ArticlePDF Available

Crystal Structures of Human Cholinesterases in Complex with Huprine W and Tacrine: Elements of Specificity for Anti-Alzheimer's Drugs Targeting Acetyl- and Butyrylcholinesterase.

Authors:
  • Ministère des Armées

Abstract and Figures

The multifunctional nature of Alzheimer's disease calls for multi-target directed ligands (MTDLs) able to act on different components of the pathology, like the cholinergic dysfunction and amyloid aggregation. Such MTDLs are usually based on cholinesterase inhibitors (e.g., tacrine or huprine) coupled to another active molecule aimed at a different target. To aid in the design of these MTDLs we report the crystal structures of human acetylcholinesterase (hAChE) in complex with FAS-2 and a hydroxylated derivative of huprine (huprine W), and of human butyrylcholinesterase (hBChE) in complex with tacrine. Huprine W in hAChE and tacrine in hBChE reside in strikingly similar positions highlighting the conservation of key interactions, namely, π-π/cation-π interactions with Trp86(82), and hydrogen bonding with the main chain carbonyl of the catalytic histidine. Huprine W forms additional interactions with hAChE which explains its superior affinity: the isoquinoline moiety is associated with a group of aromatic residues (Tyr337, Phe338 and Phe295 not present in hBChE) in addition to Trp86; the hydroxyl group is hydrogen bonded to both the catalytic serine and residues in the oxyanion hole; and the chlorine substituent is nested in a hydrophobic pocket interacting strongly with Trp439. There is no pocket in hBChE able to accommodate the chlorine substituent.
Content may be subject to copyright.
Biochem. J. (2013) 453, 393–399 (Printed in Great Britain) doi:10.1042/BJ20130013 393
Crystal structures of human cholinesterases in complex with huprine W
and tacrine: elements of specificity for anti-Alzheimer’s drugs targeting
acetyl- and butyryl-cholinesterase
Florian NACHON*1, Eug´
enie CARLETTI, Cyril RONCO, Marie TROVASLET, Yvain NICOLET, Ludovic JEANand
Pierre-Yves RENARD
*D´
epartement de Toxicologie, Institut de Recherche Biom´
edicale des Arm´
ees, 24 Avenue des Maquis du Gr´
esivaudan, BP87, 38702 La Tronche, France, Normandie Universit´
e,
COBRA, UMR 6014 and FR 3038; Universit´
e Rouen, INSA Rouen; CNRS, 1 rue Tesni`
ere, 76821 Mont-Saint-Aignan, Cedex, France, and Laboratoire de Cristallographie et
Cristallogen`
ese des Prot´
eines, Institut de Biologie Structurale ‘J.P. Ebel’, CEA, CNRS, Universit´
e Joseph Fourier, 41 rue J. Horowitz, 38027 Grenoble, France
The multifunctional nature of Alzheimer’s disease calls for
MTDLs (multitarget-directed ligands) to act on different
components of the pathology, like the cholinergic dysfunction
and amyloid aggregation. Such MTDLs are usually on the basis
of cholinesterase inhibitors (e.g. tacrine or huprine) coupled with
another active molecule aimed at a different target. To aid in
the design of these MTDLs, we report the crystal structures
of hAChE (human acetylcholinesterase) in complex with FAS-
2 (fasciculin 2) and a hydroxylated derivative of huprine (huprine
W), and of hBChE (human butyrylcholinesterase) in complex
with tacrine. Huprine W in hAChE and tacrine in hBChE reside
in strikingly similar positions highlighting the conservation of
key interactions, namely, π-π/cation-πinteractions with Trp86
(Trp82), and hydrogen bonding with the main chain carbonyl
of the catalytic histidine residue. Huprine W forms additional
interactions with hAChE, which explains its superior affinity:
the isoquinoline moiety is associated with a group of aromatic
residues (Tyr337,Phe
338 and Phe295 not present in hBChE) in
addition to Trp86; the hydroxyl group is hydrogen bonded to both
the catalytic serine residue and residues in the oxyanion hole;
and the chlorine substituent is nested in a hydrophobic pocket
interacting strongly with Trp439. There is no pocket in hBChE that
is able to accommodate the chlorine substituent.
Key words: acetylcholinesterase, Alzheimer’s disease, bu-
tyrylcholinesterase, huprine, inhibitor, tacrine, X-ray structure.
INTRODUCTION
AD (Alzheimer’s disease), which is the most common cause of
senile dementia, is a major public health issue with devastating
economic and human impacts. As of 2011, the estimated number
of patients needing treatment is 7–8 million in Europe and 4–
5 million in the U.S.A., with a total of 24 million worldwide.
Considering the aging of the world’s population, the situation
is expected to worsen, with the predicted burden of AD victims
reaching 42 million by 2020 [1].
AD results from a neurodegenerative process occurring in
the central nervous system. It is clinically characterized by
a loss of memory and cognition that is associated with
deterioration of the basal forebrain cholinergic neuronal network,
resulting in a reduction in the level of the neurotransmitter
acetylcholine [2]. AD is histologically characterized by aberrant
proteinaceous deposits: (i) from Aβ(β-amyloid) peptide outside
neurons; and (ii) from microtubule-associated Tau protein inside
neurons. The aetiology of AD is not completely understood,
yet it is clearly complex and multifactorial. Although several
treatment strategies have been proposed [3,4], most current
therapeutic options involve restoring acetylcholine levels in the
brain. hAChE [human AChE (acetylcholinesterase)] inhibitors
donepezil (Aricept®), rivastigmine (Exelon®) and galanthamine
(Reminyl®) are currently approved anti-AD drugs [5]. In previous
studies, BChE (butyrylcholinesterase) received attention owing to
its role in modulating acetylcholine levels in cholinergic neurons
under normal conditions [6] and when AChE activity decreases
[7,8]. Consequently, both enzymes are important targets in AD
treatment [9–11].
The complex aetiology of AD has prompted the development of
MTDLs (multitarget-directed ligands) that act simultaneously on
different components of AD pathology, e.g. hAChE and amyloid.
Inhibitors of hAChE have been used as scaffolds to synthesize
such MTDLs [12–15].
Rational design of more potent hAChE and hBChE (human
BChE) inhibitors would benefit from a better understanding
of how the current inhibitors bind to the active sites of these
enzymes. Structural data on hAChE in complex with anti-AD
drugs have been available only recently and have shown how
important it is to analyse complexes formed with the human
enzyme rather than with AChE from other species [16]. Yet, 3D
structures of AChE in complex with inhibitors suitable for the
design of MTDLs are available currently only for mAChE (mouse
AChE) and TcAChE (Torpedo californica AChE) [17–19]. There
is no structural information on hBChE in complex with suitable
inhibitors. In the present paper, we describe the crystal structures
of hAChE in complex with FAS-2 (fasciculin 2) and huprine W
and hBChE bound to tacrine (see Figure 1 for the structures of
these inhibitors). The stabilizing effect of FAS-2 was required to
Abbreviations used: AChE, acetylcholinesterase; AD, Alzheimer’s disease; BChE, butyrylcholinesterase; CHO, Chinese-hamster ovary;
Dm
AChE,
Drosophila melanogaster
AChE; ESRF, European Synchrotron Radiation Facility; FAS-2, fasciculin 2; hAChE, human AChE; hBChE, human BChE;
mAChE, mouse AChE; MTDL, multitarget-directed ligand; PRAD, proline-rich attachment domain;
Tc
AChE,
Torpedo californica
acetylcholinesterase;
WAT, tryptophan amphiphilic tetramerization.
1To whom correspondence should be addressed (email florian@nachon.net).
The atomic co-ordinates and structure factors for the structures of hAChE in complex with huprine W and FAS-2, and hBChE in complex with tacrine,
have been deposited in the PDB under accession codes 4BDT and 4BDS respectively.
c
The Authors Journal compilation c
2013 Biochemical Society
Biochemical Journal www.biochemj.org
394 F. Nachon and others
Figure 1 Chemical structure of tacrine, (7
S
,11
S
)-huprine W and (7
S
,11
S
)-
huprine triazole
Tacrineand huprine W are both used in the present study to create complexes with cholinesterases
for X-ray crystallography. Huprine triazole is described in the Discussion section.
obtain crystals of our preparation of full-length hAChE [20]. Both
of these structurally related ligands are suitable for the design of
MTDLs, and their complex formed with human cholinesterases
provide information that can be used in the design of inhibitors
specific to each enzyme.
EXPERIMENTAL
Chemicals
Huprine W was synthesized as a racemic mixture [21]. FAS-
2 from Bungarus venom was purchased from Latoxan. All other
chemicals including tacrine were purchased from Sigma–Aldrich.
Recombinant hAChE and hBChE
The synthetic gene (GeneArt) coding for the full cDNA of hAChE
was inserted into a pGS vector carrying the glutamine synthetase
gene marker and expressed in CHO (Chinese-hamster ovary)-
K1 cells. The cells were maintained in serum-free Ultraculture
Medium (BioWhittaker) and transfected using jetPEI following
the recommendations of the supplier (Polyplus). Transfected
clones were selected by incubation in medium containing
methionine sulfoximine. The enzymes, secreted into the culture
medium, were purified by affinity chromatography and ion-
exchange chromatography as described previously [20].
Recombinant hBChE (L530stop) is a truncated monomer
containing residues 1–529, but is missing 45 C-terminal residues
that include the tetramerization domain. Four of the nine
carbohydrate attachment sites were deleted by site-directed
mutagenesis. Mutagenesis of Asn486 resulted in glycosylation of
Asn485, an asparagine that is not glycosylated in native hBChE, so
that the recombinant hBChE contains six N-linked glycans [22]
The recombinant hBChE gene was expressed in CHO-K1 cells,
secreted into serum-free culture medium, and purified by affinity
and ion-exchange chromatographies as described previously [22].
The enzymes were concentrated to 14 mg/ml (hAChE) or
7 mg/ml (hBChE) using a Centricon-30 ultrafiltration microcon-
centrator (30000 molecular mass cut-off, Amicon, Millipore) in
10 mM Mes buffer, pH 6.5. The enzyme concentration was de-
termined from its absorbance at 280 nm using a molar absorbance
coefficient of 1.7 for 1 mg/ml hAChE [23] or 1.8 for 1 mg/ml
hBChE [24].
Crystallization of hBChE complexed with tacrine
hBChE was crystallized using the hanging drop vapour diffusion
method as described previously [22]. The mother liquor contained
1 mM tacrine and 0.1 M Mes (pH 6.5) and 2.1 M ammonium
sulfate. A similar set-up was used to make crystals of the hBChE–
huprine W complex, but failed to yield crystals large enough for
diffraction analysis (shower of microcrystals). Crystals of the
hBChE–tacrine complex grew in a couple of weeks at 20C.
Crystals were then washed for a few seconds in a cryoprotectant
solution (0.1 M Mes buffer, 1 mM tacrine, 2.3 M ammonium
sulfate and 20%glycerol) before being flash-cooled in liquid
nitrogen for data collection.
Crystallization of hAChE complexed with FAS-2 and huprine W
Purified hAChE (0.1 mM, 6.5 mg/ml) was combined with FAS-
2 (0.2 mM) and huprine W (1 mM) in 10 mM Hepes buffer,
pH 7.4. The ternary complex was crystallized at a concentration of
0.1 mM, using the hanging drop method as described previously
[25]. The mother liquor was 0.1 M Hepes buffer (pH 7.4) and
1.3 M ammonium sulfate. An equal amount of the protein
and the mother liquor were mixed to yield a 3 μl drop. Crystals
grew within a month at 10C. The crystals were transferred to
a cryoprotectant solution (0.1 M Hepes buffer, pH 7.4, 1.6 M
ammonium sulfate and 18%glycerol) for few seconds before
they were flash-cooled in liquid nitrogen.
X-ray data collection and processing, structure determination
and refinement
Diffraction data were collected at 100 K on the ID14-eh2
and ID14-eh4 beam-lines at the ESRF (European Synchrotron
Radiation Facility). All datasets were processed with XDS (X-ray
Detector Software) [26], intensities of integrated reflections were
scaled using XSCALE and structure factors were calculated using
XDSCONV. The structures were solved with the CCP4 suite [27]
using the recombinant hBChE structure (PDB code 1P0I) or the
hAChE structure (PDB code 2X8B) as starting models. The initial
models were refined by iterative cycles of model building with
Coot [28], then restrained and TLS refinement with Phenix [29].
RESULTS
X-ray data and quaternary arrangement of hAChE
hAChE was co-inhibited by FAS-2 and huprine W in solution. The
addition of FAS-2 was required to obtain diffracting crystals of
the full-length form of hAChE. The co-crystals of the ternary
complex were obtained under conditions identical to those
used previously for hAChE–FAS-2 binary complexes [20,25].
The crystals belong to space group H32 and diffract to 3.1 Å. The
unit structure consists of three pairs of canonical hAChE dimers.
Data and refinement statistics are shown in Table 1.
The long C-terminal helix of the T-splice variants of
cholinesterases contains the WAT (tryptophan amphiphilic
tetramerization) sequence involved in a four-to-one association
with one PRAD (proline-rich attachment domain) [30]. Yet,
the WAT domain has never been observed in crystal structures
of cholinesterases because it was either truncated by design
[16,22,25,31,32] or no matching electron density was visible
c
The Authors Journal compilation c
2013 Biochemical Society
Structures of human cholinesterases with huprine and tacrine 395
Table 1 Crystallographic and refinement statistics
Values in brackets refer to the highest resolution shell.
Parameters Huprine W–FAS-2–AChE Tacrine–BChE
PDB code 4BDT 4BDS
Space group
H
32
I
422
Unit cell parameters (A
˚)
a
,
b
151.6 155.7
c
246.4 127.9
Resolution (A
˚) 58–3.1 (3.2–3.1) 41–2.1 (2.4–2.1)
Completeness (%) 98.1 (93.7) 99.8 (99.9)
R
sym (%) 9.1 (66.4) 6.9 (42.4)
I
/σ
I
16.0 (3.0) 18.4 (4.8)
Number of unique reflections 19677 45776
Redundancy 3.1 (3.1) 6.4 (6.5)
R
cryst (%) 15.9 17.5
R
free (%) 21.9 20.9
RMSD bond length (A
˚) 0.008 0.009
RMSD bond angles () 1.291 1.135
Mean
B
factor (A
˚2) 48.2 40.4
Non-hydrogen atoms
Total 5057 4708
Non-solvent 4938 4430
Solvent 119 278
Figure 2 Crystal packing of hAChE in a complex with FAS-2 and huprine W
Three canonical dimers (red and green pairs) interact via extensive contacts between bound
FAS-2 (yellow and magenta, surface shown) to form a non-physiological hexamer. The long
C-terminal helices of each dimer cross at a 60angle and intertwine at the centre of the hexamer.
Huprine W is represented as spheres (grey, carbon; blue, nitrogen; red, oxygen; green, chlorine).
[20,33]. In the structure shown in Figure 2, we were fortunate
to model six previously unreported turns of the WAT domain
including residues 544–567. Although the side chains remain
poorly resolved, the main chain was clearly identified. The
C-terminal helices of the canonical dimers cross each other at
Leu546,ata60
angle. This arrangement is difficult to reconcile
with the X-ray structure of the four-to-one complex formed
between an isolated portion of WAT and PRAD, where the
helices run parallel to form a left-handed superhelix structure
around an antiparallel left-handed PRAD helix [34]. In particular,
the crossed helices do not allow the construction of a plausible
tetramer model as that proposed by Dvir et al. [34], thus strongly
indicating that this arrangement is not representative of the
conformation of the helices in the physiological tetramer.
Figure 3 X-ray crystal structures of hAChE in a complex with FAS-2 and
huprine W, and hBChE in a complex with tacrine
Top panel, hAChE in a complex with FAS-2 and huprine W. Bottom panel, hBChE in a complex
with tacrine. Key residues are represented as sticks, with nitrogen atoms in dark blue, sulfur
atoms in yellow and oxygen atoms in red. Conserved water molecules are represented as red
spheres and chlorine atoms as magenta spheres. The
F
o
F
celectron density maps, calculated
by omitting the ligand, are represented by green mesh (3 σ).
The three canonical dimers interact via FAS-2 molecules,
forming a peculiar non-physiological hexamer with a central
pore. The extremities of the C-terminal helices from each dimer
are intertwined in the pore. A hexameric arrangement for the
same recombinant hAChE (at high concentration) was observed
by electron microscopy (G. Effantin, E. Carletti and J.P. Colletier,
unpublished work).
Interactions of huprine W in the active site of hAChE
A7.5σpeak in the initial FoFcelectron density was present in the
active site of hAChE, which could be unambiguously modelled
as the (7S,11S) configuration of huprine W (Figure 3, top panel).
Met33 of FAS-2 blocks the entrance of the active site gorge by
binding to the peripheral site residues Trp286,Tyr
72,Asp
74 and
Tyr341 (not shown). FAS-2 does not penetrate deep enough into
the gorge to overlap with the binding site of huprine W. Huprine
W is 8.0 Å from FAS-2, with the closest atoms being the amine
nitrogen of huprine W (N2) and Met33-CZ of FAS-2. The absence
of overlap explains the existence of the ternary complex.
Huprine W is stabilized in hAChE by interactions similar to
those described for other huprine derivatives in complex with
c
The Authors Journal compilation c
2013 Biochemical Society
396 F. Nachon and others
Figure 4 Views of the active site structure of hBChE in a complex with tacrine
(A) Side view and (B) top view of the active site structure of hBChE in a complex with tacrine. Key residues are represented as sticks with nitrogen atoms in dark blue, oxygen atoms in red, carbon
atoms in green and dummy atoms in grey. A water molecule is represented as a red sphere. Hydrogen bonds are represented by dashes. The
F
o
F
celectron density maps, calculated by omitting
the ligands, are represented by green mesh (3 σ).
non-hAChE [36,37]. The quinolinium substructure is embedded
in a remarkable group of aromatic rings including Trp236,Phe
295,
Phe338,Tyr
337 and Trp86 (Figure 3). The central aromatic ring of
the quinolinium is facing Trp86 while the lateral aromatic ring is
facing Tyr337. Following complex formation, Tyr337-Cαshifts by
1 Å, its Cα–Cβbond rotates 60counterclockwise, and its Cβ
Cγbond rotates 30clockwise compared with its conformation in
the apo structure (PDB code 1B41). The chlorine atom is nested
in a hydrophobic pocket delimited by Tyr337,Pro
446,Tyr
449, Met443
and Trp439. Chlorine interacts the strongest with Trp439 (3.4 Å) and
induces a 1.3 Å translation of the indole ring compared with the
apo structure. This translation is accompanied by a 1 Å translation
of Tyr449 and 1.5 Å translation of the short helical portion of the
Cys69–Cys96 -loop (consisting of residues Phe80 to Met85 ) located
behind Trp439 (structure not shown). No other significant change
in the conformations of the other active site residues is observed.
Additional stabilization of the quinolinium substructure is
provided by a hydrogen bond between the aromatic nitrogen and
the main chain carbonyl of His447 (2.8 Å). Unlike in previous
huprine–AChE complexes, the closest water molecule is at least
3.4 Å from the amino group, showing that this group does not
interact strongly with the conserved water molecule network of
the active site gorge. The water molecule network was well defined
in the electron density map despite the low-resolution data. The
hydroxyl group of huprine W makes strong hydrogen bonds with
the γ-hydroxyl of Ser203 and the α-amine of Gly122 (2.3 and
2.9 Å). It occupies a similar position as that of the water molecule
that often bridges the catalytic serine residue to the oxyanion hole
in the apo forms of cholinesterases [38].
X-ray structure of the tacrine–hBChE complex
The structure of tacrine-inhibited hBChE was solved at 2.1 Å
resolution (Table 1). The clear elongated peak near Trp82 in
the initial FoFcelectron density map reveals that tacrine binds
to the catalytic site of hBChE in a very similar way to how
huprines and tacrine bind to AChE [32,37,39] (Figure 3, bottom
panel). The key interactions are aromatic stacking with Trp82,
hydrogen bonding between the aromatic nitrogen N7 and the main
chain carbonyl of His438 (3.2 Å), and hydrogen bonding between
the amino group N15 and two water molecules belonging to the
active site water network (both 3.1 Å). These water molecules
are firmly anchored to the protein by hydrogen bonds to Asp70,
Ser79 and Thr120 (2.7, 2.8 and 2.8 Å). The partially saturated ring
of tacrine is remarkably embedded in the water network. Unlike
in AChE, additional stabilization by aromatic stacking involving
the lateral aromatic ring of tacrine is not observed because in
BChE Ala328 replaces Tyr337, the residue involved in the aromatic
stacking interaction in hAChE.
An additional strong positive peak in the initial FoFcelectron
density map is found in the vicinity of the catalytic serine residue
(Figure 4). A similar peak was first observed in the structure of
recombinant hBChE crystallized in the absence of any ligand
[40]. Careful analysis of the density in that case showed that
it could be modelled as a carboxylic acid bound to Ser198 with
an unusually long bond distance of 2.16 Å between Ser198-Oγ
and the carboxylic-C. The best density fit was achieved with a
butyrate molecule. Very recently, a different recombinant hBChE
expressed in insect cells, purified using a new protocol and
crystallized under completely different conditions, displayed a
similar carboxylic acid bound to the serine residue [41]. In that
study, the density was modelled as a β-alanine rather than a
butyrate. Assignment was on the basis of the observation that
one terminal heavy atom was very close to a water molecule
and thus more likely to be a hydrogen bond donor or acceptor.
In the structure shown in Figure 4, the shape of the density is
not linear as was observed for the putative β-alanine or butyrate.
Rather, it is consistent with a small cyclic molecule. After various
attempts with substituted aromatic or saturated six- and five-
atom ring molecules, a remarkably good fit was obtained with
formyl-proline (Figure 4). Yet, we stress that this interpretation
remains a guess, especially knowing that the source of this putative
molecule remains unknown (culture medium, protein degradation
or bacterial contaminant?). The carboxylic carbon of the formyl-
proline is not covalently bonded to the serine residue (2.9 Å). The
serine residue adopts two alternative conformations. One oxygen
atom is partly engaged in the oxyanion hole and makes strong
hydrogen bonds with the Oγof Ser198 and the α-amine of Gly117
(2.7 and 2.7 Å). In addition, the formyl oxygen is 2.9 Å from a
water molecule connected to Thr120 (2.7 Å).
Weaker peaks of electron density close to Tyr332 in the initial
FoFcelectron density map were modelled as a string of dummy
c
The Authors Journal compilation c
2013 Biochemical Society
Structures of human cholinesterases with huprine and tacrine 397
atoms. Obviously, this refinement model is speculative, but we
believe that it is more satisfying to show these unidentified
densities in the PDB model, especially when they are in regions
as important as the active site.
DISCUSSION
Huprine substituents change the conformation of mammalian AChE
Since mAChE and hAChE share identical active site residues, a
comparison of the complexes they form with different huprine
derivatives provides insight into huprine specificities. Figure 5
(top panel) shows that the central three-ring scaffold of huprine
in the mAChE structure is superimposable on the three rings
of huprine W in the hAChE structure (RMSD=0.3 Å). The
presence of the triazole at position 9 vice ethoxyl in huprine
W affects the conformation of Tyr337,Tyr
341 and Asp74 (Figure 5,
top panel, where the residue numbering for mAChE is the same
as that for hAChE). In the hAChE–FAS-2–huprine W structure,
Tyr341 is hydrogen-bonded to Asp74 ,andTyr
337 has sufficient
room to point towards the hydroxyl of the ethoxyl substituent
at position 9. In the mAChE structure [37], Tyr337 must rotate
by 60clockwise away from the triazole substituent and towards
Tyr341 . In response, Tyr341 rotates away around its Cα–Cβbond
by 90counterclockwise. The rotation of Tyr341 is favoured by
an aromatic stacking interaction with the triazole ring. However,
the hydrogen bond between Tyr341 and Asp74 is lost. The loss of the
hydrogen bond changes the orientation of Asp74, which pushes a
nearby water molecule by 1.1 Å towards the amino group of
huprine allowing the formation of a strong hydrogen bond (2.8 Å)
in the mAChE structure. Thus the absence of an equivalent water
hydrogen bond in the hAChE–huprine W structure seems related
to the nature of the substituent at position 9. These conformation
changes appear independent of the presence of FAS-2 because
no rotation of Tyr341 and subsequent loss of the hydrogen bond
are observed in the structures of apo mAChE/hAChE and FAS2-
mAChE/hAChE [16,18,31].
From this analysis, it is apparent that the conformational
adaptations of active site residues involved in the interaction with
huprines are extremely difficult to predict from the apo structure
alone. Because main chain translations occur, dedicated molecular
docking software, even those implementing flexibility of the
residue side chains, could not predict the various conformational
changes induced by the binding of different huprines. The
prediction of ligand binding to such a complex and flexible active
site as that of hAChE necessitates a more elaborate strategy.
Determination of the crystal structures for representative members
of an inhibitor family seems to be a necessary first step in order to
identify correctly the rules for active site and inhibitor interactions.
To that end, we are examining the crystal structures of hAChE and
hBChE bound to various inhibitors.
Once the basic rules of engagement are known, it should be
possible to accurately dock other derivatives into the active site,
either manually or by using docking software. Water molecules
that are an integral part of the active site architecture would
be conserved at this stage. Further refinement of the binding
conformation then could be performed by means of molecular
dynamics simulations. This phase would let the active site residues
adapt their conformation to the specificities of each derivative.
We described this strategy for huprine W binding to hAChE
in an earlier report [21]. The process predicted the binding
conformation and energy with remarkably accuracy, the RMSD
between the predicted ligand co-ordinates and experimentally
measured co-ordinates being only 0.4 Å.
Figure 5 Superimposition of the active site region of huprine W–hAChE
and huprine triazole–mAChE, and huprine W–hAChE, huprine X-
Tc
AChE, and
tacrine – hBChE
Top panel, huprine W–hAChE (cyan) and huprine triazole–mAChE (yellow). Bottom panel,
huprine W–hAChE (cyan), huprine X–
Tc
AChE (slate) and tacrine–hBChE (green). Key residues
are represented by sticks with oxygen atoms in red, nitrogen atoms in dark blue, sulfur atoms
in yellow and chlorine atoms in magenta. Conserved water molecules are represented as red
spheres and chlorine atoms as magenta spheres. The van der Waals surface of the chlorine atom
of huprine W and Met447-Cεof hBChE is represented as dots. Displayed residues and their
numbering are strictly conserved for hAChE and mAChE.
Selectivity emerging from a hydrophobic pocket
Design of inhibitors that are selective for AChE from various
organisms also requires a better understanding of the inhibitor–
protein interactions than can be obtained by docking and
molecular dynamics simulations alone. This point was described
above by comparing the effect of substituting ethoxy by triazole
on position 9 of huprine. Another example is provided by the
analysis of the hydrophobic chlorine-binding subsite for huprines
complexed with hAChE and TcAChE. Application of the insights
obtained from hAChE and TcAChE can then be used to predict
the consequences of the structure in that region on the binding of
huprines to hBChE, DmAChE (Drosophila melanogaster AChE)
and Culex pipens AChE.
The structures of hAChE in a complex with huprine W, TcAChE
in a complex with huprine X and hBChE in a complex with tacrine
are shown in Figure 5 (bottom panel). The chlorine substituent
on huprine is located in a hydrophobic pocket. For hAChE, this
pocket is defined by Tyr337,Pro
446,Tyr
449, Met443 and Trp439.
For TcAChE, Pro446 is replaced by a bulkier isoleucine residue
(Ile439, note the change in residue numbering), which makes
c
The Authors Journal compilation c
2013 Biochemical Society
398 F. Nachon and others
the hydrophobic pocket hosting the chlorine atom narrower.
Narrowing of the pocket causes a 1 Å shift in the position of
the huprine in TcAChE compared with that in hAChE. Owing
to this shift, the amino group of huprine moves closer to the
water molecule network in TcAChE than in hAChE, allowing
the formation of two hydrogen bonds (3.1 and 3.0 Å), both of
which are absent in the hAChE complex. In addition, Phe330 and
Tyr337 of TcAChE have each adapted their positions to optimize
stacking interactions with the aromatic ring system of the huprine,
resulting in translations of approximately 1.5 Å for each.
Examination of the same region in hBChE reveals that the side
chain of Met437 fills the hydrophobic pocket. As shown in Figure 5
(bottom panel), the van der Waals volume of Met437-Cε(green
dots) overlaps that of the chlorine atom of huprine W bound
to hAChE (magenta dots). Thus the chlorine atom of huprine
would not be able to fit into the active site of hBChE without
a substantial rearrangement of the residues in the preferred
hydrophobic chlorine-binding subsite. This steric restriction is
reflected in the affinity of huprine W for hBChE (IC50 =1200 nM),
where it binds three orders of magnitude more weakly than it
does to hAChE (IC50 =1.1 nM) [21]. This strongly suggests that
huprine W adopts a different orientation when bound to hBChE
than when bound to hAChE. A similar trend is observed for the
IC50 of 6-chlorotacrine, which bears a chlorine atom at the same
position as in huprine W (IC50 =8 nM for hAChE compared with
IC50 =900 nM for hBChE) [42].
By contrast, tacrine is a good fit for the hydrophobic region of
hBChE with the closest atom of tacrine being 3.9 Å from Met437-
Cε(Figure 5, bottom panel). This rationalizes why the inhibition
constant of tacrine for hBChE is only 6-fold lower than that for
TcAChE (Ki=25 nM for hBChE compared with Ki=3.8 nM for
TcAChE) [43].
Another aspect of the hydrophobic chlorine subsite is found
when examining the structure of DmAChE [44]. In this enzyme, a
channel exists connecting the active site gorge to the bulk solvent
in the region of the hydrophobic chlorine-binding pocket. This
channel results from the replacement of Tyr449 in hAChE with an
aspartate residue in the Drosophila enzyme. The X-ray structure
of DmAChE in a complex with benzyl-tacrine (PDB code 1DX4)
shows that position 6 of the tacrine moiety is next to this channel
[32]. It follows that a large substituent at position 6 could avoid
steric hindrance by entering into the channel of DmAChE, whereas
it would not fit in the hydrophobic pocket of hAChE and would
therefore reduce the binding affinity.
The same TyrAsp replacement is present in C. pipens AChE
(UnitprotKB Q86GC8) and many other insect AChEs [45]. We
propose that this feature could provide the basis on which to
design specific inhibitors of insect AChE that would not interact
significantly with hAChE.
Implications for the design of MTDLs
The X-ray structure reveals that huprine is closely embedded in
the active site of hAChE. Although close embedding accounts for
great affinity for the enzyme, it limits the number of positions
available on the scaffold to introduce a linker for connecting
a ligand for the second molecular target. In particular, any
substitution on the central tricyclic substructure would lead to
unfavourable steric clashes with active site residues. Substitution
on the amine is an option that has been successfully employed,
but at the price of disrupting the active site water network in
the vicinity of the amine [19]. This is not an issue when the
second ligand targets the peripheral site of the enzyme, because
the synergy of binding to both sites is expected to largely
counterbalance the unfavourable disruption of the water network
[42]. However, this synergy does not exist for ligands that have
no affinity for the peripheral site. Substitution at position 9 of
huprine appears to be the best option as evidenced by the structure
of huprine triazole [37], because the linker can snake outside the
gorge to the bulk solvent without disruption of the active site
structure.
Despite the active site of hBChE being wider than that of
hAChE, tacrine is also tightly embedded, with no apparent
available substitution position that would be devoid of
unfavourable steric effects. The huprine scaffold with substitution
at position 9 appears again to be the best option for hBChE-
specific MTDLs, under the condition to remove the chlorine
substituent at position 3 for the reasons discussed above.
In summary, we presented structures of hAChE and hBChE
in complexes with reversible inhibitors that are of interest for
the design of MTDLs. Comparison of huprine complexes with
mAChE, TcAChE and hAChE illustrates how active site gorge
residues rearrange to better accommodate each derivative, and
highlight the necessity of using full X-ray structure analysis
to develop an accurate understanding of the protein structure
changes that occur on binding, before attempting to make binding
predictions via docking or molecular dynamics procedures. The
structure of hBChE in a complex with tacrine re-enforces this
conclusion.
AUTHOR CONTRIBUTION
Florian Nachon, Yvain Nicolet, Ludovic Jean and Pierre
-
Yves Renard conceptually
developed the project and contributed to writing the paper. Florian Nachon, Cyril Ronco,
Ludovic Jean and Pierre
-
Yves Renard designed and synthesized huprine W. Florian
Nachon, Eug´
enie Carletti, Yvain Nicolet and Marie Trovaslet made the crystals, and
collected and analysed the X
-
ray data.
ACKNOWLEDGEMENTS
We are grateful to Lawrence M Schopfer (Nebraska Medical Center) for the critical reading
of the paper and helpful suggestions before submission. We thank the ESRF for beam-time
under long-term projects (IBS BAG).
FUNDING
C.R. is funded by a Ph.D. fellowship from The French Minist`
ere de l’Enseignement
Sup´
erieur et de la Recherche (MRES). The Institut Universitaire de France (IUF), the
R´
egion Haute Normandie (Crunch program), the Agence Nationale pour la Recherche
[grant numbers ANR 06-BLAN-0163 DETOXNEURO and ANR-09-BLAN-0192 ReAChE]
and the Direction G´
en´
erale de l’Armement [grant number DGA/DSP/STTC PDH-2-NRBC-
3-C-301] are gratefully acknowledged for their financial support.
REFERENCES
1 Jellinger, K. A. (2006) Alzheimer 100: highlights in the history of Alzheimer research.
J. Neural Transm. 113, 1603–1623
2 Daulatzai, M. A. (2010) Early stages of pathogenesis in memory impairment during
normal senescence and Alzheimer’s disease. J. Alzheimer’s Dis. 20, 355–367
3 Holzgrabe, U., Kapkova, P., Alptuzun, V., Scheiber, J. and Kugelmann, E. (2007) Targeting
acetylcholinesterase to treat neurodegeneration. Expert Opin. Ther. Targets 11, 161–179
4 Castro, A., Conde, S., Rodriguez-Franco, M. I. and Martinez, A. (2002) Non-cholinergic
pharmacotherapy approaches to the future treatment of Alzheimer’s disease. Mini-Rev.
Med. Chem. 2, 37–50
5 Smith, D. A. (2009) Treatment of Alzheimer’s disease in the long-term-care setting.
Am. J. Health-Syst. Pharm. 66, 899–907
6 Mesulam, M. M., Guillozet, A., Shaw, P., Levey, A., Duysen, E. G. and Lockridge, O.
(2002) Acetylcholinesterase knockouts establish central cholinergic pathways and can
use butyrylcholinesterase to hydrolyze acetylcholine. Neuroscience 110, 627–639
7 Darvesh, S., Hopkins, D. A. and Geula, C. (2003) Neurobiology of butyrylcholinesterase.
Nat. Rev. Neurosci. 4, 131–138
c
The Authors Journal compilation c
2013 Biochemical Society
Structures of human cholinesterases with huprine and tacrine 399
8 Greig, N. H., Utsuki, T., Ingram, D. K., Wang, Y., Pepeu, G., Scali, C., Yu, Q. S., Mamczarz,
J., Holloway, H. W., Giordano, T. et al. (2005) Selective butyrylcholinesterase inhibition
elevates brain acetylcholine, augments learning and lowers Alzheimer β-amyloid peptide
in rodent. Proc. Natl. Acad. Sci. U.S.A. 102, 17213–17218
9 Venneri, A. and Lane, R. (2009) Effects of cholinesterase inhibition on brain white matter
volume in Alzheimer’s disease. NeuroReport 20, 285–288
10 Venneri, A., McGeown, W. J. and Shanks, M. F. (2005) Empirical evidence of
neuroprotection by dual cholinesterase inhibition in Alzheimer’s disease. NeuroReport
16, 107–110
11 Ballard, C. G. (2002) Advances in the treatment of Alzheimer’s disease: benefits of dual
cholinesterase inhibition. Eur. Neurol. 47, 64–70
12 Leon, R., Garcia, A. G. and Marco-Contelles, J. (2013) Recent advances in the
multitarget-directed ligands approach for the treatment of Alzheimer’s disease.
Med. Res. Rev. 33, 139–189
13 Tumiatti, V., Minarini, A., Bolognesi, M. L., Milelli, A., Rosini, M. and Melchiorre, C.
(2010) Tacrine derivatives and Alzheimer’s disease. Curr. Med. Chem. 17,
1825–1838
14 Cavalli, A., Bolognesi, M. L., Minarini, A., Rosini, M., Tumiatti, V., Recanatini, M. and
Melchiorre, C. (2008) Multi-target-directed ligands to combat neurodegenerative
diseases. J. Med. Chem. 51, 347–372
15 Van der Schyf, C. J., Gal, S., Geldenhuys,W. J. and Youdim, M. B. (2006) Multifunctional
neuroprotective drugs targeting monoamine oxidase inhibition, iron chelation, adenosine
receptors, and cholinergic and glutamatergic action for neurodegenerative diseases.
Expert Opin. Invest. Drugs 15, 873–886
16 Cheung, J., Rudolph, M. J., Burshteyn, F., Cassidy, M. S., Gary, E. N., Love, J., Franklin,
M. C. and Height, J. J. (2012) Structures of human acetylcholinesterase in complex with
pharmacologically important ligands. J. Med. Chem. 55, 10282–10286
17 Greenblatt, H. M., Dvir, H., Silman, I. and Sussman, J. L. (2003) Acetylcholinesterase: a
multifaceted target for structure-based drug design of anticholinesterase agents for the
treatment of Alzheimer’s disease. J. Mol. Neurosci. 20, 369–383
18 Bourne, Y., Taylor, P., Radic, Z. and Marchot, P. (2003) Structural insights into
ligand interactions at the acetylcholinesterase peripheral anionic site. EMBO J. 22,
1–12
19 Bourne, Y., Kolb, H. C., Radic, Z., Sharpless, K. B., Taylor, P. and Marchot, P. (2004)
Freeze-frame inhibitor captures acetylcholinesterase in a unique conformation. Proc. Natl.
Acad. Sci. U.S.A. 101, 1449–1454
20 Carletti, E., Colletier, J. P., Dupeux, F., Trovaslet, M., Masson, P. and Nachon, F. (2010)
Structural evidence that human acetylcholinesterase inhibited by tabun ages through
O-dealkylation. J. Med. Chem. 53, 4002–4008
21 Ronco, C., Foucault, R., Gillon, E., Bohn, P., Nachon, F., Jean, L. and Renard, P. Y. (2011)
New huprine derivatives functionalized at position 9 as highly potent acetylcholinesterase
inhibitors. ChemMedChem 6, 876–888
22 Nachon, F., Nicolet, Y., Viguie, N., Masson, P., Fontecilla-Camps, J. C. and Lockridge, O.
(2002) Engineering of a monomeric and low-glycosylated form of human
butyrylcholinesterase: expression, purification, characterization and crystallization. Eur. J.
Biochem. 269, 630–637
23 Rosenberry, T. L. and Scoggin, D. M. (1984) Structure of humanerythrocyte
acetylcholinesterase. Characterization of intersubunit disulfide bonding and detergent
interaction. J. Biol. Chem. 259, 5643–5652
24 Lockridge, O. and La Du, B. N. (1978) Comparison of atypical and usual human serum
cholinesterase. Purification, number of active sites, substrate affinity, and turnover
number. J. Biol. Chem. 253, 361–366
25 Kryger, G., Harel, M., Giles, K., Toker, L., Velan, B., Lazar, A., Kronman, C., Barak, D.,
Ariel, N., Shafferman, A. et al. (2000) Structures of recombinant native and E202Q mutant
human acetylcholinesterase complexed with the snake-venom toxin fasciculin-II. Acta
Crystallogr., Sect. D: Biol. Crystallogr. 56, 1385–1394
26 Kabsch, W. (2010) XDS. Acta Crystallogr., Sect. D: Biol. Crystallogr. 66, 125–132
27 Collaborative Computational Project 4 (1994) The CCP4 suite: programs for protein
crystallography. Acta Crystallogr., Sect. D: Biol. Crystallogr. 50, 760–763
28 Emsley, P., Lohkamp, B., Scott, W. G. and Cowtan, K. (2010) Features and development of
Coot. Acta Crystallogr., Sect. D: Biol. Crystallogr. 66, 486–501
29 Adams, P. D., Afonine, P. V., Bunkoczi, G., Chen, V. B., Davis, I. W., Echols, N., Headd,
J. J., Hung, L. W., Kapral, G. J., Grosse-Kunstleve, R. W. et al. (2010) PHENIX: a
comprehensive Python-based system for macromolecular structure solution. Acta
Crystallogr., Sect. D: Biol. Crystallogr. 66, 213–221
30 Simon, S., Krejci, E. and Massoulie, J. (1998) A four-to-one association between peptide
motifs: four C-terminal domains from cholinesterase assemble with one proline-rich
attachment domain (PRAD) in the secretory pathway. EMBO J. 17, 6178–6187
31 Marchot, P., Ravelli, R. B., Raves, M. L., Bourne, Y., Vellom, D. C., Kanter, J., Camp, S.,
Sussman, J. L. and Taylor, P. (1996) Soluble monomeric acetylcholinesterase from
mouse: expression, purification, and crystallization in complex with fasciculin. Protein
Sci. 5, 672–679
32 Harel, M., Kryger, G., Rosenberry, T. L., Mallender, W. D., Lewis, T., Fletcher, R. J., Guss,
J. M., Silman, I. and Sussman, J. L. (2000) Three-dimensional structures of
Drosophila
melanogaster
acetylcholinesterase and of its complexes with two potent inhibitors.
Protein Sci. 9, 1063–1072
33 Ngamelue, M. N., Homma, K., Lockridge, O. and Asojo, O. A. (2007) Crystallization and
X-ray structure of full-length recombinant human butyrylcholinesterase. Acta Crystallogr.,
Sect. F: Struct. Biol. Crystal. Commun. 63, 723–727
34 Dvir, H., Harel, M., Bon, S., Liu, W. Q., Vidal, M., Garbay, C., Sussman, J. L., Massoulie,
J. and Silman, I. (2004) The synaptic acetylcholinesterase tetramer assembles around a
polyproline II helix. EMBO J. 23, 4394–4405
35 Reference deleted
36 Dvir, H., Wong, D. M., Harel, M., Barril, X., Orozco, M., Luque, F. J., Munoz-Torrero, D.,
Camps, P., Rosenberry, T. L., Silman, I. and Sussman, J. L. (2002) 3D structure of
Torpedo californica
acetylcholinesterase complexed with huprine X at 2.1 A
˚resolution:
kinetic and molecular dynamic correlates. Biochemistry 41, 2970–2981
37 Ronco, C., Carletti, E., Colletier, J. P., Weik, M., Nachon, F., Jean, L. and Renard, P. Y.
(2012) Huprine derivatives as sub-nanomolar human acetylcholinesterase inhibitors:
from rational design to validation by X-ray crystallography. ChemMedChem 7, 400–405
38 Koellner, G., Kryger, G., Millard, C. B., Silman, I., Sussman, J. L. and Steiner, T. (2000)
Active-site gorge and buried water molecules in crystal structures of acetylcholinesterase
from
Torpedo californica
.J.Mol.Biol.296, 713–735
39 Harel, M., Schalk, I., Ehret-Sabatier, L., Bouet, F., Goeldner, M., Hirth, C., Axelsen, P. H.,
Silman, I. and Sussman, J. L. (1993) Quaternary ligand binding to aromatic residues in
the active-site gorge of acetylcholinesterase. Proc. Natl. Acad. Sci. U.S.A. 90, 9031–9035
40 Nicolet, Y., Lockridge, O., Masson, P., Fontecilla-Camps, J. C. and Nachon, F. (2003)
Crystal structure of human butyrylcholinesterase and of its complexes with substrate and
products. J. Biol. Chem. 278, 41141–41147
41 Brazzolotto, X., Wandhammer, M., Ronco, C., Trovaslet, M., Jean, L., Lockridge, O.,
Renard, P. Y. and Nachon, F. (2012) Human butyrylcholinesterase produced in insect
cells: huprine-based affinity purification and crystal structure. FEBS J. 279, 2905–2916
42 Camps, P., Formosa, X., Galdeano, C., Munoz-Torrero, D., Ramirez, L., Gomez, E.,
Isambert, N., Lavilla, R., Badia, A., Clos, M. V. et al. (2009) Pyrano[3,2-c]quinoline-6-
chlorotacrine hybrids as a novel family of acetylcholinesterase- and β-amyloid-directed
anti-Alzheimer compounds. J. Med. Chem. 52, 5365–5379
43 Berman, H. A. and Leonard, K. (1992) Interaction of tetrahydroaminoacridine with
acetylcholinesterase and butyrylcholinesterase. Mol. Pharmacol. 41, 412–418
44 Nachon, F., Stojan, J. and Fournier, D. (2008) Insights into substrate and product traffic in
the
Drosophila melanogaster
acetylcholinesterase active site gorge by enlarging a back
channel. FEBS J. 275, 2659–2664
45 Lu, Y., Pang, Y. P., Park, Y., Gao, X., Yao, J., Zhang, X. and Zhu, K. Y. (2012) Genome
organization, phylogenies, expression patterns, and three-dimensional protein models of
two acetylcholinesterase genes from the red flour beetle. PLoS ONE 7, e32288
Received 2 January 2013/15 May 2013; accepted 17 May 2013
Published as BJ Immediate Publication 17 May 2013, doi:10.1042/BJ20130013
c
The Authors Journal compilation c
2013 Biochemical Society
... Docking studies were conducted to elucidate the binding modes of the synthesized derivatives within the active site of AChE and BChE. The molecular geometries obtained from DFT calculations of the compounds were used for the docking study, while the coordinates of human AChE and BChE were obtained from the Protein Data Bank (PDB) under the identification codes 7E3H [44] and 4BDS [45], respectively. These protein structures were stripped of ligands, water molecules, heteroatoms, and co-crystallized solvents. ...
Article
Full-text available
Cholinesterase inhibitors are employed for treating different neuromuscular disorders that arise due to decreased levels of ACh in the cortical and hippocampal, such as Alzheimer’s disease. There is a need to synthesize novel drug candidates to improve therapeutic efficacy and reduce side effects due to toxicity and emerging drug resistance. Chitosan was grafted with quinolone derivatives using EDC and NHS as coupling agents. The newly synthesized quinolone-grafted chitosan derivatives were characterized by elemental analysis, UV–Vis, FTIR, SEM and TGA. The determination of substitution degree was carried out through elemental analysis, utilizing C/N ratios. The in vitro acetylcholinesterase and butyrylcholinesterase activities and antioxidant capacity of the compounds were investigated. Additionally, in silico investigations, including quantum chemistry calculations and docking studies, were conducted to gain insights into the molecular geometry, electronic properties, and interaction modes of the quinolone units. As a result, the synthesized derivatives CsMOC and CsMON exhibited a moderate inhibitory effect on AChE when compared to Donepezil with IC50 values of 0.22 ± 0.04 and 0.88 ± 0.05 µM, respectively. In contrast, CsMON displayed noteworthy activity against BChE with an IC50 values of 1.39 ± 0.22 µM. Furthermore, both derivatives showed potent antioxidant capacity.
... Auto Dock 4.2 was used to conduct the molecular docking [23]. The RCSB Protein Data Bank was retrieved for the protein targets MAO-A, MAO-B, AChE, and BuChE using the PDB IDs 2Z5Y [24], 2V5Z [25], 4BDT [26], and 4BDS [27], respectively. Chemdraw 12.0 [28] was used to create the 2D structures of chemicals. ...
Article
Full-text available
Alzheimer’s disease (AD) is the predominant etiology of dementia, impacting a global population of approximately 50 million individuals. In the field of medicinal chemistry, there have been notable advancements in the utilization of monoamine oxidase (MAO) and cholinesterase (ChE) inhibitors for the purpose of addressing the neurotransmitter shortage associated with Alzheimer’s disease (AD). A selection of previously synthesized 3-Phenylcoumarin derivatives (5a-m) were selected for examination in the pursuit of potential multi-targeting inhibitors of MAO-A, MAO-B, AChE, and BChE. The stability and reactivity of the compounds were investigated through the utilization of density functional theory (DFT) simulations. Subsequently, a CoMFA technique, grounded in 3D-QSAR principles, was employed to construct a model and predict the inhibitory properties of analogues belonging to the class of 3-phenylcoumarin derivatives. Through the application of molecular docking methodologies, we have employed predictive analyses to determine the potential binding interactions and stability of the drugs under investigation. The results obtained from the present investigation indicate that the 3-phenylcoumarin derivatives possess a reactive electronic characteristic that is crucial for their anti-cholinesterase activity. Compound 5a demonstrated a noteworthy binding score with AChE, BChE, MAO-A and MAO-B, respectively, indicating a robust binding affinity. Graphical Abstract
Article
Three undescribed compounds including two furosteroid glycosides (perfoloside and 22-O-methylperfoloside) and one stilbenedimer (perfolostilbene) together with 21 known compounds were isolated from the roots of Smilax perfoliata. The structural elucidation was established by extensive uses of HRMS, 1D and 2D spectroscopic techniques. The assignment of the stereocenters in perfolostilbene was based on NOESY data and ECD calculation. Among the isolates, two compounds showed marginal cytotoxic activity against KB and Hela cell lines while seven stilbenoids showed strong to weak antiacetylcholinesterase and antibutyrylcholinesterase activities with IC50 ranging between 2-197 µM.
Article
Investigating innovative frameworks for addressing Alzheimer's disease is a challenging goal. In this specific scenario, a selection of asymmetric biscarbothioamide derivatives (3a–l) with different substitutions has been carefully formulated and successfully synthesized.
Article
In this study, a series of novel sulfonamide derivatives (6–10) bearing an aryl sulfonate moiety was designed, and synthesized for the first time, and their inhibitory activities against human carbonic anhydrase isoenzymes (hCA II and I), acetylcholinesterase (AChE), and butyrylcholinesterase (BChE) were investigated by in vitro and silico studies. The chemical structures of the derivatives were elucidated by elemental analysis and spectral techniques. All tested hybrids showed remarkably low nanomolar inhibition with Kı values of in the range of 8.98 ± 0.10 to 14.56 ± 0.41 nM against hCA I, 7.73 ± 0.10 to 13.85 ± 0.32 nM against hCA II, 12.90 ± 0.29 to 16.38 ± 0.41 nM against AChE, and 6.21 ± 0.09 to 14.52 ± 0.38 nM against BChE. Overall, the majority of these molecules inhibited these metabolic enzymes significantly more than acetazolamide (AZA) and neostigmine. Moreover, the biological effects of the compounds produced were assessed through computational studies of target enzymes. In this regard, in silico ADMET properties were evaluated to determine the physicochemical properties of target compounds. The calculated results support the biological activity results.
Article
Neurodegenerative diseases with progressive cellular loss of the central nervous system and elusive disease etiology provide a continuous impetus to explore drug discovery programmes aiming at identifying robust and effective inhibitors of cholinesterase and monoamine oxidase enzymes. We herein present a concise library of anthranilamide derivatives involving a palladium-catalyzed Suzuki-Miyaura cross-coupling reaction to install the diverse structural diversity required for the desired biological action. Using Ellman's method, cholinesterase inhibitory activity was performed against AChE and BuChE enzymes. In vitro assay results demonstrated that anthranilamides are potent inhibitors with remarkable potency. Compound 6k emerged as the lead candidate and dual inhibitor of both enzymes with IC50 values of 0.12 ± 0.01 and 0.49 ± 0.02 μM against AChE and BuChE, respectively. Several other compounds were found as highly potent and selective inhibitors. Anthranilamide derivatives were also tested against monoamine oxidase (A and B) enzymes using fluorometric method. In vitro data revealed compound 6h as the most potent inhibitor against MAO-A, showing an IC50 value of 0.44 ± 0.02 μM, whereas, compound 6k emerged as the top inhibitor of MAO-B with an IC50 value of 0.06 ± 0.01 μM. All the lead inhibitors were analyzed for the identification of their mechanism of action using Michaelis-Menten kinetics experiments. Compound 6k and 6h depicted a competitive mode of action against AChE and MAO-A, whereas, a non-competitive and mixed-type of inhibition was observed against BuChE and MAO-B by compounds 6k. Molecular docking analysis revealed remarkable binding affinities of the potent inhibitors with specific residues inside the active site of receptors. Furthermore, molecular dynamics simulations were performed to explore the ability of potent compounds to form energetically stable complexes with the target protein. Finally, in silico ADME calculations also demonstrated that the potent compounds exhibit promising pharmacokinetic profile, satisfying the essential criteria for drug-likeness. Altogether, the findings reported in the current work clearly suggest that the identified anthranilamide derivatives have the potential to serve as effective drug candidates for future investigations.
Article
Full-text available
Butyrylcholinesterase (BChE) is a serine hydrolase that is present in all mammalian tissues. It can accommodate larger substrates or inhibitors than acetylcholinesterase (AChE), the enzyme responsible for hydrolysis of the neurotransmitter acetylcholine in the central nervous system and neuromuscular junctions. AChE is the specific target of organophosphorous pesticides and warfare nerve agents, and BChE is a stoichiometric bioscavenger. Conversion of BChE into a catalytic bioscavenger by rational design or designing reactivators specific to BChE required structural data obtained using a recombinant low‐glycosylated human BChE expressed in Chinese hamster ovary cells. This expression system yields ∼ 1 mg of pure enzyme per litre of cell culture. Here, we report an improved expression system using insect cells with a fourfold higher yield for truncated human BChE with all glycosylation sites present. We developed a fast purification protocol for the recombinant protein using huprine‐based affinity chromatography, which is superior to the classical procainamide‐based affinity. The purified BChE crystallized under different conditions and space group than the recombinant low‐glycosylated protein produced in Chinese hamster ovary cells. The crystals diffracted to 2.5 Å. The overall monomer structure is similar to the low‐glycosylated structure except for the presence of the additional glycans. Remarkably, the carboxylic acid molecule systematically bound to the catalytic serine in the low‐glycosylated structure is also present in this new structure, despite the different expression system, purification protocol and crystallization conditions. Database Structural data have been submitted to the Protein Data Bank under accession number pdb 4AQD . DNA sequence data for the synthetic gene have been submitted to GenBank under accession number JQ941878 . Structured digital abstract BChE and BChE bind by x‐raycrystallography ( View interaction )
Article
Full-text available
Since the report of a paralogous acetylcholinesterase (AChE, EC3.1.1.7) gene in the greenbug (Schizaphis graminum) in 2002, two different AChE genes (Ace1 and Ace2) have been identified in each of at least 27 insect species. However, the gene models of Ace1 and Ace2, and their molecular properties have not yet been comprehensively analyzed in any insect species. In this study, we sequenced the full-length cDNAs, computationally predicted the corresponding three-dimensional protein models, and profiled developmental stage and tissue-specific expression patterns of two Ace genes from the red flour beetle (Tribolium castaneum; TcAce1 and TcAce2), a globally distributed major pest of stored grain products and an emerging model organism. TcAce1 and TcAce2 encode 648 and 604 amino acid residues, respectively, and have conserved motifs including a choline-binding site, a catalytic triad, and an acyl pocket. Phylogenetic analysis show that both TcAce genes are grouped into two insect Ace clusters and TcAce1 is completely diverged from TcAce2, suggesting that these two genes evolve from their corresponding Ace gene lineages in insect species. In addition, TcAce1 is located on chromosome 5, whereas TcAce2 is located on chromosome 2. Reverse transcription polymerase chain reaction (PCR) and quantitative real-time PCR analyses indicate that both genes are virtually transcribed in all the developmental stages and predominately expressed in the insect brain. Our computational analyses suggest that the TcAce1 protein is a robust acetylcholine (ACh) hydrolase and has susceptibility to sulfhydryl agents whereas the TcAce2 protein is not a catalytically efficient ACh hydrolase.
Article
Full-text available
This complete study-from rational design to validation by X-ray crystallography-allowed us to discover two sub-nanomolar hAChE inhibitors (430 and 530 pM) grafted with an easily derivatized linker directed toward the AChE peripheral site. The crystal structure of mouse AChE in complex with compound 4 was solved and confirms the favorable position of the triazole in the active site gorge, paving the way for a new class of bifunctional ligands.
Article
Full-text available
A series of 24 huprine derivatives diversely functionalized at position 9 have been synthesized and evaluated for their inhibitory activity against human recombinant acetylcholinesterase (AChE). These derivatives were prepared in one to five steps from huprine 1 bearing an ester function at position 9. Ten analogues (1, 2, 6-9, 13-15, and 23) are active in the low nanomolar range (IC(50) <5 nM), very close to the parent compound huprine X. Compounds 2, 6, and 7 show a very good selectivity for AChE, with AChE inhibitory activities 700-1160-fold higher than those for butyrylcholinesterase (BChE). The inhibitory potency of these compounds decreases with the steric bulk of the substituents at position 9. According to docking simulations, small substituents fit into the acyl-binding pocket, whereas the larger ones stick out of the active site gorge of AChE. Determination of the kinetic parameters of three of the most potent huprines (2, 6, and 7) showed that most of the difference in K(D) is accounted by a decrease in k(on) , which is correlated to the increase of the substituent size. A first in vivo evaluation has been performed in mice for the most active compound 2 (IC(50) =1.1 nM) and showed a rather weak toxicity (LD(50) =40 mg kg(-1) ) and an ability to cross the blood-brain barrier with doses above 15 mg kg(-1).
Article
To date, the pharmacotherapy of Alzheimer's disease (AD) has relied on acetylcholinesterase (AChE) inhibitors (AChEIs) and, more recently, an N-methyl-D-aspartate receptor (NMDAR) antagonist. AD is a multifactorial syndrome with several target proteins contributing to its etiology. "Multi-target-directed ligands" (MTDLs) have great potential for treating complex diseases such as AD because they can interact with multiple targets. The design of compounds that can hit more than one specific AD target thus represents an innovative strategy for AD treatment. Tacrine was the first AChEI introduced in therapy. Recent studies have demonstrated its ability to interact with different AD targets. Furthermore, numerous tacrine homo- and heterodimers have been developed with the aim of improving and enlarging its biological profile beyond its ability to act as an AChEI. Several tacrine hybrid derivatives have been designed and synthesized with the same goal. This review will focus on and summarize the last two years of research into the development of tacrine derivatives able to hit AD targets beyond simple AChE inhibition.
Article
The CCP4 (Collaborative Computational Project, number 4) program suite is a collection of programs and associated data and subroutine libraries which can be used for macromolecular structure determination by X-ray crystallography. The suite is designed to be flexible, allowing users a number of methods of achieving their aims and so there may be more than one program to cover each function. The programs are written mainly in standard Fortran77. They are from a wide variety of sources but are connected by standard data file formats. The package has been ported to all the major platforms under both Unix and VMS. The suite is distributed by anonymous ftp from Daresbury Laboratory and is widely used throughout the world.
Article
The CCP4 (Collaborative Computational Project, number 4) program suite is a collection of programs and associated data and subroutine libraries which can be used for macromolecular structure determination by X-ray crystallography. The suite is designed to be flexible, allowing users a number of methods of achieving their aims and so there may be more than one program to cover each function. The programs are written mainly in standard Fortran 77. They are from a wide variety of sources but are connected by standard data file formats. The package has been ported to all the major platforms under both Unix and VMS. The suite is distributed by anonymous ftp from Daresbury Laboratory and is widely used throughout the world.
Article
Human acetylcholinesterase (AChE) is a significant target for therapeutic drugs. Here we present high resolution crystal structures of human AChE, alone and in complexes with drug ligands; donepezil, an Alzheimer's disease drug, binds differently to human AChE than it does to Torpedo AChE. These crystals of human AChE provide a more accurate platform for further drug development than previously available.
Article
We have crystallized Drosophila melanogaster acetylcholinesterase and solved the structure of the native enzyme and of its complexes with two potent reversible inhibitors, 1,2,3,4-tetrahydro-N-(phenylmethyl)-9-acridinamine and 1,2,3,4-tetrahydro-N-(3-iodophenyl-methyl)-9-acridinamine—all three at 2.7 [Angstrom capital A, ring] resolution. The refined structure of D. melanogaster acetylcholinesterase is similar to that of vertebrate acetylcholinesterases, for example, human, mouse, and fish, in its overall fold, charge distribution, and deep active-site gorge, but some of the surface loops deviate by up to 8 [Angstrom capital A, ring] from their position in the vertebrate structures, and the C-terminal helix is shifted substantially. The active-site gorge of the insect enzyme is significantly narrower than that of Torpedo californica AChE, and its trajectory is shifted several angstroms. The volume of the lower part of the gorge of the insect enzyme is [similar]50% of that of the vertebrate enzyme. Upon binding of either of the two inhibitors, nine aromatic side chains within the active-site gorge change their conformation so as to interact with the inhibitors. Some differences in activity and specificity between the insect and vertebrate enzymes can be explained by comparison of their three-dimensional structures.
Article
With 27 million cases worldwide documented in 2006, Alzheimer's disease (AD) constitutes an overwhelming health, social, economic, and political problem to nations. Unless a new medicine capable to delay disease progression is found, the number of cases will reach 107 million in 2050. So far, the therapeutic paradigm one-compound-one-target has failed. This could be due to the multiple pathogenic mechanisms involved in AD including amyloid β (Aβ) aggregation to form plaques, τ hyperphosphorylation to disrupt microtubule to form neurofibrillary tangles, calcium imbalance, enhanced oxidative stress, impaired mitochondrial function, apoptotic neuronal death, and deterioration of synaptic transmission, particularly at cholinergic neurons. Approximately 100 compounds are presently been investigated directed to single targets, namely inhibitors of β and γ secretase, vaccines or antibodies that clear Aβ, metal chelators to inhibit Aβ aggregation, blockers of glycogen synthase kinase 3β, enhancers of mitochondrial function, antioxidants, modulators of calcium-permeable channels such as voltage-dependent calcium channels, N-methyl-D-aspartate receptors for glutamate, or enhancers of cholinergic neurotransmission such as inhibitors of acetylcholinesterase or butyrylcholinesterase. In view of this complex pathogenic mechanisms, and the successful treatment of chronic diseases such as HIV or cancer, with multiple drugs having complementary mechanisms of action, the concern is growing that AD could better be treated with a single compound targeting two or more of the pathogenic mechanisms leading to neuronal death. This review summarizes the current therapeutic strategies based on the paradigm one-compound-various targets to treat AD. A treatment that delays disease onset and/or progression by 5 years could halve the number of people requiring institutionalization and/or dying from AD.  © 2011 Wiley Periodicals, Inc. Med Res Rev.