You are on page 1of 5

THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 274, No. 16, Issue of April 16, pp.

11110 –11114, 1999


© 1999 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A.

Mechanism of Triclosan Inhibition of Bacterial Fatty Acid


Synthesis*
(Received for publication, December 9, 1998, and in revised form, January 13, 1999)

Richard J. Heath‡, J. Ronald Rubin§, Debra R. Holland§, Erli Zhang§, Mark. E. Snow§, and
Charles O. Rock‡¶i
From the ‡Department of Biochemistry, St. Jude Children’s Research Hospital, Memphis, Tennessee 38105,
the §Department of Biomolecular Structure and Drug Design, Parke-Davis Pharmaceutical Research,
Ann Arbor, Michigan 48105, and the ¶Department of Biochemistry, University of Tennessee, Memphis, Tennessee 38163

Triclosan is a broad-spectrum antibacterial agent that the introduction of a cis double bond into the growing acyl
inhibits bacterial fatty acid synthesis at the enoyl-acyl chain at the b-hydroxydecanoyl-ACP step and most efficiently
carrier protein reductase (FabI) step. Resistance to tri- catalyzes dehydration of short-chain b-hydroxyacyl-ACPs,
closan in Escherichia coli is acquired through a mis- whereas the FabZ dehydratase has a broader substrate speci-
sense mutation in the fabI gene that leads to the expres- ficity (5). The last reaction in each elongation cycle is catalyzed
sion of FabI[G93V]. The specific activity and substrate by enoyl-ACP reductase (FabI). Contrary to the initial conclu-
affinities of FabI[G93V] are similar to FabI. Two differ- sion that there were two enoyl-ACP reductases, based on as-
ent binding assays establish that triclosan dramatically says in crude extracts (6), E. coli cells possess only a single
increases the affinity of FabI for NAD1. In contrast, NADH-dependent enoyl-ACP reductases encoded by the fabI
triclosan does not increase the binding of NAD1 to gene that utilizes all chain lengths (7).
FabI[G93V]. The x-ray crystal structure of the FabI-
The importance of fatty acid biosynthesis to cell growth and
NAD1-triclosan complex confirms that hydrogen bonds
function makes this pathway an attractive target for the de-
and hydrophobic interactions between triclosan and
velopment of antibacterial agents. Two important control
both the protein and the NAD1 cofactor contribute to
the formation of a stable ternary complex, with the drug points in the cycle are the condensing enzymes and the enoyl-
binding at the enoyl substrate site. These data show that ACP reductase (8, 9), and both reactions are targeted by com-
the formation of a noncovalent “bi-substrate” complex pounds that effectively inhibit fatty acid synthesis. Two natu-
accounts for the effectiveness of triclosan as a FabI in- ral products, cerulenin and thiolactomycin, are potent
hibitor and illustrates that mutations in the FabI active antibiotics that function by specifically inhibiting the condens-
site that interfere with the formation of a stable FabI- ing enzyme reactions (for reviews, see Refs. 3, 10). The diaza-
NAD1-triclosan ternary complex acquire resistance to borines, a class of heterocyclic antibacterials, inhibit fatty acid
the drug. biosynthesis by blocking the FabI step (11). Resistance to the
diazaborines arises from a missense mutation in the fabI gene
that leads to the expression of a FabI[G93S] mutant protein
The fatty acid synthase system of Escherichia coli is the (12, 13). Similarly, the fabI analog in Mycobacterium tubercu-
paradigm for the type II or dissociated fatty acid synthase losis, the inhA gene, encodes a cellular target for isoniazid and
systems (for reviews, see Refs. 1–3). Distinct genes encode each ethionamide. A point mutation in the inhA gene confers resist-
of the individual enzymes, and the same basic chemical reac- ance to the drugs (14). Both isoniazid and diazaborine bind at
tion is often catalyzed by multiple isozymes. There are four the substrate site of the respective enoyl-ACP reductases and
basic reactions that constitute a single round of elongation. The covalently react with NAD1 to form tight binding bi-substrate
first step is the condensation of malonyl-ACP1 with either complexes (15, 16). Triclosan is a broad-spectrum antibacterial
acetyl-CoA to initiate fatty acid synthesis (FabH) or with the agent that enjoys widespread applications in a multitude of
growing acyl chain to continue cycles of elongation (FabB or contemporary consumer products including, soaps, detergents,
FabF). The b-ketoacyl-ACP is reduced by an NADPH-depend- toothpastes, skin care products, cutting boards, and mattress
ent b-ketoacyl-ACP reductase (FabG). Only a single enzyme is pads (17). Triclosan is widely thought to be a nonspecific bio-
responsible for this step (4). There are two b-hydroxyacyl-ACP cide that attacks bacterial membranes (18 –20), and if triclosan
dehydrases (FabA and FabZ) capable of forming trans-2-enoyl- does not have a discrete mechanism of action, the acquisition of
ACP. The product of the fabA gene is specifically involved in cellular resistance is unlikely. However, recent work reveals
that resistant E. coli strains arise from missense mutations in
the fabI gene (21, 22) and that triclosan and other 2-hydroxy-
* This work was supported by National Institutes of Health Grant
diphenyl ethers directly inhibit fatty acid biosynthesis in vivo
GM 34496, Cancer Center (CORE) Support Grant CA 21765, and the
American Lebanese Syrian Associated Charities. The costs of publica- and FabI catalysis in vitro (22). The goal of this study was to
tion of this article were defrayed in part by the payment of page investigate the mechanism by which triclosan inhibits FabI.
charges. This article must therefore be hereby marked “advertisement”
in accordance with 18 U.S.C. Section 1734 solely to indicate this fact. EXPERIMENTAL PROCEDURES
i To whom correspondence should be addressed: Dept. of Biochemis-
try, St. Jude Children’s Research Hospital, 332 N. Lauderdale, Mem- Materials—Sources of supplies were: American Radiochemicals, Inc.,
phis, TN 38105-2794. Tel.: 901-495-3491; Fax: 901-525-8025; E-mail: [adenine-2,8-3H]NAD (specific activity 25 Ci/mmol); Sigma, NADH and
charles.rock@stjude.org. NAD1; and Millipore Corp., Ultrafree Probind centrifugal filtration
1
The abbreviations used are: ACP, acyl carrier protein; FabI, enoyl- units. The substrate analogs, 4:1-NAC and 8:1-NAC, were the generous
acyl carrier protein reductase; triclosan, 2,4,49-trichloro-29-hydroxydi- gift of Rocco Gogliotti and John Domagala (Parke-Davis). Triclosan was
phenyl ether; IC50, concentration giving 50% inhibition of activity; the gift of KIC Chemicals, Inc. (Armonk, NY). All other chemicals were
4:1-NAC, crotonyl-N-acetylcysteamine; 8:1-NAC, trans-2-octadecenoyl- of the best grade available.
N-acetylcysteamine. Purification of FabI and FabI[G93V]—The expression and purifica-

11110 This paper is available on line at http://www.jbc.org

This is an Open Access article under the CC BY license.


FabI-NAD1 Triclosan Ternary Complex 11111
G93V
tion of His-tag FabI was described previously (7). The fabI allele
was amplified from the chromosome of the triclosan-resistant strain
RJH108 (22) and ligated into the pET15b expression vector. DNA
sequencing of the expression construct confirmed the presence of the
single point mutation. The His-tagged FabI[G93V] was then expressed
and purified to homogeneity by Ni21 affinity chromatography using the
same methods as described for the wild-type FabI (7). Protein was
determined by the method of Bradford (23).
Spectrophotometric Assay of FabI—FabI activity was assayed spec-
trophotometrically by monitoring the decrease in absorption at 340 nm
using an adaptation of the spectrophotometric assay used by Bergler et
al. (24). Standard reactions contained 100 mM 8:1-NAC, 12 mg of homo-
geneous FabI (7, 8), 200 mM NADH, 0.1 M sodium phosphate, pH 7.5, in
a final volume of 300 ml. The reactions were performed at 24 °C in
semimicro quartz cuvettes. The change in optical density was continu-
ously monitored for 1 min, and the reaction rate was calculated from the
slope of the trace. Triclosan was added to the final concentrations
indicated in the figure legend from serially diluted stock solutions in
Me2SO. The Me2SO concentration in all assays was maintained at
1.66%, which did not significantly affect FabI activity. The FabI specific
activity in the absence of drugs under these experimental conditions
was 0.36 mmol/min/mg. Data points were the mean of duplicate assays,
and the individual values were within 6 5% of the average. FIG. 1. Triclosan inhibition of FabI and FabI[G93V]. The spec-
Analysis of [3H]NAD1 Binding to FabI—Two methods were used to trophotometric assay described under “Experimental Procedures” was
measure the binding of NAD1 to FabI in the presence and absence of used to determine the inhibition of the initial rate of FabI (E) or
triclosan. The first method employed gel filtration chromatography to FabI[G93V] (●) catalysis by triclosan using 8:1-NAC. The points are the
separate [3H]NAD1 bound to FabI from free [3H]NAD1. A 0.7 3 8-cm average of duplicate determinations from one experiment, and the
column (3 ml) of Sephadex G-25 was equilibrated in 0.1 M sodium experiment was replicated with the same results.
phosphate, pH 7.5, at 24 °C. Samples (50 ml) were applied, the column
eluted with the same buffer at a flow rate of approximately 0.25 ml/min, RESULTS
and 5-drop (110 ml) fractions were collected. The second method used Biochemical Characterization of the Triclosan-resistant Mu-
binding of FabI to a polyvinylidene difluoride membrane to separate
tant, FabI[G93V]—Previous work showed that FabI was inhib-
free from FabI-bound [3H]NAD1. Assays were placed into a Ultrafree-
Probind centrifugal filtration unit containing a 0.2-cm (2), 0.45-mm ited by triclosan (22) and that a mutation in the fabI gene that
polyvinylidene difluoride membrane (Millipore, Corp.). The filters were results in the expression of a FabI[G93V] protein exhibited
washed once with 100 ml of 0.1 M sodium phosphate and counted in 5 ml high level triclosan resistance (21, 22). FabI[G93V] was cloned
of ScintSafe scintillation solution. Both assays contained 0.1 M sodium into the pET15b expression vector and expressed and purified
phosphate, 1 mM [3H]NAD1 (specific activity 4 Ci/mmol), 2% Me2SO, the as described under “Experimental Procedures.” The apparent
indicated concentrations of triclosan, and 7.6 mg of FabI or FabI[G93V]
Km values for the two substrates were determined for both the
in a final volume of 50 ml. Samples were incubated at 24 °C for 15 min
and then separated using one of the two methods outlined above. wild-type and mutant protein. The wild-type protein had an
Structural Biology—Crystals of FabI were grown from hanging drops apparent Km for NADH of 15 6 1 mM, whereas the FabI[G93V]
(6 ml) containing 5 mg/ml FabI, 25 mM Tris, 150 mM NaCl, 0.5 mM exhibited an apparent Km for NADH of 8 6 2 mM. Thus, there
dithiothreitol, 5 mg/ml NADH, 50 mM sodium acetate, 4% polyethylene were not major differences between the kinetic constants for
glycol 4000 (w/v), pH 4.6. The drops were equilibrated against a reser- NADH in the wild-type and mutant protein. We also measured
voir (600 ml) containing 100 mM sodium acetate, pH 4.6, and 8% poly-
the apparent Km for the enoyl substrate using 4:1-NAC. This
ethylene glycol 4000 (w/v). Large hexagonal bipyrimidal crystals ap-
peared after equilibration for 1 day at 26 °C and reached dimensions of substrate analog was selected for these experiments because it
1.0 3 0.5 3 0.5 mm. had the higher solubility in the assay buffer than 8:1-NAC. The
Crystals of the FabI-NAD1-triclosan ternary complex were prepared apparent Km for 4:1-NAC was 25 6 5 mM for FabI compared
by soaking pregrown crystals of FabI in solutions containing 10 mg/ml with 15 6 2.2 mM for FabI[G93V]. The Vmax for FabI was 4.4
NAD1, 10 mg/ml triclosan, 100 mM sodium acetate, pH 4.6, and 10% mmol/min/mg, and for FabI[G93V], the Vmax was 3.6 mmol/min/
polyethylene glycol 4000 (w/v) for 7 days. X-ray diffraction data were
mg. These data illustrated that FabI and FabI[G93V] had
collected on the IMCA beamline (17-ID) at the Advanced Photon Source
at the Argonne National Laboratory. The crystals used for data collec- similar specific activities and substrate and cofactor binding
tion were first transferred to a cryoprotectant buffer containing 100 mM characteristics.
sodium acetate, pH 4.6, 10% polyethylene glycol 4000 (w/v), 25% glyc- Triclosan Inhibition of FabI and FabI[G93V]—The observa-
erol (v/v) for 1 h and then flash-frozen in a stream of liquid nitrogen. tion that cells expressing the FabI[G93V] protein have a min-
X-ray data to 1.8-Å resolution were collected at 100 K using synchrotron imum inhibitory concentration for triclosan that was 64-fold
X-radiation at a 1.0-Å wavelength and a Bruker 2 3 2 mosaic CCD x-ray
higher than the wild-type strain (0.5 compared with 32 mg/ml)
detector. The x-ray data were then processed using the SAINT (Bruker
x-ray, Madison, WI) program package. (22) suggested that the FabI[G93V] protein would be refractory
Crystals of the FabI-NAD1-triclosan complex were hexagonal, space to triclosan inhibition in vitro. We directly tested this assump-
group P6122 (number 178) with unit cell dimensions a 5 b 5 79.70 and tion using the spectrophotometric FabI assay and found that
c 5 327.7 Å and two molecules of the ternary complex in the asymmetric both FabI and FabI[G93V] were inhibited by triclosan (Fig. 1).
unit. These crystals were isomorphous to those of the FabI-NAD1- The IC50 for FabI (2 mM) was lower than the IC50 for FabI[G93V]
diazaborine complex reported by Baldock et al. (15). The FabI structure
(10 mM). This '5-fold increase in the IC50 for FabI[G93V] was
was initially refined as a rigid body using XPLOR, and the positions and
conformation of the bound NAD1 and triclosan molecules were deter- consistent with increased resistance of the enzyme to triclosan
mined from Fo-Fc difference electron density maps. In addition, the inhibition; however, the magnitude of this difference was sig-
positions of the 185 tightly bound water molecules were determined nificantly less than was predicted from the 64-fold difference in
from subsequent residual electron density maps. The entire structure the effectiveness of triclosan against the mutant strain.
containing 2 molecules of the FabI-NAD1-triclosan complex and 185 This unexpected discrepancy between the minimum inhibi-
water molecules was refined to a crystallographic R-factor of 0.232
tory concentration and IC50 data led us to examine the mech-
using data from 8.0 to 1.8 Å resolution. Residues 194 –210 were disor-
dered in the structure. Solvent-accessible surface areas were calculated anism of FabI inhibition by triclosan in more detail. The data in
using a 1.4-Å probe with the SOLVATION module of the InsightII Fig. 1 was obtained by starting the assay with the addition of
software (version 97.1, Molecular Simulations, Inc.). FabI and measuring the initial rate of NADH oxidation from
11112 FabI-NAD1 Triclosan Ternary Complex

FIG. 3. Triclosan-induced binding of [3H]NAD1 to FabI and


FabI[G93V]. The ability of FabI (E) or FabI[G93V] (●) to bind
[3H]NAD1 in the presence of 8 mM triclosan was determined by gel
filtration chromatography as described under “Experimental Proce-
dures.” The FabI proteins were located in the excluded volume of the
Sephadex G-25 column (fractions 12–17), and the free [3H]NAD1 eluted
in the included volume (fractions 20 –35).

binding of [3H]NAD1 to FabI[G93V] in the presence of triclosan


shown in Fig. 3 (not shown, see below). These data demon-
FIG. 2. Time-dependent inhibition of FabI and FabI[G93V] by
triclosan. The spectrophotometric assay described under “Experimen- strated that triclosan induced the high affinity binding of
tal Procedures” was used to determine the effect of triclosan on the NAD1 to FabI but not to FabI[G93V].
extended time course of the reaction FabI (panel A) or FabI[G93V] This concept was experimentally verified using a filter bind-
(panel B). The lines in the figures represent the continuous monitoring ing assay developed to quantitate the binding of [3H]NAD1 to
of the absorbance at 340 nm for 15 min. A triclosan concentration of 2
mM was used in panel A in the experiment with FabI, and a concentra- FabI in the presence and absence of triclosan. FabI exhibited
tion of 10 mM was used in panel B for the experiment with FabI[G93V]. saturable NAD1 binding in the presence of 4 mM triclosan (Fig.
These traces are representative of three independent measurements. 4A). An apparent association constant (K) of 2.5 mM–1 for NAD1
in the presence of triclosan was calculated by analysis of the
the linear portion of the trace within the first 1 min of the data using a double-reciprocal plot (Fig. 4B). There was no
reaction. Examination of the time course of the reaction over a indication of cooperative binding or the existence of more than
longer period of time (15 min) revealed a clear difference be- a single class of sites by analysis of the data using a Scatchard
tween the inhibition of FabI and FabI[G93V] by triclosan (Fig. plot (not shown). In contrast, we detected virtually no binding
2). Although there was an initial burst of activity in the pres- of [3H]NAD1 to FabI in the absence of triclosan (Fig. 4A). The
ence of triclosan, within 3 min of the start of the experiment, affinity of FabI for NAD1 was so low that we measured only
FabI catalysis essentially ceased (Fig. 2A). Thus, the inhibition trace levels of [3H]NAD1 binding to the free enzyme by this
of FabI by triclosan became irreversible as a function of time. technique. Therefore, we examined the effectiveness of NAD1
In contrast, triclosan inhibition of FabI[G93V] was consistent as a product inhibitor of FabI using the spectrophototmetric
throughout the 15-min time course, yielding a slower rate and assay. NAD1 was a poor inhibitor of the FabI reaction, with
no indication of increasing inhibition as a function of time (Fig. 50% inhibition occurring between 10 and 20 mM NAD1. These
2B). These data suggested that the difference between the data indicated a very low apparent affinity of FabI for NAD1 in
effectiveness of triclosan against wild-type strains compared the 10 –50 mM range (not shown). The filter binding assay was
with strain RJH108 (fabIG93V) was the ability of triclosan to next used to examine the effect of triclosan concentration on
irreversibly inhibit FabI, but not FabI[G93V], as a function of NAD1 binding to FabI (Fig. 5). The association of [3H]NAD1
time. with FabI was dependent on the concentration of triclosan and
Triclosan and NAD1 Binding to FabI—The data in Figs. 1 saturable (Fig. 5A). The apparent K for triclosan binding was
and 2 suggested that the irreversible inhibition of FabI by calculated as 4.2 mM–1 (Fig. 5B). Thus, in the presence of tri-
triclosan arose from the tight binding of the drug to the enzyme closan, the affinity of FabI for NAD1 increased by approxi-
in conjunction with another component in the assay. This pre- mately 4 orders of magnitude (.10 mM to 2.5 mM–1).
diction was experimentally tested by measuring the association Structure of the FabI-NAD1-triclosan Ternary Complex—
of [3H]NAD1 with either FabI or FabI[G93V] in the presence of The binding studies suggested that the basis for the potency of
triclosan by gel filtration chromatography (Fig. 3). FabI bound triclosan as a FabI inhibitor was because of the formation of a
approximately 80% of the [3H]NAD1 (0.2 mol/mol of FabI sub- high affinity FabI-NAD1-triclosan ternary complex. Determin-
unit) when triclosan was present, as evidenced from the shift in ing the structure of the FabI-NAD1-triclosan ternary complex
radioactivity from the included volume (fractions 18 – 40) to the by x-ray crystallography directly tested this idea (see “Experi-
excluded volume (fractions 11–17) of the column. In contrast, mental Procedures”). Triclosan bound noncovalently in a loca-
FabI[G93V] did not displace [3H]NAD1 from the included to tion adjacent to the nicotinamide portion of NAD1 (Fig. 6A).
the void volume of the column. The column was not equili- There was 494.7 Å2 of surface area buried between triclosan
brated with triclosan; thus the appearance of label associated and the protein and 238.5 Å2 of surface area buried at the
with FabI in the excluded volume indicated a stable association interface with NAD1. The diphenyl ether of triclosan adopted a
with the enzyme during the course of the separation (about 30 conformation with a dihedral angle of about 90° between the
min). The binding of [3H]NAD1 to the two enzymes in the two phenyl rings of the inhibitor. The 2-hydoxy-3-chlorophenyl
absence of triclosan was barely detectable and equivalent to the ring formed a parallel stack with the nicotinamide ring of the
FabI-NAD1 Triclosan Ternary Complex 11113

FIG. 4. Triclosan increases the affinity of FabI for NAD1. Panel FIG. 5. Triclosan-dependent [3H]NAD1 binding to FabI. Panel
A, binding of [3H]NAD1 to FabI in the presence (●) and absence (E) of A, dependence of [3H]NAD1 binding to FabI on the concentration of
4 mM triclosan. Panel B, double-reciprocal plot of the data in panel A. triclosan. Panel B, double-reciprocal plot of the data in panel A. The
The apparent NAD1 association constant (K) was 2.5 mM. The filter apparent triclosan association constant (K) was 4.2 mM. The filter bind-
binding assay was performed as described under “Experimental ing assay was performed as described under “Experimental Procedures.”
Procedures.”

co-factor with an interplanar stacking distance of 3.4 Å. In These residues are also not observed in the native structure of
addition to the nicotinamide ring, the hydroxychlorophenyl FabI, but the positions of these residues are evident in the
ring was surrounded by hydrophobic side chains on the protein FabI-NAD1-diazaborine structure (15). If the loop seen in the
including Tyr-146 and Tyr-156. The 29-hydroxyl group of tri- FabI-NAD1-diazaborine complex is modeled onto the triclosan
closan was involved in two strong hydrogen bonds. One bond complex structure, there are no conflicting contacts between
was with the 29-hydroxyl group of the nicotinamide ribose, and triclosan and the protein. In this model Ile-200 is in van der
the second was to the phenolic hydroxyl group of Tyr-156. The Waals contact with the 2,4-dichlorophenyl ring of triclosan, and
2,4-dichlorophenyl ring of triclosan was rotated 90° of the plane Phe-203 is in contact with the hydroxychlorophenyl group.
of the hydroxychlorophenyl group. The chlorine substituent in These data support the conclusion that triclosan forms a ter-
the 4-position of the phenyl ring accepted a hydrogen bond from nary complex that acts as a dead-end inhibitor that effectively
the backbone amide of Ala-95 and formed hydrophobic contacts removes protein from the catalytic cycle. These results lead to
with the side chain of Met-159. These data supported the the general conclusion that all 2-hydroxydiphenyl ether FabI
concept that the formation of a stable, ternary FabI-NAD1- inhibitors (22) block enzyme activity through the formation of
triclosan complex accounted for the effective inhibition of the noncovalent bisubstrate complexes with NAD1.
FabI reaction by this drug. Missense mutations, like FabI[G93V], confer resistance by
preventing the formation of the FabI-NAD1-triclosan ternary
DISCUSSION complex. Strains expressing either FabI[G93S] and
The formation of a ternary FabI-NAD1-triclosan complex FabI[G93V] have significantly elevated minimum inhibitory
accounts for the effectiveness of triclosan as an antibacterial concentrations for triclosan (21, 22). Although FabI[G93V] ac-
agent. Triclosan binds to the enoyl substrate site on FabI, and tivity is still inhibited at high triclosan concentrations (Fig.
tight binding of the drug requires interactions between both 2B), the drug is not capable of forming a stable complex with
the protein and the NAD1 cofactor (Fig. 6A). There is a strong the reductase. Triclosan is neither a dead-end inhibitor of
similarity between the mode of triclosan NAD1 binding to FabI FabI[G93V] (Fig. 2) nor does it form a FabI[G93V]-NAD1-
(Fig. 6A) and the structure of the FabI-NAD1-diazaborine com- triclosan ternary complex based on the inability of FabI[G93V]
plex described by Baldock et al. (15) (Fig. 6B). The diazaborine to bind [3H]NAD1 (Fig. 3). These biochemical results are inter-
compounds form covalent adducts with NAD1 at the FabI preted by modeling the Val-93 into the Gly-93 position in the
active site. The boron atom of the diazaborine is attached to the structure of the triclosan complex. Gly-93 lies on one side of the
29-hydroxyl of the NAD1 ribose, and this is the same 29-hy- substrate binding pocket, and the branched hydrophobic side
droxyl group that forms a hydrogen bond with the hydroxychlo- chain of Val protrudes into the pocket and occupies the same
rophenyl hydroxyl of triclosan. The FabI-NAD1-triclosan ter- space as the dichlorophenyl ring of triclosan (see Fig. 6). There-
nary complexes can be superimposed with the diazaborine fore, we conclude that the substitution of Gly-93 with bulkier
complex with a root mean square deviation of 0.3 Å for the amino acid residues imparts resistance to 2-hydroxydiphenyl
observed a-carbons. Residues 194 –210 were not well defined in ethers because of steric interference.
the electron density map of the FabI-NAD1-triclosan complex. Our identification of a ternary FabI-NAD1-triclosan complex
11114 FabI-NAD1 Triclosan Ternary Complex
that triclosan-resistant strains have missense mutations in the
fabI gene. The biochemical analysis of FabI inhibition by tri-
closan (22) and the delineation of its mechanism of binding
(this study) provide definitive evidence that enoyl-ACP reduc-
tase is the primary site for triclosan action. Thus, the reported
effects of triclosan on membrane structure and function are a
consequence of its specific inhibition of fatty acid biosynthesis
at the FabI step.
The ability of E. coli to acquire genetic resistance to triclosan
and related compounds through missense mutations in the fabI
gene suggests that the widespread use of this drug will lead to
the appearance of resistant organisms that will compromise
the usefulness of triclosan. The ubiquitous occurrence of type II
fatty acid synthase systems in bacteria and the essential na-
ture of the FabI reaction make this enzyme an attractive target
for antibacterial drugs. Accordingly, triclosan is effective
against a broad spectrum of bacteria (17), including multi-
drug-resistant Staphylococcus aureus (26, 27). The design and
development of second generation FabI inhibitors based on
their ability to form ternary FabI-NAD1-drug complexes will
supplement the arsenal against a broad spectrum of bacteria.

Acknowledgments—We thank Amy Sullivan and Magdalena


Kaminska for technical assistance, Rocco Gogliotti and John Domagala
for the trans-2-acyl-NAC substrate analogs, Tom Mueller and Craig
Banotai for the protein used in the crystal structure studies, Steve
VanderRoestand for the fabI clone, and our colleagues for their help in
editing the manuscript.

REFERENCES
1. Rock, C. O., Jackowski, S., and Cronan, J. E., Jr. (1996) in Biochemistry of
Lipids, Lipoproteins, and Membranes (Vance, D. E., and Vance, J. E., eds)
pp. 35–74, Elsevier Science Publishers B.V., Amsterdam
2. Rock, C. O., and Cronan, J. E., Jr. (1996) Biochim. Biophys. Acta 1302, 1–16
3. Cronan, J. E., Jr., and Rock, C. O. (1996) in Escherichia coli and Salmonella
typhimurium: Cellular and Molecular Biology (Neidhardt, F. C., Curtis, R.,
Gross, C. A., Ingraham, J. L., Lin, E. C. C., Low, K. B., Magasanik, B.,
Reznikoff, W., Riley, M., Schaechter, M., and Umbarger, H. E., eds) pp.
612– 636, American Society for Microbiology, Washington, D. C.
4. Zhang, Y., and Cronan, J. E., Jr. (1998) J. Bacteriol. 180, 3295–3303
FIG. 6. Structure of FabI-NAD1-inhibitor ternary complexes. 5. Heath, R. J., and Rock, C. O. (1996) J. Biol. Chem. 271, 27795–27801
Panel A, the active site region of the FabI-NAD1-triclosan ternary 6. Weeks, G., and Wakil, S. J. (1968) J. Biol. Chem. 243, 1180 –1189
complex. The hydroxychlorophenyl ring stacks with the nicotinamide 7. Heath, R. J., and Rock, C. O. (1995) J. Biol. Chem. 270, 26538 –26542
ring of the NAD1 with an interplanar distance of 3.4 Å and contacts 8. Heath, R. J., and Rock, C. O. (1996) J. Biol. Chem. 271, 1833–1836
Tyr-146 and Tyr-156 on the protein. The hydroxyl group of the ligand 9. Heath, R. J., and Rock, C. O. (1996) J. Biol. Chem. 271, 10996 –11000
forms hydrogen bonds with phenol of Tyr-156 and with the 29-hydroxyl 10. Jackowski, S. (1991) in Emerging Targets for Antibacterial and Antifungal
of the NAD1 ribose. The 2,4-dichlorophenyl ring of triclosan sits in a Chemotherapy (Sutcliffe, J. A., and Georgopapadakou, N. H., eds) pp.
hydrophobic pocket in contact with Met-159. The 4-chloro substituent 151–162, Chapman and Hall, New York
11. Bergler, H., Wallner, P., Ebeling, A., Leitinger, B., Fuchsbichler, S., Aschauer,
accepts a hydrogen bond from the amide backbone amide nitrogen of
H., Kollenz, G., Högenauer, G., and Turnowsky, F. (1994) J. Biol. Chem.
Ala-95. Panel B, binding of the diazaborine-NAD1 complex to FabI as
269, 5493–5496
described by Baldock et al. (15). The bicyclic ring of the diazaborine 12. Turnowsky, F., Fuchs, K., Jeschek, C., and Högenauer, G. (1989) J. Bacteriol.
compound stacks with the NAD1 nicotinamide ring in the same manner 171, 6555– 6565
as the hydroxychlorophenyl ring of triclosan. The diazaborine com- 13. Bergler, H., Högenauer, G., and Turnowsky, F. (1992) J. Gen. Microbiol. 138,
pounds form a covalent complex with NAD1, with the boron atom 2093–2100
covalently attached to the 29-hydroxyl of the NAD1 ribose. 14. Banerjee, A., Dubnau, E., Quémard, A., Balasubramanian, V., Um, K. S.,
Wilson, T., Collins, D., de Lisle, G., and Jacobs, W. R., Jr. (1994) Science
263, 227–230
reinforces the conclusion that FabI is a specific intracellular 15. Baldock, C., Rafferty, J. B., Sedeinikova, S. E., Baker, P. J., Stuitje, A. R.,
Slabas, A. R., Hawkes, T. R., and Rice, D. W. (1996) Science 274, 2107–2110
target for triclosan. Triclosan permeabilizes the bacterial en- 16. Rozwarski, D., Grant, G., Barton, D., Jacobs, W., and Sacchettini, J. C. (1998)
velope (18 –20), and these findings have justified the wide- Science 279, 98 –102
spread and increasing use of triclosan in consumer personal 17. Bhargava, H. N., and Leonard, P. A. (1996) Am. J. Infect. Control 24, 209 –218
18. Regös, J., Zak, O., Solf, R., Vischer, W. A., and Weirich, E. G. (1979) Derma-
care products based on the reasoning that bacteria would not
tologica (Basel) 158, 72–79
acquire resistance to a nonspecific membrane disrupter. How- 19. Vischer, W. A., and Regös, J. (1974) Zentbl. Bakt. Hyg. I. Abt. Orig. A 226,
ever, this type of evidence is not sufficient to conclude that a 376 –389
20. Regös, J., and Hitz, H. R. (1974) Zentbl. Bakt. Hyg. I. Abt. Orig. A 226,
compound has nonspecific membrane effects, because agents
390 – 401
that block fatty acid biosynthesis perturb membrane assembly 21. McMurray, L. M., Oethinger, M., and Levy, S. (1998) Nature 394, 531–532
by stopping phospholipid production. For example, the temper- 22. Heath, R. J., Yu, Y.-T., Shapiro, M. A., Olson, E., and Rock, C. O. (1998) J. Biol.
ature-sensitive fabI mutant was initially designated envM be- Chem. 273, 30316 –30320
23. Bradford, M. M. (1976) Anal. Biochem. 72, 248 –254
cause it was isolated through a genetic selection for strains 24. Bergler, H., Fuchsbichler, S., Högenauer, G., and Turnowsky, F. (1996) Eur.
with envelope-permeability defects (25). Also, the diazaborine J. Biochem. 242, 689 – 694
class of FabI inhibitors are known to perturb membrane func- 25. Egan, A. F., and Russell, R. R. B. (1973) Genet. Res. 21, 3603–3611
26. Bartzokas, C. A., Paton, J. H., Gibson, M. F., Graham, F., McLoughlin, G. A.,
tion as well (12). The first evidence that triclosan has a specific and Croton, R. S. (1984) N. Engl. J. Med. 311, 1422–1425
cellular target was provided by McMurray et al. (21) who found 27. Webster, J. (1992) J. Hosp. Infect. 21, 137–141

You might also like