You are on page 1of 15

Journal of Molecular Neuroscience

Copyright © 2003 Humana Press Inc.


All rights of any nature whatsoever reserved.
ISSN0895-8696/03/20:369–383/$25.00

ALZHEIMER’S THERAPEUTICS: Cognitive Enhancement

Acetylcholinesterase
A Multifaceted Target for Structure-Based Drug Design of Anticholinesterase
Agents for the Treatment of Alzheimer’s Disease

Harry M. Greenblatt,1 Hay Dvir,1,2 Israel Silman,2 and Joel L. Sussman*,1


Departments of 1Structural Biology and 2Neurobiology, Weizmann Institute of Science, Rehovot, Israel
Received October 15, 2002; Accepted March 24, 2003

Abstract
The structure of Torpedo californica acetylcholinesterase is examined in complex with several inhibitors that
are either in use or under development for treating Alzheimer’s disease. The noncovalent inhibitors vary greatly
in their structures and bind to different sites of the enzyme, offering many different starting points for future
drug design.

Index Entries: Alzheimer’s disease; anticholinesterase; cholinergic deficiency; drug design; Torpedo.

Introduction about the same time (Ollis et al., 1992), was a member
of a new family of proteins sharing a common fold,
Solution of the three-dimensional (3D) structure the α/β hydrolase fold, likewise opened up a fertile
of Torpedo californica acetylcholinesterase (TcAChE) field of research in which state-of-the-art techniques
in 1991 (Sussman et al.) opened up new horizons in of sequence alignment and homology modeling were
research on an enzyme that had already been the utilized (Cygler et al., 1993). The burgeoning number
subject of intensive investigation (Silman and Suss- of members of this new family resulted in founda-
man, 1998). tion of the ESTHER data base at Montpellier (http://
The unanticipated structure of this extremely www.montpellier.inra.fr:70/cholinesterase) (Cousin
rapid enzyme, in which the active site was found to et al., 1996), to collate and check the large amounts
be buried at the bottom of a deep and narrow gorge, of data accumulating, and to make them available
lined by aromatic residues, led to a revision of the and accessible to a large body of users. The fact that
views then held concerning substrate traffic, recog- the α/β hydrolase-fold family included several
nition, and hydrolysis (Botti et al., 1999). This led to members that lacked one or more of the residues in
a series of theoretical and experimental studies, the catalytic triad characteristic of the enzymes in
which took advantage of recent advances in theo- the family and had earlier been shown to be adhe-
retical techniques for treatment of proteins, such as sion proteins suggested alternative functions for
molecular dynamics and electrostatics, and of site- AChE. These, in turn, necessitated fresh approaches
directed mutagenesis, utilizing suitable expression toward analysis of its structure (Botti et al., 1998).
systems (Doctor et al., 1998). Finally, at a more practical level, the 3D structure
The realization that AChE, together with several itself, followed not long thereafter by the 3D struc-
other enzymes whose 3D structures were solved at tures of complexes with a number of ligands (Harel

*Author to whom all correspondence and reprint requests should be addressed. E-mail: Joel.Sussman@weizmann.ac.il

Journal of Molecular Neuroscience 369 Volume 20, 2003


370 Greenblatt et al.

et al., 1993), provided the basis for a rational struc- graphic studies of complexes of TcAChE with a reper-
ture-based approach to the design of drugs, such as toire of ligands (Harel et al., 1993, 1995, 1996; Raves
anti-Alzheimer’s disease (AD) medications (Fisher et al., 1997; Kryger et al., 1999); and (2) site-directed
et al., 1998), to the treatment of organophosphate mutagenesis (for literature, see Doctor et al., 1998).
(OP) intoxication (Millard and Broomfield, 1995), Crystallographic studies, using suitable quater-
and to the development of new classes of insecti- nary ligands (Harel et al., 1993, 1996), confirmed the
cides (Casida and Quistad, 1998), especially taken prediction mentioned above, made on the basis of
in conjunction with the recently solved 3D structures computerized docking of ACh, by clearly showing
of human AChE (hAChE) (Kryger et al., 1998a, 2000) that such ligands interact with Trp84 and, to a lesser
and the Drosophila melanogaster enzyme (DmAChE) degree, with Phe330. Elucidation of the structure of
(Harel et al., 2000). a complex with a powerful transition-state analog
In the following text, after giving a brief back- (Harel et al., 1996) also bore out the prediction of a
ground, we will analyze the binding modes of vari- three-pronged oxyanion hole (Sussman et al., 1991).
ous noncovalent and covalent inhibitors with TcAChE. Modeling studies (Harel et al., 1992) strongly sug-
gested that the acyl pocket for the acetyl group of
Three-Dimensional Structure ACh is provided by two more conserved aromatic
residues, Phe288 and Phe290, as was later confirmed
Crystal Structure by inspection of the complex with a transition-state
The 3D structure of TcAChE was solved in 1991 analog (Harel et al., 1996). Finally, inspection of the
(Sussman et al., 1991), subsequent to solubilization, structure of a complex with the elongated bisqua-
purification, and crystallization of the glycophos- ternary ligand, decamethonium, which was shown
phatidylinositol (GPI)-anchored dimer present in to bind along the active-site gorge, permitted iden-
large amounts in Torpedo electric organ (Sussman et tification of the peripheral anionic site at the entrance
al., 1988). Solution of the 3D structure indeed proved to the active-site gorge. The peripheral site involves
that the ChEs contain a catalytic triad, albeit with a three more conserved aromatic residues, Tyr70,
glutamate in place of the aspartate found in the serine Trp279, and Tyr121 (Harel et al., 1993).
proteases. However, the 3D structure displays a
number of unexpected features (Fig. 1). The active Correlation with Site-Directed
site is deeply buried, being located almost 20 Å from Mutagenesis Studies
the surface of the catalytic subunit, at the bottom of Extensive site-directed mutagenesis studies, car-
a long and narrow cavity. This cavity was named the ried out primarily by the laboratories of Shafferman
active-site gorge or, as >60% of its surface is lined by and Taylor (see, for example, Ordentlich et al., 1993;
the rings of conserved aromatic residues, the aro- Taylor & Radic, 1994; Shafferman et al., 1994, 1995)
matic gorge (Sussman et al., 1991; Axelsen et al., have supplemented the structural data. A complete
1994). Despite the prediction that the anionic site list of mutations is available from the ESTHER server
would contain several negative charges (Nolte et al., (see above). A few key issues addressed are as fol-
1980), in fact, only one negative charge is close to the lows. Site-directed mutagenesis of Phe288 and
catalytic site, that of Glu199, adjacent to the active- Phe290, the two aromatic residues shaping the acyl
site serine, Ser200. Based both on docking of ACh pocket, was shown to broaden specificity, thus
within the active site (Sussman et al., 1991) and on enabling AChE to hydrolyze butyrylcholine, nor-
affinity labeling (Weise et al., 1990), the quaternary mally only hydrolyzed by butyrylcholinesterase
group of ACh appears to be interacting, via a cation- (BChE), which on the basis of sequence alignment
π electron interaction (Dougherty and Stauffer, 1990; and molecular modeling, possesses a larger acyl
Dougherty, 1996), with the indole ring of one of the pocket in which these two aromatic residues are
conserved aromatic residues, Trp84. Another con- replaced by aliphatic residues (Harel et al., 1992;
served aromatic residue, Phe330, is also involved in Ordentlich et al., 1993; Vellom et al., 1993). Mutation
the interaction (Harel et al., 1993). of Trp84 to an aliphatic residue clearly demonstrated
the key role played by this residue in recognition of
Functional Binding Sites the quaternary group of ACh (Ordentlich et al., 1993).
The unexpected structure of the active site itself Mutation of the aromatic groups in the peripheral
and of the gorge leading to it have been explored site drastically reduced affinity for peripheral site
experimentally by two approaches: (1) crystallo- ligands, such as propidium (Harel et al., 1992; Radic

Journal of Molecular Neuroscience Volume 20, 2003


Acetylcholinesterase 371

Fig. 1. Ribbon diagram of the 3D structure of TcAChE. α-Helices are represented by helical coils, and β strands, by
arrows. The residues of the catalytic triad (Ser200, His440, Glu327) are rendered as bold sticks, and the arrow indicates
the entrance to the gorge. N and C termini are labeled.

et al., 1993). Radic and coworkers elegantly demon- mechanistic characteristics for phosphorylating or
strated the contribution of these residues to the bind- carbamylating the active site, with possible
ing of fasciculin (Radic et al., 1994). Thus, the triple enhancement of their affinity by recognition of suit-
mutant in which the residues in mouse AChE equiv- able groups within the substrate-binding site, such
alent to Tyr70, Trp84, and Tyr121 were all eliminated, as those that recognize the acyl group and quater-
bound fasciculin 108-fold more weakly than the wild- nary nitrogen of ACh. In contrast, the ligands that
type enzyme, with an affinity equal to that of BChE, inhibit ChEs reversibly are a heterogeneous class of
which lacks all three of these aromatics and, conse- molecules, some of which bear no obvious resem-
quently, also the peripheral site (Harel et al., 1992). blance to ACh. Consequently, if one is considering
Thus the site-directed mutagenesis studies, in gen- ligands of potential use as drugs or insecticides,
eral, confirmed the structural assignments, high- structure-based design may be essential, as QSAR
lighting the synergistic value of utilizing these or computerized docking of the ligand with the
experimental approaches in tandem. enzyme may not provide the correct orientation. This
was, indeed, the case for huperzine A(HupA), E2020,
and galanthamine (GAL), as will be discussed below.
3D Structures of Complexes with Drugs
In the following text, we will briefly review the
The drugs and toxins that display AChE action structures of conjugates with TcAChE of the anti-AD
can be broadly divided into two classes, those that drugs, Exelon™ (Bar-On et al., 1998), the physo-
bind noncovalently and reversibly, and those that stigmine analog, MF268 (Bartolucci et al., 1999),
form a covalent bond, including the OPs and car- (–)-HupA (Raves et al., 1997), its synthetic enan-
bamates, which either serve as slow substrates or tiomer (+)-HupA, HupB (Dvir et al., 2002a), the
inhibit irreversibly (Aldridge and Reiner, 1972). tacrine-huperzine hybrid, huprine X (Dvir et al.,
These latter reagents, which act by blocking the 2002a), and E2020 (Kryger et al., 1999). We will also
active-site serine, comprise a rather homogeneous present recent data on bifunctional compounds
class of inhibitors, as they effectively mimic the first derived from HupA (Wong et al., 2002), and exam-
stage of substrate hydrolysis (Silman et al., 1999). ine the complex with the AD drug, GAL (Greenblatt
Thus, they must display the correct steric and et al., 1999; Bartolucci et al., 2001; Pilger et al., 2001).

Journal of Molecular Neuroscience Volume 20, 2003


372 Greenblatt et al.

Conjugate with Rivastigmine (Exelon™)


The anti-AD drug, (+)S-N-ethyl-3-[(1-dimethyl-
amino)ethyl]-N-methylphenylcarbamate (rivastig-
mine [ENA-713]; see Fig. 2), belongs to a series of
miotine derivatives, all of which display inhibitory
action towards AChE both in vitro and in vivo (Wein-
stock et al., 1986). Compared to other clinically useful
carbamates, it has a longer duration of action in vivo
and preferentially inhibits AChE of the hippocam-
pus and cortex (Enz et al., 1993). It is already in use
for treatment of AD under the trade name of Exelon.
Rivastigmine carbamylates TcAChE very slowly,
whereas the bimolecular rate constant for inhibition
of hAChE is >1600-fold higher. Spontaneous reacti-
vation of both conjugates, is however, extremely
slow, as compared to most carbamyl-ChE conjugates,
with <10% reactivation being observed for TcAChE
after 48 h. The crystal structure of the conjugate of
rivastigmine with TcAChE (Bar-On et al., 1998)
showed that the carbamyl moiety, as expected, is
linked to the active-site serine, with the leaving
group, (–)-S-3-[1-(dimethylamino)ethyl]phenol
(NAP), being retained in the anionic site (Fig. 3). A
significant movement was observed of the active-
site histidine (H440) away from its normal hydrogen-
bonding partner, E327, to form a hydrogen bond
with E199, adjacent to the active-site serine, S200,
thus disrupting the catalytic triad. This disruption
Fig. 2. Chemical structures of inhibitors discussed in this may provide an explanation for the unusually slow
article. Note that NAP is the leaving group resulting from the kinetics of reactivation. It is of interest that a simi-
reaction of rivastigmine (ENA-713) with the active-site serine. lar disruption of the catalytic triad is observed in the

Journal of Molecular Neuroscience Volume 20, 2003


Acetylcholinesterase 373

Fig. 3. Active site of TcAChE after inhibition with rivastigmine. Amino acid residues within the active site of AChE
that may interact with the inhibitor are shown. Rivastigmine is depicted with larger spheres and thicker lines for empha-
sis. The carbamyl portion of rivastigmine is positioned to make two H-bonds (dashed lines) with the amide nitrogens of
A201 and G119, as well as nonbonded contacts (dotted lines) with F288 and F290 in the acyl pocket. NAP, the aromatic
leaving group of rivastigmine, remains in the active site of crystalline AChE. NAP is within H-bonding distance of three
water molecules (large red spheres), as well as of the amide nitrogen of G118. Nonbonding contacts and π–π interac-
tions with W84 and F330 are depicted as dotted lines.

crystal structure of the conjugate of TcAChE with not seen in the crystal structure, implying the exis-
the potent OP nerve agent, VX (Millard et al., 1999). tence of an alternative route, viz. a back door, for its
clearance.
Conjugate with the Physostigmine Analog
MF268 Complexes with the Huperzine
8-(cis-2,6-dimethylmorpholino)octylcarbamoyle- Family of Inhibitors
seroline (MF268; see Fig. 2) is a long-chain analog of (–)-HupA and (–)-HupB (Fig. 2) are two
physostigmine. Whereas AChE inhibited by Lycopodium alkaloids isolated from the herb Huperzia
physostigmine is reactivated quite rapidly, long- serrata, which grows mainly in hilly regions in the
chain physostigmine analogs form much more stable provinces of southern China (Liu et al., 1986). This
conjugates; and MF268 behaves in vitro as an almost herb has been used in traditional Chinese medicine
irreversible inhibitor (Perola et al., 1997). Bartolucci for centuries to treat contusion, strain, swelling,
and coworkers used X-ray crystallography to inves-
schizophrenia, and other ailments (Bai et al., 2000).
tigate this phenomenon by solving the crystal struc-
Both compounds are potent reversible inhibitors of
ture of the conjugate of MF268 with TcAChE. In the
crystal structure, the dimethylmorpholinooctylcar- AChE (Wang et al., 1986; Kozikowski et al., 1991).
bamic moiety of MF268 is covalently bound to the (–)-HupA, the most studied compound of the two,
active-site serine, Ser200. The alkyl group of the has already been approved as a drug for the symp-
inhibitor fills the upper part of the gorge, thus block- tomatic treatment of AD in China. A number of
ing the entrance to the active site. This prevents the studies, utilizing either computerized docking
leaving group, eseroline, from exiting by this route. techniques and/or site-directed mutagenesis
Surprisingly, however, the bulky eseroline moiety is (Ashani et al., 1994; Pang and Kozikowski, 1994b;

Journal of Molecular Neuroscience Volume 20, 2003


374 Greenblatt et al.

Saxena et al., 1994), predicted various possible ori- It was proposed that the peptide flip itself is respon-
entations for (–)-HupAwithin the active site of AChE. sible for the low on-rates observed for inhibition of
However, the crystal structure of TcAChE in com- AChE by (–)-HupA (Raves et al., 1997; Dvir et al.,
plex with (–)-HupA (Raves et al., 1997) showed an 2002a). Similarly it was hypothesized (Dvir et al.,
unexpected orientation for the inhibitor, involving 2002a) that its stabilization may contribute to the low
few strong direct interactions with the protein. rates of dissociation observed (Ashani et al., 1992).
The crystal structure of its analog, (–)-HupB, with Finally, modeling a tyrosine residue in place of
TcAChE (Dvir et al., 2002a) shows that the only sig- Phe330 in TcAChE (Dvir et al., 2002b) revealed the
nificant differences involve the interaction with the possibility of the primary amino group of (–)-HupA
extra ring of (–)-HupB and the ethylidene of (-)-HupA hydrogen bonding to the tyrosine hydroxyl. The 5- to
with the enzyme. Although (–)-HupB appears to 40-fold higher affinity of (–)-HupA for mammalian
make more intimate van der Waals or CH···π inter- AChE was attributed to this possible interaction, which
actions with Trp84 and Phe330, the latter utilizes its does not exist in the corresponding TcAChE complex.
ethylidene methyl group to make a C-H···O bond
Complex of TcAChE with Huprine X
with the main chain carbonyl of His440, which cannot
be formed with (–)-HupB. The twofold higher Superposition of the structures of TcAChE com-
potency of (–)-HupA for TcAChE (Dvir et al., 2002a) plexed with the two Alzheimer’s drugs, tacrine and
was attributed to this difference. (–)-HupA, shows that their binding sites are adja-
The effects of (–)-HupA and its synthetic enan- cent and overlap partially (Fig. 5). This suggested
tiomer, (+)-HupA, on the NMDA receptor and in that a combination of these two ligands might result
protecting against β-amyloid toxicity were shown in a hybrid compound with binding properties
to be similar. In contrast, the stereoselectivity of resembling those of both parents. Recently, the syn-
HupA for AChE is substantial, the natural com- thesis and pharmacological evaluation of a series of
pound, (–)-HupA, being almost 2 orders of magni- compounds that combine pharmacophores of
tude more potent than the synthetic (+)-HupA (–)-HupA and tacrine have been reported. Some of
(McKinney et al., 1991; Saxena et al., 1994). Never- these hybrids are potent inhibitors of AChE (Badia
theless, comparison of the crystal structures of the et al., 1998; Carlier et al., 1999b). Indeed certain
complexes of TcAChE with both enantiomers (Dvir hybrids, which contain the 4-aminoquinoline sub-
et al., 2002a) shows that the α-pyridone moieties of structure of tacrine and a carbobicyclic moiety resem-
both molecules are oriented very similarly and prac- bling that of (–)-HupAbut lack the ethylidene moiety
tically overlap. To accomplish this, (+)-HupA must of HupA, display more powerful anticholinesterase
be flipped relative to (–)-HupA, so that its ethyli- activity than either tacrine or (–)-HupA. These
dene group points away from the anionic site. hybrids were named huprines, and huprine X
Instead, the three-carbon bridge of (+)-HupA binds (Fig. 2), the most powerful of the series, inhibited
to the anionic site. This shows that the most power- hAChE with an inhibition constant of 26 pM, being
ful interactions made by HupA and its analogs can 180-fold more potent than (–)-HupA and 1200-fold
be attributed to their α-pyridone group. more potent than tacrine (Camps et al., 2000). The 3D
The most striking observation seen in the struc- structure of TcAChE in complex with huprine X (Dvir
tures of all three Hup complexes is the flip of the et al., 2002b) shows that the inhibitor binds to the
peptide bond between Gly117 and Gly118. This anionic site and also hinders access to the esteratic
conformational change appears to be induced by a site. Its aromatic portion occupies the same binding
carbonyl-carbonyl repulsion (Fig. 4). The new site as tacrine, stacking between the aromatic rings
conformation is stabilized by Gly117O making of Trp84 and Phe330, whereas the carbobicyclic unit
H-bonds with Gly119N and Ala201N, the other two occupies the same binding pocket as (–)-HupA
functional elements of the three-pronged oxyanion (Fig. 6). Its chlorine substituent was found to lie in a
hole. As a consequence, the former position of hydrophobic pocket interacting with the rings of the
Gly118N in the oxyanion hole is occupied by the car- aromatic residues Trp432 and Phe330 and with the
bonyl of Gly117. All three inhibitors would thus be methyl groups of Met436 and Ile439. Other steady-
expected to abolish hydrolysis of all ester substrates, state inhibition data (Dvir et al., 2002b) show that
even those that do not utilize the anionic site. huprine X binds to hAChE and TcAChE with Ki values

Journal of Molecular Neuroscience Volume 20, 2003


Acetylcholinesterase 375

Fig. 4. Effect of binding of (+)-HupA (light gray), (–)-HupA (dark gray), and (–)-HupB (medium gray) on the oxyanion
hole of TcAChE. The carbonyl group of all three huperzines binds in a similar position relative to the carbonyl of G117,
resulting in the close contact of ~2.8 Å shown (light gray dashed line), and presumably causes the observed conforma-
tional change.

of 0.67 and 0.13 nM, being 28- and 54-fold, respec- suggested that the relatively low association rate of
tively, more potent than tacrine. This difference stems huprine X may be caused by its slow diffusion rate
from the fact that the aminoquinoline moiety of down the active-site gorge owing to its bulky rigid
huprine X makes interactions similar to those made structure (Dvir et al., 2002b).
by tacrine, but additional interactions with the
enzyme are made by the huperzine-like substructure Complexes with Huperzine-Like
and the chlorine atom. Furthermore, both tacrine and Bivalent Compounds
huprine X bind more tightly to the Torpedo enzyme The fact that (–)-HupA is a natural product whose
than to hAChE, suggesting that their quinoline sub- chemical synthesis is relatively difficult, has dis-
structures interact better with Phe330 than with couraged its development as a prescription drug for
Tyr337, the corresponding residue in the hAChE treatment of AD. Simplified analogs of HupA, which
structure. Both (–)-HupAand huprine X display slow are more readily synthesized, have been used suc-
binding properties, but only binding of the former cessfully as building blocks for bis-acting compounds
causes a peptide flip of Gly117. Based on this, it was (Carlier et al., 1999a, 1999b; Han et al., 1999) as had

Journal of Molecular Neuroscience Volume 20, 2003


376 Greenblatt et al.

Fig. 5. Structure of TcAChE/(–)-HupA complex (dark gray) superimposed on the structure of TcAChE/tacrine (light
gray), showing the partially overlapping binding sites of these ligands.

been shown previously for bis-tacrine compounds Complex with E2020 (Aricept™)
(Pang and Kozikowski, 1994b). These compounds
are targeted to bind to both anionic sites (Trp84 and E2020 (Aricept™; Fig. 2), is a member of a large
Trp279) (Harel et al., 1993), with methylene linkers family of N-benzylpiperidine-based AChE
of varying lengths joining the two functional groups inhibitors, developed, synthesized, and evaluated
(Fig. 2) (Carlier et al., 2000). Despite the loss of bind- by the Eisai Company in Japan (Kawakami et al.,
ing affinity of the individual precursors because of 1996). It was approved for treatment of AD by the
their simplification, linkage of these building blocks FDA in 1996 (Nightingale, 1997) and shows high
results in compounds with greater affinity for AChE selectivity for AChE relative to BChE (Sugimoto et
than that of (–)-HupA. Thus, a linker of 10 methyl- al., 1995). The earlier modeling studies attributed
ene units yields E10E, which binds to TcAChE >170- this differential specificity to differences in geome-
fold more tightly than (–)-HupA. A linker of 12 units try within the active sites of AChE and BChE (Car-
yields a less effective compound, E12E, which has dozo et al., 1992). However, more recent studies,
~25-fold greater affinity than (–)-HupA. In rat AChE, carried out subsequent to determination of the 3D
however, the relative affinities are reversed; thus, structure of TcAChE, suggested that E2020 and other
E12E has greater affinity than E10E. The recently Eisai compounds orient along the active-site gorge
solved crystal structures of these compounds in and that the differential specificity can be attributed
TcAChE provide insight into the underlying causes to structural differences between AChE and BChE at
of this species specificity (Wong et al., 2002). the top of the gorge, at the peripheral anionic site

Journal of Molecular Neuroscience Volume 20, 2003


Fig. 6. Superposition of TcAChE/huprine X complex (dark gray) with TcAChE/tacrine complex (light gray) (A) and
TcAChE/(–)-HupA complex (light gray) (B).

Journal of Molecular Neuroscience Volume 20, 2003


378 Greenblatt et al.

Fig. 7. Binding mode of E2020 to TcAChE. E2020 is displayed as a ball-and-stick model. Water molecules are repre-
sented as light pink balls; standard H-bonds, as dotted orange lines; aromatic H-bonds, π-cation, and stacking interac-
tions, as dark lines. Note the multiple water-mediated contacts of E2020 atoms with the protein.

(Pang and Kozikowski, 1994a; Kawakami et al., with either the catalytic triad or the oxyanion hole,
1996). The 3D structure of the complex of E2020- but only indirectly, via solvent molecules. These loci,
TcAChE fully confirms these latter assignments (Fig. and a finger-shaped void toward the acyl pocket,
7) (Kryger et al., 1998b, 1999). It can be seen that provide spaces into which substituents could fit, thus
E2020 interacts with both the anionic subsite of the yielding analogs of E2020 with increased affinity and
active site, at the bottom of the gorge, and with the selectivity.
peripheral anionic site, near its top, via aromatic
stacking interactions with conserved aromatic Complex with GAL
residues. The piperidine nitrogen is also involved in (–)-Galanthamine (GAL) is a naturally occurring
a cation-π electron interaction with the phenyl ring alkaloid from the lily family of plants, in particular,
of Phe330. E2020 does not, however, interact directly the common snowdrop (Galanthus nivalis). Since its

Journal of Molecular Neuroscience Volume 20, 2003


Acetylcholinesterase 379

Fig. 8. Initial difference electron density map from the TcAChE/GAL complex. GAL is rendered as a ball-and-stick
model, and protein atoms are rendered as stick models.

discovery >45 years ago, it has been tested in vari- Theoretical predictions as to the binding mode of
ous neurological applications (Bretagne and Valletta, GAL to TcAChE did not produce unambiguous
1965; Gujral, 1965; Bystrzanowska, 1969; Harvey, models; indeed, the most favored model (R. Fletcher,
1995). GAL has an IC50 value of 0.35 µM for human R. Viner, and T. Lewis, unpublished results) is not
erythrocyte AChE (Thomsen and Kewitz, 1990) and in agreement with the experimentally determined
an IC50 value of 0.652 µM for TcAChE (T. Lewis and 3D structure.
C. Personeni, unpublished results). It also acts as a The crystal structure of the TcAChE/GAL com-
nicotinic activator of both ganglionic and muscle plex (Greenblatt et al., 1999; Bartolucci et al., 2001)
receptors (Storch et al., 1995; Schrattenholz et al., reveals that the inhibitor is bound at the bottom of
1996), and of nicotinic receptors in the brain (Pereira the gorge and interacts with both Trp84 and the acyl
et al., 1993). In fact, evidence indicates that the effect binding pocket (Fig. 8). Only one direct hydrogen
on the receptor also plays a role in cognitive improve- bond is formed between the inhibitor and protein
ment (Santos et al., 2002). GAL is now in use for treat- side chains: between the hydroxyl group of the
ment of AD in Europe and the United States. inhibitor and Glu199 Oε1 (2.7 Å). A second interac-

Journal of Molecular Neuroscience Volume 20, 2003


380 Greenblatt et al.

tion is possible between the hydrogen of Ser200 Oγ Bartolucci C., Perola E., Cellai L., Brufani M., and Lamba
and the O-methoxy group of GAL, but this would at D. (1999) “Back door” opening implied by the crystal
best be a bifurcated interaction, with the second structure of a carbamoylated acetylcholinesterase. Bio-
chemistry 38, 5714–5719.
acceptor being His440 N ε2 of the catalytic triad.
Bartolucci C., Perola E., Pilger C., Fels G., and Lamba D.
The rest of the possible hydrogen bonds involve water (2001) Three-dimensional structure of a complex of
molecules bound to the enzyme. The reasonably tight galanthamine (Nivalin ®) with acetylcholinesterase
association observed for this inhibitor to AChE pre- from Torpedo californica: Implications for the
sumably gains a significant contribution from favor- design of new anti-Alzheimer drugs. Proteins 42,
able entropy contributions owing to the rigid nature 182–191.
of the inhibitor and displacement of some bound Botti S. A., Felder C., Lifson S., Sussman J. L., and Silman
water molecules present in the native structure. I. (1999) A modular treatment of molecular traffic
through the active site of cholinesterases. Biophys. J. 77,
2430–2450.
Acknowledgments Botti S. A., Felder C. E., Sussman J. L., and Silman I. (1998)
This work was supported by the U.S. Army Med- Electrotactins: a class of adhesion proteins with con-
ical Research Acquisition Activity under contract no. served electrostatic and structural motifs. Protein Eng.
11, 415–420.
DAMD17-97-2-7022, the EC IVth and Vth Frame- Bretagne M. and Valletta J. (1965) Clinical trials of a new
work Quality of Life Programs, the Kimmelman anticholinesterase agent, galanthamine, in anesthesi-
Center for Biomolecular Structure and Assembly, ology. Anesth. Analg. (Paris) 22, 285–292.
Israel, the Nella and Leon Benoziyo Center for Bystrzanowska T. (1969) Trials in the treatment of
Neurosciences. I.S. is Bernstein-Mason Professor of facial nerve paralysis using nivaline. Wiad. Lek. 22,
Neurochemistry. J.L.S. is the Morton and Gladys 1233–1239.
Pickman Professor of Structural Biology. Camps P., Cusack B., Mallender W. D., Achab R. E., Morral
J., Munoz-Torrero D., and Rosenberry T. L. (2000)
Huprine X is a novel high-affinity inhibitor of acetyl-
References cholinesterase that is of interest for treatment of
Aldridge W. N. and Reiner E. (1972) Enzyme Inhibitors as Alzheimer’s disease. Mol. Pharmacol. 57, 409–417.
Substrates: Interactions of Esterases with Esters of Cardozo M. G., Kawai T., Imura Y., Sugimoto H., Yaman-
Organophosphorus and Carbamic Acids, Neuberger A. ishi Y., and Hopfinger A. J. (1992) Conformational analy-
and Tatum E. L., eds., North-Holland Publishing, ses and molecular-shape comparisons of a series of
Amsterdam. indanone-benzylpiperidine inhibitors of acetyl-
Ashani Y., Grunwald J., Kronman C., Velan B., and Shaf- cholinesterase. J. Med. Chem. 35, 590–601.
ferman A. (1994) Role of tyrosine 337 in the binding of Carlier P. R., Chow E. S., Han Y., Liu J., El Yazal J.,
huperzine A to the active site of human acetyl- and Pang Y. P. (1999a) Heterodimeric tacrine-
cholinesterase. Mol. Pharmacol. 45, 555–560. based acetylcholinesterase inhibitors: investigating
Ashani Y., Peggins J. O. III, and Doctor B. P. (1992) Mech- ligand-peripheral site interactions. J. Med. Chem. 42,
anism of inhibition of cholinesterases by huperzine A. 4225–4231.
Biochem. Biophys. Res. Commun. 184, 719–726. Carlier P. R., Du D. M., Han Y., Liu J., and Pang Y. P. (1999b)
Axelsen P. H., Harel M., Silman I., and Sussman J. L. (1994) Potent, easily synthesized huperzine A-tacrine hybrid
Structure and dynamics of the active site gorge of acetyl- acetylcholinesterase inhibitors. Bioorg. Med. Chem. Lett.
cholinesterase: synergistic use of molecular dynamics 9, 2335–2338.
simulation and X-ray crystallography. Protein Sci. 3, Carlier P. R., Du D. M., Han Y. F., Liu J., Perola E., Williams
188–197. I. D., and Pang Y. P. (2000) Dimerization of an inactive
Badia A., Baños J. E., Camps P., Contreras J., Görbig D. M., fragment of huperzine A produces a drug with twice
Muñoz-Torrero D., et al. (1998) Synthesis and evalua- the potency of the natural product. Angew. Chem. Int.
tion of tacrine-huperzine A hybrids as acetyl- Ed. Engl. 39, 1775–1777.
cholinesterase inhibitors of potential interest for the Casida J. E. and Quistad G. B. (1998) Golden age of insec-
treatment of Alzheimer’s disease. Bioorg. Med. Chem. ticide research: past, present, or future? Annu. Rev. Ento-
6, 427–440. mol. 43, 1–16.
Bai D. L., Tang X. C., and He X. C. (2000) Huperzine A: a Cousin X., Hotelier T., Lievin P., Toutant J.-P., and Cha-
potential therapeutic agent for treatment of tonnet A. (1996) A CHOLINESTERASE GENES
Alzheimer’s disease. Curr. Med. Chem. 7, 355–376. SERVER (ESTHER): A data base of cholinesterase-
Bar-On P., Harel M., Millard C. B., Enz A., Sussman J. L., related sequences for multiple alignments, phyloge-
and Silman I. (1998) Kinetic and structural studies on netic relationships, mutations and structural data
the interaction of the anti-Alzheimer drug, ENA-713, retrieval. Nucleic Acids Res. 24, 132–136.
with Torpedo californica acetylcholinesterase. J. Physiol. Cygler M., Schrag J. D., Sussman J. L., Harel M., Silman
(Paris) 92, 406–407. I., Gentry M. K., and Doctor B. P. (1993) Relationship

Journal of Molecular Neuroscience Volume 20, 2003


Acetylcholinesterase 381

between sequence conservation and three-dimensional Harel M., Schalk I., Ehret-Sabatier L., Bouet F., Goeldner
structure in a large family of esterases, lipases, and M., Hirth C., et al. (1993) Quaternary ligand binding to
related proteins. Protein Sci. 2, 366–382. aromatic residues in the active-site gorge of acetyl-
Doctor B. P., Taylor P., Quinn D. M., Rotundo R. L., and cholinesterase. Proc. Natl. Acad. Sci. U.S.A. 90, 9031–9035.
Gentry M. K. (1998) Structure and Function of Harel M., Sussman J. L., Krejci E., Bon S., Chanal P., Mas-
Cholinesterases and Related Proteins, Plenum, New York. soulié J., and Silman I. (1992) Conversion of acetyl-
Dougherty D. A. (1996) Cation-π interactions in chemistry cholinesterase to butyrylcholinesterase: modeling
and biology: a new view of benzene, phe, tyr, and trp. and mutagenesis. Proc. Natl. Acad. Sci. U.S.A. 89,
Science 271, 163–168. 10,827–10,831.
Dougherty D. A. and Stauffer D. A. (1990) Acetylcholine Harvey A. L. (1995) The pharmacology of galanthamine
binding by a synthetic receptor: implications for bio- and its analogues. Pharmacol. Ther. 68, 113–128.
logical recognition. Science 250, 1558–1560. Kawakami Y., Inoue A., Kawai T., Wakita M., Sugimoto
Dvir H., Jiang H. L., Wong D. M., Harel M., Chetrit M., H., and Hopfinger A. J. (1996) The rationale for E2020
He X. C., et al. (2002a) X-ray structures of Torpedo as a potent acetylcholinesterase inhibitor. Bioorg. Med.
californica acetylcholinesterase complexed with Chem. Lett. 4, 1429–1446.
(+)-Huperzine A and (–)-Huperzine B: Structural evi- Kozikowski A. P., Xia Y., Reddy E. R., Tuckmantel W.,
dence for an active site rearrangement. Biochemistry Hanin I., and Tang X. C. (1991). Synthesis of huperzine
41, 10,810–10,818. A and its analogues and their anticholinesterase activ-
Dvir H., Wong D. M., Harel M., Barril X., Orozco M., Luque ity. J. Org. Chem. 56, 4636–4645.
F. J., et al. (2002b) 3D structure of Torpedo californica Kryger G., Giles K., Harel M., Toker L., Velan B., Lazar A.,
acetylcholinesterase complexed with Huprine X at et al. (1998a) 3D structure at 2.7 Å resolution of native
2.1 Å resolution: kinetic and molecular dynamic corre- and G202P human acetylcholinesterase complexed
lates. Biochemistry 41, 2970–2981. with fasciculin II, in Structure and Function of Cholin-
Enz A., Amstutz R., Boddeke H., Gmelin G., and esterases and Related Proteins (Doctor B. P., Taylor P.,
Malanowski J. (1993) Brain selective inhibition of Quinn D. M., Rotundo R. L., and Gentry M. K., eds.),
acetylcholinesterase: a novel approach to therapy for Plenum, New York, pp. 323–326.
Alzheimer’s disease. Prog. Brain Res. 98, 431–438. Kryger G., Harel M., Giles K., Toker L., Velan B., Lazar A.,
Fisher A., Hanin I., and Yoshida M. (1998) Progress et al. (2000) Structures of recombinant native and E202Q
in Alzheimer’s and Parkinson’s Diseases, Plenum Press, mutant human acetylcholinesterase complexed with
New York. the snake-venom toxin fasciculin-II. Acta Crystallogr. D
Greenblatt H. M., Kryger G., Lewis T., Silman I., and Suss- 56, 1385–1394.
man J. L. (1999) Structure of acetylcholinesterase com- Kryger G., Silman I., and Sussman J. L. (1998b) Three-
plexed with (–)-galanthamine at 2.3 Å resolution. FEBS dimensional structure of a complex of E2020 with
Lett. 463, 321–326. acetylcholinesterase from Torpedo californica. J. Physiol.
Gujral V. V. (1965) Nivalin in the treatment of residual (Paris) 92, 191–194.
post-polio paralysis and pseudo-hypertrophic muscu- Kryger G., Silman I., and Sussman J. L. (1999) Structure
lar dystrophy. Indian Pediatr. 2, 89–93. of acetylcholinesterase complexed with E2020 (Ari-
Han Y. F., Li C. P., Chow E., Wang H., Pang Y. P., and Car- cept ® ): implications for the design of new anti-
lier P. R. (1999) Dual-site binding of bivalent 4-amino- Alzheimer drugs. Structure 7, 297–307.
pyridine- and 4-aminoquinoline- based AChE Liu J.-S., Zhu Y.-L., Yu C.-M., Zhou Y.-Z., Han Y.-Y., Wu
inhibitors: contribution of the hydrophobic alkylene F.-W., and Qi B.-F. (1986) The structures of huperzine
tether to monomer and dimer affinities. Bioorg. Med. A and B, two new alkaloids exhibiting marked anti-
Chem. Lett. 7, 2569–2575. cholinesterase activity. Can. J. Chem. 64, 837–839.
Harel M., Kleywegt G. J., Ravelli R. B. G., Silman I., and McKinney M., Miller J. H., Yamada F., Tuckmantel W., and
Sussman J. L. (1995) Crystal structure of an acetyl- Kozikowski A. P. (1991) Potencies and stereoselectvities
cholinesterase-fasciculin complex: interaction of a of enantiomers of huperzine A for inhibition of rat cor-
three-fingered toxin from snake venom with its target. tical acetylcholinesterase. Eur. J. Pharmacol. 203, 303–305.
Structure 3, 1355–1366. Millard C. B. and Broomfield C. A. (1995) Anti-
Harel M., Kryger G., Rosenberry T. L., Mallender cholinesterases: medical applications of neurochemi-
W. D., Lewis T., Fletcher R. J., et al. (2000) Three cal principles. J. Neurochem. 64, 1909–1918.
dimensional structures of Drosophila melanogaster Millard C. B., Koellner G., Ordentlich A., Shafferman A.,
acetylcholinesterase and of its complexes with two Silman I., and Sussman J. L. (1999) Reaction products
potent inhibitors. Protein Sci. 9, 1063–1072. of acetylcholinesterase and VX reveal a mobile histi-
Harel M., Quinn D. M., Nair H. K., Silman I., dine in the catalytic triad. J. Am. Chem. Soc. 121,
and Sussman J. L. (1996) The X-ray structure of a 9883–9884.
transition state analog complex reveals the mole- Nightingale S. L. (1997) Donepezil approved for treatment
cular origins of the catalytic power and substrate speci- of Alzheimer’s disease. JAMA 277, 10.
ficity of acetylcholinesterase. J. Am. Chem. Soc. 118, Nolte H.-J., Rosenberry T. L., and Neumann E. (1980)
2340–2346. Effective charge on acetylcholinesterase active sites

Journal of Molecular Neuroscience Volume 20, 2003


382 Greenblatt et al.

determined from the ionic strength dependence of of neuronal nicotinic acetylcholine receptors are poten-
association rate constants with cationic ligands. tiated by a novel class of allosterically acting ligands.
Biochemistry 19, 3705–3711. Mol. Pharmacol. 49, 1–6.
Ollis D. L., Cheah E., Cygler M., Dijkstra B., Frolow F., Shafferman A., Kronman C., and Ordentlich A. (1995)
Franken S. M., et al. (1992) The α/β hydrolase fold. Compilation of evaluated mutants of cholinesterases,
Protein Eng. 5, 197–211. in Enzymes of the Cholinesterase Family (Balasubraman-
Ordentlich A., Barak D., Kronman C., Flashner Y., Leitner ian A. L., Doctor B. P., Taylor P., and Quinn D. M., eds.),
M., Segall Y., et al. (1993) Dissection of the human acetyl- Plenum Press, New York, pp. 481–488.
cholinesterase active center—determinants of substrate Shafferman A., Ordentlich A., Barak D., Kronman C., Ber
specificity—identification of residues constituting the R., Bino T., et al. (1994) Electrostatic attraction by sur-
anionic site, the hydrophobic site, and the acyl pocket. face charge does not contribute to the catalytic effi-
J. Biol. Chem. 268, 17,083–17,095. ciency of acetylcholinesterase. EMBO J. 13, 3448–3455.
Pang Y.-P. and Kozikowski A. (1994a) Prediction of the Silman I. and Sussman J. L. (1998) Structure and functional
binding site of 1-benzyl-4-[(5,6-dimethoxy-1-indanon- studies on acetylcholinesterase: a perspective, in Struc-
2-yl)methyl] piperidine in acetylcholinesterase by ture and Function of Cholinesterases and Related Proteins
docking studies with the SYSDOC program. J. Comput. (Doctor B. P., Taylor P., Quinn D. M., Rotundo R. L., and
Aided Mol. Des. 8, 683–693. Gentry M. K., eds.), Plenum, New York, pp. 25–33.
Pang Y.-P. and Kozikowski A. (1994b) Prediction of the Silman I., Millard C. B., Ordentlich A., Greenblatt H. M.,
binding sites of huperzine A in acetylcholinesterase by Harel M., Barak D., et al. (1999) A preliminary com-
docking studies. J. Comput. Aided Mol. Des. 8, 669–681. parison of structural models for catalytic intermedi-
Pereira E. F., Reinhardt-Maelicke S., Schrattenholz A., ates of acetylcholinesterase. Chem. Biol. Interact.
Maelicke A., and Albuquerque E. X. (1993) Identification 119–120, 43–52.
and functional characterization of a new agonist site on Storch A., Schrattenholz A., Cooper J. C., Abdel Ghani E.
nicotinic acetylcholine receptors of cultured hippocam- M., Gutbrod O., Weber K. H., et al. (1995) Physostig-
pal neurons. J. Pharmacol. Exp. Ther. 265, 1474–1491. mine, galanthamine and codeine act as ‘noncompeti-
Perola E., Cellai L., Lamba D., Filocamo L., and Brufani M. tive nicotinic receptor agonists’ on clonal rat
(1997) Long chain analogs of physostigmine pheochromocytoma cells. Eur. J. Biochem. 290, 207–219.
as potential drugs for Alzheimer ’s disease: new Sugimoto H., Iimura Y., Yamanishi Y., and Yamatsu K.
insights into the mechanism of action in the inhibition (1995) Synthesis and structure-activity-relationships
of acetylcholinesterase. Biochim. Biophys. Acta 1343, 41–50. of acetylcholinesterase inhibitors—1-benzyl-4-
Pilger C., Bartolucci C., Lamba D., Tropsha A., and Fels [(5,6-dimethoxy-1-oxoindan-2-yl)methyl]piperidine
G. (2001) Accurate prediction of the bound conforma- hydrochloride and related compounds. J. Med. Chem.
tion of galanthamine in the active site of Torpedo cal- 38, 4821–4829.
ifornica acetylcholinesterase using molecular docking. Sussman J. L., Harel M., Frolow F., Oefner C., Goldman
J. Mol. Graphics Modelling 19, 288–296, 374–378. A., Toker L., and Silman I. (1991) Atomic structure of
Radic Z., Durán R., Vellom D. C., Li Y., Cerveñansky C., acetylcholinesterase from Torpedo californica: a proto-
and Taylor P. (1994) Site of fasciculin interaction with typic acetylcholine-binding protein. Science 253,
acetylcholinesterase. J. Biol. Chem. 269, 11,233–11,239. 872–879.
Radic Z., Pickering N. A., Vellom D. C., Camp S., and Sussman J. L., Harel M., Frolow F., Varon L., Toker L.,
Taylor P. (1993) Three distinct domains in the Futerman A. H., and Silman I. (1988) Purification and
cholinesterase molecule confer selectivity for acetyl- crystallization of a dimeric form of acetylcholinesterase
and butyrylcholinesterase inhibitors. Biochemistry 32, from Torpedo californica subsequent to solubilization
12,074–12,084. with phosphatidylinositol-specific phospholipase C. J.
Raves M. L., Harel M., Pang Y.-P., Silman I., Kozikowski Mol. Biol. 203, 821–823.
A. P., and Sussman J. L. (1997) 3D structure of acetyl- Taylor P. and Radic Z. (1994) The cholinesterases: from
cholinesterase complexed with the nootropic alkaloid, genes to proteins. Annu. Rev. Pharmacol. Toxicol. 34,
(–)-huperzine A. Nat. Struct. Biol. 4, 57–63. 281–320.
Santos M. D., Alkondon M., Pereira E. F., Aracava Y., Thomsen T. and Kewitz H. (1990) Selective inhibition of
Eisenberg H. M., Maelicke A., and Albuquerque E. X. human acetylcholinesterase by galanthamine in vitro
(2002) The nicotinic allosteric potentiating ligand galan- and in vivo. Life Sci. 46, 1553–1558.
tamine facilitates synaptic transmission in the mam- Vellom D. C., Radic Z., Li Y., Pickering N. A., Camp S.,
malian central nervous system. Mol. Pharmacol. 61, and Taylor P. (1993) Amino acid residues controlling
1222–1234. acetylcholinesterase and butyrylcholinesterase speci-
Saxena A., Qian N., Kovach I. M., Kozikowski A. P., Pang ficity. Biochemistry 32, 12–17.
Y. P., Vellom D. C., et al. (1994) Identification of amino Wang Y. E., Yue D. X., and Tang X. C. (1986) Anti-
acid residues involved in the binding of huperzine A cholinesterase activity of Huperzine A. Acta Pharma-
to cholinesterases. Protein Sci. 3, 1770–1778. col. Sinica 7, 110–113.
Schrattenholz A., Pereira E. F., Roth U., Weber K. H., Albu- Weinstock M., Razin M., Chorev M., and Tashma Z. (1986)
querque E. X., and Maelicke A. (1996) Agonist responses Pharmacological activity of novel acetylcholinesterase

Journal of Molecular Neuroscience Volume 20, 2003


Acetylcholinesterase 383

agents of potential use in the treatment of Alzheimer’s with the cationic reagent N,N-dimethyl-2-phenyl-
disease, in Advances in Behavioral Biology (Fisher A., aziridinium. EMBO J. 9, 3885–3888.
Hanin I., and Lachman P., eds.), Plenum Press, New Wong D. M., Greenblatt H. M., Dvir H., Carlier P. R., Han
York, pp. 539–549. Y.-F., Pang Y.-P., et al. (2003) Acetylcholinesterase com-
Weise C., Kreienkamp H.-J., Raba R., Pedak A., Aaviksaar plexed with bivalent ligands related to Huperzine A:
A., and Hucho F. (1990) Anionic subsites of the acetyl- experimental evidence for species-dependent protein-
cholinesterase from Torpedo californica: affinity labelling ligand complementarity. J. Am. Chem. Soc. 125, 363–373.

Journal of Molecular Neuroscience Volume 20, 2003

You might also like