2 mM), drugs (at the indicated concentrations) or carbonyl cyanide
m-chlorophenyl hydrazone (0.01 mM, Sigma-Aldrich) were added. Lactate stocks
(200 mM) were titrated with NaOH to pH 7 before experiments.
Measurement of substrate-induced proton release.Proton release was mea-
sured as described4. Purified MdfA was dialysed 4 times against 100 volumes of
unbuffered solution containing 120 mM NaCl, 10% glycerol and 0.01% DDM. The
protein was diluted to 2 mM in 120 mM NaCl, 10% glycerol, 0.1% DDM and 1 mM
fluorescein was added. The mixture was titrated to fluorescence of B550
(pHB6.5) before the experiment. During the experiment, the fluorescence was
measured (excitation and emission wavelengths of 488 and 513, respectively), while
samples were continuously stirred at ambient temperature. Substrates were added
at concentrations of 150 mM (TPP), 50 mM (Dq) or 300 mM (Pent). HCl was added
at a concentration of 2 mM.
Dq binding measured by fluorescence.Purified MdfA was concentrated using
Vivaspin ultrafiltration spin columns (100 kDa cutoff, Sartorius). MdfA and Dq
were mixed at concentra tions of 10 mM MdfA and 0.2 mM Dq. Next, this mixture
was serially diluted (1.42-fold each dilution) into a solution containing only 0.2 mM
Dq, keeping [Dq] constant while diluting MdfA. The fluorescence of each
increasingly diluted sample was measured in a quartz cuvette (excitation and
emission wavelengths of 330 and 370, respectively). To determine K
d
, the data were
fitted to the equation: fluorescence (%) ¼100% ((DF[MdfA])/(K
d
þ[MdfA]))
using nonlinear regression, where DFis the percent fluorescence difference upon
MdfA binding.
Dq-mediated cross-linking of MdfA.Everted membrane vesicles were prepared
from cells overexpressing MdfA or the functional split mutant N8-His
6
/C4-His
6
(ref. 20) as described above for proton transport measurements. Membrane vesicles
were thawed, collected by ultracentrifugation (100,000 g,30min,4°C) and
resuspended in 0.5 M NaCl, 10% glycerol, 20 mM Tris-HCl pH 8 at a total protein
concentration of 1 mg ml 1. Dq (100 mM final) or buffer wa s added to 40 ml
vesicles and was allowed to bind during 20 min incubation at ambient temperature
in the dark. Samples were transferred to ice-cold 96-well plates and irradiated on
ice for 20 min with ultraviolet at 366 nm from a distance of 2.7 cm using Model
UVGL-25 Minervalight lamp (San Gabriel, CA). Irradiated samples (40 ml) were
mixed with 13.3 ml4protein sample buffer and 10 ml samples were analysed by
SDS–PAGE and western blotting as described4.
ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms5615
8NATURE COMMUNICATIONS | 5:4615 | DOI: 10.1038/ncomms5615 | www.nature.com/naturecommunications
&2014 Macmillan Publishers Limited. All rights reserved.
References
1. Fluman, N. & Bibi, E. Bacterial multidrug transport through the lens of the
major facilitator superfamily. Biochim. Biophys. Acta 1794, 738–747 (2009).
2. Nishino, K. & Yamaguchi, A. Analysis of a complete library of putative drug
transporter genes in Escherichia coli.J. Bacteriol. 183, 5803–5812 (2001).
3. Saier, Jr M. H. & Paulsen, I. T. Phylogeny of multidrug transporters. Semin. Cell
Dev. Biol. 12, 205–213 (2001).
4. Fluman, N., Ryan, C. M., Whitelegge, J. P. & Bibi, E. Dissection of mechanistic
principles of a secondary multidrug efflux protein. Mol. Cell 47, 777–787
(2012).
5. Lewinson, O. et al. The Escherichia coli multidrug transporter MdfA catalyzes
both electrogenic and electroneutral transport reactions. Proc. Natl Acad. Sci.
USA 100, 1667–1672 (2003).
6. Mitchell, B. A., Paulsen, I. T., Brown, M. H. & Skurray, R. A. Bioenergetics of
the staphylococcal multidrug export protein QacA. Identification of distinct
binding sites for monovalent and divalent cations. J. Biol. Chem. 274,
3541–3548 (1999).
7. Schaedler, T. A. & van Veen, H. W. A flexible cation binding site in the
multidrug major facilitator superfamily transporter LmrP is associated with
variable proton coupling. FASEB J. 24, 3653–3661 (2010).
8. Tirosh, O. et al. Manipulating the drug/proton antiport stoichiometry of the
secondary multidrug transporter MdfA. Proc. Natl Acad. Sci. USA 109,
12473–12478 (2012).
9. Sigal, N. et al. 3D model of the Escherichia coli multidrug transporter MdfA
reveals an essential membrane-embedded positive charge. Biochemistry 44,
14870–14880 (2005).
10. Lewinson, O. & Bibi, E. Evidence for simultaneous binding of dissimilar
substrates by the Escherichia coli multidrug transporter MdfA. Biochemistry 40,
12612–12618 (2001).
11. Zilberstein, D., Schuldiner, S. & Padan, E. Proton electrochemical gradient in
Escherichia coli cells and its relation to active transport of lactose. Biochemistry
18, 669–673 (1979).
12. Guan, L., Murphy, F. D. & Kaback, H. R. Surface-exposed positions in the
transmembrane helices of the lactose permease of Escherichia coli determined
by intermolecular thiol cross-linking. Proc. Natl Acad. Sci. USA 99, 3475–3480
(2002).
13. Sigal, N., Lewinson, O., Wolf, S. G. & Bibi, E. E. coli multidrug transporter
MdfA is a monomer. Biochemistry 46, 5200–5208 (2007).
14. Abramson, J. et al. Structure and mechanism of the lactose permease of
Escherichia coli.Science 301, 610–615 (2003).
15. Huang, Y., Lemieux, M. J., Song, J., Auer, M. & Wang, D. N. Structure and
mechanism of the glycerol-3-phosphate transporter from Escherichia coli.
Science 301, 616–620 (2003).
16. Newstead, S. et al. Crystal structure of a prokaryotic homologue of the
mammalian oligopeptide-proton symporters, PepT1 and PepT2. EMBO J. 30,
417–426 (2011).
17. Dang, S. et al. Structure of a fucose transporter in an outward-open
conformation. Nature 467, 734–738 (2010).
18. Yin, Y., He, X., Szewczyk, P., Nguyen, T. & Chang, G. Structure of the
multidrug transporter EmrD from Escherichia coli.Science 312, 741–744
(2006).
19. Rotenberg, S. A. & Sun, X. G. Photoinduced inactivation of protein kinase C by
dequalinium identifies the RACK-1-binding domain as a recognition site.
J. Biol. Chem. 273, 2390–2395 (1998).
20. Adler, J. & Bibi, E. Promiscuity in the geometry of electrostatic interactions
between the Escherichia coli multidrug resistance transporter MdfA and
cationic substrates. J. Biol. Chem. 280, 2721–2729 (2005).
21. Brown, M. H. & Skurray, R. A. Staphylococcal multidrug efflux protein QacA.
J. Mol. Microbiol. Biotechnol. 3, 163–170 (2001).
22. Paulsen, I. T., Brown, M. H., Littlejohn, T. G., Mitchell, B. A. & Skurray, R. A.
Multidrug resistance proteins QacA and QacB from Staphylococcus aureus:
membrane topology and identification of residues involved in substrate
specificity. Proc. Natl Acad. Sci. 93, 3630–3635 (1996).
23. Mitchell, B. A., Brown, M. H. & Skurray, R. A. QacA multidrug efflux pump
from Staphylococcus aureus: comparative analysis of resistance to diamidines,
biguanidines, and guanylhydrazones. Antimicrob. Agents Chemother. 42,
475–477 (1998).
24. Rapoport, T. A. Protein translocation across the eukaryotic endoplasmic
reticulum and bacterial plasma membranes. Nature 450, 663–669 (2007).
25. Sigal, N., Fluman, N., Siemion, S. & Bibi, E. The secondary multidrug/proton
antiporter MdfA tolerates displacements of an essential negatively charged side
chain. J. Biol. Chem. 284, 6966–6971 (2009).
26. Reenstra, W. W., Patel, L., Rottenberg, H. & Kaback, H. R. Electrochemical
proton gradient in inverted membrane vesicles from Escherichia coli.
Biochemistry 19, 1–9 (1980).
Acknowledgements
This work was supported by a grant from the Israel Science Foundation (1128/06) to E.B.
and by the Weizmann Australia programme to E.B. and M.H.B.
Author contributions
N.F. and E.B. designed research and wrote the paper. N.F. performed experiments. J.A.
performed preliminary studies. M.H.B. provided QacA expressing plasmids. S.A.R. was
consulted and provided Dq and analogues.
Additional information
Competing financial interests: The authors declare no competing financial interests.
Reprints and permission information is available online at http://npg.nature.com/
reprintsandpermissions/
How to cite this article: Fluman, N. et al. Export of a single drug molecule
in two transport cycles by a multidrug efflux pump. Nat. Commun. 5:4615
doi: 10.1038/ncomms5615 (2014).
NATURE COMMUNICATIONS | DOI: 10.1038/ncomms5615 ARTICLE
NATURE COMMUNICATIONS | 5:4615 | DOI: 10.1038/ncomms5615 | www.nature.com/naturecommunications 9
&2014 Macmillan Publishers Limited. All rights reserved.
... By contrast, the addition of Cm to a solution containing E26T/D34M, which lacks A150E, was unable to evoke the release of any proton from the protein. To corroborate the essential role of A150E in the transporter-mediated export of Cm, we used the inside-out (everted) membrane vesicles to study the Cm/H + antiport 21,26 . We observed that Cm elicited the counter-movement of H + in everted vesicles expressing E26T/D34M/A150E, but not E26T/D34M or vector (Fig. 3b). ...
... fluorescence measurement of proton-release. This assay was conduct as previously performed 19,21 . ...
... Drug-proton antiport assay. This assay was carried out as previously performed 19,21 . The inside-out (everted) membrane vesicles were prepared from the BL21 (DE3) ∆acrAB∆macAB∆yojHI cells expressing the E26T/D34M/A150E variants 63,64 . ...
Article
Full-text available
The rapid increase of multidrug resistance poses urgent threats to human health. Multidrug transporters prompt multidrug resistance by exporting different therapeutics across cell membranes, often by utilizing the H+ electrochemical gradient. MdfA from Escherichia coli is a prototypical H+ -dependent multidrug transporter belonging to the Major Facilitator Superfamily. Prior studies revealed unusual flexibility in the coupling between multidrug binding and deprotonation in MdfA, but the mechanistic basis for this flexibility was obscure. Here we report the X-ray structures of a MdfA mutant E26T/D34M/A150E, wherein the multidrug-binding and protonation sites were revamped, separately bound to three different substrates at resolutions up to 2.0 Å. To validate the functional relevance of these structures, we conducted mutational and biochemical studies. Our data elucidated intermediate states during antibiotic recognition and suggested structural changes that accompany the substrate-evoked deprotonation of E26T/D34M/A150E. These findings help to explain the mechanistic flexibility in drug/H+ coupling observed in MdfA and may inspire therapeutic development to preempt efflux-mediated antimicrobial resistance.
... Whilst MdfA exchanges a single proton with a single monovalent cationic drug molecule, MdfA can also efflux divalent cations with a unique architecture where the two charged moieties are separated by a long linker. Compounds such as chlorhexidine, dequalinium and pentamidine are exchanged for two protons in two consecutive transport cycles, where each cationic moiety is transported as if it were a separate substrate [139,140]. Structures of a MdfA mutant (E26T/D34M/A150E) had revamped multidrug-binding and protonation sites that separately bound to three different substrates [141], and a molecular dynamics investigation revealed an intermediate state between its inward and outward conformations [142]. Other bacterial MFS multidrug efflux proteins with high-resolution structures available are EmrD from E. coli (PDB 2GFP, 3.50 Å) [143], YajR from E. coli (PDB 3WDO, 3.74 Å) [144], LmrP from Lactococcus lactis (PDB 6T1Z, 2.90 Å) [145], SotB from E. coli (PDB 6KKJ, 3.38 Å) [146,147], NorC from S. aureus (PDB 7D5P, 3.65 Å) [148] and NorA from S. aureus (PDB 2LO7, 3.74 Å) [149]. ...
Preprint
Full-text available
Multidrug efflux proteins, also known as efflux pumps, are one of the major mechanisms that bacteria have evolved for their resistance against antimicrobial agents. Gram-negative bacteria are intrinsically more resistant to many antibiotics and biocides due to their cell structure and the activity of multidrug efflux proteins. These transporters actively extrude antibiotics and other xenobiotics from the cytoplasm or surrounding membranes of cells to the external environment. Based on amino acid sequence similarity, substrate specificity and the energy source used to export their substrates, there are seven major families of distinct bacterial multidrug efflux proteins: ABC, RND, MFS, SMR, MATE, PACE, AbgT. Individual proteins may be highly specialized for one compound or highly promiscuous, transporting a broad range of structurally dissimilar substrates. Protein structural organization in a large majority of the families, including the number of transmembrane helices, has been confirmed by high-resolution structure determination for at least one member. In this book chapter, we provide an updated review on the families of bacterial multidrug efflux proteins, including basic properties, energization, structural organization and molecular mechanism. Using representative proteins from each family, we also performed analyses of transmembrane helices, amino acid composition and distribution of charged residues. Ongoing characterization of structure-function relationships and regulation of bacterial multidrug efflux proteins are necessary for contributing new knowledge to assist drug development and strategies that will overcome antimicrobial resistance.
... However, various MFS transporters bind in different sequences as well, such as initiated with binding with proton, then substrate, or with a substrate first and then proton. Additionally, it has been observed that the transportation cycle of MFS protein generally involves the binding (as an initial step) followed by the release of substrate and proton, however, variations in the stoichiometric drug-proton interconversion within family members have been reported (Schaedler and Van Veen, 2010;Fluman et al., 2014). Small multidrug resistance (SMR) efflux pump system utilizes proton motive force as an energy source was previously reported via an experiment with S. aureus Smr (Sau-Smr) conducted first by Grinius andGoldberg in 1994 (Littlejohn et al., 1992;Grinius and Goldberg, 1994). ...
Article
Full-text available
Efflux pumps function as an advanced defense system against antimicrobials by reducing the concentration of drugs inside the bacteria and extruding the substances outside. Various extraneous substances, including antimicrobials, toxic heavy metals, dyes, and detergents, have been removed by this protective barrier composed of diverse transporter proteins found in between the cell membrane and the periplasm within the bacterial cell. In this review, multiple efflux pump families have been analytically and widely outlined, and their potential applications have been discussed in detail. Additionally, this review also discusses a variety of biological functions of efflux pumps, including their role in the formation of biofilms, quorum sensing, their survivability, and the virulence in bacteria, and the genes/proteins associated with efflux pumps have also been explored for their potential relevance to antimicrobial resistance and antibiotic residue detection. A final discussion centers around efflux pump inhibitors, particularly those derived from plants.
... It is anticipated that the export of these kinds of divalent drug molecules is in need of two consecutive transport cycles, thereby consuming 2 protons. 393 In Gram-negative bacteria, this large superfamily comprises not only single component transporters but also tripartite efflux machineries. Tripartite multidrug pumps of the MFS family are encoded by chromosomal or plasmid genes, and usually PAP and MFS are coded on a single operon (e.g., emrAB in E. coli, salAB in Rhyzobium leguminosarum, or emrKY in Shigella f lexneri), while the OMF is borrowed from another operon. ...
Article
Full-text available
Tripartite efflux pumps and the related type 1 secretion systems (T1SSs) in Gram-negative organisms are diverse in function, energization, and structural organization. They form continuous conduits spanning both the inner and the outer membrane and are composed of three principal components-the energized inner membrane transporters (belonging to ABC, RND, and MFS families), the outer membrane factor channel-like proteins, and linking the two, the periplasmic adaptor proteins (PAPs), also known as the membrane fusion proteins (MFPs). In this review we summarize the recent advances in understanding of structural biology, function, and regulation of these systems, highlighting the previously undescribed role of PAPs in providing a common architectural scaffold across diverse families of transporters. Despite being built from a limited number of basic structural domains, these complexes present a staggering variety of architectures. While key insights have been derived from the RND transporter systems, a closer inspection of the operation and structural organization of different tripartite systems reveals unexpected analogies between them, including those formed around MFS- and ATP-driven transporters, suggesting that they operate around basic common principles. Based on that we are proposing a new integrated model of PAP-mediated communication within the conformational cycling of tripartite systems, which could be expanded to other types of assemblies.
... Each of the two positively charged aminomethylquinolinium moieties of DQ binds to two vicinal pockets on the QacA protein while the decamethylene linker adjusts the orientation to provide an optimal fit [13]. DQ can be transported by different types of efflux pumps, such as AcrB from Escherichia coli (Fig. 3) and other transporters [14][15][16]. In addition, the drug can directly bind to the transcriptional repressor RamR, which regulates the bacterial efflux and thus participates in multidrug resistance [17]. ...
Article
For more than 60 years dequalinium chloride (DQ) has been used as anti-infective drug, mainly to treat local infections. It is a standard drug to treat bacterial vaginosis and an active ingredient of sore-throat lozenges. As a lipophilic bis-quaternary ammonium molecule, the drug displays membrane effects and selectively targets mitochondria to deplete DNA and to block energy production in cells. But beyond its mitochondriotropic property, DQ can interfere with the correct functioning of diverse proteins. A dozen of DQ protein targets have been identified and their implication in the antibacterial, antiviral, antifungal, antiparasitic and anticancer properties of the drug is discussed here. The anticancer effects of DQ combine a mitochondrial action, a selective inhibition of kinases (PKC-α/β, Cdc7/Dbf4), and a modulation of Ca²⁺-activated K⁺ channels. At the bacterial level, DQ interacts with different multidrug transporters (QacR, AcrB, EmrE) and with the transcriptional regulator RamR. Other proteins implicated in the antiviral (MPER domain of gp41 HIV-1) and antiparasitic (chitinase A from Vibrio harveyi) activities have been identified. DQ also targets α -synuclein oligomers to restrict protofibrils formation implicated in some neurodegenerative disorders. In addition, DQ is a typical bolaamphiphile molecule, well suited to form liposomes and nanoparticules useful for drug entrapment and delivery (DQAsomes and others). Altogether, the review highlights the many pharmacological properties and therapeutic benefits of this old 'multi-talented' drug, which may be exploited further. Its multiple sites of actions in cells should be kept in mind when using DQ in experimental research.
... Advances in in vitro studies of transporter function and structure have expanded our understanding of the range of conformational changes that can produce alternating access (26, 27), and the mechanisms controlling stoichiometry and transport (28). Recently, two transporters, MdfA and PepT ST , have been found to use different proton/substrate transport stoichiometries when transporting substrates of different size or charge (29,30). Here we show that EmrE can transport a single substrate with multiple proton/substrate transport stoichiometries. ...
Preprint
EmrE is a small multidrug resistance transporter found in E. coli that confers resistance to toxic polyaromatic cations due to its proton-coupled antiport of these substrates. Here we show that EmrE breaks the rules generally deemed essential for coupled antiport. NMR spectra reveal that EmrE can simultaneously bind and cotransport proton and drug. The functional consequence of this finding is an exceptionally promiscuous transporter: Not only can EmrE export diverse drug substrates, it can couple antiport of a drug to either one or two protons, performing both electrogenic and electroneutral transport of a single substrate. We present a new kinetically-driven free exchange model for EmrE antiport that is consistent with these results and recapitulates ΔpH-driven concentrative drug uptake. Our results suggest that EmrE sacrifices coupling efficiency for initial transport speed and multidrug specificity. SIGNIFICANCE EmrE facilitates E. coli multidrug resistance by coupling drug efflux to proton import. This antiport mechanism has been thought to occur via a pure exchange model which achieves coupled antiport by restricting when the single binding pocket can alternate access between opposite sides of the membrane. We test this model using NMR titrations and transport assays and find it cannot account for EmrE antiport activity. We propose a new kinetically-driven free exchange model of antiport with fewer restrictions that better accounts for the highly promiscuous nature of EmrE drug efflux. This model expands our understanding of coupled antiport and has implications for transporter design and drug development.
... Since these transporters utilize electrochemical potential as energy to transport substrate and require the involvement of cations (Na ? or H ? ) in the transport of drugs (Fluman et al. 2014), the addition of sea salt might facilitate these transporters and accelerate the efflux of HYG to confer fungal resistance. ...
Article
Full-text available
Objectives To determine the effect of sea salt on the resistance of Trichoderma harzianum LZDX-32-08 to hygromycin B and speculate the possible mechanisms involved via transcriptome analysis.ResultsSea salt addition in media to simulate marine environment significantly increased the tolerance of marine-derived fungus Trichoderma harzianum LZDX-32-08 to hygromycin B from 40 to 500 μg/ml. Meanwhile, sea salt addition also elicited the hygromycin B resistance of 5 other marine or terrestrial fungi. Transcriptomic analyses of T. harzianum cultivated on PDA, PDA supplemented with sea salt and PDA with both sea salt and hygromycin B revealed that genes coding for P-type ATPases, multidrug resistance related transporters and acetyltransferases were up-regulated, while genes coding for Ca2+/H+ antiporter and 1,3-glucosidase were down-regulated, indicating probable increased efflux and inactivation of hygromycin B as well as enhanced biofilm formation, which could jointly contribute to the drug resistance.Conclusions Marine environment or high ion concentration in the environment could be an importance inducer for antifungal resistance. Possible mechanisms and related key genes were proposed for understanding the molecular basis and overcoming this resistance.
... Based on the results of the previous campaign, a second targeted screening was carried out by paying attention on known marketed bis-cationic molecules. The selection was based on the previous docking simulations as well as on a conceptually similar study on MdfA ligands [24] in order to collect molecules able to explore the role of the linker length and of the arrangement of the positively charged moieties. As shown in Figure 1, such a selection identified three promising bis-cationic ligands (i.e., chlorhexidine, pentamidine and diminazene) to which a fourth very rigid molecule was added (paraquat) to better investigate the role of linker flexibility. ...
Article
Full-text available
Structure-based virtual screening is a truly productive repurposing approach provided that reliable target structures are available. Recent progresses in the structural resolution of the G-Protein Coupled Receptors (GPCRs) render these targets amenable for structure-based repurposing studies. Hence, the present study describes structure-based virtual screening campaigns with a view to repurposing known drugs as potential allosteric (and/or orthosteric) ligands for the hM2 muscarinic subtype which was indeed resolved in complex with an allosteric modulator thus allowing a precise identification of this binding cavity. First, a docking protocol was developed and optimized based on binding space concept and enrichment factor optimization algorithm (EFO) consensus approach by using a purposely collected database including known allosteric modulators. The so-developed consensus models were then utilized to virtually screen the DrugBank database. Based on the computational results, six promising molecules were selected and experimentally tested and four of them revealed interesting affinity data; in particular, dequalinium showed a very impressive allosteric modulation for hM2. Based on these results, a second campaign was focused on bis-cationic derivatives and allowed the identification of other two relevant hM2 ligands. Overall, the study enhances the understanding of the factors governing the hM2 allosteric modulation emphasizing the key role of ligand flexibility as well as of arrangement and delocalization of the positively charged moieties.
Article
Full-text available
Multidrug efflux pumps function at the frontline to protect bacteria against antimicrobials by decreasing the intracellular concentration of drugs. This protective barrier consists of a series of transporter proteins, which are located in the bacterial cell membrane and periplasm and remove diverse extraneous substrates, including antimicrobials, organic solvents, toxic heavy metals, etc., from bacterial cells. This review systematically and comprehensively summarizes the functions of multiple efflux pumps families and discusses their potential applications. The biological functions of efflux pumps including their promotion of multidrug resistance, biofilm formation, quorum sensing, and survival and pathogenicity of bacteria are elucidated. The potential applications of efflux pump-related genes/proteins for the detection of antibiotic residues and antimicrobial resistance are also analyzed. Last but not least, efflux pump inhibitors, especially those of plant origin, are discussed.
Article
Bacterial multidrug efflux pumps have come to prominence in human and veterinary pathogenesis because they help bacteria protect themselves against the antimicrobials used to overcome their infections. However, it is increasingly realized that many, probably most, such pumps have physiological roles that are distinct from protection of bacteria against antimicrobials administered by humans. Here we undertake a broad survey of the proteins involved, allied to detailed examples of their evolution, energetics, structures, chemical recognition, and molecular mechanisms, together with the experimental strategies that enable rapid and economical progress in understanding their true physiological roles. Once these roles are established, the knowledge can be harnessed to design more effective drugs, improve existing microbial production of drugs for clinical practice and of feedstocks for commercial exploitation, and even develop more sustainable biological processes that avoid, for example, utilization of petroleum.
Article
Full-text available
Multidrug transporters are integral membrane proteins that use cellular energy to actively extrude antibiotics and other toxic compounds from cells. The multidrug/proton antiporter MdfA from Escherichia coli exchanges monovalent cationic substrates for protons with a stoichiometry of 1, meaning that it translocates only one proton per antiport cycle. This may explain why transport of divalent cationic drugs by MdfA is energetically unfavorable. Remarkably, however, we show that MdfA can be easily converted into a divalent cationic drug/≥ 2 proton-antiporter, either by random mutagenesis or by rational design. The results suggest that exchange of divalent cationi c drugs with two (or more) protons requires an additional acidic residue in the multidrug recognition pocket of MdfA. This outcome further illustrates the exceptional promiscuous capabilities of multidrug transporters.
Article
Full-text available
The staphylococcal multidrug efflux pump QacA mediates resistance to a broad spectrum of monovalent and divalent antimicrobial cations. Resistance toward various classes of these compounds identified features of the substrate that may be important for interaction with QacA. Analysis of combinations of two substrates suggested that the same mechanism is used for the extrusion of different classes of compounds.
Article
Full-text available
PepT1 and PepT2 are major facilitator superfamily (MFS) transporters that utilize a proton gradient to drive the uptake of di- and tri-peptides in the small intestine and kidney, respectively. They are the major routes by which we absorb dietary nitrogen and many orally administered drugs. Here, we present the crystal structure of PepT(So), a functionally similar prokaryotic homologue of the mammalian peptide transporters from Shewanella oneidensis. This structure, refined using data up to 3.6 Å resolution, reveals a ligand-bound occluded state for the MFS and provides new insights into a general transport mechanism. We have located the peptide-binding site in a central hydrophilic cavity, which occludes a bound ligand from both sides of the membrane. Residues thought to be involved in proton coupling have also been identified near the extracellular gate of the cavity. Based on these findings and associated kinetic data, we propose that PepT(So) represents a sound model system for understanding mammalian peptide transport as catalysed by PepT1 and PepT2.
Article
Full-text available
The multidrug major facilitator superfamily transporter LmrP from Lactococcus lactis mediates protonmotive-force dependent efflux of amphiphilic ligands from the cell. We compared the role of membrane-embedded carboxylates in transport and binding of divalent cationic propidium and monovalent cationic ethidium. D235N, E327Q, and D142N replacements each resulted in loss of electrogenicity in the propidium efflux reaction, pointing to electrogenic 3H(+)/propidium(2+) antiport. During ethidium efflux, single D142N and D235N replacements resulted in apparent loss of electrogenicity, whereas the E327Q substitution did not affect the energetics, consistent with electrogenic 2H(+)/ethidium(+) antiport. Different roles of carboxylates were also observed in fluorescence anisotropy-based ligand-binding assays. Whereas D235 and E327 were both involved in propidium binding, the loss of one of these carboxylates could be compensated for by the other in ethidium binding. The D142N replacement did not affect the binding of either ligand. These data point to the presence of a dedicated proton binding site containing D142, and a flexible proton/ligand binding site containing D235 and E327, the contributions to proton and ligand binding of which depend on the chemical structure of the bound ligand. Our findings provide the first evidence that multidrug transport by secondary-active transporters can be associated with variable ion coupling.
Article
Full-text available
The largest family of solute transporters includes ion motive force-driven secondary transporters. Several well characterized solute-specific transport systems in this group have at least one irreplaceable acidic residue that plays a critical role in energy coupling during transport. Previous studies have established the importance of acidic residues in substrate recognition by major facilitator superfamily secondary multidrug transporters, but their role in the transport mechanism remained unknown. We have been investigating the involvement of acidic residues in the mechanism of MdfA, an Escherichia coli secondary multidrug/proton antiporter. We demonstrated that no single negatively charged side chain plays an irreplaceable role in MdfA. Accordingly, we hypothesized that MdfA might be able to utilize at least two acidic residues alternatively. In this study, we present evidence that indeed, unlike solute-specific secondary transporters, MdfA tolerates displacements of an essential negative charge to various locations in the putative drug translocation pathway. The results suggest that MdfA utilizes a proton translocation strategy that is less sensitive to perturbations in the geometry of the proton-binding site, further illustrating the exceptional structural promiscuity of multidrug transporters.
Article
Multidrug transporters are ubiquitous efflux pumps that provide cells with defense against various toxic compounds. In bacteria, which typically harbor numerous multidrug transporter genes, the majority function as secondary multidrug/proton antiporters. Proton-coupled secondary transport is a fundamental process that is not fully understood, largely owing to the obscure nature of proton-transporter interactions. Here we analyzed the substrate/proton coupling mechanism in MdfA, a model multidrug/proton antiporter. By measuring the effect of protons on substrate binding and by directly measuring proton binding and release, we show that substrates and protons compete for binding to MdfA. Our studies strongly suggest that competition is an integral feature of secondary multidrug transport. We identified the proton-binding acidic residue and show that, surprisingly, the substrate binds at a different site. Together, the results suggest an interesting mode of indirect competition as a mechanism of multidrug/proton antiport.
Article
The major facilitator superfamily (MFS) transporters are an ancient and widespread family of secondary active transporters. In Escherichia coli, the uptake of l-fucose, a source of carbon for microorganisms, is mediated by an MFS proton symporter, FucP. Despite intensive study of the MFS transporters, atomic structure information is only available on three proteins and the outward-open conformation has yet to be captured. Here we report the crystal structure of FucP at 3.1 Å resolution, which shows that it contains an outward-open, amphipathic cavity. The similarly folded amino and carboxyl domains of FucP have contrasting surface features along the transport path, with negative electrostatic potential on the N domain and hydrophobic surface on the C domain. FucP only contains two acidic residues along the transport path, Asp 46 and Glu 135, which can undergo cycles of protonation and deprotonation. Their essential role in active transport is supported by both in vivo and in vitro experiments. Structure-based biochemical analyses provide insights into energy coupling, substrate recognition and the transport mechanism of FucP.
Article
Multidrug transporters are membrane proteins that expel a wide spectrum of cytotoxic compounds from the cell. Through this function, they render cells resistant to multiple drugs. These transporters are found in many different families of transport proteins, of which the largest is the major facilitator superfamily. Multidrug transporters from this family are highly represented in bacteria and studies of them have provided important insight into the mechanism underlying multidrug transport. This review summarizes the work carried out on these interesting proteins and underscores the differences and similarities to other transport systems.
Article
The electrochemical potential gradients of protons (Δμ̃H+) and lactose (Δμ̃lac) maintained by respiring Escherichia coli ML 308-225 (i-z-y+a+) cells have been measured as a function of the external pH. The proton gradient (ΔpH) was determined from the distribution across the cell membrane of the weak acid 5,5-dimethyloxazolidine-2,4-dione (DMO) after a rapid centrifugation step. The accumulation of tetraphenylphosphonium (TPP+; a lipophylic cation) and 86Rb+ in the presence of valinomycin by EDTA-treated cells incubated in a flow dialysis system has been used to calculate the membrane potential (ΔΨ)- Simultaneous measurements of Δμ̃H+ and lactose gradient were conducted using [14C]lactose and [3H]TPP+. Respiring EDTA-treated cells maintain, at neutrality, a Δμ̃H+ of 170 mV (ΔpH = 35 mV, ΔΨ = 135 mV, interior negative and alkaline). When the external pH was changed from 6.0 to 8.0, ΔΨ increased from 95 to 150 mV, while ΔpH decreased from 1.8 to -0.2 units (at pH 6 interior alkaline and at pH 8 interior acid). Thus, ΔΨ significantly compensates for the decrease in ΔpH, yielding a Δμ̃H+ of 200 mV at pH 6 and 140 mV at pH 8. In contrast to intact cells, inverted membrane vesicles from E. coli maintain a large ΔpH at external pH 8.0, while ΔΨ was undetectable. Lactose steady-state gradients were determined in cells maintaining different Δμ̃H+ levels in the presence of various concentrations of the uncoupler FCCP at pH 8. The results indicate that one proton is translocated with each molecule of lactose at pH 8 similar to the case at pH 6. This conclusion is supported by direct measurements of H+ and lactose fluxes induced by a lactose pulse in nonmetabolizing cells.