Skip to main content
British Journal of Pharmacology logoLink to British Journal of Pharmacology
. 2009 Jul 15;158(1):104–145. doi: 10.1111/j.1476-5381.2009.00317.x

Prostanoid receptor antagonists: development strategies and therapeutic applications

RL Jones 1, MA Giembycz 2, DF Woodward 3
PMCID: PMC2795261  PMID: 19624532

Abstract

Identification of the primary products of cyclo-oxygenase (COX)/prostaglandin synthase(s), which occurred between 1958 and 1976, was followed by a classification system for prostanoid receptors (DP, EP1, EP2 …) based mainly on the pharmacological actions of natural and synthetic agonists and a few antagonists. The design of potent selective antagonists was rapid for certain prostanoid receptors (EP1, TP), slow for others (FP, IP) and has yet to be achieved in certain cases (EP2). While some antagonists are structurally related to the natural agonist, most recent compounds are ‘non-prostanoid’ (often acyl-sulphonamides) and have emerged from high-throughput screening of compound libraries, made possible by the development of (functional) assays involving single recombinant prostanoid receptors. Selective antagonists have been crucial to defining the roles of PGD2 (acting on DP1 and DP2 receptors) and PGE2 (on EP1 and EP4 receptors) in various inflammatory conditions; there are clear opportunities for therapeutic intervention. The vast endeavour on TP (thromboxane) antagonists is considered in relation to their limited pharmaceutical success in the cardiovascular area. Correspondingly, the clinical utility of IP (prostacyclin) antagonists is assessed in relation to the cloud hanging over the long-term safety of selective COX-2 inhibitors. Aspirin apart, COX inhibitors broadly suppress all prostanoid pathways, while high selectivity has been a major goal in receptor antagonist development; more targeted therapy may require an intermediate position with defined antagonist selectivity profiles. This review is intended to provide overviews of each antagonist class (including prostamide antagonists), covering major development strategies and current and potential clinical usage.

Keywords: prostaglandin; thromboxane A2; prostacyclin; prostamide; prostanoid receptor antagonist; development strategy; high-throughput screening; acyl-sulphonamide, pA2 values; therapeutic applications

Introduction

The pharmacological classification for prostanoid receptors was developed on the basis that each receptor preferentially recognizes one of the major natural prostaglandins (PGs). Thus, PGD2 preferentially activates the DP receptor, PGE2 preferentially activates the EP receptor, with the same applying to PGF/FP receptor and PGI2 (prostacyclin)/IP receptor (Coleman et al., 1994b). In the case of the TP receptor, both thromboxane A2 (TXA2) and its precursor PGH2 are potent agonists. There are two distinct subtypes of DP receptor, DP1 and DP2; the latter has also been called CRTh2 (chemoattractant receptor-homologous molecule expressed on T helper 2 cells). Of the four EP receptor subtypes, EP1 and EP3 generally elicit excitatory actions, while EP2 and EP4 elicit inhibitory actions on cell function. All prostanoid receptors belong to the G protein-coupled receptor superfamily of cell-surface receptors. Each has seven transmembrane (TM)-spanning segments and may couple to one or more signal-transduction processes. In addition, mRNA splicing variants have been discovered for DP1, EP1, EP3, EP4, FP and TP receptors (see Pierce and Regan, 1998).

The repertoire of bioactive oxygenated lipids and signal transduction mechanisms is expanded by homodimerization and heterodimerization between different prostanoid receptors. For example, an isoprostane binding site is created by IP/TPα heterodimerization (Wilson et al., 2004). Moreover, the prostamide (prostaglandin-ethanolamide) recognition site appears to result from heterodimerization of wild type and alternatively spliced FP receptor variants (Liang et al., 2008). Prostanoid receptors may also complex with non-prostanoid receptors, for example, EP1 receptor with β2-adrenoceptor, but discussion of these interactions is outside the scope of this review. The receptor/second messenger nomenclature used in the review conforms to this journal's Guide to Receptors and Channels (Alexander et al., 2008).

Assay systems and antagonist development

The initial pharmacological differentiation of prostanoid receptors relied heavily on isolated tissue studies, with ileum, trachea and vas deferens of the guinea pig being especially important (Jones et al., 1982; Coleman et al., 1984; 1987; 1994a; Dong et al., 1986; see Chen et al., 2001 for experimental details). Isolated tissue preparations are still used today as they often reflect phenomena that occur in the integrated, living mammal. In addition, they still provide evidence for new receptor entities, as shown by the critical role of the cat iris preparation in the elucidation of prostamide pharmacology (Matias et al., 2004; Woodward et al., 2007; 2008;). Studies on isolated and cultured cells have also been important in building the pharmacological classification of prostanoid receptors (Coleman et al., 1984; Eglen and Whiting, 1988; Woodward et al., 1995a,b;).

Although of low throughput, isolated tissue studies have been critical to the discovery of prostanoid antagonists, particularly for DP1, EP1 and TP receptors. However, it is remarkable that antagonists for other receptors (EP2, EP3, FP and IP) have been slow to emerge following discovery of the natural ligand/receptor; several reasons may account for this situation. Partial agonism is found in closely related analogues of the natural ligands for DP1 and TP receptors (Table 1), thereby providing inroads to pure antagonists. However, partial agonism may not always be readily recognized. For example, taprostene, an early analogue of PGI2 (Müller et al., 1983), was only shown to be an IP partial agonist in 2004 (Chan and Jones, 2004). Furthermore, partial agonism may not always translate into pure antagonism. Thus, some of the many ‘non-prostanoid prostacyclin mimetics’ synthesized (see later) show IP partial agonism (Merritt et al., 1991a; Jones et al., 1997; Seiler et al., 1997; Kam et al., 2001), but IP antagonists have apparently not emerged from this grouping. Moreover, a large range of compounds block TP receptors (see later); this expansiveness may simply not apply to other prostanoid receptors. Finally, there has not been widespread commercial drive to develop EP2 antagonists, for example, owing to a perceived lack of therapeutic utility (see later).

Table 1.

Prostanoid receptor agonists relevant to defining antagonist profiles

Prostanoid receptor Full agonist
Partial agonist
High selectivity Moderate selectivity
DP1 BW-245C BW-192C86a
DP2 15(R) PGD2, 15(R)-15-methyl PGD213,14-dihydro-15-oxo PGD2
EP1 ONO-DI-004 17-Phenyl PGE2 Iloprostb,c
EP2 ONO-AE1-259, CAY-10399d Butaprost-FA, CP-533536e19(R)-hydroxy PGE2f
EP3 ONO-AE-248, SC-46275 Sulprostone, MB-28767 ONO-AP-324g
EP4 ONO-AE1-329, tetrazolo PGE1h (PGE2)i
FP Fluprostenol, latanoprost-FA Cloprostenol AL-8810j,k,l
IP Cicaprost AFP-07, iloprost Octimibatem, taprostenen
TP STA2, U-46619 CTA2o, PTA2o,p, U-44069p

Information on the non-referenced agonists may be obtained from Jones (2004) and this journal's Guide to Receptors and Channels edited by Alexander et al. (2008). In older publications, fluprostenol = ICI-81008, cloprostenol = ICI-80996, cicaprost = ZK-96480 and iloprost = ZK-36374.

FA, free acid.

a

Series of bicyclic-hydantoin prostanoids (Leff and Giles, 1992).

h

Analogue 19a in Billot et al. (2003).

i

Utility on high-sensitivity EP4 systems.

Dramatic progress in antagonist development was seen following cloning of the various prostanoid receptors in the early 1990s: DP1 (Boie et al., 1995), DP2 (Hirai et al., 2001), EP1 (Funk et al., 1993a), EP2 (Regan et al., 1994), EP3 (Yang et al., 1994), EP4 (Bastien et al., 1994), FP (Abramovitz et al., 1994), IP (Namba et al., 1994), TP (Hirata et al., 1991). Stable over-expression of each prostanoid receptor in carrier cell lines allowed high-throughput radioligand binding and functional studies using 96- or even 384-well plate format. Thus, chemical library screening resulted in the discovery of new non-prostanoid scaffolds as leads, from which potent and selective agonists and antagonists were designed. Inspection of Figures 18 will reveal the prevalence of aryl-sulphonamido linkages in the various antagonist classes. Non-prostanoid structures provide a practical approach to obtaining new prostanoid-based therapies, as the issues of bioavailability, metabolic stability and cost of synthesis that surround PG analogues are avoided.

Figure 1.

Figure 1

DP1 receptor antagonists. The natural ligand PGD2 is shown in the box; the trans-orientation of the α (upper) and ω (lower) side-chains and (S) configuration at C15 are found in primary products of all COX/synthase systems. BW-A868C and ZK-138357 are each composed of four diastereoisomers (chiral centres at C8/C15 and C10/C15 respectively); compound 1 is racemic. Indomethacin is a lead compound for the non-prostanoid antagonists shown in the lower row; N-benzoyl-2-methyl-indol-3-yl-acetic acid templates are shown in red. The 3(S)-enantiomer of MK-0524 has 320-fold lower affinity for the human DP1 receptor (Sturino et al., 2007).

Figure 8.

Figure 8

TP receptor antagonists. Conversion of PGH2 to TXA2 by thromboxane synthase (TXS) is shown in the box; α and ω represent natural 2-series side-chains. The pinane-thromboxane residue (related to 1(S)-α-pinene) is shown in blue; the 6-oxabicyclo(2.2.1)heptane system is in red. AH-23848 has the same α- and ω-chains as GR-32191. Benzenesulphonamide residues present in both prostanoid and non-prostanoid antagonists are shown in cerise. TP antagonists with two types of additional activity are presented. (A) IP agonism is conferred by the diphenylmethyl-heteroatomic unit in the bicyclo[2.2.2]octene analogue EP-157. (B) TXS inhibitory activity is conferred by the pyridin-3-yl residue (green) in isbogrel and ZD-1542 and by a similar replacement for ring A in relatives of GR-32191 (e.g. GR-83783; see text). Additionally, the broken arrows (lower right) typically indicate attachment of part of a TP antagonist to isbogrel (or ridogrel) to generate novel combined TP antagonist/TXS inhibitors; the tether (0–11 carbon units) has also been attached to the left-hand phenyl ring of ICI-192605 (2–8 carbon units) (Ackerley et al., 1995).

Selective agonists for prostanoid receptors

Selective agonists are the pharmacological counterparts of antagonists and, as such, fulfil an important role in receptor characterization. In terms of prostanoid receptors, three of the natural ligands, PGH2/TXA2 and PGI2, are unstable under physiological conditions and are usually replaced by U-46619 and cicaprost respectively (Table 1). In addition, there are potential problems with certain EP agonists synthesized as C1 methyl esters, for example, butaprost (EP2) and SC-46275 (EP3); full agonist potency is only realized after (enzymatic) hydrolysis of the ester within the tissue. Ono Pharmaceuticals have synthesized selective agonists for EP1 (ONO-DI-004), EP2 (ONO-AE1-259), EP3 (ONO-AE-248) and EP4 (ONO-AE1-329) receptors (Suzawa et al., 2000). However, the rather modest potencies of ONO-DI-004 and ONO-AE-248 may restrict their utility in full Schild antagonism protocols (R.L. Jones et al., 2008, submitted). Partial agonists for DP1, EP1, EP3, FP, IP and TP receptors are known (Table 1); their use in antagonist protocols may present difficulties of interpretation.

As drugs, notably for systemic administration, selective prostanoid mimetics are a high-risk proposition because of the myriad of unwanted side effects that may occur. As such, the future of prostanoid-based therapies appears to reside in the main in selective PG synthase inhibitors and prostanoid receptor antagonists.

Antagonist protocols

The Schild protocol for inferring the nature of competition and determining the affinity constant of an antagonist remains the gold standard (see Colquhoun, 2007). It is applicable to both isolated tissue and the ever-increasing number of recombinant (rc) receptor/cell-based assay methods usually involving Ca2+ mobilization or cAMP generation. While the latter methods are given to high throughput and precision, care must be taken with high-affinity antagonists in the Ca2+ assays as the agonist response is usually measured as the peak of the transient Ca2+ signal, which may occur before re-equilibration of antagonist occupancy is complete. The majority of the data given in Table 2 relate to pA2 values derived using the Schild protocol, with emphasis on human, guinea pig and rat isolated preparations. Binding data (preferably pKi) are given where functional information is not available. There are considerable binding data on mouse prostanoid receptors, but little affinity data in functional systems.

Table 2.

Affinities of prostanoid receptor antagonists in functional isolated preparations

Antagonist Species Tissue system Agonist pA2 Reference
DP1 receptor
AH-6809 Human Neutrophil/superoxide release BW-245C 6.55 Wheeldon and Vardey (1993)
BW-245C 6.59 Lydford et al. (1996a)
Platelet/aggregation PGD2 6.27a Keery and Lumley (1988)
Cow Embryonic tracheal fibroblast/cAMP PGD2 6.36 Ito et al. (1990)
Rabbit Saphenous vein BW-245C 5.93 Lydford et al. (1996c)
BW-A868C Human Neutrophil/superoxide release BW-245C 9.46 Lydford et al. (1996a)
Platelet/aggregation BW-245C 9.26 Giles et al. (1989)
Pulmonary vein PGD2 7.84 Walch et al. (1999)
Uterus (non-pregnant) BW-245C 8.3 Senior et al. (1992)
Uterus (pregnant) BW-245C 8.6 Senior et al. (1993)
Cow Embryonic tracheal fibroblast/cAMP BW-245C 8.0 Crider et al. (1999)
Dog Nasal vein BW-245C 7.3 Liu et al. (1996a)
Tracheal epithelium/Cl- secretion BW-245C 8.16 Liu et al. (1996b)
Rabbit Jugular vein BW-245C 8.73 Giles et al. (1989)
Saphenous vein BW-245C 8.50 Lydford et al. (1996c)
MK-0524 (Laropiprant) Human rc-DP1/HEK-293E/binding [3H]-PGD2 10.5b Sturino et al. (2007)
Platelet/cAMP PGD2 10.05c Sturino et al. (2007)
ONO-AE3-237 Human rc-DP1/CHO/binding [3H]-PGD2 7.74 Torisu et al. (2004c)
S-5751 Human Platelet/cAMP PGD2 9.02b,c Arimura et al. (2001)
Guinea pig Platelet/cAMP PGD2 7.50b,c Arimura et al. (2001)
ZK-138357 Human Neutrophil/superoxide release BW-245C 7.25 Lydford et al. (1996a)
Rabbit Saphenous vein BW-245C 5.05 Lydford et al. (1996a)
Rat Peritoneal mast cell BW-245C ∼6.0 Chan et al. (2000)
Compound 1 Human Platelet/binding [3H]-PGD2 6.22 Mitsumori et al. (2003a)
Compound 2 Human Platelet/binding [3H]-PGD2 7.62 Mitsumori et al. (2003a)
Compound 3 Human rc-DP1/HEK-293-Gα15/Ca2+ BW-245C [∼7.4] Krauss et al. (2005)
Compound 4 Human rc-DP1/CHO/binding [3H]-PGD2 8.27 Torisu et al. (2004c)
Compound 5 Human rc-DP1/HEK-293/binding [3H]-PGD2 9.0 Beaulieu et al. (2008)
DP2 receptor
BAY-u3405 (Ramatroban) Human rc-DP2/CHO/GTPγS binding PGD2 7.44 Mathiesen et al. (2006)
Eosinophil/shape change PGD2 ∼8.0a Mathiesen et al. (2006)
K-117 Human rc-DP2/HEK-293/binding [3H]-PGD2 8.26 Mimura et al. (2005)
K-604 Human rc-DP2/HEK-293/binding [3H]-PGD2 7.96 Mimura et al. (2005)
TM-30089d (CAY-10471) Human rc-DP2/HEK-293/binding [3H]-PGD2 8.74 Mathiesen et al. (2006)
9.22 Ulven and Kostenis (2005)
Compound 6 Human rc-DP2/pre-B L1.2/Ca2+ PGD2 [6.8] Bauer et al. (2002)
Compound 7 Human rc-DP2/HEK-293/binding [3H]-PGD2 8.64c Birkinshaw et al. (2006)
Compound 8 Human rc-DP2/HEK-293/binding [3H]-PGD2 [9.40]c Bonnert and Rasul (2004)
Compound 9 Human rc-DP2/CHO/binding [3H]-PGD2 7.17 Armer et al. (2005)
Eosinophil/shape change PGD2 7.13c Armer et al. (2005)
Th2-lymphocyte/chemotaxis PGD2 7.17c Armer et al. (2005)
Compound 10 Human rc-DP2/CHO/Ca2+ PGD2 [8.53] Fretz et al. (2005)
Compound 11 Human rc-DP2/HEK-293/binding [3H]-PGD2 [9.0] Bonnert et al. (2005c)
EP1 receptor
AH-6809 Human rc-EP1/HEK-293E/reporter gene Iloprost ∼6.4 Durocher et al. (2000)
Pulmonary vein Sulprostone 5.52 Walch et al. (2001)
Guinea pig Ileum PGE2 6.8 Coleman et al. (1987)
Ileum PGE2 7.39 Eglen and Whiting (1988)
Ileum PGE1 7.42 Eglen and Whiting (1988)
Ileum 16,16-DM PGE2 7.59 Eglen and Whiting (1988)
Trachea 16,16-DM PGE2 7.48 Eglen and Whiting (1988)
Trachea 17-Phenyl PGE2 7.35 Lawrence et al. (1992)
GW-848687 Human rc-EP1/not given/reporter gene PGE2 9.1 Giblin et al. (2007)
MF-266-1 Human rc-EP1/HEK-293/Ca2+ PGE2 7.8 Clark et al. (2008)
ONO-8711 Mouse rc-EP1/CHO/binding [3H]-PGE2 8.77 Watanabe et al. (1999)
Human rc-EP1/CHO/binding [3H]-PGE2 9.22 Watanabe et al. (1999)
ONO-8713 Mouse rc-EP1/not given/binding [3H]-PGE2 9.5 Narumiya and Fitzgerald (2001)
SC-19220 Guinea pig Ileum PGE2 5.5 Sanner (1969)
Ileum PGE2 5.6 Bennett and Posner (1971)
Trachea PGF 6.6 Farmer et al. (1974)
SC-51089 Human rc-EP1/HEK-293E/reporter gene Iloprost 6.94 Durocher et al. (2000)
Guinea pig Ileum PGE2 6.5 Hallinan et al. (1993)
Ileum PGE2 6.7 Sametz et al. (2000)
SC-51322 Human rc-EP1/HEK-293E/reporter gene Iloprost 8.80 Durocher et al. (2000)
Guinea pig Ileum PGE2 8.1 Hallinan et al. (1994)
Trachea 17-Phenyl PGE2 8.45 Hung et al. (2006)
Compound 13 Mouse rc-EP1/CHO/Ca2+ PGE2 8.25c Naganawa et al. (2006)
Compound 14 Human rc-EP1/HEK-293E/binding [3H]-PGE2 8.0 Ruel et al. (1999)
Compound 15 Human rc- EP1/CHO/Ca2+ PGE2 8.2 Hall et al. (2007b)
EP2 receptor
AH-6809 Human rc-EP2/COS-7/cAMP PGE2 ∼6.5 Woodward et al. (1995)
Bronchus PGE2 5.78 Norel et al. (1999)
Guinea pig Trachea PGE2 5.7e KJ Ong and RL Jones (unpublished)
EP3 receptor
DG-041 Human rc-EP3/Chem-1/Ca2+ PGE2 8.09c Singh et al. (2009)
L-798106 Guinea pig Aorta 17-Phenyl PGE2f 7.96 Jones et al. (2008)
Trachea Sulprostone 7.82 Clarke et al. (2004)
Vas deferens Sulprostone 7.48 Clarke et al. (2004)
L-826266 Human rc-EP3/HEK-293E/Ca2+ PGE2 7.97 R.L. Jones et al. (2008, submitted)
Erythroleukaemia cell/cAMP Sulprostone 8.35 Clark et al. (2008)
Guinea pig Aorta 17-Phenyl PGE2 7.58 R.L. Jones et al. (2008, submitted)
ONO-AE3-240 Mouse rc-EP3/not given/Ca2+ PGE2 8.8 Amano et al. (2003)
Compound 17 Human Erythroleukaemia cell/cAMP Sulprostone 6.89 Gallant et al. (2002)
Compound 18 Human rc-EP3/HEK-293E/binding [3H]-PGE2 7.7 Juteau et al. (2001)
Compound 19 Human rc-EP3/HEK-293E/cAMP PGE2 8.22 Belley et al. (2005)
EP4 receptor
AH-23848 Sheep Ductus arteriosus PGE2 ∼5.2a Bouayad et al. (2001)
Human Middle cerebral artery PGE2 5.7 Davis et al. (2004)
Pig Saphenous vein PGE2 5.0 Coleman et al. (1994a)
Rabbit Saphenous vein PGE2 4.96 Lydford et al. (1996b)
Mouse rc-EP4/CHO/cAMP PGE2 5.3 Nishigaki et al. (1995)
BGC-20-1531 Human rc-EP4/HEK-293E/cAMP PGE2 7.6 Maubach et al. (2009)
Middle cerebral artery PGE2 7.8 Maubach et al. (2009)
Dog Middle meningeal artery PGE2 7.7 Maubach et al. (2009)
CJ-023423 Human rc-EP4/HEK-293/cAMP PGE2 8.3 Nakao et al. (2007)
Rat rc-EP4/HEK-293/cAMP PGE2 8.2 Nakao et al. (2007)
CJ-042794 Human rc-EP4/HEK-293/cAMP PGE2 8.6 Murase et al. (2008b)
Rat rc-EP4/HEK-293/cAMP PGE2 8.7 Murase et al. (2008a)
GW-627368 Human rc-EP4/HEK-293/cAMP PGE2b 7.9 Wilson et al. (2006)
Pulmonary vein ONO-AE1-329 7.06 Foudi et al. (2008)
Pig Saphenous vein PGE2 9.2 Wilson et al. (2006)
Rabbit Saphenous vein PGE2g ≥8.5 Jones and Chan. (2005)
L-161982 Human rc-EP4/HEK-293/cAMP PGE2 ∼8.5 Machwate et al. (2001)
BEAS-2B cell/CRE reporter ONO-AE1-329 9.14 L.M. Ayer and M.A. Giembycz (unpublished)
Middle cerebral artery PGE2 8.4 Davis et al. (2004)
Rat rc-EP4/HEK-293/binding [3H]-PGE2 7.50 Machwate et al. (2001)
Periosteal cell/cAMP PGE2 7.0c Machwate et al. (2001)
MF-498 Human rc-EP4/HEK-293/cAMP PGE2 8.77c Clark et al. (2008)
ONO-AE2-227 Mouse rc-EP4/CHO/cAMP PGE2 8.0c Mutoh et al. (2002)
ONO-AE3-208 Mouse rc-EP4/not given/binding [3H]-PGE2 8.89 Kabashima et al. (2002)
Compound 20 Human rc-EP4/HEK-293E/cAMP PGE2 8.49c Burch et al. (2008)
FP receptor
AS-604872 Human rc-FP/HEK-293E/PI PGF 7.33c Cirillo et al. (2007)
THG-113 Pig Retinal blood vessel PGF [6.34]c Peri et al. (2006)
THG-113.31 Pig Retinal blood vessel PGF [8.00]c Peri et al. (2006)
THG-113.824 Pig Retinal blood vessel PGF [8.96]c Peri et al. (2006)
THG-113.825 Pig Retinal blood vessel PGF 7.21c Peri et al. (2006)
Prostamide F receptor
AGN-204396 Cat Iris sphincter Prostamide F 5.64 Woodward et al. (2007)
IP receptor
RO-1138452 Human rc-IP/CHO/cAMP Carbacyclin 9.0 Bley et al. (2006)
Pulmonary artery Cicaprost 8.20 Jones et al. (2006)
Guinea pig Aorta Cicaprost 8.39 Jones et al. (2006)
Rabbit Mesenteric artery Cicaprost 8.12 Jones et al. (2006)
RO-3244794 Human rc-IP/CHO/cAMP Carbacyclin 8.5 Bley et al. (2006)
BEAS-2B cell/CRE reporter Taprostene 9.24 L.A. Ayer and M.A. Giembycz (unpublished)
Compound 21 Rat rc-IP/not given/cAMP (Not given) [8.12]h Keitz et al. (2004)
Compound 22 Rat UMR-106 osteosarcoma cell/cAMP Iloprost 6.41 Nakae et al. (2005)
Compound 23 Rat UMR-106 osteosarcoma cell/cAMP Iloprost 6.32 Nakae et al. (2005)
Compound 24 Human Platelet membrane/cAMP Iloprost 7.8c Brescia et al. (2007)
TP receptor
AA-2414 (Seratrodast) Human Bronchus U-46619 7.7 Itoh et al. (1993)
Pig Coronary artery U-44069 9.0 Imura et al. (1990)
Guinea pig Aorta U-46619 8.5 Zhang et al. (1996)
Trachea U-46619 7.69 Ashida et al. (1989)
Rat Aorta U-46619 7.8 Zhang et al. (1996)
AH-23848 Human Bronchus U-46619 8.5 Coleman and Sheldrick (1989)
Bronchus U-46619 6.9 McKenniff et al. (1988)
Hand vein U-46619 8.4 Arner et al. (1991)
Lung parenchyma U-46619 8.7 McKenniff et al. (1988)
Platelet/aggregation U-46619 8.05 Tymkewycz et al. (1991)
Guinea pig Lung parenchyma U-46619 8.7 McKenniff et al. (1988)
Trachea U-46619 8.7 McKenniff et al. (1988)
Trachea U-46619 9.76 Tymkewycz et al. (1991)
Rat Aorta U-46619 8.47 Tymkewycz et al. (1991)
Lung parenchyma U-46619 6.9 Norman et al. (1992)
Platelet/aggregation U-46619 8.19 Tymkewycz et al. (1991)
BAY-u3405 (Ramatroban) Human Bronchus U-46619 8.8 McKenniff et al. (1991)
Pulmonary vein U-46619 8.94 Walch et al. (2001)
Guinea pig Lung parenchyma U-46619 7.7 Norman et al. (1992)
Trachea U-46619 8.7 McKenniff et al. (1991)
Rat Lung parenchyma U-46619 8.6 McKenniff et al. (1991)
BM-13177 (Sulotroban) Human Platelet/aggregation U-46619 6.31 Karasawa et al. (1991a)
Guinea pig Aorta U-46619 5.58 Karasawa et al. (1991b)
Rabbit Jugular vein U-46619 6.01 Giles et al. (1989)
BM-13505 Human Hand vein U-46619 7.9 Arner et al. (1991)
Platelet/aggregation U-46619 7.75 Karasawa et al. (1991a)
Uterus (non-pregnant) U-46619 7.4 Senchyna and Crankshaw (1996)
Guinea pig Aorta U-46619 6.89 Dubéet al. (1992)
Aorta U-46619 7.22 Karasawa et al. (1991b)
Aorta U-46619 7.3 Zhang et al. (1996)
Lung parenchyma U-46619 7.0 Norman et al. (1992)
Trachea U-46619 7.73 Dubéet al. (1992)
Trachea U-46619 7.5 Ogletree and Allen (1992)
Rat Aorta U-46619 8.2 Zhang et al. (1996)
Aorta U-46619 8.6 Ogletree and Allen (1992)
Lung parenchyma U-46619 7.5 Norman et al. (1992)
BMS-180291 (Ifetroban) Human Platelet membrane/binding [3H]-SQ-29548 8.4 Ogletree et al. (1993)
Guinea pig Aorta U-46619 9.8 Zhang et al. (1996)
Rat Aorta U-46619 9.5 Zhang et al. (1996)
CV-4151 Human Platelet/aggregation U-46619 5.2 Watts et al. (1991)
Rabbit Aorta U-44069 5.9 Imura et al. (1988)
EP-092 Human Bronchus U-46619 7.37 Featherstone et al. (1990)
Bronchus U-46619 6.8 McKenniff et al. (1988)
Lung parenchyma U-46619 8.9 McKenniff et al. (1988)
Platelet /aggregation U-46619 7.73 Tymkewycz et al. (1991)
Uterine artery U-46619 8.5 Baxter et al. (1995)
Guinea pig Lung parenchyma U-46619 8.7 McKenniff et al. (1988)
Trachea U-46619 7.29 Featherstone et al. (1990)
Trachea U-46619 8.7 McKenniff et al. (1988)
Trachea U-46619 8.02 Tymkewycz et al. (1991)
Rat Aorta U-46619 8.55 Tymkewycz et al. (1991)
Lung parenchyma U-46619 7.1 Norman et al. (1992)
Platelet aggregation U-46619 7.80 Tymkewycz et al. (1991)
EP-169 Human Platelet/aggregation U-46619 8.30 Tymkewycz et al. (1991)
Pulmonary artery U-46619 ∼8.2 Qian et al. (1994)
Guinea pig Trachea U-46619 8.77 Tymkewycz et al. (1991)
Rat Aorta U-46619 8.73 Tymkewycz et al. (1991)
Platelet/aggregation U-46619 8.48 Tymkewycz et al. (1991)
Glibenclamide Human Internal mammary artery U-46619 6.3i Stanke et al. (1998)
Saphenous vein 6.7i Stanke et al. (1998)
Dog Coronary artery U-46619 6.2 Cocks et al. (1990)
Rabbit Aorta U-46619 6.08a Pfister et al. (2004)
Guinea pig Aorta U-46619 <5.0 Kemp and McPherson (1998)
Rat Aorta U-46619 6.13 Kemp and McPherson (1998)
GR-32191 (Vapiprost) Human Bladder (detrusor) U-46619 8.27 Palea et al. (1998)
Bronchus U-46619 8.77 Featherstone et al. (1990)
Bronchus U-46619 8.40 Armour et al. (1989)
Platelet/binding [3H]-GR-32191 8.66 Armstrong et al. (1993)
Pulmonary artery U-46619 8.18 Lumley et al. (1989)
Saphenous vein U-46619 8.93 Furci et al. (1991)
Umbilical artery U-46619 8.0 Boersma et al. (1999)
Uterine artery U-46619 8.5 Baxter et al. (1995)
Uterus (non-pregnant) U-46619 8.6 Senchyna and Crankshaw (1996)
Uterus (pregnant) U-46619 8.5a Senior et al. (1993)
Guinea pig Aorta U-46619 8.77 Lumley et al. (1989)
Aorta U-46619 9.4 Ogletree and Allen (1992)
Trachea U-46619 8.26 Featherstone et al. (1990)
Trachea U-46619 9.43 Tymkewycz et al. (1991)
Trachea U-46619 10.0 Ogletree and Allen (1992)
Rat Aorta U-46619 7.87 Lumley et al. (1989)
Aorta U-46619 7.49 Furci et al. (1991)
Aorta U-46619 8.41 Tymkewycz et al. (1991)
Aorta U-46619 8.3 Ogletree and Allen (1992)
Trachea U-46619 8.31 Lydford and McKechnie (1994)
GR-83783j Rat Aorta U-46619 7.5 Campbell et al. (1991a)
GR-108774j Rat Aorta U-46619 9.2 Campbell et al. (1991b)
ICI-192605 Human Platelet/aggregation U-46619 8.16 Brewster et al. (1988)
Umbilical artery U-46619 8.1 Boersma et al. (1999)
Umbilical vein U-46619 9.07 Daray et al. (2003)
Uterus (non-pregnant) U-46619 9.2 Senchyna and Crankshaw (1996)
Rat Aorta U-46619 8.4 Brewster et al. (1988)
I-PTA-OH Guinea pig Lung parenchyma U-46619 5.6 Norman et al. (1992)
Rat Lung parenchyma U-46619 5.8 Norman et al. (1992)
I-SAP Human Platelet/aggregation U-46619 8.01k Naka et al. (1992)
KW-3635 Human Platelet/aggregation U-46619 8.88 Karasawa et al. (1991a)
Guinea pig Aorta U-46619 7.74 Karasawa et al. (1991b)
L-655240 Guinea pig Aorta U-44069 8.0 Hall et al. (1987)
Aorta U-44069 8.0 Hall et al. (1987)
ONO-11120 Human Platelet/binding [125I]-PTA-0H 7.71c Narumiya et al. (1986)
Platelet/aggregation U-46619 7.49 Tymkewycz et al. (1991)
Guinea pig Trachea U-46619 8.07 Tymkewycz et al. (1991)
Rat Aorta U-46619 7.14 Tymkewycz et al. (1991)
Platelet/aggregation U-46619 7.38 Tymkewycz et al. (1991)
ONO-NT-126 Human Astrocytoma cell/PI STA2 10.0 Nakahata et al. (1990)
Ridogrelj Human Platelet/aggregation U-46619 5.7 Watts et al. (1991)
Rat Tail artery U-46619 5.5 Janssens et al. (1990)
(±)-S-145 (Domitroban)l Human Astrocytoma cell/PI STA2 8.48 Nakahata et al. (1990)
Platelet membrane/binding [3H]-(+)-S-145 9.35 Kishino et al. (1991)
Rat Aorta smooth muscle cell/binding [3H]-SQ-29548 9.5a Hanasaki et al. (1988)
S-18886 (Terutroban) Rabbit Saphenous vein U-46619 8.9 Cimetière et al. (1998)
SQ-29548 Human Astrocytoma cell/PI STA2 8.08 Nakahata et al. (1990)
Immortalized ciliary epithelial cell/PI U-46619 7.7m Sharif et al. (2002)
Corpus cavernosum U-46619 9.0 Angulo et al. (2002)
Umbilical artery U-46619 7.6 Boersma et al. (1999)
Umbilical vein U-46619 7.96 Daray et al. (2003)
Uterus (non-pregnant) U-46619 8.2 Senchyna and Crankshaw (1996)
Pig Coronary artery U-46619 8.8a Kromer and Tippins (1996)
Rabbit Aorta U-46619 7.95 Yoshida et al. (2007)
Guinea pig Aorta U-46619 7.96 Dubéet al. (1992)
Aorta U-46619 8.9 Ogletree and Allen (1992)
Aorta U-46619 8.5 Zhang et al. (1996)
Lung parenchyma U-46619 7.7 Norman et al. (1992)
Trachea U-46619 8.70 Dubéet al. (1992)
Trachea U-46619 8.9 Ogletree and Allen (1992)
Rat Aorta U-46619 9.2 Zhang et al. (1996)
Lung parenchyma U-46619 7.2 Norman et al. (1992)
SQ-30741 Human Coronary artery U-46619 7.54 Maassen VanDenBrink et al., 1996)
Human Umbilical artery U-46619 7.0 Boersma et al. (1999)
Guinea pig Aorta U-46619 8.1 Ogletree and Allen (1992)
Trachea U-46619 8.6 Ogletree and Allen (1992)
Rat Aorta U-46619 7.9 Ogletree and Allen (1992)
YM-158 Guinea pig Trachea U-46619 8.81n Arakida et al. (1998)
Z-335 Human Platelet membrane/binding [3H]-SQ-29548 7.52 Tanaka et al. (1998)
Platelet/shape changeo U-46619 8.02 Yoshida et al. (2007)
Rabbit Aorta U-46619 8.64 Yoshida et al. (2007)
ZD-1542j Guinea pig Lung parenchyma U-46619 8.5 Brownlie et al. (1993)
Trachea U-46619 8.3 Brownlie et al. (1993)
Rat Aorta U-46619 8.51 Brownlie et al. (1993)

pA2 values relate to functional assays. Recombinant (rc-) systems: prostanoid receptor followed by the carrier cell line and second messenger measurement. Smooth muscle preparations: contraction or relaxation of induced tone. Platelets: all data from plasma-free platelet suspensions. Where pA2 values are not available, pKi/pKD values derived from radioligand binding are given (italics). Affinity values in square brackets derive from the patent literature.

Structures of compounds: 1–5, Figure 1; 6–12, Figure 2; 13–15, Figure 3; 16–19, Figure 4; 20, Figure 5; 21–24, Figure 7.

BEAS, human bronchial epithelium; CHO, Chinese hamster ovary; COS-7, African green monkey kidney; HEK, human embryonic kidney; CRE, cAMP response element; 16,16-DM PGE2, 16,16-dimethyl PGE2; PI, phosphoinositide.

a

Our calculation.

b

Appreciable affinity for corresponding TP receptor.

c

pIC50.

d

Insurmountable antagonism in functional DP2 systems.

e

1 µM SC-51322 present.

f

Replacement for sulprostone, which had slow onset and offset.

g

Possible interference by EP2 system.

h

pKi.

i

pKb for non-competitive antagonism.

j

Combined TP antagonist/TXS inhibitor.

k

Shape change seen.

l

Data for (+)- and (−)-enantiomers in Kishino et al. (1991).

m

Non-competitive antagonism.

n

pA2= 8.87 for LTD4 antagonism.

o

Suppression of maximum aggregation response.

Inhibition-curve (or Cheng-Prusoff) protocols have been used infrequently in prostanoid receptor studies. Bley et al. (2006) estimated the pA2 of the IP antagonists RO-1138452 and RO-3244794 using carbacyclin as the fixed-concentration agonist in a human rc-IP receptor – cAMP assay, but failed to use the modified form of the Cheng-Prusoff equation (Craig, 1993; Lazareno and Birdsall, 1993; Leff and Dougall, 1993). These protocols have the advantages of operating over a lower agonist concentration range than Schild protocols and providing direct observation of the rate of onset of antagonism. The latter is important in recognizing the slow approach to steady state that occurs with high-affinity antagonists at low concentration, and also with highly lipophilic antagonists (Jones et al., 2008) that regularly emerge from combinatorial chemistry – high-throughput screening. Ultimately, it is important and even preferable to define the pharmacology in the human target tissue; human rc-receptor assays are a useful accompaniment. Successful drugs require the correct pharmacological attributes, but physical chemical properties are also important.

DP1 receptor antagonists

Development

Antagonists for, what we now know to be, the DP1 receptor subtype, were first described in the 1970s and early 1980s. Examples include N-0164 (MacIntyre and Gordon, 1977), diphloretin phosphate (Westwick and Webb, 1978) and desacetyl-1-nantradol (Horne, 1984), but none of these compounds exhibit the potency and selectivity essential for unambiguous receptor classification. The simple xanthone-carboxylic acid AH-6809 (Keery and Lumley, 1988) has sufficient DP1 affinity (pA2= 5.9–6.6, Table 2), but it has been mainly employed as an EP1 antagonist (see later). Indeed, the hydantoin derivative, BW-A868C (Figure 1) has been the only selective, surmountable and competitive DP1 antagonist (pA2 > 9 for the human subtype) readily available to pharmacologists (Giles et al., 1989; Lydford et al., 1996c). The N-benzyl substituent is crucial for antagonist activity; in a related series of bicyclic-hydantoin analogues a progression from full agonism to virtually pure antagonism is seen with hydrogen, methyl, ethyl and n-propyl substituents on N10 (Giles and Leff, 1992). BW-A868C also has low affinity (pA2= 5.1) for the EP4 subtype (Lydford et al., 1996c). ZK-138357 (Schering AG) is a moderate-affinity DP1 antagonist (Table 2) with some structural similarity to BW-A868C.

Recently, the potential pathological role of PGD2, especially in allergic disorders, has been revived resulting in the discovery and evaluation of highly selective DP1 antagonists of several structural classes. The most important of these (Figure 1) and their therapeutic applications are described below.

Bicycloheptanes

Chemists in Shionogi have synthesized selective DP1 antagonists containing a bicyclo[2.2.1]heptane ring system akin to that present in PGH2 (see inset in Figure 8) (Tsuri et al., 1997; Honma et al., 1998; Mitsumori et al., 2003a). An initial lead was the racemic compound 1 previously shown to be a TP antagonist (Narisada et al., 1988). Subsequent structure–activity relationship (SAR) studies revealed that a 6,6-dimethylbicyclo[3.1.1]heptane (pinane) system could substitute for the bicycloheptane ring (Tsuri et al., 1997; Mitsumori et al., 2003b; Yoshikawa et al., 2005) and carbonylamino or sulphonylamino linkages to the ω-aryl moiety were required for potent DP1 antagonism (Tsuri et al., 1997; Honma et al., 1998; Mitsumori et al., 2003b). Accordingly, S-5751 has high affinity for the DP1 receptor (pKi= 8.8) and is orally active in models of allergy and inflammation in the guinea pig (Tsuri et al., 1997; Arimura et al., 2001; Mitsumori et al., 2003b; Yasui et al., 2008). In addition, lead optimization of the (+)-isomer of compound 1, which has much higher DP1 selectivity than its mirror-image, led to the 1-methoxy-dibenzo[b,d]furan 2. This antagonist is potent, orally bioavailable and efficacious in guinea pig models of conjunctivitis and allergen-induced bronchoconstriction (Mitsumori et al., 2003a).

Allergan have also filed patents claiming DP1 receptor antagonism based on a 1(S),4(S)-7-oxabicyclo[2.2.1]heptane scaffold (Krauss et al., 2005). One of these, compound 3, has a pA2 of ∼7.4 for the human rc-DP1 receptor and represents a logical structure for lead optimization.

Indole acetic acids

A non-prostanoid exploited for DP1 antagonism at Ono Pharmaceuticals is the cyclo-oxygenase (COX) inhibitor/non-steroidal anti-inflammatory drug (NSAID), indomethacin (Figure 1). Initial studies showed that the acetic acid moiety could be switched to position 4 on the indole ring (Torisu et al., 2004a). Optimization led to the discovery of two benzoxazines, ONO-AE3-237 and compound 4, with high DP1 antagonist selectivity and (sub)nanomolar affinity (Torisu et al., 2004b,c,d; Torisu et al., 2005). Administered orally, both compounds effectively suppressed PGD2- and allergen-induced vascular permeability in the guinea pig conjunctiva (Torisu et al., 2004c). Pharmacokinetic studies on ONO-AE3-237 given by the oral (10 mg·kg−1) and intravenous (1 mg·kg−1) routes to fasted rats afforded plasma half-lives of 7.8 and 9.2 h respectively. The compound has a high volume of distribution indicating good tissue penetration and is 48% bioavailable at a dose of 10 mg·kg−1 p.o. (Torisu et al., 2004c).

Merck Frosst (Wang et al., 2002; Berthelette et al., 2003) and Sanofi-Aventis (Yang et al., 2008) also filed patents claiming DP1 antagonists with further variation of the acetic acid position on the indole template. Screening of the Merck compound collection identified a difluoro-indole that had low-nanomolar affinity for DP1 and TP receptors (Sturino et al., 2006) and this, and other analogues, were optimized resulting in two 7-methylsulphone derivatives with high DP1 affinity (Ki∼2 nM) and at least 100-fold selectivity over other prostanoid receptors. However, both compounds displayed poor pharmacokinetics in the rat, in particular extensive biliary excretion (Sturino et al., 2007). This liability was overcome when the 5-substituent and the 7-methylsulphone on the indole ring were replaced by methylsulphone and fluorine, respectively, to give MK-0542 (laropiprant; Figure 1). Laropiprant has very high affinity for the human rc-DP1 receptor (pKi= 10.5), with ∼300-fold lower affinity for the corresponding TP receptor (Sturino et al., 2007). Excellent pharmacokinetic profiles have been found in the rat, dog, monkey and man (Chang et al., 2007; Karanam et al., 2007; Sturino et al., 2007; Lai et al., 2008b). Given orally to healthy male volunteers, laropiprant at single doses up to 900 mg and multiple doses up to 450 mg is rapidly absorbed (Tmax= 0.8–2 h), demonstrates dose-proportional systemic exposure, has a half-life of 12–18 h and is generally well tolerated; this pharmacokinetic profile is unaffected by food (Karanam et al., 2007; Lai et al., 2008b). At a dose of 6 mg, laropiprant was effective in antagonizing PGD2-induced cAMP accumulation in human platelets ex vivo, indicating an interaction with the desired molecular target. Evidence for TP receptor blockade was also detected, but this effect was deemed not to be clinically relevant (Lai et al., 2008b).

Merck Frosst has also disclosed ‘backup’ DP1 antagonists, in which the indole template present in laropiprant is inverted. The tetrahydropyridoindole 5 (Figure 1) exhibited the best profile (pKi= 9.0 and 6.8 at DP1 and TP receptors respectively) and is considered a suitable candidate for development (Beaulieu et al., 2008).

Other structural classes

Certain aminopyrimidines have also been claimed in the patent literature to be DP1 antagonists but neither in vitro nor in vivo pharmacological data are yet available (Langevin et al., 2007; Stefany et al., 2007).

Therapeutic applications

PGD2 is an established mediator of allergic disease. It is the major prostanoid released from mast cells (Lewis et al., 1982; Peters et al., 1982) and is also secreted, albeit in lower amounts, by T-lymphocytes of the Th2 subset (Tanaka et al., 2000). In asthma, dermatitis and rhinitis, allergen challenge leads to the rapid production of PGD2 (Naclerio et al., 1983; Murray et al., 1986; Charlesworth et al., 1991) and PGD2, itself, can reproduce many symptoms associated with allergic phenomena (see Pettipher, 2008). However, therapeutic/commercial success has not yet been attained with DP1 antagonists probably because functional and subsequently molecular evidence emerged for a second subtype of PGD2-sensitive receptor that is strongly implicated in several manifestations of allergic disease including eosinophil infiltration, mucus hyper-secretion and plasma extravasation (see next section).

In terms of allergic inflammation, activation of DP1 receptors is known to mediate pathological changes in blood flow. In allergic rhinitis, vessels within the nasal mucosa become engorged leading to congestion and the release of plasma proteins, which contribute to enhanced nasal secretions. These effects are mimicked by PGD2, which explains the limited efficacy of histamine H1 antagonists in allergic rhinitis (see Pettipher, 2008). Sturino et al. (2007) have shown that laropiprant abolishes the marked increase in nasal airway resistance induced by intranasal instillation of PGD2 in conscious sheep. Significantly, comparable data have also been obtained in 15 healthy, non-smoking male volunteers in whom laropiprant (25 mg or 100 mg q.d. for 3 days) significantly suppressed PGD2-induced nasal congestion (Van Hecken et al., 2007). Interestingly, PGD2 fails to lower diastolic blood pressure (BP) in human volunteers (Heavey et al., 1984) although (BW)-192C86, a DP1 partial agonist (Gray et al., 1992), is an effective depressor and BW-245C evokes adverse cardiovascular effects consistent with vasodilatation (Al Sinawi et al., 1985). The reason for this discrepancy is unknown, but given that PGD2 is a potent vasoconstrictor in several species (Jones, 1976; 1978;) and the DP2 receptor is expressed in the aorta (Nagata and Hirai, 2003) and potentially other blood vessels, its activation may oppose the BP-lowering activity of PGD2 acting via the DP1 receptor.

The ability of PGD2 to promote adverse vasodilatation is also associated with the therapeutic use of niacin (vitamin B3), which, in high doses, is used clinically to lower plasma cholesterol. Niacin may also have anti-oxidant and anti-inflammatory activity and is used, in conjunction with statins, to treat dyslipidaemia (see Kamanna et al., 2008). The adverse effects of niacin, particularly flushing, are due to receptor (GPR109A)-mediated release from Langherhans' cells of PGD2 and PGE2, which promote vasodilatation of dermal and cerebrovascular capillaries though activation of DP1, EP2 and EP4 receptors. Merck (Paolini et al., 2008) and Sanofi-Aventis (Harris, 2008) have claimed that a DP1 antagonist could limit the cardiovascular liability of niacin if given as a combination therapy, thereby providing a better-tolerated drug. Indeed, laropiprant significantly reduced niacin-induced flushing in normal and dyslipidaemic subjects when compared with niacin alone (Cheng et al., 2006; Lai et al., 2007; Paolini et al., 2008). Accordingly, these and other findings led Merck to develop the investigational combination product Cordaptive, which significantly reduced the vascular side effects of niacin in patients with primary hypercholesterolaemia and mixed dyslipidaemia (see Kamanna et al., 2008). However, in April 2008 the US Food and Drug administration did not approve Merck's application to market Cordaptive (see http://www.merck.com/newsroom/press_releases/research_and_development/2008_0428.html), and at the same time, rejected the name of the combination therapy. The US Food and Drug administration's decision not to approve Cordaptive, now renamed Tredaptive, is unclear given that the European Medicines Agency has approved this new extended-release combination therapy (see http://www.emea.europa.eu/humandocs/Humans/EPAR/tredaptive/tredaptive.htm). One possibility is that the beneficial effects of laropiprant are not superior to aspirin (Kamanna et al., 2008) and the risk/benefit ratio is a primary consideration when the long-term effects of a new drug class are unknown.

DP1 antagonists may also have utility in the treatment of allergic asthma, although, currently, this is controversial (see Pettipher, 2008). Persuasive evidence is available that activated mast cells, through their ability to generate PGD2, promote T-lymphocyte polarization towards a Th2 phenotype (Faith et al., 2005; Hammad and Lambrecht, 2008). Mechanistically, it is believed that mast cell-derived PGD2 activates DP1 receptors on dendritic cells within the respiratory mucosa. This interaction then suppresses the generation of interleukin (IL)-12 (which normally favours a Th1 cell bias) leading to Th2 dominance (Kitawaki et al., 2006; Theiner et al., 2006). A cycle of chronic immunological activation then ensues through further production of IgE and mast cell activation. Indeed, these data are consistent with allergic airway responses being less intense in DP1 receptor-deficient mice when compared with wild-type animals (Matsuoka et al., 2000). However, the role of DP1 receptors in regulating allergic reactions is complex. For example, the administration to wild-type mice of the DP1 agonist, BW-245C, paradoxically reduced pulmonary allergic responses whereas DP1 receptor null mice were unaffected (Hammad et al., 2007). To explain this apparently contradictory result it has been proposed that DP1 receptor-mediated inhibition of IL-12 release from dendritic cells during host sensitization promotes Th2 polarization (i.e. is pro-inflammatory). In contrast, during maintained airway inflammation, PGD2 by suppressing dendritic cell function reduces Th2 cell function (i.e. is anti-inflammatory). Indeed, the DP1 antagonist, S-5751, attenuated rather than exacerbated allergen-induced inflammation in sensitized guinea pigs (Arimura et al., 2001). Thus, the harmful/protective actions of PGD2 may depend on when and where it is produced (Pettipher, 2008).

DP2 receptor antagonists

Development

Studies conducted in the 1970s showed that PGD2 elicited peripheral vasoconstrictor responses that were unlikely to be due to activation of, what we now understand to be, DP1, FP or TP receptors (Jones, 1976; 1978;). In particular, 15-oxo PGD analogues (Jones and Wilson, 1978) had unexpectedly high agonist potency. Moreover, evidence for DP receptor heterogeneity was provided in 1985 by the finding that PGD2 and six related analogues evoked functional responses across a variety of PGD2-sensitive systems that were not mimicked by the selective DP agonist, BW-245C (Narumiya and Toda, 1985). The results of several other studies also indicated the existence of multiple DP receptors (Woodward et al., 1990a; 1993b; Rangachari and Betti, 1993; Fernandes and Crankshaw, 1995; Rangachari et al., 1995). However, it was not until 1999 that the idea of multiple DP-receptors really began to gain general acceptance. Nagata et al. (1999b) identified a novel molecule expressed on the surface of minor populations of CD4+ T-lymphocytes, which resembled activated Th2 cells in that they released IL-4, IL-5 and/or IL-13 but not interferon-γ on stimulation. This orphan site was named ‘chemoattractant receptor-homologous molecule expressed on Th2 cells’, or CRTh2, as primary sequence analysis showed significant amino acid homology to receptors that mediate chemoattraction (Hirai et al., 2001). Indeed, despite the ability of PGD2 to interact with the DP1 subtype and CRTh2 with comparably high affinity (Ki= 45 and 61 nM respectively), the two cognate receptors are quite dissimilar at the amino acid level (Hirai et al., 2001). A follow-up study by the same investigators discovered a mast cell-derived factor that could increase the cytosolic free Ca2+ concentration in CRTh2-expressing cells (Nagata et al., 1999a), which was identified as PGD2 (Hirai et al., 2001). This latter finding coincided with the publication of a pharmacological study in which Monneret et al. (2001) found that PGD2 was chemotactic for eosinophils and also up-regulated the expression of CD11b and L-selectin by a mechanism that was neither mimicked by BW-245C nor blocked by the DP1 antagonist, BW-A868C. Thus, a second DP receptor subtype (DP2, aka CD294) was identified on Th2 cells and eosinophils that mediated responses diametrically opposite to the inhibitory effects classically associated with DP1 agonism. Given that CRTh2 is now known to be ubiquitously expressed within (Nagata et al., 1999a) and outwith (Hirai et al., 2001; Nagata and Hirai, 2003; Kostenis and Ulven, 2006; Kim and Luster, 2007) the immune system (Sawyer et al., 2002; Nagata and Hirai, 2003), the term ‘DP2’ is a more appropriate designation and is used throughout this review.

The ability of PGD2 to act as a chemoattractant for pro-inflammatory cells and to release Th2-like cytokines has resulted in a concerted effort by the pharmaceutical industry to synthesize selective DP2 antagonists. Indeed, such compounds may be useful in suppressing a myriad of Th2-driven inflammatory pathologies including asthma, otitis, contact dermatitis and rhinitis. At the time of writing, in excess of 90 patents had been filed claiming selective DP2 antagonists. In the sections below the main structural classes are described and affinity estimates of lead antagonists (Figure 2) are given in Table 2.

Figure 2.

Figure 2

DP2 receptor antagonists. Inverted 2-methyl-indole-acetic acid residues (compare with 6) are highlighted in red; ramatroban has an extra methylene (C2a). The phenylacetic acid moiety is shown in blue in fenclofenac, a lead molecule for compound 11. Compound 12, K-117 and K-604 contain a tetrahydroquinoline residue (green).

Indole acetic acids

In addition to providing a scaffold for the development of DP1 antagonists, indomethacin (Figure 1) is also a selective, albeit weak, DP2 agonist (Hirai et al., 2002; Stubbs et al., 2002). Exploiting this property, Pfizer first reported a benzothiazole derivative 6 that had a pA2 of 6.8 and was ∼40-fold selective for the DP2 receptor (Bauer et al., 2002). Subsequently, several patents describing highly potent and selective DP2 antagonists were filed by AstraZeneca (Baxter et al., 2003a,b; Birkinshaw et al., 2003; Bonnert et al., 2003; 2004; 2005a,b,c; Bonnert and Rasul, 2004). An initial hit was a 7-chloroquinoline derivative of indomethacin (pA2∼7 for human rc-DP2 receptor), which also inhibited COX-1 with high potency (Birkinshaw et al., 2006). Inversion of the indole template (Figure 2) and substitution of the 5-methoxy moiety by methyl increased antagonist potency by 23-fold and reduced COX-1 inhibition by a factor of 10 (Birkinshaw et al., 2006). Addition of chlorine at position 8 of the quinoline to give 7 also increased DP2 affinity by 13-fold (IC50 for inhibition of [3H]-PGD2 binding = 2.3 nM). This derivative has a preferred biological profile with relatively weak binding to plasma proteins and good bioavailability in rats (76%) and dogs (100%), with half-lives of 1.7 and 5.3 h respectively (Birkinshaw et al., 2006). Compounds having an arylthio substituent on position 3 of the indole nucleus also display very potent antagonism at the DP2 receptor; compound 8 has a binding IC50 of 0.4 nM (Bonnert and Rasul, 2004).

Related 1-acetic acid derivatives from Oxagen containing methylene or sulphonyl spacers between the aromatic moieties have been reported as potent DP2 antagonists (Middlemiss et al., 2005a,b,c,d; Armer et al., 2006; Lovell, 2007), including compound 9, which has a Ki of 68 nM and a DP2/DP1-selectivity ratio of approximately 150. Functionally, 9 potently inhibits DP2 receptor-mediated human eosinophil shape change and Th2 cell chemotaxis with IC50 values of 74 and 67 nM respectively (Armer et al., 2005). Moreover, this compound is metabolically stable, has no inhibitory effect on five of the major cytochrome P450 enzymes (1A2, 2C19, 2C9, 2D6, 3A4) and fails to induce CYP3A4, CYP1A and CYP2C9. In rats, 9 is 56% bioavailable and has a half-life of 5.5 h following oral administration. Oxagen has reported the development of a lead compound, ODC9101 (aka OC459), which is in Phase IIa clinical trials for asthma. According to the company's website, ODC9101 has completed safety evaluations, is orally active and suitable for once-a-day dosing. Oxagen has also reported the development of a pre-clinical back-up molecule, OC499, and DP2 antagonists for non-oral delivery (OC1768) and topical administration (OC2125, OC2184; see http://www.oxagen.co.uk/pdfs/CRTH2summary.pdf). The structures of these compounds have not been disclosed.

Athersys has also described a series of indole acetic acid derivatives with potent DP2 antagonist activity (Bennani et al., 2006) including substituted 3-benzylphthalazin-1(2H)-ones which have radioligand binding IC50 values in the low nanomolar range; their development status is unknown.

Ramatroban and analogues

Ramatroban (Bay u3405) was originally described as a TP antagonist with a pA2 of ∼8.8 on human tissues (McKenniff et al., 1991). Later studies revealed DP2 antagonism (Sugimoto et al., 2003), albeit of lower affinity (pA2= 7.44; Mathiesen et al., 2006). The structural similarity of ramatroban to compounds 79 is clear. These observations were the impetus for the synthesis of compounds with increased selectivity for the DP2 receptor (Arimura et al., 2003). An example is 10, in which the amide nitrogen atom is included in the tricyclic system. This compound has high affinity for the DP2 receptor (pA2= 8.53) with significantly reduced TP receptor-blocking activity (Fretz et al., 2005).

Other minor changes to ramatroban also resulted in compounds with a high degree of DP2 selectivity. 7TM Pharma reported that N-methylating the sulphonamide or truncating the propionate moiety to acetate produces very selective (>1000-fold over DP1 and TP), high-affinity DP2 antagonists with Ki values of 1.9 (TM-30642) and 0.51 nM (TM-30643) respectively (Ulven and Kostenis, 2005). Furthermore, making both modifications to produce TM-30089 (aka CAY-10471) preserved the DP2 affinity of TM-30643 and further increased the DP2/TP-selectivity ratio to >10 000 (Ulven and Kostenis, 2005). Interestingly, in functional studies (e.g. [35S]-GTPγS binding/inositol phosphate accumulation; PGD2-induced eosinophil shape change), ramatroban and TM-30642 are surmountable competitive antagonists whereas TM 30643 and TM 30089 suppress the maximal response in a concentration-dependent manner (Mathiesen et al., 2006). The insurmountable behaviour of TM-30089 and TM-30643 may be due to its slow dissociation from the DP2 receptor, which also results in long-lasting antagonism (Mathiesen et al., 2006). Whether such pharmacological behaviour occurs in vivo and would be therapeutically advantageous is currently unclear. However, this could be a desirable property as slowly dissociating drugs should act much longer than would be predicted from their plasma half-lives (Mathiesen et al., 2006). 7TM Pharma, in partnership with Ortho-McNeil-Janssen Pharmaceuticals, has a compound in late-stage lead optimization although neither the structure nor the profile of the antagonist has been disclosed (see http://www.7tm.com/News.aspx?M=News&PID=42&NewsID=39).

Phenyl acetic acids

Another NSAID, fenclofenac (Figure 2), provided the starting point for the synthesis of DP2 antagonists based on a phenylacetic acid template. An initial hit claimed in the original patent filed by Pfizer (Bauer et al., 2002) had a 4-chlorophenylthio substituent resulting in a functional DP2/DP1-selectivity ratio of 40. Several other companies, including AstraZeneca, have since filed patents for bis-ether derivatives such as 11, which has a binding pIC50 of 9.0 (Bonnert et al., 2005c).

Tetrahydroquinolines

Millennium and Warner-Lambert (now Pfizer) were the first to disclose DP2 antagonists within the tetrahydroquinoline class (Awad et al., 2004; Ghosh et al., 2004; 2005; Kuhn et al., 2004). These compounds are unique in that they are non-acidic indicating that a carboxylic acid moiety is not essential for DP2 antagonism, as previously assumed (see Pettipher et al., 2007). The 4-amino-tetrahydroquinoline 12 (Figure 2) is reported to gain access to the cerebrospinal fluid after oral dosing and also is efficacious in animal models of inflammation at an oral dose of 25 mg·kg−1 (Corradini et al., 2005). However, at the time of writing detailed pharmacological data on non-acidic DP2 antagonists is sparse. Researchers at Kyowa Hakko Kogyo have reported Ki values of 5.5 and 11 nM for K-117 and K-604 respectively, with minimal interaction with TP or DP1 receptors at concentrations up to 1 µM (Mimura et al., 2005).

Therapeutic applications

Arguably, allergic inflammation is the primary indication for antagonists that selectively block the DP2 receptor. Indeed, the gene encoding this receptor shows a particularly strong association with asthma in Chinese and African-American populations (Huang et al., 2004). Moreover, there is good evidence from in vitro and in vivo studies in laboratory animals that PGD2, acting via the DP2 receptor, can mediate many of the cardinal features of allergic airways inflammation (see Ulven and Kostenis, 2006; Pettipher, 2008 for detailed reviews). The most important observations that have led to this view can be summarized as follows:

  1. PGD2 and selective DP2 agonists promote chemotaxis of eosinophils, basophils and CD4+ T-lymphocytes of the Th2 subset and this effect is abolished by a neutralizing anti-DP2 receptor antibody (Hirai et al., 2001; Monneret et al., 2001).

  2. PGD2 promotes pulmonary eosinophilia in rats; this effect is mimicked by selective DP2, but not DP1, agonists and is abolished by ramatroban (Almishri et al., 2005; Shiraishi et al., 2005).

  3. In guinea pigs, the DP2 agonist, Δ12-PGJ2, mobilizes eosinophils from the bone marrow (Heinemann et al., 2003).

  4. The DP2 agonist, 13,14-dihydro-15-oxo PGD2, promotes pulmonary eosinophilia and exacerbates histopathology in a murine model of allergic asthma (Spik et al., 2005).

  5. Ramatroban and DP2 antagonists devoid of TP receptor-blocking activity reduce pulmonary eosinophilia in several animal species in response to allergen challenge (Nagai et al., 1995; Uller et al., 2007; Pettipher, 2008).

  6. PGD2 promotes the production of Th2 cytokines in vivo including IL-4, IL-5 and IL-13 (Fujitani et al., 2002) and in vitro, this can occur in the absence of allergen or co-stimulatory molecules (Xue et al., 2005).

  7. The expression of the DP2 receptor on eosinophils is up-regulated in atopic individuals (see Kostenis and Ulven, 2006).

  8. High concentrations of PGD2 are present in the airways of asthmatic subjects after antigen challenge (Murray et al., 1986).

  9. The expression of the DP2 receptor on Th2 T-lymphocytes is up-regulated in individuals sensitized to pollen or house dust mite or who have atopic dermatitis (Iwasaki et al., 2002).

The proof of this line of argument is the marketing of ramatroban in Japan under the trade name Baynas for the treatment of perennial allergic rhinitis; its clinical efficacy (e.g. reduction of symptoms and of chronic nasal swelling) has been attributed to DP2 receptor blockade (Terada et al., 1998).

In addition to anti-allergic indications, DP2 antagonists may have utility in combating neuropathic pain (Corradini et al., 2005) where the up-regulation of COX-2 and the subsequent formation of PGs are central to disease pathophysiology (Camu et al., 2003).

Hybrid DP1/DP2 receptor antagonists

In considering the SAR data described in the preceding sections, indole acetic acids may be a fruitful starting point for antagonists that block both DP1 and DP2 receptors. Indeed, in the context of allergic diseases, a hybrid antagonist may exert clinically relevant, beneficial effects that are not achieved when just one DP receptor is targeted. Thus, blockade of the DP1 subtype would prevent PGD2 from inhibiting the generation of IL-12 from dendritic cells, thereby inhibiting the polarization of T-lymphocytes to a Th2 phenotype that occurs during host sensitization. Antagonism of the DP2 receptor would, at the same time, suppress the chemotactic activity of PGD2 towards eosinophils, basophils and T-lymphocytes and so reduce pulmonary leukocyte burden and inflammatory status. Theoretically, this latter action would overcome any DP1 receptor-mediated anti-inflammatory effect of PGD2 on dendritic cells during maintained inflammation (Pettipher, 2008). It is also tempting to speculate that a non-selective DP antagonist that also has TP receptor-blocking activity could be efficacious in allergic asthma. In this scenario, TP receptor blockade would reduce the ability of PGD2 to induce bronchoconstriction, which in humans is mediated through the TP receptor (Beasley et al., 1989). Again, indole acetic acids and ramatroban analogues may provide good templates for optimization.

EP1 receptor antagonists

Development

Figure 3 shows the commonly used EP1 antagonists; while some may be considered as prostanoids, none is structurally close to PGE2. The first EP1 antagonist was SC-19220, a dibenzoxazepine hydrazide (Sanner, 1969). Although of low affinity (pA2= 5.5), SC-19220 proved useful in the early characterization and elucidation of the roles of EP receptors. For example, on guinea pig trachea, 3–50 µM SC-19220 (and indomethacin) suppressed inherent tone, indicating that PGE2 acting via EP1 receptors is the likely mediator (Farmer et al., 1974). Also, SC-19220 at 10 µM equally inhibited matching contractions of guinea pig trachea induced by 16,16-dimethyl PGE2, iloprost and isocarbacyclin, but had no effect on contractions induced by U-46619 or histamine; these results demonstrated that certain PGI2 analogues could potently activate EP1 receptors (Dong et al., 1986).

Figure 3.

Figure 3

EP1 receptor antagonists. The natural ligand PGE2 is shown in the box. The dibenzoxazepine residue in SC-51322 is shown in blue. Aryl-sulphonamido residues in antagonists with prostanoid and non-prostanoid structures are shown in cerise; ONO-NT-012 contains a styryl-sulphonamido moiety. The 1,2-biaryl-cyclopentene pharmacophores in GW-848687 and MF-266-1 are shown in red. Ring A in GW-848687 is part of a picolinic acid (pyridine-2-carboxylic acid) residue.

SC-19220 (7.5–30 mg·kg−1 i.p.) inhibited carrageenan-induced inflammation in the rat without suppressing PGE2 levels at the injury site (Barbieri et al., 1977). These and other observations were the stimulus for the synthesis of higher-affinity EP1 antagonists with potential as anti-inflammatory/analgesic drugs in man; alteration of the acetyl group was the most rewarding strategy. SC-51089 was unusual in the series in releasing hydrazine and was dropped from development (Hallinan et al., 1993). The thioether SC-51322, which is much more potent than the corresponding sulphone (Hallinan et al., 1994), has become the agent of choice for receptor characterization. It behaved competitively over the range 25–625 nM in a human rc-EP1 receptor – reporter gene assay (pA2= 8.8; Schild plot slope = 0.91; Durocher et al. (2000). In rat hepatocytes, SC-51322 at 100 nM abolished DNA synthesis induced by PGE2 or 17-phenyl PGE2 (Table 1), while 1 µM SC-51322 did not affect responses to either PGF or PGI2; the selective involvement of EP1 receptors in the PGE response is clear (Kimura et al., 2000; 2001;). However, not all inferences about EP1 receptor involvement are as secure. In a study on rat progenitor Leydig cells, 3 and 30 µM SC-51322 suppressed IL-1β expression induced by 10 µM 17-phenyl PGE2 by 28% and 59% respectively (Walch et al., 2003). Given that this system responded well to 100 nM cloprostenol, a potent FP agonist (Table 1), the action of 17-phenyl PGE2 and its inhibition by SC-51322 may also have involved FP receptors. SC-51322 at 10 µM did not inhibit PGE2-mediated inhibition of superoxide generation in human blood neutrophils (EP2 system) (Kanamori et al., 1997) or inhibit sulprostone-induced contraction of rat femoral artery (EP3 system) at 1 µM (Hung et al., 2006).

Another widely used EP1 antagonist is AH-6809. At 0.1–10 µM, it blocked EP1-mediated actions of PGE2, 16,16-dimethyl PGE2 or 17-phenyl PGE2 in an apparently competitive manner (pA2= 7.4) (Coleman et al., 1987; Eglen and Whiting, 1988; Lawrence et al., 1992). It did not block EP3 receptor-mediated contraction of guinea pig ileum (Lawrence et al., 1992) and human pulmonary artery (Qian et al., 1994) at 2 and 5 µM respectively and had no effect on the pre-synaptic EP3 action of sulprostone on rat trachea at 3 µM (Rackéet al., 1992). However, AH-6809 blocks EP2 receptors (see later) as well as DP1 and TP receptors in human-washed platelets (Keery and Lumley, 1988) with pA2 values of ∼6.3 and ∼5.9 respectively (our calculation). Of further concern is the inhibitory effect of AH-6809 (3–10 µM) on platelet-activating factor (PAF)- and ADP-induced aggregation, which was attributed to inhibition of phosphodiesterase(s) (Keery and Lumley, 1988). In our hands, AH-6809 at 3–10 µM caused similar partial block of the contractile actions of phenylephrine (α1), histamine (H1), U-46619 (TP) and ONO-AE-248 (EP3) on guinea pig aorta (R.L. Jones et al., 2009, submitted). AH-6809 has often been used at even higher concentrations, possibly because of its high water solubility. For example, 30–300 µM AH-6809 inhibited the contractile action of PGE2 in pig large cerebral artery; EP1 receptor involvement was inferred (Jadhav et al., 2004). However, the pA2 corresponding to 30 µM AH-6809 is only 5.2 (our calculation) and the block was insurmountable at the higher concentrations. With more potent and selective EP1 antagonists now available, it is time to relegate AH-6809 to its place in the historical development of prostanoid antagonists.

The EP1 antagonists developed by Ono Pharmaceuticals (Figure 3, middle row) demonstrate an interesting progression from the TP antagonist ONO-11120 (see Figure 8; Katsura et al., 1983) to a related pinane analogue (ONO-NT-012) showing EP1, FP and TP antagonism (and EP3 agonism), to a bicyclo[2.2.2]octane analogue (ONO-8711) showing EP1/EP3 antagonism, and, finally, to the non-prostanoids ONO-8713 and compound 13 with high selectivity for the EP1 receptor. ONO-8711, ONO-8713 and 13 have KD values for mouse rc-EP1 receptors of 1.7, 0.3 and 0.14 nM respectively (Watanabe et al., 1999; 2000; Naganawa et al., 2006). Small modifications to 13 can restore EP3 antagonist affinity. In rat-cultured mesangial cells, ONO-8713 at 1 µM abolished induction of the transforming growth factor-β-fibronectin cascade elicited by PGE2 under high-glucose conditions; a COX-2-PGE2-EP1 receptor drive was postulated to contribute to deleterious changes in diabetes (Makino et al., 2002). Ohnishi et al. (2001) showed that 10 µM ONO-8713 partially inhibited PGE2-induced exocytosis in mucous cells from guinea pig antrum (IC50∼1 µM). However, 17-phenyl PGE2 was a very weak agonist. The authors postulated that EP1 and EP4 receptors co-operate to sustain the high exocytotic response to PGE2. Norel et al. (2004) also showed that 10 µM ONO-8713 partially suppressed the contractile action of sulprostone on human pulmonary vein; functional EP1 receptors were postulated, even though the selective EP1 agonist ONO-DI-004 was a very weak agonist. Prostanoid receptor binding data alone do not guarantee functional selectivity at these high antagonist concentrations; inclusion of control agonists (both prostanoid and non-prostanoid) in the system under test is essential.

Other pharmacophores for potent EP1 antagonism, with inactivity against cytochrome P450 enzymes and good penetration into the central nervous system being secondary goals (see later), have emerged within the last 10 years. The series reported by Merck (Ruel et al., 1999) contains a tricyclic system akin to that in the Searle series. Again EP1/EP3 selectivity can be readily modulated, as shown by replacement of the terminal phenyl group in compound 14 (Ki values for human EP1 and EP3 receptors = 10 and 4000 nM) by methyl (770 and 1000 nM). EP1 antagonists with aryl groups attached to adjacent carbons of a 5-membered ring (cyclopentene, thiophene, pyrrole) have been reported by research groups at GlaxoSmithKline and Merck Frosst. GW-848687 has nanomolar affinity for the human EP1 receptor, 30-fold lower affinity for the human TP receptor and >400 times lower affinity for other prostanoid receptors (Giblin et al., 2007). The picolinic acid residue in GW-848687 (ring A) is highly acidic (pKa ∼1.0) and there has been considerable SAR work by both research groups on modulating the acidity of this region. The Merck antagonist MF-266-1, with a m-C(CF3)-(OH)2 substituent (Figure 3), is a weak acid (pKa = 7.5 for first ionization) that retains high EP1 affinity (Ducharme et al., 2005; Clark et al., 2008). A m-C(CF3)2-OH substituent in the Merck series and a p-C(CF3)2-OH substituent in the GlaxoSmithKline series resulted in lower EP1 affinity (Ducharme et al., 2005; Hall et al., 2007a). Further modifications to ring A (e.g. m-NH(C=O)CH2Ph/p-(C=O)NHCH(CH3)Ph substituents) resulted in non-acidic relatives with high EP1 potency. Compound 15 is also a non-acidic EP1 antagonist (Hall et al., 2007b).

Therapeutic applications

The upper portion of Table 3 shows that parenteral administration of EP1 antagonists of different chemical classes suppressed the allodynic/hyperalgesic signs of inflammation in the rat and mouse. Intraplantar injection of these antagonists also opposed pain-producing stimuli, although at the mouse EP1 receptor AH-6809 had minimal affinity, which suggests an off-target effect. That said, Khasar et al. (1993) showed that PGE2 and SC-19220 are only mutually antagonistic when they are both injected intradermally and not subcutaneously. More recently, attention has focused on the role of EP receptors in central nociception (see Svensson and Yaksh, 2002; Hefferan et al., 2003; Mebane et al., 2003). Minami et al. (2001) showed that PGE2 could induce hyperalgesia in the mouse when injected intrathecally over the dose-range 0.00035–350 pmol. Deletion of the EP3 receptor gene removed the more sensitive component of the hyperalgesia. In addition, the selective EP3 agonist ONO-AE-248 induced hyperalgesia at relatively low doses. Surprisingly, the EP1 knock-out mouse showed a hyperalgesic response in the hot-plate test, thereby confounding the role of EP1 receptors in mediating hyperalgesia (Minami et al., 2001). EP1 antagonists given intrathecally suppress flinching, and mechanical allodynia and hyperalgesia (Table 3), although in certain cases the doses required for these effects are high. For example, in the studies by Omote et al. (2002), 100 µg ONO-8711 was injected intrathecally in a volume of 10 µl, representing an injectate concentration of 23 mM! Even after dilution in the cerebrospinal fluid, EP3 receptors are likely to be blocked as well.

Table 3.

Anti-nociceptive and anti-inflammatory activities of EP receptor antagonists

Route/species Experimental model Noxious stimulus/site Antagonist (nominal receptor) Antagonist dosing Efficacy Reference
Oral
Guinea pig Joint pain (mono-osteoarthritis) Iodoacetate/shoulder joint MF-498 (EP4) 30 mg·kg−1 +++ Clark et al. (2008)
Thermal hyperalgesia L-902688a/paw MF-498 (EP4) 0.1–30 mg·kg−1 +++ Clark et al. (2008)
Rat Writhing response Acetic acid/peritoneal cavity SC-19220 (EP1) 50–300 mg·kg−1 ++ Drower et al. (1987)
Behavioural response Formalin/paw SC-19220 (EP1) 50–150 mg·kg−1 ++ Drower et al. (1987)
Chronic mechanical/thermal hyperalgesia/allodynia Nerve constriction/sciatic ONO-8711 (EP1) 10–100 mg·kg−1 per day +++ Kawahara et al. (2001)
Chronic hyperalgesia (weight-bearing) Freund's adjuvant/knee joint GW-848687 (EP1) 30 mg·kg−1 per day +++ Giblin et al. (2007)
Chronic arthritis (oedema/radiology) Freund's adjuvant/paw Compound 20 (EP4) 0.005 mg·kg−1 per day +++ Burch et al. (2008)
Local oedema Freund's adjuvant/paw MF-266-1 (EP1) Each 0.008–20 mg·kg−1 daily for 10 days 0 Clark et al. (2008)
MF-266-3 (EP3) 0
MF-498 (EP4) +++
Mechanical hyperalgesia Carrageenan/paw CJ-042794 (EP4) 1–30 mg·kg−1 +++ Murase et al. (2008a)
Local oedema Freund's adjuvant/paw CJ-042794 (EP4) 1–30 mg·kg−1 +++ Murase et al. (2008a)
Mechanical hyperalgesia Carrageenan/paw CJ-023423 (EP4) 3–100 mg·kg−1 +++ Nakao et al. (2007)
Thermal hyperalgesia PGE2/paw CJ-023423 (EP4) 1–29 mg·kg−1 +++ Nakao et al. (2007)
Chronic inflammation (weight-bearing) Freund's adjuvant/paw CJ-023423 (EP4) 10–57 mg·kg−1 +++ Nakao et al. (2007)
Chronic inflammation (oedema, histology) Freund's adjuvant/paw CJ-023423 (EP4) 29–96 mg·kg−1 +++ Okumura et al. (2008)
Mouse Writhing response Phenylbenzylquinone/peritoneal cavity SC-51089 (EP1) 1–30 mg·kg−1 +++ Hallinan et al. (1993)
Writhing response Phenylbenzylquinone/peritoneal cavity SC-51322 (EP1) 1–30 mg·kg−1 +++ Hallinan et al. (1994)
Migration of Langerhans cells into lymph node Fluorescein isothiocyanate/skin ONO-AE3-208 (EP4) 10 mg·kg−1 twice daily ++ Kabashima et al. (2003)
Ear swelling/histology UVB/skin of ear ONO-AE3-208 (EP4) 10 mg·kg−1 ++ Kabashima et al. (2007)
Intravenous
Rat Sensory discharge Acetic acid/bladder lumen ONO-8711 (EP1) 1–3 mg·kg−1 ++ Ikeda et al. (2006)
Visceromotor reflex Distension/bladder DG-041 (EP3) 10 mg·kg−1 +++ Su et al. (2008a)
Intraperitoneal
Rat Mechanical hyperalgesia Freund's adjuvant/paw AH-23848 (EP4) 0.1–10 mg·kg−1 +++ Lin et al. (2006)
Thermal hyperalgesia Freund's adjuvant/paw AH-23848 (EP4) 0.1–10 mg·kg−1 +++ Lin et al. (2006)
Mouse Mechanical allodynia Sulprostone/intrathecal ZM-325802 (EP1) 0.03 µg·kg−1 +++ Gil et al. (2008)
Topical to skin
Mouse Oedema Chronic UVB ONO-8713 (EP1) 96 nmol ++ Tober et al. (2006)
Intraplantar
Rat Paw-withdrawal PGE2/pawb SC-19220 (EP1) 2.3 nmol ++ Khasar et al. (1993)
Mechanical hyperalgesia Incision wound/paw ONO-8711 (EP1) 4.5–114 nmol +++ Omote et al. (2001)
Chronic mechanical hyperalgesia Partial nerve transection/sciatic SC-19220 (EP1) 2.3 nmol +++ Syriatowicz et al. (1999)
Chronic thermal hyperalgesia Partial nerve transection/sciatic SC-19220 (EP1) 2.3 nmol ++ Syriatowicz et al. (1999)
Mouse Mechanical allodynia PGE2/paw AH-6809 (EP1) 100 nmol 0 Kassuya et al. (2007)
L-826266 (EP3) 0.1–10 nmol +++
L-161982 (EP4) 10 nmol 0
Carrageenan/paw L-826266 (EP3) 10 nmol ++ Kassuya et al. (2007)
L-161982 (EP4) 10 nmol ++
Paw-licking PGE2/paw AH-6809 (EP1) 10–100 nmol 0 Kassuya et al. (2007)
L-826266 (EP3) 3–30 nmol ++
L-161982 (EP4) 1–30 nmol +++
Intrathecal
Rat Mechanical allodynia Nerve ligation/L5 spinal root SC-51322 (EP1) 22–220 nmol ++ Hefferan et al. (2003); O'Reilly and Loomis (2007)
Mechanical hyperalgesia Incision wound/foot ONO-8711 (EP1) 114–228 nmol ++ Omote et al. (2002)
Thermal hyperalgesia Incision wound/paw ONO-8711 (EP1) 114–128 nmol 0 Ikeda et al. (2006)
Flinching (late phase) Formalin/paw SC-58109 (EP1) 65–650 nmol ++ Malmberg et al. (1994)
Mechanical hyperalgesia (late phase) Carrageenan/paw ONO-8711 (EP1) 2.28–228 nmol +++ Nakayama et al. (2002)
Mechanical hyperalgesia (late phase) PGE2/intrathecal ONO-8711 (EP1) 0.11–0.45 nmol +++ Nakayama et al. (2004)
Visceromotor reflex Distension/bladder L-798106 (EP3) 10–300 nmol +++ Su et al. (2008b)
DG-041 (EP3) 10–100 nmol +++ Su et al. (2008b)
Mouse Thermal hyperalgesia PGE2/intrathecal AH-6809 (EP1) 0.29–2.9 nmol +++ Uda et al. (1990)
Writhing response PGE2/intrathecal AH-6809 (EP1) 0.29–2.9 nmol +++ Uda et al. (1990)
Mechanical allodynia PGE2/intrathecal AH-6809 (EP1) 4.2 nmol 0 Minami et al. (1995)
Thermal hyperalgesia PGE2/intrathecalc AH-6809 (EP1) 4.2 nmol +++ Minami et al. (1995)
Mechanical allodynia PGE2/intrathecal ONO-NT-012 (?)d 3 × 10−5–0.03 nmol +++ Minami et al. (1995)
Thermal hyperalgesia PGE2/intrathecal ONO-NT-012 (?) 3 × 10−5–0.03 nmol 0 Minami et al. (1995)
Topical to spinal tissue
Rat Mechanical allodynia Bicuculline/L5–L6 spine SC-51322 (EP1) 0.22–17.5 nmol +++ Zhang et al. (2001)

Upper and lower panels relate to parenteral and topical administration of antagonist.

a

Selective EP4 agonist.

b

Intradermal.

c

About 30-fold higher dose of PGE2 required compared with mechanical allodynia.

d

See EP1 antagonist section for specificity; reported pA2 of 9.96 in allodynia model is unlikely to be valid because antagonist concentrations in biophase are unknown.

Intravenous ONO-8711 also suppressed afferent nerve discharge to distension of the rat bladder sensitized with acetic acid (Ikeda et al., 2006). The inference from these studies was that PG(E2) generated locally activates EP1 receptors on peripheral sensory neurones. Again EP1/EP3 selectivity is critical, given that EP3 agonists augment bradykinin-induced sensory nerve discharge (Kumazawa et al., 1996). Topical ONO-8713 (like celecoxib) inhibited UV-B-induced skin inflammation and tumour development in the mouse (Tober et al., 2006); the higher selectivity of ONO-8713 supports EP1 receptor involvement. PGE2 production, but not COX-1/COX-2 expression, was suppressed by ONO-8713; the mechanism is not clear.

Despite continuing synthesis of potent EP1 antagonists with testing in analgesic/anti-inflammatory models, there has been little solid evidence of clinical efficacy. Sarkar et al. (2003) reported that ZD-6416, which is related to ZM-325802 (Figure 3; Shaw et al., 1999; Jenkins et al., 2001), inhibited upper oesophageal pain threshold to electrical stimulation in human volunteers. However, ZD-6416 does not appear to be particularly useful in the clinical setting (development profile obtained from Pharmaprojects; http://www.pharmaprojects.com).

While the preferred indication for EP1 antagonists has been for pain, additional therapeutic uses in cancer, osteoporosis, arthritis, and neurodegenerative and renal disorders have been suggested. COX-2 inhibitors are of potential value in reducing colorectal adenomas which, in turn, has created interest in using prostanoid antagonists as an alternative. The cardiovascular risk associated with celecoxib in clinical trials involving colorectal adenoma prevention (Solomon et al., 2005) would intensify interest in using prostanoid antagonists for this indication. ONO-8711 inhibited formation of colonic crypts (Kawamori et al., 2001) and reduced the frequency of polyp formation in APC1309 mice (Watanabe et al., 1999; Kitamura et al., 2003b), and aberrant crypt foci in oxymetazaline-treated mice (Watanabe et al., 1999). Correspondingly, aberrant crypt foci were reduced by 60% in EP1−/− receptor mice (Watanabe et al., 1999). COX-2 up-regulation has also been considered a target for drug treatment of pathologies involving neurological injury and neurodegeneration. The downstream effects of COX-2 neurotoxicity have been reported to be EP1 receptor-mediated (Kawano et al., 2006).

Blood pressure in the spontaneously hypertensive rat was reduced by SC-51322 (10 mg−1·kg−1·day−1, gavage) (Guan et al., 2007). EP1 receptors appear to regulate BP in the male but not the female mouse (Audoly et al., 1999; Stock et al., 2001). In a more detailed study, Guan et al. (2007) showed that SC-51322 and EP1-receptor gene deletion blunted pressor responses to 17-phenyl PGE2 and sulprostone, whereas the pressor response to the PGE1 analogue MB-28767 (claimed to be ‘a pure EP3 agonist’) was the same in EP1+/+ and EP1−/− mice. While we agree with the authors' contention that both EP1 and EP3 receptors contribute to the pressor effects, the utility of MB-28767 may be compromised by its moderate TP agonism (Lawrence and Jones, 1992) EP1-receptor gene deletion also reduced the elevated BP and cardiac hypertrophy following a 4-week infusion of angiotensin (Ang) II in the mouse, and 1 µM SC-51322 markedly attenuated the contractile action of Ang II on the isolated pre-glomerular arteriole (Guan et al., 2007). How these findings for EP1 receptors integrate with the modest hypertension and reduction of antihypertensive efficacy associated with NSAIDs and COX-2 inhibitors in the human setting (Johnson et al., 1994; Ishiguro et al., 2008) is not clear. Species and gender appear to be highly influential factors.

EP2 receptor antagonists

Selective EP2 receptor antagonists are essentially unavailable. This may be partly due to a reluctance to inhibit the potential anti-inflammatory actions of endogenous PGE2 (see Teixeira et al., 1997) mediated via EP2 receptors (Noguchi et al., 1999; Nataraj et al., 2001).

Woodward et al. (1995b) showed that AH-6809 has modest affinity for human rc-EP2 receptors and antagonized PGE2-induced activation of adenylyl cyclase with a pA2 of about 6.5. Lower affinities were found for inhibition of PGE2-induced relaxation of human bronchus (pA2= 5.78; Norel et al., 1999) and guinea pig trachea (pA2= 5.6, KJ Ong and RL Jones, unpubl. obs.) (Table 2). This is consistent with its utility for distinguishing EP1 receptors from other EP subtypes in non-primate pharmacological preparations (Coleman et al., 1987; Eglen and Whiting, 1988; Lawrence et al. 1992). However, given the poor selectivity of AH-6809 in the low micromolar range, findings arising from its use in high concentration as an EP2 antagonist (e.g. 100 µM in Aronoff et al., 2004) should be approached with caution. Ki values for rc-EP2 receptors of about 1 µM have been found for some compounds in combinatorial studies (see Murase et al., 2008b); it should be possible to build on these observations.

EP3 receptor antagonists

Development

EP3 antagonism in a series of biaryl-acylsulphonamides was reported in 2002 by Merck (Gallant et al., 2002). The lead compound was an AT1 receptor antagonist 16 (Figure 4) with a binding KD of 7 µM for the human EP3 receptor. In a combinatorial approach, reversal of the acylsulphonamide and insertion of an ethylene spacer yielded 17 with an EP3KD of 25 nM and minimal binding to other prostanoid receptors. A related analogue, L-798106, at 0.2 µM blocked the pre-synaptic (EP3) inhibitory actions of sulprostone on guinea pig vas deferens and trachea with pA2 values of 7.5 and 7.8 respectively (Clarke et al., 2004). L-798106 (0.2–1 µM) also caused parallel displacement of log concentration–response curves for contraction of rat femoral artery by sulprostone (in synergy with phenylephrine) (Hung et al., 2006). L-826266, a chloro analogue of L-798106, has been used in other studies despite the lack of detailed information on its basic pharmacology. Schlemper et al. (2005) showed that L-826266 at 30 µM inhibited both PGE2- and bradykinin-induced relaxation of guinea pig trachea, and proposed that bradykinin induces de novo synthesis of PGE2, which then activates EP3 receptors. However, the specificity of L-826266 at this high concentration is unknown. An alternative explanation is that L-826266 modestly blocks EP2 receptors thereby allowing the action of PGE2 on the contractile EP1 system in the trachea to dominate. Oliva et al. (2006) reported that L-826266 injected (as 0.1 µL of 1.25–5 mM solutions in 20% DMSO/PSS) into the periaqueductal grey matter of the mouse brain suppressed the late hyperalgesic response to intradermal formalin. However, the concentration of L-826266 at the site(s) of action is indeterminate and interpretation of the finding is difficult because similar high ‘doses’ of EP1 and EP4 antagonists and AH-6809 also suppressed the formalin response.

Figure 4.

Figure 4

EP3 receptor antagonists. The AT1 receptor antagonist, compound 16, is a lead molecule for the biaryl-ene-acyl-sulphonamide antagonists (pharmacophore in red). The left-hand portion of this pharmacophore corresponds to the cinnamic acid moiety in compounds 18 and 19 (see broken brackets). L-826266 is a chloro analogue of L-798106. The lower-middle brackets show modifications to the indole nucleus in the EP3 antagonist series of DeCode Genetics.

DeCode Genetics has described a series of related molecules containing an indole nucleus, from which DG-041 (Figure 4) was selected for clinical investigation (Singh et al., 2009). DG-041 had a IC50 of 8.1 nM in a EP3/Ca2+ flux FLIPR assay; corresponding values in DP1 and DP2 assays were 131 and >10 000 nM respectively (Singh et al., 2009). SAR studies involved modifications to the terminal aryl moieties, together with the indole unit. Compounds with an inverted indole nucleus retain high EP3 affinity (Zhou et al., 2009a,b;), as do indolones and hexahydro-indolones (O'Connell et al., 2009). Saturation of the α.β-double bond also produced highly potent EP3 antagonists, while further saturation of the remaining double bond in the hexahydro-indolone residue resulted in marked reduction in affinity. Much of DeCode's work was directed towards improving water solubility: predicted n-octanol/water partition coefficients (ClogP) for DG-041, L-798106 and L-826266 are 6.6, 6.9 and 7.4 respectively (ChemAxon freeware). In this context, our recent studies on L-798106 and L-826266 have shown a slowly developing block of EP3 agonist-induced contraction on guinea pig aorta, affording pA2 values of 7.96 and 7.58 respectively after 3-h contact (Jones et al., 2008). As expected, highly potent antagonists such as BMY-180291 (TP, pA2= 9.8) and doxepin (histamine H1, pA2= 9.6) also had slow onsets at low-nanomolar concentrations. However, the slow onsets of L-798106 and L-826266 may be related to their high lipophilicity rather than their (moderate) receptor affinity.

Merck-Frosst researchers have also identified EP3 antagonism in ortho-substituted cinnamic acid derivatives, which correspond to the left-hand portions of the ene-acyl-sulphonamide antagonists shown in Figure 4. Compound 18 has a binding KD for the human EP3 receptor of 20 nM (Juteau et al., 2001), while compound 19 has the highest binding affinity (3 nM) and behaved as a pure antagonist in a human EP3/adenylyl cyclase assay (pA2= 8.22; Belley et al., 2005). Compound 19 is also highly lipophilic (ClogP = 8.29).

ONO-AE3-240 is reported to be a highly selective EP3 antagonist (mouse EP3/EP1 selectivity ratio = 2500; Amano et al., 2003), but its structure has not been disclosed.

Therapeutic applications

Like most prostanoid receptors, the EP3 receptor has been implicated in pain of various aetiologies. These include allodynia produced by HIV-1 glycoprotein gp 120 (Minami et al., 2003) and PGE2 (Kassuya et al., 2007), acute herpetic pain (Takasaki et al., 2005), thermal hyperalgesia (Oka et al., 1994) and formalin-induced hyperalgesia (Oliva et al., 2006). A major role for EP3 (and IP) receptors has been claimed in endotoxin-induced enhancement of pain perception (Ueno et al., 2001). Intravenous DG-041 also suppressed the visceromotor reflex to bladder distension in the rat (Su et al., 2008a); the authors defined this antagonist as brain non-penetrant, but no pharmacokinetic evidence was presented. Intrathecal application of L-798106 and DG-041 also produced a long-lasting suppression of the visceromotor reflex, while intracerebroventricular administration produced only a transient reduction (Su et al., 2008b). However, EP3 agonists are known to exert gastrointestinal cytoprotection and attenuate gastric acid secretion in animal models (Bunce et al., 1990; Kunikata et al., 2002). Although it is widely assumed that there is a similar involvement of EP3 receptors in man, an extensive search of the literature provides only circumstantial evidence for this. For example, while misoprostol is clinically useful in suppressing gastroduodenal erosion (see Hawkey, 2000), its selectivity is not high enough to infer the involvement of EP3 receptors (EP3∼ EP2∼ EP4 >> EP1 for misoprostol-free acid; Abramovitz et al., 2000). A reduction in myocardial ischaemic damage was also achieved with EP3 agonist treatment in the rat (Zacharowski et al., 1999). Thus, it is feasible that the therapeutic use of EP3 antagonists will result in a side effect profile at least comparable to that of COX inhibitors. This possibility does not seem to have hindered development work on other uses of EP3 antagonists as discussed below.

It has been suggested that endogenous PGE2 activates EP3 receptors on stromal cells surrounding a tumour causing the release of vascular endothelial growth factor, which then promotes angiogenesis and tumour growth (Amano et al., 2003). Injection of the EP3 antagonist ONO-AE3-240 around sarcoma-180 tumours in the mouse markedly suppressed these effects in a manner similar to EP3 receptor gene-deletion; ONO-8711 (EP1 antagonist) and ONO-AE3-208 (EP4 antagonist) were ineffective. In contrast, EP3 agonists acting on all three mouse EP3 receptor isoforms expressed in HEK-293 cells caused cell clustering and inhibited their proliferation via a G12-RhoA pathway (Macias-Perez et al., 2008).

A further possible therapeutic application of an EP3 antagonist is the treatment of pre-term labour. Based on the known activity of misoprostol (Sanchez-Ramoz et al., 1997) and sulprostone (Fruzzetti et al., 1988), it appears that EP3 receptor stimulation produces cervical ripening, a critical event that precedes parturition. An EP3 antagonist could be effectively combined with a tocolytic, such as an EP2 agonist (Senior et al., 1993), to provide therapy for pre-term labour. The EP3 receptor has been uniquely associated with febrile responses (Ushikubi et al., 1998). However, development of an EP3 antagonist for treating fever seems unlikely, given that low-cost COX inhibitors are highly effective in reducing body temperature.

Finally, DG-041 has shown promise in the treatment of peripheral cardiovascular disease. Activation of EP3 receptors on human platelets enhances aggregation induced by a variety of agents (in the presence of a TP antagonist); inhibition of adenylyl cyclase/priming of protein kinase C is thought to be the mechanism (Matthews and Jones, 1993; Vezza et al., 1993). In addition, there is increased bleeding tendency and increased susceptibility to thromboembolism in the EP3 receptor knock-out mouse (Ma et al., 2001). DG-041 at 0.03–3 µM inhibited the enhancement by sulprostone of ADP- or collagen-induced aggregation in human platelet-rich plasma (PRP) (Heptinstall et al., 2008; Singh et al., 2009). The pA2 value of 8.3 (our calculation from Heptinstall et al., 2008 data) is probably an underestimate of the affinity owing to plasma protein binding of DG-041. In the rat, DG-041 at 5 or 60 mg·kg−1 (by gavage; co-administration with clopidogrel) inhibited enhancement of platelet aggregation induced by PGE2ex vivo; there was no increase in bleeding time compared with clopidogrel alone (Singh et al., 2009). A similar profile was obtained with DG-041 in Phase I trials in healthy subjects (reported in Heptinstall et al., 2008).

EP4-receptor antagonists

Development

The first EP4 antagonist to be reported was AH-23848 (Coleman et al., 1994a), a close relative of the selective TP antagonist GR-32191 (Figure 8). Ligand binding studies on human rc-receptors indicated low EP selectivity for AH-23848, with Ki values for EP1, EP2, EP3, EP4 and IP receptors being 45, 50, 4.4, 14 and >100 µM respectively (Abramovitz et al., 2000). However, its selectivity in functional assays appears to be higher and it has been of considerable utility in differentiating the EP2, EP4 and IP agonist activities of prostanoid ligands. For example, Jones and Chan (2001) used AH-23848 at 30 µM to demonstrate that the PGI2 analogues cicaprost and AFP-07 relax certain vascular preparations by activating both EP4 and IP receptors. Moreover, Lai et al. (2008a) showed that pulmonary artery smooth muscle cells from monocrotaline-treated rats have a reduced IP receptor density and that iloprost-induced cAMP elevation is blocked by AH-23848, and therefore likely to be due to activation of EP4 receptors.

AH-23848 has since been overtaken by more potent agents. Antagonists in the major group contain an acyl-sulphonamide unit (Figure 5) and show some similarity to the ene-acyl-sulphonamide EP3 antagonists. Indeed, L-161982 is a methyl analogue of the lead molecule 16 for the Merck EP3 antagonists. L-161982 was reported to have a pA2 of about 8.5 in rc-EP4 receptor – adenylyl cyclase assays (Machwate et al., 2001). In a rat native EP4 assay, L-161982 inhibited PGE2-induced cAMP accumulation with an IC50 of about 30 nM, while forskolin-induced cAMP accumulation was unaffected by 10 µM L-161982. In similar functional assays, CJ-023423 afforded pA2 values of 8.3 and 8.2 for human and rat rc-EP4 receptors using the Schild protocol (Nakao et al., 2007). In binding assays, CJ-023423 showed weak affinity for human EP2 receptors and PAF receptors, while interactions with other prostanoid receptors and a range of non-prostanoid receptors were minimal.

Figure 5.

Figure 5

EP4 receptor antagonists. L-161982 is a methyl analogue of compound 16 in Figure 4. Acyl-sulphonamido residues are shown in red. The bonds indicated by asterisks in MF-498 are subject to oxidative/hydrolytic attack in vivo; the corresponding substituents in compound 20 prevent these transformations.

GW-627368 has pKi values for binding of 7.0 and 6.8 for human rc-EP4 and TP receptors respectively; binding to other prostanoid receptors is minimal (Wilson et al., 2006). Correspondingly, a pA2 of 7.9 was obtained for GW-627368 in a human rc-EP4 receptor – adenylyl cyclase assay, with good evidence for competition. On human pulmonary vein, GW-627368 had a pA2 of 7.06 against ONO-AE1-329 (Table 2; Foudi et al., 2008); its higher affinity against PGE2 may have been due to opposing contractile activity. GW-627368 had a higher affinity on piglet saphenous vein (pA2= 9.2) with a linear Schild plot up to a concentration ratio of about 60; further rightward shift of the agonist curve was insignificant owing to PGE2 activating a less sensitive EP2 relaxant system (Wilson et al., 2006). GW-627368 at 10 µM did not affect TP receptor-induced contraction under the same conditions. Results for rabbit saphenous vein were discrepant: Wilson et al. (2006) showed that 10 µM GW-627368 did not affect PGE2-induced relaxation implying the presence of an EP2 system only, while Jones and Chan (2005) found a right-shift of about 1 log unit with 1 µM GW-627368, consistent with the presence of EP2 and EP4 systems; GW-627368 did not antagonize relaxation induced by either the selective EP2 agonist ONO-AEI-259 or the PGI2 analogue, taprostene in the latter experiments. MF-498 (Clark et al., 2008), which is quite similar in structure to GW-627368, undergoes oxidative/hydrolytic metabolism at the three regions indicated in Figure 5; compound 20 was considerably more resistant to attack while retaining high EP4 antagonist affinity (Burch et al., 2008). The most recent addition to this group is BGC-20-1531 (Maubach et al., 2009). It exhibits surmountable antagonism of PGE2-induced relaxation of human cerebral and middle meningeal and dog carotid and middle meningeal arteries in vitro, while having no effect on PGE2 (probably EP3)-induced contraction of human coronary, pulmonary and renal arteries.

A second looser group of amide-containing EP4 antagonists is represented by ONO-AE2-227, ONO-AE3-208 and CJ-042794 (Figure 5). Binding studies indicate that ONO-AE2-227 and ONO-AE3-208 retain considerable affinity for EP3 receptors (Ki = 21 and 30 nM; Mutoh et al. 2002; Kabashima et al. 2003) and this must be borne in mind when interpreting in vivo data (see later). CJ-042794 behaved competitively in functional assays (Schild protocol) involving human and rat rc-EP4 receptors (Murase et al., 2008a,b;). Its EP4/EP3 selectivity ratio is very high, but it still retains measurable affinity for human EP2 receptors (pKi= 6.2).

Therapeutic applications

Recent studies involving parenteral administration of several EP4 antagonists (AH-23848, CJ-023423, CJ-042794, MF-498, ONO-AE3-208) have clearly demonstrated a major involvement of EP4 receptors in small-animal models of inflammation (Table 3). Joint pain, mechanical and thermal hyperalgesia and oedema were markedly suppressed, often equivalent to the efficacy of selective COX-2 inhibitors such as rofecoxib. A peripheral site of action seems likely given that intraplantar administration of L-161982 suppressed carrageenan-induced mechanical allodynia in the mouse (Kassuya et al., 2007). Whether block of the EP4 receptor alone produces an effective anti-inflammatory drug in man remains to be seen. Moreover, PGE2 may have a protective role in inflammation. Takayama et al. (2002) showed that PGE2 suppressed chemokine production stimulated by lipopolysaccharide in human macrophages; L-161982 at 100 nM blocked this action.

Prostanoid EP4 receptors, indeed nearly all EP receptor subtypes, have been implicated as contributors to colon tumorigenesis caused by excessive production of PGE2 (Fujino and Regan, 2003; Majima et al., 2003; Masataka et al., 2003; Mutoh et al., 2006). The selective EP4 agonist ONO-AE1-329 (0.1–1 µM) increased colony formation in the human colon cancer cell line HCA-7 (Mutoh et al., 2002) and L-161982 blocked PGE2-induced proliferation of HCA-7 cells (Cherukuri et al., 2007). In both the oxymetazoline model of aberrant crypt foci (putative preneoplastic lesions) and the Min mouse model of intestinal polyp development, the EP4 antagonist ONO-AE2-227 produced about a 67% reduction in the appropriate scores (Mutoh et al., 2002). Aberrant crypt foci were similarly reduced in EP4−/− mice. In the APC1309 mouse, ONO-AE2-227 had a preferential effect on polyp size, while the EP1 antagonist ONO-8711 had a more pronounced effect on polyp number, and a combination of antagonists behaved additively (Kitamura et al., 2003b). In the context of (tumour) vascularity, ONO-AE3-208 reduced IL-1β-induced angiogenesis in the mouse cornea at an oral dose of 1 mg·kg−1 o.d. (Kuwano et al., 2004). Finally, pretreatment of mouse mammary tumour cells with AH-23848 and ONO-AE3-208 followed by washing and immediate injection into immunologically compatible mice reduced pulmonary tumour score (Fulton et al., 2006). These comprehensive studies provide encouragement that EP4 receptor antagonists may provide a safer replacement for COX-2 inhibitors in treating colon cancer. As, however, all EP receptor subtypes have been implicated in colon cancer, a pan-PGE synthase inhibitor may be more effective in preventing the key cell proliferative and angiogenic events. Arguably, and somewhat paradoxically, the EP4 receptor has also been shown to participate in the maintenance of intestinal homeostasis by preserving mucosal integrity. In both EP4−/− mice and wild-type mice treated with an EP4 antagonist (ONO-AE3-208), susceptibility to the development of colitis was reported in a model of inflammatory bowel disease (Kabashima et al., 2003; Narumiya, 2003).

There has been interest in using EP4 agonists for their anabolic effects on bone (Raisz, 2006). The EP4 antagonist L-161982 at a dose of 10 mg·kg−1·day−1 reversed bone formation induced by PGE2 in the rat, without affecting its diarrhoeal action (Machwate et al., 2001). Also, human mesenchymal stem cells in culture secreted PGE2 via COX-2, which was associated with production of bone morphogenetic protein-2 (BMP-2), a factor that stimulates differentiation of precursor mesenchymal cells into mature bone. The selective COX-2 inhibitor, NS-398, and the EP4 antagonist, ONO-AE3-208, (concentration not specified) suppressed BMP-2 expression (Arikawa et al., 2004). These findings identify bone loss as a potential side effect of EP4 antagonists.

In vitro, EP4 vasodilator systems typically exhibit high sensitivity to PGE2 and this property is reflected in the role of PGE2 in maintaining the open state of the ductus arteriosus during gestation in human and animal species. Strong expression of EP4 (and IP) receptors is found in ductus tissue from the newborn infant and child (Leonhardt et al., 2003). Towards the end of gestation, smooth muscle cells within the ductus migrate to the endothelial lining where they form intimal cushions. Declining PG levels at birth result in ductus constriction, thereby bringing the intimal cushions into close contact and effecting permanent closure. Intimal cushion formation is also driven by EP4 receptor activation (Yokoyama et al., 2006). The COX inhibitor, indomethacin, is commonly used to treat failure of ductus closure, but its efficacy is poor in a substantial proportion of cases, perhaps related to suppression of intimal cushion development (see Ivey and Srivastava, 2006). AH-23848 blocked PGE2-induced relaxation of the rabbit and sheep isolated ductus arteriosus preparations (Smith et al., 1994; Bouayad et al., 2001), while ONO-AE3-208 constricted the ductus of fetal and neonatal rats in vivo (Momma et al., 2005). Whether an EP4 antagonist would be better than a COX inhibitor in treating patent ductus in premature infants is not yet clear.

A second potential role for an EP4 vasodilator system is the genesis of vascular headache in migraine. EP4 antagonists block PGE2-induced relaxation of human-isolated middle cerebral artery (Davis et al., 2004; Maubach et al., 2009) and the picture has been enlarged to include the interaction of endogenous PGE2 with calcitonin gene-related peptide release from trigeminal nerves (Maubach et al., 2009). The use of AH-23848 and L-161982 also has provided evidence for a role of EP4 receptors in substance P release by stretching of the renal pelvic wall (volume expansion) leading to activation of renal sensory afferents and subsequent diuresis/natriuresis (reno-renal reflex) (Kopp et al., 2004).

FP receptor and prostamide receptor antagonists

Development

A variety of FP receptor antagonists have been reported in the past, but none has stood up to rigorous analysis. Thus, PGF-1-dimethylamine and PGF-1-dimethylamide (Maddox et al., 1978; Stinger et al., 1982) exhibited no meaningful FP antagonist or agonist activities (Sharif et al., 2000; Woodward et al., 2008). Phloretin (Kitanaka et al., 1993) was non-selective and very weak in blocking fluprostenol-mediated phosphoinoside turnover in rat A7r5 vascular smooth muscle cells (Sharif et al., 2000). Similarly glibenclamide (Delaey and Van de Voorde, 1995) was a weak, non-selective prostanoid antagonist (Sharif et al., 2000).

AL-3138 and AL-8810, both PGF analogues (Figure 6), have received a degree of acceptance as FP antagonists (Griffin et al., 1999; Sharif et al., 2000). AL-8810 proved useful in studying PGF-mediated up-regulation of the orphan nuclear receptor Nur 77 (Liang et al., 2004). However, further studies reveal that they are neither potent nor selective. For example, AL-8810 appears to block TP receptors (Hutchinson et al., 2003), an effect that has been confirmed in human rc-TP receptor stable transfectants (A.H. Krauss and D.F. Woodward, unpublished). In addition, both agents are FP partial agonists. This can be seen in the original report of AL-8810 on phosphoinositide turnover (Griffin et al., 1999) and is corroborated by the finding of myogenic activity in the mouse uterus (Hutchinson et al., 2003). AL-8810 induced a more pronounced Ca2+ signal in human rc-FP stable transfectants (Y. Liang and D.F. Woodward, unpublished) and was a full agonist in the cat isolated iris preparation (Woodward et al., 2007); the latter action was not blocked by a prostamide antagonist (Woodward et al., 2007; see later), so it presumably reflects FP receptor agonism.

Figure 6.

Figure 6

FP receptor and prostamide receptor antagonists. The natural ligand PGF and its 1-ethanolamide derivative, prostamide F, are shown in the box. AL-3138 and AL-8810 are FP partial agonists in many systems; α= corresponding side-chain in PGF. The THG analogues are peptides: amide (CO-NH) residues are shown as red bars. The AGN analogues (upper right) are prostamide receptor antagonists; C1-amide residues are shown in blue.

A series of octapeptides (THG-131 derivatives, Figure 6) have been claimed to possess selective FP antagonist activity (Chemtob and Peri, 2006; Peri et al., 2006). In particular, THG-113.31 at 1 µM markedly inhibited PGF-induced contraction of pig retinal blood vessels, while having minimal effect on contraction to 17-phenyl PGE2, U-46619, phenylephrine, Ang II and endothelin-1; the inhibition of PGF contraction was insurmountable. In addition, THG-131.31 inhibited the associated phosphoinositide hydrolysis with an IC50 of about 30 nM. However, THG-131.31 competed poorly with [3H]-PGF for binding to human rc-FP receptors (∼13% at 10 µM). Several other studies have cast doubt on the utility of THG-131.31. At 10 µM, THG-113.31 showed fairly weak antagonism of the contractility of longitudinal and circular strips of sheep myometrium induced by PGF; there was no effect on PGE2 contractions (Hirst et al., 2005). Also, 10 µM THG-113.31 had no effect on PGF-induced contraction of human pregnant myometrium, while inhibiting spontaneous and oxytocin-induced contractions at much lower concentrations (Friel et al., 2005). Finally, Doheny et al. (2007) showed that THG-113.31 at 10–50 µM enhanced BKCa channel opening in isolated myocytes from human uterus, an effect that was reversed by iberiotoxin. Several chemically simpler peptidomimetics in the THG series showed more potent block of PGF-induced contraction than THG-113.31 (Figure 6, Table 2; Peri et al., 2006); it would be of interest to have full pharmacological profiles.

Currently, the most convincing FP antagonist is the non-prostanoid AS-604872 (Cirillo et al., 2007). It has Ki values of 35, 158 and 323 nM for human, rat and mouse rc-FP receptors and its selectivity was 20-fold for EP2 receptors and greater than 300-fold for other prostanoid receptors. AS-604872 showed no agonist activity in a human FP receptor – inositol phosphate assay and had an IC50 of 47 nM against PGF. In vivo, AS-604872 (1–30 mg·kg−1, i.v.) inhibited PGF-induced uterine contraction in the non-pregnant rat; inhibition of oxytocin-induced contraction was slight.

Prostamide research originated from studies on neutral PGF analogues and notably bimatoprost (17-phenyl PGF-1-ethylamide), an effective anti-glaucoma drug (Woodward et al., 2003; 2004;). As previously mentioned, neutral PGF analogues show only weak affinity for FP receptors (Maddox et al., 1978; Schaaf and Hess, 1979), a profile subsequently confirmed for −OH, −OCH3 and −CON(R)2 C1-substitutes (Woodward et al., 2000; 2008; Matias et al., 2004). The pharmacology of bimatoprost was similar, with no meaningful activity at FP receptors but pronounced activity in certain preparations such as cat lung strip, cat iris, rabbit uterus, and human ciliary smooth muscle cells (Liang et al., 2003; Woodward et al., 2003b; Matias et al. 2004; Chen et al. 2005). The pharmacology of bimatoprost appeared indistinguishable from that of PGF-1-amides, but this could not be rationalized until the discovery that anandamide (arachidonic acid-1-ethanolamide) was a substrate for COX-2 (Yu et al., 1997).

PGE2-1-ethanolamide (prostamide E2) was the first prostamide to be discovered and was identified as the major product following addition of anandamide to rc-COX-2 or cells expressing COX-2, but not COX-1 (Yu et al., 1997). Subsequently, more extensive studies demonstrated that COX-2 oxidizes anandamide to endoperoxide intermediates, which are converted by specific PG synthases to the various prostamides (Kozak et al., 2002; Koda et al., 2004; Yang et al., 2005; Moriuchi et al., 2008). Despite being different terminal biosynthetic products, the evidence to date suggests that prostamides D2, E2 and F interact with a single receptor to exert their effects. Prostamide F and its analogues are, however, about 10 times more potent than prostamides D2 and E2 (Woodward et al., 2007).

Initial pharmacological characterization of the prostamides, for example prostamide E2 (Ross et al., 2002), relied on agonist studies. In the context of PGF analogues, FP agonists (17-phenyl PGF and PGF) and bimatoprost produced Ca2+ signals in entirely different cells in a cat iris smooth muscle cell preparation (Spada et al., 2005), suggesting the existence of a receptor with a distinct preference for prostamide F agonists. The receptor structure appears to involve heterodimerization of the wild-type and alternative mRNA splicing variants of the FP receptor, both encoded by PTGFR, the FP receptor gene (Liang et al., 2008). This FP/alt-FP co-expression is analogous to the isoprostane binding site which is formed as a result of IP/TP receptor heterodimerization (Wilson et al., 2004).

The strategy for discovering a prostamide antagonist was to identify antagonists in cluster 2 of the prostanoid receptor evolutionary tree (Narumiya et al., 1999) and then form corresponding C1-amides. Cluster 2 contains the TP receptor, whose stable agonists have been a starting point for antagonists at other receptors (DP1, see Figure 1; EP1, see Figure 3). Based on the oxabicycloheptane analogue BMS-180291 (Figure 8) (Webb et al., 1993), two prototype prostamide antagonists AGN-204396 (Figure 6) and AGN-204397 were identified (Woodward et al., 2007; 2008;). These agents showed good prostamide F/FP selectivity, but were of low affinity (pA2∼5.5) and also blocked TP receptors. Substituting oxygen at C3 dramatically enhanced prostamide affinity, by as much as 100-fold for AGN-211334 and AGN-211335 (Wan et al., 2007; Liang et al., 2008; Woodward et al., 2008). AGN-211334 and AGN-211335 potently inhibited prostamide F and bimatoprost responses in cat iris preparations, but did not alter responses to FP agonists (Wan et al., 2007; Liang et al., 2008). AGN-211334 blocked the increase in conventional aqueous humour outflow produced by bimatoprost in the human perfused anterior segment preparation (Wan et al., 2007), thereby demonstrating that the effects of bimatoprost in the human eye are prostamide receptor-mediated; previously, it has been suggested that bimatoprost's activity is dependent on deamidation to the FP-active free acid in ocular tissue (Camras et al., 2004). AGN-211335 blocked the secondary Ca2+ wave, myosin light chain phosphorylation, and Cyr 61 up-regulation induced by bimatoprost in the FP/alt-FP prostamide system (Liang et al., 2008). These second-generation prostamide antagonists are likely to be sufficiently potent for in vivo studies.

Therapeutic applications

Prevention of pre-term labour (tocolysis) is probably the only therapeutic modality where a FP antagonist may be of value. This is a serious and unmet medical need given that premature birth accounts for 60–80% of perinatal deaths (Goldenberg, 2002). Parturition is prevented in FP−/− mice (Sugimoto et al., 1997) confirming a significant role for PGF (see Challis et al., 2002 for a review of PG involvement). THG-113.31 is tocolytic in the pregnant sheep (Hirst et al., 2005) and may be useful for delaying pre-term birth (Olson, 2005). AS-604872 suppressed spontaneous uterine contractions in late-term pregnant rats and delayed preterm birth caused by mefipristone in pregnant mice; AS-604872 appeared to be more effective than the β2-adrenoceptor agonist ritodrine (Chollet et al., 2007). It is not yet clear whether these findings will translate into an effective drug in human reproduction.

Prostanoid FP receptors have been implicated in cardiomyocyte hypertrophy (Pönicke et al., 2000; Xu et al., 2008) and cancer (Fujino and Regan, 2001) and may play a role in regulating water and solute transport based on in situ hybridization studies in the mouse kidney (Saito et al., 2003). The utility of a selective FP antagonist is, however, unclear.

The therapeutic utility of prostamide antagonists is similarly uncertain. Increased anandamide levels may result in the formation of prostamides as major products in inflammation and infection (Glass et al., 2005). Testing of AGN 211334, or a close congener, in relevant animal models should provide evidence for or against a functional role of prostamides.

IP receptor antagonists

Development

IP receptor antagonists are a recent development based on the potential role of PGI2 in mediating pain (Bley et al., 1998). In an extensive series of studies, two structurally distinct classes of selective IP antagonist emerged from focused chemical library screening and synthetic chemistry (Bley et al., 2006) (Figure 7). The 2-(phenylamino)-imidazoline series is represented by RO-1138452 and compound 21 (Clark et al., 2004; Keitz et al., 2004), while the N-substituted phenylalanine series has a ‘traditional’ carboxylate in the phenylalanine residue and is typified by RO-3244019 (Fitch et al., 2004) and its difluoro analogue RO-3244794 (Bley et al., 2006).

Figure 7.

Figure 7

IP receptor antagonists. The natural ligand PGI2 (prostacyclin) is shown in the box. 2-(Phenylamino)-imidazoline moieties are shown in blue and phenylalanine residues in red (S-configuration in compound 24). RO-3244794 is a difluoro analogue of RO-3244019.

RO-1138452 has high affinity for human native (platelet) and rc-IP receptors, with pKi values of 9.3 and 8.7 respectively in studies utilizing [3H]-iloprost, although it also displays considerable affinity for PAF (7.9) and imidazoline (8.3) receptors (Bley et al., 2006). In functional studies in platelets (Jones et al., 2006), pA2 values were lower than Ki values obtained in radioligand competition studies with platelet membrane preparations and for inhibition of carbacyclin-induced cAMP formation in cells over-expressing rc-IP receptors (Bley et al., 2006); the difference was attributed to protein binding in studies involving PRP. The pA2 values obtained in isolated blood vessel preparations (human pulmonary artery 8.20, guinea pig aorta 8.39 and rabbit mesenteric artery 8.12) were intermediate between values obtained in PRP and plasma membranes/cells in buffer (Jones et al., 2006). The slight suppression of the cicaprost maximum response seen with higher concentrations of RO-1138452 was attributed to functional antagonism emanating from the (albeit weak) EP3 agonist action of cicaprost. However, RO-1138452 displayed an insurmountable antagonist profile in studies of chemokine release from human airway epithelial cells using taprostene (see Table 1) as IP agonist (Ayer et al., 2008). Moreover, RO-1138452 inhibition of taprostene-induced cAMP response element-dependent transcription was not reversed over a 20-h ‘washout’ period. These data could not be ascribed to covalent receptor inactivation, allosterism or a state of antagonist hemi-equilibrium and may be due to a pseudo-irreversible interaction with the IP receptor (Ayer et al., 2008).

The N-substituted phenylalanines 22 and 23 (Nakae et al., 2005) are weak IP antagonists, while the greater potency of 24 is associated with as S-configuration in the phenylalanine residue (Brescia et al., 2007). RO-3244794 has no meaningful activity at EP1, EP3, EP4 and TP receptors; no data were provided for DP1, DP2, EP2 or FP receptors (Bley et al., 2006). While RO-1138452 did not block EP2 receptor-mediated relaxation in guinea pig aorta (pA2 < 6.0; Jones et al., 2006), RO-3244794 had a pA2 of 6.92 for the human rc-EP2 receptor expressed in HEK-293 cells (S.M. Hill and M.A. Giembycz, unpubl. data). Compound 24 did not bind to human EP2 and EP4 receptors (Brescia et al., 2007). Strictly speaking, the missing data related to these compounds (e.g. RO-1138452 at DP2 and FP receptors; Jones et al., 2006) need to be addressed. In the context of systemic drug design, RO-3244794 has much greater oral bioavailability than RO-1138452 (51% vs. 0.7% in the rat) (Bley et al., 2006).

Therapeutic applications

Pain has been the initial focus for IP antagonists based on IP receptor agonist effects and IP receptor distribution (reviewed by Bley et al., 1998). Altered pain perception and inflammation were observed in IP−/− mice (Murata et al., 1997). Subsequently, IP antagonists were shown to reduce pain responses in models where prostanoids have been implicated. These include acetic acid-induced abdominal constriction, mechanical hyperalgesia produced by carrageenan and pain associated with models of osteoarthritis and inflammatory arthritis (Bley et al., 2006; Pulichino et al., 2006). Such analgesic effects are accompanied by anti-inflammatory properties, which is invariably the case. In a collagen-induced arthritis model in mice, 21 produced effective inhibition when administered as a pretreatment, but was inactive when given after the initiation of the arthritis (Pulichino et al., 2006). RO-3244794 and indomethacin were equi-effective in reducing carrageenan-induced rat paw oedema and more effective than rofecoxib in inhibiting the foot weight distribution change associated with intra-articular injection of monoiodoacetate (Bley et al., 2006). Ostensibly, these results imply that IP receptors provide a singular target that would result in drugs that are at least as effective as NSAIDs and COX-2 inhibitors. The role of EP receptors in pain/inflammation has already been discussed. Despite a satisfactory preclinical profile, strong circumstantial evidence suggests that EP1 antagonists were a failure in clinical trials. This, in turn, tends to cast doubt over the clinical prognosis for IP antagonists. Given the number of prostanoids that may be released locally and the even greater number of target receptors available, it may be regarded as counter-intuitive to propose that inflammation can be attributed to only one type of prostanoid receptor.

A role for PGI2 in mediating the sensitized release of substance P from rat dorsal root ganglion neurons has been suggested based on studies with 23 (Nakae et al., 2005). Related to this is the potential use of IP antagonists for bladder disorders. In rat models of bladder function, RO-3244019 dose-dependently decreased bladder contraction frequency and increased micturition threshold and voiding interval (Cefalu et al., 2007). RO-3244019 was also effective in treating neurogenic detrusor overactivity arising from spinal cord injury in the rat (Khera et al., 2007). Again, these results are of uncertain predictive value for clinical success, because indomethacin exhibited good activity in these models. The IP antagonist, BAY-73-1449 (Figure 7), was effective in acutely reducing shunt vessel blood flow in a rat model of portal hypertension (Bexis et al., 2008); full details of its pharmacology are unavailable.

Finally, side effects originating from perturbation of the TXA2/PGI2 balance could be greater with IP antagonists than with COX-2 inhibitors, including hypertension, stroke, myocardial infarction and atherosclerosis. Nephrotoxicity and K+ and Na+ retention also may occur (Nasrallah and Hébert, 2005), because prostacyclin synthase (PGIS) and COX-2 deficiencies produce similar renal toxicity. This suggests a protective role for PGI2, but it is important to note that no marked renal phenotype occurs in TP−/− or IP−/− mice (Breyer and Breyer, 2000; Yokoyama et al., 2002; Nasrallah and Hébert, 2005).

TP receptor antagonists

Development

Elucidation of the structures of PGH2 and TXA2 (Figure 8) was soon followed by a range of carba/thia bicyclic analogues (see Wilson and Jones, 1985). One aim was to produce a chemically stable TP agonist (e.g. U-46619, STA2); another was to investigate the potential for TP receptor antagonism. Thus, pinane-TXA2 (PTA2) was reported to block both constriction of cat coronary artery and aggregation of human platelets elicited by U-46619 (Nicolaou et al., 1979). However, this analogue often behaves as a partial agonist, inducing a 60% maximal activation in some TP systems (Jones et al., 1982; Tymkewycz et al., 1991). More useful TP antagonists emerged from additional modification of the ω-chain. The pinane analogue ONO-11120 (Figure 8) is a true antagonist (Katsura et al., 1983), while EP-045 (Jones et al., 1982) and EP-092 (Armstrong et al., 1985) were developed from the PG endoperoxide analogue, 9,11-etheno PGH2, also a partial agonist (Jones et al., 1982). SQ-29548 (Ogletree et al., 1985) and BMS-180291 (Ogletree et al., 1993) have a hybrid bicyclic system (oxabicyclo[2.2.1]heptane) and differ from the natural agonists in the cis-orientation of the α and ω-chains. While GR-32191 (Lumley et al., 1989) and ICI-192605 (Brewster et al., 1988) still retain a prostanoid skeleton, other TP antagonists do not, including L-655240 (related to indomethacin; Hall et al., 1987), BM-13505 (daltroban; Yanagisawa et al., 1987), KW-3635 (Karasawa et al., 1991a,b;), AA-2414 (seratrodast; Ashida et al., 1989) and even the KATP-channel blocker glibenclamide (Cocks et al., 1990). The attractively simple chemistry involved in the synthesis of daltroban has spurred the development of higher-potency, non-prostanoid antagonists containing a (p-halo)-benzenesulphonylaminomethyl residue at a critical distance from the carboxylate, for example, Z-335 (Tanaka et al., 1998) and S-18886 (terutroban; Cimetière et al., 1998). A similar residue in prostanoid molecules such as S-145 (domitroban; Mihara et al., 1989), I-SAP (Naka et al., 1992) (Figure 8), ONO-NT-126 (p-bromo; Nakahata et al., 1990) and oxa- and thia-bicyclo[3.1.0]hexane derivatives (Kamata et al., 1990) also confers high TP affinity.

Differences in antagonist affinity constants between platelet and vascular smooth systems have stimulated much debate about the existence of TP receptor subtypes (Mais et al., 1985; 1988; Swayne et al., 1988; Morinelli et al., 1989; Masuda et al., 1991; Tymkewycz et al., 1991; Folger et al., 1992). Species heterology, the difficulty of distinguishing partial agonism from functional antagonism, and incomplete equilibration of high-affinity antagonists have been confounding factors. For example, antagonism of U-46619 by BMS-180291 on human platelets was surmountable for the shape-change response, but insurmountable for aggregation; the rate of aggregation was also slowed (Ogletree et al., 1993). Similar profiles had been reported previously for EP-092 (Armstrong et al., 1985) and GR-32191 (Lumley et al., 1989) on human platelets, and for EP-169 and AH-23848 on human and rat platelets, but not on rabbit platelets where their affinities are lower (Tymkewycz et al., 1991). However, against the slow-acting high-affinity TP agonist EP-171 (Jones et al., 1989), GR-32191 did not alter the aggregation rate (Lumley et al., 1989). It is likely that slow dissociation of a high-affinity antagonist from the TP receptor retards U-46619 occupancy in the early stage of the aggregation response thereby favouring the disaggregation process and insurmountability; in contrast, shape change, which does not fade, allows a true measure of the equilibrium state. In pig platelets, longer pre-incubation times were required for S-145 than SQ-29548 owing to the smaller association rate constant of the former antagonist (Mihara et al., 1989). BMS-180291 also had a Schild slope greater than unity on guinea pig aorta (Zhang et al., 1996). However, in the low nanomolar range BMS-180291 requires up to 2 h to reach steady state on the aorta (Jones et al., 2008) and the non-ideality may simply reflect retarded diffusion of a high-affinity ligand (pA2= 9.8, Table 2) through the extracellular space.

Radioligand binding studies of TP receptors coincided with these functional measurements. Two saturable binding sites were identified on human platelets using [3H]-9,11-epoxymethano PGH2, the first radioligand developed for the TP receptor (Armstrong et al., 1983; Pollock et al., 1984). Binding to the more abundant site was displaced by EP-045 at concentrations similar to those required to inhibit [3H]-phosphatidate formation, Ca2+ elevation and aggregation induced by the TP agonist. A less abundant higher-affinity binding site was also identified, but it did not show the characteristic preference for a 15(S) configuration in the natural prostanoid ω-chain. The presence of high- and low-affinity binding sites for agonist (but not antagonist) ligands was also evident using [3H]-trimetoquinol (Ahn et al., 1988) and [3H]-SQ-29548 (Hedberg et al., 1988). Further studies showed that the high-affinity site was associated with the platelet shape change (and increase in cytosolic Ca2+), while the lower-affinity site was associated with aggregation [and activation of phospholipase C (PLC)] (Dorn, 1989; Takahara et al., 1990). [3H]-GR-32191 played an important role in the elucidation of these relationships by binding reversibly to the ‘shape change site’ and irreversibly to the ‘aggregation site’ (Takahara et al., 1990). It is difficult to explain this irreversibility given that GR-32191 does not obviously contain a chemically reactive group (Figure 8). Detailed studies have shown that 30-min exposure of human platelets to GR-32191 resulted in about 50% loss of binding sites for either [3H]-GR-32191 or [3H]-SQ-29548, while neither SQ-29548 nor BM-13177 affected Bmax. It was speculated that GR-32191 binds to internalized TP receptors (Armstrong et al., 1993); the zwitterionic nature of GR-32191 at neutral pH may be relevant. A light-activated, covalent-bonding TP antagonist, azido-BSP, also discriminated these platelet sites by blocking aggregation but not shape change induced by U-46619 (Zehender et al., 1988). The subsequent identification of a second TP receptor isoform (TPβ) from a human umbilical vein endothelial cDNA library (Raychowdhury et al., 1994) and the detection of mRNA for the α and β isoforms in human platelets (Hirata et al., 1996) would appear to complete the argument. However, these isoforms, which arise by alternative gene splicing and differ only in their cytoplasmic tails, do not show the ligand discrimination typical of the high- and low-affinity binding sites. Finally, significant expression of the TPα isoform only was found in human platelets (Habib et al., 1999).

The seventh transmembrane domain (TM-7), which is strictly conserved in all the TP receptors characterized to date, is critical to TP agonist and antagonist function. Point mutations in this domain in the human TP receptor severely suppressed the binding of SQ-29548 (Funk et al., 1993b). Chimeric substitutions of the human TP receptor with the corresponding TM 1, 2 and 4 from rat resulted in modest suppression of SQ-29548 binding, lesser suppression of I-BOP (TP agonist) binding, and a poor correlation between the data sets (Dorn et al., 1997).

Several different inhibitory properties have been combined with specific TP antagonism, either by chance or deliberately. For example, TP antagonists based on PGH2 with diphenylmethyl-oxime (e.g. EP-157, Figure 8) or diphenylmethyl-azine residues in the ω-chain were found to activate IP receptors in platelet and vascular systems (Armstrong et al., 1986; 1989; Jones et al., 1993). A diaryl-hetero(cyclic) moiety is critical to the IP agonism (Jones et al., 1993). Related compounds lacking a prostanoid ring system (e.g. octimibate) showed similar profiles (Merritt et al., 1991a,b;), with BMY-45778 being the most potent of a large series of non-prostanoid prostacyclin mimetics synthesized by Bristol-Myers Squibb (Meanwell et al., 1994; Seiler et al., 1997). Accurate estimation of the TP antagonist affinities of many of these compounds is difficult owing to their high lipophilicity, slow onset/offset (Jones et al., 1997) and, in certain instances, an ability to inhibit (non-prostanoid) Gq-PLC-driven responses (Chow et al., 2001).

Combining TP receptor antagonism with thromboxane synthase (TXS) inhibition has been an extensively investigated strategy, with the aim of balancing antagonistic/inhibitory activities several hours after dosing. TXS inhibition is expected to divert PGH2 to PGD2 and PGI2 (Vermylen et al., 1981; Smith, 1982), which both inhibit human platelet activation, and usually requires the presence of either a (N)-imidazole as found in dazoxiben (Randall et al., 1981) or a pyridin-3-yl group as in ridogrel (Hoet et al., 1990) and the related CV-4151 (isbogrel, Figure 8, Imura et al., 1988). ZD-1542 (Brownlie et al., 1993) is a pyridin-3-yl derivative related to ICI-192605, while GR-83783 (Campbell et al., 1991a), a relative of GR-32191, has a 4-(pyridin-3-yl)-phenyl moiety (Figure 8). The (N)-benzimidazole KW-3635 did not inhibit cow platelet TXS at 100 µM (Miki et al., 1992). Several groups have successfully expanded the biaryl region of ridogrel (Cozzi et al., 1994) or combined sulotroban/daltroban moieties with ridogrel/isbogrel moieties (Figure 8), resulting in GR-108774 (Campbell et al., 1991b), CGS-22652 (Bhagwat et al., 1993) and compound 36 in Soyka et al. (1993). In addition, Zeneca have used various tethers to connect the whole or part of the ICI-192605 nucleus to either a dazoxiben or an isbogrel nucleus (Figure 8) (Ackerley et al., 1995). BM-531 and BM-573 are combined TP antagonist/TXS inhibitors lacking a carboxylic acid group (Dognéet al., 2001; Rolin et al., 2001).

Finally, YM-158 has similar high affinity for TP and cys-LT receptors in guinea pig trachea (Arakida et al., 1998) (see later).

Therapeutic applications

The discovery of thromboxanes was Nobel prize-winning research. Thromboxane A2 is undoubtedly important in regulating cardiovascular homeostasis. Its biosynthesis by platelets and other tissues and highly potent actions on platelets and blood vessels provided a strong impetus for the development of TP antagonists (Patrono, 1990; Patscheke, 1990; Davis-Bruno and Halushka, 1994). However, these agents have not been a success to date. Clear evidence of benefit was not established in early studies (Misra, 1994), but economics was also a major factor. Low-dose aspirin, which targets platelet TXA2 synthesis by irreversibly acetylating COX-1, has proven benefits (Patrignani et al., 1982; Fitzgerald et al., 1983; Reilly and Fitzgerald, 1987). More potent TP antagonists, with superior pharmacokinetic profiles, were developed, but these also do not appear successful at the commercial level. The reasons are not entirely clear. At the pharmacological level, a TP antagonist would appear preferable to low-dose aspirin, especially in the light of recent events surrounding COX-2 inhibitors.

  1. The clinical side effects associated with COX-2 inhibitor therapy reveal the TXA2/PGI2 balance as more delicate to perturbation and more important than was previously believed.

  2. Low-dose aspirin does not always have the required TXS/PGIS selectivity (Knapp et al., 1988; Patscheke, 1990).

  3. Isoprostanes are alternative endogenous human TP receptor agonists and hence blockable by TP antagonists: their non-enzymatic formation would not be affected by aspirin, other COX inhibitors or TXS inhibitors (Kawikova et al., 1996; Gardan et al., 2000; Janssen et al., 2001).

  4. TP antagonists exhibit cardio-protective effects that are not shared by aspirin (Gomoll and Ogletree, 1994; Grover et al., 1994).

Given these considerations, a potent, highly selective TP antagonist may be worth revisiting in large-scale clinical trials on cardiovascular disease. As oxidative stress and resultant formation of isoprostanes is now linked to atherogenesis (Dognéet al., 2005), an initially unappreciated dimension to the TXA2/PGI2 balance and cardiovascular risk is made manifest. Specifically, the clinical outcomes would be related to stroke, heart attack and angina. Thromboxane A2 has also been implicated in hypertension occurring in pregnancy and TP antagonists (/TXS inhibitors) have been proposed as treatments for pre-eclampsia (Keith et al., 1993; Dognéet al., 2006).

In addition to the vasculature, TXA2 potently stimulates other smooth muscles to contract. Thus, TP antagonists have been proposed as potential therapeutic modalities for asthma. Seratrodast (Figure 8) has received marketing approval for treatment of asthma in Japan (see Rolin et al., 2006). Ramatroban and seratrodast have also been evaluated in Phase III clinical trials in the USA (Dognéet al., 2002). The therapeutic rationale is to ameliorate the marked bronchoconstriction produced by TXA2 and its involvement in bronchial hyperesponsiveness. However, TXA2 and other prostanoids that activate TP receptors (e.g. PGD2) are not the only powerful bronchoconstrictors generated in asthma and it is unlikely that a TP antagonist alone would be adequate therapy in most patients. In this context, SQ-29548 and the cys-LT antagonist montelukast acting alone inconsistently inhibited contraction of human lung slices in culture challenged with antigen; a combination of the antagonists was much more effective, while an H1 antagonist was ineffective (Wohlsen et al., 2003). A combined TP/cys-LT antagonist (e.g. YM-158) may have greater clinical efficacy.

Activation of EP3, FP, TP and possibly EP1 receptors causes contraction of the pregnant human myometrium, while DP1, EP2 and IP receptors mediate relaxation (Senior et al., 1993). Given this scenario, the potential of selective TP antagonists for treating labour-associated disorders, such as pre-term labour, would seem low. Recent studies on human myometrial specimens obtained at parturition provide a new perspective. It appears that there is a marked alteration in prostanoid receptor functional dynamics at term, prior to and following the onset of labour (Fischer et al., 2008). Comparing the effects of U-46619, PGE2 and PGF, a loss of FP but not TP responsiveness was apparent following the onset of labour (Fischer et al., 2008). As PGE2 produces a net inhibition of myogenic activity, this leaves TXA2 as the only COX-derived product capable of exerting a contractile effect during labour. The role of TXA2 in human parturition may be greater than previously envisaged.

The involvement of TXA2 in inflammatory bowel disease may be significant (Rampton and Collins, 1993). The aetiology of ulcerative colitis and Crohns' disease is not fully understood, but initiating factors in pre-disposed individuals elicit severe and prolonged inflammation of the gut mucosa. Early studies with ridogrel and picotamide, combined TP antagonist/TXS inhibitors, claimed some clinical improvement associated with reduced TXA2 (TXB2) release (Rampton and Collins, 1993). Given the multifactorial nature of these diseases and the number of other eicosanoids purported to be involved, pronounced beneficial effects of TP antagonists alone seem unlikely.

Evidence is available that activation of the TP receptor is implicated in chronic persistent cough (CPC). This is a detrimental and debilitating condition that serves no obvious function (French et al., 2002), afflicts 9–33% of the population in Europe and North America and for which no satisfactory treatment is currently available (see Chung and Pavord, 2008). Thus, the mechanisms underlying CPC and the identification of new anti-tussive agents present a grossly neglected and unmet clinical need.

In humans, PGs have high tussive potency suggesting that they may be released locally in a variety of respiratory diseases where cough is a characteristic symptom. In subjects with asthma, both indomethacin and ozagrel (OKY-046; a TXS inhibitor) increase the threshold for cough when compared with placebo, suggesting that TXA2 may be one of the cyclo-oxygenase products that sensitize the cough reflex (Fujimura et al., 1995). Indomethacin and ozagrel also attenuate cough in subjects with hypertension treated with angiotensin-converting enzyme inhibitors (Fogari et al., 1992; Malini et al., 1997; Umemura et al., 1997). Thus, the potential application for TP antagonists in alleviating CPC is clear. Indeed, in an animal model of asthma-related cough, the TP antagonist, seratrodast (Figure 8) significantly suppressed cough induced by capsaicin (Xiang et al., 2002). Moreover, in 16 patients with stable chronic bronchitis, seratrodast (80 mg b.i.d. for 28 days) significantly increased the threshold for capsaicin-induced cough (Ishiura et al., 2003). Finally, ramatroban is reported to attenuate cough in subjects with cough variant asthma (Kitamura et al., 2003), which could be due to blockade of both TP and DP2 receptors (Gardiner and Browne, 1984).

A role for TXA2 in cancer of the female breast and genital tract has also been suggested. Findings are reminiscent of the TXA2/PGI2 balance in the cardiovascular system: TXA2 promotes tumour growth and metastasis, while PGI2 exerts a protective effect by maintaining vascular and platelet homeostasis (Honn and Meyer, 1981; Nigam et al., 1992). A significant role in ovarian, cervical and other gynaecological cancers may exist (Nigam et al., 1992), but there remains insufficient evidence to assign a pivotal role.

Thromboxanes have also been implicated in a number of other conditions, including glomerulonephritis (Patrono, 1990; Wardle, 1999), allergic conjunctivitis (Woodward et al., 1990b), allergic rhinitis (Misra, 1994), diabetes (Dognéet al., 2006), septic shock and pulmonary embolism (Ghuysen et al., 2005). In summary, while numerous potential uses have been proposed over several years, clinical and economic success has been modest to date despite the availability of extremely potent and long-acting TP antagonists. What are the future therapeutic prospects? An argument could be made for superiority over low-dose aspirin. In order to establish this, extensive clinical trials would be essential. The bigger concern for the pharmaceutical industry would be ‘how much better would a TP antagonist perform compared with low-dose aspirin?’ If the difference is small, economics will dictate the fate of excellent TP antagonists. Perhaps the brightest new star in the galaxy of therapeutic options is the treatment of pre-term labour, given that TXA2 is now known to be the major prostanoid involved in parturition. The only potential side effect that may occur with TP antagonist therapy is possible enhancement of hypersensitivity reactions (Narumiya, 2003). This would not be a major burden under most circumstances.

Therapeutic future of prostanoid receptor antagonists

The future for selective prostanoid antagonists does not appear to provide an encouraging scenario. Potent and selective antagonists for every prostanoid receptor are now available, with the exception of the EP2 subtype. Some selective antagonists have received regulatory approval as drugs but these instances tend to be few and not worldwide. In this category are DP2 and TP antagonists. Drug approval filings/development have been discontinued on numerous TP antagonists, EP1 antagonists and, most recently, an EP4 antagonist. Arguments in favour of more potent and irreversible antagonists have been put forward but, in the global scheme of drug research, the current drugs seem adequate at the very least.

DP2 antagonists apart, of those drugs designed as selective for a single prostanoid receptor, the TP blockers still appear to be the best prospect. This would require revisiting cardiovascular studies and daring to directly compare with low-dose aspirin. The use of the TP antagonists for treating pre-term labour and even perhaps primary and secondary dysmenorrhoea should be contemplated in the light recent evidence (Fischer et al., 2008). This is, however, a rather narrow spectrum of utility and does not fulfil three decades of high expectation.

Straightforward consideration of the current situation does not favour the widespread therapeutic utility of highly selective prostanoid receptor antagonists. Stated simply, where COX inhibitors are clinically effective, selective antagonists are not. This is not really astonishing. There are five major prostanoids biosynthesized by COX and additional active metabolites (e.g. 13,14-dihydro-15-oxo PGD2, 19(R)-OH PGE2). The isoprostanes are formed by non-enzymatic oxidation of arachidonic acid (Morrow et al., 1990; 1994;). Finally, there are the PG-ethanolamides (prostamides) and PG-glyceryl esters, which are COX-2 products of the endocannabinoids anandamide and 2-arachidonyl glycerol. Expectations that one prostanoid and one dedicated receptor play a dominant, all-important role in most disease processes are lofty. It is unlikely that this all distils down to one important receptor activity in most diseases: a receptor widely controlling synergy or solely mediating a critical pathophysiological event.

Individual PGs and their receptors may play compensatory, fail-safe roles. Individual prostanoids may act sequentially to initiate and sustain disease states. They may subserve complementary roles. The role of DP1 and DP2 receptors in allergy, notably allergic rhinitis, provides an excellent example of two receptors behaving in a complementary manner to initiate and maintain the disease state. The combination of DP1/DP2 antagonism in a single molecule appears a very promising therapeutic approach (Pettipher, 2008) and one that appears feasible. Equally so, combining TP and EP3 antagonistic properties to prevent both platelet activation and vasoconstriction in cardiovascular disease states appears to be an achievable goal. Prostanoid-based drugs for pre-term labour may require even more versatility, with ideal therapy perhaps embodying. the following attributes in a single molecule: (i) TP antagonism, (ii) EP3 antagonism to block cervical ripening and (iii) EP2 agonism to provide a tocolytic effect. Such a molecule would be a formidable medicinal chemistry challenge. Further challenges relate to diseases where COX inhibitors are widely and successfully used, because the ideal spectrum of antagonist properties may not be entirely clear. If the promise of prostanoid-based therapeutics is ever to be fulfilled, the role of prostaglandins may need to be carefully thought-out on a disease-by-disease basis.

Glossary

Abbreviations:

Ang II

angiotensin II

BMP

bone morphogenetic protein

BP

blood pressure

ClogP

predicted n-octanol/water partition coefficient

COX

cyclo-oxygenase

CPC

chronic persistent cough

PG

prostaglandin

EP+/+/EP−/−

system involving wild-type/homozygous gene-deleted EP receptor

FA

free acid

NSAID

non-steroidal anti-inflammatory agent

PAF

platelet-activating factor

PGIS

prostacyclin synthase

PLC

phospholipase C

PRP

platelet-rich plasma

rc

recombinant

SAR

structure–activity relationships

TM

transmembrane

TXA2

thromboxane A2

TXS

thromboxane synthase

Conflicts of interest

The authors affirm no conflicts of interest.

References

  1. Abramovitz M, Boie Y, Nguyen T, Rushmore TH, Bayne MA, Metters KM, et al. Cloning and expression of a cDNA for the human prostanoid FP receptor. J Biol Chem. 1994;269:2632–2636. [PubMed] [Google Scholar]
  2. Abramovitz M, Adam M, Boie Y, Carrière M, Denis D, Godbout C, et al. The utilization of recombinant prostanoid receptors to determine the affinities and selectivities of prostaglandins and related analogs. Biochim Biophys Acta. 2000;1483:285–293. doi: 10.1016/s1388-1981(99)00164-x. [DOI] [PubMed] [Google Scholar]
  3. Ackerley N, Brewster AG, Brown GR, Clarke DS, Foubister AJ, Griffin SJ, et al. A novel approach to dual-acting thromboxane receptor antagonist/synthase inhibitors based on the link of 1,3-dioxane-thromboxane receptor antagonists and thromboxane synthase inhibitors. J Med Chem. 1995;38:1608–1628. doi: 10.1021/jm00010a005. [DOI] [PubMed] [Google Scholar]
  4. Ahn CH, Wallace LJ, Miller DD, Feller DR. Use of [3H]trimetoquinol as a radioligand in human platelets: interaction with putative endoperoxide/thromboxane A2 receptor sites. Thromb Res. 1988;50:387–399. doi: 10.1016/0049-3848(88)90268-x. [DOI] [PubMed] [Google Scholar]
  5. Alexander SPH, Mathie A, Peters JA. Guide to Receptors and Channels (GRAC), 3rd edn. Br J Pharmacol. 2008;153(Suppl. 2):S1–S209. doi: 10.1038/sj.bjp.0707746. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Almishri W, Cossette C, Rokach J, Martin JG, Hamid Q, Powell WS. Effects of prostaglandin D2, 15-deoxy-Δ12,14 prostaglandin J2, and selective DP1 and DP2 receptor agonists on pulmonary infiltration of eosinophils in Brown Norway rats. J Pharmacol Exp Ther. 2005;313:64–69. doi: 10.1124/jpet.104.079079. [DOI] [PubMed] [Google Scholar]
  7. Al Sinawi LA, Mekki QA, Hassan S, Hedges A, Burke C, Moody SG, et al. Effect of a hydantoin prostaglandin analogue, BW 245C, during oral dosing in man. Prostaglandins. 1985;29:99–111. doi: 10.1016/0090-6980(85)90155-8. [DOI] [PubMed] [Google Scholar]
  8. Amano H, Hayashi I, Endo H, Kitasato H, Yamashina S, Maruyama T, et al. Host prostaglandin E2-EP3 signaling regulates tumor-associated angiogenesis and tumor growth. J Exp Med. 2003;197:221–232. doi: 10.1084/jem.20021408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Angulo J, Cuevas P, La Fuente JM, Pomerol JM, Ruiz-Castañé E, Puigvert A, et al. Regulation of human penile smooth muscle tone by prostanoid receptors. Br J Pharmacol. 2002;136:23–30. doi: 10.1038/sj.bjp.0704675. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Arikawa T, Omura K, Morita I. Regulation of bone morphogenetic protein-2 expression by endogenous prostaglandin E2 in human mesenchymal stem cells. J Cell Physiol. 2004;200:400–406. doi: 10.1002/jcp.20031. [DOI] [PubMed] [Google Scholar]
  11. Arakida Y, Suwa K, Ohga K, Yokota M, Miyata K, Yamada T, et al. In vitro pharmacologic profile of YM158, a new dual antagonist for LTD4 and TXA2 receptors. J Pharmacol Exp Ther. 1998;287:633–639. [PubMed] [Google Scholar]
  12. Aronoff DM, Canetti C, Peters-Golden M. Prostaglandin E2 inhibits alveolar macrophage phagocytosis through an E-prostanoid 2 receptor-mediated increase in intracellular cyclic AMP. J Immunol. 2004;173:559–565. doi: 10.4049/jimmunol.173.1.559. [DOI] [PubMed] [Google Scholar]
  13. Arimura A, Yasui K, Kishino J, Asanuma F, Hasegawa H, Kakudo S, et al. Prevention of allergic inflammation by a novel prostaglandin receptor antagonist, S-5751. J Pharmacol Exp Ther. 2001;298:411–419. [PubMed] [Google Scholar]
  14. Arimura A, Kishino J, Tanimoto N. PGD2 receptor antagonist. WO2003/03097042, Shionogi & Co.
  15. Armer RE, Ashton MR, Boyd EA, Brennan CJ, Brookfield FA, Gazi L, et al. Indole-3-acetic acid antagonists of the prostaglandin D2 receptor CRTH2. J Med Chem. 2005;48:6174–6177. doi: 10.1021/jm050519b. [DOI] [PubMed] [Google Scholar]
  16. Armer RE, Boyd AE, Hay PA. 1-Acetic acid-indole derivatives with PGD2 antagonist activity. WO2006/095183, Oxagen Ltd.
  17. Armour CL, Johnson PR, Alfredson ML, Black JL. Characterization of contractile prostanoid receptors on human airway smooth muscle. Eur J Pharmacol. 1989;165:215–222. doi: 10.1016/0014-2999(89)90715-2. [DOI] [PubMed] [Google Scholar]
  18. Armstrong RA, Jones RL, Wilson NH. Ligand binding to thromboxane receptors on human platelets: correlation with biological activity. Br J Pharmacol. 1983;79:953–964. doi: 10.1111/j.1476-5381.1983.tb10541.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Armstrong RA, Jones RL, Pesssapati V, Will SG, Wilson NH. Competitive antagonism at thromboxane receptors in human platelets. Br J Pharmacol. 1985;84:595–607. doi: 10.1111/j.1476-5381.1985.tb16139.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Armstrong RA, Jones RL, MacDermot J, Wilson NH. Prostaglandin endoperoxide analogues which are both thromboxane receptor antagonists and prostacyclin mimetics. Br J Pharmacol. 1986;87:543–551. doi: 10.1111/j.1476-5381.1986.tb10196.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Armstrong RA, Lawrence RA, Jones RL, Wilson NH, Collie A. Functional and ligand binding studies suggest heterogeneity of platelet prostacyclin receptors. Br J Pharmacol. 1989;97:657–658. doi: 10.1111/j.1476-5381.1989.tb12001.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Armstrong RA, Humphrey PPA, Lumley P. Characteristics of the binding of [3H]-GR32191 to the thromboxane (TP-) receptor of human platelets. Br J Pharmacol. 1993;110:539–547. doi: 10.1111/j.1476-5381.1993.tb13844.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Arner M, Högestätt ED, Uski TK. Characterization of contraction-mediating prostanoid receptors in human hand veins: effects of the thromboxane receptor antagonists BM13,505 and AH23848. Acta Physiol Scand. 1991;141:79–86. doi: 10.1111/j.1748-1716.1991.tb09047.x. [DOI] [PubMed] [Google Scholar]
  24. Ashida Y, Matsumoto T, Kuriki H, Shiraishi M, Kato K, Terao S. A novel anti-asthmatic quinone derivative, AA-2414 with a potent antagonistic activity against a variety of spasmogenic prostanoids. Prostaglandins. 1989;38:91–112. doi: 10.1016/0090-6980(89)90019-1. [DOI] [PubMed] [Google Scholar]
  25. Audoly LP, Tilley SA, Goulet J, Key M, Nguyen H, Stock JL, et al. Identification of specific EP receptors responsible for the hemodynamic effects of PGE2. Am J Physiol. 1999;277:H924–H930. doi: 10.1152/ajpheart.1999.277.3.H924. [DOI] [PubMed] [Google Scholar]
  26. Awad MMA, Bazin M, Feru F, Goldstein SW, Kuhn CF. Tetrahydroquinoline derivatives as CRTH2 antagonists. WO2004/035543, Warner-Lambert & Co.
  27. Ayer LM, Wilson SM, Traves SL, Proud D, Giembycz MA. 4,5-Dihydro-1H-imidazol-2-yl)-[4-(4-isopropoxy-benzyl)-phenyl]-amine (RO1138452) is a selective, pseudo-irreversible orthosteric antagonist at the prostacyclin (IP)-receptor expressed by human airway epithelial cells: IP-receptor-mediated inhibition of CXCL9 and CXCL10 release. J Pharmacol Exp Ther. 2008;324:815–826. doi: 10.1124/jpet.107.129312. [DOI] [PubMed] [Google Scholar]
  28. Barbieri EJ, Orzechowski RF, Rossi GV. Measurement of prostaglandin E2 in an inflammatory exudate: effects of nonsteroidal anti-inflammatory agents. J Pharmacol Exp Ther. 1977;201:769–777. [PubMed] [Google Scholar]
  29. Bastien L, Sawyer N, Grygorczyk R, Metters KM, Adam M. Cloning, functional expression, and characterization of the human prostaglandin E2 receptor EP2 subtype. J Biol Chem. 1994;269:11873–11877. [PubMed] [Google Scholar]
  30. Bauer PH, Cheng JB, Gladue RP, Li B, Neote KS, Zhang J. Methods for the identification of compounds useful for the treatment of disease states mediated by prostaglandin D2. EP1170594, Pfizer Products Inc.
  31. Baxter GS, Clayton JK, Coleman RA, Marshall K, Sangha R, Senior J. Characterization of the prostanoid receptors mediating constriction and relaxation of human isolated uterine artery. Br J Pharmacol. 1995;116:1692–1696. doi: 10.1111/j.1476-5381.1995.tb16393.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Baxter A, Steele J, Teague S. Use of indole-3-acetic acids in the treatment of asthma, COPD and other diseases. WO2003/066046, AstraZeneca AB.
  33. Baxter A, Steele J, Teague S. Use of indole-3-acetic acids in the treatment of asthma, COPD and other diseases. WO2003/066047, AstraZeneca AB.
  34. Beasley RC, Featherstone RL, Church MK, Rafferty P, Varley JG, Harris A, et al. Effect of a thromboxane receptor antagonist on PGD2- and allergen-induced bronchoconstriction. J Appl Physiol. 1989;66:1685–1693. doi: 10.1152/jappl.1989.66.4.1685. [DOI] [PubMed] [Google Scholar]
  35. Beaulieu C, Guay D, Wang Z, Leblanc Y, Roy P, Dufresne C, et al. Identification of prostaglandin D2 receptor antagonists based on a tetrahydropyridoindole scaffold. Bioorg Med Chem Lett. 2008;18:2696–2700. doi: 10.1016/j.bmcl.2008.03.015. [DOI] [PubMed] [Google Scholar]
  36. Belley M, Gallant M, Roy B, Houde K, Lachance N, Labelle M, et al. Structure-activity relationship studies on ortho-substituted cinnamic acids, a new class of selective EP3 antagonists. Bioorg Med Chem Lett. 2005;15:527–530. doi: 10.1016/j.bmcl.2004.11.051. [DOI] [PubMed] [Google Scholar]
  37. Bennani YL, Tumey LN, Gleason EA, Robarge MJ. Indole acetic acids exhibiting CRTh2 receptor antagonism and uses thereof. US2006/0100425, Athersys Inc.
  38. Bennett A, Posner J. Studies on prostaglandin antagonists. Br J Pharmacol. 1971;42:584–594. doi: 10.1111/j.1476-5381.1971.tb07142.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Berthelette C, Lachance N, Li L, Sturino CF, Wang Z. Fluoro substituted cycloalkanoindoles, compositions containing such compounds and methods of treatment. WO2003/062200, Merck Frosst Canada & Co.
  40. Bexis S, McCormick PA, Docherty JR. Vascular actions of the prostacyclin receptor antagonist BAY 73-1449 in the portal hypertensive rat. Eur J Pharmacol. 2008;590:322–326. doi: 10.1016/j.ejphar.2008.06.006. [DOI] [PubMed] [Google Scholar]
  41. Bhagwat SS, Gude C, Cohen DS, Dotson R, Mathis J, Lee W, et al. Thromboxane receptor antagonism combined with thromboxane synthase inhibition. 5. Synthesis and evaluation of enantiomers of 8-[[(4-chlorophenyl)sulfonyl]amino]-4-(3-pyridinylalkyl)octanoic acid. J Med Chem. 1993;36:205–210. doi: 10.1021/jm00054a003. [DOI] [PubMed] [Google Scholar]
  42. Billot X, Chateauneuf A, Chauret N, Denis D, Greig G, Mathieu M-C, et al. Discovery of a potent and selective agonist of the prostaglandin EP4 receptor. Bioorg Med Chem Lett. 2003;13:1129–1132. doi: 10.1016/s0960-894x(03)00042-8. [DOI] [PubMed] [Google Scholar]
  43. Birkinshaw TN, Bonnert R, Cook A, Rasul R, Sanganee H, Teague S. Novel substituted indoles. WO2003/03101981, AstraZeneca AB.
  44. Birkinshaw TN, Teague SJ, Beech C, Bonnert RV, Hill S, Patel A, et al. Discovery of potent CRTh2 (DP2) receptor antagonists. Bioorg Med Chem Lett. 2006;16:4287–4290. doi: 10.1016/j.bmcl.2006.05.062. [DOI] [PubMed] [Google Scholar]
  45. Bley KR, Hunter JC, Eglen RM, Smith JA. The role of IP prostanoid receptors in inflammatory pain. Trends Pharmacol Sci. 1998;19:141–147. doi: 10.1016/s0165-6147(98)01185-7. [DOI] [PubMed] [Google Scholar]
  46. Bley KR, Bhattacharya A, Daniels DV, Gever J, Jahangir A, O'Yang C, et al. RO1138452 and RO3244794: characterization of structurally distinct, potent and selective IP (prostacyclin) receptor antagonists. Br J Pharmacol. 2006;147:335–345. doi: 10.1038/sj.bjp.0706554. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Boersma JI, Janzen KM, Oliveira L, Crankshaw DJ. Characterization of excitatory prostanoid receptors in the human umbilical artery in vitro. Br J Pharmacol. 1999;128:1505–1512. doi: 10.1038/sj.bjp.0702965. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Boie Y, Sawyer N, Slipetz DM, Metters KM, Abramovitz M. Molecular cloning and characterization of the human prostanoid DP receptor. J Biol Chem. 1995;270:18910–18916. doi: 10.1074/jbc.270.32.18910. [DOI] [PubMed] [Google Scholar]
  49. Boie Y, Stocco R, Sawyer N, Slipetz DM, Ungrin MD, Neuschafer-Rube F, et al. Molecular cloning and characterization of the four rat prostaglandin E2 prostanoid receptor subtypes. Eur J Pharmacol. 1997;340:227–241. doi: 10.1016/s0014-2999(97)01383-6. [DOI] [PubMed] [Google Scholar]
  50. Bonnert R, Rasul R. Novel substituted 3-sulfur indoles. WO2004/106302, AstraZeneca AB.
  51. Bonnert R, Brough S, Cook T, Dickinson M, Rasul R, Sanganee H, et al. Novel substituted indoles. WO2003/101961, AstraZeneca AB.
  52. Bonnert R, Dickinson M, Rasul R, Sanganee H, Teague S. Indole-3-sulphur derivatives. WO2004/007451, AstraZeneca AB.
  53. Bonnert R, Cook AR, Luker TJ, Mohammed RS, Thom S. Substituted indole derivates for pharmaceutical composition for treating respiratory diseases. WO2005/019171, AstraZeneca AB.
  54. Bonnert R, Mohammed RT, Teague S. 1-Acetic acid-indole, indazole and benzimidazole derivatives useful for the treatment of respiratory disorders. WO2005/054232, AstraZeneca AB.
  55. Bonnert RV, Patel A, Thom S. Phenoxyacetic acid derivatives. WO2005/018529A3, AstraZeneca AB.
  56. Bouayad A, Kajino H, Waleh N, Fouron J-C, Andelfinger G, Varma DR, et al. Characterization of PGE2 receptors in fetal and newborn lamb ductus arteriosus. Am J Physiol. 2001;280:H2342–H2349. doi: 10.1152/ajpheart.2001.280.5.H2342. [DOI] [PubMed] [Google Scholar]
  57. Brescia MR, Rokosz LL, Cole AG, Stauffer TM, Lehrach JM, Auld DS. Discovery and preliminary evaluation of 5-(4-phenylbenzyl)oxazole-4-carboxamides as prostacyclin receptor antagonists. Bioorg Med Chem Lett. 2007;17:1211–1215. doi: 10.1016/j.bmcl.2006.12.025. [DOI] [PubMed] [Google Scholar]
  58. Brewster AG, Brown GR, Foubister AJ, Jessup R, Smithers MJ. The synthesis of a novel thromboxane receptor antagonist 4(Z)-6-(2-O-chlorophenyl-4-O-hydroxyphenyl-1,3-dioxan-cis-5-yl) hexenoic acid, ICI 192605. Prostaglandins. 1988;36:173–178. doi: 10.1016/0090-6980(88)90304-8. [DOI] [PubMed] [Google Scholar]
  59. Breyer MD, Breyer RM. Prostaglandin receptors: their role in regulating renal function. Curr Opin Nephrol Hypertens. 2000;9:23–29. doi: 10.1097/00041552-200001000-00005. [DOI] [PubMed] [Google Scholar]
  60. Brownlie RP, Brownrigg NJ, Butcher HM, Garcia R, Jessup R, Lee VJ, et al. ZD1542, a potent thromboxane A2 synthase inhibitor and receptor antagonist in vitro. Br J Pharmacol. 1993;110:1600–1606. doi: 10.1111/j.1476-5381.1993.tb14007.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Bunce KT, Clayton NM, Coleman RA, Collington EW, Finch H, Humphray JM, et al. GR-63799X – a novel prostanoid with selectivity for EP3 receptors. Adv Prost Throm Leuk Res. 1990;21A:379–382. [PubMed] [Google Scholar]
  62. Burch JD, Belley M, Fortin R, Deschênes D, Girard M, Colucci J, et al. Structure-activity relationships and pharmacokinetic parameters of quinoline acylsulfonamides as potent and selective antagonists of the EP4 receptor. Bioorg Med Chem Lett. 2008;18:2048–2054. doi: 10.1016/j.bmcl.2008.01.103. [DOI] [PubMed] [Google Scholar]
  63. Campbell IB, Collington EW, Finch H, Hayes R, Lumley P, Mills K, et al. Synthesis and pharmacological evaluation of combined thromboxane receptor antagonist/synthase inhibitors: pyridine-containing amino-prostanoids. Bioorg Med Chem Lett. 1991a;1:695–698. [Google Scholar]
  64. Campbell IB, Collington EW, Finch H, Hayes R, Lumley P, Mills K, et al. Synthesis and pharmacological evaluation of combined thromboxane receptor antagonist/synthase inhibitors: pyridine-containing sulfonamido acids. Bioorg Med Chem Lett. 1991b;1:699–704. [Google Scholar]
  65. Camras CB, Toris CB, Sjoquist B, Milleson M, Thorngren JO, Hejkal TW, et al. Detection of the free acid of bimatoprost in aqueous humor samples from human eyes treated with bimatoprost before cataract surgery. Ophthalmology. 2004;111:2193–2198. doi: 10.1016/j.ophtha.2004.06.028. [DOI] [PubMed] [Google Scholar]
  66. Camu F, Shi L, Vanlersberghe C. The role of COX-2 inhibitors in pain modulation. Drugs. 2003;63(Suppl. 1):1–7. doi: 10.2165/00003495-200363001-00002. [DOI] [PubMed] [Google Scholar]
  67. Cefalu JS, Zhu QM, Eggers AC, Kaan TK, Ho MJ, Jett MF, et al. Effects of the selective prostacyclin receptor antagonist RO3244019 on the micturition reflex in rats. J Urol. 2007;178:2683–2686. doi: 10.1016/j.juro.2007.07.122. [DOI] [PubMed] [Google Scholar]
  68. Challis JR, Sloboda DM, Alfaidy N, Lye SJ, Gibb W, Patel FA, et al. Prostaglandins and mechanisms of preterm birth. Reproduction. 2002;124:1–17. doi: 10.1530/rep.0.1240001. [DOI] [PubMed] [Google Scholar]
  69. Chan CL, Jones RL, Lau HY. Characterization of prostanoid receptors mediating inhibition of histamine release from anti-IgE-activated rat peritoneal mast cells. Br J Pharmacol. 2000;129:589–597. doi: 10.1038/sj.bjp.0703072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Chan KM, Jones RL. Partial agonism of taprostene at prostanoid IP receptors in vascular preparations from guinea-pig, rat, and mouse. J Cardiovasc Pharmacol. 2004;43:795–807. doi: 10.1097/00005344-200406000-00009. [DOI] [PubMed] [Google Scholar]
  71. Chang SW, Reddy V, Pereira T, Dean BJ, Xia YQ, Seto C, et al. The pharmacokinetics and disposition of MK-0524, a prostaglandin D2 receptor 1 antagonist, in rats, dogs and monkeys. Xenobiotica. 2007;37:514–533. doi: 10.1080/00498250601175565. [DOI] [PubMed] [Google Scholar]
  72. Charlesworth EN, Kagey-Sobotka A, Schleimer RP, Norman PS, Lichtenstein LM. Prednisone inhibits the appearance of inflammatory mediators and the influx of eosinophils and basophils associated with the cutaneous late-phase response to allergen. J Immunol. 1991;146:671–676. [PubMed] [Google Scholar]
  73. Chemtob S, Peri AG. Peptide antagonists of prostaglandin F2α receptor. US2006/6984719, Hosptial Sainte-Justine, Montreal.
  74. Chen J, Woodward DF, Coleman RA, Jones RL, Lydford SJ. Prostanoid receptor assays. In: Enna SJ, Williams M, Ferkany T, Kenakin T, Persolt RD, Sullivan JP, editors. Current Protocols in Pharmacology. New York: John Wiley & Sons; 2001. p. 4.18.1–4.18.41. [DOI] [PubMed] [Google Scholar]
  75. Chen J, Senior J, Marshall K, Abbas F, Dinh H, Dinh T, et al. Studies using isolated uterine and other preparations show bimatoprost and prostanoid FP agonists have different activity profiles. Br J Pharmacol. 2005;144:493–501. doi: 10.1038/sj.bjp.0706044. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Cheng K, Wu TJ, Wu KK, Sturino C, Metters K, Gottesdiener K, et al. Antagonism of the prostaglandin D2 receptor 1 suppresses nicotinic acid-induced vasodilation in mice and humans. Proc Natl Acad Sci USA. 2006;103:6682–6687. doi: 10.1073/pnas.0601574103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Cherukuri DP, Chen XB, Goulet AC, Young RN, Han Y, Heimark RL, et al. The EP4 receptor antagonist, L-161,982, blocks prostaglandin E2-induced signal transduction and cell proliferation in HCA-7 colon cancer cells. Exp Cell Res. 2007;313:2969–2979. doi: 10.1016/j.yexcr.2007.06.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Chollet A, Tos EG, Cirillo R. Tocolytic effect of a selective FP receptor antagonist in rodent models reveals an innovative approach to the treatment of preterm labor. BMC Pregnancy Childbirth. 2007;7(Suppl. 1):S16. doi: 10.1186/1471-2393-7-S1-S16. open access. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Chow KB, Wong YH, Wise H. Prostacyclin receptor-independent inhibition of phospholipase C activity by non-prostanoid prostacyclin mimetics. Br J Pharmacol. 2001;134:1375–1384. doi: 10.1038/sj.bjp.0704388. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Chung KF, Pavord ID. Prevalence, pathogenesis, and causes of chronic cough. Lancet. 2008;371:1364–1374. doi: 10.1016/S0140-6736(08)60595-4. [DOI] [PubMed] [Google Scholar]
  81. Cimetière B, Dubuffet T, Muller O, Descombes JJ, Simonet S, Laubie M, et al. Synthesis and biological evaluation of new tetrahydronaphthalene derivatives as thromboxane receptor antagonists. Bioorg Med Chem Lett. 1998;8:1375–1380. doi: 10.1016/s0960-894x(98)00220-0. [DOI] [PubMed] [Google Scholar]
  82. Cirillo R, Tos EG, Page P, Missotten M, Quattropani A, Scheer A, et al. Arrest of preterm labor in rat and mouse by an oral and selective nonprostanoid antagonist of the prostaglandin F2α receptor (FP) Am J Obstet Gynecol. 2007;197(54):e1–e9. doi: 10.1016/j.ajog.2007.02.010. [DOI] [PubMed] [Google Scholar]
  83. Clark P, Rowland SE, Denis D, Mathieu MC, Stocco R, Poirier H, et al. MF498 [N-{[4-(5,9-Diethoxy-6-oxo-6,8-dihydro-7H-pyrrolo[3,4-g]quinolin-7-yl)-3-methylbenzyl]sulfonyl}-2-(2-methoxyphenyl)acetamide], a selective E prostanoid receptor 4 antagonist, relieves joint inflammation and pain in rodent models of rheumatoid and osteoarthritis. J Pharmacol Exp Ther. 2008;325:425–434. doi: 10.1124/jpet.107.134510. [DOI] [PubMed] [Google Scholar]
  84. Clark RD, Jahangir A, Severance D, Salazar R, Chang T, Chang D, et al. Discovery and SAR development of 2-(phenylamino) imidazolines as prostacyclin receptor antagonists. Bioorg Med Chem Lett. 2004;14:1053–1056. doi: 10.1016/j.bmcl.2003.10.070. [DOI] [PubMed] [Google Scholar]
  85. Clarke DL, Giembycz MA, Patel HJ, Belvisi MG. E-ring 8-isoprostanes inhibit ACh release from parasympathetic nerves innervating guinea-pig trachea through agonism of prostanoid receptors of the EP3-subtype. Br J Pharmacol. 2004;141:600–609. doi: 10.1038/sj.bjp.0705648. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Cocks TM, King SJ, Angus JA. Glibenclamide is a competitive antagonist of the thromboxane A2 receptor in dog coronary artery in vitro. Br J Pharmacol. 1990;100:375–378. doi: 10.1111/j.1476-5381.1990.tb15812.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Coleman RA, Sheldrick RL. Prostanoid-induced contraction of human bronchial smooth muscle is mediated by TP-receptors. Br J Pharmacol. 1989;96:688–692. doi: 10.1111/j.1476-5381.1989.tb11869.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Coleman RA, Humphrey PPA, Kennedy I, Lumley P. Prostanoid receptors: the development of a working classification. Trends Pharmacol Sci. 1984;5:303–306. [Google Scholar]
  89. Coleman RA, Kennedy I, Sheldrick RLG. New evidence with selective agonists and antagonists for the subclassification of PGE2-senstive (EP) receptors. Adv Prost Thromb Leukot Res. 1987;17:467–470. [PubMed] [Google Scholar]
  90. Coleman RA, Grix SP, Head SA, Louttit JB, Mallett A, Sheldrick RL. A novel inhibitory prostanoid receptor in piglet saphenous vein. Prostaglandins Leukot Essent Fatty Acids. 1994a;47:151–168. doi: 10.1016/0090-6980(94)90084-1. [DOI] [PubMed] [Google Scholar]
  91. Coleman RA, Smith WL, Narumiya S. VIII International union of pharmacology classification of prostanoid receptors: properties distribution, and structure of the receptors and their subtypes. Pharmacol Rev. 1994b;46:205–229. [PubMed] [Google Scholar]
  92. Colquhoun D. Why the Schild method is better than Schild realised. Trends Pharmacol Sci. 2007;28:608–614. doi: 10.1016/j.tips.2007.09.011. [DOI] [PubMed] [Google Scholar]
  93. Corradini L, Field MJ, Kinloch RA, Williams-Jones BI. Methods of treating neuropathic pain using a CRTH2 receptor antagonist. WO2005/102338, Warner-Lambert & Co.
  94. Cozzi P, Giordani A, Menichincheri M, Pillan A, Pinciroli V, Rossi A, et al. Agents combining thromboxane receptor antagonism with thromboxane synthase inhibition: [[[2-(1H-imidazol-1-yl)ethylidene]amino]oxy]alkanoic acids. J Med Chem. 1994;37:3588–3604. doi: 10.1021/jm00047a016. [DOI] [PubMed] [Google Scholar]
  95. Craig DA. The Cheng-Prusoff relationship: something lost in the translation. Trends Pharmacol Sci. 1993;14:89–91. doi: 10.1016/0165-6147(93)90070-z. [DOI] [PubMed] [Google Scholar]
  96. Crider JY, Griffin BW, Sharif NA. Prostaglandin DP receptors positively coupled to adenylyl cyclase in embryonic bovine tracheal (EBTr) cells: pharmacological characterization using agonists and antagonists. Br J Pharmacol. 1999;127:204–210. doi: 10.1038/sj.bjp.0702490. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Daray FM, Minvielle AI, Puppo S, Rothlin RP. Pharmacological characterization of prostanoid receptors mediating vasoconstriction in human umbilical vein. Br J Pharmacol. 2003;139:1409–1426. doi: 10.1038/sj.bjp.0705375. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Davis RJ, Murdoch CE, Ali M, Purbrick S, Ravid R, Baxter GS, et al. EP4 prostanoid receptor-mediated vasodilatation of human middle cerebral arteries. Br J Pharmacol. 2004;141:580–585. doi: 10.1038/sj.bjp.0705645. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Davis-Bruno KL, Halushka PV. Molecular pharmacology and therapeutic potential of thromboxane A2 receptor antagonists. Adv Drug Res. 1994;25:173–202. [Google Scholar]
  100. Delaey C, Van de Voorde J. Prostanoid-induced contractions are blocked by sulfonylureas. Eur J Pharmacol. 1995;280:179–184. doi: 10.1016/0014-2999(95)00201-u. [DOI] [PubMed] [Google Scholar]
  101. Dogné J-M, Rolin S, de Leval X, Benoit P, Neven P, Delarge J, et al. Pharmacology of the thromboxane receptor antagonist and thromboxane synthase inhibitor BM-531. Cardiovasc Drug Rev. 2001;19:87–96. doi: 10.1111/j.1527-3466.2001.tb00057.x. [DOI] [PubMed] [Google Scholar]
  102. Dogné J-M, de Leval X, Benoit P, Rolin S, Pirotte B, Masereel B. Therapeutic potential of thromboxane inhibitors in asthma. Expert Opin Investig Drug. 2002;11:275–281. doi: 10.1517/13543784.11.2.275. [DOI] [PubMed] [Google Scholar]
  103. Dogné J-M, Hanson J, Pratico D. Thromboxane, prostacyclin and isoprostanes: therapeutic targets in atherogenesis. Trends Pharmacol Sci. 2005;26:639–644. doi: 10.1016/j.tips.2005.10.001. [DOI] [PubMed] [Google Scholar]
  104. Dogné J-M, Hanson J, de Leval X, Pratico D, Pace-Asciak CR, Drion P, et al. From the design to the clinical application of thromboxane modulators. Curr Pharm Des. 2006;12:903–923. doi: 10.2174/138161206776055921. [DOI] [PubMed] [Google Scholar]
  105. Doheny HC, O'Reilly MJ, Sexton DJ, Morrison JJ. THG113.31, a specific PGF2alpha receptor antagonist, induces human myometrial relaxation and BKCa channel activation. Reprod Biol Endocrinol. 2007;5:10. doi: 10.1186/1477-7827-5-10. open access. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Dong YJ, Jones RL, Wilson NH. Prostaglandin E receptor subtypes in smooth muscle: agonist activities of stable prostacyclin analogues. Br J Pharmacol. 1986;87:97–107. doi: 10.1111/j.1476-5381.1986.tb10161.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Dorn GW. Distinct platelet thromboxane A2/prostaglandin H2 receptor subtypes. A radioligand binding study of human platelets. J Clin Invest. 1989;84:1883–1891. doi: 10.1172/JCI114375. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Dorn GW, Davis MG, D'Angelo DD. Structural determinants for agonist binding affinity to thromboxane/prostaglandin endoperoxide (TP) receptors. Analysis of chimeric rat/human TP receptors. J Biol Chem. 1997;272:12399–12405. doi: 10.1074/jbc.272.19.12399. [DOI] [PubMed] [Google Scholar]
  109. Drower EJ, Stapelfeld A, Mueller RA, Hammond DL. The antinociceptive effects of prostaglandin antagonists in the rat. Eur J Pharmacol. 1987;133:249–256. doi: 10.1016/0014-2999(87)90020-3. [DOI] [PubMed] [Google Scholar]
  110. Dubé GP, Mais DE, Jakubowski JA, Brune KA, Utterback BG, True TA, et al. In vitro characterization of a novel TXA2/PGH2 receptor ligand (S-145) in platelets and vascular and airway smooth muscle. J Pharmacol Exp Ther. 1992;262:784–791. [PubMed] [Google Scholar]
  111. Ducharme Y, Blouin M, Carrière MC, Chateauneuf A, Côté B, Denis D, et al. 2,3-Diarylthiophenes as selective EP1-receptor antagonists. Bioorg Med Chem Lett. 2005;15:1155–1160. doi: 10.1016/j.bmcl.2004.12.005. [DOI] [PubMed] [Google Scholar]
  112. Durocher Y, Perret S, Thibaudeau E, Gaumond MH, Kamen A, Stocco R, et al. A reporter gene assay for high-throughput screening of G-protein-coupled receptors stably or transiently expressed in HEK293 EBNA cells grown in suspension culture. Anal Biochem. 2000;284:316–326. doi: 10.1006/abio.2000.4698. [DOI] [PubMed] [Google Scholar]
  113. Eglen RM, Whiting RL. The action of prostanoid receptor agonists and antagonists on smooth muscle and platelets. Br J Pharmacol. 1988;94:591–601. doi: 10.1111/j.1476-5381.1988.tb11565.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Faith A, McDonald J, Peek E, Richards D, Caulfield J, Chevretton E, et al. Functional plasticity of human respiratory tract dendritic cells: GM-CSF enhances Th2 development. J Allergy Clin Immunol. 2005;116:1136–1143. doi: 10.1016/j.jaci.2005.08.002. [DOI] [PubMed] [Google Scholar]
  115. Farmer JB, Farrar DG, Wilson J. Antagonism of tone and prostaglandin-mediated responses in a tracheal preparation by indomethacin and SC-19220. Br J Pharmacol. 1974;52:559–565. doi: 10.1111/j.1476-5381.1974.tb09724.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Featherstone RL, Robinson C, Holgate ST, Church MK. Evidence for thromboxane receptor mediated contraction of guinea-pig and human airways in vitro by prostaglandin (PG) D2, 9α,11β-PGF2 and PGF2α. Naunyn Schmied Arch Pharmacol. 1990;341:439–443. doi: 10.1007/BF00176337. [DOI] [PubMed] [Google Scholar]
  117. Fernandes B, Crankshaw D. Functional characterization of the prostanoid DP receptor in human myometrium. Eur J Pharmacol. 1995;283:73–81. doi: 10.1016/0014-2999(95)00288-v. [DOI] [PubMed] [Google Scholar]
  118. Fischer DP, Hutchinson JA, Farrar D, O'Donovan PJ, Woodward DF, Marshall KM. Loss of prostanoid F2α, but not thromboxane, responsiveness in pregnant human myometrium during labour. Endocrinology. 2008;197:171–179. doi: 10.1677/JOE-07-0494. [DOI] [PubMed] [Google Scholar]
  119. Fitch WL, Berry PW, Tu Y, Tabatabaei A, Lowrie L, Lopez-Tapia F, et al. Identification of glutathione-derived metabolites from an IP receptor antagonist. Drug Metab Dispos. 2004;32:1482–1490. doi: 10.1124/dmd.104.000471. [DOI] [PubMed] [Google Scholar]
  120. Fitzgerald GA, Oates JA, Hawiger J, Maas RL, Roberts LJ, 2nd, Lawson JA, et al. Endogenous biosynthesis of prostacyclin and thromboxane and platelet function during chronic administration of aspirin in man. J Clin Invest. 1983;71:676–688. doi: 10.1172/JCI110814. [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Fogari R, Zoppi A, Tettamanti F, Malamani GD, Tinelli C, Salvetti A. Effects of nifedipine and indomethacin on cough induced by angiotensin-converting enzyme inhibitors: a double-blind, randomized, cross-over study. J Cardiovasc Pharmacol. 1992;19:670–673. [PubMed] [Google Scholar]
  122. Folger WH, Halushka PV, Wilcox CS, Guzman NJ. Characterization of rat glomerular thromboxane A2 receptors: comparison to rat platelets. Eur J Pharmacol. 1992;227:71–78. doi: 10.1016/0922-4106(92)90144-k. [DOI] [PubMed] [Google Scholar]
  123. Foudi N, Kotelevets L, Louedec L, Leséche G, Henin D, Chastre E, et al. Vasorelaxation induced by prostaglandin E2 in human pulmonary vein: role of the EP4 receptor subtype. Br J Pharmacol. 2008;154:1631–1639. doi: 10.1038/bjp.2008.214. [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. French CT, Irwin RS, Fletcher KE, Adams TM. Evaluation of a cough-specific quality-of-life questionnaire. Chest. 2002;121:1123–1131. doi: 10.1378/chest.121.4.1123. [DOI] [PubMed] [Google Scholar]
  125. Fretz H, Fecher A, Hilpert K, Riederer M. Tetrahydropyridoindole derivatives. WO2005/095397, Actelion Pharmaceuticals.
  126. Friel AM, O'Reilly MW, Sexton DJ, Morrison JJ. Specific PGF2α receptor (FP) antagonism and human uterine contractility in vitro. BJOG Int. 2005;112:1034–1042. doi: 10.1111/j.1471-0528.2005.00658.x. [DOI] [PubMed] [Google Scholar]
  127. Fruzzetti F, Melis GB, Strigini F, Vettori C, Ricci C, Fioreit P. Use of sulprostone for induction of preoperative cervical dilation or uterine evacuation, a comparison among the effects of different treatment schedules. Obstet Gynecol. 1988;72:704–708. [PubMed] [Google Scholar]
  128. Fujimura M, Kamio Y, Kasahara K, Bando T, Hashimoto T, Matsuda T. Prostanoids and cough response to capsaicin in asthma and chronic bronchitis. Eur Resp J. 1995;8:1499–1505. [PubMed] [Google Scholar]
  129. Fujino H, Regan JW. FP Prostanoid receptor activation of a T-cell factor/β-catenin signaling pathway. J Biol Chem. 2001;276:12489–12492. doi: 10.1074/jbc.C100039200. [DOI] [PubMed] [Google Scholar]
  130. Fujino H, Regan JW. Prostanoid receptors and phosphatidylinositol 3-kinase: a pathway to cancer. Trends Pharmacol Sci. 2003;24:335–340. doi: 10.1016/S0165-6147(03)00162-7. [DOI] [PubMed] [Google Scholar]
  131. Fujitani Y, Kanaoka Y, Aritake K, Uodome N, Okazaki-Hatake K, Urade Y. Pronounced eosinophilic lung inflammation and Th2 cytokine release in human lipocalin-type prostaglandin D synthase transgenic mice. J Immunol. 2002;168:443–449. doi: 10.4049/jimmunol.168.1.443. [DOI] [PubMed] [Google Scholar]
  132. Fulton AM, Ma X, Kundu N. Targeting prostaglandin E EP receptors to inhibit metastasis. Cancer Res. 2006;66:9794–9797. doi: 10.1158/0008-5472.CAN-06-2067. [DOI] [PubMed] [Google Scholar]
  133. Funk CD, Furci L, Fitzgerald GA, Grygorczyk R, Rochette C, Bayne MA, et al. Cloning and expression of a cDNA for the human prostaglandin E receptor EP1 subtype. J Biol Chem. 1993a;268:26767–26772. [PubMed] [Google Scholar]
  134. Funk CD, Furci L, Moran N, Fitzgerald GA. Point mutation in the seventh hydrophobic domain of the human thromboxane A2 receptor allows discrimination between agonist and antagonist binding sites. Mol Pharmacol. 1993b;44:934–949. [PubMed] [Google Scholar]
  135. Furci L, Fitzgerald DJ, Fitzgerald GA. Heterogeneity of prostaglandin H2/thromboxane A2 receptors: distinct subtypes mediate vascular smooth muscle contraction and platelet aggregation. J Pharmacol Exp Ther. 1991;258:74–81. [PubMed] [Google Scholar]
  136. Gallant M, Carrière MC, Chateauneuf A, Denis D, Gareau Y, Godbout C, et al. Structure-activity relationship of biaryl acylsulfonamide analogues on the human EP3 prostanoid receptor. Bioorg Med Chem Lett. 2002;12:583–2586. doi: 10.1016/s0960-894x(02)00518-8. [DOI] [PubMed] [Google Scholar]
  137. Gardan B, Crakowski JL, Sessa C, Hunt M, Stanke-Labesque F, Devillier P, et al. Vasoconstrictor effects of iso-prostaglandin F2α type-III (8-iso-prostaglandin F2α) on human saphenous veins. J Cardiovasc Pharmacol. 2000;35:729–734. doi: 10.1097/00005344-200005000-00008. [DOI] [PubMed] [Google Scholar]
  138. Gardiner PJ, Browne JL. Tussive activity of inhaled PGD2 in the cat and characterisation of the receptor(s) involved. Prostaglandins Leukot Med. 1984;14:153–159. doi: 10.1016/0262-1746(84)90034-9. [DOI] [PubMed] [Google Scholar]
  139. Ghosh S, Elder AM, Carson KG, Sprott K, Harrison S. PGD2 receptor antagonists for the treatment of inflammatory diseases. WO2004/032848, Millennium Pharmaceuticals.
  140. Ghosh S, Elder AM, Carson KG, Sprott KT, Harrison SJ, Hicks FA, et al. PGD2 receptor antagonists for the treatment of inflammatory diseases. WO2005/100321, Millennium Pharmaceuticals.
  141. Ghuysen A, Dogné JM, Chiap P, Rolin S, Masereel B, Lambermont B. Pharmacological profile and therapeutic potential of BM-573, a combined thromboxane receptor antagonist and synthase inhibitor. Cardiovasc Drug Rev. 2005;23:1–14. doi: 10.1111/j.1527-3466.2005.tb00153.x. [DOI] [PubMed] [Google Scholar]
  142. Giblin GM, Bit RA, Brown SH, Chaignot HM, Chowdhury A, Chessell IP, et al. The discovery of 6-[2-(5-chloro-2-{[(2,4-difluorophenyl)methyl]oxy}phenyl)-1-cyclopenten-1-yl]-2-pyridinecarboxylic acid, GW848687X, a potent and selective prostaglandin EP1 receptor antagonist for the treatment of inflammatory pain. Bioorg Med Chem Lett. 2007;17:385–389. doi: 10.1016/j.bmcl.2006.10.041. [DOI] [PubMed] [Google Scholar]
  143. Gil DW, Cheevers CV, Donello JE. Transient allodynia pain models in mice for early assessment of analgesic activity. Br J Pharmacol. 2008;153:769–774. doi: 10.1038/sj.bjp.0707412. [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Giles H, Leff P. Classification of platelet and vascular prostaglandin D2 (DP) receptors: estimation of affinities and relative efficacies for a series of novel bicyclic ligands. With an appendix on goodness-of-fit analyses. Br J Pharmacol. 1992;106:996–1003. doi: 10.1111/j.1476-5381.1992.tb14447.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Giles H, Leff P, Bolofo ML, Kelly MG, Robertson AD. The classification of prostaglandin DP-receptors in platelets and vasculature using BW A868C, a novel, selective and potent competitive antagonist. Br J Pharmacol. 1989;96:291–300. doi: 10.1111/j.1476-5381.1989.tb11816.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Glass M, Hong J, Sato TA, Mitchell MD. Misidentification of prostamides as prostaglandins. J Lipid Res. 2005;46:1364–1368. doi: 10.1194/jlr.C500006-JLR200. [DOI] [PubMed] [Google Scholar]
  147. Goldenberg RL. The management of preterm labor. Obstet Gynecol. 2002;100:1020–1037. doi: 10.1016/s0029-7844(02)02212-3. [DOI] [PubMed] [Google Scholar]
  148. Gomoll AW, Ogletree ML. Failure of aspirin to interfere with the cardioprotective effects of ifetroban. Eur J Pharmacol. 1994;271:471–479. doi: 10.1016/0014-2999(94)90808-7. [DOI] [PubMed] [Google Scholar]
  149. Gray SJ, Giles H, Posner J. The effect of a prostaglandin DP-receptor partial agonist (192C86) on platelet aggregation and the cardiovascular system in healthy volunteers. Br J Clin Pharmacol. 1992;34:344–351. doi: 10.1111/j.1365-2125.1992.tb05641.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Griffin BW, Klimko P, Crider JY, Sharif NA. AL-8810: a novel prostaglandin F2 alpha analog with selective antagonist effects at the prostaglandin F2 alpha (FP) receptor. J Pharmacol Exp Ther. 1999;290:1278–1284. [PubMed] [Google Scholar]
  151. Grover GJ, Schumacher WA, Ogletree ML. Thromboxane receptor antagonist BMS-180291, but not aspirin, reduces the severity of pacing-induced ischaemia in dogs. J Cardiovasc Pharmacol. 1994;24:493–499. doi: 10.1097/00005344-199409000-00019. [DOI] [PubMed] [Google Scholar]
  152. Guan Y, Zhang Y, Wu J, Qi Z, Yang G, Dou D, et al. Antihypertensive effects of selective prostaglandin E2 receptor subtype 1 targeting. J Clin Invest. 2007;117:2496–2505. doi: 10.1172/JCI29838. [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Habib A, Fitzgerald GA, Maclouf J. Phosphorylation of the thromboxane receptor alpha, the predominant isoform expressed in human platelets. J Biol Chem. 1999;274:2645–2651. doi: 10.1074/jbc.274.5.2645. [DOI] [PubMed] [Google Scholar]
  154. Hall A, Atkinson S, Brown SH, Chessell IP, Chowdhury A, Giblin GM, et al. Discovery of novel, non-acidic 1,5-biaryl pyrrole EP1 receptor antagonists. Bioorg Med Chem Lett. 2007a;17:4450–4455. doi: 10.1016/j.bmcl.2006.12.021. [DOI] [PubMed] [Google Scholar]
  155. Hall A, Brown SH, Chowdhury A, Giblin GM, Gibson M, Healy MP, et al. Identification and optimization of novel 1,3,4-oxadiazole EP1 receptor antagonists. Bioorg Med Chem Lett. 2007b;17:1200–1205. doi: 10.1016/j.bmcl.2007.06.014. [DOI] [PubMed] [Google Scholar]
  156. Hall RA, Gillard J, Guindon Y, Letts G, Champion E, Ethier D, et al. Pharmacology of L-655,240 (3-[1-(4-chlorobenzyl)-5-fluoro-3-methyl-indol-2-yl]2,2-dimethylpropanoic acid); a potent, selective thromboxane/prostaglandin endoperoxide antagonist. Eur J Pharmacol. 1987;135:193–201. doi: 10.1016/0014-2999(87)90611-x. [DOI] [PubMed] [Google Scholar]
  157. Hallinan EA, Hagen TJ, Husa RK, Tsymbalov S, Rao SN, vanHoeck J-P, et al. N-substituted dibenzoxazepines as analgesic PGE2 antagonists. J Med Chem. 1993;36:3293–3299. doi: 10.1021/jm00074a010. [DOI] [PubMed] [Google Scholar]
  158. Hallinan EA, Stapelfeld A, Savage MA, Reichman M. 8-Chlorodibenz[b,f][1,4]oxazepine-10(11H)-carboxylic acid, 2-[3-[2-(furanylmethyl)thio]-1-oxopropyl]hydrazide (SC-51322): a potent PGE2 antagonist and analgesic. Bioorg Med Chem Lett. 1994;4:509–514. [Google Scholar]
  159. Hammad H, Lambrecht BN. Dendritic cells and epithelial cells: linking innate and adaptive immunity in asthma. Nat Rev Immunol. 2008;8:193–204. doi: 10.1038/nri2275. [DOI] [PubMed] [Google Scholar]
  160. Hammad H, Kool M, Soullie T, Narumiya S, Trottein F, Hoogsteden HC, et al. Activation of the D prostanoid 1 receptor suppresses asthma by modulation of lung dendritic cell function and induction of regulatory T cells. J Exp Med. 2007;204:357–367. doi: 10.1084/jem.20061196. [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Hanasaki K, Nakano K, Kasai H, Arita H, Ohtani K, Doteuchi M. Specific receptors for thromboxane A2 in cultured vascular smooth muscle cells of rat aorta. Biochem Biophys Res Commun. 1988;150:1170–1175. doi: 10.1016/0006-291x(88)90752-8. [DOI] [PubMed] [Google Scholar]
  162. Harris KJ. A combination of niacin and a prostaglandin D2 receptor antagonist. WO2008/039882, Sanofi-Aventis.
  163. Hawkey CJ. Management of gastroduodenal ulceration caused by non-steroidal anti-inflammatory drugs. Baillieres Best Pract Res Clin Gastroenterol. 2000;14:173–192. doi: 10.1053/bega.1999.0067. [DOI] [PubMed] [Google Scholar]
  164. Heavey DJ, Lumley P, Barrow SE, Murphy MB, Humphrey PP, Dollery CT. Effects of intravenous infusions of prostaglandin D2 in man. Prostaglandins. 1984;28:755–767. doi: 10.1016/0090-6980(84)90033-9. [DOI] [PubMed] [Google Scholar]
  165. Hedberg A, Hall SE, Ogletree ML, Harris DN, Liu EC. Characterization of [5,6-3H]SQ 29,548 as a high affinity radioligand, binding to thromboxane A2/prostaglandin H2-receptors in human platelets. J Pharmacol Exp Ther. 1988;245:786–792. [PubMed] [Google Scholar]
  166. Hefferan MP, Carter C, Haley M, Loomis CW. Spinal nerve injury activates prostaglandin synthesis in the spinal cord that contributes to early maintenance of tactile allodynia. Pain. 2003;101:139–147. doi: 10.1016/s0304-3959(02)00322-6. [DOI] [PubMed] [Google Scholar]
  167. Heinemann A, Schuligoi R, Sabroe I, Hartnell A, Peskar BA. Δ12-Prostaglandin J2, a plasma metabolite of prostaglandin D2, causes eosinophil mobilization from the bone marrow and primes eosinophils for chemotaxis. J Immunol. 2003;170:4752–4758. doi: 10.4049/jimmunol.170.9.4752. [DOI] [PubMed] [Google Scholar]
  168. Heptinstall S, Espinosa DI, Manolopoulos P, Glenn JR, White AE, Johnson A, et al. DG-041 inhibits the EP3 prostanoid receptor – a new target for inhibition of platelet function in atherosclerotic disease. Platelets. 2008;19:605–613. doi: 10.1080/09537100802351073. [DOI] [PubMed] [Google Scholar]
  169. Hirai H, Tanaka K, Yoshie O, Ogawa K, Kenmotsu K, Takamori Y, et al. Prostaglandin D2 selectively induces chemotaxis in T helper type 2 cells, eosinophils, and basophils via seven-transmembrane receptor CRTH2. J Exp Med. 2001;193:255–261. doi: 10.1084/jem.193.2.255. [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. Hirai H, Tanaka K, Takano S, Ichimasa M, Nakamura M, Nagata K. Agonistic effect of indomethacin on a prostaglandin D2 receptor, CRTH2. J Immunol. 2002;168:981–985. doi: 10.4049/jimmunol.168.3.981. [DOI] [PubMed] [Google Scholar]
  171. Hirata M, Hayashi Y, Ushikubi F, Yokota Y, Kageyama R, Nakanishi S, et al. Cloning and expression of cDNA for a human thromboxane A2 receptor. Nature. 1991;349:617–620. doi: 10.1038/349617a0. [DOI] [PubMed] [Google Scholar]
  172. Hirata T, Ushikubi F, Kakizuka A, Okuma M, Narumiya S. Two thromboxane A2 receptor isoforms in human platelets. Opposite coupling to adenylyl cyclase with different sensitivity to Arg60 to Leu mutation. J Clin Invest. 1996;97:949–956. doi: 10.1172/JCI118518. [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Hirst JR, Parkington HC, Young IR, Palliser HK, Peri KG, Olson DM. Delay of preterm birth in sheep by PHG 113.31, a prostaglandin F2α receptor antagonist. Am J Obstet Gynecol. 2005;193:256–266. doi: 10.1016/j.ajog.2004.11.009. [DOI] [PubMed] [Google Scholar]
  174. Hoet B, Falcon C, De Reys S, Arnout J, Deckmyn H, Vermylen J. R68070, a combined thromboxane/endoperoxide receptor antagonist and thromboxane synthase inhibitor, inhibits human platelet activation in vitro and in vivo: a comparison with aspirin. Blood. 1990;75:646–653. [PubMed] [Google Scholar]
  175. Honma T, Hiramatsu K, Arimura A. Benzothiophenecarboxamide derivatives and PGD2 antagonists comprising them. WO98/25919, Shionogi & Co Ltd.
  176. Honn KV, Meyer J. Thromboxanes and prostacyclin: positive and negative modulators of tumour growth. Biochem Biophys Res Comm. 1981;102:1122–1129. doi: 10.1016/s0006-291x(81)80128-3. [DOI] [PubMed] [Google Scholar]
  177. Horne WC. Desacetyl-1-nantradol: a selective prostaglandin antagonist. Prostaglandins Leukot Med. 1984;15:129–136. doi: 10.1016/0262-1746(84)90063-5. [DOI] [PubMed] [Google Scholar]
  178. Huang JL, Gao PS, Mathias RA, Yao TC, Chen LC, Kuo ML, et al. Sequence variants of the gene encoding chemoattractant receptor expressed on Th2 cells (CRTH2) are associated with asthma and differentially influence mRNA stability. Hum Mol Genet. 2004;13:2691–2697. doi: 10.1093/hmg/ddh279. [DOI] [PubMed] [Google Scholar]
  179. Hung GH, Jones RL, Lam FF, Chan KM, Hidaka H, Suzuki M, et al. Investigation of the pronounced synergism between prostaglandin E2 and other constrictor agents on rat femoral artery. Prostaglandins Leukot Essent Fatty Acids. 2006;74:401–405. doi: 10.1016/j.plefa.2006.04.002. [DOI] [PubMed] [Google Scholar]
  180. Hutchinson J, Marshall KM, Senior J. Preliminary studies using a putative FP-receptor antagonist, Al-8810, on isolated mouse uterus. Proceedings of the British Pharmacological Society at http://www.pa2online.org/Vol1Issue3abst038P.html.
  181. Ikeda M, Kawatani M, Maruyama T, Ishihama H. Prostaglandin facilitates afferent nerve activity via EP1 receptors during urinary bladder inflammation in rats. Biomed Res. 2006;27:49–54. doi: 10.2220/biomedres.27.49. [DOI] [PubMed] [Google Scholar]
  182. Imura Y, Terashita Z, Shibouta Y, Nishikawa K. The thromboxane A2/prostaglandin endoperoxide receptor antagonist activity of CV-1451, a thromboxane A2 synthetase inhibitor. Eur J Pharmacol. 1988;147:359–365. doi: 10.1016/0014-2999(88)90169-0. [DOI] [PubMed] [Google Scholar]
  183. Imura Y, Terashita Z, Shibouta Y, Inada Y, Nishikawa K. Antagonistic action of AA-2414 on thromboxane A2/prostaglandin endoperoxide receptor in platelets and blood vessels. Jpn J Pharmacol. 1990;52:35–43. doi: 10.1254/jjp.52.35. [DOI] [PubMed] [Google Scholar]
  184. Ishiguro C, Fujita T, Omori T, Fujii Y, Mayama T, Sato T. Assessing the effects of non-steroidal anti-inflammatory drugs on antihypertensive drug therapy using post-marketing surveillance database. J Epidemiol. 2008;18:119–124. doi: 10.2188/jea.JE2007413. [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Ishiura Y, Fujimura Y, Yamamori C, Nobata K, Myou S, Kurashima K, et al. Thromboxane antagonism and cough in chronic bronchitis. Ann Med. 2003;35:135–139. doi: 10.1080/07853890310010032. [DOI] [PubMed] [Google Scholar]
  186. Ito S, Okuda E, Sugama K, Negishi M, Hayaishi O. Evaluation of ZK110841 and AH6809, an agonist and an antagonist of prostaglandin DP-receptors on human platelets, with a PGD2-responsive cell line from bovine embryonic trachea. Br J Pharmacol. 1990;99:13–14. doi: 10.1111/j.1476-5381.1990.tb14645.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  187. Itoh Y, Shindoh J, Horiba M, Kohno S, Ohata K, Ashida Y, et al. Inhibitory effects of AA-2414, a thromboxane (Tx) A2 receptor antagonist, on U-46619-, prostaglandin (PG) D2- and 9α,11β PGF2-induced contractions of guinea-pig tracheas and isolated human bronchi (Japanese) Arerugi. 1993;42:1670–1676. [PubMed] [Google Scholar]
  188. Ivey KN, Srivastava D. The paradoxical patent ductus arteriosus. J Clin Invest. 2006;116:2863–2866. doi: 10.1172/JCI30349. [DOI] [PMC free article] [PubMed] [Google Scholar]
  189. Iwasaki M, Nagata K, Takano S, Takahashi K, Ishii N, Ikezawa Z. Association of a new-type prostaglandin D2 receptor CRTH2 with circulating T helper 2 cells in patients with atopic dermatitis. J Invest Dermatol. 2002;119:609–616. doi: 10.1046/j.1523-1747.2002.01862.x. [DOI] [PubMed] [Google Scholar]
  190. Jadhav V, Jabre A, Lin SZ, Lee TJ. EP1- and EP3-receptors mediate prostaglandin E2-induced constriction of porcine large cerebral arteries. J Cereb Blood Flow Metab. 2004;24:1305–1316. doi: 10.1097/01.WCB.0000139446.61789.14. [DOI] [PubMed] [Google Scholar]
  191. Janssen LJ, Premji M, Netherton S, Coruzzi J, Lu-Chao H, Cox PG. Vasoconstrictor actions of isoprostanes via tyrosine kinase and Rho kinase in human and canine pulmonary vascular smooth muscles. Br J Pharmacol. 2001;132:127–134. doi: 10.1038/sj.bjp.0703784. [DOI] [PMC free article] [PubMed] [Google Scholar]
  192. Janssens WJ, Cools FJ, Hoskens LA, Van Neuten JM. Effect of ridogrel on vascular contractions caused by vasoactive substances released during platelet activation. Thromb Haemost. 1990;64:91–96. [PubMed] [Google Scholar]
  193. Jenkins DW, Feniuk W, Humphrey PPA. Characterization of the prostanoid receptor types involved in mediating calcitonin gene-related peptide release from cultured rat trigeminal neurones. Br J Pharmacol. 2001;134:1296–1302. doi: 10.1038/sj.bjp.0704357. [DOI] [PMC free article] [PubMed] [Google Scholar]
  194. Johnson AG, Nguyen TV, Day RO. Do nonsteroidal anti-inflammatory drugs affect blood pressure? A meta-analysis. Ann Intern Med. 1994;121:289–300. doi: 10.7326/0003-4819-121-4-199408150-00011. [DOI] [PubMed] [Google Scholar]
  195. Jones R, Woodward D, Wang J. Investigation of the slow kinetics of the prostanoid EP3 receptor antagonists L-798106 and L-826266 on guinea-pig aorta. Fundam Clin Pharmacol. 2008;22(Suppl. 2):P078. [Google Scholar]
  196. Jones RL. Cardiovascular actions of prostaglandins D and E in the sheep: evidence for two distinct receptors. Adv Prostaglandin Thromboxane Res. 1976;1:221–230. [PubMed] [Google Scholar]
  197. Jones RL. Definition of prostaglandin-sensitive arterial constrictor systems. Acta Biol Med Ger. 1978;37:837–844. [PubMed] [Google Scholar]
  198. Jones RL. An update on ligands for prostanoid receptors. Sigma-RBI Newsletter for Cell Signallng and Neurosci Res. 2004;20:1–9. http://www.sigmaaldrich.com/sigma/general%20information/prostanoidreceptors.pdf. [Google Scholar]
  199. Jones RL, Chan K. Distinction between relaxations induced via prostanoid EP4 and IP1 receptors in pig and rabbit blood vessels. Br J Pharmacol. 2001;134:313–324. doi: 10.1038/sj.bjp.0704252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  200. Jones RL, Chan K. Investigation of the agonist activity of prostacyclin analogues on prostanoid EP4 receptors using GW 627368 and taprostene: evidence for species differences. Prost Leukot Essent Fatty Acids. 2005;72:289–299. doi: 10.1016/j.plefa.2004.12.002. [DOI] [PubMed] [Google Scholar]
  201. Jones RL, Wilson NH. Tautomerism of the enedione system of 15-oxoprostaglandin D2. J Org Chem (Perkin 1) 1978:209–214. [Google Scholar]
  202. Jones RL, Peesapati V, Wilson NH. Antagonism of the thromboxane-sensitive contractile systems of the rabbit aorta, dog saphenous vein and guinea-pig trachea. Br J Pharmacol. 1982;76:423–438. doi: 10.1111/j.1476-5381.1982.tb09236.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  203. Jones RL, Wilson NH, Lawrence RA. EP 171: a high affinity thromboxane A2-mimetic, the actions of which are slowly reversed by receptor blockade. Br J Pharmacol. 1989;96:875–887. doi: 10.1111/j.1476-5381.1989.tb11898.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Jones RL, Wilson NH, Marr CG, Muir G, Armstrong RA. Diphenylmethylazine prostanoids with prostacyclin-like actions on human platelets. J Lipid Mediat. 1993;6:405–410. [PubMed] [Google Scholar]
  205. Jones RL, Qian YM, Wise H, Wong HN, Lam WL, Chan HW, et al. Relaxant actions of non-prostanoid prostacyclin mimetics on human pulmonary artery. J Cardiovasc Pharmacol. 1997;29:525–535. doi: 10.1097/00005344-199704000-00015. [DOI] [PubMed] [Google Scholar]
  206. Jones RL, Qian YM, Chan KM, Yip AP. Characterization of a prostanoid EP3-receptor in guinea-pig aorta: partial agonist action of the non-prostanoid ONO-AP-324. Br J Pharmacol. 1998;125:1288–1296. doi: 10.1038/sj.bjp.0702189. [DOI] [PMC free article] [PubMed] [Google Scholar]
  207. Jones RL, Wise H, Clark R, Whiting RL, Bley KR. Investigation of the prostacyclin (IP) receptor antagonist RO1138452 on isolated blood vessel and platelet preparations. Br J Pharmacol. 2006;149:110–120. doi: 10.1038/sj.bjp.0706841. [DOI] [PMC free article] [PubMed] [Google Scholar]
  208. Juteau H, Gareau Y, Labelle M, Sturino CF, Sawyer N, Tremblay N, et al. Structure-activity relationship of cinnamic acylsulfonamide analogues on the human EP3 prostanoid receptor. Bioorg Med Chem. 2001;9:1977–1984. doi: 10.1016/s0968-0896(01)00110-9. [DOI] [PubMed] [Google Scholar]
  209. Kabashima K, Saji T, Murata T, Nagamachi M, Matsuoka T, Segi E, et al. The prostaglandin receptor EP4 suppresses colitis, mucosal damage and CD4 cell activation in the gut. J Clin Invest. 2002;109:883–893. doi: 10.1172/JCI14459. [DOI] [PMC free article] [PubMed] [Google Scholar]
  210. Kabashima K, Sakata D, Nagamachi M, Miyachi Y, Inaba K, Narumiya S. Prostaglandin E2-EP4 signaling initiates skin immune responses by promoting migration and maturation of Langerhans cells. Nature Med. 2003;9:744–749. doi: 10.1038/nm872. [DOI] [PubMed] [Google Scholar]
  211. Kabashima K, Nagamachi M, Honda T, Nishigori C, Miyachi Y, Tokura Y. Prostaglandin E2 is required for ultraviolet B-induced skin inflammation via EP2 and EP4 receptors. Lab Invest. 2007;87:49–55. doi: 10.1038/labinvest.3700491. [DOI] [PubMed] [Google Scholar]
  212. Kam Y, Chow KB, Wise H. Factors affecting prostacyclin receptor agonist efficacy in different cell types. Cell Signal. 2001;13:841–847. doi: 10.1016/s0898-6568(01)00210-8. [DOI] [PubMed] [Google Scholar]
  213. Kamanna VS, Vo A, Kashyap ML. Nicotinic acid: recent developments. Curr Opin Cardiol. 2008;23:393–398. doi: 10.1097/HCO.0b013e3283021c82. [DOI] [PubMed] [Google Scholar]
  214. Kanamori Y, Niwa M, Kohno K, Al-Essa LY, Matsuno H, Kozawa O, et al. Migration of neutrophils from blood to tissue: alteration of modulatory effects of prostanoid on superoxide generation in rabbits and humans. Life Sci. 1997;60:1407–1417. doi: 10.1016/s0024-3205(97)00086-6. [DOI] [PubMed] [Google Scholar]
  215. Kamata S, Haga N, Tsuri T, Uchida K, Kakushi H, Arita H. Synthesis of thromboxane receptor antagonists with bicyclo[3.1.0]hexane ring systems. J Med Chem. 1990;33:229–239. doi: 10.1021/jm00163a038. [DOI] [PubMed] [Google Scholar]
  216. Karanam B, Madeira M, Bradley S, Wenning L, Desai R, Soli E, et al. Absorption, metabolism, and excretion of [14C]MK-0524, a prostaglandin D2 receptor antagonist, in humans. Drug Metab Dispos. 2007;35:1196–1202. doi: 10.1124/dmd.107.014696. [DOI] [PubMed] [Google Scholar]
  217. Karasawa A, Kawakage M, Shirakura S, Higo K, Kubo K, Ohshima E, et al. Antiplatelet effects of the novel thromboxane A2 receptor antagonist sodium (E)-11-[2-(5,6-dimethyl-1-benzimidazolyl)-ethylidene]-6,11-dihydrodibenz[b,e] oxepine-2-carboxylate monohydrate. Arzneimittelforschung. 1991a;41:1230–1236. [PubMed] [Google Scholar]
  218. Karasawa A, Shirakura S, Higo K, Kubo K. Actions of the novel thromboxane A2 receptor antagonist sodium (E)-11-[2-(5,6-dimethyl-1-benzimidazolyl)-ethylidene]-6,11-dihydrodibenz[b,e]oxepine-1-carboxylate monohydrate on smooth muscle preparations. Arzneimittelforschung. 1991b;41:1237–1241. [PubMed] [Google Scholar]
  219. Kassuya CA, Ferreira J, Claudino RF, Calixto JB. Intraplantar PGE2 causes nociceptive behaviour and mechanical allodynia: the role of prostanoid E receptors and protein kinases. Br J Pharmacol. 2007;150:727–737. doi: 10.1038/sj.bjp.0707149. [DOI] [PMC free article] [PubMed] [Google Scholar]
  220. Katsura M, Miyamoto T, Hamanaka N, Kondo K, Terada T, OHgaki Y, et al. In vitro and in vivo effects of new powerful thromboxane antagonists (3-alkylamino pinane derivatives) Adv Prost Thromb Leukot Res. 1983;11:351–357. [PubMed] [Google Scholar]
  221. Kawahara H, Sakamoto A, Takeda S, Onodera H, Imaki J, Ogawa R. A prostaglandin E2 receptor subtype EP1 receptor antagonist (ONO-8711) reduces hyperalgesia, allodynia, and c-fos gene expression in rats with chronic nerve constriction. Anaesth Analg. 2001;93:1012–1017. doi: 10.1097/00000539-200110000-00043. [DOI] [PubMed] [Google Scholar]
  222. Kawamori T, Uchiya N, Kitamura T. Evaluation of a selective prostaglandin E receptor EP1 antagonist for potential properties in colon carcinogenesis. Anticaner Res. 2001;21:3865–3869. [PubMed] [Google Scholar]
  223. Kawano T, Anrather J, Zhou P, Park L, Wang G, Frys KA, et al. Prostaglandin E2 EP1 receptors: downstream effectors of COX-2 neurotoxicity. Nat Med. 2006;12:225–229. doi: 10.1038/nm1362. [DOI] [PubMed] [Google Scholar]
  224. Kawikova I, Barnes PJ, Takahashi T, Tadjkarimi S, Yacoub MH, Belvisi MG. 8-Epi-PGF2α, a novel noncyclooxygenase-derived prostaglandin, constricts airways in vitro. Am J Respir Crit Care Med. 1996;153:590–596. doi: 10.1164/ajrccm.153.2.8564103. [DOI] [PubMed] [Google Scholar]
  225. Keery RJ, Lumley P. AH6809, a prostaglandin DP-receptor blocking drug on human platelets. Br J Pharmacol. 1988;94:745–754. doi: 10.1111/j.1476-5381.1988.tb11584.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  226. Keith JC, Spitz B, Van Assche FA. Thromboxane synthetase inhibition as a new therapy for preeclampsia: animal and human studies minireview. Prostaglandins. 1993;45:3–13. doi: 10.1016/0090-6980(93)90085-l. [DOI] [PubMed] [Google Scholar]
  227. Keitz PF, Jahangir A, Lopez-Tapia FJ, O'Yang C. Substituted arylamides as IP antagonists US 2004/0220247A1 Roche Palo Alto.
  228. Kemp BK, McPherson GA. Interspecies differences in thromboxane A2 receptors are distinguished by glibenclamide. Eur J Pharmacol. 1998;354:173–178. doi: 10.1016/s0014-2999(98)00450-6. [DOI] [PubMed] [Google Scholar]
  229. Khasar SG, Green PG, Levine PD. Comparison of intradermal and subcutaneous hyperalgesic effects of inflammatory mediators in the rat. Neurosci Lett. 1993;153:215–218. doi: 10.1016/0304-3940(93)90325-f. [DOI] [PubMed] [Google Scholar]
  230. Khera M, Boone TB, Salas N, Jett MF, Somogyi GT. The role of the prostacyclin receptor antagonist RO3244019 in treating neurogenic detrusor overactivity after spinal cord injury in rats. BJU Int. 2007;99:442–446. doi: 10.1111/j.1464-410X.2007.06615.x. [DOI] [PubMed] [Google Scholar]
  231. Kim N, Luster AD. Regulation of immune cells by eicosanoid receptors. Scientif World J. 2007;7:1307–1328. doi: 10.1100/tsw.2007.181. [DOI] [PMC free article] [PubMed] [Google Scholar]
  232. Kimura M, Osumi S, Ogihara M. Stimulation of DNA synthesis and proliferation by prostaglandins in primary cultures of adult rat hepatocytes. Eur J Pharmacol. 2000;404:259–271. doi: 10.1016/s0014-2999(00)00594-x. [DOI] [PubMed] [Google Scholar]
  233. Kimura M, Osumi S, Ogihara M. Prostaglandin E2 EP1 receptor agonist-induced DNA synthesis and proliferation in primary cultures of adult rat hepatocytes: the involvement of TGFα. Endocrinology. 2001;142:4428–4440. doi: 10.1210/endo.142.10.8450. [DOI] [PubMed] [Google Scholar]
  234. Kishino J, Hanasaki K, Nagasaki T, Arita H. Kinetic studies on stereospecific recognition by the thromboxane A2/prostaglandin H2 receptor of the antagonist, S-145. Br J Pharmacol. 1991;103:1883–1888. doi: 10.1111/j.1476-5381.1991.tb12346.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  235. Kitamura N, Hukuda R, Majima T, Horie T, Sugihara T. The 3 effective cases of cough variant asthma with ramatroban. Jpn J Allergol. 2003a;52:1089–1092. [PubMed] [Google Scholar]
  236. Kitamura T, Itoh M, Nada T, Tani K, Kobayashi M, Maruyama T, et al. Combined effects of prostaglandin E receptor subtype E1 and subtype EP4 antagonists on intestinal tumorigenesis in adenomatous polyposis coli gene knockout mice. Comcer Sci. 2003b;94:618–621. doi: 10.1111/j.1349-7006.2003.tb01492.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  237. Kitanaka J, Ishibashi T, Baba A. Phloretin as an antagonist of prostaglandin F2α receptor in cultured rat astrocytes. J Neurochem. 1993;60:704–708. doi: 10.1111/j.1471-4159.1993.tb03204.x. [DOI] [PubMed] [Google Scholar]
  238. Kitawaki T, Kadowaki N, Sugimoto N, Kambe N, Hori T, Miyachi Y, et al. IgE-activated mast cells in combination with pro-inflammatory factors induce Th2-promoting dendritic cells. Int Immunol. 2006;18:1789–1799. doi: 10.1093/intimm/dxl113. [DOI] [PubMed] [Google Scholar]
  239. Knapp HR, Heally C, Lawson J, FitzGerald GA. Effects of low-dose aspirin on endogenous eicosanoid formation in normal and atherosclerotic men. Thromb Res. 1988;50:377–386. doi: 10.1016/0049-3848(88)90267-8. [DOI] [PubMed] [Google Scholar]
  240. Koda N, Tsutsui Y, Niwa H, Ito S, Woodward DF, Watanabe K. Synthesis of prostaglandin F ethanolamide by prostaglandin F synthase and identification of bimatoprost as a potent inhibitor of the enzyme. New enzyme method by LC/GSI/MS. Arch Biochem Biophys. 2004;424:128–136. doi: 10.1016/j.abb.2004.02.009. [DOI] [PubMed] [Google Scholar]
  241. Kopp UC, Cicha MZ, Nakamura K, Nusing RM, Smith LA, Hokfelt T. Activation of EP4 receptors contributes to prostaglandin E2-mediated stimulation of renal sensory nerves. Am J Physiol Renal Physiol. 2004;287:F1269–F1282. doi: 10.1152/ajprenal.00230.2004. [DOI] [PubMed] [Google Scholar]
  242. Kostenis E, Ulven T. Emerging roles of DP and CRTH2 in allergic inflammation. Trends Mol Med. 2006;12:148–158. doi: 10.1016/j.molmed.2006.02.005. [DOI] [PubMed] [Google Scholar]
  243. Kozak KR, Crews BC, Morrow JD, Wang LH, Ma YH, Weinander R, et al. Metabolism of the endocannabinoids, 2-arachidonylglycerol and anandamide, into prostaglandin, thromboxane, and prostacyclin glycerol esters and ethanolamides. J Biol Chem. 2002;277:44877–44885. doi: 10.1074/jbc.M206788200. [DOI] [PubMed] [Google Scholar]
  244. Krauss AH, Woodward DF, Donde Y, Burk RM. Prostaglandin D2 antagonist. WO2005/079793, Allergan Inc.
  245. Kromer BM, Tippins JR. Coronary artery constriction by the isoprostane 8-epi prostaglandin F2α. Br J Pharmacol. 1996;119:1276–1280. doi: 10.1111/j.1476-5381.1996.tb16033.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  246. Kuhn CF, Feru F, Bazin M, Awad MMA, Goldstein SW. Quinoline derivatives as CRTH2 antagonists. EP1435356, Warner-Lambert & Co.
  247. Kumazawa T, Mizumura K, Koda H, Fukusako H. EP receptor subtypes implicated in the PGE2-induced sensitization of polymodal receptors in response to bradykinin and heat. J Neurophysiol. 1996;75:2361–2368. doi: 10.1152/jn.1996.75.6.2361. [DOI] [PubMed] [Google Scholar]
  248. Kunikata T, Tanaka A, Miyazawa T, Kato S, Takeuchi K. 16,16-dimethyl prostaglandin E2 inhibits indomethacin-induced small intestinal lesions through EP3 and EP4 receptors. Dig Dis Sci. 2002;47:894–904. doi: 10.1023/a:1014725024519. [DOI] [PubMed] [Google Scholar]
  249. Kuwano T, Nakao S, Yamamoto H, Tsuneyoshi M, Yamamoto T, Kuwano M, et al. Cyclooxygenase 2 is a key enzyme for inflammatory cytokine-induced angiogenesis. FASEB J. 2004;18:300–310. doi: 10.1096/fj.03-0473com. [DOI] [PubMed] [Google Scholar]
  250. Lai E, De Lepeleire I, Crumley TM, Liu F, Wenning LA, Michiels N, et al. Suppression of niacin-induced vasodilation with an antagonist to prostaglandin D2 receptor subtype 1. Clin Pharmacol Ther. 2007;81:849–857. doi: 10.1038/sj.clpt.6100180. [DOI] [PubMed] [Google Scholar]
  251. Lai E, Wenning LA, Crumley TM, De LI, Liu F, de Hoon JN, et al. Pharmacokinetics, pharmacodynamics, and safety of a prostaglandin D2 receptor antagonist. Clin Pharmacol Ther. 2008b;83:840–847. doi: 10.1038/sj.clpt.6100345. [DOI] [PubMed] [Google Scholar]
  252. Lai YJ, Pullamsetti SS, Dony E, Weissmann N, Butrous G, Banat GA, et al. Role of prostanoid EP4 receptor in iloprost-mediated vasodilatation in pulmonary hypertension. Am J Respir Crit Care Med. 2008a;178:188–196. doi: 10.1164/rccm.200710-1519OC. [DOI] [PubMed] [Google Scholar]
  253. Langevin B, Orton E, Sherer D. Dihydrogen phosphate salt of a prostaglandin D2 receptor antagonist. WO2007/047378, Sanofi-Aventis.
  254. Lawrence RA, Jones RL. Investigation of the prostaglandin E (EP-) receptor subtype mediating relaxation of the rabbit jugular vein. Br J Pharmacol. 1992;105:817–824. doi: 10.1111/j.1476-5381.1992.tb09063.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  255. Lawrence RA, Jones RL, Wilson NH. Characterization of receptors involved in the direct and indirect actions of prostaglandins E and I on the guinea-pig ileum. Br J Pharmacol. 1992;105:271–278. doi: 10.1111/j.1476-5381.1992.tb14245.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  256. Lazareno S, Birdsall NJ. Estimation of competitive antagonist affinity from functional inhibition curves using the Gaddum, Schild and Cheng-Prusoff equations. Br J Pharmacol. 1993;109:1110–1119. doi: 10.1111/j.1476-5381.1993.tb13737.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  257. Leff P, Dougall IG. Further concerns over Cheng-Prusoff analysis. Trends Pharmacol Sci. 1993;14:110–112. doi: 10.1016/0165-6147(93)90080-4. [DOI] [PubMed] [Google Scholar]
  258. Leff P, Giles H. Classification of platelet and vascular prostaglandin D2 (DP) receptors: estimation of affinities and relative efficacies for a series of novel bicyclic ligands. With an appendix on goodness-of-fit analyses. Br J Pharmacol. 1992;106:996–1003. doi: 10.1111/j.1476-5381.1992.tb14447.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  259. Leonhardt A, Glaser A, Wegmann M, Schranz D, Seyberth H, Nüsing R. Expression of prostanoid receptors in human ductus arteriosus. Br J Pharmacol. 2003;138:655–659. doi: 10.1038/sj.bjp.0705092. [DOI] [PMC free article] [PubMed] [Google Scholar]
  260. Lewis RA, Soter NA, Diamond PT, Austen KF, Oates JA, Roberts LJ. Prostaglandin D2 generation after activation of rat and human mast cells with anti-IgE. J Immunol. 1982;129:1627–1631. [PubMed] [Google Scholar]
  261. Liang Y, Li C, Guzman VM, Evinger AJ, 3rd, Protzman CE, Krauss AH, Woodward DF. Comparison of prostaglandin F2α, bimatoprost (prostamide), and butaprost (EP2 agonist) on Cyr61 and connective tissue growth factor gene expression. J Biol Chem. 2003;278:27267–27277. doi: 10.1074/jbc.M301009200. [DOI] [PubMed] [Google Scholar]
  262. Liang Y, Li C, Guzman VM, Chang WW, Evinger AJ, Pablo JV, Woodward DF. Upregulation of orphan nuclear receptor Nur77 following PGF2α, Bimatoprost, and Butaprost treatments. Essential role of a protein kinase C pathway involved in EP2 receptor activated Nur77 gene transcription. Br J Pharmacol. 2004;142:737–748. doi: 10.1038/sj.bjp.0705829. [DOI] [PMC free article] [PubMed] [Google Scholar]
  263. Liang Y, Woodward DF, Guzman VM, Li C, Scott DF, Wang JW, et al. Identification and pharmacological characterization of the prostaglandin FP receptor and FP receptor variant complexes. Br J Pharmacol. 2008;154:1079–1093. doi: 10.1038/bjp.2008.142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  264. Lin CR, Amaya F, Barrett L, Wang H, Takada J, Samad TA, et al. Prostaglandin E2 receptor EP4 contributes to inflammatory pain hypersensitivity. J Pharmacol Exp Ther. 2006;319:1096–1103. doi: 10.1124/jpet.106.105569. [DOI] [PubMed] [Google Scholar]
  265. Liu YJ, Jackson DM, Blackham A. Effects of BW A868C, a selective prostaglandin DP receptor antagonist, in dog isolated vascular preparations. Eur J Pharmacol. 1996a;303:187–1192. doi: 10.1016/0014-2999(96)00037-4. [DOI] [PubMed] [Google Scholar]
  266. Liu YJ, Jackson DM, Blackham A, Leff P. Partial agonist effects of BW A868C, a selective DP receptor antagonist, on Cl- secretion in dog tracheal epithelium. Eur J Pharmacol. 1996b;304:117–122. doi: 10.1016/0014-2999(96)00129-x. [DOI] [PubMed] [Google Scholar]
  267. Lovell JM. Salts with CRTh2 antagonist activity. WO2007/107772, Oxagen Ltd.
  268. Lumley P, White BP, Humphrey PPA. GR32191, a highly potent and specific thromboxane A2 receptor blocking drug on platelets and vascular and airways smooth muscle in vitro. Br J Pharmacol. 1989;97:783–794. doi: 10.1111/j.1476-5381.1989.tb12017.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  269. Lydford SJ, McKechnie KC. Characterization of the prostaglandin E2 sensitive (EP)-receptor in the rat isolated trachea. Br J Pharmacol. 1994;112:133–136. doi: 10.1111/j.1476-5381.1994.tb13042.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  270. Lydford SJ, Li SW, McKechnie KC. Comparison of prostanoid DP receptors in the rabbit saphenous vein and human neutrophil. Br J Pharmacol. 1996a;117:190P. doi: 10.1111/j.1476-5381.1996.tb15148.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  271. Lydford SJ, McKechnie KC, Dougall IG. Pharmacological studies on prostanoid receptors in the rabbit isolated saphenous vein: a comparison with the rabbit isolated ear artery. Br J Pharmacol. 1996b;117:13–20. doi: 10.1111/j.1476-5381.1996.tb15148.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  272. Lydford SJ, McKechnie KC, Leff P. Interaction of BW A868C, a prostanoid DP-receptor antagonist, with two receptor subtypes in the rabbit isolated saphenous vein. Prostaglandins. 1996c;52:125–139. doi: 10.1016/0090-6980(96)00058-5. [DOI] [PubMed] [Google Scholar]
  273. Ma H, Hara A, Xiao CY, Okada Y, Takahata O, Nakaya K, et al. Increased bleeding tendency and decreased susceptibility to thromboembolism in mice lacking the prostaglandin E receptor subtype EP3. Circulation. 2001;104:1176–1180. doi: 10.1161/hc3601.094003. [DOI] [PubMed] [Google Scholar]
  274. Maassen VanDenBrink A, Bax WA, Ferrari MD, Zijlstra FJ, Bos E, et al. Augmented contraction of the human isolated coronary artery by sumatriptan: a possible role for endogenous thromboxane. Br J Pharmacol. 1996;119:855–862. doi: 10.1111/j.1476-5381.1996.tb15751.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  275. Machwate M, Harada S, Leu CT, Seedor G, Labelle M, Gallant M, et al. Prostaglandin receptor EP4 mediates the bone anabolic effects of PGE2. Mol Pharmacol. 2001;60:36–41. doi: 10.1124/mol.60.1.36. [DOI] [PubMed] [Google Scholar]
  276. Macias-Perez IM, Zent R, Carmosino M, Breyer MD, Breyer RM, Pozzi A. Mouse EP3α, β and γ receptor variants reduce tumor cell proliferation and tumorigenesis in vivo. J Biol Chem. 2008;283:12538–12545. doi: 10.1074/jbc.M800105200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  277. MacIntyre DE, Gordon JL. Discrimination between platelet prostaglandin receptors with a specific antagonist of bisenoic prostaglandins. Thromb Res. 1977;11:705–713. doi: 10.1016/0049-3848(77)90099-8. [DOI] [PubMed] [Google Scholar]
  278. McKenniff MG, Norman P, Cuthbert NJ, Gardiner PJ. BAY u3405, a potent and selective thromboxane A2 receptor antagonist on airway smooth muscle in vitro. Br J Pharmacol. 1991;104:585–590. doi: 10.1111/j.1476-5381.1991.tb12473.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  279. McKenniff M, Rodger IW, Norman P, Gardiner PJ. Characterisation of receptors mediating the contractile effects of prostanoids in guinea-pig and human airways. Eur J Pharmacol. 1988;153:149–159. doi: 10.1016/0014-2999(88)90601-2. [DOI] [PubMed] [Google Scholar]
  280. Maddox YT, Ramwell PW, Shiner CS, Corey EJ. Amide and 1-amino derivatives of F prostaglandins as prostaglandin antagonists. Nature. 1978;273:549–552. doi: 10.1038/273549a0. [DOI] [PubMed] [Google Scholar]
  281. Mais DE, Saussy DLJ, Chaikhouni A, Kochel PJ, Knapp DR, Hamanaka N, et al. Pharmacologic characterization of human and canine thromboxane A2/prostaglandin H2 receptors in platelets and blood vessels: evidence for different receptors. J Pharmacol Exp Ther. 1985;233:418–424. [PubMed] [Google Scholar]
  282. Mais DE, DeHoll D, Sightler H, Halushka PV. Different pharmacologic activities for 13-azapinane thromboxane A2 analogs in platelets and blood vessels. Eur J Pharmacol. 1988;148:309–315. doi: 10.1016/0014-2999(88)90108-2. [DOI] [PubMed] [Google Scholar]
  283. Majima M, Amano H, Hayashi I. Prostanoid receptor signaling relevant to tumor growth and angiogenesis. Trends Pharmacol Sci. 2003;24:524–529. doi: 10.1016/j.tips.2003.08.005. [DOI] [PubMed] [Google Scholar]
  284. Makino H, Tanaka I, Mukoyama M, Sugawara A, Mori K, Muro S, et al. Prevention of diabetic nephropathy in rats by prostaglandin E receptor EP1-selective antagonist. J Am Soc Nephrol. 2002;13:1757–1765. doi: 10.1097/01.asn.0000019782.37851.bf. [DOI] [PubMed] [Google Scholar]
  285. Malini PL, Strocchi E, Zanardi M, Milani M, Ambrosioni E. Thromboxane antagonism and cough induced by angiotensin-converting-enzyme inhibitor. Lancet. 1997;350:15–18. doi: 10.1016/S0140-6736(96)12045-6. [DOI] [PubMed] [Google Scholar]
  286. Malmberg AB, Rafferty MF, Yaksh TL. Antinociceptive effect of spinally delivered prostaglandin E receptor antagonists in the formalin test on the rat. Neurosci Letts. 1994;173:193–196. doi: 10.1016/0304-3940(94)90181-3. [DOI] [PubMed] [Google Scholar]
  287. Masataka M, Amano H, Izumi H. Prostanoid receptor signaling relevant to tumor growth and angiogenesis. Trends Pharmacol Sci. 2003;24:524–529. doi: 10.1016/j.tips.2003.08.005. [DOI] [PubMed] [Google Scholar]
  288. Masuda A, Mais DE, Oatis JEJ, Halushka PV. Platelet and vascular thromboxane A2/prostaglandin H2 receptors. Evidence for different subclasses in the rat. Biochem Pharmacol. 1991;42:537–544. doi: 10.1016/0006-2952(91)90316-w. [DOI] [PubMed] [Google Scholar]
  289. Matias I, Chen J, De Petrocellis L, Bisogno T, Ligresti A, Fezza F, et al. Prostaglandin ethanolamides (prostamides): in vitro pharmacology and metabolism. J Pharmacol Exp Ther. 2004;309:745–757. doi: 10.1124/jpet.103.061705. [DOI] [PubMed] [Google Scholar]
  290. Mathiesen JM, Christopoulos A, Ulven T, Royer JF, Campillo M, Heinemann A, et al. On the mechanism of interaction of potent surmountable and insurmountable antagonists with the prostaglandin D2 receptor CRTH2. Mol Pharmacol. 2006;69:1441–1453. doi: 10.1124/mol.105.017681. [DOI] [PubMed] [Google Scholar]
  291. Matsuoka T, Hirata M, Tanaka H, Takahashi Y, Murata T, Kabashima K, et al. Prostaglandin D2 as a mediator of allergic asthma. Science. 2000;287:2013–2017. doi: 10.1126/science.287.5460.2013. [DOI] [PubMed] [Google Scholar]
  292. Matthews JS, Jones RL. Potentiation of aggregation and inhibition of adenylate cyclase in human platelets by prostaglandin E analogues. Br J Pharmacol. 1993;108:363–369. doi: 10.1111/j.1476-5381.1993.tb12810.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  293. Maubach KA, Clark DE, Fenton G, Lockey PM, Clark KL, Oxford AW, et al. BCG20-1531, a novel, potent and selective prostanboid EP4 receptor antagonist; a putative new treatment for migraine headache. Br J Pharmacol. 2009;156:316–327. doi: 10.1111/j.1476-5381.2009.00027.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  294. Meanwell NA, Romine JL, Seiler SM. Non-prostanoid prostacyclin mimetics. Drugs Future. 1994;19:361–385. [Google Scholar]
  295. Mebane H, Turnbach ME, Randich A. Spinal EP receptors mediating prostaglandin E2-induced mechanical hyperalgesia, thermal hyperalgesia, and touch-evoked allodynia in rats. J Pain. 2003;4:392–399. doi: 10.1016/s1526-5900(03)00715-6. [DOI] [PubMed] [Google Scholar]
  296. Merritt JE, Brown AM, Bund S, Cooper DG, Egan JW, Hallam TJ, et al. Primate vascular responses to octimibate, a non-prostanoid agonist at the prostacyclin receptor. Br J Pharmacol. 1991a;102:260–266. doi: 10.1111/j.1476-5381.1991.tb12163.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  297. Merritt JE, Hallam TJ, Brown AM, Boyfield I, Cooper DG, Hickey DM, et al. Octimibate, a potent non-prostanoid inhibitor of platelet aggregation, acts via the prostacyclin receptor. Br J Pharmacol. 1991b;102:251–259. doi: 10.1111/j.1476-5381.1991.tb12162.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  298. Middlemiss D, Ashton MR, Boyd AE, Brookfield FA. Compounds with PGD2 antagonist activity. WO2005/040112, Oxagen Ltd.
  299. Middlemiss D, Ashton MR, Boyd AE, Brookfield FA. Substituted indol-3-yl acetic acid derivatives. GB 2 407 318, Oxagen Ltd.
  300. Middlemiss D, Ashton MR, Boyd AE, Brookfield FA. Use of CRTh2 antagonist compounds in therapy. WO2005/044260, Oxagen Ltd.
  301. Middlemiss D, Ashton MR, Boyd AE, Brookfield FA, Armer RE. Compounds having CRTh2 antagonist activity. WO2005/040114, Oxagen Ltd.
  302. Mihara S, Hara S, Ueda M, Ide M, Fujimoto M. Antagonistic actions of S-145 on vascular and platelet thromboxane A2 receptors. Eur J Pharmacol. 1989;171:179–187. doi: 10.1016/0014-2999(89)90106-4. [DOI] [PubMed] [Google Scholar]
  303. Miki I, Kishibayashi N, Nonaka H, Ohshima E, Takami H, Obase H, et al. Effects of KW-3635, a novel dibenzoxepin derivative of a selective thromboxane A2 antagonist, on human, guinea pig and rat platelets. Jpn J Pharmacol. 1992;59:357–364. doi: 10.1254/jjp.59.357. [DOI] [PubMed] [Google Scholar]
  304. Mimura H, Ikemura T, Kotera O, Sawada M, Tashiro S, Fuse E, et al. Inhibitory effect of the 4-aminotetrahydroquinoline derivatives, selective chemoattractant receptor-homologous molecule expressed on T helper 2 cell antagonists, on eosinophil migration induced by prostaglandin D2. J Pharmacol Exp Ther. 2005;314:244–251. doi: 10.1124/jpet.104.081539. [DOI] [PubMed] [Google Scholar]
  305. Minami T, Nishihara I, Sakamoto K, Ito S, Hyodo M, Hayashi O. Blockade by ONO-NT-012, a unique prostanoid analogue, of prostaglandin E2-induced allodynia in conscious mice. Br J Pharmacol. 1995;115:73–76. doi: 10.1111/j.1476-5381.1995.tb16321.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  306. Minami T, Nakano H, Kobayashi T, Sugimoto Y, Ushikubi F, Ichikawa A, et al. Characterization of EP receptor subtypes responsible for prostaglandin E2-induced pain responses by use of EP1 and EP3 receptor knockout mice. Br J Pharmacol. 2001;133:438–444. doi: 10.1038/sj.bjp.0704092. [DOI] [PMC free article] [PubMed] [Google Scholar]
  307. Minami T, Matsumura S, Mabuchi T, Kobayashi T, Sugimoto Y, Ushikubi F, et al. Functional evidence for interaction between prostaglandin EP3 and kappa-opioid receptor pathways in tactile pain induced by human immunodeficiency virus type-1 (HIV-1) glycoprotein gp120. Neuropharmacology. 2003;45:96–105. doi: 10.1016/s0028-3908(03)00133-3. [DOI] [PubMed] [Google Scholar]
  308. Misra RA. Recent progress in the clinical development of thromboxane A2 receptor antagonists. Exp Opin Invest Drugs. 1994;3:469–480. [Google Scholar]
  309. Mitsumori S, Tsuri T, Honma T, Hiramatsu Y, Okada T, Hashizume H, et al. Synthesis and biological activity of various derivatives of a novel class of potent, selective, and orally active prostaglandin D2 receptor antagonists. 1. Bicyclo[2.2.1]heptane derivatives. J Med Chem. 2003a;46:2436–2445. doi: 10.1021/jm020517g. [DOI] [PubMed] [Google Scholar]
  310. Mitsumori S, Tsuri T, Honma T, Hiramatsu Y, Okada T, Hashizume H, et al. Synthesis and biological activity of various derivatives of a novel class of potent, selective, and orally active prostaglandin D2 receptor antagonists. 1. 6,6-Dimethylbicyclo[3.1.1]heptane derivatives. J Med Chem. 2003b;46:2446–2455. doi: 10.1021/jm0205189. [DOI] [PubMed] [Google Scholar]
  311. Momma K, Toyoshima K, Takeuchi D, Imamura S, Nakanishi T. In vivo constriction of the fetal and neonatal ductus arteriosus by a prostanoid EP4-receptor antagonist. Pediatr Res. 2005;58:971–975. doi: 10.1203/01.pdr.0000182182.49476.24. [DOI] [PubMed] [Google Scholar]
  312. Monneret G, Gravel S, Diamond M, Rokach J, Powell WS. Prostaglandin D2 is a potent chemoattractant for human eosinophils that acts via a novel DP receptor. Blood. 2001;98:1942–1948. doi: 10.1182/blood.v98.6.1942. [DOI] [PubMed] [Google Scholar]
  313. Morinelli TA, Okwu AK, Mais DE, Halushka PV, John V, Chen CK, et al. Difluorothromboxane A2 and stereoisomers: stable derivatives of thromboxane A2 with differential effects on platelets and blood vessels. Proc Natl Acad Sci USA. 1989;86:5600–5604. doi: 10.1073/pnas.86.14.5600. [DOI] [PMC free article] [PubMed] [Google Scholar]
  314. Moriuchi H, Koda N, Okuda-Ashitaka E, Daiyasu H, Ogasarawa K, Toh H, et al. Molecular characterization of a novel type of prostamide/prostaglandin F synthase, belonging to the thioredoxin-like superfamily. J Biol Chem. 2008;283:792–801. doi: 10.1074/jbc.M705638200. [DOI] [PubMed] [Google Scholar]
  315. Morrow DJ, Hill KE, Burk RF, Nammour TM, Badr KF, Roberts LJ. A series of prostaglandin F2-like compounds are produced in vivo in humans by a non-cyclooxygenase, free radical-catalysed mechanism. Proc Nat Acad Sci USA. 1990;87:9383–9387. doi: 10.1073/pnas.87.23.9383. [DOI] [PMC free article] [PubMed] [Google Scholar]
  316. Morrow JD, Minton TA, Mukundan CR, Campbell MD, Zackert WE, Daniel VC, et al. Free radical-induced generation of isoprostanes in vivo. Evidence for the formation of D-ring and E-ring isoprostanes. J Biol Chem. 1994;269:4317–4326. [PubMed] [Google Scholar]
  317. Müller B, Schneider J, Wilsmann K, Lintz W, Flohé L. Role of renin release in the hemodynamic, renal and dipsogenic actions of the prostacyclin analogue CG 4203 in conscious rats. Prostaglandins Leukot Med. 1983;11:361–372. doi: 10.1016/0262-1746(83)90088-4. [DOI] [PubMed] [Google Scholar]
  318. Murase A, Okumura T, Sakakibara A, Tonai-Kachi H, Nakao S, Takada J. Effect of prostanoid EP4 receptor antagonist, CJ 042,794 in rat models of pain and inflammation. Eur J Pharmacol. 2008a;580:116–121. doi: 10.1016/j.ejphar.2007.10.054. [DOI] [PubMed] [Google Scholar]
  319. Murase A, Taniguchi Y, Tonai-Kachi H, Nakao S, Takada J. In vitro pharmacological characterization of CJ-042794, a novel, potent, and selective prostaglandin EP4 receptor antagonist. Life Sci. 2008b;82:226–232. doi: 10.1016/j.lfs.2007.11.002. [DOI] [PubMed] [Google Scholar]
  320. Murata T, Ushikubi F, Matsuoka T, Hirata M, Yamasaki A, Sugimoto Y, et al. Altered pain perception and inflammatory response in mice lacking prostacyclin receptor. Nature. 1997;388:678–682. doi: 10.1038/41780. [DOI] [PubMed] [Google Scholar]
  321. Murray JJ, Tonnel AB, Brash AR, Roberts LJ, Gosset P, Workman R, et al. Release of prostaglandin D2 into human airways during acute antigen challenge. N Engl J Med. 1986;315:800–804. doi: 10.1056/NEJM198609253151304. [DOI] [PubMed] [Google Scholar]
  322. Mutoh M, Watanabe K, Kitamura T, Shoji Y, Takahashi M, Kawamori T, et al. Involvement of prostaglandin E receptor subtype EP4 in colon carcinogenesis. Cancer Res. 2002;62:28–32. [PubMed] [Google Scholar]
  323. Mutoh M, Takahashi M, Wakabayashi K. Roles of prostanoids in colon carcinogenesis and their potential targeting for cancer chemoprevention. Curr Pharm Des. 2006;12:2375–2382. doi: 10.2174/138161206777698972. [DOI] [PubMed] [Google Scholar]
  324. Naclerio RM, Meier HL, Kagey-Sobotka A, Adkinson NF, Jr, Norman DA, Meyers PS, et al. Mediator release after nasal airway challenge with allergen. Am Rev Respir Dis. 1983;128:597–602. doi: 10.1164/arrd.1983.128.4.597. [DOI] [PubMed] [Google Scholar]
  325. Nagai H, Takeda H, Yamaguchi S, Tanaka H, Matsuo A, Inagaki N. The effect of a thromboxane A2 receptor antagonist BAY-u3405 on experimental allergic reactions. Prostaglandins. 1995;50:75–87. doi: 10.1016/0090-6980(95)00111-5. [DOI] [PubMed] [Google Scholar]
  326. Naganawa A, Matsui T, Ima M, Yoshida K, Tsuruta H, Yamamoto S, et al. Optimization of sulfonamide derivatives as highly selective EP1 receptor antagonists. Bioorg Med Chem. 2006;14:7774–7789. doi: 10.1016/j.bmc.2006.08.001. [DOI] [PubMed] [Google Scholar]
  327. Nagata K, Hirai H. The second PGD2 receptor CRTH2: structure, properties, and functions in leukocytes. Prostaglandins Leukot Essent Fatty Acids. 2003;69:169–177. doi: 10.1016/s0952-3278(03)00078-4. [DOI] [PubMed] [Google Scholar]
  328. Nagata K, Hirai H, Tanaka K, Ogawa K, Aso T, Sugamura K, et al. CRTH2, an orphan receptor of T-helper-2-cells, is expressed on basophils and eosinophils and responds to mast cell-derived factor(s) FEBS Lett. 1999a;459:195–199. doi: 10.1016/s0014-5793(99)01251-x. [DOI] [PubMed] [Google Scholar]
  329. Nagata K, Tanaka K, Ogawa K, Kemmotsu K, Imai T, Yoshie O, et al. Selective expression of a novel surface molecule by human Th2 cells in vivo. J Immunol. 1999b;162:1278–1286. [PubMed] [Google Scholar]
  330. Naka M, Mais DE, Morinelli TA, Hamanaka N, Oatis JE, Jr, Halushka PV. 7-[(1R,2S,3S,5R)-6,6-dimethyl-3-(4-iodobenzenesulfonylamino)bicyclo[3.1.1]hept-2-4yl]-5(Z)-heptenoic acid: a novel high-affinity radiolabeled antagonist for platelet thromboxane A2/prostaglandin H2 receptors. J Pharmacol Exp Ther. 1992;262:632–637. [PubMed] [Google Scholar]
  331. Nakae K, Saito K, Iino T, Yamamoto N, Wakabayashi M, Yoshikawa S, et al. A prostacyclin receptor antagonist inhibits the sensitized release of substance P from rat sensory neurones. J Pharmacol Exp Ther. 2005;315:1136–1142. doi: 10.1124/jpet.105.091967. [DOI] [PubMed] [Google Scholar]
  332. Nakahata N, Sato K, Abe MT, Nakanishi H. ONO NT-126 is a potent and selective thromboxane A2 antagonist in human astrocytoma cells. Eur J Pharmacol. 1990;184:233–238. doi: 10.1016/0014-2999(90)90614-c. [DOI] [PubMed] [Google Scholar]
  333. Nakao K, Murase A, Ohshiro H, Okumura T, Taniguchi K, Murata Y, et al. CJ-023, 423, a novel, potent and selective prostaglandin EP4 receptor antagonist with anti-hyperalgesic properties. J Pharm Exp Ther. 2007;322:686–694. doi: 10.1124/jpet.107.122010. [DOI] [PubMed] [Google Scholar]
  334. Nakayama Y, Omote K, Namiki A. Role of prostaglandin receptor EP1 in the spinal dorsal horn in carrageenan-induced inflammatory pain. Anesthesiology. 2002;97:1254–1262. doi: 10.1097/00000542-200211000-00032. [DOI] [PubMed] [Google Scholar]
  335. Nakayama Y, Omote K, Kawamata T, Namiki A. Role of prostaglandin receptor subtype EP1 in prostaglandin E2-induced nociceptive transmission in the rat spinal dorsal horn. Brain Res. 2004;1010:62–68. doi: 10.1016/j.brainres.2004.03.002. [DOI] [PubMed] [Google Scholar]
  336. Namba T, Oida H, Sugimoto Y, Kakizuka A, Negishi M, Ichikawa A, et al. cDNA Cloning of a mouse prostacyclin receptor. J Biol Chem. 1994;269:9986–9992. [PubMed] [Google Scholar]
  337. Narisada M, Ohtani M, Watanabe F, Uchida K, Arita H, Doteuchi M, et al. Synthesis and in vitro activity of various derivatives of a novel thromboxane receptor antagonist, (±)-(5Z)-7-[3-endo-[(phenylsulfonyl)amino]bicyclo[2.2.1]hept-2-exo-yl]heptenoic acid. J Med Chem. 1988;31:1847–1854. doi: 10.1021/jm00117a028. [DOI] [PubMed] [Google Scholar]
  338. Narumiya S. Prostanoids in immunity: roles revealed by mice deficient in their receptors. Life Sci. 2003;74:391–395. doi: 10.1016/j.lfs.2003.09.025. [DOI] [PubMed] [Google Scholar]
  339. Narumiya S, Fitzgerald GA. Genetic and pharmacological analysis of prostanoid receptor function. J Clin Invest. 2001;108:25–30. doi: 10.1172/JCI13455. [DOI] [PMC free article] [PubMed] [Google Scholar]
  340. Narumiya S, Toda N. Different responsiveness of prostaglandin D2-sensitive systems to prostaglandin D2 and its analogues. Br J Pharmacol. 1985;85:367–375. doi: 10.1111/j.1476-5381.1985.tb08870.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  341. Narumiya S, Okuma S, Ushikubi F. Binding of a radioiodinated 13-azapinane thromboxane antagonist to platelets: correlation with antiaggregatory activity in different species. Br J Pharmacol. 1986;88:323–331. doi: 10.1111/j.1476-5381.1986.tb10208.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  342. Narumiya S, Sugimoto Y, Ushikubi B. Prostanoid receptors: structure, properties and functions. Physiol Rev. 1999;79:1193–1226. doi: 10.1152/physrev.1999.79.4.1193. [DOI] [PubMed] [Google Scholar]
  343. Nasrallah R, Hébert RL. Prostaglandin signaling in the kidney: implications for health and disease. Am J Physiol Renal Physiol. 2005;289:F235–F246. doi: 10.1152/ajprenal.00454.2004. [DOI] [PubMed] [Google Scholar]
  344. Nataraj C, Thomas DW, Tilley SL, Nguyen MT, Mannon R, Koller BH, et al. Receptors for prostaglandin E2 that regulate cellular immune responses in the mouse. J Clin Invest. 2001;108:1229–1235. doi: 10.1172/JCI13640. [DOI] [PMC free article] [PubMed] [Google Scholar]
  345. Nicolaou KC, Magolda RL, Smith JB, Aharony D, Smith EF, Lefer AM. Synthesis and biological properties of pinane-thromboxane A2, a selective inhibitor of coronary artery constriction, platelet aggregation, and thromboxane formation. Proc Natl Acad Sci USA. 1979;76:2566–2570. doi: 10.1073/pnas.76.6.2566. [DOI] [PMC free article] [PubMed] [Google Scholar]
  346. Nishigaki N, Negishi M, Honda A, Sugimoto Y, Namba T, Narumiya S, et al. Identification of prostaglandin E receptor ‘EP2’ cloned from mouse mastocytoma cells as EP4. FEBS Lett. 1995;364:339–341. doi: 10.1016/0014-5793(95)00421-5. [DOI] [PubMed] [Google Scholar]
  347. Nigam S, Zakrzewicz A, Eskafi S, Roscher A. Clinical significance of prostacyclin and thromboxane in cancer of the female breast and genital tract. Cancer Metastasis Rev. 1992;11:411–420. doi: 10.1007/BF01307190. [DOI] [PubMed] [Google Scholar]
  348. Noguchi K, Iwasaki K, Shitashiqe M, Murato S, Ishikawa I. Prostaglandin E2 downregulates interferon-gamma-induced intercellular adhesion molecule-1 expression via EP2 receptors in human gingival fibroblasts. Inflammation. 1999;23:481–493. doi: 10.1023/a:1021921211559. [DOI] [PubMed] [Google Scholar]
  349. Norel X, Walch L, Labat C, Gascard JP, Dulmet E, Brink C. Prostanoid receptors involved in the relaxation of human bronchial preparations. Br J Pharmacol. 1999;126:867–872. doi: 10.1038/sj.bjp.0702392. [DOI] [PMC free article] [PubMed] [Google Scholar]
  350. Norel X, De Montpreville V, Brink C. Vasoconstriction induced by activation of EP1 and EP3 receptors in human lung: effects of ONO-AE-248, ONO-DI-004, ONO-8711 or ONO-8713. Prostaglandins Other Lipid Mediat. 2004;74:101–112. doi: 10.1016/j.prostaglandins.2004.07.003. [DOI] [PubMed] [Google Scholar]
  351. Norman P, Cuthbert NJ, McKenniff MG, Gardiner PJ. The thromboxane receptors of rat and guinea-pig lung. Eur J Pharmacol. 1992;229:171–178. doi: 10.1016/0014-2999(92)90552-f. [DOI] [PubMed] [Google Scholar]
  352. O'Connell M, Zeller W, Burgeson J, Mishra RK, Ramirez J, Kiselyov AS, et al. Peri-substituted hexahydro-indolones as novel, potent and selective human EP3 receptor antagonists. Bioorg Med Chem Lett. 2009;19:778–782. doi: 10.1016/j.bmcl.2008.12.027. [DOI] [PubMed] [Google Scholar]
  353. Ogletree ML, Allen GT. Interspecies differences in thromboxane receptors: studies with thromboxane receptor antagonists in rat and guinea pig smooth muscles. J Pharmacol Exp Ther. 1992;260:789–794. [PubMed] [Google Scholar]
  354. Ogletree ML, Harris DN, Greenberg R, Haslanger MF, Nakane M. Pharmacological actions of SQ 29,548, a novel selective thromboxane antagonist. J Pharmacol Exp Ther. 1985;234:435–441. [PubMed] [Google Scholar]
  355. Ogletree ML, Harris DN, Schumacher WA, Webb ML, Mistra RN. Pharmacological profile of BMS 180,291: a potent, long-acting, orally active thromboxane A2/prostaglandin endoperoxide receptor antagonist. J Pharmacol Exp Ther. 1993;264:570–578. [PubMed] [Google Scholar]
  356. Ohnishi A, Shimamoto C, Katsu K, Ito S, Imai Y, Nakahari T. EP1 and EP4 receptors mediate exocytosis evoked by prostaglandin E2 in guinea-pig antral mucous cells. Exp Physiol. 2001;86:451–460. doi: 10.1113/eph8602160. [DOI] [PubMed] [Google Scholar]
  357. Oka T, Aou S, Hori T. Intracerebroventricular injection of prostaglandin E2 induces thermal hyperalgesia in rats: the possible involvement of EP3 receptors. Brian Res. 1994;663:287–292. doi: 10.1016/0006-8993(94)91275-0. [DOI] [PubMed] [Google Scholar]
  358. Okumura T, Murata Y, Taniguchi K, Murase A, Nii A. Effects of the selective EP4 antagonist, CJ-023,423 on chronic inflammation and bone destruction in rat adjuvant-induced arthritis. J Pharm Pharmacol. 2008;60:723–730. doi: 10.1211/jpp.60.6.0007. [DOI] [PubMed] [Google Scholar]
  359. Oliva P, Berrino L, De Novellis V, Palazzo E, Marabesse I, Siniscalco D, et al. Role of periaqueductal grey prostaglandin receptors in formalin-induced hyperalgesia. Eur J Pharmacol. 2006;530:40–47. doi: 10.1016/j.ejphar.2005.11.025. [DOI] [PubMed] [Google Scholar]
  360. Olson DM. The promise of prostaglandins: have they fulfilled their potential as therapeutic targets for the delay of preterm birth? J Soc Gynecol Investig. 2005;12:466–478. doi: 10.1016/j.jsgi.2005.06.004. [DOI] [PubMed] [Google Scholar]
  361. Omote K, Kawamata T, Nakayama Y, Kawamata M, Hazama K, Namiki A. The effects of peripheral administration of a novel selective antagonist for prostaglandin E receptor subtype EP1, ONO-8711, in a rat model of postoperative pain. Anesth Analg. 2001;92:233–238. doi: 10.1097/00000539-200101000-00045. [DOI] [PubMed] [Google Scholar]
  362. Omote K, Yamamoto H, Kawamata T, Nakayama Y, Namiki A. The effects of intrathecal administration of an antagonist for prostaglandin E receptor subtype EP1 on mechanical and thermal hyperalgesia in a rat model of postoperative pain. Anesth Analg. 2002;95:1708–1712. doi: 10.1097/00000539-200212000-00044. [DOI] [PubMed] [Google Scholar]
  363. O'Reilly DD, Loomis CW. Spinal prostaglandins facilitate exaggerated A- and C-fiber-mediated reflex responses and are critical to the development of allodynia early after L5–L6 spinal nerve ligation. Anesthesiology. 2007;106:795–805. doi: 10.1097/01.anes.0000264777.94662.d6. [DOI] [PubMed] [Google Scholar]
  364. Palea S, Toson G, Pietra C, Trist DG, Artibani W, Romano O, et al. Pharmacological characterization of thromboxane and prostanoid receptors in human isolated urinary bladder. Br J Pharmacol. 1998;124:865–872. doi: 10.1038/sj.bjp.0701903. [DOI] [PMC free article] [PubMed] [Google Scholar]
  365. Paolini JF, Mitchel YB, Reyes R, Kher U, Lai E, Watson DJ, et al. Effects of laropiprant on nicotinic acid-induced flushing in patients with dyslipidemia. Am J Cardiol. 2008;101:625–630. doi: 10.1016/j.amjcard.2007.10.023. [DOI] [PubMed] [Google Scholar]
  366. Paralkar VM, Borovecki F, Ke HZ, Cameron KO, Lefker B, Grasser WA, et al. An EP2 receptor-selective prostaglandin E2 agonist induces bone healing. Proc Natl Acad Sci USA. 2003;100:6736–6740. doi: 10.1073/pnas.1037343100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  367. Patrignani P, Filabozzi P, Patrono C. Selective cumulative inhibition of platelet thromboxane production by low-dose aspirin in healthy subjects. J Clin Invest. 1982;69:1366–1372. doi: 10.1172/JCI110576. [DOI] [PMC free article] [PubMed] [Google Scholar]
  368. Patrono C. Thromboxane synthesis inhibitors and receptor antagonists. Thromb Res Suppl. 1990;11:15–23. doi: 10.1016/0049-3848(90)90387-r. [DOI] [PubMed] [Google Scholar]
  369. Patscheke H. Thromboxane A2/prostaglandin H2 receptor antagonists, a new therapeutic principle. Stroke. 1990;21:139–142. [PubMed] [Google Scholar]
  370. Peri K, Polyak F, Lubell W, Thouin E, Chemtob S. Peptides and peptidomimetics useful for inhibiting the activity of prostaglandin F2α receptor. US 2006/0211626, Quebec.
  371. Peters SP, Schleimer RP, Kagey-Sobotka A, Naclerio RM, MacGlashan DW, Schulman ES, et al. The role of prostaglandin D2 in IgE-mediated reactions in man. Trans Assoc Am Physicians. 1982;95:221–228. [PubMed] [Google Scholar]
  372. Pettipher R. The roles of the prostaglandin D2 receptors DP1 and CRTH2 in promoting allergic responses. Br J Pharmacol. 2008;153:S191–S199. doi: 10.1038/sj.bjp.0707488. [DOI] [PMC free article] [PubMed] [Google Scholar]
  373. Pettipher R, Hansel TT, Armer R. Antagonism of the prostaglandin D2 receptors DP1 and CRTH2 as an approach to treat allergic diseases. Nat Rev Drug Discov. 2007;6:313–325. doi: 10.1038/nrd2266. [DOI] [PubMed] [Google Scholar]
  374. Pfister SL, Pratt PE, Kurian J, Campbell WB. Glibenclamide inhibits thromboxane-mediated vasoconstriction by thromboxane receptor blockade. Vascul Pharmacol. 2004;40:285–292. doi: 10.1016/j.vph.2004.02.001. [DOI] [PubMed] [Google Scholar]
  375. Pierce KL, Regan JW. Prostanoid receptor heterogeneity through alternative mRNA splicing. Life Sci. 1998;62:1479–1483. doi: 10.1016/s0024-3205(98)00093-9. [DOI] [PubMed] [Google Scholar]
  376. Pollock WK, Armstrong RA, Brydon LJ, Jones RL, MacIntyre DE. Thromboxane-induced phosphatidate formation in human platelets. Relationship to receptor occupancy and to changes in cytosolic free calcium. Biochem J. 1984;219:833–842. doi: 10.1042/bj2190833. [DOI] [PMC free article] [PubMed] [Google Scholar]
  377. Pönicke K, Giessler C, Grapow M, Heinroth-Hoffman I, Becker K, Osten B, et al. FP-receptor mediated trophic effects of prostanoids in rat ventricular cardiomyocytes. Br J Pharmacol. 2000;129:1723–1731. doi: 10.1038/sj.bjp.0703243. [DOI] [PMC free article] [PubMed] [Google Scholar]
  378. Pulichino AM, Rowland S, Wu T, Clark P, Xu D, Mathieu MC, et al. Prostacyclin antagonism reduces pain and inflammation in rodent models of hyperalgesia and chronic arthritis. J Pharmacol Exp Ther. 2006;319:1043–1050. doi: 10.1124/jpet.106.110387. [DOI] [PubMed] [Google Scholar]
  379. Qian YM, Jones RL, Chan KM, Stock AI, Ho JK. Potent contractile actions of prostanoid EP3-receptor agonists on human isolated pulmonary artery. Br J Pharmacol. 1994;113:369–374. doi: 10.1111/j.1476-5381.1994.tb16997.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  380. Racké K, Bähring J, Langer C, Bräutigam M, Wessler I. Prostanoids inhibit release of endogenous norepinephrine from rat isolated trachea. Am Rev Respir Dis. 1992;146:1182–1186. doi: 10.1164/ajrccm/146.5_Pt_1.1182. [DOI] [PubMed] [Google Scholar]
  381. Raisz LG. Anabolic effect of prostaglandins. Clin Rev Bone Min. 2006;4:123–128. [Google Scholar]
  382. Rampton DS, Collins CE. Thromboxane in inflammatory bowel disease – pathogenic and therapeutic implications. Aliment Pharmacol Ther. 1993;7:357–367. [PubMed] [Google Scholar]
  383. Randall MJ, Parry MJ, Hawkeswood E, Cross PE, Dickinson RP. UK-37 248, a novel, selective thromboxane synthetase inhibitor with platelet anti-aggregatory and anti-thrombotic activity. Thromb Res. 1981;23:145–162. doi: 10.1016/0049-3848(81)90247-4. [DOI] [PubMed] [Google Scholar]
  384. Rangachari PK, Betti PA. Biological activity of metabolites of PGD2 on canine proximal colon. Am J Physiol. 1993;264:G886–G894. doi: 10.1152/ajpgi.1993.264.5.G886. [DOI] [PubMed] [Google Scholar]
  385. Rangachari PK, Betti P-A, Prior ET, Roberts LJ. Effects of a selective DP receptor agonist (BW 245C) and antagonist (BW A868C) on the canine colonic epithelium: an argument for a different DP receptor? J Pharmacol Exp Ther. 1995;275:611–617. [PubMed] [Google Scholar]
  386. Raychowdhury MK, Yukawa M, Collins LJ, McGrail SH, Kent KC, Ware JA. Alternative splicing produces a divergent cytoplasmic tail in the human endothelial thromboxane A2 receptor. J Biol Chem. 1994;269:19256–19261. [PubMed] [Google Scholar]
  387. Regan JW, Bailey TJ, Pepperl DJ, Pierce KL, Bogardus AM, Donello JE, et al. Cloning of a novel human prostaglandin receptor with characteristics of the pharmacologically defined EP2 subtype. Mol Pharmacl. 1994;46:213–220. [PubMed] [Google Scholar]
  388. Reilly IA, FitzGerald GA. Inhibition of thromboxane formation in vivo and ex vivo: implications for therapy with platelet inhibitory drugs. Blood. 1987;69:180–186. [PubMed] [Google Scholar]
  389. Rolin S, Dogné J-M, Michaux C, Delarge J, Masereel B. Activity of a novel dual thromboxane A2-receptor antagonist and thromboxane synthase inhibitor (BM-573) on platelet function and isolated smooth muscles. Prostaglandins Leukot Essent Fatty Acids. 2001;65:67–72. doi: 10.1054/plef.2001.0290. [DOI] [PubMed] [Google Scholar]
  390. Rolin S, Masereel B, Dogné J-M. Prostanoids as pharmacological targets in COPD and asthma. Eur J Pharmacol. 2006;533:89–100. doi: 10.1016/j.ejphar.2005.12.058. [DOI] [PubMed] [Google Scholar]
  391. Ross RA, Craib SJ, Stevenson LA, Pertwee RG, Henderson A, Toole J, et al. Pharmacological characterization of the anandamide cyclooxygenase metabolite: prostaglandin E2 ethanolamide. J Pharmacol Exp Ther. 2002;301:900–907. doi: 10.1124/jpet.301.3.900. [DOI] [PubMed] [Google Scholar]
  392. Ruel R, Lacombe P, Abramovitz M, Godbout C, Lamontagne S, Rochette C, et al. New class of biphenylene dibenzazocinones as potent ligands for the human EP1 prostanoid receptor. Bioorg Med Chem Lett. 1999;9:2699–2704. doi: 10.1016/s0960-894x(99)00465-5. [DOI] [PubMed] [Google Scholar]
  393. Saito O, Guan Y, Qi Z, Davis LS, Koernhoff M, Sugimoto Y, et al. Expression of the prostaglandin F receptor (FP) gene along the mouse genitourinary tract. Am J Physiol Renal Physiol. 2003;284:F1164–F1170. doi: 10.1152/ajprenal.00441.2002. [DOI] [PubMed] [Google Scholar]
  394. Sametz W, Hennerbichler S, Glaser S, Wintersteiger R, Juan H. Characterization of prostanoid receptors mediating actions of the isoprostanes, 8-iso-PGE2 and 8-iso-PGF2α, in some isolated smooth muscle preparations. Br J Pharmacol. 2000;130:1903–1910. doi: 10.1038/sj.bjp.0703522. [DOI] [PMC free article] [PubMed] [Google Scholar]
  395. Sanchez-Ramoz L, Kaunitz AM, Wears RL, Delke I, Gaudier FL. Misoprostol for cervical ripening and labor induction: a meta-analysis. Obstet Gynecol. 1997;89:633–642. doi: 10.1016/S0029-7844(96)00374-2. [DOI] [PubMed] [Google Scholar]
  396. Sanner JH. Antagonism of prostaglandin E2 by 1-acetyl-2-(8-chloro-10,11-dihydrobibenz[b,f][1,4]oxazepine-10-carbonyl) hydrazine (SC-19220) Arch Int Pharmacodyn Ther. 1969;180:46–56. [PubMed] [Google Scholar]
  397. Sarkar S, Hobson AR, Hughes A, Growcott J, Woolf CJ, Thompson DG, et al. The prostaglandin E2 receptor-1 (EP1) mediates acid-induced visceral pain hypersensitivity in humans. Gastroenterology. 2003;124:18–25. doi: 10.1053/gast.2003.50022. [DOI] [PubMed] [Google Scholar]
  398. Sawyer N, Cauchon E, Chateauneuf A, Cruz RP, Nicholson DW, Metters KM, et al. Molecular pharmacology of the human prostaglandin D2 receptor, CRTH2. Br J Pharmacol. 2002;137:1163–1172. doi: 10.1038/sj.bjp.0704973. [DOI] [PMC free article] [PubMed] [Google Scholar]
  399. Schlemper V, Medeiros R, Ferreira J, Campos MM, Calixto JB. Mechanisms underlying the relaxation response induced by bradykinin in the epithelium-intact guinea-pig trachea in vitro. Br J Pharmacol. 2005;145:740–750. doi: 10.1038/sj.bjp.0706222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  400. Schaaf TK, Hess HJ. Synthesis and biological activity of carboxyl-terminus modified prostaglandin analogues. J Med Chem. 1979;22:1340–1346. doi: 10.1021/jm00197a012. [DOI] [PubMed] [Google Scholar]
  401. Seiler SM, Brassard CL, Federici ME, Romini J, Meanwell NA. 3-[4-(4,5-Diphenyl-2-oxazolyl)-5-oxazolyl]phenoxy]acetic acid (BMY 45778) is a potent non-prostanoid prostacyclin partial agonist: effects on platelet aggregation, adenylyl cyclase, cAMP levels, protein kinase, and iloprost binding. Prostaglandins. 1997;53:21–35. doi: 10.1016/s0090-6980(96)00138-4. [DOI] [PubMed] [Google Scholar]
  402. Senchyna M, Crankshaw DJ. Characterization of the prostanoid TP receptor population in human nonpregnant myometrium. J Pharmacol Exp Ther. 1996;279:262–270. [PubMed] [Google Scholar]
  403. Senior J, Sangha R, Baxter GS, Marshall K, Clayton JK. In vitro characterization of prostanoid FP-, DP-, IP- and TP-receptors on the non-pregnant human myometrium. Br J Pharmacol. 1992;107:215–221. doi: 10.1111/j.1476-5381.1992.tb14489.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  404. Senior J, Marshall K, Sangha R, Clayton JK. In vitro characterization of prostanoid receptors on human myometrium at term pregnancy. Br J Pharmacol. 1993;108:501–506. doi: 10.1111/j.1476-5381.1993.tb12832.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  405. Sharif NA, Crider JY, Davis TL. AL-3138 antagonizes FP prostanoid receptor-mediated inositol phosphates generation: comparison with some purported FP antagonists. J Pharm Pharmacol. 2000;52:1229–1239. doi: 10.1211/0022357001777586. [DOI] [PubMed] [Google Scholar]
  406. Sharif NA, Senchyna M, Xu SX. Pharmacological and molecular biological (RT-PCR) characterization of functional TP prostanoid receptors in immortalized human non-pigmented ciliary epithelial cells. J Ocul Pharmacol Ther. 2002;18:141–162. doi: 10.1089/108076802317373905. [DOI] [PubMed] [Google Scholar]
  407. Sharif NA, Kaddour-Djebbar I, Abdel-Latif A. Cat iris sphincter smooth-muscle contraction: comparison of FP-class prostaglandin analog agonist activities. J Ocul Pharmacol Ther. 2008;24:152–163. doi: 10.1089/jop.2007.0076. [DOI] [PubMed] [Google Scholar]
  408. Shaw JS, Barker AJ, Breault G, Griffen EJ, Heapy CG, McLaughlin S. ZM325082: a potent and selective antagonist at the prostanoid EP1 receptor. 9th World Congress on Pain, Vienna, 1999. Abstract Book, p. 272.
  409. Shiraishi Y, Asano K, Nakajima T, Oguma T, Suzuki Y, Shiomi T, et al. Prostaglandin D2-induced eosinophilic airway inflammation is mediated by CRTH2 receptor. J Pharmacol Exp Ther. 2005;312:954–960. doi: 10.1124/jpet.104.078212. [DOI] [PubMed] [Google Scholar]
  410. Singh J, Zeller W, Zhou N, Hategen G, Mishra R, Polozov A, et al. Antagonists of the EP3 receptor for prostaglandin E2 are novel antiplatelet agents that do not prolong bleeding. ACS Chem Biol. 2009;4:115–126. doi: 10.1021/cb8002094. [DOI] [PubMed] [Google Scholar]
  411. Smith GCS, Coleman RA, McGrath JC. Characterization of dilator prostanoid receptors in the fetal rabbit ductus arteriosus. J Pharmacol Exp Ther. 1994;271:390–396. [PubMed] [Google Scholar]
  412. Smith JB. Effect of thromboxane synthetase inhibitors on platelet function: enhancement by inhibition of phosphodiesterase. Thromb Res. 1982;28:477–485. doi: 10.1016/0049-3848(82)90164-5. [DOI] [PubMed] [Google Scholar]
  413. Solomon SD, McMurray JJV, Pfeffer MA, Wittes J, Fowler R, Finn P, et al. Cardiovascular risk associated with celecoxib in a clinical trail for colorectal adenoma prevention. N Engl J Med. 2005;352:1071–1080. doi: 10.1056/NEJMoa050405. [DOI] [PubMed] [Google Scholar]
  414. Soyka R, Heckel A, Nickl J, Eisert W, Muller TH, Weisenberger H. 6,6-Disubstituted hex-5-enoic acid derivatives as combined thromboxane A2 receptor antagonists and synthase inhibitors. J Med Chem. 1993;37:26–39. doi: 10.1021/jm00027a004. [DOI] [PubMed] [Google Scholar]
  415. Spada CS, Krauss AH, Woodward DF, Chen J, Protzman CE, Nieves AL, et al. Bimatoprost and prostaglandin F2α selectivity stimulate intracellular calcium signaling in different cat iris sphincter cells. Exp Eye Res. 2005;80:135–145. doi: 10.1016/j.exer.2004.08.019. [DOI] [PubMed] [Google Scholar]
  416. Spik I, Brenuchon C, Angeli V, Staumont D, Fleury S, Capron M, et al. Activation of the prostaglandin D2 receptor DP2/CRTH2 increases allergic inflammation in mouse. J Immunol. 2005;174:3703–3708. doi: 10.4049/jimmunol.174.6.3703. [DOI] [PubMed] [Google Scholar]
  417. Stanke F, Cracowski JL, Chavanon O, Magne JL, Blin D, Bessard G, et al. Glibenclamide inhibits thromboxane A2-induced contraction in human internal mammary artery and saphenous vein. Eur J Pharmacol. 1998;341:65–71. doi: 10.1016/s0014-2999(97)01458-1. [DOI] [PubMed] [Google Scholar]
  418. Stefany D, Harris KJ, Gillespy TA, Gardner CJ, Aguiar JC. 2,6-Substituted-4-monosubstituted aminopyrimidine as prostaglandin D2 receptor antagonists. WO2007/121280, Sanofi-Aventis.
  419. Stinger RB, Fitzpatrick TM, Corey EJ, Ramwell PW, Rose JC, Kot PA. Selective antagonism of prostaglandin F2α-mediated vascular responses by N-dimethylamino substitution of prostaglandin F2α. J Pharmacol Exp Ther. 1982;220:521–525. [PubMed] [Google Scholar]
  420. Stock JL, Shinjo K, Burkhardt J, Roach M, Taniguchi K, Ishikawa T, et al. The prostaglandin E2 EP1 receptor mediates pain perception and regulates blood pressure. J Clin Invest. 2001;107:325–331. doi: 10.1172/JCI6749. [DOI] [PMC free article] [PubMed] [Google Scholar]
  421. Stubbs VE, Schratl P, Hartnell A, Williams TJ, Peskar BA, Heinemann A, et al. Indomethacin causes prostaglandin D2-like and eotaxin-like selective responses in eosinophils and basophils. J Biol Chem. 2002;277:26012–26020. doi: 10.1074/jbc.M201803200. [DOI] [PubMed] [Google Scholar]
  422. Sturino CF, Lachance N, Boyd M, Berthelette C, Labelle M, Li L, et al. Identification of an indole series of prostaglandin D2 receptor antagonists. Bioorg Med Chem Lett. 2006;16:3043–3048. doi: 10.1016/j.bmcl.2006.02.062. [DOI] [PubMed] [Google Scholar]
  423. Sturino CF, O'Neill G, Lachance N, Boyd M, Berthelette C, Labelle M, et al. Discovery of a potent and selective prostaglandin D2 receptor antagonist, [(3R)-4-(4-chloro-benzyl)-7-fluoro-5-(methylsulfonyl)-1,2,3,4-tetrahydrocyclopenta[b] indol- 3-yl]-acetic acid (MK-0524) J Med Chem. 2007;50:794–806. doi: 10.1021/jm0603668. [DOI] [PubMed] [Google Scholar]
  424. Su X, Lashinger ES, Leon LA, Hoffman BE, Hieble JP, Gardner SD, et al. An excitatory role for peripheral EP3 receptors in bladder afferent function. Am J Physiol Renal Physiol. 2008a;295:F585–F594. doi: 10.1152/ajprenal.90273.2008. [DOI] [PubMed] [Google Scholar]
  425. Su X, Leon LA, Wu CW, Morrow DM, Jaworski JP, Hieble JP, et al. Modulation of bladder function by prostaglandin EP3 receptors in the central nervous system. Am J Physiol Renal Physiol. 2008b;295:F984–F994. doi: 10.1152/ajprenal.90373.2008. [DOI] [PubMed] [Google Scholar]
  426. Sugimoto H, Shichijo M, Iino T, Manabe Y, Watanabe A, Shimazaki M, et al. An orally-bioavailable small molecule antagonist of CRTH2, ramatroban (BAY u3405), inhibits PGD2-induced eosinophil migration in vitro. J Pharmacol Exp Ther. 2003;305:347–352. doi: 10.1124/jpet.102.046748. [DOI] [PubMed] [Google Scholar]
  427. Sugimoto Y, Yamasaki A, Segi E, Tsuboi K, Aze Y, Nishimura T, et al. Failure of parturition in mice lacking the prostaglandin F-receptor. Science. 1997;277:681–683. doi: 10.1126/science.277.5326.681. [DOI] [PubMed] [Google Scholar]
  428. Suzawa T, Miyaura C, Inada M, Maruyama T, Sugimoto Y, Ushikubi F, et al. The role of prostaglandin E receptor subtypes (EP1, EP2, EP3, and EP4) in bone resorption; an analysis using specific agonists for the respective EPs. Endocrinology. 2000;141:1554–1559. doi: 10.1210/endo.141.4.7405. [DOI] [PubMed] [Google Scholar]
  429. Svensson CI, Yaksh TL. The spinal phospholipase-cyclooxygenase-prostanoid cascade in nociceptive processing. Annu Rev Pharmacol Toxicol. 2002;42:553–583. doi: 10.1146/annurev.pharmtox.42.092401.143905. [DOI] [PubMed] [Google Scholar]
  430. Swayne GT, Maguire J, Dolan J, Raval P, Dane G, Greener M, et al. Evidence for homogeneity of thromboxane A2 receptor using structurally different antagonists. Eur J Pharmacol. 1988;152:311–319. doi: 10.1016/0014-2999(88)90726-1. [DOI] [PubMed] [Google Scholar]
  431. Syriatowicz JP, Hu D, Walker JS, Tracey DJ. Hyperalgesia due to nerve injury: role of prostaglandins. Neuroscience. 1999;94:587–594. doi: 10.1016/s0306-4522(99)00365-6. [DOI] [PubMed] [Google Scholar]
  432. Takahara K, Murray R, Fitzgerald GA, Fitzgerald DJ. The response to thromboxane A2 analogues in human platelets. Discrimination of two binding sites linked to distinct effector systems. J Biol Chem. 1990;265:6836–6844. [PubMed] [Google Scholar]
  433. Takasaki I, Nojima H, Shiraki K, Sugimoto Y, Ichikawa A, Ushikubi F, et al. Involvement of cyclooxygenase-2 and EP3 prostaglandin receptor in acute herpetic but not postherpetic pain in mice. Neuropharmacology. 2005;49:283–292. doi: 10.1016/j.neuropharm.2004.12.025. [DOI] [PubMed] [Google Scholar]
  434. Takayama K, Garcia-Cardena G, Sukhova GK, Comander J, Gimbrone MA, Jr, Libby P. Prostaglandin E2 suppresses chemokine production in human macrophages through the EP4 receptor. J Biol Chem. 2002;277:44147–44154. doi: 10.1074/jbc.M204810200. [DOI] [PubMed] [Google Scholar]
  435. Tanaka T, Fukuta Y, Higashino R, Sato R, Nomura Y, Fukuda Y, et al. Antiplatelet effect of Z-335, a new orally active and long-lasting thromboxane receptor antagonist. Eur J Pharmacol. 1998;357:53–60. doi: 10.1016/s0014-2999(98)00540-8. [DOI] [PubMed] [Google Scholar]
  436. Tanaka K, Ogawa K, Sugamura K, Nakamura M, Takano S, Nagata K. Differential production of prostaglandin D2 by human helper T cell subsets. J Immunol. 2000;164:2277–2280. doi: 10.4049/jimmunol.164.5.2277. [DOI] [PubMed] [Google Scholar]
  437. Tani K, Naganawa A, Ishida A, Sagawa K, Harada H, Ogawa M, et al. Development of a highly selective EP2-receptor agonist. Part 1: identification of 16-hydroxy-17,17-trimethylene PGE2 derivatives. Bioorg Med Chem Lett. 2002;10:1093–1106. doi: 10.1016/s0968-0896(01)00369-8. [DOI] [PubMed] [Google Scholar]
  438. Teixeira MM, al-Rashed S, Rossi AG, Hellewell PG. Characterization of the prostanoid receptors mediating inhibition of PAF-induced aggregation of guinea-pig eosinophils. Br J Pharmacol. 1997;121:77–82. doi: 10.1038/sj.bjp.0701107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  439. Terada N, Yamakoshi T, Hasegawa M, Tanikawa H, Maesako K, Ishikawa K, et al. The effect of ramatroban (BAY u3405), a thromboxane A2 receptor antagonist, on nasal cavity volume and minimum cross-sectional area and nasal mucosal hemodynamics after nasal mucosal allergen challenge in patients with perennial allergic rhinitis. Acta Otolaryngol Suppl. 1998;537:32–37. doi: 10.1080/00016489850182323. [DOI] [PubMed] [Google Scholar]
  440. Theiner G, Gessner A, Lutz MB. The mast cell mediator PGD2 suppresses IL-12 release by dendritic cells leading to Th2 polarized immune responses in vivo. Immunobiology. 2006;211:463–472. doi: 10.1016/j.imbio.2006.05.020. [DOI] [PubMed] [Google Scholar]
  441. Tober KL, Wilgus TA, Kusewitt DF, Thomas-Ahner JM, Maruyama T, Oberyszyn TM. Importance of the EP1-receptor in cutaneous UVB-induced inflammation and tumor development. J Invest Dermatol. 2006;126:205–211. doi: 10.1038/sj.jid.5700014. [DOI] [PubMed] [Google Scholar]
  442. Tsuri T, Honma T, Hiramatsu Y, Okada T, Hashizume H, Mitsumori S, et al. Bicyclo[2.2.1]heptane and 6,6-dimethylbicyclo[3.1.1]heptane derivatives: orally active, potent, and selective prostaglandin D2 receptor antagonists. J Med Chem. 1997;40:3504–3507. doi: 10.1021/jm970343g. [DOI] [PubMed] [Google Scholar]
  443. Torisu K, Kobayashi K, Iwahashi M, Egashira H, Nakai Y, Okada Y, et al. Discovery of new chemical leads for prostaglandin D2 receptor antagonists. Bioorg Med Chem Lett. 2004a;14:4557–4562. doi: 10.1016/j.bmcl.2004.06.006. [DOI] [PubMed] [Google Scholar]
  444. Torisu K, Kobayashi K, Iwahashi M, Nakai Y, Onoda T, Nagase T, et al. Development of prostaglandin D2 receptor antagonist: discovery of highly potent antagonists. Bioorg Med Chem. 2004b;12:4685–4700. doi: 10.1016/j.bmc.2004.06.024. [DOI] [PubMed] [Google Scholar]
  445. Torisu K, Kobayashi K, Iwahashi M, Nakai Y, Onoda T, Nagase T, et al. Discovery of a new class of potent, selective, and orally active prostaglandin D2 receptor antagonists. Bioorg Med Chem. 2004c;12:5361–5378. doi: 10.1016/j.bmc.2004.07.048. [DOI] [PubMed] [Google Scholar]
  446. Torisu K, Kobayashi K, Iwahashi M, Nakai Y, Onoda T, Nagase T, et al. ). Discovery of orally active prostaglandin D2 receptor antagonists. Bioorg Med Chem Lett. 2004d;14:4891–4895. doi: 10.1016/j.bmcl.2004.07.039. [DOI] [PubMed] [Google Scholar]
  447. Torisu K, Kobayashi K, Iwahashi M, Egashira H, Nakai Y, Okada Y, et al. Development of a prostaglandin D2 receptor antagonist: discovery of a new chemical lead. Eur J Med Chem. 2005;40:505–519. doi: 10.1016/j.ejmech.2004.11.011. [DOI] [PubMed] [Google Scholar]
  448. Tymkewycz PM, Jones RL, Wilson NH, Marr CG. Heterogeneity of thromboxane A2 (TP-) receptors: evidence from antagonist but not agonist potency measurements. Br J Pharmacol. 1991;102:607–614. doi: 10.1111/j.1476-5381.1991.tb12220.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  449. Uda R, Horiguchi S, Ito S, Hyodo M, Hayashi O. Nociceptive effects induced by intrathecal administration of prostaglandin D2, E2, or F2α to conscious mice. Brain Res. 1990;510:26–32. doi: 10.1016/0006-8993(90)90723-o. [DOI] [PubMed] [Google Scholar]
  450. Ueno A, Matsumoto H, Naraba H, Ikeda Y, Ushikubi F, Matsuoka T, et al. Major roles of prostanoid receptors IP and EP3 in endotoxin-induced enhancement of pain perception. Biochem Pharmacol. 2001;62:157–160. doi: 10.1016/s0006-2952(01)00654-2. [DOI] [PubMed] [Google Scholar]
  451. Uller L, Mathiesen JM, Alenmyr L, Korsgren M, Ulven T, Hogberg T, et al. Antagonism of the prostaglandin D2 receptor CRTH2 attenuates asthma pathology in mouse eosinophilic airway inflammation. Respir Res. 2007;8(16):1–10. doi: 10.1186/1465-9921-8-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  452. Ulven T, Kostenis E. Minor structural modifications convert the dual TP/CRTH2 antagonist ramatroban into a highly selective and potent CRTH2 antagonist. J Med Chem. 2005;48:897–900. doi: 10.1021/jm049036i. [DOI] [PubMed] [Google Scholar]
  453. Ulven T, Kostenis E. Targeting the prostaglandin D2 receptors DP and CRTH2 for treatment of inflammation. Curr Top Med Chem. 2006;6:1427–1444. doi: 10.2174/15680266106061427. [DOI] [PubMed] [Google Scholar]
  454. Umemura K, Nakashima M, Saruta T. Thromboxane A2 synthetase inhibition suppresses cough induced by angiotensin converting enzyme inhibitors. Life Sci. 1997;60:1583–1588. doi: 10.1016/s0024-3205(97)00124-0. [DOI] [PubMed] [Google Scholar]
  455. Ushikubi F, Segi E, Sugimoto Y, Murata T, Matsuoka T, Kobayashi T, et al. Impaired febrile response in mice lacking the prostaglandin E receptor subtype EP3. Nature. 1998;395:281–284. doi: 10.1038/26233. [DOI] [PubMed] [Google Scholar]
  456. Van Hecken A, Depre M, De LI, Thach C, Oeyen M, Van Effen J, et al. The effect of MK-0524, a prostaglandin D2 receptor antagonist, on prostaglandin D2-induced nasal airway obstruction in healthy volunteers. Eur J Clin Pharmacol. 2007;63:135–141. doi: 10.1007/s00228-006-0211-2. [DOI] [PubMed] [Google Scholar]
  457. Vermylen J, Defreyn G, Carreras LO, Machin SJ, Van Schaeren J, Verstraete M. Thromboxane synthetase inhibition as antithrombotic strategy. Lancet. 1981;1(8229):1073–1075. doi: 10.1016/s0140-6736(81)92241-8. [DOI] [PubMed] [Google Scholar]
  458. Vezza R, Roberti R, Nenci GG, Gresele P. Prostaglandin E2 potentiates platelet aggregation by priming protein kinase C. Blood. 1993;82:2704–2713. [PubMed] [Google Scholar]
  459. Walch L, Labat C, Gascard JP, de Montpreville V, Brink C, Norel X. Prostanoid receptors involved in the relaxation of human pulmonary vessels. Br J Pharmacol. 1999;126:859–866. doi: 10.1038/sj.bjp.0702393. [DOI] [PMC free article] [PubMed] [Google Scholar]
  460. Walch L, de Montpreville V, Brink C, Norel X. Prostanoid EP1- and TP-receptors involved in the contraction of human pulmonary veins. Br J Pharmacol. 2001;134:1671–1678. doi: 10.1038/sj.bjp.0704423. [DOI] [PMC free article] [PubMed] [Google Scholar]
  461. Walch L, Clavarini E, Morris PL. Prostaglandin (PG) FP and EP1 receptors mediate PGF2α and PGE2 regulation of interleukin-1β expression in Leydig cell progenitors. Endocrinology. 2003;144:1284–1291. doi: 10.1210/en.2002-220868. [DOI] [PubMed] [Google Scholar]
  462. Wan Z, Woodward DF, Cornell CL, Fliri HG, Martos JS, Pettit SN, et al. Bimatoprost, prostamide activity, and conventional drainage. Invest Ophthalmol Vis Sci. 2007;48:4107–4115. doi: 10.1167/iovs.07-0080. [DOI] [PMC free article] [PubMed] [Google Scholar]
  463. Wang Z, Beaulieu C, Dufresne C, Guay D, Leblanc Y. Dihyrdopyrrolo[1,2-a]indole and tetrahydropyrido[1,2-a]indole derivatives as prostaglandin D2 receptor antagonists. WO2002/094830, Merck Frosst Canada & Co.
  464. Wardle EN. Thromboxane in glomerulonephritis: what about therapy? Am J Ther. 1999;6:111–114. doi: 10.1097/00045391-199903000-00008. [DOI] [PubMed] [Google Scholar]
  465. Watanabe K, Kawamori T, Nakatsugi S, Ohta T, Ohuchida S, Yamamoto H, et al. Role of the prostaglandin E receptor subtype EP1 in colon carcinogenesis. Cancer Res. 1999;59:5093–5096. [PubMed] [Google Scholar]
  466. Watanabe K, Kawamori T, Nakatsugi S, Ohta T, Ohuchida S, Yamamoto H, et al. Inhibitory effect of a prostaglandin E receptor subtype EP1 selective antagonist, ONO-8713, on development of azoxymethane-induced aberrant crypt foci in mice. Cancer Lett. 2000;156:57–61. doi: 10.1016/s0304-3835(00)00440-7. [DOI] [PubMed] [Google Scholar]
  467. Watts IS, Wharton KA, White BP, Lumley P. Thromboxane (Tx) A2 receptor blockade and TxA2 synthase inhibition alone and in combination: comparison of anti-aggegatory efficacy in human platelets. Br J Pharmacol. 1991;102:487–505. doi: 10.1111/j.1476-5381.1991.tb12200.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  468. Webb ML, Liu EC, Monshizadegan H, Hedberg A, Misra RN, Goldenberg H, et al. Binding and function of a potent new thromboxane receptor antagonist, BMS 180,291, in human platelets. J Pharmacol Exp Ther. 1993;264:1387–1394. [PubMed] [Google Scholar]
  469. Westwick J, Webb H. Selective antagonism of prostaglandin (PG) E1, PGD2 and prostacyclin (PGI2) on human and rabbit platelets by di-4-phloretin phosphate (DPP) Thromb Res. 1978;12:973–978. doi: 10.1016/0049-3848(78)90053-1. [DOI] [PubMed] [Google Scholar]
  470. Wheeldon A, Vardey CJ. Characterization of the inhibitory prostanoid receptors on human neutrophils. Br J Pharmacol. 1993;108:1051–1054. doi: 10.1111/j.1476-5381.1993.tb13504.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  471. Wilson NH, Jones RL. Prostaglandin endoperoxide and thromboxane A2 analogues. Adv Prostaglandin Thromboxane Leukot Res. 1985;14:393–425. [PubMed] [Google Scholar]
  472. Wilson RJ, Giblin GM, Roomans S, Rhodes SA, Cartwright KA, Shield VJ, et al. GW627368X ((N-{2-[4-(4,9-diethoxy-1-oxo-1,3-dihydro-2H-benzo[f]isoindol-2-yl)phenyl]acetyl} benzene sulphonamide): a novel, potent and selective prostanoid EP4 receptor antagonist. Br J Pharmacol. 2006;148:326–339. doi: 10.1038/sj.bjp.0706726. [DOI] [PMC free article] [PubMed] [Google Scholar]
  473. Wilson SJ, Roche AM, Kostetskaia E, Smyth EM. Dimerization of the human receptors for prostacyclin and thromboxane facilitates thromboxane receptor-mediated cAMP generation. J Biol Chem. 2004;279:53036–53047. doi: 10.1074/jbc.M405002200. [DOI] [PubMed] [Google Scholar]
  474. Wohlsen A, Martin C, Vollmer E, Branscheid D, Magnussen H, Becker W-M, et al. The early allergic response in small airways of human precison-cut lung slices. Eur Respir J. 2003;21:1024–1032. doi: 10.1183/09031936.03.00027502. [DOI] [PubMed] [Google Scholar]
  475. Woodward DF, Hawley SB, Williams LS, Ralston TL, Protzman CE, Spada CS, et al. Studies on the ocular pharmacology of prostaglandin D2. Invest Ophthalmol Vis Sci. 1990a;31:138–146. [PubMed] [Google Scholar]
  476. Woodward DF, Nieves AL, Williams LS. Thromboxane A2-mimetics are potent microvascular permeability factors in the conjunctiva. J Pharmacol Exp Ther. 1990b;255:23–27. [PubMed] [Google Scholar]
  477. Woodward DF, Protzman CE, Krauss AH, Williams LS. Identification of 19(R)-OH prostaglandin E2 as a selective prostanoid EP2-receptor agonist. Prostaglandins. 1993a;46:371–383. doi: 10.1016/0090-6980(93)90102-d. [DOI] [PubMed] [Google Scholar]
  478. Woodward DF, Spada CS, Hawley SB, Williams LS, Protzman CE, Nieves AL. Further studies on ocular responses to DP-receptor stimulation. Eur J Pharmacol. 1993b;230:327–333. doi: 10.1016/0014-2999(93)90569-4. [DOI] [PubMed] [Google Scholar]
  479. Woodward DF, Fairbairn CE, Krauss H-P, Lawrence RA, Protzman CE. Radioligand binding analysis of receptor subtypes in two FP receptors preparations that exhibit different functional rank orders of potency in response to prostaglandin. J Pharmacol Exp Ther. 1995a;273:285–291. [PubMed] [Google Scholar]
  480. Woodward DF, Pepperl DJ, Burkley TH, Regan JW. 6-Isopropoxy-9-oxoxanthene-2-carboxylic acid (AH 6809), a human EP2 receptor antagonist. Biochem Pharmacol. 1995b;50:1731–1733. doi: 10.1016/0006-2952(95)02035-7. [DOI] [PubMed] [Google Scholar]
  481. Woodward DF, Krauss AH-P, Chen J, Gil DW, Kedzie KM, Protzman CE, et al. Replacement of the carboxylic acid group of prostaglandin F2α with a hydroxyl or methoxy substitute provides biologically unique compounds. Br J Pharmacol. 2000;130:1933–1943. doi: 10.1038/sj.bjp.0703462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  482. Woodward DF, Krauss AH, Chen J, Liang Y, Li C, Protzman CE, et al. Pharmacological characterization of a novel antiglaucoma agent, Bimatoprost (AGN 192024) J Pharmacol Exp Ther. 2003;305:772–785. doi: 10.1124/jpet.102.047837. [DOI] [PubMed] [Google Scholar]
  483. Woodward DF, Phelps RL, Krauss AH, Weber A, Short B, Chen J, et al. Bimatoprost: a novel antiglaucoma agent. Cardiovasc Drug Rev. 2004;22:103–120. doi: 10.1111/j.1527-3466.2004.tb00134.x. [DOI] [PubMed] [Google Scholar]
  484. Woodward DF, Krauss AH-P, Wang JW, Protzman CE, Nieves AL, Liang Y, et al. Identification of an antagonist that selectively blocks the activity of prostamides (prostaglandin-ethanolamides) in the feline iris. Br J Pharmacol. 2007;150:342–352. doi: 10.1038/sj.bjp.0706989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  485. Woodward DF, Liang Y, Krauss AH. Prostamides (prostaglandin-ethanolamides) and their pharmacology. Br J Pharmacol. 2008;153:410–419. doi: 10.1038/sj.bjp.0707434. [DOI] [PMC free article] [PubMed] [Google Scholar]
  486. Xiang A, Uchida Y, Nomura A, Iijima H, Sakamoto T, Ishii Y, et al. Involvement of thromboxane A2 in airway mucous cells in asthma-related cough. J Appl Physiol. 2002;92:763–770. doi: 10.1152/jappl.2002.92.2.763. [DOI] [PubMed] [Google Scholar]
  487. Xu W, Chou C-L, Sun H, Fujino H, Chen QM, Regan JW. FP prostanoid receptor-mediated induction of the expression of early growth response factor-1 by activation of a Ras/Raf/mitogen-activated protein kinase signaling cascade. Mol Pharmacol. 2008;73:111–118. doi: 10.1124/mol.107.038778. [DOI] [PubMed] [Google Scholar]
  488. Xue L, Gyles SL, Wettey FR, Gazi L, Townsend E, Hunter MG, et al. Prostaglandin D2 causes preferential induction of pro-inflammatory Th2 cytokine production through an action on chemoattractant receptor-like molecule expressed on Th2 cells. J Immunol. 2005;175:6531–6536. doi: 10.4049/jimmunol.175.10.6531. [DOI] [PubMed] [Google Scholar]
  489. Yanagisawa A, Smith JA, Brezinski ME, Lefer AM. Mechanism of antagonism of thromboxane receptors in vascular smooth muscle. Eur J Pharmacol. 1987;133:89–96. doi: 10.1016/0014-2999(87)90209-3. [DOI] [PubMed] [Google Scholar]
  490. Yang J, Xia M, Goetzl EJ, An S. Cloning and expression of the EP3-subtype of human receptors for prostaglandin E2. Biochem Biophys Res Commun. 1994;198:999–1006. doi: 10.1006/bbrc.1994.1142. [DOI] [PubMed] [Google Scholar]
  491. Yang W, Ni J, Woodward DF, Tang-Liu DD, Ling KH. Enzymatic formation of prostamide F2α from anandamide involves as newly identified intermediate metabolite, prostamide H2. J Lipid Res. 2005;46:2745–2751. doi: 10.1194/jlr.M500374-JLR200. [DOI] [PubMed] [Google Scholar]
  492. Yang Z, Reiling S, Nieduzak TR, Mathew RM, Jackson S, Harris KJ. 2-Phenylindoles as prostaglandin D2 receptor antagonists. WO2008/014186, Sanofi-Aventis.
  493. Yasui K, Asanuma F, Hirano Y, Shichijo M, Deguchi M, Arimura A. Contribution of prostaglandin D2 via prostanoid DP receptor to nasal hyperresponsiveness in guinea pigs repeatedly exposed to antigen. Eur J Pharmacol. 2008;578:286–291. doi: 10.1016/j.ejphar.2007.09.014. [DOI] [PubMed] [Google Scholar]
  494. Yokoyama C, Yabuki T, Shimonishi M, Wada M, Hatae T, Ohkwara S, et al. Prostacyclin-deficient mice develop ischaemic renal disorders, including nephrosclerosis and renal infarction. Circulation. 2002;106:2397–2403. doi: 10.1161/01.cir.0000034733.93020.bc. [DOI] [PubMed] [Google Scholar]
  495. Yokoyama U, Minamisawa S, Quan H, Ghatak S, Akaike T, Segi-Nishida E, et al. Chronic activation of the prostaglandin receptor EP4 promotes hyaluronan-mediated neointimal formation in the ductus arteriosus. J Clin Invest. 2006;116:3026–3034. doi: 10.1172/JCI28639. [DOI] [PMC free article] [PubMed] [Google Scholar]
  496. Yoshida M, Sato Y, Shimura T, Ohkubo S, Honma S, Tanaka T, et al. Distinct effects of Z-335, a new thromboxane A2 receptor antagonist, on rabbit platelets and aortic smooth muscle. Pharmacology. 2007;79:50–56. doi: 10.1159/000097632. [DOI] [PubMed] [Google Scholar]
  497. Yoshikawa M, Takagi K, Ohinata K, Inui A, Asakawa A, Kakudo S. Novel uses of prostaglandin D2, prostaglandin D2 agonist and prostaglandin D2 antagonist. US2005/0215609, Shionogi & Co.
  498. Yu M, Ives D, Ramesha CS. Synthesis of prostaglandin E2 ethanolamide from anandamide by cyclooxygenase-2. J Biol Chem. 1997;272:21181–21186. doi: 10.1074/jbc.272.34.21181. [DOI] [PubMed] [Google Scholar]
  499. Zacharowski K, Olbrich A, Otto M, Hafner GS, Thiemermann C. Effects of the prostanoid EP3-receptor agonists M&B 28767 and GR 63799X on infarct size caused regional myocardial ischemia in the anaesthetized rat. Br J Pharmacol. 1999;126:849–858. doi: 10.1038/sj.bjp.0702348. [DOI] [PMC free article] [PubMed] [Google Scholar]
  500. Zehender H, Witte EC, Wolff HP, Patscheke H. Irreversible inhibition of the TXA2/PGH2 receptor of human platelets by a photoaffinity ligand. Biochem Pharmacol. 1988;37:491–495. doi: 10.1016/0006-2952(88)90219-5. [DOI] [PubMed] [Google Scholar]
  501. Zhang R, Ogletree ML, Moreland S. Characterization of thromboxane A2/prostaglandin endoperoxide receptors in aorta. Eur J Pharmacol. 1996;317:91–96. doi: 10.1016/s0014-2999(96)00697-8. [DOI] [PubMed] [Google Scholar]
  502. Zhang Z, Hefferan MP, Loomis CW. Topical bicuculline to the rat spinal cord induces highly localized allodynia that is mediated by spinal prostaglandins. Pain. 2001;92:351–361. doi: 10.1016/S0304-3959(01)00276-7. [DOI] [PubMed] [Google Scholar]
  503. Zhou N, Zeller W, Krohn M, Anderson H, Zhang J, Onua E, et al. 3,4-Disubstituted indole acylsulfonamides: a novel series of potent and selective human EP3 receptor antagonists. Bioorg Med Chem Lett. 2009a;19:123–126. doi: 10.1016/j.bmcl.2008.11.007. [DOI] [PubMed] [Google Scholar]
  504. Zhou N, Zeller W, Zhang J, Onua E, Kiselyov AS, Ramirez J, et al. 3-Acrylamide-4-aryloxyindoles: synthesis, biological evaluation and metabolic stability of potent and selective EP3 receptor antagonists. Bioorg Med Chem Lett. 2009b;19:1528–1531. doi: 10.1016/j.bmcl.2008.12.112. [DOI] [PubMed] [Google Scholar]

Articles from British Journal of Pharmacology are provided here courtesy of The British Pharmacological Society

RESOURCES