SlideShare ist ein Scribd-Unternehmen logo
1 von 12
Downloaden Sie, um offline zu lesen
Antagonists of the EP3 Receptor for
Prostaglandin E2 Are Novel Antiplatelet
Agents That Do Not Prolong Bleeding
Jasbir Singh†
, Wayne Zeller†
, Nian Zhou†
, Georgeta Hategen†
, Rama Mishra†
, Alex Polozov†
, Peng Yu†
,
Emmanuel Onua†
, Jun Zhang†
, David Zembower†
, Alex Kiselyov†
, Jose´ L. Ramı´rez‡
,
Gudmundur Sigthorsson‡
, Jon Mar Bjornsson‡
, Margret Thorsteinsdottir‡
, Thorkell Andre´sson‡
,
Maria Bjarnadottir‡
, Olafur Magnusson‡
, Jean-Etienne Fabre§
, Kari Stefansson‡
, and
Mark E. Gurney†,‡,
*
†
deCODE Chemistry, 2501 Davey Road, Woodridge, Illinois 60517, ‡
deCODE Genetics, Sturlugata 8, IS-101 Reykjavik,
Iceland, and §
Institut de Ge´ne´tique et de Biologie Mole´culaire et Cellulaire, Institut National de la Sante´ et de la Recherche
Me´dicale U596, Centre National de la Recherche Scientifique UMR7104, Universite´ Louis Pasteur, 67400 Illkirch, France
M
yocardial infarction (MI) and stroke result
from acute thrombosis over ruptured or de-
nuded atherosclerotic plaques (atherothrom-
bosis). Atherosclerotic plaques form in the arterial ves-
sel wall through deposition of lipids and consequent
inflammatory cell infiltration (1). Treatment of acute
thrombosis is designed either to reopen occluded arter-
ies through mechanical revascularization and/or fibrin-
olysis or to prevent further atherothrombotic events by
blunting platelet responses to vascular wall injury. Al-
though the potent antiplatelet drug acetylsalicylic acid
(ASA) was synthesized in the 1850s, first in France and
then in Germany, and had entered widespread use as
an antipyretic and analgesic by the early 1900s, recog-
nition of its benefit for reducing risk of heart attack was
slow to emerge. It was not until 1988 that the Physicians
Health Study established the benefit of ASA for the pre-
vention of heart attack (2, 3). This established the clini-
cal concept that antagonism of platelet function could
decrease risk of atherothrombosis while prompting the
search for even more potent antiplatelet agents. Clopi-
dogrel, an antagonist of the platelet purinergic P2Y12 re-
ceptor for adenosine diphosphate (ADP), provides addi-
tional benefit over ASA for the prevention of recurrent
heart attack (4). A more potent, second generation P2Y12
antagonist, prasugrel, recently was shown to lower risk
for myocardial infarction still further when compared to
ASA or clopidogrel (5).
Unfortunately, P2Y12 antagonists and ASA greatly im-
pair normal platelet function, as shown by an increase in
the risk of severe or fatal bleeding. Patients treated
with prasugrel have fewer cardiac events than those tak-
*Corresponding author:
mgurney@decode.com.
Received for review August 25, 2008
and accepted January 9, 2009.
Published online February 5, 2009
10.1021/cb8002094 CCC: $40.75
© 2009 American Chemical Society
ABSTRACT Myocardial infarction and stroke are caused by blood clots forming
over a ruptured or denuded atherosclerotic plaque (atherothrombosis). Production
of prostaglandin E2 (PGE2) by an inflamed plaque exacerbates atherothrombosis
and may limit the effectiveness of current therapeutics. Platelets express multiple
G-protein coupled receptors, including receptors for ADP and PGE2. ADP can mobi-
lize Ca2ϩ
and through the P2Y12 receptor can inhibit cAMP production, causing
platelet activation and aggregation. Clopidogrel (Plavix), a selective P2Y12 antago-
nist, prevents platelets from clotting but thereby increases the risk of severe or fa-
tal bleeding. The platelet EP3 receptor for PGE2, like the P2Y12 receptor, also inhib-
its cAMP synthesis. However, unlike ADP, facilitation of platelet aggregation via the
PGE2/EP3 pathway is dependent on co-agonists that can mobilize Ca2ϩ
. We used
a ligand-based design strategy to develop peri-substituted bicylic acylsulfon-
amides as potent and selective EP3 antagonists. We show that DG-041, a selec-
tive EP3 antagonist, inhibits PGE2 facilitation of platelet aggregation in vitro and
ex vivo. PGE2 can resensitize platelets to agonist even when the P2Y12 receptor has
been blocked by clopidogrel, and this can be inhibited by DG-041. Unlike clopi-
dogrel, DG-041 does not affect bleeding time in rats, nor is bleeding time further in-
creased when DG-041 is co-administered with clopidogrel. This indicates that EP3
antagonists potentially have a superior safety profile compared to P2Y12 antago-
nists and represent a novel class of antiplatelet agents.
ARTICLE
www.acschemicalbiology.org VOL.4 NO.2 • ACS CHEMICAL BIOLOGY 115
ing clopidogrel, but additional patients suffer severe
bleeding (5). For every additional cardiovascular death
prevented by prasugrel, there is approximately one ad-
ditional death from fatal bleeding (5, 6). Thus, antiplate-
let therapy to reduce risk of MI and stroke could be im-
proved if there was a means to prevent acute
thrombosis over the damaged atherosclerotic plaque
without prolonging bleeding.
Atherosclerotic plaque rupture or denudation of the
endothelium exposes subendothelial collagen and re-
leases von Willebrand factor, thereby providing a sub-
strate to which platelets can adhere. Engagement of
platelet glycoprotein receptors by collagen and von Wil-
lebrand factor triggers Ca2ϩ
mobilization and activation
of protein kinase C. Ca2ϩ
mobilization in turn stimulates
platelet production of thromboxane A2 (TxA2) and plate-
let degranulation to release ADP. Platelets aggregate if
Ca2ϩ
mobilization is accompanied by a decrease in
cAMP. Platelets express multiple G-protein coupled re-
ceptors that regulate Ca2ϩ
mobilization and cAMP syn-
thesis (Figure 1). The TP receptor for thromboxane A2
(TxA2) and the P2Y1 receptor for ADP act through the Gq
pathway to release intracellular Ca2ϩ
stores. When Ca2ϩ
mobilization is accompanied by inhibition of cAMP syn-
thesis, platelet aggregation is induced, for example, by
ADP acting through the P2Y12 receptor and the inhibi-
tory Gi protein to inhibit adenyl cyclase.
Prostanoids are formed from arachidonic acid
through the concerted action of a cyclooxygenase
(COX-1 or COX-2) and an appropriate synthase (7).
When activated, platelets generate TxA2 through COX-1
functionally coupled to thromboxane synthase. ASA pre-
vents TxA2 synthesis by covalently modifying COX-1
though acetylation of a serine residue in the active site.
While platelets are the source of TxA2, the healthy arte-
rial wall is a source of prostacyclin (PGI2). PGI2 is pro-
duced by COX-2 functionally coupled to prostacyclin
synthase. PGI2 prevents platelet activation and aggrega-
tion in response to most agonists by acting through the
IP receptor and the stimulatory Gs protein to increase ad-
enyl cyclase activity and elevate platelet cAMP
(Figure 1). The synthesis of PGI2 by the arterial wall is in-
hibited by selective COX-2 inhibitors (8). Indeed, the im-
portance of PGI2 for platelet homeostasis is under-
scored by the increased cardiovascular risk associated
with Vioxx (rofecoxib) that led to the voluntary market
withdrawal of the drug in 2006.
In contrast to the healthy arterial wall, the chronic in-
flammatory condition of the atherosclerotic plaque
leads to increased content of COX-2 and microsomal
PGE2 synthase, mPGES-1 (9, 10), with consequent in-
creased production of PGE2 (11, 12). PGE2 acts at low
concentrations to facilitate platelet responses to mul-
tiple agonists such as collagen, TxA2, and ADP but has
no effect on platelet aggregation in the absence of a co-
agonist as it does not mobilize platelet Ca2ϩ
(13). Four
PGE2 receptors, EP1Ϫ4 (14−16), have been identified and
of these, EP1, EP3, and EP4 have been shown to be
present in platelets (17). Studies of gene-deleted mice
have shown that only EP3 mediates PGE2 facilitation of
platelet co-agonist responses (13). This is because, like
P2Y12, EP3 couples to the inhibitory Gi protein to inhibit
adenyl cyclase and thereby decreases platelet cAMP
(13, 18).
Gross et al. (11) recently showed that EP3 plays a
key role in atherothrombosis. In three different models
of inflammatory-mediated arterial thrombosis, arachi-
donic acid superfusion, FeCl3-mediated endothelial
damage, and mechanical rupture of atherosclerotic
plaque, PGE2-promoted thrombosis was impaired if
platelets lacked EP3. Since the healthy arterial wall pro-
duces negligible PGE2 (11, 12), the platelet EP3 system
should minimally affect bleeding time, although contra-
Figure 1. Adenosine diphosphate (ADP), thromboxane A2
(TxA2), and prostaglandin E2 (PGE2) act in concert to mobi-
lize platelet Ca2؉
and decrease adenyl cyclase activity,
thereby lowering platelet cAMP and consequently trigger-
ing platelet aggregation. ADP is able to cause platelet ag-
gregation as a sole agonist as it acts through the P2Y1 re-
ceptor to mobilize Ca2؉
and through the P2Y12 receptor
(the target of clopidogrel) to decrease cAMP. PGE2, which
is a negative regulator of adenylyl cyclase through the EP3
receptor, is not able to cause platelet aggregation in the
absence of a co-agonist that also mobilizes Ca2؉
. Prosta-
cyclin (PGI2) acting through the IP receptor, elevates cAMP
and thereby is able to block the platelet agonist effects of
ADP, TxA2, and PGE2.
116 VOL.4 NO.2 • 115–126 • 2009 www.acschemicalbiology.orgSINGH ET AL.
dictory results have been reported for bleeding times in
EP3 gene-deleted mice (13, 18).
We now report the identification of a novel and po-
tent EP3 antagonist that has antiplatelet activity with
minimal impact on bleeding time. These data point to
EP3 as a novel antiplatelet target in cardiovascular dis-
ease in the context of inflammatory production of PGE2,
with reduced risk of bleeding.
RESULTS AND DISCUSSION
Ligand-Based Design of EP3 Antagonists. Our initial
medicinal chemistry effort was focused on the synthe-
sis of 1,3-disubstituted five-membered heterocycles. We
reasoned that these derivatives combined features of
the endogenous ligand, PGE2, a potent and selective EP3
agonist, sulprostone (15), and the cinnamic acid based
EP3 antagonists reported by Juteau et al. (19) (Figure 2).
After limited structureϪactivity relationship (SAR) stud-
ies, however, we failed to discover compounds with po-
tency better than 1Ϫ10 ␮M.
We subsequently developed a ligand-based design
strategy in which we attempted to recapitulate PGE2
pharmacophores. Specifically, we explored chemotypes
featuring binding elements (15) that overlapped with
the C7/C13 or C8/C14 atoms of PGE2 (Figure 2). The ap-
proach yielded multiple compounds with low nanomo-
lar potency in EP3 binding assays as exemplified by 1,7-
substituted indole 4. In developing our
structureϪactivity relationship (SAR) strategy, we simul-
taneously addressed potential metabolic liabilities of
the unsubstituted indole core and developed feasible
synthetic protocols. The key aspects of this are pre-
sented in Figure 3 (Areas IϪIV).
To investigate structural requirements for Area I, we
prepared a set of diverse bicyclic systems. Derivatives
of these showed good to excellent selectivity when pro-
filed in a panel of other EP receptors and the IP receptor.
We particularly wanted at least 1000 times less po-
tency for binding to IP versus EP3 based on the increased
cardiovascular risk found with rofecoxib (8). In prioritiz-
ing our chemistry efforts, we focused on 1,7-substituted
indoles because of their (i) synthetic feasibility and (ii)
metabolic stability upon proper substitution (Supple-
mentary Table ST1) (20). We selected for further stud-
ies molecule 7 where both C3 and C5 metabolic sites of
the indole were blocked with CH3 and F, respectively.
Compound 7 retained activity against EP3 and displayed
adequate stability in the rat liver microsomes assay. Of
the N-substituents for Area II (Figure 3), the 2,4-
dichlorobenzyl group was identified as the optimal sub-
stituent featuring Ͼ100Ϫ1000 selectivity for EP3 ver-
sus prostanoid receptor panels. In studying binding ele-
ments within Area III, we prioritized acylsulfonamide
derivatives over the parent cinnamic acid due to the ex-
Figure 2. Use of ligand conformation to aid design of small molecule EP3 antagonists. Top row: (1) structure of the endog-
enous ligand, prostaglandin E2 (PGE2) (14); (2) a potent and selective EP3 agonist, sulprostone; (3) cinnamic acid based
EP3 antagonists reported by Juteau et al. (13). Bottom row: energy minimized conformation of PGE2; schematic represen-
tation of the pharmacophore hypothesis and its overlay on PGE2; compound 4, the initial hEP3 antagonist hit with IC50 ‫؍‬
4.5 nM; overlay of the geometry optimized structures of PGE2 (cyan) and compound 4a, (2E)-3-(1-benzyl-1H-indol-7-yl)-N-
(thiophen-2-ylsulfonyl)prop-2-enamide) (green), based on the charge similarity index (30, 31).
ARTICLE
www.acschemicalbiology.org VOL.4 NO.2 • 115–126 • 2009 117
tensive plasma protein binding of the latter (19). Further-
more, a 4,5-dichloro-2-thiophene sulfonamide group
consistently featured high EP3 potency and low fold shift
(Ͻ10X) in binding assays performed in the presence of
plasma proteins. Consequently, this functionality was
selected as a key substituent for inclusion in subse-
quent compounds.
For comparison, we evaluated a cinnamic acylsulfon-
amide EP3 receptor antagonist (compound 6d) reported
by Juteau et al. (19) by in vitro binding assay using mem-
brane preparations derived from CHO cells stably ex-
pressing hEP3D or mEP3a. Compound 6d displayed IC50
of 18 nM and 563 nM for human and mouse EP3 recep-
tors, respectively. In addition, the hEP3 IC50 in the pres-
ence of 10% and 20% mouse serum was 1.19 ␮M (66-
fold IC50 shift) and 4.72 ␮M (260-fold IC50 shift),
respectively. Therefore, because of 30-fold lower affin-
ity for mouse receptor and the high fold shift in the pres-
ence of plasma proteins, compound 6d was not evalu-
ated further.
A number of aliphatic, acyclic, and heteroaryl linkers
for Area IV have been evaluated (Figure 3, Supplemen-
tary Table ST-2). In studying the optimum linker length,
we found that compounds bearing a two-atom spacer
(CAC, CϪC, OϪC, etc.) between the core and the acyl-
sulfonamide group consistently displayed potent EP3
antagonism. Compounds with longer linkers (3-atom)
were less potent. Analogs with the acylsulfonamide
group directly attached to the C7 of the indole did not
fully displace 3
H-PGE2 up to 20 ␮M. Disappointingly, al-
though molecules with saturated linkers yielded good
receptor affinity, their metabolic stability was consis-
tently low (20). In addition, these displayed very high
plasma protein binding. Insertion of aromatic linkers led
to inferior potency of the resulting molecules (Figure 3).
In our hands, derivatives of acrylamide (cinnamides)
featured the best activity and adequate metabolic stabil-
ity. We were concerned about the acrylamide double
bond, as it might have the potential to covalently modify
biological targets through the Michael addition reac-
tion. We reasoned however that the chemical nature of
the double bond in compound 12 (Supplementary Table
ST-2) should not be a liability (21). Relevant synthetic
papers (22−25) demonstrate that cinnamide is a sub-
optimal Michael acceptor. Also, the aryl acrylic acid an-
ion is not likely to feature chemical or in vivo reactivity.
Accordingly, the analog containing ␣,␤-unsaturated
acylsulfonamide provided significantly higher meta-
bolic stability for CYP-mediated oxidative reactions, as
shown in Supplementary Table ST-2.
Figure 3. Strategic representation of the structure؊activity relationship for different areas of the lead 1,7-indole template
as discussed in the text.
118 VOL.4 NO.2 • 115–126 • 2009 www.acschemicalbiology.orgSINGH ET AL.
To further confirm in vivo the low potential for chemi-
cal reactivity of the selected cinnamide linker, we ex-
plored the potential immunogenicity and systemic clear-
ance of compound 12. It was inactive in the local lymph
node assay (LLNA) following topical exposure, suggest-
ing that it did not provoke immune reactivity in vivo. We
also labeled compound 12 with 14
C at the benzylic car-
bon for in vivo radiolabel distribution studies in rats test-
ing the potential of the compound for chemical reactiv-
ity. Both male and female rats were exposed to the
radioligand. At 1 h after exposure, 14
C radioactivity was
localized to a limited number of tissues including liver,
GI, and renal systems, as expected for a molecule
mainly metabolized by the liver. Within 24 h radioactiv-
ity was limited to the GI tract and liver. Most of the radio-
activity was excreted within 72 h with mean overall re-
covery of Ͼ90%. Residual radioactivity was not
detected in blood, plasma, or cellular fractions. Re-
sidual radioactivity in the liver was negligibly small
(0.03Ϫ0.17% between males and females), indicating
the low potential of the compound for accumulation or
retention through metabolic transformation into a
chemically reactive species.
Pharmacokinetic data for compound 12 following in-
travenous and oral administration in rats are shown in
Table 1 and Supplementary Figure SF-6. The apparent
half-life of elimination was ϳ2.7 h (iv) with a systemic
clearance (CL) of ϳ1250 mL h kgϪ1
. Clearance is greater
than the estimated hepatic plasma flow rate in rats,
and the volume of distribution (Vss) was ϳ1790 mL
kgϪ1
, which is larger than the estimated volume of total
body water. Compound 12 is absorbed rapidly following
oral administration with tmax at ϳ0.42 h and absolute
oral bioavailability (%F) of ϳ27%.
Therefore, detailed SAR studies identified compound
12 as the lead compound for further evaluation.
In Vitro EP3 Pharmacology. Compound 12 (hence-
forth designated DG-041) displayed Ͼ1000 times selec-
tivity in radioligand displacement binding assays
against other PGE2 receptors and against the IP recep-
tor (Supplementary Figure SF-1A). When profiled against
CHO-K1 cells expressing human or mouse EP3, DG-041
blocked inhibition by PGE2 of forskolin-stimulated pro-
duction of cAMP (Supplementary Figure SF-1B). DG-041
also antagonized Ca2ϩ
responses in FLIPR assays (Milli-
pore) in which EP3 activation by PGE2 was coupled to
Ca2ϩ
influx (Table 2). The IC50 value for DG-041 against
EP3 measured in the FLIPR assay (8.1 nM) was in good
agreement with the IC50 measured in the binding assay
(4.6 nM). The compound was a full antagonist in both
assays. DG-041 also was equipotent in the FLIPR assay
against either PGE2 or sulprostone, evaluating each EP3
agonist at their EC80. Binding of DG-041 to EP3 was non-
TABLE 1. Pharmacokinetic parameters of DG-041 following intravenous and oral administration to male
Sprague؊Dawley ratsa
Route Dose (mg kg؊1
) tmax (h) Cmax (␮M) t1/2 (h) AUC0؊ؕ (ng h mL؊1
) Vss (mL kg؊1
) CL (mL h؊1
kg؊1
) % F
Intravenous 1.78 na 9.46 Ϯ 0.76 2.7 Ϯ 0.4 1447 Ϯ 175 1790 Ϯ 169 1250 Ϯ 154 na
Oral 9.62 0.42 Ϯ 0.14 2.74 Ϯ 0.16 4.06 Ϯ 0.99 2082 Ϯ 453 na na 27 Ϯ 5.6
a
All values are mean Ϯ SD with N ϭ 3. tmax ϭ time of maximum concentration, Cmax ϭ maximum concentration measured, t1/2 ϭ terminal elimination
half-life, AUC0Ϫϱ ϭ area under the curve extrapolated to infinity, Vss ϭ volume of distribution at steady-state, CL ϭ plasma clearance.
TABLE 2. Prostanoid receptor selectivity
for DG-041
Receptor Binding assaya
IC50 (nM) FLIPR assayb
IC50 (nM)
EP1 Ͼ20,000 486
EP2 4169 Ͼ10000
EP3 4.6 8.1
EP4 8039 Ͼ10000
IP 14414 Ͼ10000
FP nd Ͼ10000
TP nd 742
DP1 nd 131
CRTH2 nd Ͼ10000
a
The following ligands were used for displacement bind-
ing assays: 3
H-PGE2 and 3
H-Iloprost (IP). b
The following
agonists were used for FLIPR assays: EP1Ϫ4, PGE2; IP,
Iloprost; FP, PGF2␣; TP, U-46619; DP and CRTH2, PGD2.
nd ϭ not determined.
ARTICLE
www.acschemicalbiology.org VOL.4 NO.2 • 115–126 • 2009 119
competitive with PGE2 (Supplementary Figure SF-2). DG-
041 was a less potent antagonist of the DP1, EP1, and
TP receptors with IC50 ϭ 131, 486, and 742 nM, respec-
tively (Table 2). The free or unbound fraction of DG-041
is 0.67% in human and 1.1% in rat plasma, respectively
(data not shown). The relatively high plasma protein
binding for DG-041 lessens the likelihood of the com-
pound displaying pharmacology at the DP1, EP1, or TP re-
ceptors. DG-041 was inactive against an extended
panel of more than 50 other G-protein coupled recep-
tors (MDS Pharma Services).
The antiplatelet activity of DG-041 was measured by
light transmission aggregometry using platelet-rich
plasma (PRP) from human and rat (26). Because PGE2
or other EP3 agonists alone do not induce platelet aggre-
gation, collagen was used as a co-aggregant at sub-
optimal concentration, and PGE2 or sulprostone, a selec-
tive EP3 agonist, were used to potentiate the aggregation
response (15, 27). Human platelets are more sensitive
to collagen than rat. This is illustrated in Figure 4. Colla-
gen at 0.25 ␮g mLϪ1
fails to induce human platelet ag-
gregation, an 8-fold higher concentration fully induces
platelet aggregation, and exposure to sulprostone with
the sub-optimal concentration of collagen induces the
full aggregation response (Figure 4, panel a). We fo-
cused on collagen as an agonist in our experiments to
mimick conditions of plaque rupture, which exposes the
collagen-rich subendothelial matrix of the plaque to
the platelets (1). PGE2 or sulprostone also facilitate ag-
gregation responses to multiple other platelet agonists
Figure 4. In vitro and in vivo pharmacology of DG-041. a) EP3 agonists potentiate human and rat platelet aggregation in
the presence of a co-agonist as measured by light transmission aggregometry of platelet-rich-plasma (PRP). Human
platelets were exposed to either high or low collagen (2 or 0.25 ␮g mL؊1
, respectively) in the presence of varying concen-
trations of sulprostone (1, 10, or 100 nM). Sulprostone dose-dependently potentiated platelet aggregation in the pres-
ence of sub-optimal collagen. Full aggregation was obtained at 100 nM sulprostone ؉ 0.25 ␮g mL؊1
collagen, similar to
collagen alone at 2 ␮g mL؊1
. Rat platelet aggregation similarly is facilitated by PGE2 at sub-optimal collagen concentra-
tions, and this is blocked by DG-041. b) DG-041 inhibits rat platelet aggregation ex vivo. Female Long Evans rats were ga-
vaged with ascending doses of DG-041 in vehicle, blood samples were taken 2 h post-dose, and platelet aggregation was
assessed in PRP using collagen (2 ␮g mL؊1
) and PGE2 (5 ␮M). Above a dose of 10 mg kg؊1
(plasma exposure ‫؍‬ 100 nM),
rat platelet aggregation in the ex vivo assay was completely inhibited. The ED50 was 0.15 mg kg؊1
. Data are expressed
as mean ؎ SEM, N ‫؍‬ 3. c) DG-041 does not increase bleeding time. Impact on bleeding time after surgical tail incision
was assessed at doses of 10, 30, and 100 mg kg؊1
as compared to vehicle in male Sprague؊Dawley rats.
120 VOL.4 NO.2 • 115–126 • 2009 www.acschemicalbiology.orgSINGH ET AL.
at sub-optimal concentration such as ADP, the throm-
boxane analog U-46619, thrombin receptor activating
peptide, serotonin, and platelet activating factor (27,
28). In assays utilizing a sub-optimal concentration of
collagen and either PGE2 or sulprostone to facilitate the
collagen response, DG-041 inhibited human and rat
platelet aggregation as shown for rat in Figure 4, panel a.
The IC50 for DG-041 in human and rat platelet aggrega-
tion assays with the different EP3 agonists are reported
in Table 3.
In Vivo EP3 Pharmacology and Pharmacokinetics.
We next wanted to compare and contrast an EP3 an-
tagonist with a P2Y12 antagonist with respect to their im-
pact on platelet function and bleeding time. Our expec-
tation was that both receptor antagonists would
decrease platelet aggregation but that only the P2Y12 an-
tagonist would prolong bleeding.
We first tested DG-041 for effects on platelet aggrega-
tion ex vivo and on bleeding time after surgical incision.
A steep doseϪresponse curve was observed for inhibi-
tion of platelet aggregation in rats gavaged with DG-041
with an ED50 ϭ 0.2 mg kgϪ1
(Figure 4, panel b). Platelet
aggregation was completely inhibited at a dose of 10 mg
kgϪ1
or higher. DG-041 measured in the PRP sample
was 0.10 Ϯ 0.007 ␮M at 2 h post-dosing (10 mg kgϪ1
).
No increase in bleeding time after calibrated surgical in-
cision was noted at doses up to 100 mg kgϪ1
(Figure 4,
panel c). Absorption of DG-041 became nonlinear at the
top dose, reaching a plasma concentration of 19 Ϯ
2.5 ␮M at 2 h or 190 times the plasma concentration
needed for inhibition of platelet aggregation ex vivo.
Thus, despite very high exposure to DG-041, we did not
see an increase in bleeding time consistent with the
lack of an effect on bleeding times in EP3 null mice re-
ported by Fabre et al. (13). While the platelet is very sen-
sitive to the facilitation of aggregation by PGE2 and this
can be blocked by an EP3 antagonist, this underscores
how little blockade of EP3 affects bleeding.
Clopidogrel undergoes hepatic metabolism to a
chemically reactive metabolite that inactivates the plate-
let P2Y12 receptor through covalent modification. We
therefore wanted to learn if PGE2 would restore platelet
responsiveness to collagen and ADP even when the
P2Y12 receptor was chemically blocked. Range finding
experiments with clopidogrel dosed by gavage at 2.5,
5, 10, 20, and 30 mg kgϪ1
showed that at 10 mg kgϪ1
and above clopidogrel completely blocked the ex vivo
platelet response to varying collagen concentration
(Figure 5, panel a and Supplementary Figure SF-3). At
10 mg kgϪ1
, clopidogrel also prevented platelet aggre-
gation induced by ADP at 0.5Ϫ1 ␮M, but at higher ADP
concentrations, efficacy was reduced (Figure 5, panel b
and Supplementary Figure SF-3). Our data are similar to
the clopidogrel doseϪresponses for collagen and ADP
previously reported by Sugidachi et al. (29). Even at the
minimum effective dose to prevent platelet aggregation
ex vivo, clopidogrel significantly increased bleeding time
from 80 Ϯ 4 s in rats gavaged with vehicle to 123 Ϯ
14 s in rats gavaged with clopidogrel (P ϭ 0.0149;
Figure 5, panel c). Despite the substantial blockade of
P2Y12 by clopidogrel at 10 mg kgϪ1
, PGE2 restores the
platelet aggregation response to both collagen and ADP
(Figure 5, panels a and b). PGE2 facilitation of the plate-
let response to collagen and ADP can be reversed by in-
creasing the dose of clopidogrel to 30 mg kgϪ1
, but
this further prolongs bleeding time to 4 times greater
than controls (237 Ϯ 67 s, P ϭ 0.0003) (Figure 5,
panel c). In the context of the inflamed plaque, produc-
tion of PGE2 may similarly limit the efficacy of antiplate-
let therapy with P2Y12 antagonists. Higher doses of clopi-
TABLE 3. Inhibition of human and rat platelet aggregation by DG-041 measured in vitro
Species Collagen (␮g mL؊1
)a
% Plasma EP3 agonist IC50 (nM)b
Nc
Human 0.125Ϫ0.250 100 Sulprostone (0.1 ␮M) 218 Ϯ 78 8
Human 0.125Ϫ0.250 100 PGE2 (1 ␮M) 130 Ϯ 33 9
Rat 2.0 20 Sulprostone (0.2 ␮M) 83 1
Rat 2.0 20 PGE2 (5 ␮M) 298 Ϯ 82 3
a
Collagen served as the primary aggregant; PGE2 and/or sulprostone were used to facilitate platelet responses to sub-
optimal collagen. b
IC50 mean Ϯ SEM were calculated across independent experiments. c
N indicates the number of indepen-
dent experiments performed.
ARTICLE
www.acschemicalbiology.org VOL.4 NO.2 • 115–126 • 2009 121
dogrel, or a more potent P2Y12 antagonist such as
prasugrel, may reduce PGE2-facilitated platelet re-
sponses over the inflamed atherosclerotic plaque, but
at greater risk of severe or fatal bleeding.
We then explored if DG-041 could be co-administered
with clopidogrel without further prolonging bleeding,
since in the human clinical setting, initial clinical trials
in cardiovascular disease would likely require co-
administration of DG-041 with current antiplatelet drugs.
Coadministration of DG-041 at 5 or 60 mg kgϪ1
with clo-
pidogrel efficiently inhibited PGE2 facilitation of platelet
responses to collagen and ADP (Figure 5, panels a and b
and Supplementary Figure SF-4). (Doses of DG-041
greater than 60 mg kgϪ1
do not provide increased expo-
sure due to nonlinear absorption of the compound.)
PGE2 facilitation of platelet responses to collagen were
completely blocked by adding DG-041 in comparison to
clopidogrel alone (P Ͻ 0.01), while facilitation of the
platelet response to ADP was significantly reduced (P
Ͻ 0.05). A lower dose of DG-041 (5 mg kgϪ1
) with clopi-
dogrel also completely suppresses PGE2 facilitation of
platelet responses to collagen ex vivo (Supplementary
Figure SF-4). Even at the deliberately high dose of 60
mg kgϪ1
, despite potentiation of the antiplatelet effect
of clopidogrel, there was no further increase in bleeding
time in rats gavaged with clopidogrel and DG-041
(Figure 5, panel c). This again underscores that PGE2 ac-
tivation of EP3 plays little role in hemostasis in response
to vascular breach even in the context of effective P2Y12
blockade. Taken together these results suggest that co-
administration of DG-041, perhaps with a lower dose of
clopidogrel that has minimal effect on bleeding, could
be a useful strategy for improving antiplatelet therapy in
cardiovascular disease.
We successfully employed a ligand-based design
strategy to develop potent antagonists of PGE2 binding
to EP3. Early in our chemistry program, we realized that
proper spatial and electronic overlap for the synthetic
molecule required peri-substituted bicyclic aromatic
templates. As confirmation of our ligand-based model,
a broad diversity of heterocycles were tolerated by the
receptor and provided potent EP3 antagonists. Because
of both synthetic considerations and potential for meta-
bolic transformation of the central core, we focused our
Figure 5. Effects of clopidogrel with and without co-administration of DG-041 on rat ex vivo platelet aggregation and on bleeding times. PRP was
harvested 4 h (clopidogrel) and 2 h (DG-041) after oral administration. a) Rat platelet aggregation tested with collagen (5 ␮g mL؊1
) ؎ PGE2
(1 ␮M). Clopidogrel fully inhibited collagen-induced platelet aggregation, whereas the addition of PGE2 restored the aggregatory effect of colla-
gen. Co-administration of DG-041 completely blocked PGE2-facilitated aggregation (96% reduction compared to clopidogrel alone). A similar re-
duction in platelet aggregation toward collagen and PGE2 was achieved with a high dose of clopidogrel (30 mg kg؊1
). b) Rat platelet aggregation
tested with ADP (0.5 ␮M) ؎ PGE2 (1 ␮M). Clopidogrel provided partial protection toward ADP-mediated platelet aggregation (83% reduction),
while addition of PGE2 reversed the inhibitory effect of clopidogrel. Combination of clopidogrel and DG-041 yielded a decrease in platelet aggrega-
tion compared to clopidogrel alone (61% reduction), whereas a high dose of clopidogrel (30 mg kg؊1
) fully blocked aggregation. c) Clopidogrel at
10 mg kg؊1
caused increased bleeding compared to control (from 80 ؎ 4 s to 129 ؎ 16 s), while addition of a high dose of DG-041 (60 mg kg؊1
)
did not further increase bleeding times (95 ؎ 13 s). Increasing the dose of clopidogrel to 30 mg kg؊1
causes a 4-fold increase in bleeding com-
pared to controls (237 ؎ 67 s).*P < 0.05, **P < 0.01, ***P < 0.001.
122 VOL.4 NO.2 • 115–126 • 2009 www.acschemicalbiology.orgSINGH ET AL.
efforts on SAR studies of the 1,7-disubstituted indole
core. Introduction of 3-methyl and 5-fluoro substituents
in the aromatic core addressed metabolic liabilities
while preserving potency of the respective molecules.
Addressing this issue at the earlier stages of the SAR
studies allowed us to select the optimal central core for
further optimization. In prioritizing our synthetic efforts,
we focused on achieving optimal selectivity while reduc-
ing plasma protein binding as these parameters were re-
ported to be a challenge for similar programs in the in-
dustry (19). Indole derivatives combining a lipophilic
meta/para-substituted benzylic group at N1 with an
acylsulfonamide moiety at C7 of the ring furnished the
best results across multiple in vitro and ex vivo assay
panels. Additional heterocyclic cores featuring the non-
vicinal, peri-substitution pattern and preferred substitu-
ents resulted in active compounds with favorable in vitro
and ex vivo profiles. Proper selection of the linker
yielded additional advantage with regard to the meta-
bolic stability of the final molecule. Specifically, a vinyl
acylsulfonamide was identified as the optimized C7
substituent. In a comprehensive series of experiments,
we confirmed the suitability of this functionality, e.g.,
low potential for chemical reactivity, lack of antigenicity
in the local lymph node assay, and lack of accumulation
in 14
C radiolabel distribution studies in the rat. The com-
bined SAR and in vitro and ex vivo studies yielded com-
pound 12 (designated DG-041) as our candidate for hu-
man clinical studies. DG-041 was well tolerated in
preclinical safety pharmacology and in 1 and 3 month
duration chronic toxicity studies in the rat (up to 150 mg
kgϪ1
) and dog (up to 20 mg kgϪ1
) following oral dosing
(unpublished data). There was no evidence of bleeding
in any vascular bed during the in-life phase of the
chronic toxicity studies or by histopathological examina-
tion. DG-041 was not genotoxic in AMES or in vivo mam-
malian (mouse) micronucleus assays (unpublished
data).
Availability of a potent and selective EP3 antagonist
allowed us to explore the pharmacology of the EP3 sig-
naling pathway in vitro and in vivo. Consistent with stud-
ies of platelets harvested from EP3 null mice (13, 18),
DG-041 is a potent inhibitor of PGE2-facilitated re-
sponses to multiple platelet agonists when those ago-
nists are present at sub-optimal levels. Since PGE2 does
not elicit calcium mobilization, PGE2 is not able to trig-
ger platelet aggregation in the absence of a co-agonist.
DG-041 potently suppresses rat platelet aggregation ex
vivo at doses above 10 mg kgϪ1
that achieve plasma lev-
els of Ն0.1 ␮M. Even at a dose of 100 mg kgϪ1
, which
gave a 190 times higher plasma concentration, no effect
on bleeding time was observed after calibrated surgical
incision of the tail vein. Our results in rats are consistent
with the previous report by Fabre et al. (13) showing
that bleeding times are unaffected in EP3 null mice but
differ from the report by Ma et al. (18), who claim that
bleeding time was increased in EP3 null mice. Since the
healthy arterial wall contains only low amounts of PGE2
(11) and since platelets do not produce PGE2, blockade
of EP3 should minimally affect hemostasis and that is in
fact observed. Agonists released by platelets (such as
ADP and TxA2) that amplify the response of surround-
ing platelets to the local environment of the fibrin clot
are of overwhelming importance for hemostasis, as
shown by the impact of P2Y12 antagonists and COX-1 in-
hibitors on bleeding.
Fabre and co-workers hypothesize that production of
PGE2 differentiates acute thrombosis in the context of
atherosclerotic plaque rupture from hemostasis in re-
sponse to vascular breach (Fabre et al., personal com-
munication). COX-2 and mPGES-1, the biosynthetic en-
zymes for producing PGE2 are expressed in
atherosclerotic plaque (9, 10), and an increase in PGE2
content in plaque has been reported (11, 12). We mod-
eled the pro-thrombotic environment of the inflamed
plaque by using PGE2 to facilitate platelet responses to
collagen. Even at the minimum effective dose of clopi-
dogrel (10 mg kgϪ1
) that blocks ex vivo platelet re-
sponses to both collagen and ADP, PGE2 is able to re-
store the response of clopidogrel-inhibited platelets to
those agonists. We further showed that higher doses of
clopidogrel can block PGE2-facilitated responses to col-
lagen and ADP, but at a cost. Bleeding times are even
further increased, illustrating that the efficacy of clopi-
dogrel at controlling acute thrombosis in the context of
PGE2 is limited by increasingly severe impact on hemo-
stasis. DG-041 efficiently blocks PGE2 facilitation of
platelet aggregation, especially with respect to facilita-
tion of a matrix signal such as collagen present in the ar-
terial wall, and when dosed with clopidogrel, does not
further prolong bleeding.
In separate studies, we show that as reported previ-
ously for EP3 gene deletion (11), pharmacological block-
ade of EP3 with DG-041 reduces arterial thrombosis in
multiple murine models: superfusion of arachidonic
acid over the carotid artery, FeCl3 damage to the carotid
ARTICLE
www.acschemicalbiology.org VOL.4 NO.2 • 115–126 • 2009 123
endothelium, and atherothrombosis in response to me-
chanical plaque rupture (Fabreet al., personal communi-
cation). In each of these conditions, PGE2 is produced lo-
cally and precipitates intra-arterial thrombosis in
association with platelet agonists present in the vessel
wall. We further show in human phase I clinical studies
parallel to the preclinical studies reported herein that
DG-041 does not prolong bleeding time in human sub-
jects at doses that block PGE2-facilitated platelet re-
sponses ex vivo (Human Clinical Trial DG-041-CV-013,
deCODE Genetics Inc., unpublished data). Thus, an EP3
antagonist has the potential to improve antiplatelet
therapy in myocardial infarction and stroke, where con-
trol of atherothrombosis is the therapeutic goal, while
minimally impacting bleeding, which limits the benefit
of current antiplatelet therapy.
METHODS
Materials. Sulprostone and PGE2 were from Cayman Chemi-
cal. Equine collagen and ADP was from Chrono-Log. 3.2% Na-
citrate blood collection tubes were from Greiner Bio-One. Hy-
droxylpropyl ␤-cyclodextrin (HP␤CD) was from Roquette, France.
Clopidogrel/Plavix pharmacy grade tablets (75 mg) were from
Sanofi Pharma-Bristol Myers Squibb. DG-041 was synthesized
by deCODE Chemistry, Woodridge IL. All other buffers, salts, and
reagents were from Sigma and were used as supplied.
Human in Vitro Platelet Aggregation Assay. Fresh blood was
collected in citrated tubes by venipuncture from healthy male
and female volunteers with informed consent. The subjects had
not taken any nonsteroidal anti-inflammatory drugs for at least
one week prior to blood collection.
Platelet-rich plasma (PRP) isolated from citrate-collected hu-
man blood was obtained after centrifugation at 100g for 20 min.
Platelet-poor plasma (PPP) was obtained by centrifugation of
PRP at 2400g for 10 min at 25 °C. Platelet aggregation was ini-
tiated by equine collagen (0.125Ϫ0.25 ␮g mLϪ1
) and PGE2
(1 ␮M) or sulprostone (0.1 ␮M) and measured by light absor-
bance in a Chrono-Log model 490 stirred cell aggregometer at
37 °C for 10 min. In samples with DG-041, the antagonist was
preincubated at 25 °C for 10 min.
Animals. Experiments were performed using either female
Long Evans or male SpragueϪDawley rats obtained from Tac-
onic M&B, Denmark. After shipment the animals were allowed
to acclimatize for at least 5 days before use. The animals were
housed under a standard 12 h light-dark cycle with free access
to water and standard laboratory chow. All animal experiments
were approved by the Laboratory Animal Ethical Committee in
Iceland.
Dosing and Formulations. Clopidogrel/Plavix tablets (75 mg)
were ground in a morter and dissolved in deionized water. The
resulting supension was sonicated at 37 °C for 20 min followed
by centrifugation to remove poorly soluble excipients. DG-041
was formulated in a 40% HP␤CD solution in 50 mM Na-
phosphate buffer, pH 7.2 for the ex vivo platelet aggregation ex-
periments. For the oral route in Table 1, DG-041 was formu-
lated in 40% HP␤CD, 0.25% polyvinylpyrrolidone, 0.9% lysine
hydrochloride in 50 mM Na-phosphate buffer, pH 7.2. Dosing so-
lutions were administered orally by gavage to non-fasting rats
in a dosing volume of 5Ϫ10 mL kgϪ1
. Stainless steel feeding
needles (0.9 ϫ 38 mm) and 1 mL Omnifit syringes were used for
the orogastric gavage. Clopidogrel was dosed 4 h before har-
vesting platelets or determination of bleeding times, whereas
DG-041 was dosed either 1 or 2 h before. Intravenous dosing in
Table 1 was with DG-041 formulated in a 20% HP␤CD solution
in 50 mM Na-phosphate buffer, pH 7.2 in a dosing volume of
1 mL kgϪ1
.
Rat in Vitro/ex Vivo Platelet Aggregation Assay. Blood samples
were collected by cardiac puncture while rats were under isoflu-
ran anesthesia. Blood was drawn into 10 mL syringes (Braun)
with 3.2% sodium citrate buffer from blood collection tubes.
PRP was obtained by centrifugation at 150g for 15 min at 25 °C.
PPP was obtained as described above for human samples. Both
PRP and PPP were diluted 1:4 with saline before use. Platelet ag-
gregation experiments with DG-041 in vitro and ex vivo were per-
formed as described above for human platelets, except colla-
gen was used at 2 ␮g mLϪ1
and PGE2 and sulprostone were used
at 5 and 0.2 ␮M, respectively. Experiments with clopidogrel
and a combination of clopidogrel and DG-041 were performed
with collagen (2, 3.5, and 5 ␮g mLϪ1
) or ADP (0.5, 1, and 5 ␮M)
as primary agonists in the absence or presence of PGE2 (1 ␮M).
Bleeding Times. Bleeding time was determined in either con-
scious or mildly anesthetized male SpragueϪDawley rats that
were immobilized in a restrainer. Anesthesia was performed us-
ing a combination of ketamine (10 mg kgϪ1
) and xylazine (10
mg kgϪ1
) in saline, administered intraperitoneally. The tail of
each immobilized rat was warmed for 1 min at 40 °C and then
dried. A small transverse incision was made in the middle of the
tail with a scalpel. Recording of the bleeding time started when
the first drop of blood touched a circular filter paper placed be-
low the cut. Bleeding was monitored every 10 s until cessation.
Data Analysis. Platelet aggregation and bleeding time data
was analyzed using Prism 4.03 for Windows (GraphPad Soft-
ware). Comparison between individual groups was performed
using Student’s t test with data following normal distribution or
the non-parametric MannϪWhitney test for data displaying non-
gaussian distribution. Statistical significance was determined
at the 5% level.
Pharmacokinetic parameters were calculated using the soft-
ware WinNonlin version 4.1 for Windows. Parameters were de-
rived using non-compartmental analysis. Data is reported as
mean and standard deviation for each parameter as calculated
for individual animals, N ϭ 3.
General Chemistry Methods. All reagents and anhydrous sol-
vents were obtained from commercial sources and used with-
out further purification unless otherwise noted. NMR spectra
were recorded at either 400 or 500 MHz in the solvent indicated
with TMS as an internal reference. Coupling constants (J) are
given in Hz. Column chromatography was carried out in the sol-
vents indicated with silica gel. HPLC purity of compounds was
measured with a reversed-phase HPLC (Zorbax SB-C18 column,
4.6 ϫ 150 mm, 5 ␮m, 254 nm) with two diverse solvent sys-
tems: in system 1, compounds were eluted using a gradient elu-
tion 95/5 to 5/95 A/B over 20 min at a flow rate of 1.0 mL minϪ1
where solvent A was aqueous 0.05% TFA and solvent B was ace-
tonitrile (0.05% TFA); in system 2, compounds were eluted using
a gradient of 95/5 to 5/95 A/C over 20 min at a flow rate of
1.0 mL minϪ1
where solvent A was aqueous 0.05% TFA and sol-
vent C was methanol. Elemental analyses were carried out by
Galbraith Laboratories, Inc. (Knoxville, TN) or Midwest Microlab,
LLC (Indianapolis, IN). Mass spectra were obtained using either
124 VOL.4 NO.2 • 115–126 • 2009 www.acschemicalbiology.orgSINGH ET AL.
APCI or electrospray ionization. Schemes and synthesis of com-
pounds are described in Supporting Information.
Acknowledgment: The authors wish to acknowledge techni-
cal contributions of J. Lin for providing guidance and overall ana-
lytical support, D. Anderson for running mass spectral analy-
ses, and N. Zhao for acquiring NMR spectra. We also thank G.
Halldo´rsdo´ttir, A. G. Hansdo´ttir, I. Gylfado´ttir, R. Spilliaert, G. A.
Pa´lsdo´ttir, B. N. Nielsen, T. Dagbjartsdottir, B. Gudmundsdo´ttir,
S. Hrafnsdo´ttir, and S. Gunnarsdo´ttir for performance of in vitro
and in vivo pharmacological assays.
Supporting Information Available: This material is available
free of charge via the Internet at http://pubs.acs.org.
REFERENCES
1. Ross, R. (1999) AtherosclerosisϪan inflammatory disease, N. Engl.
J. Med. 340, 115–126.
2. The Physicians’ Health Study: aspirin for the primary prevention of
myocardial infarction, (1988) N. Engl. J. Med. 318, 924Ϫ926.
3. Collaborative overview of randomised trials of antiplatelet
therapyϪI: prevention of death, myocardial infarction, and stroke by
prolonged antiplatelet therapy in various categories of patients.
Antiplatelet Trialists’ Collaboration, (1994) Br. Med. J. 308, 81Ϫ106.
4. Bhatt, D., Hirsch, A., Ringleb, P., Hacke, W., and Topol, E. (2000) Re-
duction in the need for hospitalization for recurrent ischemic events
and bleeding with clopidogrel instead of aspirin. CAPRIE investiga-
tors, Am. Heart J. 140, 67–73.
5. Wiviott, S., Braunwald, E., McCabe, C., Montalescot, G., Ruzyllo, W.,
Gottlieb, S., Neumann, F., Ardissino, D., De Servi, S., Murphy, S.,
Riesmeyer, J., Weerakkody, G., Gibson, M., and Antman, E. M.
(2007) The TRITON-TIMI 38 Investigators. Prasugrel versus clopi-
dogrel in patients with acute coronary syndromes, N. Engl. J. Med.
357, 2001–2015.
6. Bhatt, D. (2007) Intensifying platelet inhibitionϪnavigating be-
tween Scylla and Charybdis, N. Engl. J. Med. 357, 2078–2081.
7. Funk, C. D. (2001) Prostaglandins and leukotrienes: advances in ei-
cosanoid biology, Science 294, 1871–1875.
8. Grosser, T, Fries, S, and FitzGerald, G A. (2006) Biological basis for
the cardiovascular consequences of COX-2 inhibition: therapeutic
challenges and opportunities, J. Clin. Invest. 116, 4–15.
9. Cipollone, F., Cesaria, P., Barbara, P., Matteo, M., Maria, F., Do-
menico, De C., Annalisa, I., Sante, U., Gianfranco, B., Vittorio, S.,
Francesco, C., Franco, C., and Andrea, M. (2001) Overexpression of
functionally coupled cyclooxygenase-2 and prostaglandin E syn-
thase in symptomatic atherosclerotic plaques as a basis of prostag-
landin E(2)-dependent plaque instability, Circulation 104, 921–
927.
10. Belton, O., Duffy, A., Toomey, S., and Fitzgerald, D. (2003) Cyclooxy-
genase isoforms and platelet vessel wall interactions in the apoli-
poprotein E knockout mouse model of atherosclerosis, Circulation
108, 3017–3023.
11. Gross, S., Tilly, P., Hentsch, D., Vonesch, J., and Fabre, J. E. (2007)
Vascular wall-produced prostaglandin E2 exacerbates arterial
thrombosis and atherothrombosis through platelet EP3 receptors,
J. Exp. Med. 204, 3113–3120.
12. Cipollone, F., Fazia, M., Iezzi, A., Cuccurullo, C., Cesare, D., Ucchino,
S., Spigonardo, F., Marchetti, A., Buttitta, F., Paloscia, L., Mascel-
lanti, M., Cuccurullo, F., and Mezzetti, A. (2005) Association be-
tween prostaglandin E receptor subtype EP4 overexpression and un-
stablephenotypeinatheroscleroticplaquesinhuman,Arterioscler.,
Thromb., Vasc. Biol. 25, 1925–19331.
13. Fabre, J., Nguyen, M., Athirakul, K., Coggins, K., McNeish, J., Austin,
S., Parise, L., FitzGerald, G., Coffman, T., and Koller, B. (2001) Activa-
tion of the murine EP3 receptor for PGE2 inhibits cAMP production
and promotes platelet aggregation, J. Clin. Invest. 107, 603–610.
14. Sugimoto, Y, and Narumiya, S. (2007) Prostaglandin E receptors,
J. Biol. Chem. 282, 11613–11617.
15. Breyer, R. M., Bagdassarian, C. K., Myers, S. A., and Breyer, M. D.
(2001) Prostanoid receptors: subtypes and signaling, Ann. Rev.
Pharmacol. Toxicol. 41, 661–690.
16. Takeuchi, K, Ukawa, H, Furukawa, O, Kawauchi, S, Araki, H, Sugi-
moto, Y, Ishikawa, A, Ushikubi, F, and Narumiya, S. (1999) Prostag-
landin E receptor subtypes involved in stimulation of gastroduode-
nalbicarbonatesecretioninratsandmice,J.Physiol.Pharmacol.50,
155–167.
17. Paul, B., Ashby, B., and Sheth, S. (1998) Distribution of prostag-
landin IP and EP receptor subtypes and isoforms in platelets and hu-
man umbilical artery smooth muscle cells, Br. J. Haematol. 102,
1204–1211.
18. Ma, H., Hara, A., Xiao, C. Y., Okada, Y., Takahata, O., Nakaya, K.,
Sugimoto, Y., Ichikawa, A., Narumiya, S., and Ushikubi, F. (2001) In-
creased bleeding tendency and decreased susceptibility to throm-
boembolism in mice lacking the prostaglandin E receptor subtype
EP(3), Circulation 104, 1176–1180.
19. Juteau, H., Gareau, Y., Labelle, M., Sturino, C., Sawyer, N., Trem-
blay, N., Lamontagne, S., Carriere, M., Denis, D., and Metters, K.
(2001) Structure-activity relationship of cinnamic acylsulfonamide
analogues on the human EP3 prostanoid receptor, Bioorg. Med.
Chem. 9, 1977–1984.
20. Selected compounds were incubated with rat liver microsomes and
incubateswerefollowedusingLC-MS-MSanalysisformetabolicpro-
file. For example, analog 4 (Supplementary Table ST-1) showed
Mϩ16 and Mϩ32 metabolites, consistent with mono- and bis-
oxidation derived metabolites.
21. Tranilast containing an ␣,␤-unsaturated amide functionality has
been reported to be launched in Japan and Korea for the treatment
of allergic rhinitis, asthma, and atopic dermatitis by Kissei Phar-
maceuticals Co., Ltd. Data source http://www.iddb3.com.
22. For a report that substituents that are particularly effective in stabi-
lizing nucleophilic addition to ␣,␤-unsaturated carbonyl com-
pounds include aldehydes, ketone, esters and other carboxylic acid
derivatives except amide, see:Carroll, F. (1998) Perspectives on
Structure and Mechanism in Organic Chemistry, pp 628Ϫ629,
Brooks/Cole Publishing Co., Pacific Grove, CA
23. Perlmutter P. (1992) Conjugate Addition Reaction in Organic Synthe-
sis, Pergamon Press, Oxford, England.
24. Palomo, C., Oiarbide, M., Dias, F., Ortiz, A., and Linden, A. (2001)
J. Am. Chem. Soc. 123, 5602–5603. This source reports that reac-
tion of N-enoly oxazolidine-2-thione with Lewis acids yields, after hy-
drolytic workup, the corresponding ␤-metcapto carbonyl adducts.
This reaction represents intramolecular addition of thiol to a tertiary
amide surrogate.
25. Reaction of methyl cinnamate and O-tert-butyldimethyl silyl cin-
namate with N-BOC cystamine in MeOH/NaOMe for 12 h at RT pro-
vided 68% and 64% yield of the corresponding Michael adducts,
respectively, whereas under similar reactions conditions neither cin-
namide nor cinnamic acid yielded any desired adduct.
26. Kawasaki, T., Ozeki, Y., Igawa, T., and Kambayashi, J. (2000) In-
creased platelet sensitivity to collagen in individuals resistant to
low-dose aspirin, Stroke 31, 591–595.
27. Heptinstall, S., Iue Espinosa, D., Manolopoulos, P., Glenn, J. R.,
White, A. E., Johnson, A., Dovlatova, N., Fox, S. C., May, J. A., Her-
mann, D., Magnusson, O., Stefansson, K., Hartman, D., and Gur-
ney, M. (2008) DG-041 inhibits the EP3 prostanoid receptorϪa new
target for inhibition of platelet function in atherothrombotic dis-
ease, Platelets 19, 605–613.
28. Wise, H., Wong, Y., and Jones, R. (2002) Prostanoid signal integra-
tion and cross talk, Neurosignals 11, 20–28.
29. Sugidachi, A., Asai, F., Ogawa, T., Inoue, T., and Koike, H. (2000) The
in vivo pharmacological profile of CS-747, a novel antiplatelet agent
with platelet ADP receptor antagonist properties, Br. J. Pharmacol.
129, 1439–1446.
ARTICLE
www.acschemicalbiology.org VOL.4 NO.2 • 115–126 • 2009 125
30. Burt, C., Richards, W. G., and Huxley, P. (1990) The application of
molecular similarity calculations, J. Comput. Chem. 11, 1139–
1146.
31. Carbo, R., Ledya, L., and Arnau, M. (1980) How similar is a mole-
cule to another? An electron density measure of similarity between
two molecular structures, Int. J. Quantum Chem. 17, 1185–1189.
126 VOL.4 NO.2 • 115–126 • 2009 www.acschemicalbiology.orgSINGH ET AL.

Weitere ähnliche Inhalte

Was ist angesagt?

Seminar 112108
Seminar 112108Seminar 112108
Seminar 112108
sheanfu
 
Circulation-2003-Spragg-929-32
Circulation-2003-Spragg-929-32Circulation-2003-Spragg-929-32
Circulation-2003-Spragg-929-32
Morteza Loghmani
 
Clopidogrel & Epidurals (Powerpoint)
Clopidogrel & Epidurals (Powerpoint)Clopidogrel & Epidurals (Powerpoint)
Clopidogrel & Epidurals (Powerpoint)
Pritesh Vyas
 
Surgery
SurgerySurgery
Surgery
axix
 

Was ist angesagt? (20)

Nsaid cardiotoxicity 1
Nsaid cardiotoxicity 1Nsaid cardiotoxicity 1
Nsaid cardiotoxicity 1
 
Goldschmidt1
Goldschmidt1Goldschmidt1
Goldschmidt1
 
Ezzahti sijbrands mulder roeters van lennep familial hypercholesterolemia new...
Ezzahti sijbrands mulder roeters van lennep familial hypercholesterolemia new...Ezzahti sijbrands mulder roeters van lennep familial hypercholesterolemia new...
Ezzahti sijbrands mulder roeters van lennep familial hypercholesterolemia new...
 
1 ffp octaplas massive transfusion 35 slides
1 ffp octaplas massive transfusion 35 slides1 ffp octaplas massive transfusion 35 slides
1 ffp octaplas massive transfusion 35 slides
 
Hypertension and HTPR (high on treatment platelet reactivity measured by aggr...
Hypertension and HTPR (high on treatment platelet reactivity measured by aggr...Hypertension and HTPR (high on treatment platelet reactivity measured by aggr...
Hypertension and HTPR (high on treatment platelet reactivity measured by aggr...
 
Clopidogril as prodrug
Clopidogril as prodrugClopidogril as prodrug
Clopidogril as prodrug
 
Seminar 112108
Seminar 112108Seminar 112108
Seminar 112108
 
Trauma induced coagulopathy
Trauma induced coagulopathyTrauma induced coagulopathy
Trauma induced coagulopathy
 
217 c reactive protein
217 c reactive protein217 c reactive protein
217 c reactive protein
 
Esv2n40
Esv2n40Esv2n40
Esv2n40
 
Interdepartment Wafarin
Interdepartment WafarinInterdepartment Wafarin
Interdepartment Wafarin
 
Circulation-2003-Spragg-929-32
Circulation-2003-Spragg-929-32Circulation-2003-Spragg-929-32
Circulation-2003-Spragg-929-32
 
Agents used in cytopenias
Agents used in cytopeniasAgents used in cytopenias
Agents used in cytopenias
 
Clopidogrel & Epidurals (Powerpoint)
Clopidogrel & Epidurals (Powerpoint)Clopidogrel & Epidurals (Powerpoint)
Clopidogrel & Epidurals (Powerpoint)
 
Ascorbic acid, corticosteroids, and thiamine
Ascorbic acid, corticosteroids, and thiamineAscorbic acid, corticosteroids, and thiamine
Ascorbic acid, corticosteroids, and thiamine
 
Holley on Coagulation Management
Holley on Coagulation ManagementHolley on Coagulation Management
Holley on Coagulation Management
 
Surgery
SurgerySurgery
Surgery
 
1090116-二型糖尿病用藥預防併發症提供器官保護作用的重要性!
1090116-二型糖尿病用藥預防併發症提供器官保護作用的重要性!1090116-二型糖尿病用藥預防併發症提供器官保護作用的重要性!
1090116-二型糖尿病用藥預防併發症提供器官保護作用的重要性!
 
Does Our Compound Show Efficacy in Heart Failure?
Does Our Compound Show Efficacy in Heart Failure?Does Our Compound Show Efficacy in Heart Failure?
Does Our Compound Show Efficacy in Heart Failure?
 
Genetic Rat Models for Hypertension:Relationship to Human Hypertension
Genetic Rat Models for Hypertension:Relationship to Human HypertensionGenetic Rat Models for Hypertension:Relationship to Human Hypertension
Genetic Rat Models for Hypertension:Relationship to Human Hypertension
 

Andere mochten auch

Photosynthesis
PhotosynthesisPhotosynthesis
Photosynthesis
Ash G
 
Evaluation question 4 media dtechnology
Evaluation question 4 media dtechnologyEvaluation question 4 media dtechnology
Evaluation question 4 media dtechnology
RachelElizaaa
 
Pln reports progress
Pln reports progress Pln reports progress
Pln reports progress
MargauxATX
 
1336333055 php tutorial_from_beginner_to_master
1336333055 php tutorial_from_beginner_to_master1336333055 php tutorial_from_beginner_to_master
1336333055 php tutorial_from_beginner_to_master
jeeva indra
 
Genetics-2
Genetics-2Genetics-2
Genetics-2
Ash G
 
Traumatic brain injury and memory
Traumatic brain injury and memoryTraumatic brain injury and memory
Traumatic brain injury and memory
LearningCog2013
 

Andere mochten auch (20)

Avoin koulutus
Avoin koulutusAvoin koulutus
Avoin koulutus
 
Portfolio final
Portfolio finalPortfolio final
Portfolio final
 
Photosynthesis
PhotosynthesisPhotosynthesis
Photosynthesis
 
PRODENSA SOURCING SERVICES
PRODENSA SOURCING SERVICESPRODENSA SOURCING SERVICES
PRODENSA SOURCING SERVICES
 
Farmacologie: Antidepresive
Farmacologie: AntidepresiveFarmacologie: Antidepresive
Farmacologie: Antidepresive
 
Evaluation question 4 media dtechnology
Evaluation question 4 media dtechnologyEvaluation question 4 media dtechnology
Evaluation question 4 media dtechnology
 
James’s life
James’s lifeJames’s life
James’s life
 
Pln reports progress
Pln reports progress Pln reports progress
Pln reports progress
 
Portable Rebar Tying Machine 2014
Portable Rebar Tying Machine 2014Portable Rebar Tying Machine 2014
Portable Rebar Tying Machine 2014
 
CV_Sahadat
CV_SahadatCV_Sahadat
CV_Sahadat
 
Towards blended learning: a case study
Towards blended learning: a case studyTowards blended learning: a case study
Towards blended learning: a case study
 
Insegnare ipad nero
Insegnare ipad neroInsegnare ipad nero
Insegnare ipad nero
 
Idj 4 slide
Idj 4 slideIdj 4 slide
Idj 4 slide
 
руководство для кондиционеров Cooper&hunter cozy www.climatmontage.com.ua
руководство для кондиционеров Cooper&hunter cozy www.climatmontage.com.uaруководство для кондиционеров Cooper&hunter cozy www.climatmontage.com.ua
руководство для кондиционеров Cooper&hunter cozy www.climatmontage.com.ua
 
travelling BALI Vol. XI
travelling BALI Vol. XItravelling BALI Vol. XI
travelling BALI Vol. XI
 
1336333055 php tutorial_from_beginner_to_master
1336333055 php tutorial_from_beginner_to_master1336333055 php tutorial_from_beginner_to_master
1336333055 php tutorial_from_beginner_to_master
 
Genetics-2
Genetics-2Genetics-2
Genetics-2
 
Traumatic brain injury and memory
Traumatic brain injury and memoryTraumatic brain injury and memory
Traumatic brain injury and memory
 
Parazitologie: Helmintiaze - Nematodoze
Parazitologie: Helmintiaze - NematodozeParazitologie: Helmintiaze - Nematodoze
Parazitologie: Helmintiaze - Nematodoze
 
Intravenous immune globulin for prevention and treatment of neonatal sepsis, ...
Intravenous immune globulin for prevention and treatment of neonatal sepsis, ...Intravenous immune globulin for prevention and treatment of neonatal sepsis, ...
Intravenous immune globulin for prevention and treatment of neonatal sepsis, ...
 

Ähnlich wie EP3 ACS Chem Bio 2009

Ähnlich wie EP3 ACS Chem Bio 2009 (20)

Antithrombotic therapy for patients with chronic
Antithrombotic therapy for patients with chronicAntithrombotic therapy for patients with chronic
Antithrombotic therapy for patients with chronic
 
Eicosanoids
EicosanoidsEicosanoids
Eicosanoids
 
Prostaglandins
ProstaglandinsProstaglandins
Prostaglandins
 
eicosanoids_1.ppt
eicosanoids_1.ppteicosanoids_1.ppt
eicosanoids_1.ppt
 
PROSTAGLANDINS
PROSTAGLANDINSPROSTAGLANDINS
PROSTAGLANDINS
 
Impact of sodium glucose cotransporter 2 (SGLT2) inhibitors on atherosclerosi...
Impact of sodium glucose cotransporter 2 (SGLT2) inhibitors on atherosclerosi...Impact of sodium glucose cotransporter 2 (SGLT2) inhibitors on atherosclerosi...
Impact of sodium glucose cotransporter 2 (SGLT2) inhibitors on atherosclerosi...
 
Prostaglandins
ProstaglandinsProstaglandins
Prostaglandins
 
Sglt2 inhibitor (1)
Sglt2 inhibitor (1)Sglt2 inhibitor (1)
Sglt2 inhibitor (1)
 
PROSTAGLANDINS
PROSTAGLANDINSPROSTAGLANDINS
PROSTAGLANDINS
 
Prostaglandins & NSAIDS
Prostaglandins & NSAIDSProstaglandins & NSAIDS
Prostaglandins & NSAIDS
 
Prostaglandins,Angiotensin,Bradykinin-Dr.Jibachha Sah,M.V.Sc,Lecturer
Prostaglandins,Angiotensin,Bradykinin-Dr.Jibachha Sah,M.V.Sc,LecturerProstaglandins,Angiotensin,Bradykinin-Dr.Jibachha Sah,M.V.Sc,Lecturer
Prostaglandins,Angiotensin,Bradykinin-Dr.Jibachha Sah,M.V.Sc,Lecturer
 
Eocosanoids
EocosanoidsEocosanoids
Eocosanoids
 
Κλούρας Ε, Λυμπερόπουλος Ε, Ελισάφ Μ. Αναστολή της PCSK9: Μια νέα προσέγγιση...
Κλούρας Ε, Λυμπερόπουλος Ε, Ελισάφ Μ.  Αναστολή της PCSK9: Μια νέα προσέγγιση...Κλούρας Ε, Λυμπερόπουλος Ε, Ελισάφ Μ.  Αναστολή της PCSK9: Μια νέα προσέγγιση...
Κλούρας Ε, Λυμπερόπουλος Ε, Ελισάφ Μ. Αναστολή της PCSK9: Μια νέα προσέγγιση...
 
Zingerone alleviates the delayed ventricular repolarization and AV conduction...
Zingerone alleviates the delayed ventricular repolarization and AV conduction...Zingerone alleviates the delayed ventricular repolarization and AV conduction...
Zingerone alleviates the delayed ventricular repolarization and AV conduction...
 
Zingerone alleviates the delayed ventricular repolarization and AV conduction...
Zingerone alleviates the delayed ventricular repolarization and AV conduction...Zingerone alleviates the delayed ventricular repolarization and AV conduction...
Zingerone alleviates the delayed ventricular repolarization and AV conduction...
 
05_IJPBA_1929_21.pdf
05_IJPBA_1929_21.pdf05_IJPBA_1929_21.pdf
05_IJPBA_1929_21.pdf
 
Anti Platelet Agents
Anti Platelet AgentsAnti Platelet Agents
Anti Platelet Agents
 
NSAIDs - an introduction
NSAIDs - an introductionNSAIDs - an introduction
NSAIDs - an introduction
 
Antiplatelet agents 250419
Antiplatelet agents 250419Antiplatelet agents 250419
Antiplatelet agents 250419
 
05_IJPBA_1929_21.pdf
05_IJPBA_1929_21.pdf05_IJPBA_1929_21.pdf
05_IJPBA_1929_21.pdf
 

Mehr von Alex Kiselyov

ASK_Modeling_PaperBMC2016
ASK_Modeling_PaperBMC2016ASK_Modeling_PaperBMC2016
ASK_Modeling_PaperBMC2016
Alex Kiselyov
 
1-s2.0-S2214647415000719-main
1-s2.0-S2214647415000719-main1-s2.0-S2214647415000719-main
1-s2.0-S2214647415000719-main
Alex Kiselyov
 
Nat_Chem_Biol_GPR30_2006
Nat_Chem_Biol_GPR30_2006Nat_Chem_Biol_GPR30_2006
Nat_Chem_Biol_GPR30_2006
Alex Kiselyov
 
Polyalkoxybenzenes JMC2011
Polyalkoxybenzenes JMC2011Polyalkoxybenzenes JMC2011
Polyalkoxybenzenes JMC2011
Alex Kiselyov
 
caspase6 FMK paper BMCL2011
caspase6 FMK paper BMCL2011caspase6 FMK paper BMCL2011
caspase6 FMK paper BMCL2011
Alex Kiselyov
 
Wityak 2015 JMC Lead optimization towards PoC tools for HD with a 4-pyrazol-...
Wityak 2015 JMC Lead optimization  towards PoC tools for HD with a 4-pyrazol-...Wityak 2015 JMC Lead optimization  towards PoC tools for HD with a 4-pyrazol-...
Wityak 2015 JMC Lead optimization towards PoC tools for HD with a 4-pyrazol-...
Alex Kiselyov
 
PDE4 Nature Biotech 2010
PDE4 Nature Biotech 2010PDE4 Nature Biotech 2010
PDE4 Nature Biotech 2010
Alex Kiselyov
 
DG041 2009 JMC jm9005912
DG041 2009 JMC jm9005912DG041 2009 JMC jm9005912
DG041 2009 JMC jm9005912
Alex Kiselyov
 
Loss of Cilia ACS ChemBio 2008 cb700163q
Loss of Cilia ACS ChemBio 2008 cb700163qLoss of Cilia ACS ChemBio 2008 cb700163q
Loss of Cilia ACS ChemBio 2008 cb700163q
Alex Kiselyov
 
DG051 JMC 2010 jm900838g
DG051 JMC 2010 jm900838gDG051 JMC 2010 jm900838g
DG051 JMC 2010 jm900838g
Alex Kiselyov
 

Mehr von Alex Kiselyov (16)

ASK_Modeling_PaperBMC2016
ASK_Modeling_PaperBMC2016ASK_Modeling_PaperBMC2016
ASK_Modeling_PaperBMC2016
 
Isothiazoles 2016
Isothiazoles 2016Isothiazoles 2016
Isothiazoles 2016
 
semenov2016
semenov2016semenov2016
semenov2016
 
ijc 29918
ijc 29918ijc 29918
ijc 29918
 
1-s2.0-S2214647415000719-main
1-s2.0-S2214647415000719-main1-s2.0-S2214647415000719-main
1-s2.0-S2214647415000719-main
 
Nat_Chem_Biol_GPR30_2006
Nat_Chem_Biol_GPR30_2006Nat_Chem_Biol_GPR30_2006
Nat_Chem_Biol_GPR30_2006
 
Polyalkoxybenzenes JMC2011
Polyalkoxybenzenes JMC2011Polyalkoxybenzenes JMC2011
Polyalkoxybenzenes JMC2011
 
LTA4H FBDD 2009 JMC
LTA4H FBDD 2009 JMCLTA4H FBDD 2009 JMC
LTA4H FBDD 2009 JMC
 
caspase6 FMK paper BMCL2011
caspase6 FMK paper BMCL2011caspase6 FMK paper BMCL2011
caspase6 FMK paper BMCL2011
 
EurJMC 2015
EurJMC 2015EurJMC 2015
EurJMC 2015
 
Wityak 2015 JMC Lead optimization towards PoC tools for HD with a 4-pyrazol-...
Wityak 2015 JMC Lead optimization  towards PoC tools for HD with a 4-pyrazol-...Wityak 2015 JMC Lead optimization  towards PoC tools for HD with a 4-pyrazol-...
Wityak 2015 JMC Lead optimization towards PoC tools for HD with a 4-pyrazol-...
 
PDE4 Nature Biotech 2010
PDE4 Nature Biotech 2010PDE4 Nature Biotech 2010
PDE4 Nature Biotech 2010
 
DG041 2009 JMC jm9005912
DG041 2009 JMC jm9005912DG041 2009 JMC jm9005912
DG041 2009 JMC jm9005912
 
Loss of Cilia ACS ChemBio 2008 cb700163q
Loss of Cilia ACS ChemBio 2008 cb700163qLoss of Cilia ACS ChemBio 2008 cb700163q
Loss of Cilia ACS ChemBio 2008 cb700163q
 
DG051 JMC 2010 jm900838g
DG051 JMC 2010 jm900838gDG051 JMC 2010 jm900838g
DG051 JMC 2010 jm900838g
 
casp6 paper JMB2011
casp6 paper JMB2011casp6 paper JMB2011
casp6 paper JMB2011
 

EP3 ACS Chem Bio 2009

  • 1. Antagonists of the EP3 Receptor for Prostaglandin E2 Are Novel Antiplatelet Agents That Do Not Prolong Bleeding Jasbir Singh† , Wayne Zeller† , Nian Zhou† , Georgeta Hategen† , Rama Mishra† , Alex Polozov† , Peng Yu† , Emmanuel Onua† , Jun Zhang† , David Zembower† , Alex Kiselyov† , Jose´ L. Ramı´rez‡ , Gudmundur Sigthorsson‡ , Jon Mar Bjornsson‡ , Margret Thorsteinsdottir‡ , Thorkell Andre´sson‡ , Maria Bjarnadottir‡ , Olafur Magnusson‡ , Jean-Etienne Fabre§ , Kari Stefansson‡ , and Mark E. Gurney†,‡, * † deCODE Chemistry, 2501 Davey Road, Woodridge, Illinois 60517, ‡ deCODE Genetics, Sturlugata 8, IS-101 Reykjavik, Iceland, and § Institut de Ge´ne´tique et de Biologie Mole´culaire et Cellulaire, Institut National de la Sante´ et de la Recherche Me´dicale U596, Centre National de la Recherche Scientifique UMR7104, Universite´ Louis Pasteur, 67400 Illkirch, France M yocardial infarction (MI) and stroke result from acute thrombosis over ruptured or de- nuded atherosclerotic plaques (atherothrom- bosis). Atherosclerotic plaques form in the arterial ves- sel wall through deposition of lipids and consequent inflammatory cell infiltration (1). Treatment of acute thrombosis is designed either to reopen occluded arter- ies through mechanical revascularization and/or fibrin- olysis or to prevent further atherothrombotic events by blunting platelet responses to vascular wall injury. Al- though the potent antiplatelet drug acetylsalicylic acid (ASA) was synthesized in the 1850s, first in France and then in Germany, and had entered widespread use as an antipyretic and analgesic by the early 1900s, recog- nition of its benefit for reducing risk of heart attack was slow to emerge. It was not until 1988 that the Physicians Health Study established the benefit of ASA for the pre- vention of heart attack (2, 3). This established the clini- cal concept that antagonism of platelet function could decrease risk of atherothrombosis while prompting the search for even more potent antiplatelet agents. Clopi- dogrel, an antagonist of the platelet purinergic P2Y12 re- ceptor for adenosine diphosphate (ADP), provides addi- tional benefit over ASA for the prevention of recurrent heart attack (4). A more potent, second generation P2Y12 antagonist, prasugrel, recently was shown to lower risk for myocardial infarction still further when compared to ASA or clopidogrel (5). Unfortunately, P2Y12 antagonists and ASA greatly im- pair normal platelet function, as shown by an increase in the risk of severe or fatal bleeding. Patients treated with prasugrel have fewer cardiac events than those tak- *Corresponding author: mgurney@decode.com. Received for review August 25, 2008 and accepted January 9, 2009. Published online February 5, 2009 10.1021/cb8002094 CCC: $40.75 © 2009 American Chemical Society ABSTRACT Myocardial infarction and stroke are caused by blood clots forming over a ruptured or denuded atherosclerotic plaque (atherothrombosis). Production of prostaglandin E2 (PGE2) by an inflamed plaque exacerbates atherothrombosis and may limit the effectiveness of current therapeutics. Platelets express multiple G-protein coupled receptors, including receptors for ADP and PGE2. ADP can mobi- lize Ca2ϩ and through the P2Y12 receptor can inhibit cAMP production, causing platelet activation and aggregation. Clopidogrel (Plavix), a selective P2Y12 antago- nist, prevents platelets from clotting but thereby increases the risk of severe or fa- tal bleeding. The platelet EP3 receptor for PGE2, like the P2Y12 receptor, also inhib- its cAMP synthesis. However, unlike ADP, facilitation of platelet aggregation via the PGE2/EP3 pathway is dependent on co-agonists that can mobilize Ca2ϩ . We used a ligand-based design strategy to develop peri-substituted bicylic acylsulfon- amides as potent and selective EP3 antagonists. We show that DG-041, a selec- tive EP3 antagonist, inhibits PGE2 facilitation of platelet aggregation in vitro and ex vivo. PGE2 can resensitize platelets to agonist even when the P2Y12 receptor has been blocked by clopidogrel, and this can be inhibited by DG-041. Unlike clopi- dogrel, DG-041 does not affect bleeding time in rats, nor is bleeding time further in- creased when DG-041 is co-administered with clopidogrel. This indicates that EP3 antagonists potentially have a superior safety profile compared to P2Y12 antago- nists and represent a novel class of antiplatelet agents. ARTICLE www.acschemicalbiology.org VOL.4 NO.2 • ACS CHEMICAL BIOLOGY 115
  • 2. ing clopidogrel, but additional patients suffer severe bleeding (5). For every additional cardiovascular death prevented by prasugrel, there is approximately one ad- ditional death from fatal bleeding (5, 6). Thus, antiplate- let therapy to reduce risk of MI and stroke could be im- proved if there was a means to prevent acute thrombosis over the damaged atherosclerotic plaque without prolonging bleeding. Atherosclerotic plaque rupture or denudation of the endothelium exposes subendothelial collagen and re- leases von Willebrand factor, thereby providing a sub- strate to which platelets can adhere. Engagement of platelet glycoprotein receptors by collagen and von Wil- lebrand factor triggers Ca2ϩ mobilization and activation of protein kinase C. Ca2ϩ mobilization in turn stimulates platelet production of thromboxane A2 (TxA2) and plate- let degranulation to release ADP. Platelets aggregate if Ca2ϩ mobilization is accompanied by a decrease in cAMP. Platelets express multiple G-protein coupled re- ceptors that regulate Ca2ϩ mobilization and cAMP syn- thesis (Figure 1). The TP receptor for thromboxane A2 (TxA2) and the P2Y1 receptor for ADP act through the Gq pathway to release intracellular Ca2ϩ stores. When Ca2ϩ mobilization is accompanied by inhibition of cAMP syn- thesis, platelet aggregation is induced, for example, by ADP acting through the P2Y12 receptor and the inhibi- tory Gi protein to inhibit adenyl cyclase. Prostanoids are formed from arachidonic acid through the concerted action of a cyclooxygenase (COX-1 or COX-2) and an appropriate synthase (7). When activated, platelets generate TxA2 through COX-1 functionally coupled to thromboxane synthase. ASA pre- vents TxA2 synthesis by covalently modifying COX-1 though acetylation of a serine residue in the active site. While platelets are the source of TxA2, the healthy arte- rial wall is a source of prostacyclin (PGI2). PGI2 is pro- duced by COX-2 functionally coupled to prostacyclin synthase. PGI2 prevents platelet activation and aggrega- tion in response to most agonists by acting through the IP receptor and the stimulatory Gs protein to increase ad- enyl cyclase activity and elevate platelet cAMP (Figure 1). The synthesis of PGI2 by the arterial wall is in- hibited by selective COX-2 inhibitors (8). Indeed, the im- portance of PGI2 for platelet homeostasis is under- scored by the increased cardiovascular risk associated with Vioxx (rofecoxib) that led to the voluntary market withdrawal of the drug in 2006. In contrast to the healthy arterial wall, the chronic in- flammatory condition of the atherosclerotic plaque leads to increased content of COX-2 and microsomal PGE2 synthase, mPGES-1 (9, 10), with consequent in- creased production of PGE2 (11, 12). PGE2 acts at low concentrations to facilitate platelet responses to mul- tiple agonists such as collagen, TxA2, and ADP but has no effect on platelet aggregation in the absence of a co- agonist as it does not mobilize platelet Ca2ϩ (13). Four PGE2 receptors, EP1Ϫ4 (14−16), have been identified and of these, EP1, EP3, and EP4 have been shown to be present in platelets (17). Studies of gene-deleted mice have shown that only EP3 mediates PGE2 facilitation of platelet co-agonist responses (13). This is because, like P2Y12, EP3 couples to the inhibitory Gi protein to inhibit adenyl cyclase and thereby decreases platelet cAMP (13, 18). Gross et al. (11) recently showed that EP3 plays a key role in atherothrombosis. In three different models of inflammatory-mediated arterial thrombosis, arachi- donic acid superfusion, FeCl3-mediated endothelial damage, and mechanical rupture of atherosclerotic plaque, PGE2-promoted thrombosis was impaired if platelets lacked EP3. Since the healthy arterial wall pro- duces negligible PGE2 (11, 12), the platelet EP3 system should minimally affect bleeding time, although contra- Figure 1. Adenosine diphosphate (ADP), thromboxane A2 (TxA2), and prostaglandin E2 (PGE2) act in concert to mobi- lize platelet Ca2؉ and decrease adenyl cyclase activity, thereby lowering platelet cAMP and consequently trigger- ing platelet aggregation. ADP is able to cause platelet ag- gregation as a sole agonist as it acts through the P2Y1 re- ceptor to mobilize Ca2؉ and through the P2Y12 receptor (the target of clopidogrel) to decrease cAMP. PGE2, which is a negative regulator of adenylyl cyclase through the EP3 receptor, is not able to cause platelet aggregation in the absence of a co-agonist that also mobilizes Ca2؉ . Prosta- cyclin (PGI2) acting through the IP receptor, elevates cAMP and thereby is able to block the platelet agonist effects of ADP, TxA2, and PGE2. 116 VOL.4 NO.2 • 115–126 • 2009 www.acschemicalbiology.orgSINGH ET AL.
  • 3. dictory results have been reported for bleeding times in EP3 gene-deleted mice (13, 18). We now report the identification of a novel and po- tent EP3 antagonist that has antiplatelet activity with minimal impact on bleeding time. These data point to EP3 as a novel antiplatelet target in cardiovascular dis- ease in the context of inflammatory production of PGE2, with reduced risk of bleeding. RESULTS AND DISCUSSION Ligand-Based Design of EP3 Antagonists. Our initial medicinal chemistry effort was focused on the synthe- sis of 1,3-disubstituted five-membered heterocycles. We reasoned that these derivatives combined features of the endogenous ligand, PGE2, a potent and selective EP3 agonist, sulprostone (15), and the cinnamic acid based EP3 antagonists reported by Juteau et al. (19) (Figure 2). After limited structureϪactivity relationship (SAR) stud- ies, however, we failed to discover compounds with po- tency better than 1Ϫ10 ␮M. We subsequently developed a ligand-based design strategy in which we attempted to recapitulate PGE2 pharmacophores. Specifically, we explored chemotypes featuring binding elements (15) that overlapped with the C7/C13 or C8/C14 atoms of PGE2 (Figure 2). The ap- proach yielded multiple compounds with low nanomo- lar potency in EP3 binding assays as exemplified by 1,7- substituted indole 4. In developing our structureϪactivity relationship (SAR) strategy, we simul- taneously addressed potential metabolic liabilities of the unsubstituted indole core and developed feasible synthetic protocols. The key aspects of this are pre- sented in Figure 3 (Areas IϪIV). To investigate structural requirements for Area I, we prepared a set of diverse bicyclic systems. Derivatives of these showed good to excellent selectivity when pro- filed in a panel of other EP receptors and the IP receptor. We particularly wanted at least 1000 times less po- tency for binding to IP versus EP3 based on the increased cardiovascular risk found with rofecoxib (8). In prioritiz- ing our chemistry efforts, we focused on 1,7-substituted indoles because of their (i) synthetic feasibility and (ii) metabolic stability upon proper substitution (Supple- mentary Table ST1) (20). We selected for further stud- ies molecule 7 where both C3 and C5 metabolic sites of the indole were blocked with CH3 and F, respectively. Compound 7 retained activity against EP3 and displayed adequate stability in the rat liver microsomes assay. Of the N-substituents for Area II (Figure 3), the 2,4- dichlorobenzyl group was identified as the optimal sub- stituent featuring Ͼ100Ϫ1000 selectivity for EP3 ver- sus prostanoid receptor panels. In studying binding ele- ments within Area III, we prioritized acylsulfonamide derivatives over the parent cinnamic acid due to the ex- Figure 2. Use of ligand conformation to aid design of small molecule EP3 antagonists. Top row: (1) structure of the endog- enous ligand, prostaglandin E2 (PGE2) (14); (2) a potent and selective EP3 agonist, sulprostone; (3) cinnamic acid based EP3 antagonists reported by Juteau et al. (13). Bottom row: energy minimized conformation of PGE2; schematic represen- tation of the pharmacophore hypothesis and its overlay on PGE2; compound 4, the initial hEP3 antagonist hit with IC50 ‫؍‬ 4.5 nM; overlay of the geometry optimized structures of PGE2 (cyan) and compound 4a, (2E)-3-(1-benzyl-1H-indol-7-yl)-N- (thiophen-2-ylsulfonyl)prop-2-enamide) (green), based on the charge similarity index (30, 31). ARTICLE www.acschemicalbiology.org VOL.4 NO.2 • 115–126 • 2009 117
  • 4. tensive plasma protein binding of the latter (19). Further- more, a 4,5-dichloro-2-thiophene sulfonamide group consistently featured high EP3 potency and low fold shift (Ͻ10X) in binding assays performed in the presence of plasma proteins. Consequently, this functionality was selected as a key substituent for inclusion in subse- quent compounds. For comparison, we evaluated a cinnamic acylsulfon- amide EP3 receptor antagonist (compound 6d) reported by Juteau et al. (19) by in vitro binding assay using mem- brane preparations derived from CHO cells stably ex- pressing hEP3D or mEP3a. Compound 6d displayed IC50 of 18 nM and 563 nM for human and mouse EP3 recep- tors, respectively. In addition, the hEP3 IC50 in the pres- ence of 10% and 20% mouse serum was 1.19 ␮M (66- fold IC50 shift) and 4.72 ␮M (260-fold IC50 shift), respectively. Therefore, because of 30-fold lower affin- ity for mouse receptor and the high fold shift in the pres- ence of plasma proteins, compound 6d was not evalu- ated further. A number of aliphatic, acyclic, and heteroaryl linkers for Area IV have been evaluated (Figure 3, Supplemen- tary Table ST-2). In studying the optimum linker length, we found that compounds bearing a two-atom spacer (CAC, CϪC, OϪC, etc.) between the core and the acyl- sulfonamide group consistently displayed potent EP3 antagonism. Compounds with longer linkers (3-atom) were less potent. Analogs with the acylsulfonamide group directly attached to the C7 of the indole did not fully displace 3 H-PGE2 up to 20 ␮M. Disappointingly, al- though molecules with saturated linkers yielded good receptor affinity, their metabolic stability was consis- tently low (20). In addition, these displayed very high plasma protein binding. Insertion of aromatic linkers led to inferior potency of the resulting molecules (Figure 3). In our hands, derivatives of acrylamide (cinnamides) featured the best activity and adequate metabolic stabil- ity. We were concerned about the acrylamide double bond, as it might have the potential to covalently modify biological targets through the Michael addition reac- tion. We reasoned however that the chemical nature of the double bond in compound 12 (Supplementary Table ST-2) should not be a liability (21). Relevant synthetic papers (22−25) demonstrate that cinnamide is a sub- optimal Michael acceptor. Also, the aryl acrylic acid an- ion is not likely to feature chemical or in vivo reactivity. Accordingly, the analog containing ␣,␤-unsaturated acylsulfonamide provided significantly higher meta- bolic stability for CYP-mediated oxidative reactions, as shown in Supplementary Table ST-2. Figure 3. Strategic representation of the structure؊activity relationship for different areas of the lead 1,7-indole template as discussed in the text. 118 VOL.4 NO.2 • 115–126 • 2009 www.acschemicalbiology.orgSINGH ET AL.
  • 5. To further confirm in vivo the low potential for chemi- cal reactivity of the selected cinnamide linker, we ex- plored the potential immunogenicity and systemic clear- ance of compound 12. It was inactive in the local lymph node assay (LLNA) following topical exposure, suggest- ing that it did not provoke immune reactivity in vivo. We also labeled compound 12 with 14 C at the benzylic car- bon for in vivo radiolabel distribution studies in rats test- ing the potential of the compound for chemical reactiv- ity. Both male and female rats were exposed to the radioligand. At 1 h after exposure, 14 C radioactivity was localized to a limited number of tissues including liver, GI, and renal systems, as expected for a molecule mainly metabolized by the liver. Within 24 h radioactiv- ity was limited to the GI tract and liver. Most of the radio- activity was excreted within 72 h with mean overall re- covery of Ͼ90%. Residual radioactivity was not detected in blood, plasma, or cellular fractions. Re- sidual radioactivity in the liver was negligibly small (0.03Ϫ0.17% between males and females), indicating the low potential of the compound for accumulation or retention through metabolic transformation into a chemically reactive species. Pharmacokinetic data for compound 12 following in- travenous and oral administration in rats are shown in Table 1 and Supplementary Figure SF-6. The apparent half-life of elimination was ϳ2.7 h (iv) with a systemic clearance (CL) of ϳ1250 mL h kgϪ1 . Clearance is greater than the estimated hepatic plasma flow rate in rats, and the volume of distribution (Vss) was ϳ1790 mL kgϪ1 , which is larger than the estimated volume of total body water. Compound 12 is absorbed rapidly following oral administration with tmax at ϳ0.42 h and absolute oral bioavailability (%F) of ϳ27%. Therefore, detailed SAR studies identified compound 12 as the lead compound for further evaluation. In Vitro EP3 Pharmacology. Compound 12 (hence- forth designated DG-041) displayed Ͼ1000 times selec- tivity in radioligand displacement binding assays against other PGE2 receptors and against the IP recep- tor (Supplementary Figure SF-1A). When profiled against CHO-K1 cells expressing human or mouse EP3, DG-041 blocked inhibition by PGE2 of forskolin-stimulated pro- duction of cAMP (Supplementary Figure SF-1B). DG-041 also antagonized Ca2ϩ responses in FLIPR assays (Milli- pore) in which EP3 activation by PGE2 was coupled to Ca2ϩ influx (Table 2). The IC50 value for DG-041 against EP3 measured in the FLIPR assay (8.1 nM) was in good agreement with the IC50 measured in the binding assay (4.6 nM). The compound was a full antagonist in both assays. DG-041 also was equipotent in the FLIPR assay against either PGE2 or sulprostone, evaluating each EP3 agonist at their EC80. Binding of DG-041 to EP3 was non- TABLE 1. Pharmacokinetic parameters of DG-041 following intravenous and oral administration to male Sprague؊Dawley ratsa Route Dose (mg kg؊1 ) tmax (h) Cmax (␮M) t1/2 (h) AUC0؊ؕ (ng h mL؊1 ) Vss (mL kg؊1 ) CL (mL h؊1 kg؊1 ) % F Intravenous 1.78 na 9.46 Ϯ 0.76 2.7 Ϯ 0.4 1447 Ϯ 175 1790 Ϯ 169 1250 Ϯ 154 na Oral 9.62 0.42 Ϯ 0.14 2.74 Ϯ 0.16 4.06 Ϯ 0.99 2082 Ϯ 453 na na 27 Ϯ 5.6 a All values are mean Ϯ SD with N ϭ 3. tmax ϭ time of maximum concentration, Cmax ϭ maximum concentration measured, t1/2 ϭ terminal elimination half-life, AUC0Ϫϱ ϭ area under the curve extrapolated to infinity, Vss ϭ volume of distribution at steady-state, CL ϭ plasma clearance. TABLE 2. Prostanoid receptor selectivity for DG-041 Receptor Binding assaya IC50 (nM) FLIPR assayb IC50 (nM) EP1 Ͼ20,000 486 EP2 4169 Ͼ10000 EP3 4.6 8.1 EP4 8039 Ͼ10000 IP 14414 Ͼ10000 FP nd Ͼ10000 TP nd 742 DP1 nd 131 CRTH2 nd Ͼ10000 a The following ligands were used for displacement bind- ing assays: 3 H-PGE2 and 3 H-Iloprost (IP). b The following agonists were used for FLIPR assays: EP1Ϫ4, PGE2; IP, Iloprost; FP, PGF2␣; TP, U-46619; DP and CRTH2, PGD2. nd ϭ not determined. ARTICLE www.acschemicalbiology.org VOL.4 NO.2 • 115–126 • 2009 119
  • 6. competitive with PGE2 (Supplementary Figure SF-2). DG- 041 was a less potent antagonist of the DP1, EP1, and TP receptors with IC50 ϭ 131, 486, and 742 nM, respec- tively (Table 2). The free or unbound fraction of DG-041 is 0.67% in human and 1.1% in rat plasma, respectively (data not shown). The relatively high plasma protein binding for DG-041 lessens the likelihood of the com- pound displaying pharmacology at the DP1, EP1, or TP re- ceptors. DG-041 was inactive against an extended panel of more than 50 other G-protein coupled recep- tors (MDS Pharma Services). The antiplatelet activity of DG-041 was measured by light transmission aggregometry using platelet-rich plasma (PRP) from human and rat (26). Because PGE2 or other EP3 agonists alone do not induce platelet aggre- gation, collagen was used as a co-aggregant at sub- optimal concentration, and PGE2 or sulprostone, a selec- tive EP3 agonist, were used to potentiate the aggregation response (15, 27). Human platelets are more sensitive to collagen than rat. This is illustrated in Figure 4. Colla- gen at 0.25 ␮g mLϪ1 fails to induce human platelet ag- gregation, an 8-fold higher concentration fully induces platelet aggregation, and exposure to sulprostone with the sub-optimal concentration of collagen induces the full aggregation response (Figure 4, panel a). We fo- cused on collagen as an agonist in our experiments to mimick conditions of plaque rupture, which exposes the collagen-rich subendothelial matrix of the plaque to the platelets (1). PGE2 or sulprostone also facilitate ag- gregation responses to multiple other platelet agonists Figure 4. In vitro and in vivo pharmacology of DG-041. a) EP3 agonists potentiate human and rat platelet aggregation in the presence of a co-agonist as measured by light transmission aggregometry of platelet-rich-plasma (PRP). Human platelets were exposed to either high or low collagen (2 or 0.25 ␮g mL؊1 , respectively) in the presence of varying concen- trations of sulprostone (1, 10, or 100 nM). Sulprostone dose-dependently potentiated platelet aggregation in the pres- ence of sub-optimal collagen. Full aggregation was obtained at 100 nM sulprostone ؉ 0.25 ␮g mL؊1 collagen, similar to collagen alone at 2 ␮g mL؊1 . Rat platelet aggregation similarly is facilitated by PGE2 at sub-optimal collagen concentra- tions, and this is blocked by DG-041. b) DG-041 inhibits rat platelet aggregation ex vivo. Female Long Evans rats were ga- vaged with ascending doses of DG-041 in vehicle, blood samples were taken 2 h post-dose, and platelet aggregation was assessed in PRP using collagen (2 ␮g mL؊1 ) and PGE2 (5 ␮M). Above a dose of 10 mg kg؊1 (plasma exposure ‫؍‬ 100 nM), rat platelet aggregation in the ex vivo assay was completely inhibited. The ED50 was 0.15 mg kg؊1 . Data are expressed as mean ؎ SEM, N ‫؍‬ 3. c) DG-041 does not increase bleeding time. Impact on bleeding time after surgical tail incision was assessed at doses of 10, 30, and 100 mg kg؊1 as compared to vehicle in male Sprague؊Dawley rats. 120 VOL.4 NO.2 • 115–126 • 2009 www.acschemicalbiology.orgSINGH ET AL.
  • 7. at sub-optimal concentration such as ADP, the throm- boxane analog U-46619, thrombin receptor activating peptide, serotonin, and platelet activating factor (27, 28). In assays utilizing a sub-optimal concentration of collagen and either PGE2 or sulprostone to facilitate the collagen response, DG-041 inhibited human and rat platelet aggregation as shown for rat in Figure 4, panel a. The IC50 for DG-041 in human and rat platelet aggrega- tion assays with the different EP3 agonists are reported in Table 3. In Vivo EP3 Pharmacology and Pharmacokinetics. We next wanted to compare and contrast an EP3 an- tagonist with a P2Y12 antagonist with respect to their im- pact on platelet function and bleeding time. Our expec- tation was that both receptor antagonists would decrease platelet aggregation but that only the P2Y12 an- tagonist would prolong bleeding. We first tested DG-041 for effects on platelet aggrega- tion ex vivo and on bleeding time after surgical incision. A steep doseϪresponse curve was observed for inhibi- tion of platelet aggregation in rats gavaged with DG-041 with an ED50 ϭ 0.2 mg kgϪ1 (Figure 4, panel b). Platelet aggregation was completely inhibited at a dose of 10 mg kgϪ1 or higher. DG-041 measured in the PRP sample was 0.10 Ϯ 0.007 ␮M at 2 h post-dosing (10 mg kgϪ1 ). No increase in bleeding time after calibrated surgical in- cision was noted at doses up to 100 mg kgϪ1 (Figure 4, panel c). Absorption of DG-041 became nonlinear at the top dose, reaching a plasma concentration of 19 Ϯ 2.5 ␮M at 2 h or 190 times the plasma concentration needed for inhibition of platelet aggregation ex vivo. Thus, despite very high exposure to DG-041, we did not see an increase in bleeding time consistent with the lack of an effect on bleeding times in EP3 null mice re- ported by Fabre et al. (13). While the platelet is very sen- sitive to the facilitation of aggregation by PGE2 and this can be blocked by an EP3 antagonist, this underscores how little blockade of EP3 affects bleeding. Clopidogrel undergoes hepatic metabolism to a chemically reactive metabolite that inactivates the plate- let P2Y12 receptor through covalent modification. We therefore wanted to learn if PGE2 would restore platelet responsiveness to collagen and ADP even when the P2Y12 receptor was chemically blocked. Range finding experiments with clopidogrel dosed by gavage at 2.5, 5, 10, 20, and 30 mg kgϪ1 showed that at 10 mg kgϪ1 and above clopidogrel completely blocked the ex vivo platelet response to varying collagen concentration (Figure 5, panel a and Supplementary Figure SF-3). At 10 mg kgϪ1 , clopidogrel also prevented platelet aggre- gation induced by ADP at 0.5Ϫ1 ␮M, but at higher ADP concentrations, efficacy was reduced (Figure 5, panel b and Supplementary Figure SF-3). Our data are similar to the clopidogrel doseϪresponses for collagen and ADP previously reported by Sugidachi et al. (29). Even at the minimum effective dose to prevent platelet aggregation ex vivo, clopidogrel significantly increased bleeding time from 80 Ϯ 4 s in rats gavaged with vehicle to 123 Ϯ 14 s in rats gavaged with clopidogrel (P ϭ 0.0149; Figure 5, panel c). Despite the substantial blockade of P2Y12 by clopidogrel at 10 mg kgϪ1 , PGE2 restores the platelet aggregation response to both collagen and ADP (Figure 5, panels a and b). PGE2 facilitation of the plate- let response to collagen and ADP can be reversed by in- creasing the dose of clopidogrel to 30 mg kgϪ1 , but this further prolongs bleeding time to 4 times greater than controls (237 Ϯ 67 s, P ϭ 0.0003) (Figure 5, panel c). In the context of the inflamed plaque, produc- tion of PGE2 may similarly limit the efficacy of antiplate- let therapy with P2Y12 antagonists. Higher doses of clopi- TABLE 3. Inhibition of human and rat platelet aggregation by DG-041 measured in vitro Species Collagen (␮g mL؊1 )a % Plasma EP3 agonist IC50 (nM)b Nc Human 0.125Ϫ0.250 100 Sulprostone (0.1 ␮M) 218 Ϯ 78 8 Human 0.125Ϫ0.250 100 PGE2 (1 ␮M) 130 Ϯ 33 9 Rat 2.0 20 Sulprostone (0.2 ␮M) 83 1 Rat 2.0 20 PGE2 (5 ␮M) 298 Ϯ 82 3 a Collagen served as the primary aggregant; PGE2 and/or sulprostone were used to facilitate platelet responses to sub- optimal collagen. b IC50 mean Ϯ SEM were calculated across independent experiments. c N indicates the number of indepen- dent experiments performed. ARTICLE www.acschemicalbiology.org VOL.4 NO.2 • 115–126 • 2009 121
  • 8. dogrel, or a more potent P2Y12 antagonist such as prasugrel, may reduce PGE2-facilitated platelet re- sponses over the inflamed atherosclerotic plaque, but at greater risk of severe or fatal bleeding. We then explored if DG-041 could be co-administered with clopidogrel without further prolonging bleeding, since in the human clinical setting, initial clinical trials in cardiovascular disease would likely require co- administration of DG-041 with current antiplatelet drugs. Coadministration of DG-041 at 5 or 60 mg kgϪ1 with clo- pidogrel efficiently inhibited PGE2 facilitation of platelet responses to collagen and ADP (Figure 5, panels a and b and Supplementary Figure SF-4). (Doses of DG-041 greater than 60 mg kgϪ1 do not provide increased expo- sure due to nonlinear absorption of the compound.) PGE2 facilitation of platelet responses to collagen were completely blocked by adding DG-041 in comparison to clopidogrel alone (P Ͻ 0.01), while facilitation of the platelet response to ADP was significantly reduced (P Ͻ 0.05). A lower dose of DG-041 (5 mg kgϪ1 ) with clopi- dogrel also completely suppresses PGE2 facilitation of platelet responses to collagen ex vivo (Supplementary Figure SF-4). Even at the deliberately high dose of 60 mg kgϪ1 , despite potentiation of the antiplatelet effect of clopidogrel, there was no further increase in bleeding time in rats gavaged with clopidogrel and DG-041 (Figure 5, panel c). This again underscores that PGE2 ac- tivation of EP3 plays little role in hemostasis in response to vascular breach even in the context of effective P2Y12 blockade. Taken together these results suggest that co- administration of DG-041, perhaps with a lower dose of clopidogrel that has minimal effect on bleeding, could be a useful strategy for improving antiplatelet therapy in cardiovascular disease. We successfully employed a ligand-based design strategy to develop potent antagonists of PGE2 binding to EP3. Early in our chemistry program, we realized that proper spatial and electronic overlap for the synthetic molecule required peri-substituted bicyclic aromatic templates. As confirmation of our ligand-based model, a broad diversity of heterocycles were tolerated by the receptor and provided potent EP3 antagonists. Because of both synthetic considerations and potential for meta- bolic transformation of the central core, we focused our Figure 5. Effects of clopidogrel with and without co-administration of DG-041 on rat ex vivo platelet aggregation and on bleeding times. PRP was harvested 4 h (clopidogrel) and 2 h (DG-041) after oral administration. a) Rat platelet aggregation tested with collagen (5 ␮g mL؊1 ) ؎ PGE2 (1 ␮M). Clopidogrel fully inhibited collagen-induced platelet aggregation, whereas the addition of PGE2 restored the aggregatory effect of colla- gen. Co-administration of DG-041 completely blocked PGE2-facilitated aggregation (96% reduction compared to clopidogrel alone). A similar re- duction in platelet aggregation toward collagen and PGE2 was achieved with a high dose of clopidogrel (30 mg kg؊1 ). b) Rat platelet aggregation tested with ADP (0.5 ␮M) ؎ PGE2 (1 ␮M). Clopidogrel provided partial protection toward ADP-mediated platelet aggregation (83% reduction), while addition of PGE2 reversed the inhibitory effect of clopidogrel. Combination of clopidogrel and DG-041 yielded a decrease in platelet aggrega- tion compared to clopidogrel alone (61% reduction), whereas a high dose of clopidogrel (30 mg kg؊1 ) fully blocked aggregation. c) Clopidogrel at 10 mg kg؊1 caused increased bleeding compared to control (from 80 ؎ 4 s to 129 ؎ 16 s), while addition of a high dose of DG-041 (60 mg kg؊1 ) did not further increase bleeding times (95 ؎ 13 s). Increasing the dose of clopidogrel to 30 mg kg؊1 causes a 4-fold increase in bleeding com- pared to controls (237 ؎ 67 s).*P < 0.05, **P < 0.01, ***P < 0.001. 122 VOL.4 NO.2 • 115–126 • 2009 www.acschemicalbiology.orgSINGH ET AL.
  • 9. efforts on SAR studies of the 1,7-disubstituted indole core. Introduction of 3-methyl and 5-fluoro substituents in the aromatic core addressed metabolic liabilities while preserving potency of the respective molecules. Addressing this issue at the earlier stages of the SAR studies allowed us to select the optimal central core for further optimization. In prioritizing our synthetic efforts, we focused on achieving optimal selectivity while reduc- ing plasma protein binding as these parameters were re- ported to be a challenge for similar programs in the in- dustry (19). Indole derivatives combining a lipophilic meta/para-substituted benzylic group at N1 with an acylsulfonamide moiety at C7 of the ring furnished the best results across multiple in vitro and ex vivo assay panels. Additional heterocyclic cores featuring the non- vicinal, peri-substitution pattern and preferred substitu- ents resulted in active compounds with favorable in vitro and ex vivo profiles. Proper selection of the linker yielded additional advantage with regard to the meta- bolic stability of the final molecule. Specifically, a vinyl acylsulfonamide was identified as the optimized C7 substituent. In a comprehensive series of experiments, we confirmed the suitability of this functionality, e.g., low potential for chemical reactivity, lack of antigenicity in the local lymph node assay, and lack of accumulation in 14 C radiolabel distribution studies in the rat. The com- bined SAR and in vitro and ex vivo studies yielded com- pound 12 (designated DG-041) as our candidate for hu- man clinical studies. DG-041 was well tolerated in preclinical safety pharmacology and in 1 and 3 month duration chronic toxicity studies in the rat (up to 150 mg kgϪ1 ) and dog (up to 20 mg kgϪ1 ) following oral dosing (unpublished data). There was no evidence of bleeding in any vascular bed during the in-life phase of the chronic toxicity studies or by histopathological examina- tion. DG-041 was not genotoxic in AMES or in vivo mam- malian (mouse) micronucleus assays (unpublished data). Availability of a potent and selective EP3 antagonist allowed us to explore the pharmacology of the EP3 sig- naling pathway in vitro and in vivo. Consistent with stud- ies of platelets harvested from EP3 null mice (13, 18), DG-041 is a potent inhibitor of PGE2-facilitated re- sponses to multiple platelet agonists when those ago- nists are present at sub-optimal levels. Since PGE2 does not elicit calcium mobilization, PGE2 is not able to trig- ger platelet aggregation in the absence of a co-agonist. DG-041 potently suppresses rat platelet aggregation ex vivo at doses above 10 mg kgϪ1 that achieve plasma lev- els of Ն0.1 ␮M. Even at a dose of 100 mg kgϪ1 , which gave a 190 times higher plasma concentration, no effect on bleeding time was observed after calibrated surgical incision of the tail vein. Our results in rats are consistent with the previous report by Fabre et al. (13) showing that bleeding times are unaffected in EP3 null mice but differ from the report by Ma et al. (18), who claim that bleeding time was increased in EP3 null mice. Since the healthy arterial wall contains only low amounts of PGE2 (11) and since platelets do not produce PGE2, blockade of EP3 should minimally affect hemostasis and that is in fact observed. Agonists released by platelets (such as ADP and TxA2) that amplify the response of surround- ing platelets to the local environment of the fibrin clot are of overwhelming importance for hemostasis, as shown by the impact of P2Y12 antagonists and COX-1 in- hibitors on bleeding. Fabre and co-workers hypothesize that production of PGE2 differentiates acute thrombosis in the context of atherosclerotic plaque rupture from hemostasis in re- sponse to vascular breach (Fabre et al., personal com- munication). COX-2 and mPGES-1, the biosynthetic en- zymes for producing PGE2 are expressed in atherosclerotic plaque (9, 10), and an increase in PGE2 content in plaque has been reported (11, 12). We mod- eled the pro-thrombotic environment of the inflamed plaque by using PGE2 to facilitate platelet responses to collagen. Even at the minimum effective dose of clopi- dogrel (10 mg kgϪ1 ) that blocks ex vivo platelet re- sponses to both collagen and ADP, PGE2 is able to re- store the response of clopidogrel-inhibited platelets to those agonists. We further showed that higher doses of clopidogrel can block PGE2-facilitated responses to col- lagen and ADP, but at a cost. Bleeding times are even further increased, illustrating that the efficacy of clopi- dogrel at controlling acute thrombosis in the context of PGE2 is limited by increasingly severe impact on hemo- stasis. DG-041 efficiently blocks PGE2 facilitation of platelet aggregation, especially with respect to facilita- tion of a matrix signal such as collagen present in the ar- terial wall, and when dosed with clopidogrel, does not further prolong bleeding. In separate studies, we show that as reported previ- ously for EP3 gene deletion (11), pharmacological block- ade of EP3 with DG-041 reduces arterial thrombosis in multiple murine models: superfusion of arachidonic acid over the carotid artery, FeCl3 damage to the carotid ARTICLE www.acschemicalbiology.org VOL.4 NO.2 • 115–126 • 2009 123
  • 10. endothelium, and atherothrombosis in response to me- chanical plaque rupture (Fabreet al., personal communi- cation). In each of these conditions, PGE2 is produced lo- cally and precipitates intra-arterial thrombosis in association with platelet agonists present in the vessel wall. We further show in human phase I clinical studies parallel to the preclinical studies reported herein that DG-041 does not prolong bleeding time in human sub- jects at doses that block PGE2-facilitated platelet re- sponses ex vivo (Human Clinical Trial DG-041-CV-013, deCODE Genetics Inc., unpublished data). Thus, an EP3 antagonist has the potential to improve antiplatelet therapy in myocardial infarction and stroke, where con- trol of atherothrombosis is the therapeutic goal, while minimally impacting bleeding, which limits the benefit of current antiplatelet therapy. METHODS Materials. Sulprostone and PGE2 were from Cayman Chemi- cal. Equine collagen and ADP was from Chrono-Log. 3.2% Na- citrate blood collection tubes were from Greiner Bio-One. Hy- droxylpropyl ␤-cyclodextrin (HP␤CD) was from Roquette, France. Clopidogrel/Plavix pharmacy grade tablets (75 mg) were from Sanofi Pharma-Bristol Myers Squibb. DG-041 was synthesized by deCODE Chemistry, Woodridge IL. All other buffers, salts, and reagents were from Sigma and were used as supplied. Human in Vitro Platelet Aggregation Assay. Fresh blood was collected in citrated tubes by venipuncture from healthy male and female volunteers with informed consent. The subjects had not taken any nonsteroidal anti-inflammatory drugs for at least one week prior to blood collection. Platelet-rich plasma (PRP) isolated from citrate-collected hu- man blood was obtained after centrifugation at 100g for 20 min. Platelet-poor plasma (PPP) was obtained by centrifugation of PRP at 2400g for 10 min at 25 °C. Platelet aggregation was ini- tiated by equine collagen (0.125Ϫ0.25 ␮g mLϪ1 ) and PGE2 (1 ␮M) or sulprostone (0.1 ␮M) and measured by light absor- bance in a Chrono-Log model 490 stirred cell aggregometer at 37 °C for 10 min. In samples with DG-041, the antagonist was preincubated at 25 °C for 10 min. Animals. Experiments were performed using either female Long Evans or male SpragueϪDawley rats obtained from Tac- onic M&B, Denmark. After shipment the animals were allowed to acclimatize for at least 5 days before use. The animals were housed under a standard 12 h light-dark cycle with free access to water and standard laboratory chow. All animal experiments were approved by the Laboratory Animal Ethical Committee in Iceland. Dosing and Formulations. Clopidogrel/Plavix tablets (75 mg) were ground in a morter and dissolved in deionized water. The resulting supension was sonicated at 37 °C for 20 min followed by centrifugation to remove poorly soluble excipients. DG-041 was formulated in a 40% HP␤CD solution in 50 mM Na- phosphate buffer, pH 7.2 for the ex vivo platelet aggregation ex- periments. For the oral route in Table 1, DG-041 was formu- lated in 40% HP␤CD, 0.25% polyvinylpyrrolidone, 0.9% lysine hydrochloride in 50 mM Na-phosphate buffer, pH 7.2. Dosing so- lutions were administered orally by gavage to non-fasting rats in a dosing volume of 5Ϫ10 mL kgϪ1 . Stainless steel feeding needles (0.9 ϫ 38 mm) and 1 mL Omnifit syringes were used for the orogastric gavage. Clopidogrel was dosed 4 h before har- vesting platelets or determination of bleeding times, whereas DG-041 was dosed either 1 or 2 h before. Intravenous dosing in Table 1 was with DG-041 formulated in a 20% HP␤CD solution in 50 mM Na-phosphate buffer, pH 7.2 in a dosing volume of 1 mL kgϪ1 . Rat in Vitro/ex Vivo Platelet Aggregation Assay. Blood samples were collected by cardiac puncture while rats were under isoflu- ran anesthesia. Blood was drawn into 10 mL syringes (Braun) with 3.2% sodium citrate buffer from blood collection tubes. PRP was obtained by centrifugation at 150g for 15 min at 25 °C. PPP was obtained as described above for human samples. Both PRP and PPP were diluted 1:4 with saline before use. Platelet ag- gregation experiments with DG-041 in vitro and ex vivo were per- formed as described above for human platelets, except colla- gen was used at 2 ␮g mLϪ1 and PGE2 and sulprostone were used at 5 and 0.2 ␮M, respectively. Experiments with clopidogrel and a combination of clopidogrel and DG-041 were performed with collagen (2, 3.5, and 5 ␮g mLϪ1 ) or ADP (0.5, 1, and 5 ␮M) as primary agonists in the absence or presence of PGE2 (1 ␮M). Bleeding Times. Bleeding time was determined in either con- scious or mildly anesthetized male SpragueϪDawley rats that were immobilized in a restrainer. Anesthesia was performed us- ing a combination of ketamine (10 mg kgϪ1 ) and xylazine (10 mg kgϪ1 ) in saline, administered intraperitoneally. The tail of each immobilized rat was warmed for 1 min at 40 °C and then dried. A small transverse incision was made in the middle of the tail with a scalpel. Recording of the bleeding time started when the first drop of blood touched a circular filter paper placed be- low the cut. Bleeding was monitored every 10 s until cessation. Data Analysis. Platelet aggregation and bleeding time data was analyzed using Prism 4.03 for Windows (GraphPad Soft- ware). Comparison between individual groups was performed using Student’s t test with data following normal distribution or the non-parametric MannϪWhitney test for data displaying non- gaussian distribution. Statistical significance was determined at the 5% level. Pharmacokinetic parameters were calculated using the soft- ware WinNonlin version 4.1 for Windows. Parameters were de- rived using non-compartmental analysis. Data is reported as mean and standard deviation for each parameter as calculated for individual animals, N ϭ 3. General Chemistry Methods. All reagents and anhydrous sol- vents were obtained from commercial sources and used with- out further purification unless otherwise noted. NMR spectra were recorded at either 400 or 500 MHz in the solvent indicated with TMS as an internal reference. Coupling constants (J) are given in Hz. Column chromatography was carried out in the sol- vents indicated with silica gel. HPLC purity of compounds was measured with a reversed-phase HPLC (Zorbax SB-C18 column, 4.6 ϫ 150 mm, 5 ␮m, 254 nm) with two diverse solvent sys- tems: in system 1, compounds were eluted using a gradient elu- tion 95/5 to 5/95 A/B over 20 min at a flow rate of 1.0 mL minϪ1 where solvent A was aqueous 0.05% TFA and solvent B was ace- tonitrile (0.05% TFA); in system 2, compounds were eluted using a gradient of 95/5 to 5/95 A/C over 20 min at a flow rate of 1.0 mL minϪ1 where solvent A was aqueous 0.05% TFA and sol- vent C was methanol. Elemental analyses were carried out by Galbraith Laboratories, Inc. (Knoxville, TN) or Midwest Microlab, LLC (Indianapolis, IN). Mass spectra were obtained using either 124 VOL.4 NO.2 • 115–126 • 2009 www.acschemicalbiology.orgSINGH ET AL.
  • 11. APCI or electrospray ionization. Schemes and synthesis of com- pounds are described in Supporting Information. Acknowledgment: The authors wish to acknowledge techni- cal contributions of J. Lin for providing guidance and overall ana- lytical support, D. Anderson for running mass spectral analy- ses, and N. Zhao for acquiring NMR spectra. We also thank G. Halldo´rsdo´ttir, A. G. Hansdo´ttir, I. Gylfado´ttir, R. Spilliaert, G. A. Pa´lsdo´ttir, B. N. Nielsen, T. Dagbjartsdottir, B. Gudmundsdo´ttir, S. Hrafnsdo´ttir, and S. Gunnarsdo´ttir for performance of in vitro and in vivo pharmacological assays. Supporting Information Available: This material is available free of charge via the Internet at http://pubs.acs.org. REFERENCES 1. Ross, R. (1999) AtherosclerosisϪan inflammatory disease, N. Engl. J. Med. 340, 115–126. 2. The Physicians’ Health Study: aspirin for the primary prevention of myocardial infarction, (1988) N. Engl. J. Med. 318, 924Ϫ926. 3. Collaborative overview of randomised trials of antiplatelet therapyϪI: prevention of death, myocardial infarction, and stroke by prolonged antiplatelet therapy in various categories of patients. Antiplatelet Trialists’ Collaboration, (1994) Br. Med. J. 308, 81Ϫ106. 4. Bhatt, D., Hirsch, A., Ringleb, P., Hacke, W., and Topol, E. (2000) Re- duction in the need for hospitalization for recurrent ischemic events and bleeding with clopidogrel instead of aspirin. CAPRIE investiga- tors, Am. Heart J. 140, 67–73. 5. Wiviott, S., Braunwald, E., McCabe, C., Montalescot, G., Ruzyllo, W., Gottlieb, S., Neumann, F., Ardissino, D., De Servi, S., Murphy, S., Riesmeyer, J., Weerakkody, G., Gibson, M., and Antman, E. M. (2007) The TRITON-TIMI 38 Investigators. Prasugrel versus clopi- dogrel in patients with acute coronary syndromes, N. Engl. J. Med. 357, 2001–2015. 6. Bhatt, D. (2007) Intensifying platelet inhibitionϪnavigating be- tween Scylla and Charybdis, N. Engl. J. Med. 357, 2078–2081. 7. Funk, C. D. (2001) Prostaglandins and leukotrienes: advances in ei- cosanoid biology, Science 294, 1871–1875. 8. Grosser, T, Fries, S, and FitzGerald, G A. (2006) Biological basis for the cardiovascular consequences of COX-2 inhibition: therapeutic challenges and opportunities, J. Clin. Invest. 116, 4–15. 9. Cipollone, F., Cesaria, P., Barbara, P., Matteo, M., Maria, F., Do- menico, De C., Annalisa, I., Sante, U., Gianfranco, B., Vittorio, S., Francesco, C., Franco, C., and Andrea, M. (2001) Overexpression of functionally coupled cyclooxygenase-2 and prostaglandin E syn- thase in symptomatic atherosclerotic plaques as a basis of prostag- landin E(2)-dependent plaque instability, Circulation 104, 921– 927. 10. Belton, O., Duffy, A., Toomey, S., and Fitzgerald, D. (2003) Cyclooxy- genase isoforms and platelet vessel wall interactions in the apoli- poprotein E knockout mouse model of atherosclerosis, Circulation 108, 3017–3023. 11. Gross, S., Tilly, P., Hentsch, D., Vonesch, J., and Fabre, J. E. (2007) Vascular wall-produced prostaglandin E2 exacerbates arterial thrombosis and atherothrombosis through platelet EP3 receptors, J. Exp. Med. 204, 3113–3120. 12. Cipollone, F., Fazia, M., Iezzi, A., Cuccurullo, C., Cesare, D., Ucchino, S., Spigonardo, F., Marchetti, A., Buttitta, F., Paloscia, L., Mascel- lanti, M., Cuccurullo, F., and Mezzetti, A. (2005) Association be- tween prostaglandin E receptor subtype EP4 overexpression and un- stablephenotypeinatheroscleroticplaquesinhuman,Arterioscler., Thromb., Vasc. Biol. 25, 1925–19331. 13. Fabre, J., Nguyen, M., Athirakul, K., Coggins, K., McNeish, J., Austin, S., Parise, L., FitzGerald, G., Coffman, T., and Koller, B. (2001) Activa- tion of the murine EP3 receptor for PGE2 inhibits cAMP production and promotes platelet aggregation, J. Clin. Invest. 107, 603–610. 14. Sugimoto, Y, and Narumiya, S. (2007) Prostaglandin E receptors, J. Biol. Chem. 282, 11613–11617. 15. Breyer, R. M., Bagdassarian, C. K., Myers, S. A., and Breyer, M. D. (2001) Prostanoid receptors: subtypes and signaling, Ann. Rev. Pharmacol. Toxicol. 41, 661–690. 16. Takeuchi, K, Ukawa, H, Furukawa, O, Kawauchi, S, Araki, H, Sugi- moto, Y, Ishikawa, A, Ushikubi, F, and Narumiya, S. (1999) Prostag- landin E receptor subtypes involved in stimulation of gastroduode- nalbicarbonatesecretioninratsandmice,J.Physiol.Pharmacol.50, 155–167. 17. Paul, B., Ashby, B., and Sheth, S. (1998) Distribution of prostag- landin IP and EP receptor subtypes and isoforms in platelets and hu- man umbilical artery smooth muscle cells, Br. J. Haematol. 102, 1204–1211. 18. Ma, H., Hara, A., Xiao, C. Y., Okada, Y., Takahata, O., Nakaya, K., Sugimoto, Y., Ichikawa, A., Narumiya, S., and Ushikubi, F. (2001) In- creased bleeding tendency and decreased susceptibility to throm- boembolism in mice lacking the prostaglandin E receptor subtype EP(3), Circulation 104, 1176–1180. 19. Juteau, H., Gareau, Y., Labelle, M., Sturino, C., Sawyer, N., Trem- blay, N., Lamontagne, S., Carriere, M., Denis, D., and Metters, K. (2001) Structure-activity relationship of cinnamic acylsulfonamide analogues on the human EP3 prostanoid receptor, Bioorg. Med. Chem. 9, 1977–1984. 20. Selected compounds were incubated with rat liver microsomes and incubateswerefollowedusingLC-MS-MSanalysisformetabolicpro- file. For example, analog 4 (Supplementary Table ST-1) showed Mϩ16 and Mϩ32 metabolites, consistent with mono- and bis- oxidation derived metabolites. 21. Tranilast containing an ␣,␤-unsaturated amide functionality has been reported to be launched in Japan and Korea for the treatment of allergic rhinitis, asthma, and atopic dermatitis by Kissei Phar- maceuticals Co., Ltd. Data source http://www.iddb3.com. 22. For a report that substituents that are particularly effective in stabi- lizing nucleophilic addition to ␣,␤-unsaturated carbonyl com- pounds include aldehydes, ketone, esters and other carboxylic acid derivatives except amide, see:Carroll, F. (1998) Perspectives on Structure and Mechanism in Organic Chemistry, pp 628Ϫ629, Brooks/Cole Publishing Co., Pacific Grove, CA 23. Perlmutter P. (1992) Conjugate Addition Reaction in Organic Synthe- sis, Pergamon Press, Oxford, England. 24. Palomo, C., Oiarbide, M., Dias, F., Ortiz, A., and Linden, A. (2001) J. Am. Chem. Soc. 123, 5602–5603. This source reports that reac- tion of N-enoly oxazolidine-2-thione with Lewis acids yields, after hy- drolytic workup, the corresponding ␤-metcapto carbonyl adducts. This reaction represents intramolecular addition of thiol to a tertiary amide surrogate. 25. Reaction of methyl cinnamate and O-tert-butyldimethyl silyl cin- namate with N-BOC cystamine in MeOH/NaOMe for 12 h at RT pro- vided 68% and 64% yield of the corresponding Michael adducts, respectively, whereas under similar reactions conditions neither cin- namide nor cinnamic acid yielded any desired adduct. 26. Kawasaki, T., Ozeki, Y., Igawa, T., and Kambayashi, J. (2000) In- creased platelet sensitivity to collagen in individuals resistant to low-dose aspirin, Stroke 31, 591–595. 27. Heptinstall, S., Iue Espinosa, D., Manolopoulos, P., Glenn, J. R., White, A. E., Johnson, A., Dovlatova, N., Fox, S. C., May, J. A., Her- mann, D., Magnusson, O., Stefansson, K., Hartman, D., and Gur- ney, M. (2008) DG-041 inhibits the EP3 prostanoid receptorϪa new target for inhibition of platelet function in atherothrombotic dis- ease, Platelets 19, 605–613. 28. Wise, H., Wong, Y., and Jones, R. (2002) Prostanoid signal integra- tion and cross talk, Neurosignals 11, 20–28. 29. Sugidachi, A., Asai, F., Ogawa, T., Inoue, T., and Koike, H. (2000) The in vivo pharmacological profile of CS-747, a novel antiplatelet agent with platelet ADP receptor antagonist properties, Br. J. Pharmacol. 129, 1439–1446. ARTICLE www.acschemicalbiology.org VOL.4 NO.2 • 115–126 • 2009 125
  • 12. 30. Burt, C., Richards, W. G., and Huxley, P. (1990) The application of molecular similarity calculations, J. Comput. Chem. 11, 1139– 1146. 31. Carbo, R., Ledya, L., and Arnau, M. (1980) How similar is a mole- cule to another? An electron density measure of similarity between two molecular structures, Int. J. Quantum Chem. 17, 1185–1189. 126 VOL.4 NO.2 • 115–126 • 2009 www.acschemicalbiology.orgSINGH ET AL.