当前位置: 首页 > 医学版 > 期刊论文 > 内科学 > 内分泌进展 > 2005年 > 第1期 > 正文
编号:11168576
Proteinase-Activated Receptors: Transducers of Proteinase-Mediated Signaling in Inflammation and Immune Response
http://www.100md.com 内分泌进展 2005年第1期
     Department of Dermatology and Ludwig Boltzmann Institute for Cell and Immunobiology of the Skin (M.S., J.B., V.S., A.R., C.M., T.A.L.), University of Münster, 48149 Münster, Germany

    Department of Pharmacology (N.V., M.D.H.), University of Calgary, Calgary, Canada T2N 4N1

    Abstract

    Serine proteinases such as thrombin, mast cell tryptase, trypsin, or cathepsin G, for example, are highly active mediators with diverse biological activities. So far, proteinases have been considered to act primarily as degradative enzymes in the extracellular space. However, their biological actions in tissues and cells suggest important roles as a part of the body’s hormonal communication system during inflammation and immune response. These effects can be attributed to the activation of a new subfamily of G protein-coupled receptors, termed proteinase-activated receptors (PARs). Four members of the PAR family have been cloned so far. Thus, certain proteinases act as signaling molecules that specifically regulate cells by activating PARs. After stimulation, PARs couple to various G proteins and activate signal transduction pathways resulting in the rapid transcription of genes that are involved in inflammation. For example, PARs are widely expressed by cells involved in immune responses and inflammation, regulate endothelial-leukocyte interactions, and modulate the secretion of inflammatory mediators or neuropeptides. Together, the PAR family necessitates a paradigm shift in thinking about hormone action, to include proteinases as key modulators of biological function. Novel compounds that can modulate PAR function may be potent candidates for the treatment of inflammatory or immune diseases.

    I. Introduction

    II. PAR1 in Inflammation and Immune Response

    A. Vasculature and heart: PAR1 antagonists—a novel potential approach for the treatment of cardiovascular diseases

    B. Platelets

    C. Immune cells

    D. Airways

    E. Gastrointestinal tract

    F. Kidneys and urogenital tract

    G. Brain and peripheral nervous system

    H. Signaling by proteinases via PAR1

    III. PAR2 in Inflammation and Immune Response

    A. Vasculature

    B. Immune cells

    C. Arthritis

    D. Skin

    E. Airways

    F. Brain and peripheral nervous system

    G. Digestive tract and pancreas

    H. Signaling by proteinases via PAR2

    IV. PAR3 and PAR4

    A. Biology and distribution of PAR3 and PAR4

    B. Signaling by proteinases via PAR3 and PAR4

    V. Conclusions

    I. Introduction

    SERINE PROTEINASES CONSTITUTE a family of proteolytic enzymes characterized by a unique catalytic triad consisting of Ser, His, and Asp. These residues are able to hydrolyze peptide bonds (1). Serine proteinases are produced as inactive precursors or zymogens. Subsequent zymogen conversion into a mature physiologically active enzyme is mediated via a process called "limited proteolysis" or zymogen activation (2, 3, 4). In mammals, serine proteases, for example, regulate the hemostatic and fibrinolytic balance, degrade neuropeptides involved in neurogenic inflammation or serve as modulators of immune response during inflammation (5). Three different types of serine proteinase inhibitors can be distinguished based on their mechanism of action: canonical, noncanonical inhibitors, and serpins (6). An imbalance between these inhibitors and their targeted proteinases can affect immune/inflammatory responses and may result in disease. Moreover, the presence of proteinase inhibitors regulates and limits interactions between proteinases and their receptors (7, 8, 9, 10, 11).

    Recent studies elucidated the ability of certain serine proteinases to regulate cell function via G protein-coupled receptors (GPCRs). At least two different types of proteinase receptors have been identified involving proteolytic cleavage in their activation mechanism: urokinase receptors and proteinase-activated receptors (PARs) (12, 13, 14).

    PARs belong to a new subfamily of GPCRs with seven transmembrane domains activated via proteolytic cleavage by serine proteinases (13, 14, 15, 16). PAR1, PAR3, and PAR4 are targets for thrombin, trypsin, or cathepsin G (17, 18, 19, 20). In contrast, PAR2 is resistant to thrombin, but can be activated by trypsin, mast cell tryptase, factor Xa, acrosin, gingipain, and neuronal serine proteinases (21, 22, 23, 24, 25, 26, 27, 28) (Table 1). Interestingly, PARs are activated by a unique mechanism: proteinases activate PARs by proteolytic cleavage within the extracellular N terminus of their receptors, thereby exposing a novel "cryptic" receptor-activating N-terminal sequence that, remaining tethered, binds to and activates the receptor (Fig. 1) within the same receptor (17, 21, 22). Specific residues (about six amino acids) within this tethered ligand domain are believed to interact with extracellular loop 2 and other domains of the cleaved receptor (29), resulting in activation. This intramolecular activation process is followed by coupling to G proteins and the triggering of a variety of downstream signal transduction pathways (see Sections II.H, III.H, and IV.B; also see Table 2 and Fig. 2, and Refs.13 , 14 , and 16). Thus, PARs are not activated like "classical" receptors because the specific receptor-activating ligand is part of the receptor, whereas the circulating agonist is a relatively nonspecific serine proteinase that does not behave like a traditional hormonal regulator akin to insulin.

    Several studies during the past few years have also demonstrated that several mechanisms exist to regulate stimulation and termination of PAR-initiated signaling (13, 14, 15, 16, 30). Importantly, the availability of PARs at the cell surface is governed by trafficking of the receptor from intracellular stores, and the signaling properties depend on the presence of G proteins and G protein-coupled receptor kinases (GRKs) that modify activity. For PAR1, PAR2, and PAR4, it is well established that short synthetic peptides [PAR-activating peptides (PAR-APs)] designed on their proteolytically revealed tethered ligand sequences can serve as selective receptor agonists (Table 1). Some PAR-APs activate more than one PAR, and they activate receptors at concentrations in the micromolar range as compared with nanomolar potencies of the proteinases themselves (31, 32). Although the PAR1-AP, SFLLRN-NH2 also activates PAR2, PAR2-APs, like SLIGRL-NH2 are not capable of activating other PARs. Unfortunately, the relatively low potency (10 to 100 μM EC50) and susceptibility to aminopeptidases (33) limit the utility of the PAR-APs in some bioassay systems. Recently, modified synthetic agonist peptides with higher potency, resistance to aminopeptidases, and greater receptor selectivity have been developed and characterized. These receptor-selective agonists are of use to study the consequences of activating PARs in vivo (Table 1) (13, 14, 16, 34). So far, antagonists for PAR1 (35, 36, 37) and PAR4 (38) have been synthesized, but are not yet available for PAR2. PAR3, as will be elaborated upon in Section IV, remains a puzzle, in that studies with the murine receptor indicate that it cannot be activated either by its cognate synthetic tethered ligand peptide or by thrombin (39, 40). Rather than acting as an independent cell regulator, PAR3 appears to function as a cofactor for the activation of PAR4 (39). Complementary data documenting PAR-mediated effects in various tissues have been obtained using PAR-deficient (PAR–/–) mice.

    Taken together, data obtained using the enzyme activators themselves (trypsin, thrombin), the PAR-APs and using PAR gene-deficient mice provide compelling evidence that PARs play a critical role in the regulation of various physiological and pathophysiological functions in mammals, including humans. This review focuses on the biology and signaling properties of PARs in various mammalian tissues and highlights the current knowledge about the role of PARs during inflammation and immune response. To complement the information summarized in the sections that follow, the reader is encouraged to consult a number of other recent reviews concerning the activation mechanisms and cell biological aspects of PARs (13, 14, 16, 34, 41, 42, 43, 44, 45).

    II. PAR1 in Inflammation and Immune Response

    Thrombin is an important effector proteinase of the coagulation cascade that leads to formation of a hemostatic plug. Thrombin is thought to act near the site at which it is generated and it is activated when circulating coagulation factors in the blood plasma make contact with tissue factor. Tissue factor is a membrane protein that is usually produced by cells that are separated from blood (i.e., epithelial cells). However, it is also expressed at low levels on circulating monocytes and microparticles from leukocytes. Tissue factor is associated with the activation of zymogen factor X by factor VIIa. Factor Xa together with its cofactor Va subsequently converts prothrombin to the active enzyme. Thus, plasma coagulation can only take place when the vascular integrity is damaged (15). Thrombin causes shape change of endothelial cells and increased permeability of endothelial cell layers.

    However, recent evidence revealed that thrombin is not only a clotting proteinase serving as both a pro- and anticoagulant molecule but also appears to play multifunctional roles related to inflammation, allergy, tumor growth, metastasis, tissue remodeling, thrombosis, and probably wound healing (13, 15, 42, 43). Subsequent to the cloning of PAR1 (17, 46), it was realized that many of the cellular actions of thrombin (e.g., platelet aggregation, angiogenesis, endothelial cell permeability, vasoregulation, gene regulation, leukocyte trafficking, immunomodulation) could be attributed to the activation of its GPCR (40, 47, 48, 49, 50, 51, 52, 53).

    PAR1 has been detected in a variety of tissues, including platelets; endothelial cells; fibroblasts; monocytes; T cells positive for CD 8, CD 16, and either CD 56 or CD 57 (54, 55); natural killer (NK) cells (48); CD 34+ hematopoietic progenitor cells (56); dental pulp cells (57); smooth muscle cells (SMC); epithelial cells; neurons; glial cells; mast cells; and certain tumor cell lines (17, 54, 58, 59, 60, 61). A receptor with high affinity for thrombin has also been detected in rat peritoneal macrophages (62). It should be pointed out, however, that PAR1 very likely does not represent the only target for thrombin, in that other (non-PAR) high-affinity binding sites for thrombin [e.g., on platelets or macrophages (62)] and other conserved thrombin sequences apart from the catalytic domain (63) may also play a role in the cellular actions of thrombin in a variety of target tissues.

    A. Vasculature and heart: PAR1 antagonists—a novel potential approach for the treatment of cardiovascular diseases

    PAR1 can potentially regulate vascular function under both physiological as well as pathophysiological conditions (15, 42). A number of studies have revealed that thrombin and other agonists of PAR1 can affect the vascular tone. For example, before the discovery of PAR1, it was observed that thrombin can regulate vascular tone by an endothelial-dependent mechanism involving the release of nitric oxide (NO) (64). It is now recognized that this effect of thrombin is due to the activation of PAR1. Moreover, thrombin and PAR1 agonists can contract vascular SMC (VSMC) by a direct effect that requires extracellular Ca2+ (64). Thus, in isolated coronary artery and aorta preparations, PAR1 mediates relaxation (64, 65). In contrast, PAR1 agonists contract human placental and umbilical arteries (66). Furthermore, PAR1 mediates prostanoid generation and secretion, cellular contraction and barrier dysfunction, and enhanced expression of platelet-derived growth factor (PDGF) in human and bovine endothelial cells (67). Interestingly, progestins such as progesterone or gestodene up-regulate PAR1 expression and thereby stimulate thrombin-induced tissue factor-dependent surface procoagulant activity in the rat vascular system, suggesting a role of thrombin in hormone-induced thrombosis via PAR1 activation (68).

    Recent observations support a role of thrombin and PAR1 in regulation of functions normal (69, 70, 71, 73) and atherosclerotic (75) endothelium. In normal human arteries, PAR1 is mostly confined to the endothelium, whereas during atherogenesis, its expression is enhanced in regions of inflammation associated with macrophage influx, smooth muscle cell proliferation, and an increase in mesenchymal-like intimal cells (75). In vivo, a neutralizing antibody to PAR1 has been observed to reduce expression of mRNA for the proliferating cell nuclear antigen, an index of intimal and neointimal smooth muscle cell accumulation in rat arteries during balloon angioplasty. These data suggest that PAR1 regulates proliferation and accumulation of neointimal SMC during tissue repair (76).

    Thrombin and PAR1 agonists cause a rapid but transient contraction of endothelial cells in various tissues resulting in gap formation and increased permeability of plasma proteins and inflammatory cells. Several mediators are involved in this PAR1 modulated process such as cytokines, kinins, and biogenic amines (67, 72, 77, 78, 79, 80, 81, 82, 83, 84, 85, 86, 87, 88, 89, 90, 91, 92, 93, 94, 95).

    It was also revealed that PAR1 plays a crucial role during vascular ontogenesis. Interestingly, about 50% of the PAR1-deficient mouse embryos die at midgestation with bleeding from multiple sites. However, a PAR1 transgene driven by an endothelial-specific promoter prevented death of PAR1-deficient embryos (96), indicating that PAR1 modulates endothelial cell function in developing blood vessels, thereby contributing to vascular development and homeostasis in mice (88, 96).

    Vascular wall cells respond to the procoagulant factor Xa by an increase in intracellular Ca2+ ([Ca2+]i) and by assembly of this factor into prothrombinase complexes that even enhance this effect. Additionally, factor Xa stimulation of PAR1 leads to an increased production of tissue factor, a prothrombotic agent, underlining the important role of PAR1 for thrombosis (86). Together, these results point to a pivotal role of PAR1 in vascular homeostasis and thrombosis.

    PAR1 agonists are also mitogenic, stimulating proliferation of endothelial cells (97, 98), mediating endothelium-dependent relaxation to thrombin and trypsin in human pulmonary arteries (99), and causing the release of IL-6 from human microvascular endothelial cells (HMEC) (72). Because PAR1 up-regulates 1(I)-procollagen synthesis in VSMC, one may speculate that PAR1 plays a role in vascular wound healing (100).

    Finally, factors that regulate PAR1 function on endothelial cells have also been studied. For example, PAR1 is down-regulated by shear stress (101) and inflammatory mediators such as TNF- (102), and is directly modulated by thrombomodulin (103) in human endothelial cells. On the other hand, cyclic mechanical stress leads to an up-regulation of PAR1 in VSMC (104). In contrast to PAR2, little is known about the inflammatory mediators that regulate PAR1 expression in unstimulated or stimulated endothelial cells.

    As of yet, only a few studies exist about the role of GRKs in regulating PAR1 in endothelial cells. Only recently, Tiruppathi et al. (105) found a crucial role of the isoform GRK-5 on thrombin-induced desensitization of PAR1.

    It is also necessary to mention the recently demonstrated role of PAR1 in vascular associated pathological processes in which thrombin is involved. The role of PAR1 in thrombus formation was recently investigated in an animal model. Cook et al. (106) examined the role of PAR1 in intravascular thrombus formation in an experimental model of arterial thrombosis in the African green monkey. Using a blocking antibody to PAR1, this group demonstrated a dramatically diminished thrombin-stimulated aggregation and secretion of human platelets, whereas platelet activation induced by the PAR1 agonist SFLLR-NH2 was not affected. These results demonstrated that a specific blockade of PAR1 activation by thrombin can prevent arterial thrombosis in this animal model without significantly altering hemostatic parameters. The data suggest that PAR1 is crucially involved in this disease and is an attractive antithrombotic therapeutic target, although it probably has to be inhibited in concert with the low-affinity thrombin receptor PAR4 to fully prevent platelet activation (see Section II.A).

    A demonstrated role of PAR1 in intravascular thrombus formation and knowledge about side effects of thrombin inhibitors made appealing an idea to investigate the potency of PAR1 antagonists as antithrombotic therapeutic drugs. Indirect thrombin inhibitors like heparin and warfarin, which prevent the circulation of thrombin, or the newly developed direct thrombin inhibitor ximelagatran (465) are currently used for the treatment of cardiovascular diseases, e.g., myocardial infarction, atherosclerosis, and thrombosis. However, they carry potential bleeding liabilities as an undesirable side effect when administered in chronic patients. Because it is now well known that human platelet responses to thrombin are mainly mediated via PAR1 (19), direct inhibition of this receptor instead of the proteinase represents an attractive way to circumvent possible side effects by thrombin inhibitors.

    PAR1 possesses several extra- and intracellular sites that are crucial for its function and that might represent possible targets for antagonists (107). On the extracellular side, blocking of the cleavage site or the hirudin-like domain could inhibit N-terminal cleavage. Bradykinin or peptides derived thereof, also known as thrombostatins, have been shown to prevent thrombin-induced platelet aggregation (108). In addition, antibodies directed against both the cleavage site and the hirudin-binding site have been generated (106, 109). Binding of the antagonist to either the tethered ligand or extracellular loop 2 prevents interaction of the novel N terminus with the extracellular loop. Several small synthetic antagonists have been developed that block the binding site of the tethered ligand. However, peptide antagonists based on modifications of the tethered ligand were relatively unstable or showed only limited inhibitory activity (35). Thus, a second generation of indole or indazole-based peptide-mimetic antagonists was created that block binding of both the tethered ligand and the proteinase (110). One of these was recently shown to prevent thrombus formation in nonhuman primates (107). In addition, several nonpeptide inhibitors were developed. However, it was not reported whether these compounds were PAR1-specific (111). On the intracellular side, the G protein binding sites are possible targets for pharmacological intervention. One approach used the transfection of endothelial cells with a G minigene (112). However, this might suppress all signaling pathways that involve G. Another more specific approach used so-called "pepducins," peptides that were derived from the third intracellular loop of PAR1. These peptides were able to permeate the cells and also carried a membrane anchor (113, 114).

    In summary, two preclinical studies with primates demonstrated that PAR1 antagonists have indeed a therapeutic potential for the treatment of cardiovascular diseases (106, 107). However, extensive research needs to be done for the development of orally administered PAR1 antagonists because only those drugs are thought to be suitable for chronic patients (111).

    B. Platelets

    Activation of platelets by thrombin or APs specific for PAR1 is characterized by calcium influx; cytoskeletal reorganization; platelet aggregation; degranulation (17, 115, 116); thromboxane production (117); mobilization of the adhesion molecule P-selectin and the CD40 ligand to the platelet surface (119); stimulation of serotonin and epinephrine release (120); enhanced expression of CD62, PDGF (AB) and PDGF(BB) (121); and exposure of anionic phospholipids (phosphatidylserine, phosphatidylethanolamine) that support blood clotting (122). Thrombin has also been demonstrated to stimulate vascular endothelial growth factor release in megakaryocytes and platelets (123). The growth factor in turn is able to induce proliferation of endothelial cells. In addition, tissue inhibitor of matrix metalloproteinase (MMP)-4 that colocalized with MMP-2 in resting platelets was released upon platelet aggregation induced by collagen and thrombin. However, a direct effect using specific PAR1 agonists has not been shown yet (124). The density of human PAR1 on the platelet cell surface was estimated to be about 1500 molecules per resting platelet (125). A recently reported intronic polymorphism in the PAR1 gene leads to decreased expression on the platelet surface and thus to a lower response to the AP (126, 127). PAR1 is a high-affinity thrombin receptor in humans: receptor blocking using an antagonist, domain specific antibodies, or simply by desensitization led to inhibition of responses at 1 nM thrombin but only to attenuated responses at 30 nM (128).

    High-affinity thrombin binding of platelets is possibly associated via the membrane glycoprotein (GP)Ib-IX-V complex. This interaction is lost in patients suffering from the Bernard-Soulier syndrome, a disease where GPIb-IX-V is not expressed (129). The same condition can be mimicked using either monoclonal antibodies directed against GPIb or proteolytic removal of this protein from the platelet surface (130). Under these conditions, platelet aggregation is either delayed or requires higher thrombin concentrations.

    Recently, Schmidt et al. (131) reported that the gene encoding IQ motif containing GTPase activating protein-2, a scaffolding protein for filopodal extension during platelet activation, was identified within the human PAR gene cluster at 5q13, flanked by the PAR1 gene and encompassing the PAR3 gene. Only thrombin- or PAR1-AP-activated platelets showed a rapid translocation of IQ motif containing GTPase activating protein-2 to the platelet cytoskeleton. This suggests that a functional genomic unit evolved to mediate thrombin signaling events in humans (132).

    PAR1 on platelets also seems to play an important role during bacterial invasion: the cysteine proteinase gingipain from Porphyromonas gingivalis was reported to activate the receptor leading to uncontrolled activation of the host cells (28). The enzyme streptokinase derived from Streptococcus sp. is widely used in the treatment of coronary thrombosis. It functions as a plasminogen activator that forms an active complex with plasminogen of the host that is able to cleave PAR1 (133). In mouse platelets, PAR1 does not seem to play a role: PAR1 expression was hardly detectable, and a specific AP did not activate rodent platelets (23, 134, 135). Moreover, platelets derived from PAR1-deficient mice responded to thrombin-like wild-type platelets (135).

    C. Immune cells

    For some time, an important role for serine proteinases and PARs for the modulation of leukocyte effector functions has been proposed (136). As of yet, only limited data about the regulation of immune and inflammatory responses by PARs are available. It is well documented that PARs influence monocyte motility and chemotaxis, modulate pleiotropic cytokine responses, contribute to mononuclear cell proliferation, and induce apoptosis in various immune cells (84, 137). Recently, it was shown that PAR1 is capable of stimulating elastase secretion from macrophages (138). Moreover, functional thrombin receptors are expressed on human T lymphoblastoid cells (139). However, the receptor subtype(s) (PAR1, PAR3, and PAR4) have not been characterized as of yet. Human -thrombin stimulated five different T lymphoblastoid cell lines to increase intracellular free Ca2+ concentrations and to activate protein kinase C (PKC), whereas thrombin receptors were absent in B cell lines. Thus, PARs may regulate T cell function during inflammation and immune responses, but the precise mechanism is still unknown. Granzyme A from cytotoxic and helper T lymphocytes appears to interact with PAR1 in astrocytes to regulate thrombin function (140). Interestingly, granzyme A blocked the thrombin-induced platelet aggregation in a dose-dependent manner by cleaving PAR1, presumably downstream from its thrombin targeted activation site, thereby reducing the response to subsequent challenge with thrombin; but granzyme A itself did not induce a signal in thrombin-stimulated platelets. Thus, granzyme A may interact with PAR1 in a manner that is insufficient to cause aggregation, but sufficient to disarm the ability of the receptor to respond to thrombin. Finally, these observations demonstrate that granzyme A release occurring during immune responses within blood vessels would not directly cause platelet aggregation. Thus, the T cell-derived proteinase granzyme A seems to inhibit responses triggered by thrombin during inflammation or tissue injury.

    PAR1 modulates chemotaxis in inflammatory cells. Besides IL-8 secretion (71), thrombin induces production of monocyte chemoattractant protein-1 (MCP-1) in monocytes, probably via PAR1 (58). However, other PAR1-specific agonists were not used in this study. Therefore, it cannot be excluded that other thrombin receptors are involved. IL-8 secretion is up-regulated by interferon- (IFN-) and diminished by prostaglandin (PG)E2 (51), suggesting a role in cytokine modulation. Furthermore, PAR1 is capable of inducing IL-1 as well as IL-6 production in monocytes (82). These cytokines are known to be proangiogenic, implicating PAR1 in angiogenesis and tissue repair. However, IFN--differentiated growth-arrested U937 cells also respond to PAR1 agonist administration by overcoming cell arrest and revert to a high proliferation rate via regulation of p21CIP1/WAF1 and cyclin D1. Together, these results may help to explain how thrombin promotes tissue repair and unrestricted proliferation in malignant tissues (137, 141).

    Because thrombin, via PAR1, is chemotactic for large granular lymphocytes in humans (55), large granular lymphocytes can enhance the effects of PAR1 in patients with inflammatory disorders (142, 143, 144). Moreover, thrombin modulates activity of NK cells (48). For example, thrombin can enhance NK cell-mediated cytotoxicity and IL-2 production in rheumatoid arthritis, and PAR1 may therefore play an important role in this inflammatory process, as well as in NK cell responsiveness to IL-2.

    D. Airways

    Various cells in mammalian airways such as epithelial cells, SMC, and fibroblasts abundantly express PAR1 on the cell surface. Several recent studies also suggest that PAR1 may play an important role in inflammatory lung diseases such as neutrophilic alveolitis, pulmonary fibrosis, and asthma (50, 77, 91, 145, 146, 147, 148, 149).

    In the airways of different species, PAR1 exerts a dual role leading to stimulation of contraction on the one hand and relaxation on the other hand. For example, thrombin stimulates contraction of human bronchial rings (150) and constriction in guinea pig bronchi (as does AP) (151), but activates constriction as well as relaxation in mouse tracheal SMC (152, 153). In these mouse studies, trypsin, thrombin, and peptide agonists of PAR1, PAR2, and PAR4 induced relaxant responses of isolated tracheal smooth muscle preparations, which were mediated by a prostanoid, probably PGE2. This effect was abolished by indomethacin, the cyclooxygenase (COX)-2 inhibitor, nimesulide, and a prostanoid receptor antagonist (AH6809). PAR1 and PAR4 synthetic peptides induced a rapid, transient, contractile response that preceded the relaxant response. Interestingly, the relaxations but not the contractions were inhibited by indomethacin, indicating that this response is mediated by cyclooxygenase products.

    E. Gastrointestinal tract

    PAR1 was detected in the lamina propria, the submucosa, endothelial cells, and nerves of the gastrointestinal (GI) tract (154, 155, 156, 157). The GI tract expresses relatively high levels of PAR1 mRNA compared with other tissues, both in mice and humans. The functional role of PAR1 as a receptor mediating thrombin-induced effects in the GI tract is far from being resolved because thrombin can also interact with PAR3 and PAR4 (18, 20, 74). Several studies suggest a role for PAR1 in regulating GI motility. In guinea pigs, PAR1 mediates contraction of longitudinal smooth muscle tissue, dependent on extracellular Ca2+ (158, 159). In mouse intestine, PAR1 agonists modulate the function of L-type Ca-channels and also cause contraction (161). Similar results have been observed in rats (162, 163). Irradiation by fractionated X-radiation leads to up-regulation of PAR1 at the protein level, indicating a regulatory role of external trigger factors such as irradiation on PAR1 expression (164). Stimulation of PAR1 upon adhesion of pancreatic carcinoma cells to extracellular matrix proteins such as laminin, collagen IV, and fibronectin has been observed in a pancreatic carcinoma cell line (MIA PaCa-2) (165, 166). Other studies have pointed out a role for PAR1 in intestinal secretory pathways. Buresi et al. (156) have shown that selective PAR1 agonists stimulated Ca2+-dependent chloride secretion in intestinal epithelial cells. This PAR1-induced intestinal chloride secretion involves protein tyrosin kinase Src (Src), epidermal growth factor (EGF) receptor (EGFR) transactivation, activation of a MAPK pathway, phosphorylation of cytosolic phospholipase A2 (cPLA2), and cyclooxygenase activity. The presence of functional PAR1 on intestinal epithelium and the fact that PAR1 activation leads to ion secretion suggest that PAR1 might have important implications for intestinal barrier functions. PAR1 activation on intestinal surfaces could lead to a secretory response, thus contributing to diarrhea, a symptom of intestinal inflammation. More recent observations further suggest a role of PAR1 in intestinal inflammation. We have shown that intracolonic administration of PAR1 agonists caused inflammation and disruption of intestinal barrier integrity (53). Taken together, these results suggest a proinflammatory role of PAR1 in the gut (reviewed in Ref.167). However, additional studies using PAR1 antagonists and/or PAR1-deficient mice will have to verify the role of PAR1 in the pathogenesis of inflammatory bowel diseases.

    F. Kidneys and urogenital tract

    Recently, a crucial role of PAR1 in the cell-mediated renal inflammation of crescentic glomerulonephritis has been demonstrated in vivo (49). In wild-type mice treated with hirudin (a direct thrombin inhibitor, characterized by a bifunctional mechanism of inhibition, exclusive specificity and strong ability to bind the enzyme) and in PAR1-deficient animals, the proinflammatory effect of thrombin was significantly reduced. Moreover, treatment of wild-type mice with the PAR1 peptide agonist (SFLLRN-NH2) augmented the inflammatory response, suggesting that PAR1 plays an important role in renal inflammation in vivo. Unfortunately, as mentioned above, the peptide SFLLRN-NH2 can also activate PAR2 with a comparable potency, and therefore a role of PAR2 in renal inflammation cannot be ruled out by this study. Surprisingly, although highly expressed in human kidneys, data clarifying a role for PAR1 in inflammation of human kidneys are still lacking.

    Recent data revealed that PAR1 is involved in cellular invasion of a transfected canine kidney cell line. PAR1 agonists abrogated G(olf)-mediated invasion of MDCKts.src cells in collagen gels, indicating an important role for PAR1 in tumor metastasis (168, 169). Using the human urogenital cell line RT4, studies done in vitro have revealed that PAR1 and PAR2 agonists are capable of activating iPLA2 accompanied by release of PGE2, which may provide cytoprotection during an acute inflammatory reaction (170).

    G. Brain and peripheral nervous system

    Recent studies are in favor of an important role of thrombin and PARs in the brain under normal and pathophysiological conditions such as trauma, inflammation, or tumorigenesis (171, 172). Under pathophysiological conditions, i.e., during breakdown of the blood-brain barrier, circulating thrombin in the bloodstream may enter the central nervous system (CNS), leading to activation of PAR1 or PAR4. Additionally, neurons and glia cells are capable of generating prothrombin (173). Most data so far have been achieved by investigating PAR1 (13, 14, 174). In rat brain, PAR1 is expressed by neurons of the neocortex, cingulate cortex, subsets of thalamic and hypothalamic nuclei, discrete layers of the hippocampus, cerebellum, and olfactory bulb, as well as by astroglia (60). Striggow et al. (175) showed that all four PAR subtypes are expressed in rat brain. PAR1 expression was most abundant in the hippocampus, amygdala, and cortex. Interestingly, the expression of PAR1, PAR2, and PAR3 was up-regulated during experimentally induced ischemia.

    In human brain, PAR1 is expressed in neurons and astrocytes (175). Cultured rat glia cells (C6) express both functional PAR1 and PAR2 (176, 177). Moreover, PAR1 agonists induce up-regulation of inducible NO synthase (iNOS) in these cells (85) and promote neuronal survival after ischemia (178) or brain trauma (179). PAR1 also protects astrocytes and neurons from apoptosis induced by hypoglycemia and oxidative stress during inflammation (180). In contrast, thrombin may also exert cytotoxic effects on neurons and induce neurite retraction (181, 182), which may be at least in part due to PAR1. Functional studies further revealed that PAR1 agonists cause retraction of neurites by neuroblastoma cells (181) and induce Ca2+ mobilization by hippocampal neurons (178).

    In astrocytes, thrombin stimulates aggregation, morphological changes, and proliferation via PAR1 and induces intracellular caspase pathways leading to apoptosis in a cultured motor neuron cell line (NSC19) (183, 184). Both PAR1 and PAR2 stimulate enhanced proliferation of astrocytes (185). Friedmann et al. (186) reported a key role of thrombin in PAR1-mediated postinjury neuron survival. They demonstrated that the toxicity of thrombin on neurons can be controlled by down-regulation of PAR1 and/or release of antithrombin III by T cells. Interestingly, prothrombin expression was enhanced 24 h after injury, whereas PAR1, PAR3, and nexin-1 mRNA expression was unchanged (187). Nexin-1 is a thrombin inhibitor that also appears to control thrombin-PAR1 interactions in a rat trauma model (171, 187, 188). Thus, the whole factory needed to regulate and modulate thrombin function, including receptor as well as activating and inhibiting proteinases, can be generated in the brain. However, PAR1 also induces the reversal of astrocyte stellation in mice and rat (189, 190, 191). A comparable process is caused by exogenous or endogenous injuries of the CNS, triggering astrogliosis.

    Taken together, these data clearly indicate a functional role of PAR1 during inflammation and injury in the CNS. However, in some studies using thrombin as an agonist, activation of other PARs and a role of other molecules like thrombomodulin cannot be excluded. For example, thrombin itself up-regulates thrombomodulin in astrocytes in a dose-dependent manner (192). Thus, future studies taking into account all CNS-derived PARs, their proteinases, and proteinase inhibitors are necessary to fully explore the role of PARs in the brain. Finally, some authors suggest a role of thrombin in Alzheimer’s disease, amyotrophic lateral sclerosis, or HIV encephalitis, probably via activation of PARs (182, 193, 194).

    Accumulating data also indicate the role of PAR1 in the peripheral nervous system. In rat peripheral nerves, PAR1 is expressed by primary afferent neurons (195). From these observations, one may speculate that thrombin may also regulate inflammatory responses in the peripheral nervous system via PAR1. Indeed, very recently it has been demonstrated that PAR1 is expressed by a large proportion of primary spinal afferent neurons (155). Thus, thrombin directly signals to sensory neurons by cleaving PAR1. Moreover, administration of PAR1 agonists can induce plasma extravasation and edema, which were blocked by ablation of sensory nerves and administration of antagonists to the neurokinin-1 receptor, supporting the idea that thrombin cleaves PAR1 on sensory nerves to stimulate release of SP, which in turn interacts with the NK1 receptor to induce neurogenic inflammation. Thus, thrombin triggers neurogenic inflammation via PAR1 (155). However, because PAR2 and PAR4 mRNA are also expressed in the CNS, it cannot be excluded that proinflammatory effects of thrombin in the nervous system may, in addition, be mediated by other PARs. Functional studies further revealed that PAR1 agonists induce Ca2+ mobilization in enteric neurons, and regulate both excitatory and inhibitory neurons of the myenteric plexus in guinea pig small intestine, e.g., by releasing neuropeptides (196).

    Recently, Vergnolle and co-workers (197) as well as other groups (198) examined the effects of PAR1 agonists on nociceptive responses to mechanical and thermal noxious stimuli. Interestingly, intraplantar injection of selective PAR1 agonists induced an enhanced nociceptive threshold and withdrawal latency resulting in mechanical and thermal analgesia. However, thrombin was analgesic in response to mechanical, but not to thermal, stimuli. Moreover, application of PAR1-AP with carrageenan significantly reduced the hyperalgesia. Thus, thrombin may play a dual nociceptive-analgesic role, and future studies will be required to determine whether PAR1 agonists might be of therapeutic use for the treatment of pain.

    H. Signaling by proteinases via PAR1

    Investigations during the last few years revealed that the above-mentioned PAR1-mediated effects are transduced by various signaling pathways leading to diverse functions of PAR1 under physiological and pathophysiological conditions. These findings were recently reviewed (13). However, in the current work we focus on the signaling events mediated via PAR1 in different tissues and cell types (Table 2).

    1. Platelets.

    After its generation from prothrombin, thrombin plays multiple roles in the blood coagulation cascade that are mediated by interaction with a number of physiological substrates, effectors, and inhibitors. The accumulation of thrombin at sites of vascular injury provides the recruitment of platelets into a growing hemostatic plug. When added to human platelets in vitro, thrombin causes platelets to change shape, stick to each other, and secrete the contents of their storage granules. How this is accomplished is still not fully understood, but a major step forward occurred since the identification of PAR1 in the 1990s. This finding shed light on the way by which an extracellular protease may initiate intracellular events.

    Mammalian GTP-binding proteins (G proteins) fall into four families that are typically referred to by the designation of the -subunit: Gs, Gi, Gq/11, and G12/13. Human platelets express at least one member of the Gs family and four members of the Gi family (Gi1, Gi2, Gi3, and Gz), which, among other functions, stimulate or inhibit cAMP formation by adenylyl cyclase (199). In addition, platelets express one or more members of the Gq/11 family and members of the G12/13 family (G12 and G13) (199).

    A number of different G protein -subunits have been shown to bind to PAR1, including members of the Gi, Gq/11 and G12/13 families (200). However, the knowledge of these interactions in humans under physiological and pathophysiological conditions is still far from completion.

    Kim et al. (201) have demonstrated that thrombin and PAR1-AP, as well as PAR4-AP, induce Gi pathway stimulation in human platelets. Additionally, the authors demonstrated that thrombin, PAR1-AP, and PAR4-AP cause platelet aggregation independently of Gi stimulation (201). Offermanns et al. (202) also demonstrated a direct interaction between G12, G13, and PAR1. Additionally, it was reported that thrombin-mediated cleavage of the PAR1 (and PAR4; see Section IV.B) receptor leads to calcium-dependent and calcium-independent shape changes of human platelets in consequence of direct activation of both Gq and G12/13 pathways, respectively (115, 202). Therefore, current evidence suggests that PAR1 interacts with Gq, G12/13, and possibly Gi protein family members in human platelets. In turn, these data suggest an activation of downstream signaling cascade members after PAR1 stimulation on human platelets. Among such members are phospholipase C (PLC)-?, phosphoinositide 3 (PI3)-kinase, and monomeric G proteins.

    Indeed, PI3 kinase was demonstrated to play an important role in some PAR1 or thrombin-mediated cellular effects such as cytoskeletal reorganization, alterations in cell motility, cell survival, and mitogenesis. For example, thrombin is able to activate multiple PI3-kinase isoforms, including the recently discovered 110-kDa isoforms that can be directly activated by G protein ?-subunits (145, 150, 203). Stimulation of platelets by PAR1 leads to the activation of PI3 kinase, which is dependent on the small G protein Rho (204). Recently, Trumel et al. (205) have demonstrated that PI3 kinase plays an important role in the PAR1-dependent reorganization of the platelet cytoskeleton via myosin heavy chain translocation and stable association of signaling complexes with the actin cytoskeleton. Interestingly, the activity of small G proteins such as Rac and cdc42 in platelets may be regulated through PAR1 stimulation, but the role of PI3 kinase in these events remains to be determined (206). In their recent study, Vaidyula and Rao (207) provided evidence that in human platelets PLC-?2 plays a major role in responses to PAR1 and PAR4 activation, and that PLC-?2 is required for the sustained rise in [Ca2+]i concentration upon thrombin activation.

    Putting all of this together, current data suggest that thrombin activates human platelets by cleaving and activating PAR1 and PAR4. In turn, PAR1 activates the members of Gq/11, G12/13, and perhaps Gi families, leading to the activation of PI3 kinase, PLC-?, and monomeric G proteins (Rho, Rac, and possibly RapI), and also causes an increase of cytosolic Ca2+ concentration and inhibition of cAMP formation.

    Additionally, it is important to note that despite the fact that human platelets express both PAR1 and PAR4, these receptors appear to be activated by different thrombin concentrations. Cleavage of human PAR4 requires a higher concentration of thrombin than does cleavage of PAR1, and it is likely that PAR1 is the predominant signaling receptor at low thrombin concentrations (208, 209).

    Mouse platelets provide an interesting contrast to human platelets: whereas human platelets express functional PAR1 and PAR4, mouse platelets express PAR3 and PAR4, although in the latter case signaling appears to be mediated entirely by PAR4, with PAR3 serving to facilitate PAR4 cleavage at low thrombin concentrations (for details, see Section IV.A) (39, 40).

    2. Cells in the nervous system.

    PAR1, as mentioned above, appears to affect various processes in both the central and peripheral nervous systems (45, 210). Among such processes are: neuroinflammation and neurodegeneration, neuritogenesis, astrocyte proliferation, and synaptic plasticity. However, here we would like to focus on data concerning the involvement of PAR1 in intracellular signaling cascades in neuronal cells.

    PAR1 may have an effect on various intracellular signaling cascades within neuronal cells (171, 181, 189). By coupling to different G proteins, PAR1 affects a wide range of neuronal cell functions. For example, the necessity of PAR1 coupling to G12 was demonstrated for thrombin-stimulated DNA synthesis in 1321N1 astrocytoma cells (211). Interestingly, injection of antibodies directed against G12 abolished the thrombin-stimulated DNA synthesis (212). Additional studies performed on 1321N1 astrocytoma cells revealed that thrombin treatment causes a concentration-dependent rounding. One may speculate that such an effect of thrombin could be Rho-dependent. The Rho family of small GTPases is known to be involved in the control of cytoskeletal changes via modulation of actin polymerization. Indeed, this thrombin-induced rounding of 1321N1 astrocytoma cells was Rho-dependent and mediated via G12 (213).

    Moreover, LaMorte et al. (214) demonstrated binding of PAR1 to Gq by using the same cell line. The authors also showed that thrombin stimulation induces Ras-GTP complex formation and that Ras is required for PAR1-mediated activation of PLC in 1321N1 astrocytoma cells (214). This is especially interesting because thrombin is known to be a factor regulating the proliferation of astrocytes. This effect of thrombin appears to be associated with Go/i- and Gq-mediated pathways. Wang and coworkers recently shed some light on the downstream cascade events of PAR1 signaling (174, 210). Gq-Mediated PLC activity results in Ca2+ mobilization and activation of PKC, which phosphorylates a nonreceptor tyrosine kinase, proline-rich tyrosine kinase (Pyk2). Interestingly, the activation of other nonreceptor tyrosine kinases like Src and focal adhesion kinase (FAK) was also demonstrated after thrombin stimulation. Pyk2 is a factor that is able to connect GPCRs to ERK1/2 activation. Therefore, Pyk2, acting together with Src-tyrosine kinase, causes the activation of the ERK/MAPK pathway, which mediates the proliferative effect of thrombin. Additionally, the same group of authors demonstrated that the Go/i-mediated PI3 kinase pathway is also involved in thrombin-induced astrocyte proliferation (174).

    Thrombin is also known to exhibit both beneficial and unfavorable effects in hippocampal neurons and astrocytes (178, 180, 215). Donovan and colleagues (184, 216) found that tyrosine kinases, serine/threonine kinases, and the actin cytoskeleton are involved in both pro- and antiapoptotic effects of thrombin in neuronal cells. However, there was no involvement of Go/i and the PI3 kinase pathway observed in these thrombin effects.

    Additionally, Zieger et al. (217) demonstrated the existence of a novel PAR1-associated signaling pathway in the nervous system. According to these data, PAR1 participates in cAMP-independent PKA activation in SNB-19 glioblastoma cells. Moreover, PAR1 stimulation causes the activation of transcriptional factor nuclear factor B (NFB) in this cell type (217).

    In summary, besides morphological and proliferative effects described above that are mediated via G12/13, Gq, and Go/i, respectively, thrombin also affects activation of transcriptional factors such as activator protein-1 (AP-1) in neuronal cells (214).

    3. SMC.

    Thrombin exerts direct effects on vascular cells such as SMC and endothelial cells via interactions with PARs (218). As mentioned in Sections II.H.1 and II.H.2, PAR1 interacts with G12/13, Gq, and Gi to elicit diverse downstream signaling events. For example, thrombin-stimulated DNA synthesis and cell migration are associated with activation of the G13 signaling pathway in SMC. The G13 signaling cascade includes the activation of Rho and thus induction of cytoskeletal changes affecting cell migration (219). The Gq-dependent signaling pathway includes the activation of PLC, which in turn leads to MAPK phosphorylation and receptor tyrosine kinase transactivation, both necessary events in thrombin-mediated proliferation. It is interesting that the thrombin-mediated activation of MAPKs such as ERK1/2 was recently demonstrated in SMC. As noticed, PAR1 caused rapid phosphorylation, whereas PAR4 induced prolonged phosphorylation of these kinases in SMC (220). Additionally, Ghosh et al. (221) have shown that thrombin activates p38 MAPK in a time-dependent manner in VSMC. Furthermore, Sabri et al. (222) demonstrated that PAR1 agonists induce activation of Jun N-terminal kinase (JNK) and Akt/PKB in rat ventricular cardiomyocytes.

    After dissociation of the G protein heterodimer, G? interactions activate phosphoinositide 3-kinase, which promotes [Ca2+]i release that is required for SMC growth in response to thrombin stimulation (223).

    It is also interesting that PAR1 signaling can modulate gene transcription induced by cytokines in SMC. Thrombin, acting via PAR1, can block IL-6-induced signal transducer and activator of transcription 3/Sis-inducible factor-A (Stat3/SIF-A) activation (224).

    In summary, since the discovery of PARs, considerable progress in our understanding of thrombin signaling in SMC was achieved. This allowed some light to be shed on signaling events associated with SMC proliferation, migration, and synthesis of extracellular matrix proteins (e.g., collagen) after thrombin stimulation. However, additional studies are necessary to reveal the role of thrombin in SMC apoptosis, vascular lesion formation, and wound-healing associated pathways.

    4. Endothelial cells.

    Thrombin is known to affect various functions of endothelial cells. Among these are cell rounding, changes of cell-cell junctions, proliferation, barrier function, and permeability. These thrombin-induced effects appear to be mediated via PARs, particularly PAR1 (225, 226). A major step forward was reached in recent studies, which revealed important intracellular signaling events underlying PAR1-mediated effects in endothelial cells (227).

    PAR1 was demonstrated to interact with Go/i, Gq, and probably G12/13 in endothelial cells (228, 229). As well as it was demonstrated for neuronal cells, PAR1-mediated cytoskeletal changes in endothelial cells (cell rounding) are associated with the activation of RhoA (225). Additionally, in the same work it was demonstrated that PAR1-AP stimulation rapidly enhanced vascular permeability in a mouse skin assay (225). Moreover, it was found that binding of PAR1 to pertussis toxin (PTX)-sensitive G proteins (however only to Go, but not to Gi) is also necessary for thrombin-induced changes of endothelial barrier permeability (229). The activation of the MAPK signaling pathway by thrombin in endothelial cells was also demonstrated. This activation plays a crucial role in thrombin-induced effects of endothelial cell functions such as chemokine and cytokine production as well as the expression of cell adhesion molecules (89, 230, 231).

    The role of thrombin and PAR1 in the activation of transcriptional factor networks was intensively investigated in recent publications. Malik and coauthors (228, 231, 232) studied the involvement of PAR1 in the activation of NFB in endothelial cells. Their studies showed that Gq and the G? dimer are responsible for NFB activation and intercellular adhesion molecule-1 (ICAM-1) transcription in endothelial cells induced by the PAR1 agonist thrombin and the PAR1-specific AP TFLLRNPNDK. Furthermore, transfection experiments strongly supported simultaneous activation of G?/PKC-/p38 and G?/PI3-kinase pathways that converge into the Akt pathway, leading to subsequent NFB activation and ICAM-1 expression (228, 231, 232). Moreover, thrombin-induced stimulation of vascular cell adhesion molecule-1 (VCAM-1) production involves the inducible binding of p65 NFB to a tandem NFB motif in the 5' flanking region (233). Taken together, these findings suggest that in the case of ICAM-1 and VCAM-1, p65 NFB is necessary for transducing the thrombin response in endothelial cells.

    Additionally, thrombin-mediated induction of VCAM-1 was shown to involve the inducible binding of GATA-2 to a tandem GATA motif in the upstream promoter region (234). Interestingly, the effect of thrombin on GATA-2 DNA binding and transcriptional activity was found to be mediated by a PI3K, PKC--dependent signaling pathway (233).

    It is important to note that thrombin effects on transcriptional factor networks have been more deeply investigated in endothelial cells than in other cell types. These data explain, at least in part, thrombin-mediated effects on the production of cell adhesion molecules and some chemokines in endothelial cells. Subsequently, this accounts for thrombin effects at leukocyte migration via endothelial barrier (ICAM-1 serves as a ligand for leukocyte ?2 integrins and promotes leukocyte adhesion) and leukocyte recruitment.

    5. Immune cells.

    The participation of PARs in T cell signaling pathways has also been demonstrated (235). Recently, Bar-Shavit et al. (236) have examined a possible involvement of Vav 1 in PAR-mediated signaling in human Jurkat T cells. The Vav family has three known members in mammalian cells (Vav, Vav2, and Vav3) and one in nematodes (CelVav) (237). Tyrosine phosphorylation of Vav1 regulates its activity as a guanine-nucleotide exchange factor for the Rho-like small GTPases RhoA, Rac1, and cdc 42, which affect cytoskeletal reorganization and activation of stress-activated protein kinases/JNKs. Bar-Shavit et al. (236) clearly showed that activation of PARs induces tyrosine phosphorylation of Vav1 in Jurkat T-leukemic cells. Because -chain-associated protein kinase of 70 kDa (ZAP-70) and SH2-domain containing leukocyte-specific phosphoprotein of 76 kDa (SLP-76) have been shown to associate physically with Vav1 and because this association is critical for normal functioning of T cells, it was tempting to investigate whether ZAP-70 and SLP-76 are involved in PAR-induced signaling cascades. Indeed, an increase of tyrosine phosphorylation of ZAP-70 and SLP-76 was observed after activation of Jurkat cells with PAR-APs. Moreover, phosphorylation of Vav1 in response to PAR stimulation was dependent on p56lck (236). Unfortunately, because nonselective PAR-APs were used in the studies done with the Jurkat cells (236), it is difficult to distinguish between the effects of PAR1 and PAR2 in this T cell model system. Furthermore, because of the lack of the ability of PAR3-derived peptides to activate PAR3 (39), the role of PAR3 in T cell signaling remains unknown.

    6. Factor Xa signaling mediated via PAR1.

    Most of the PAR1-associated signaling events have been observed subsequent to cell stimulation by thrombin. However, recently a possible role of PAR1 in coagulation factor Xa-mediated signaling has been demonstrated (238, 239, 240). The coagulation proteinase factor Xa is generated at sites of vascular injury and inflammation after formation of the tissue factor/VIIa (TF/VIIa) complex. Coagulation factor Xa has been shown to be mitogenic for SMC (238) and elicits inflammatory responses in endothelial cells. Riewald and Ruf (241) have presented new data indicating the involvement of PAR1 in Xa-mediated signaling. They used HeLa cells expressing only PAR1 and demonstrated that factor Xa induces NFB activation and MAPK phosphorylation in these cells. Inhibition studies with specific antibodies revealed that factor Xa responses were mediated via PAR1 (241). Thus, factor Xa (or potentially other serine proteinases) may substitute for thrombin in proteolytic signaling via PAR activation. In endothelial cells, which also express PAR2, Camerer et al. (240) have shown that factor Xa signaling was mediated not only by PAR1, but also to a large extent via PAR2. Together, PAR-1 and PAR-2 appear to account for more than 90% of factor Xa signaling in endothelial cells (240).

    In summary, the multiple biological as well as inflammatory and immune responses of thrombin that are mediated by PAR1 [including 1) vasoregulation; 2) increased vascular permeability; 3) cellular adhesion and infiltration of leukocytes; 4) angiogenesis; 5) stimulation of the production of inflammatory mediators such as cytokines, neuropeptides, NO and prostanoids, for example; 6) regulation of extracellular matrix proteins; and 7) induction of signal transduction pathways which are involved in immunomodulation] suggest an important role of PAR1 during inflammation and immune response.

    III. PAR2 in Inflammation and Immune Response

    PAR2 is expressed in brain, lia (DRG), eye, airway, heart, GI tract, pancreas, kidney, liver, prostate, ovary, testes, and skin (21, 22, 24, 154, 242, 243, 244, 245) and is found in various cell types such as epithelial cells, endothelial cells, SMC, osteoblasts, as well as immune cells such as T cells, neutrophils, mast cells, or eosinophils (97, 154, 246, 247, 248, 249, 250, 251, 252, 253, 254). On the other hand, platelets do not express PAR2 (248). Recent findings point to an important role for PAR2 under physiological and pathophysiological conditions in many tissues (Table 2). However, the endogenous enzymes responsible for activating PAR2 in many tissues remain to be determined. Many endogenous or exogenous trypsin-like enzymes may cleave and activate PAR2. Expression of trypsinogen-2 mRNA and its translation product has been demonstrated in endothelial cells (255). Interestingly, various types of human cancer cells secrete enzymes with trypsin-like specificity (255) that may activate PAR2. In human skin, trypsinogen-4 generated by keratinocytes and trypsinogen-2 from human dermal microvascular endothelial cells can activate PAR2 in vitro (M. Steinhoff, unpublished observation). Another candidate is mast cell tryptase, a major secretory protein of human mast cells. Mast cells from mice and rats are more heterogeneous regarding their protease content, although they also produce proteases with tryptic specificity. Tryptase has been shown to activate PAR2 on epithelial as well as endothelial cells and neurons (23, 196, 243, 251, 256). The observation that tryptase can activate PAR2 suggests a role of this receptor in humans under circumstances when mast cells are involved, e.g., during inflammation, hypersensitivity reactions, and wound repair (24, 257). However, the ability of human tryptase to activate PAR2 appears to be restricted by receptor glycosylation at an N-terminal residue just proximal to the receptor’s cleavage activation site (258, 259). Notwithstanding, the effects of tryptase on cells in vitro often mimic those of PAR2 activation: tryptase up-regulates IL-1? and IL-8 secretion, enhances the presence of intracellular adhesion molecules/selectins on endothelial cells, mediates accumulation of neutrophils and eosinophils, produces vascular leakage, and is mitogenic for epithelial cells, fibroblasts, and SMC (260, 261, 262, 263). That said, direct proof that PAR2 mediates the effects of tryptase during inflammation in vivo is still lacking.

    A. Vasculature

    In the vasculature, PAR2 has many effects that are proinflammatory. Agonists of PAR2 induce relaxation in the rings of rat aorta or porcine coronary artery dependent on endothelial NO synthase activity (246, 248, 250). This effect is abolished in the absence of endothelium (246, 250). In contrast, trypsin stimulates contraction of the rabbit aorta in the absence of endothelium (264). In the intact rat, iv injection of SLIGRL-NH2 produces a marked fall in blood pressure, consistent with release of NO from endothelial cells (248). Furthermore, PAR2 agonists increase IL-6 production (72), induce von Willebrand factor release, and serve as a mitogen for human umbilical vein endothelial cells (HUVEC) (97, 265, 266).

    Moreover, it was demonstrated that some inflammatory mediators are able to affect the expression of PAR2 in endothelial cells. In HUVEC, PAR2 mRNA is up-regulated by TNF- and IL-1, cytokines that act together to orchestrate the acute inflammatory response (25, 265).

    In summary, PAR2 mediates vasodilation, plasma protein extravasation, as well as endothelial cell proliferation in the cardiovascular system. Thus, this receptor can be regarded as an important factor in neovascularization. Moreover, PAR2 can be considered as a vascular sensor for trypsin-like proteinases of the coagulation cascade, which play an important role in cardiovascular medicine.

    B. Immune cells

    Although information has been acquired about the role of PAR2 in epithelial and endothelial function, relatively little is known so far about the role of this receptor in the immune system. As was recently demonstrated, PAR2 is expressed by various cells involved in immune response, such as T cell lines, eosinophils, neutrophils, and mast cells (147, 160, 249, 252, 267). Accumulating evidence points to a role of PAR2 in the regulation of leukocyte function. Some of the PAR2- mediated effects on leukocyte behavior have been observed in vivo in rodents (262). In this model, it has been reported that the PAR2-AP (SLIGRL-NH2) caused a significant increase in leukocyte migration into the peritoneal cavity after ip injection. Furthermore, PAR2-APs induced a significant increase in leukocyte rolling and adherence by a mechanism depending on the release of platelet-activating factor (262). Lindner et al. (268) have also demonstrated the importance of PAR2 for leukocyte adhesion and rolling. In a model of acute inflammation induced by tissue trauma, they showed that PAR2-deficient mice have a decreased leukocyte rolling capacity compared with wild-type mice. Conversely, activation of PAR2 by a specific PAR2-AP in wild-type mice induced a significant reduction of leukocyte rolling velocity, an increased leukocyte rolling flux as well as increased leukocyte adhesion (268).

    Howells et al. (252) reported that a specific PAR2-AP as well as trypsin are able to activate PAR2 on human neutrophils. PAR2 activation caused shape changes and up-regulation of CD11b/CD18 in these cells. In a coculture of human neutrophils with endothelial cells, PAR2 agonists induced L-selectin shedding and CD11b/CD18 up-regulation in neutrophils (268). We demonstrated that PAR2 agonists induce [Ca2+]i release in human neutrophils and enhance neutrophil motility in 3-D collagen gel lattices (269). Moreover, Lourbakos et al. (27) demonstrated that a bacterial proteinase, gingipain-R from P. gingivalis, activates PAR2 on human neutrophils. Because many bacteria with pathogenicity in humans produce serine proteinases with trypsin-like activity, it can be assumed that certain pathogenic effects of bacteria in many tissues can be mediated through PAR2. For example, Lourbakos et al. (28) reported that activation of PAR2 by bacterial gingipain R leads to IL-6 secretion in human oral epithelial cells. In mouse airways, PAR2 agonists were capable of inhibiting the recruitment of neutrophils induced by bacterial proteinase (270), suggesting a direct role of PAR2 on neutrophil function in vivo. From these data, one may speculate that PAR2 may serve as a receptor for serine proteinases derived from bacteria, viruses, or even fungi and may thus directly trigger inflammatory or immune responses in vivo induced by microorganisms. However, this intriguing hypothesis remains to be verified.

    Recent work has shed some light on the role of PAR2 on eosinophil function. Miike et al. (267) clearly demonstrated that human eosinophils express functional PAR2. Furthermore, trypsin as well as a specific PAR2-AP was able to induce degranulation and superoxide production in human eosinophils. More recently, Temkin et al. (79) reported that human mast cell tryptase induced up-regulation of IL-8 mRNA and caused IL-8 release from human eosinophils. A comparable effect of tryptase on eosinophil IL-6 expression and release was also observed. This effect of tryptase appears to be activated via the protein kinase/AP-1 signaling pathway (79). Schmidlin et al. (271) have also demonstrated that PAR2 mediates infiltration of eosinophils and hyperreactivity in allergic inflammation of the airway. Furthermore, PAR2 activation releases survival factors such as granulocyte-macrophage colony-stimulating factor (GM-CSF) from eosinophils (272). Together, these results are clearly in favor of an important role of PAR2 in eosinophil regulation and in inflammatory/allergic diseases in which eosinophils are involved.

    Several T cell lines also express PAR2 (244). Trypsin as well as a synthetic PAR2-AP induce [Ca2+]i mobilization in the Jurkat and HPB.ALL T cell lines (249). In Jurkat cells, the involvement of PARs, possibly PAR2, in Vav1-mediated signaling was clearly demonstrated (236).

    The localization of PAR1 and PAR2 on human mast cells was first reported by D’Andrea et al. (147). As shown by immunohistochemistry, PAR2 was localized not only on the cell membrane of mast cells, but also on the membrane of the intracellular tryptase-positive granules (147). Moreover, rat peritoneal mast cells also express PAR2 mRNA (273). However, the precise role of PAR2 in mast cells especially with regard to a potential autocrine regulatory mechanism via tryptase activation requires additional investigation. In summary, these results provide evidence for a role of serine proteinases in directly regulating immune cells and immune responses via PARs. An understanding of the molecular mechanisms by which PAR2 is involved in immune regulation will aid the design of new antiinflammatory drugs that can target proteinase-activated receptors.

    C. Arthritis

    Arthritis is a chronic inflammatory disease that is characterized by joint swelling, vasodilatation, edema, hyperemia, pain, and recruitment of inflammatory cells, especially neutrophils and macrophages, but also lymphocytes. Very recently, it was demonstrated that PAR2 participates in the pathophysiology of arthritis (274). Using PAR2-deficient mice, the authors showed that PAR2 deficiency results in a marked reduction of swelling responses of experimentally induced monoarthritis. However, no differences were observed concerning joint damage when compared with wild-type mice. Using an enzymatic approach (?-galactosidase), they also showed positive ?-gal staining in endothelial cells of blood vessels. Two weeks after induction of monoarthritis, however, a significant increase of ?-gal staining of extravascular inflammatory cells was observed in PAR2+/+ mice compared with PAR2–/– mice. Examination of the knee joint showed dramatic arthritic changes in PAR2+/+ mice characterized by synovial hyperplasia, an enhanced inflammatory infiltrate, and cartilage damage. Thirty days after the induction of monoarthritis, the cartilage was completely replaced by pannus in PAR2+/+ mice. In contrast, PAR2–/– mice showed an intact appearance of the cartilage similar to normal and control joints. This is in favor of an important role of PAR2 as a mediator of proinflammatory responses in the joints. Moreover, a novel synthetic PAR2 peptide agonist (ASKH95, phenylacetyl-LIGKV-OH) revealed proinflammatory effects after intraarticular injection like synovial hyperemia and joint swelling. These results clearly demonstrate that ASKH95 induces signs of chronic inflammation such as long-lasting swelling responses, vasodilatation, and tissue destruction. Interestingly, ASKH95 showed a longer-lasting proinflammatory effect in the tissue compared with the classical SLIGRL PAR2 agonist peptide. This may be due to different rates of degradation, longer receptor activation by ASKH95, or most likely different lipophilicity. The underlying mechanisms of the PAR2-mediated inflammatory effect in this chronic disease, however, are not fully explored yet. Potential mediators may be cytokines or prostanoids, for example. Thus, PAR2 may play an essential role in the pathophysiology of arthritis, and PAR2 antagonists may be beneficial for the treatment of this chronic inflammatory disease.

    D. Skin

    PAR2 may play an important role in cutaneous inflammation. Keratinocytes and dermal endothelial cells express functional PAR2 and are important targets of inflammatory mediators (247, 275, 276). In humans, PAR2 is also expressed by dermal immunocompetent cells, which so far have not been characterized (24), and by dermal sensory nerves (277). Potential endogenous activators of PAR2 in human skin are mast cell tryptase, proteinases of the trypsin-family produced by keratinocytes (trypsinogen-4), and proteinases of the fibrinolysis cascade, such as factor VII/Xa, which could be released during inflammation and wound healing. Exogenous activators of PAR2 may be serine proteinases generated by bacteria, fungi, and house dust mites, although direct evidence on keratinocytes is still lacking. Indirect evidence for bacterial-induced activation of PAR2 is coming from the observation that bacterial gingipain activates the receptor on buccal keratinocytes (28). At certain stages of atopic dermatitis or psoriasis, mast cells are in close contact to basal keratinocytes or are recruited into the epidermal layer (278). Numerous mast cells are found in the dermis in close proximity to blood vessels and at the dermal-epidermal border in atopic dermatitis and psoriasis. Thus, tryptase or other skin-derived proteinases may activate PAR2 to induce inflammatory changes and thereby contribute to the pathophysiology of atopic dermatitis and psoriasis. It is well known that intradermal injection of tryptase into the skin results in vasodilatation and erythema, followed by leukocyte infiltration and local induration (279), indicating a role of tryptase in cutaneous inflammation. Moreover, tryptase is an important long-term marker in body fluids for systemic anaphylaxis and other immediate hypersensitivity allergic reactions (280). Thus, it is possible that PAR2 mediates the systemic effects of tryptase in certain skin diseases.

    Serine proteinases other than tryptase may also activate PAR2. Cytotoxic T cells, which express PAR2 (249), also produce the trypsin-like enzyme granzyme A (281), suggesting a potential role for PAR2 in regulating T cells in inflammatory dermatoses. In summary, proteinases in the skin are signaling molecules that may directly regulate the function of target cells by activating PAR2. PAR2 as well as PAR1 agonists regulate cytokine production such as IL-8 (282), IL-6, and GM-CSF (283) in cultured human keratinocytes. Furthermore, PAR2 may also regulate proliferation and differentiation in the skin. Similar to TGF-?, PAR2 agonists inhibit proliferation or differentiation of human neonatal keratinocytes, whereas PAR1 agonists stimulate proliferation (275). In contrast, agonists of both PAR2 and PAR1 are mitogenic for endothelial cells (248). Thus, our observations in normal and inflamed human skin suggest a functional role of PAR2 in different cell types of human skin during inflammation.

    In human skin, PAR2 is only weakly expressed by microvascular endothelial cells. Studies done in vitro revealed that PAR2 is functionally expressed by human dermal microvascular endothelial cells and that PAR2 agonists induce up-regulation and release of IL-6 and IL-8 in a concentration-dependent fashion (263). Moreover, PAR2 regulates expression of cell adhesion molecules in primary cultures of human endothelial cells (284) and cell lines (268). These findings are supported by in vivo findings demonstrating a role of PAR2 in the adhesion, rolling, and extravasation of leukocytes in rat venules. Finally, PAR2 appears to play a role in the regulation of leukocyte/endothelial interactions in humans and mice in vivo, as shown for atopic dermatitis or experimentally induced contact dermatitis (285). Together, these results strongly support the idea that PAR2 plays an important role in leukocyte/endothelial interactions during inflammation and immune response.

    The knowledge about the signaling cascades involved in PAR2-induced inflammatory and immune responses by keratinocytes is still incomplete. Recently, Kanke et al. (286) have shown that PAR2 agonists stimulate activation of inhibitory B kinases (IKK) and IKK?, and also stimulate NFB-DNA binding. Our own data using primary human keratinocytes also clearly demonstrate that PAR2 activates NFB (287). Moreover, PAR2 is directly involved in the NFB-mediated regulation of keratinocyte ICAM-1, which is an important molecule for the regulation of keratinocyte immune responses such as host defense, allergy, and recruitment of inflammatory cells into the epidermis (287).

    Finally, recent data suggest a role of PAR2 during cutaneous inflammation in vivo. In a murine model of experimentally induced allergic contact dermatitis, a proinflammatory role of PAR2 agonists was demonstrated (285, 288). Stimulation of PAR2 resulted in increased ear swelling responses with a maximum after 48 h. Independently, our group confirmed the proinflammatory effects of PAR2 agonists in a model of experimentally induced allergic and irritant contact dermatitis using PAR2-deficient mice. Moreover, we examined the underlying mechanisms responsible for PAR2-induced effects during contact dermatitis (285). These results clearly indicate that PAR2 is involved in edema formation, plasma extravasation, up-regulation of cytokines (IL-6), cell adhesion molecules (ICAM-1), and selectins (E-selectin) in mice. Intravital microscopy studies further showed that velocity and adhesion of leukocytes is impaired in PAR2-deficient mice compared with controls. Moreover, in vivo studies by microdialysis confirmed a role of PAR2 in mediating vascular responses such as edema formation and plasma extravasation also in human skin. Finally, ex vivo studies in skin biopsies of human volunteers are in favor of a role for PAR2 in selectin regulation of dermal blood vessels because PAR2 agonists induced a marked increase of E- selectin immunoreactivity in dermal endothelial cells in comparison to control tissues. Together, these results support a regulatory role of PAR2 during cutaneous inflammation and immune response.

    E. Airways

    Several observations are in favor of an important role of PAR2 in airway inflammation. Firstly, tryptase (251, 289) seems to play an important role in airway inflammation and hyperresponsiveness (253). Secondly, PAR2 is markedly up-regulated after exposure to proinflammatory stimuli or cytokines (265) that have been shown to play a role in chronic airway diseases. Morphological studies have demonstrated PAR2 immunoreactivity in endothelium and smooth muscle of bronchial vessels. These observations are consistent with a similar distribution in other tissues and with the known ability of PAR2 agonists to cause arterial relaxation as well as contraction in certain arteries (290). Moreover, PAR2 is expressed by bronchial, tracheal, epithelial and SMC which may result in direct or indirect bronchomotor effects leading to pathophysiological conditions in the airways such as asthma or chronic obstructive pulmonary disease. Interestingly, PAR2 agonists are capable of causing bronchoconstriction (291). Whether this response is dependent on secondary effects such as the generation of bronchoactive peptides or other mediators remains to be determined. However, studies done in vitro suggest further that PAR2 agonists also produce a relaxant effect in isolated main bronchi (292). Thus, PAR2 agonists may cause dose-dependent bronchoconstrictor or bronchodilatator effects in the airways, with high agonist concentrations favoring bronchoconstriction. NO and prostanoids like PGE2 may be involved in this process, because potentiation of PAR2-induced bronchoconstriction has been observed upon inhibiting NO synthase or after prostanoid generation and COX-2 activation (153). In conclusion, there is strong evidence that PAR2 stimulation activates contractile and dilator mechanisms in the airways. The reason why the predominant effect in guinea pig airways in vivo is bronchoconstriction, however, is not known at present. In addition to the potentially bronchoprotective effects observed by Cocks et al. (292) in rat bronchial preparations, others have shown that PAR2 may also mediate bronchoprotection in guinea pig airways (293). However, other observations show that PAR2 can act as an enhancer of histamine-mediated contraction (294). The study of Cocks et al. (292) showed that trypsin colocalized with PAR2 in airway epithelial cells, where PAR2 activation resulted in relaxation of airway preparations by the release of cytoprotective PGE2. Prostanoids may be important mediators of PAR2 activity in the airways not only in the lung, but also in the GI tract (152, 153, 253, 295). It is tempting to speculate that tryptase and/or other proteinases with trypsin-like activity, released in close proximity to the cells expressing PAR2, may stimulate this receptor during airway inflammation. Finally, intensive staining of PAR2 in the apical region of the airway epithelium might suggest an activating role of proteinases, for instance bacterial proteinases, present in the airway lumen (296). Very recently, PAR2 was shown to stimulate the release of MMP-9 in a human epithelial cell line (297), providing a strong hint that PAR2 is involved in the orchestration and reorganization of lung extracellular matrix proteins.

    It was also revealed that PAR2 stimulation of airway epithelial cells leads to the release of eosinophil survival-promoting factors such as GM-CSF (272). Moreover, in human respiratory epithelial cells, PAR2, like PAR1, stimulates IL-6, IL-8, and PGE2 release (77).

    Profound evidence for an involvement of PAR2 in allergic inflammation of the airways is provided by the recent work of Schmidlin et al. (271). To study the involvement of PAR2 in airway inflammation, they used PAR2-deficient mice and mice overexpressing human PAR2 (PAR2tg). PAR2 is known to be up-regulated by inflammatory agents. Sensitization of wild-type mice by injection of ovalbumin led to infiltration of immune cells into the lumen of the airways and induced hyperreactivity, whereas both infiltration of eosinophils into the lumen and airway hyperreactivity was exacerbated in PAR2tg mice. In contrast, infiltration of immune cells and airway hyperreactivity was markedly diminished in PAR2-deficient mice. Additionally, in PAR2–/– mice the IgE response was strongly reduced, implicating a role of PAR2 in immune response. Remarkably, intranasal administration of PAR2-AP in wild-type mice stimulated increased recruitment of macrophages, confirming the proinflammatory role of PAR2 in the airways, whereas altered PAR2 expression mainly influenced eosinophil infiltration.

    In summary, evidence exists for a proinflammatory as well as antiinflammatory role of PAR2 in airway inflammation. This dual role may be explained by the various tissues, cells, methodological approaches, and inflammatory model systems (acute vs. chronic, mouse strains) that have been used to evaluate this role of PAR2 in the airways. So far, it appears that PAR2 is a sensor receptor that releases proinflammatory mediators in the early phase and antiinflammatory molecules in the late phase of "regulated inflammation." Under "dysregulated" inflammatory situations such as chronic disease states, PAR2 may exert antiinflammatory effects, as shown in a colitis model (298), or proinflammatory effects, as demonstrated in an asthma model (271). Thus, additional studies using human and mouse (knockout) models of pulmonary dysfunction are necessary to fully explain the role of PAR2 during inflammation and immune response.

    F. Brain and peripheral nervous system

    By immunohistochemistry, PAR2 has been localized in various compartments of the nervous system such as brain, spinal cord, DRG, and peripheral nerves, with different receptor densities found during various stages of mouse development (299). At d 14, PAR2-staining was observed throughout the mouse brain. Among the hippocampal formation, PAR2 was localized in the subiculum, in pyramidal cells throughout the CA1, CA2, CA3, and hilus region, and in granular cells of the dentate area. Additionally, staining for PAR2 was observed in all cortical layers, amygdaloid nuclei, striatum, thalamus, and hypothalamus (175). Peripheral nerves were intensively stained after d 14. Similarly, PAR2 has also been detected in human brain (276). In rat brain, PAR2 is also expressed by neurons of the hippocampus (175, 300), meningeal cells (301), as well as astrocytes (176, 185). PAR2 is abundantly expressed in various subareas and is transiently up-regulated during oxygen and glucose deprivation (175). Moreover, PAR2 activation is cytotoxic for isolated rat hippocampal neurons in a concentration-dependent manner (300).

    Domotor et al. (302) recently reported that agonists of PAR2 induce Ca2+ responses in brain microvascular endothelial cells. This effect may be modulated by elastase and plasmin, which regulate PAR2 signaling. Studies of the actions of PAR2 agonists on CNS targets suggest an important role of PAR2 during injury, growth, apoptosis, and probably memory.

    Very recently, PAR2 has been localized on rat sensory neurons (243). Moreover, functional data strongly support the idea that the peripheral nervous system is directly regulated by PAR2 during neurogenic inflammation. Neuropeptides such as calcitonin gene-related peptide and substance P (SP) from primary spinal afferent neurons are known as important mediators of neurogenic inflammation in many organs. Because of the close proximity of tryptase-containing mast cells to spinal afferent fibers, and because agonists of PAR2 cause effects similar to those of tryptase in many tissues, comprising many of the characteristics of neurogenic inflammation, it was speculated that serine proteinases or other PAR2 agonists may activate PAR2 on sensory neurons to stimulate neuropeptide release and may thus regulate inflammation. Indeed, it was shown that agonists of PAR2 can induce inflammation by a neurogenic mechanism that depends on the release of calcitonin-gene related peptide (CGRP) and SP from primary spinal afferent neurons. However, there also appears to be a nonneurogenic component of PAR2-induced inflammation because granulocyte infiltration triggered by PAR2 in a paw edema inflammation model was unaffected by sensory denervation or by neuropeptide receptor antagonists. Thus, PAR2 agonists may also act directly on endothelial cells or on neutrophils themselves to stimulate inflammation-related granulocyte adhesion and infiltration (14, 250). It is possible that serine proteinases may activate PAR2 on spinal afferent neurons under pathophysiological inflammatory circumstances. In a similar manner, serine proteinases may regulate enteric neuronal function by cleaving PAR2 in a Ca2+-dependent manner (196). Trypsin, which is released by airway epithelial cells, but not the PAR2 agonist peptide sequence SLIGRL, induced a marked contraction in guinea pig bronchi via SP release. Thus, non-PAR mediated effects of trypsin may also regulate neuropeptide function in the lung (303).

    Several reports strongly suggest a role of PAR2 in the central transmission of pain and nociception in rats and mice (304, 305, 306). Intraplantar administration of low doses of PAR2 agonists, which do not induce inflammation, resulted in a long-lasting hyperalgesia. Interestingly, these effects of PAR1 agonists were more efficacious in comparison with other inducers of hyperalgesia such as PGE2 (0.3 μg). In the spinal cord, PAR2 agonists induced an up-regulation of c-fos, a marker of activated nociceptive neurons (305, 307, 308). Another study also reported a role of PAR2 in a rat visceral pain model (308). PAR2 agonists administered in vivo clearly increased abdominal colonic contractions. This effect was inhibited by an NK1 receptor antagonist, but not by the PG inhibitor indomethacin (308). However, additional studies are necessary to fully explain the underlying direct or indirect effects of PAR2-induced nociception. For example, neuropeptides released from neurons upon PAR2 stimulation may activate the release of nociceptive mast cell mediators such as kinins or prostanoids (309).

    The observation that PAR2 agonists are involved in the transmission of neurogenic inflammation and pain subsequently draws the question whether PAR2 may be involved in the central transmission of pruritus. Itching is one of the most frequent symptoms in dermatological diseases and accompanies inflammatory and immune responses of many diseases such as hypersensitivity reactions or urticaria, for example. Indeed, neuronal PAR2 appears to be involved in the induction of pruritus in human skin (277). Moreover, the endogenous PAR2 agonist tryptase was increased up to 4-fold in atopic dermatitis patients, and PAR2 expression was markedly enhanced on primary afferent nerve fibers in skin biopsies of atopic dermatitis patients. Intracutaneous injection of specific PAR2 agonists provoked enhanced and prolonged itch when applied intralesionally. Thus, PAR2 activation on cutaneous sensory nerves may be a novel pathway for the transmission of itch and inflammatory responses during atopic dermatitis and probably other skin diseases. PAR2 antagonists may be promising therapeutic targets for the treatment of cutaneous neurogenic inflammation and pruritus (277).

    In summary, PAR2 may play an important regulatory role in the central and peripheral nervous system in normal and disease states, including neurogenic inflammation.

    G. Digestive tract and pancreas

    In rats, Kawabata et al. (167, 310, 311) detected PAR2 mRNA in the sublingual, submaxillary, and parotid salivary glands. Of these three distinct salivary glands, the sublingual exhibited the strongest responses to PAR2 activation resulting in the secretion of mucin. This response appeared to be mediated in part via a tyrosine kinase signal pathway, because genistein, an inhibitor of several tyrosine and other protein kinases, attenuated the effect. Pretreatment with the alkaloid capsaicin, which activates the transient receptor potential of vanilloid type 1 and which, in turn, is known to play a major role in inflammatory thermal nociception, failed to abrogate the secretion of mucin from salivary glands but abolished the PAR2-triggered cytoprotective secretion of mucus in the stomach. These results provide both a sensory neuron-independent and -dependent mechanism for PAR2-regulated exocrine secretion (312). Additional support for a role of PAR2 in salivary gland secretion was provided by data showing that a PAR2 agonist can stimulate amylase secretion from rat parotid gland slices in vitro (311).

    In the GI tract, PAR2 is highly expressed by enterocytes, where it is localized on the apical and basolateral membranes (253, 313). Moreover, myocytes of the muscularis mucosae and muscularis externa as well as neuronal elements are immunoreactive for PAR2. Recent observations indicate that trypsin may regulate enterocytes by cleaving and triggering PAR2 at the apical membrane (253) to induce the generation of PGE2 and PGF1a, suggesting that PAR2 may have a cytoprotective as well as an inflammatory effect on the GI tract. Moreover, mucosal mast cell proteinases may possibly activate PAR2 on enterocytes and colonic myocytes. PAR2-mediated inhibition of intestinal motility may contribute to inflammatory conditions in which mast cells are involved.

    Trypsin may activate PAR2 in the pancreas itself under physiological and pathophysiological conditions because trypsin can be prematurely activated in the inflamed pancreas (314). PAR2 seems to play an important role also in pancreatic nociception because injection of an AP specific for PAR2 can activate and sensitize pancreas-specific afferent neurons in vivo (314). Moreover, Hoogerwerf et al. (314) found that stimulating DRG with a PAR2-AP resulted in enhanced capsaicin- and KCl-stimulated release of calcitonin gene-related peptide, which is a marker for nociceptive signaling.

    PAR2 is expressed by both acinar cells, which release digestive enzymes, and duct cells, which produce fluid and bicarbonate. In isolated pancreatic acini, trypsin and PAR2 agonists stimulate amylase release (154) (N. W. Bunnett, personal communication). In vivo studies revealed the secretion of pancreatic juice after PAR2 activation (167). Interestingly, PAR2 agonists applied to the basolateral but not apical membrane of monolayers of pancreatic duct cells increase short circuit currents due to activation of Ca2+-sensitive Cl– and K+ channels (315). These effects may be of relevance in pancreatitis when trypsin is released across the basolateral membrane.

    H. Signaling by proteinases via PAR2

    There are relatively few studies examining the involvement of PAR2 in signaling cascades compared with the relatively large number of studies concerning PAR1-mediated intracellular signaling (13, 14). So far, there are only indirect data indicating that PAR2 interacts with Gq/G11 (activation of calcium signaling) and possibly with G0/Gi (316). Whether or not PAR2 binds to other G proteins, such as G12 or G13, remains unclear.

    Interaction of PAR2 with Gi and Gq suggests a subsequent activation of PLC, PKC, and MAPK pathways. These signaling pathways can affect various cell activities including cell proliferation, morphological changes, motility and survival, and gene transcription regulation (Table 2 and Fig. 3). Indeed, as demonstrated for neuronal cells and SMC, stimulation by tryptase as well as PAR2-AP leads to subsequent activation of PLC and PKC (317, 318). Kanke et al. (286) demonstrated that PAR2 agonists (trypsin as well as the AP SLIGKV-NH2) stimulate JNK and p38 MAPK activation in a human keratinocyte cell line (NCTC2544). Moreover, Vouret-Craviari et al. (225) described PAR2-induced activation of RhoA in HUVEC revealing a possible mechanism underlying PAR2-mediated effects at cytoskeleton in macrovascular endothelial cells. Together, these data shed some light on the potential signaling mechanisms underlying PAR2-associated cellular effects such as depolarization response in neuronal cells, proliferation and mitogenic response in SMC, proliferation and differentiation of keratinocytes, and cytoskeletal changes in endothelial cells (225, 275, 319, 320).

    Additionally, the effects of trypsin as well as PAR2-AP on the activation of nuclear transcriptional factors have recently been demonstrated. It was shown that PAR2 agonists stimulate NFB-DNA binding activity and activation of upstream kinases IKK and IKK? (286). The effects of PAR2 stimulation on NFB or IKK also have been described in other studies performed on different cell types (13, 263, 287, 297, 321).

    It is also important to pay attention to factor Xa signaling mediated via PAR2. It was mentioned in Section II.H.6 that this factor induces signaling events via PAR1; the potential role of PAR2 in such signaling remained unclear for a long time. Recently, however, the involvement of PAR2 activation in coagulation factor Xa signaling has been reported. In HUVEC, factor Xa interacts with a protein called effector cell proteinase receptor-1 (EPR-1). This interaction is associated with signal transduction, generation of intracellular second messengers, and modulation of cytokine gene expression. Inhibitors of factor Xa blocked these responses (322). Additionally, direct cleavage of PAR2 by factor Xa has been demonstrated, reflecting the complexity of PAR2-induced signaling. These data allowed the authors to suggest that factor Xa induces endothelial cell activation via a novel cascade of receptor activation involving docking to EPR-1 and proteolytic cleavage of PAR2 (322).

    As already mentioned extensively in Section II.H, PAR1 and PAR2 both account for around 90% of endothelial factor Xa-mediated signaling in mice. In contrast, PAR1 virtually accounts for all factor Xa-induced activation in fibroblasts, indicating potential cell-specific synergistic effects of PAR receptors on target cells (25).

    Moreover, Koo et al. (323) demonstrated an involvement of PAR2 in factor Xa signaling by analyzing factor Xa-induced stimulation of coronary artery SMC using cell proliferation and ERK1/2 activation as indices of response. Furthermore, factor Xa-induced ERK1/2 activation was not desensitized by preincubation of the cells with thrombin. However, ERK1/2 activation was markedly attenuated by prior exposure of the cells to PAR2-AP (SLIGKV-NH2). The mitogenic effect of factor Xa was significantly reduced in the presence of an anti-PAR2 monoclonal antibody that attenuates receptor activation, demonstrating the specificity of these effects (323). Together, these observations suggest that various signaling cascades are involved in PAR2-mediated signaling. Clearly, we are just beginning to understand the variety of PAR2-induced cell signaling pathways regulated under physiological conditions and during disease.

    IV. PAR3 and PAR4

    A. Biology and distribution of PAR3 and PAR4

    As already mentioned, murine platelets do not express PAR1. The observation that thrombin was still capable of inducing Ca responses in these cells led to the identification of PAR3 (18). In humans, PAR3 is expressed in bone marrow, heart, brain, placenta, liver, pancreas, thymus, small intestine, stomach, lymph nodes, and trachea, although the cell types remain to be identified. In mouse, PAR3 is expressed by megakaryocytes and platelets, among other cell types. The receptor is necessary for normal thrombin signaling in mouse platelets because blocking of the hirudin-like domain of PAR3 using a specific antibody prevented mouse platelet activation by low but not by high concentrations of thrombin. The same result was observed with PAR3-deficient platelets (19, 324). This also shows that PAR3 is a high-affinity thrombin receptor in mouse that is activated by proteolytic cleavage. Interestingly, murine PAR3 (mPAR3) itself does not lead to thrombin signaling even when overexpressed, indicating that the receptor has lost its ability to function autonomously during evolution (39, 40). Knocking out the PAR3 gene in mice leads to protection of the animals against thrombosis but has a relatively mild effect on hemostasis (325) (Table 2). However, analysis of PAR3 expression in human platelets showed that the receptor is not produced or is hardly produced by these cells (128). This suggests that in humans PAR3 does not play a major role for platelet activation in contrast to the mouse system.

    Antibodies specific for PAR1 inhibited human platelet activation by low but not by high concentrations of thrombin (125, 326). In mice, PAR3 is necessary for normal thrombin signaling in platelets. This indicates the presence of more than one thrombin receptor on the surface of these cells. Indeed, a fourth receptor was cloned, named PAR4 (19, 20, 128). So far, PAR4 has been cloned from human, mouse, and rat tissues (19, 20, 327). In humans, PAR4 is widely expressed in brain (175), testes, placenta, lung, liver, pancreas, thyroid, skeletal muscle, and small intestine (20). In rats, PAR4 is expressed in esophagus, stomach, duodenum, jejunum, distal colon, spleen, and brain (327, 328).

    Both in mice and humans, platelets utilize two thrombin receptors. PAR4 is a low-affinity receptor in platelets both in humans and mice. However, mouse platelets use PAR3 and PAR4 instead of PAR1 and PAR4 to respond to thrombin (19). In humans, platelet effects of thrombin appear to be mediated predominantly by PAR1. Only at high concentrations and when PAR1 activation has been inhibited is platelet activation by thrombin dependent on PAR4. Signals mediated by PAR4 result in calcium influx (19, 122, 329), thromboxane production (117), endostatin secretion in rat platelets (118), and platelet aggregation (122) (Table 2). However, selective activation of human PAR4 (hPAR4) resulted in a weaker response compared with hPAR1-mediated signaling because anionic phospholipids were not exposed on the surface of hPAR4-activated platelets. Interestingly, hPAR4 can also be activated by cathepsin G, a neutrophil granule proteinase (330). Cathepsin G mediates neutrophil-platelet interactions at sites of vascular injury or inflammation. Inhibition of hPAR1 had no effect on platelet responses to cathepsin G, indicating a specific activation of PAR4.

    Patients with Hermansky-Pudlak syndrome, an autosomal recessive disorder, lack platelet dense granules and have no ADP autocrine response. However, these patients show only a mild bleeding phenotype (331, 332). It was hypothesized that the defect of ADP autocrine response was compensated by signaling through PAR4 because the activation of this receptor occurs well after ADP release in normal individuals (333). Therefore, the ADP-autocrine response does not seem to be necessary for platelet aggregation as long as PAR4 is strongly activated. However, it was observed by another laboratory that PAR4-induced, but not PAR1-induced, aggregation was entirely ADP-dependent using a specific AP for PAR4 (334). Moreover, the authors found that subthreshold concentrations of an AP-activating PAR1 potentiated the effects of a PAR4-AP to stimulate maximal aggregation. In addition, both prostacyclin (PGI2) and S-nitroso-glutathione, an NO-releasing agent, reduced AP-stimulated aggregation and fibrinogen-receptor up-regulation.

    PAR4–/– mice had markedly increased bleeding times (40). In addition, the platelets of these mice failed to change shape, mobilize calcium, or aggregate in response to thrombin. Ma et al. (118) observed that a specific AP for PAR4 could induce endostatin release in rat platelets. A selective PAR4 antagonist prevented endostatin release. In human platelets, specific agonists for PAR1 or PAR4 stimulated thromboxane production (117). Thromboxane produced by the combined stimulation of PAR1 and PAR4 was additive, suggesting the presence of two pathways for thrombin-induced thromboxane production in platelets. Blocking PAR1 function with a domain-specific antibody resulted in substantial inhibition of thrombin signaling. However, PAR4 can mediate platelet activation in response to high concentrations of thrombin (128). hPAR4 itself showed no pharmacological effect at either 1 nM or 30 nM thrombin. However, blocking of both hPAR1 and hPAR4 resulted in a profound inhibitory response even at high concentrations of the proteinase (128). This shows that hPAR4 activation is not necessary for robust responses in platelets when hPAR1 function is intact. Thus, inhibition of hPAR1 alone is probably not sufficient when seeking new antithrombotic therapies; it might be necessary to block both receptors simultaneously (106, 128). hPAR4 seems to play a role as a backup signaling device that might mediate thrombin signaling to distinct effectors or with different kinetics compared with PAR1. It might also allow platelets to respond to proteinases other than thrombin. Moreover, both receptors might be able to directly interact with each other.

    The fact that mPAR3 alone does not result in thrombin signaling on mouse platelets showed that mPAR3 functions as a cofactor that promotes cleavage and activation of PAR4 at low concentrations of thrombin (39, 40). Interestingly, knocking out the PAR3 or the PAR4 gene leads to a similar degree of protection against thrombosis in mice (40, 325). In humans, both hPAR1 and hPAR4 may independently mediate thrombin signaling (20, 128, 177). Thus, the mouse system does not have a direct analog in the human platelet. However, mPAR3 promoted cleavage and activation of hPAR4 as effectively as mPAR4. Thus, hPAR4 can be "cofactored" by PAR3 in vitro, but it remains unclear whether or not cofactors such as PAR1 play a direct role in hPAR4 activation in vivo (39).

    This reservoir of multiple thrombin receptors including PAR1, PAR3, and PAR4 may allow for a precise regulation of thrombin-induced inflammatory stimuli under different pathophysiological conditions and the fine-tuned induction of different signal transduction pathways.

    In endothelial cells, thrombin induces a rapid but transient activation of endothelial cells by stimulating the secretion of PGI2 or platelet-activating factor (PAF) and by inducing the production of cell adhesion molecules (P-selectin, E-selectin) (47, 69, 70). Thrombin and a nonselective AP also induce synthesis and release of cytokines from endothelial cells such as IL-1, IL-6, and IL-8 (71, 72). Moreover, they regulate the cytokine-independent expression of ICAM-1 and VCAM-1 on human endothelial cells and cause an increased adhesion of monocytes to endothelial cells. This effect can be diminished by blocking antibodies (anti-CD 18, anti-CD 49d) (73). In the past, the effects observed above had been attributed to the activation of PAR1. However, a recent in vivo study has shown that thrombin-induced leukocyte rolling and adhesion to vascular walls is mediated via PAR4 (74). Although thrombin can trigger the recruitment of leukocytes to sites of inflammation, the PAR1 antagonist RWJ-56110 does not block this effect. In contrast, a PAR4-AP is able to reproduce the effects of thrombin on leukocyte rolling and adherence, indicating that this proinflammatory effect of thrombin is due to the activation of PAR4 and not PAR1 (74). Thus, PAR4 appears to function in early events of inflammatory reaction, in terms of recruiting leukocytes to the site of injury.

    In contrast to a potential participation of PAR1 in stimulating the angiogenic process (335), PAR4 may play an opposing role. Apart from its antiangiogenic role via the activation of platelets, PAR4 can also affect the arterial system by stimulating vascular smooth muscle mitogenesis (220).

    PAR4, like PAR1 and PAR2, also appears to be linked to signal transduction processes that modulate airway smooth muscle tone, because a PAR4-AP caused a rapid transient contractile response in a murine tracheal preparation. This contraction was followed by a transient relaxant phase. The underlying molecular mechanism is still unclear, but the relaxation appears to be dependent on PAR4-stimulated and COX-2-generated PGE2 production, which activates the E-type prostanoid 2 receptor (152, 153).

    In conclusion, preliminary data strongly suggest an important pathophysiological role of PAR4 in vascular homeostasis and platelet function. However, the role of PAR3 as a signaling molecule and the precise function of PAR4 during inflammation and immune response remain to be clarified.

    B. Signaling by proteinases via PAR3 and PAR4

    PAR3 and PAR4 are the most recent members of the proteinase-activated receptor family. They are activated by thrombin. Subsequent to the cloning of PAR3 (18) and PAR4 (19, 20), relatively few publications have appeared examining the involvement of these receptors in signaling cascades. One puzzle that is yet to be resolved, as already mentioned, is the inability of PAR3 to signal in response to its tethered ligand-derived peptide or to thrombin (39, 40). Until this issue is resolved unequivocally, a meaningful discussion about PAR3 signaling is not possible. In contrast, PAR4 appears to be capable of activating both G12/G13 and Gq pathways (201, 336, 337).

    As mentioned above, PAR4 appears to be involved in the same signaling cascades as PAR1 in human platelets. In contrast, mouse platelets express PAR3 and PAR4, but only PAR4 appears to serve as a real signaling receptor, and PAR3 serves solely to facilitate cleavage of PAR4 by thrombin (39).

    Studying downstream signaling events in which PAR1 and PAR4 could be involved in VSMC, Bretschneider et al. (220) revealed that these receptors have distinct downstream signaling kinetics. Later, activation of MAPKs after stimulation of PAR4 was demonstrated. In their recent work, Sabri et al. (338) investigated PAR4-mediated signaling in cardiomyocytes derived from PAR1–/– mice. Using AYPGKF-NH2, a modified PAR4 agonist with an increased binding potency to PAR4, they were able to demonstrate p38 phosphorylation as well as slight activation of PLC and ERK1/2. Additionally, thrombin and PAR4-AP, but not PAR1-AP, were able to activate Src in these cells, clearly indicating that the action of thrombin on Src activation is mediated by PAR4, and not by PAR1 in these cells. Further studies implicated the involvement of Src and EGFR kinase activity in the PAR4-dependent p38 signaling pathway (338).

    Recently, the involvement of PAR4 in factor Xa-mediated signaling has been demonstrated. Camerer et al. (25) revealed that expression of PAR1, PAR2, and PAR4 in Xenopus oocytes confers calcium signaling in response to factor Xa. They further showed that PAR4-AP (AYPGKF-NH2) stimulated increased phosphoinositide hydrolysis in endothelial cells and that responses to AYPGKF-NH2 are absent in PAR4–/– endothelial cells (240). Accordingly, PAR4-mediated phosphoinositide hydrolysis in response to factor Xa has been clearly demonstrated (25).

    V. Conclusions

    There is no doubt that PARs play an important regulatory role during inflammation and immune response. Recent findings consolidate the concept that serine proteinases act as autocrine, paracrine, or endocrine mediators that talk directly to cells (13, 14, 15, 30). Some of these proteinase-stimulated effects are mediated through activation of proteinase-activated receptors, resulting in signal transduction pathways that are involved in inflammation and immune response. In many cases, PARs appear to play a proinflammatory role due to activation of proinflammatory mediators and cytokines (28, 59, 67, 72, 77, 78, 79, 80, 81, 82, 83, 84, 85, 86, 87, 88, 89, 90, 94, 99, 263, 272, 274, 282, 283, 295, 339, 340, 341, 342, 343, 344, 345). In other instances, a protective and antiinflammatory role of PARs has been observed (52, 292, 298, 312, 346, 347, 348). Various serine proteinases, which serve as PAR activators, are relevant mediators during inflammation (Table 2). However, the precise role of PARs and serine proteinases in different tissues, cell types, and states of inflammation remains to be determined.

    Open questions include: 1) For the several proteinases that are released in various tissues from different cell types during inflammation, what is their mechanism of regulation and what is their functional relevance? 2) What is the significance of the existence of multiple receptor subtypes for one proteinase, and how might these common proteinase targets be differentially regulated? 3) Which immune cells express functional PARs in humans in vivo, and what is their biological relevance? 4) Which are the endogenous proteinases that activate PAR2 in human inflammation in vivo? 5) Which factors influence the regulation of PARs during inflammation or host defense? 6) Which effects of PAR-activating proteinases are non-PAR-mediated effects? 7) Which are the cell-specific signaling pathways and molecular mechanisms (transcription factors) after selective PAR activation in different inflammatory states and tissues? 8) What role do PARs play in neuronal transmission in humans? 9) Can PAR agonists or antagonists be used as therapeutic agents during inflammation or immune response in human diseases?

    Thus, an integrative understanding of a regulatory role of serine proteinases as extracellular degradative enzymes as well as hormone-like signaling messengers in part via PARs and their counterregulation by extracellular proteinase inhibitors and intracellular molecules should lead to effective therapeutic approaches for various inflammatory/immune diseases such as thrombosis, sepsis, bacterial infections, gingivitis, asthma, hyperreactivity reaction, lung fibrosis, renal inflammation, rheumatoid arthritis, colitis ulcerosa, Crohn’s disease, pancreatitis, Alzheimer’s disease, amyotrophic lateral sclerosis, HIV encephalitis, atopic dermatitis, contact dermatitis, rosacea, wound repair, and infertility, for example, as well as pathophysiological symptoms such as pain and pruritus, in the future.

    Footnotes

    This work was supported by grants from the Federal Ministry of Education and Research (Interdisciplinary Center for Clinical Research, Münster F?. II-703-04; German Research Association, STE 1014; Collaborative Research Center (SFB) 293; SFB 492 to M.S.), Schering Foundation (to M.S., V.S.), C.E.R.I.E.S., Paris, France, and Boltzmann-Institute, Münster, Germany (to T.A.L., M.S.); Novartis Research Foundation (Vienna, Austria); Rosacea Foundation (to M.S.), and Boehringer-Ingelheim Fonds (to J.B.). The work of M.D.H. and N.V. is supported by funds from the Canadian Institutes for Health Research, the Kidney Foundation of Canada (to M.D.H.), the Heart and Stroke Foundation of Canada (to M.D.H.), by a Johnson & Johnson Focused Giving Grant, and by the Ileitis Colitis Foundation (to N.V.).

    First Published Online October 12, 2004

    Abbreviations: AP, Activating peptide; AP-1, activator protein-1; [Ca2+]i, intracellular calcium; CGRP, calcitonin-gene related peptide; CNS, central nervous system; COX, cyclooxygenase; cPLA2, cytosolic phospholipase A2; DRG, dorsal root ganglion; EGF, epidermal growth factor; EGFR, EGF receptor; EPR-1, effector cell proteinase receptor-1; FAK, focal adhesion kinase; GI, gastrointestinal; GM-CSF, granulocyte-macrophage colony-stimulating factor; GP, glycoprotein; GPCR, G protein-coupled receptor; GRK, G protein-coupled receptor kinase; HMEC, human microvascular endothelial cells; hPAR, human PAR; HUVEC, human umbilical vein endothelial cells; ICAM-1, intercellular adhesion molecule-1; IFN-, interferon ; IKK, inhibitor of NFB kinase; iNOS, inducible NO synthase; JNK, Jun N-terminal kinase; MCP-1, monocyte chemoattractant protein-1; MDCK cells, Madin-Darby canine kidney cells; MMP, matrix metalloproteinase; mPAR, murine PAR; NFB, nuclear factor B; NK cells, natural killer cells; NK receptor, neurokinin receptor; NO, nitric oxide; PAF, platelet-activating factor; PAR, proteinase-activated receptor; PAR-AP, proteinase-activated receptor activating peptide; PDGF, platelet-derived growth factor; PG, prostaglandin; PI3, phosphoinositide 3; PKA, protein kinase A; PKB, protein kinase B; PKC, protein kinase C; PLC, phospholipase C; PTX, pertussis toxin; Pyk2, proline-rich tyrosine kinase 2; SLP-76, SH2-domain containing leukocyte-specific phosphoprotein of 76 kDa; SMC, smooth muscle cells; SP, substance P; Src, protein tyrosin kinase Src; TK, tyrosine kinase; VCAM-1 , vascular cell adhesion molecule-1; VSMC, vascular SMC; ZAP-70, -chain-associated protein kinase of 70 kDa.

    References

    Barrett AJ, Rawlings ND, Woessner JF 1998 Proteolytic enzymes: nomenclature and classification. Handbook of proteolytic enzymes. New York: Academic Press; 13–29

    Rinderknecht H 1986 Activation of pancreatic zymogens: normal activation, premature intrapancreatic activation, protective mechanism against inappropriate activation. Dig Dis Sci 31:314–321

    Neurath H, Walsh K 1976 Role of proteolytic enzymes in biological regulation (a review). Proc Natl Acad Sci USA 73:3825–3832

    Walsh PN, Ahmad S 2002 Proteases in blood clotting. Essays Biochem 38:95–111

    Thomas CA, Yost Jr FJ, Snyderman R, Hatcher VB, Lazarus GS 1977 Cellular serine proteinase induces chemotaxis by complement activation: proteinases of human epidermis; a possible mechanism for polymorphonuclear leukocyte chemotaxis. Nature 269:521–522

    Krowarsch D, Cierpicki T, Jelen F, Otlewski J 2003 Canonical protein inhibitors of serine proteases. Cell Mol Life Sci 60:2427–2444

    Cumashi A, Ansuini H, Celli N, De Blasi A, O’Brien PJ, Brass LF, Molino M 2001 Neutrophil proteases can inactivate human PAR3 and abolish the co-receptor function of PAR3 on murine platelets. Thromb Haemost 85:533–538

    Muehlenweg B, Assfalg-Machleidt I, Parrado SG, Burgle M, Creutzburg S, Schmitt M, Auerswald EA, Machleidt W, Magdolen V 2000 A novel type of bifunctional inhibitor directed against proteolytic activity and receptor/ligand interaction. Cystatin with a urokinase receptor binding site. J Biol Chem 275:33562–33566

    Kinlough-Rathbone RL, Perry DW, Rand ML, Packham MA 1999 Effects of cathepsin G pretreatment of platelets on their subsequent responses to aggregating agents. Thromb Res 95:315–323

    Hayes KL, Tracy PB 1999 The platelet high affinity binding site for thrombin mimics hirudin, modulates thrombin-induced platelet activation, and is distinct from the glycoprotein Ib-IX-V complex. J Biol Chem 274:972–980

    Lu B, Figini M, Emanueli C, Geppetti P, Grady EF, Gerard NP, Ansell J, Payan DG, Gerard C, Bunnett N 1997 The control of microvascular permeability and blood pressure by neutral endopeptidase. Nat Med 3:904–907

    Blasi FCP 2002 uPAR: a versatile signalling orchestrator. Nat Rev Mol Cell Biol 3:932–943

    Macfarlane SR, Seatter MJ, Kanke T, Hunter GD, Plevin R 2001 Proteinase-activated receptors. Pharmacol Rev 53:245–282

    Déry O, Corvera CU, Steinhoff M, Bunnett NW 1998 Proteinase-activated receptors: novel mechanisms of signaling by serine proteases. Am J Physiol 274:C1429–C1452

    Coughlin SR 2000 Thrombin signalling and protease-activated receptors. Nature 407:258–264

    Hollenberg MD, Compton SJ 2002 International union of pharmacology. XXVIII. Proteinase-activated receptors. Pharmacol Rev 54:203–217

    Vu TK, Hung DT, Wheaton VI, Coughlin SR 1991 Molecular cloning of a functional thrombin receptor reveals a novel proteolytic mechanism of receptor activation. Cell 64:1057–1068

    Ishihara H, Connolly AJ, Zeng D, Kahn ML, Zheng YW, Timmons C, Tram T, Coughlin SR 1997 Protease-activated receptor 3 is a second thrombin receptor in humans. Nature 386:502–506

    Kahn ML, Zheng YW, Huang W, Bigornia V, Zeng D, Moff S, Farese Jr RV, Tam C, Coughlin SR 1998 A dual thrombin receptor system for platelet activation. Nature 394:690–694

    Xu WF, Andersen H, Whitmore TE, Presnell SR, Yee DP, Ching A, Gilbert T, Davie EW, Foster DC 1998 Cloning and characterization of human protease-activated receptor 4. Proc Natl Acad Sci USA 95:6642–6646

    Nystedt S, Larsson AK, Aberg H, Sundelin J 1995 The mouse proteinase-activated receptor-2 cDNA and gene. Molecular cloning and functional expression. J Biol Chem 270:5950–5955

    Nystedt S, Emilsson K, Larsson AK, Strombeck B, Sundelin J 1995 Molecular cloning and functional expression of the gene encoding the human proteinase-activated receptor 2. Eur J Biochem 232:84–89

    Molino M, Barnathan ES, Numerof R, Clark J, Dreyer M, Cumashi A, Hoxie JA, Schechter N, Woolkalis M, Brass LF 1997 Interactions of mast cell tryptase with thrombin receptors and PAR-2. J Biol Chem 272:4043–4049

    Steinhoff M, Corvera CU, Thoma MS, Kong W, McAlpine BE, Caughey GH, Ansel JC, Bunnett NW 1999 Proteinase-activated receptor-2 in human skin: tissue distribution and activation of keratinocytes by mast cell tryptase. Exp Dermatol 8:282–294

    Camerer E, Huang W, Coughlin SR 2000 Tissue factor- and factor X-dependent activation of protease-activated receptor 2 by factor VIIa. Proc Natl Acad Sci USA 97:5255–5260

    Smith R, Jenkins A, Lourbakos A, Thompson P, Ramakrishnan V, Tomlinson J, Deshpande U, Johnson DA, Jones R, Mackie EJ, Pike RN 2000 Evidence for the activation of PAR-2 by the sperm protease, acrosin: expression of the receptor on oocytes. FEBS Lett 484:285–290

    Lourbakos A, Chinni C, Thompson P, Potempa J, Travis J, Mackie EJ, Pike RN 1998 Cleavage and activation of proteinase-activated receptor-2 on human neutrophils by gingipain-R from Porphyromonas gingivalis. FEBS Lett 435:45–48

    Lourbakos A, Potempa J, Travis J, D’Andrea MR, Andrade-Gordon P, Santulli R, Mackie EJ, Pike RN 2001 Arginine-specific protease from Porphyromonas gingivalis activates protease-activated receptors on human oral epithelial cells and induces interleukin-6 secretion. Infect Immun 69:5121–5130

    Verrall S, Ishii M, Chen M, Wang L, Tram T, Coughlin SR 1997 The thrombin receptor second cytoplasmic loop confers coupling to Gq-like G proteins in chimeric receptors. Additional evidence for a common transmembrane signaling and G protein coupling mechanism in G protein-coupled receptors. J Biol Chem 272:6898–6902

    Coughlin SR 1999 How the protease thrombin talks to cells. Proc Natl Acad Sci USA 96:11023–11027

    Scarborough RM, Naughton MA, Teng W, Hung DT, Rose J, Vu TK, Wheaton VI, Turck CW, Coughlin SR 1992 Tethered ligand agonist peptides. Structural requirements for thrombin receptor activation reveal mechanism of proteolytic unmasking of agonist function. J Biol Chem 267:13146–13149

    Blackhart BD, Emilsson K, Nguyen D, Teng W, Martelli AJ, Nystedt S, Sundelin J, Scarborough RM 1996 Ligand cross-reactivity within the protease-activated receptor family. J Biol Chem 271:16466–16471

    Coller BS, Springer KT, Scudder LE, Kutok JL, Ceruso M, Prestwich GD 1993 Substituting isoserine for serine in the thrombin receptor activation peptide SFLLRN confers resistance to aminopeptidase M-induced cleavage and inactivation. J Biol Chem 268:20741–20743

    Hollenberg MD 2002 PARs in the stars: proteinase-activated receptors and astrocyte function. Focus on "Thrombin (PAR-1)-induced proliferation in astrocytes via MAPK involves multiple signaling pathways." Am J Physiol Cell Physiol 283:C1347–C1350

    Bernatowicz MS, Klimas CE, Hartl KS, Peluso M, Allegretto NJ, Seiler SM 1996 Development of potent thrombin receptor antagonist peptides. J Med Chem 39:4879–4887

    Andrade-Gordon P, Maryanoff BE, Derian CK, Zhang HC, Addo MF, Darrow AL, Eckardt AJ, Hoekstra WJ, McComsey DF, Oksenberg D, Reynolds EE, Santulli RJ, Scarborough RM, Smith CE, White KB 1999 Design, synthesis, and biological characterization of a peptide-mimetic antagonist for a tethered-ligand receptor. Proc Natl Acad Sci USA 96:12257–12262

    Ahn HS, Foster C, Boykow G, Stamford A, Manna M, Graziano M 2000 Inhibition of cellular action of thrombin by N3-cyclopropyl-7-[[4-(1-methylethyl)phenyl]methyl]-7H-pyrrolo[3, 2-]quinazoline-1,3-diamine (SCH 79797), a nonpeptide thrombin receptor antagonist. Biochem Pharmacol 60:1425–1434

    Wu CC, Hwang TL, Liao CH, Kuo SC, Lee FY, Lee CY, Teng CM 2002 Selective inhibition of protease-activated receptor 4-dependent platelet activation by YD-3. Thromb Haemost 87:1026–1033

    Nakanishi-Matsui M, Zheng YW, Sulciner DJ, Weiss EJ, Ludeman MJ, Coughlin SR 2000 PAR3 is a cofactor for PAR4 activation by thrombin. Nature 404:609–613

    Sambrano GR, Weiss EJ, Zheng YW, Huang W, Coughlin SR 2001 Role of thrombin signalling in platelets in haemostasis and thrombosis. Nature 413:74–78

    Schmidlin F, Bunnett NW 2001 Protease-activated receptors: how proteases signal to cells. Curr Opin Pharmacol 1:575–582

    Coughlin SR 2001 Protease-activated receptors in vascular biology. Thromb Haemost 86:298–307

    Cirino G, Napoli C, Bucci M, Cicala C 2000 Inflammation-coagulation network: are serine protease receptors the knot? Trends Pharmacol Sci 21:170–172

    Vergnolle N, Ferazzini M, D’Andrea MR, Buddenkotte J, Steinhoff M 2003 Proteinase-activated receptors: novel signals for peripheral nerves. Trends Neurosci 26:496–500

    Noorbakhsh F, Vergnolle N, Hollenberg MD, Power C 2003 Proteinase-activated receptors in the nervous system. Nat Rev Neurosci 4:981–990

    Rasmussen UB, Vouret-Craviari V, Jallat S, Schlesinger Y, Pages G, Pavirani A, Lecocq JP, Pouyssegur J, Obberghen-Schilling E 1991 cDNA cloning and expression of a hamster -thrombin receptor coupled to Ca2+ mobilization. FEBS Lett 288:123–128

    Sugama Y, Tiruppathi C, Offakidevi K, Andersen TT, Fenton JW, Malik AB 1992 Thrombin-induced expression of endothelial P-selectin and intercellular adhesion molecule-1: a mechanism for stabilizing neutrophil adhesion. J Cell Biol 119:935–944

    Naldini A, Carney DH 1996 Thrombin modulation of natural killer activity in human peripheral lymphocytes. Cell Immunol 172:35–42

    Cunningham MA, Rondeau E, Chen X, Coughlin SR, Holdsworth SR, Tipping PG 2000 Protease-activated receptor 1 mediates thrombin-dependent, cell-mediated renal inflammation in crescentic glomerulonephritis. J Exp Med 191:455–462

    Ludwicka-Bradley A, Tourkina E, Suzuki S, Tyson E, Bonner M, Fenton JW, Hoffman S, Silver RM 2000 Thrombin upregulates interleukin-8 in lung fibroblasts via cleavage of proteolytically activated receptor-I and protein kinase C- activation. Am J Respir Cell Mol Biol 22:235–243

    Suk K, Cha S 1999 Thrombin-induced interleukin-8 production and its regulation by interferon- and prostaglandin E2 in human monocytic U937 cells. Immunol Lett 67:223–227

    Vergnolle N, Hollenberg MD, Wallace JL 1999 Pro- and anti-inflammatory actions of thrombin: a distinct role for proteinase-activated receptor-1 (PAR1). Br J Pharmacol 126:1262–1268

    Chin AC, Vergnolle N, MacNaughton WK, Wallace JL, Hollenberg MD, Buret AG 2003 Proteinase-activated receptor 1 activation induces epithelial apoptosis and increases intestinal permeability. Proc Natl Acad Sci USA 100:11104–11109

    Howells GL, Macey M, Curtis MA, Stone SR 1993 Peripheral blood lymphocytes express the platelet-type thrombin receptor. Br J Haematol 84:156–160

    Howells GL, Macey M, Murrell AM, Green AR, Curtis MA, Stone SR 1994 Thrombin receptor expression and function in large granular lymphocyte proliferative disorders. Br J Haematol 88:383–388

    DiCuccio MN, Shami P, Hoffman M 1996 A functional tethered ligand thrombin receptor is present on human hematopoietic progenitor cells. Exp Hematol 24:914–918

    Chang MC, Lin CP, Huang TF, Lan WH, Lin YL, Hsieh CC, Jeng JH 1998 Thrombin-induced DNA synthesis of cultured human dental pulp cells is dependent on its proteolytic activity and modulated by prostaglandin E2. J Endod 24:709–713

    Colotta F, Sciacca FL, Sironi M, Luini W, Rabiet MJ, Mantovani A 1994 Expression of monocyte chemotactic protein-1 by monocytes and endothelial cells exposed to thrombin. Am J Pathol 144:975–985

    Grandaliano G, Valente AJ, Abboud HE 1994 A novel biologic activity of thrombin: stimulation of monocyte chemotactic protein production. J Exp Med 179:1737–1741

    Weinstein JR, Gold SJ, Cunningham DD, Gall CM 1995 Cellular localization of thrombin receptor mRNA in rat brain: expression by mesencephalic dopaminergic neurons and codistribution with prothrombin mRNA. J Neurosci 15:2906–2919

    Arena CS, Quirk SM, Zhang YQ, Henrikson KP 1996 Rat uterine stromal cells: thrombin receptor and growth stimulation by thrombin. Endocrinology 137:3744–3749

    Kudahl K, Fisker S, Sonne O 1991 A thrombin receptor in resident rat peritoneal macrophages. Exp Cell Res 193:45–53

    Bar-Shavit R, Wilner GD 1986 Biologic activities of nonenzymatic thrombin: elucidation of a macrophage interactive domain. Semin Thromb Hemost 12:244–249

    Muramatsu I, Laniyonu A, Moore GJ, Hollenberg MD 1992 Vascular actions of thrombin receptor peptide. Can J Physiol Pharmacol 70:996–1003

    Ku DD 1986 Mechanism of thrombin-induced endothelium-dependent coronary vasodilation in dogs: role of its proteolytic enzymatic activity. J Cardiovasc Pharmacol 8:29–36

    Tay-Uyboco J, Poon MC, Ahmad S, Hollenberg MD 1995 Contractile actions of thrombin receptor-derived polypeptides in human umbilical and placental vasculature: evidence for distinct receptor systems. Br J Pharmacol 115:569–578

    Garcia JG, Patterson C, Bahler C, Aschner J, Hart CM, English D 1993 Thrombin receptor activating peptides induce Ca2+ mobilization, barrier dysfunction, prostaglandin synthesis, and platelet-derived growth factor mRNA expression in cultured endothelium. J Cell Physiol 156:541–549

    Herkert O, Kuhl H, Sandow J, Busse R, Schini-Kerth VB 2001 Sex steroids used in hormonal treatment increase vascular procoagulant activity by inducing thrombin receptor (PAR-1) expression: role of the glucocorticoid receptor. Circulation 104:2826–2831

    Weksler BB, Ley CW, Jaffe EA 1978 Stimulation of endothelial cell prostacyclin production by thrombin, trypsin, and the ionophore A 23187. J Clin Invest 62:923–930

    Kameda H, Morita I, Handa M, Kaburaki J, Yoshida T, Mimori T, Murota S, Ikeda Y 1997 Re-expression of functional P-selectin molecules on the endothelial cell surface by repeated stimulation with thrombin. Br J Haematol 97:348–355

    Kaplanski G, Fabrigoule M, Boulay V, Dinarello CA, Bongrand P, Kaplanski S, Farnarier C 1997 Thrombin induces endothelial type II activation in vitro: IL-1 and TNF--independent IL-8 secretion and E-selectin expression. J Immunol 158:5435–5441

    Chi L, Li Y, Stehno-Bittel L, Gao J, Morrison DC, Stechschulte DJ, Dileepan KN 2001 Interleukin-6 production by endothelial cells via stimulation of protease-activated receptors is amplified by endotoxin and tumor necrosis factor-. J Interferon Cytokine Res 21:231–240

    Kaplanski G, Marin V, Fabrigoule M, Boulay V, Benoliel AM, Bongrand P, Kaplanski S, Farnarier C 1998 Thrombin-activated human endothelial cells support monocyte adhesion in vitro following expression of intercellular adhesion molecule-1 (ICAM-1; CD54) and vascular cell adhesion molecule-1 (VCAM-1; CD106). Blood 92:1259–1267

    Vergnolle N, Derian CK, D’Andrea MR, Steinhoff M, Andrade-Gordon P 2002 Characterization of thrombin-induced leukocyte rolling and adherence: a potential proinflammatory role for proteinase-activated receptor-4. J Immunol 169:1467–1473

    Nelken NA, Soifer SJ, O’Keefe J, Vu TK, Charo IF, Coughlin SR 1992 Thrombin receptor expression in normal and atherosclerotic human arteries. J Clin Invest 90:1614–1621

    Takada M, Tanaka H, Yamada T, Ito O, Kogushi M, Yanagimachi M, Kawamura T, Musha T, Yoshida F, Ito M, Kobayashi H, Yoshitake S, Saito I 1998 Antibody to thrombin receptor inhibits neointimal smooth muscle cell accumulation without causing inhibition of platelet aggregation or altering hemostatic parameters after angioplasty in rat. Circ Res 82:980–987

    Asokananthan N, Graham PT, Fink J, Knight DA, Bakker AJ, McWilliam AS, Thompson PJ, Stewart GA 2002 Activation of protease-activated receptor (PAR)-1, PAR-2, and PAR-4 stimulates IL-6, IL-8, and prostaglandin E2 release from human respiratory epithelial cells. J Immunol 168:3577–3585

    Houliston RA, Keogh RJ, Sugden D, Dudhia J, Carter TD, Wheeler-Jones CP 2002 Protease-activated receptors upregulate cyclooxygenase-2 expression in human endothelial cells. Thromb Haemost 88:321–328

    Temkin V, Kantor B, Weg V, Hartman ML, Levi-Schaffer F 2002 Tryptase activates the mitogen-activated protein kinase/activator protein-1 pathway in human peripheral blood eosinophils, causing cytokine production and release. J Immunol 169:2662–2669

    Hirano F, Kobayashi A, Hirano Y, Nomura Y, Fukawa E, Makino I 2002 Thrombin-induced expression of RANTES mRNA through protease activated receptor-1 in human synovial fibroblasts. Ann Rheum Dis 61:834–837

    Riewald M, Petrovan RJ, Donner A, Mueller BM, Ruf W 2002 Activation of endothelial cell protease activated receptor 1 by the protein C pathway. Science 296:1880–1882

    Naldini A, Carney DH, Pucci A, Pasquali A, Carraro F 2000 Thrombin regulates the expression of proangiogenic cytokines via proteolytic activation of protease-activated receptor-1. Gen Pharmacol 35:255–259

    Suo Z, Wu M, Ameenuddin S, Anderson HE, Zoloty JE, Citron BA, Andrade-Gordon P, Festoff BW 2002 Participation of protease-activated receptor-1 in thrombin-induced microglial activation. J Neurochem 80:655–666

    Szaba FM, Smiley ST 2002 Roles for thrombin and fibrin(ogen) in cytokine/chemokine production and macrophage adhesion in vivo. Blood 99:1053–1059

    Meli R, Raso GM, Cicala C, Esposito E, Fiorino F, Cirino G 2001 Thrombin and PAR-1 activating peptide increase iNOS expression in cytokine-stimulated C6 glioma cells. J Neurochem 79:556–563

    McLean K, Schirm S, Johns A, Morser J, Light DR 2001 FXa-induced responses in vascular wall cells are PAR-mediated and inhibited by ZK-807834. Thromb Res 103:281–297

    Bassus S, Herkert O, Kronemann N, Gorlach A, Bremerich D, Kirchmaier CM, Busse R, Schini-Kerth VB 2001 Thrombin causes vascular endothelial growth factor expression in vascular smooth muscle cells: role of reactive oxygen species. Arterioscler Thromb Vasc Biol 21:1550–1555

    Carmeliet P 2001 Biomedicine. Clotting factors build blood vessels. Science 293:1602–1604

    Marin V, Farnarier C, Gres S, Kaplanski S, Su MS, Dinarello CA, Kaplanski G 2001 The p38 mitogen-activated protein kinase pathway plays a critical role in thrombin-induced endothelial chemokine production and leukocyte recruitment. Blood 98:667–673

    Martin CB, Mahon GM, Klinger MB, Kay RJ, Symons M, Der CJ, Whitehead IP 2001 The thrombin receptor, PAR-1, causes transformation by activation of Rho-mediated signaling pathways. Oncogene 20:1953–1963

    Chow JM, Moffatt JD, Cocks TM 2000 Effect of protease-activated receptor (PAR)-1, -2 and -4-activating peptides, thrombin and trypsin in rat isolated airways. Br J Pharmacol 131:1584–1591

    Kawabata A, Saifeddine M, Al Ani B, Leblond L, Hollenberg MD 1999 Evaluation of proteinase-activated receptor-1 (PAR1) agonists and antagonists using a cultured cell receptor desensitization assay: activation of PAR2 by PAR1-targeted ligands. J Pharmacol Exp Ther 288:358–370

    Cheung WM, Andrade-Gordon P, Derian CK, Damiano BP 1998 Receptor-activating peptides distinguish thrombin receptor (PAR-1) and protease activated receptor 2 (PAR-2) mediated hemodynamic responses in vivo. Can J Physiol Pharmacol 76:16–25

    Johnson K, Choi Y, DeGroot E, Samuels I, Creasey A, Aarden L 1998 Potential mechanisms for a proinflammatory vascular cytokine response to coagulation activation. J Immunol 160:5130–5135

    Anrather D, Millan MT, Palmetshofer A, Robson SC, Geczy C, Ritchie AJ, Bach FH, Ewenstein BM 1997 Thrombin activates nuclear factor-B and potentiates endothelial cell activation by TNF. J Immunol 159:5620–5628

    Griffin CT, Srinivasan Y, Zheng YW, Huang W, Coughlin SR 2001 A role for thrombin receptor signaling in endothelial cells during embryonic development. Science 293:1666–1670

    Mirza H, Yatsula V, Bahou WF 1996 The proteinase activated receptor-2 (PAR-2) mediates mitogenic responses in human vascular endothelial cells. J Clin Invest 97:1705–1714

    Schaeffer P, Riera E, Dupuy E, Herbert JM 1997 Nonproteolytic activation of the thrombin receptor promotes human umbilical vein endothelial cell growth but not intracellular Ca2+, prostacyclin, or permeability. Biochem Pharmacol 53:487–491

    Hamilton JR, Moffatt JD, Frauman AG, Cocks TM 2001 Protease-activated receptor (PAR) 1 but not PAR2 or PAR4 mediates endothelium-dependent relaxation to thrombin and trypsin in human pulmonary arteries. J Cardiovasc Pharmacol 38:108–119

    Dabbagh K, Laurent GJ, McAnulty RJ, Chambers RC 1998 Thrombin stimulates smooth muscle cell procollagen synthesis and mRNA levels via a PAR-1 mediated mechanism. Thromb Haemost 79:405–409

    Nguyen KT, Eskin SG, Patterson C, Runge MS, McIntire LV 2001 Shear stress reduces protease activated receptor-1 expression in human endothelial cells. Ann Biomed Eng 29:145–152

    Shinohara T, Suzuki K, Takada K, Okada M, Ohsuzu F 2002 Regulation of proteinase-activated receptor 1 by inflammatory mediators in human vascular endothelial cells. Cytokine 19:66–75

    Bhattacharya A, Cohen ML 2000 Vascular contraction and relaxation to thrombin and trypsin: thrombomodulin preferentially attenuates thrombin-induced contraction. J Pharmacol Exp Ther 295:284–290

    Nguyen KT, Frye SR, Eskin SG, Patterson C, Runge MS, McIntire LV 2001 Cyclic strain increases protease-activated receptor-1 expression in vascular smooth muscle cells. Hypertension 38:1038–1043

    Tiruppathi C, Yan W, Sandoval R, Naqvi T, Pronin AN, Benovic JL, Malik AB 2000 G protein-coupled receptor kinase-5 regulates thrombin-activated signaling in endothelial cells. Proc Natl Acad Sci USA 97:7440–7445

    Cook JJ, Sitko GR, Bednar B, Condra C, Mellott MJ, Feng DM, Nutt RF, Shafer JA, Gould RJ, Connolly TM 1995 An antibody against the exosite of the cloned thrombin receptor inhibits experimental arterial thrombosis in the African green monkey. Circulation 91:2961–2971

    Derian CK, Damiano BP, Addo MF, Darrow AL, D’Andrea MR, Nedelman M, Zhang HC, Maryanoff BE, Andrade-Gordon P 2003 Blockade of the thrombin receptor protease-activated receptor-1 with a small-molecule antagonist prevents thrombus formation and vascular occlusion in nonhuman primates. J Pharmacol Exp Ther 304:855–861

    Hasan AA, Amenta S, Schmaier AH 1996 Bradykinin and its metabolite, Arg-Pro-Pro-Gly-Phe, are selective inhibitors of -thrombin-induced platelet activation. Circulation 94:517–528

    Bahou WF, Coller BS, Potter CL, Norton KJ, Kutok JL, Goligorsky MS 1993 The thrombin receptor extracellular domain contains sites crucial for peptide ligand-induced activation. J Clin Invest 1405–1413

    Zhang HC, Derian CK, Andrade-Gordon P, Hoekstra WJ, McComsey DF, White KB, Poulter BL, Addo MF, Cheung WM, Damiano BP, Oksenberg D, Reynolds EE, Pandey A, Scarborough RM, Maryanoff BE 2001 Discovery and optimization of a novel series of thrombin receptor (par-1) antagonists: potent, selective peptide mimetics based on indole and indazole templates. J Med Chem 44:1021–1024

    Derian CK, Maryanoff BE, Zhang HC, Andrade-Gordon P 2003 Therapeutic potential of protease-activated receptor-1 antagonists. Expert Opin Investig Drugs 12:209–221

    Gilchrist A, Vanhauwe JF, Li A, Thomas TO, Voyno-Yasenetskaya T, Hamm HE 2001 G minigenes expressing C-terminal peptides serve as specific inhibitors of thrombin-mediated endothelial activation. J Biol Chem 276:25672–25679

    Covic L, Misra M, Badar J, Singh C, Kuliopulos A 2002 Pepducin-based intervention of thrombin-receptor signaling and systemic platelet activation. Nat Med 8:1161–1165

    Covic L, Gresser AL, Talavera J, Swift S, Kuliopulos A 2002 Activation and inhibition of G protein-coupled receptors by cell-penetrating membrane-tethered peptides. Proc Natl Acad Sci USA 99:643–648

    Hung DT, Wong YH, Vu TK, Coughlin SR 1992 The cloned platelet thrombin receptor couples to at least two distinct effectors to stimulate phosphoinositide hydrolysis and inhibit adenylyl cyclase. J Biol Chem 267:20831–20834

    Vassallo Jr RR, Kieber-Emmons T, Cichowski K, Brass LF 1992 Structure-function relationships in the activation of platelet thrombin receptors by receptor-derived peptides. J Biol Chem 267:6081–6085

    Henriksen RA, Hanks VK 2002 PAR-4 agonist AYPGKF stimulates thromboxane production by human platelets. Arterioscler Thromb Vasc Biol 22:861–866

    Ma L, Hollenberg MD, Wallace JL 2001 Thrombin-induced platelet endostatin release is blocked by a proteinase activated receptor-4 (PAR4) antagonist. Br J Pharmacol 134:701–704

    Henn V, Slupsky JR, Grafe M, Anagnostopoulos I, Forster R, Muller-Berghaus G, Kroczek RA 1998 CD40 ligand on activated platelets triggers an inflammatory reaction of endothelial cells. Nature 391:591–594

    Smith CC, Wilson AP, Prichard BN, Betteridge DJ 1986 Stimulus-induced release of endogenous catecholamines from human washed platelets. Clin Sci (Lond) 70:495–500

    Graff J, Klinkhardt U, Schini-Kerth VB, Harder S, Franz N, Bassus S, Kirchmaier CM 2002 Close relationship between the platelet activation marker CD62 and the granular release of platelet-derived growth factor. J Pharmacol Exp Ther 300:952–957

    Andersen H, Greenberg DL, Fujikawa K, Xu W, Chung DW, Davie EW 1999 Protease-activated receptor 1 is the primary mediator of thrombin-stimulated platelet procoagulant activity. Proc Natl Acad Sci USA 96:11189–11193

    Mohle R, Green D, Moore MA, Nachman RL, Rafii S 1997 Constitutive production and thrombin-induced release of vascular endothelial growth factor by human megakaryocytes and platelets. Proc Natl Acad Sci USA 94:663–668

    Radomski A, Jurasz P, Sanders EJ, Overall CM, Bigg HF, Edwards DR, Radomski MW 2002 Identification, regulation and role of tissue inhibitor of metalloproteinases-4 (TIMP-4) in human platelets. Br J Pharmacol 137:1330–1338

    Brass LF, Vassallo Jr RR, Belmonte E, Ahuja M, Cichowski K, Hoxie JA 1992 Structure and function of the human platelet thrombin receptor. Studies using monoclonal antibodies directed against a defined domain within the receptor N terminus. J Biol Chem 267:13795–13798

    Arnaud E, Nicaud V, Poirier O, Rendu F, Alhenc-Gelas M, Fiessinger JN, Emmerich J, Aiach M 2000 Protective effect of a thrombin receptor (protease-activated receptor 1) gene polymorphism toward venous thromboembolism. Arterioscler Thromb Vasc Biol 20:585–592

    Dupont A, Fontana P, Bachelot-Loza C, Reny JL, Bieche I, Desvard F, Aiach M, Gaussem P 2003 An intronic polymorphism in the PAR-1 gene is associated with platelet receptor density and the response to SFLLRN. Blood 101:1833–1840

    Kahn ML, Nakanishi-Matsui M, Shapiro MJ, Ishihara H, Coughlin SR 1999 Protease-activated receptors 1 and 4 mediate activation of human platelets by thrombin. J Clin Invest 103:879–887

    De Marco L, Mazzucato M, Masotti A, Fenton JW, Ruggeri ZM 1991 Function of glycoprotein Ib in platelet activation induced by -thrombin. J Biol Chem 266:23776–23783

    Greco NJ, Jones GD, Tandon NN, Kornhauser R, Jackson B, Jamieson GA 1996 Differentiation of the two forms of GPIb functioning as receptors for -thrombin and von Willebrand factor: Ca2+ responses of protease-treated human platelets activated with -thrombin and the tethered ligand peptide. Biochemistry 35:915–921

    Schmidt VA, Scudder L, Devoe CE, Bernards A, Cupit LD, Bahou WF 2003 IQGAP2 functions as a GTP-dependent effector protein in thrombin-induced platelet cytoskeletal reorganization. Blood 101:3021–3028

    Schmidt VA, Vitale E, Bahou WF 1996 Genomic cloning and characterization of the human thrombin receptor gene. Structural similarity to the proteinase activated receptor-2 gene. J Biol Chem 271:9307–9312

    McRedmond JP, Harriott P, Walker B, Fitzgerald DJ 2000 Streptokinase-induced platelet activation involves antistreptokinase antibodies and cleavage of protease-activated receptor-1. Blood 95:1301–1308

    Derian CK, Santulli RJ, Tomko KA, Haertlein BJ, Andrade-Gordon P 1995 Species differences in platelet responses to thrombin and SFLLRN. Receptor-mediated calcium mobilization and aggregation and regulation by protein kinases. Thromb Res 78:505–519

    Connolly AJ, Ishihara H, Kahn ML, Farese Jr RV, Coughlin SR 1996 Role of the thrombin receptor in development and evidence for a second receptor. Nature 381:516–519

    Altieri DC 1995 Proteases and protease receptors in modulation of leukocyte effector functions. J Leukoc Biol 58:120–127

    Naldini A, Sower L, Bocci V, Meyers B, Carney DH 1998 Thrombin receptor expression and responsiveness of human monocytic cells to thrombin is linked to interferon-induced cellular differentiation. J Cell Physiol 177:76–84

    Raza SL, Nehring LC, Shapiro SD, Cornelius LA 2000 Proteinase-activated receptor-1 regulation of macrophage elastase (MMP-12) secretion by serine proteinases. J Biol Chem 275:41243–41250

    Tordai A, Fenton JW, Andersen T, Gelfand EW 1993 Functional thrombin receptors on human T lymphoblastoid cells. J Immunol 150:4876–4886

    Suidan HS, Niclou SP, Monard D 1996 The thrombin receptor in the nervous system. Semin Thromb Hemost 22:125–133

    Naldini A, Carney DH, Pucci A, Carraro F 2002 Human -thrombin stimulates proliferation of interferon- differentiated, growth-arrested U937 cells, overcoming differentiation-related changes in expression of p21CIP1/WAF1 and cyclin D1. J Cell Physiol 191:290–297

    Morris R, Winyard PG, Brass LF, Blake DR, Morris CJ 1996 Thrombin receptor expression in rheumatoid and osteoarthritic synovial tissue. Ann Rheum Dis 55:841–843

    Shin H, Nakajima T, Kitajima I, Shigeta K, Abeyama K, Imamura T, Okano T, Kawahara K, Nakamura T, Maruyama I 1995 Thrombin receptor-mediated synovial proliferation in patients with rheumatoid arthritis. Clin Immunol Immunopathol 76:225–233

    Macey MG, McCarthy DA, Gaffney K, Perry JD, Brown KA, Newland AC 1998 CD11b positive blood lymphocytes correlate with disease activity in patients with RA. Clin Exp Rheumatol 16:108–109

    Walker TR, Moore SM, Lawson MF, Panettieri Jr RA, Chilvers ER 1998 Platelet-derived growth factor-BB and thrombin activate phosphoinositide 3-kinase and protein kinase B: role in mediating airway smooth muscle proliferation. Mol Pharmacol 54:1007–1015

    Chambers RC, Dabbagh K, McAnulty RJ, Gray AJ, Blanc-Brude OP, Laurent GJ 1998 Thrombin stimulates fibroblast procollagen production via proteolytic activation of protease-activated receptor 1. Biochem J 333:121–127

    D’Andrea MR, Rogahn CJ, Andrade-Gordon P 2000 Localization of protease-activated receptors-1 and -2 in human mast cells: indications for an amplified mast cell degranulation cascade. Biotech Histochem 75:85–90

    Chambers RC, Leoni P, Blanc-Brude OP, Wembridge DE, Laurent GJ 2000 Thrombin is a potent inducer of connective tissue growth factor production via proteolytic activation of protease-activated receptor-1. J Biol Chem 275:35584–35591

    Shimizu S, Gabazza EC, Hayashi T, Ido M, Adachi Y, Suzuki K 2000 Thrombin stimulates the expression of PDGF in lung epithelial cells. Am J Physiol Lung Cell Mol Physiol 279:L503–L510

    Hauck RW, Schulz C, Schomig A, Hoffman RK, Panettieri Jr RA 1999 -Thrombin stimulates contraction of human bronchial rings by activation of protease-activated receptors. Am J Physiol 277:L22–L29

    Cicala C, Bucci M, De Dominicis G, Harriot P, Sorrentino L, Cirino G 1999 Bronchoconstrictor effect of thrombin and thrombin receptor activating peptide in guinea-pigs in vivo. Br J Pharmacol 126:478–484

    Lan RS, Stewart GA, Henry PJ 2000 Modulation of airway smooth muscle tone by protease activated receptor-1,-2, -3 and -4 in trachea isolated from influenza A virus-infected mice. Br J Pharmacol 129:63–70

    Lan RS, Knight DA, Stewart GA, Henry PJ 2001 Role of PGE(2) in protease-activated receptor-1, -2 and -4 mediated relaxation in the mouse isolated trachea. Br J Pharmacol 132:93–100

    B?hm SK, Kong W, Bromme D, Smeekens SP, Anderson DC, Connolly A, Kahn M, Nelken NA, Coughlin SR, Payan DG, Bunnett NW 1996 Molecular cloning, expression and potential functions of the human proteinase-activated receptor-2. Biochem J 314:1009–1016

    de Garavilla L, Vergnolle N, Young SH, Ennes H, Steinhoff M, Ossovskaya VS, D’Andrea MR, Mayer EA, Wallace JL, Hollenberg MD, Andrade-Gordon P, Bunnett NW 2001 Agonists of proteinase-activated receptor 1 induce plasma extravasation by a neurogenic mechanism. Br J Pharmacol 133:975–987

    Buresi MC, Schleihauf E, Vergnolle N, Buret A, Wallace JL, Hollenberg MD, Macnaughton WK 2001 Protease-activated receptor-1 stimulates Ca(2+)-dependent Cl(–) secretion in human intestinal epithelial cells. Am J Physiol Gastrointest Liver Physiol 281:G323–G332

    Vergnolle N 2000 Review article: proteinase-activated receptors—novel signals for gastrointestinal pathophysiology. Aliment Pharmacol Ther 14:257–266

    Hollenberg MD, Yang SG, Laniyonu AA, Moore GJ, Saifeddine M 1992 Action of thrombin receptor polypeptide in gastric smooth muscle: identification of a core pentapeptide retaining full thrombin-mimetic intrinsic activity. Mol Pharmacol 42:186–191

    Saifeddine M, Al Ani B, Sandhu S, Wijesuriya SJ, Hollenberg MD 2001 Contractile actions of proteinase-activated receptor-derived polypeptides in guinea-pig gastric and lung parenchymal strips: evidence for distinct receptor systems. Br J Pharmacol 132:556–566

    Nishikawa H, Kawabata A, Kuroda R, Nishida M, Kawai K 2000 Characterization of protease-activated receptors in rat peritoneal mast cells. Jpn J Pharmacol 82:74–77

    Kawabata A, Kuroda R, Nagata N, Kawao N, Masuko T, Nishikawa H, Kawai K 2001 In vivo evidence that protease-activated receptors 1 and 2 modulate gastrointestinal transit in the mouse. Br J Pharmacol 133:1213–1218

    Mule F, Baffi MC, Falzone M, Cerra MC 2002 Signal transduction pathways involved in the mechanical responses to protease-activated receptors in rat colon. J Pharmacol Exp Ther 303:1265–1272

    Mule F, Baffi MC, Cerra MC 2002 Dual effect mediated by protease-activated receptors on the mechanical activity of rat colon. Br J Pharmacol 136:367–374

    Wang J, Zheng H, Ou X, Fink LM, Hauer-Jensen M 2002 Deficiency of microvascular thrombomodulin and up-regulation of protease-activated receptor-1 in irradiated rat intestine: possible link between endothelial dysfunction and chronic radiation fibrosis. Am J Pathol 160:2063–2072

    Rudroff C, Striegler S, Schilli M, Scheele J 2001 Thrombin enhances adhesion in pancreatic cancer in vitro through the activation of the thrombin receptor PAR 1. Eur J Surg Oncol 27:472–476

    Rudroff C, Seibold S, Kaufmann R, Zetina CC, Reise K, Schafer U, Schneider A, Brockmann M, Scheele J, Neugebauer EA 2002 Expression of the thrombin receptor PAR-1 correlates with tumour cell differentiation of pancreatic adenocarcinoma in vitro. Clin Exp Metastasis 19:181–189

    Kawabata A, Kuroda R, Hollenberg MD 1999 Physiology of protease-activated receptors (PARs): involvement of PARs in digestive functions. Nippon Yakurigaku Zasshi 114(Suppl 1):173P–179P

    Faivre S, Regnauld K, Bruyneel E, Nguyen QD, Mareel M, Emami S, Gespach C 2001 Suppression of cellular invasion by activated G-protein subunits Go, Gi1, Gi2, and Gi3 and sequestration of G?. Mol Pharmacol 60:363–372

    Regnauld K, Nguyen QD, Vakaet L, Bruyneel E, Launay JM, Endo T, Mareel M, Gespach C, Emami S 2002 G-protein (olf) subunit promotes cellular invasion, survival, and neuroendocrine differentiation in digestive and urogenital epithelial cells. Oncogene 21:4020–4031

    McHowat J, Creer MH, Rickard A 2001 Stimulation of protease activated receptors on RT4 cells mediates arachidonic acid release via Ca2+ independent phospholipase A2. J Urol 165:2063–2067

    Festoff BW, Smirnova IV, Ma J, Citron BA 1996 Thrombin, its receptor and protease nexin I, its potent serpin, in the nervous system. Semin Thromb Hemost 22:267–271

    Grand RJ, Turnell AS, Grabham PW 1996 Cellular consequences of thrombin-receptor activation. Biochem J 313:353–368

    Dihanich M, Kaser M, Reinhard E, Cunningham D, Monard D 1991 Prothrombin mRNA is expressed by cells of the nervous system. Neuron 6:575–581

    Wang H, Ubl JJ, Stricker R, Reiser G 2002 Thrombin (PAR-1)-induced proliferation in astrocytes via MAPK involves multiple signaling pathways. Am J Physiol Cell Physiol 283:C1351–C1364

    Striggow F, Riek-Burchardt M, Kiesel A, Schmidt W, Henrich-Noack P, Breder J, Krug M, Reymann KG, Reiser G 2001 Four different types of protease-activated receptors are widely expressed in the brain and are up-regulated in hippocampus by severe ischemia. Eur J Neurosci 14:595–608

    Ubl JJ, Vohringer C, Reiser G 1998 Co-existence of two types of [Ca2+]i-inducing protease-activated receptors (PAR-1 and PAR-2) in rat astrocytes and C6 glioma cells. Neuroscience 86:597–609

    Kaufmann R, Bone B, Westermann M, Nowak G, Ramakrishnan V 1998 Investigation of PAR-1-type thrombin receptors in rat glioma C6 cells with a novel monoclonal anti-PAR-1 antibody (Mab COR7–6H9). J Neurocytol 27:661–666

    Striggow F, Riek M, Breder J, Henrich-Noack P, Reymann KG, Reiser G 2000 The protease thrombin is an endogenous mediator of hippocampal neuroprotection against ischemia at low concentrations but causes degeneration at high concentrations. Proc Natl Acad Sci USA 97:2264–2269

    Xi G, Keep RF, Hua Y, Xiang J, Hoff JT 1999 Attenuation of thrombin-induced brain edema by cerebral thrombin preconditioning. Stroke 30:1247–1255

    Vaughan PJ, Pike CJ, Cotman CW, Cunningham DD 1995 Thrombin receptor activation protects neurons and astrocytes from cell death produced by environmental insults. J Neurosci 15:5389–5401

    Gurwitz D, Cunningham DD 1988 Thrombin modulates and reverses neuroblastoma neurite outgrowth. Proc Natl Acad Sci USA 85:3440–3444

    Suidan HS, Stone SR, Hemmings BA, Monard D 1992 Thrombin causes neurite retraction in neuronal cells through activation of cell surface receptors. Neuron 8:363–375

    Smirnova IV, Zhang SX, Citron BA, Arnold PM, Festoff BW 1998 Thrombin is an extracellular signal that activates intracellular death protease pathways inducing apoptosis in model motor neurons. J Neurobiol 36:64–80

    Donovan FM, Pike CJ, Cotman CW, Cunningham DD 1997 Thrombin induces apoptosis in cultured neurons and astrocytes via a pathway requiring tyrosine kinase and RhoA activities. J Neurosci 17:5316–5326

    Wang H, Ubl JJ, Reiser G 2002 Four subtypes of protease-activated receptors, co-expressed in rat astrocytes, evoke different physiological signaling. Glia 37:53–63

    Friedmann I, Hauben E, Yoles E, Kardash L, Schwartz M 2001 T cell-mediated neuroprotection involves antithrombin activity. J Neuroimmunol 121:12–21

    Riek-Burchardt M, Striggow F, Henrich-Noack P, Reiser G, Reymann KG 2002 Increase of prothrombin-mRNA after global cerebral ischemia in rats, with constant expression of protease nexin-1 and protease-activated receptors. Neurosci Lett 329:181–184

    Niclou SP, Suidan HS, Pavlik A, Vejsada R, Monard D 1998 Changes in the expression of protease-activated receptor 1 and protease nexin-1 mRNA during rat nervous system development and after nerve lesion. Eur J Neurosci 10:1590–1607

    Beecher KL, Andersen TT, Fenton JW, Festoff BW 1994 Thrombin receptor peptides induce shape change in neonatal murine astrocytes in culture. J Neurosci Res 37:108–115

    Pindon A, Festoff BW, Hantai D 1998 Thrombin-induced reversal of astrocyte stellation is mediated by activation of protein kinase C ?-1. Eur J Biochem 255:766–774

    Debeir T, Vige X, Benavides J 1997 Pharmacological characterization of protease-activated receptor (PAR-1) in rat astrocytes. Eur J Pharmacol 323:111–117

    Pindon A, Berry M, Hantai D 2000 Thrombomodulin as a new marker of lesion-induced astrogliosis: involvement of thrombin through the G-protein-coupled protease-activated receptor-1. J Neurosci 20:2543–2550

    Festoff BW, D’Andrea MR, Citron BA, Salcedo RM, Smirnova IV, Andrade-Gordon P 2000 Motor neuron cell death in wobbler mutant mice follows overexpression of the G-protein-coupled, protease-activated receptor for thrombin. Mol Med 6:410–429

    Boven LA, Vergnolle N, Henry SD, Silva C, Imai Y, Holden J, Warren K, Hollenberg MD, Power C 2003 Up-regulation of proteinase-activated receptor 1 expression in astrocytes during HIV encephalitis. J Immunol 170:2638–2646

    Gill JS, Pitts K, Rusnak FM, Owen WG, Windebank AJ 1998 Thrombin induced inhibition of neurite outgrowth from dorsal root ganglion neurons. Brain Res 797:321–327

    Corvera CU, Déry O, McConalogue K, Gamp P, Thoma M, Al Ani B, Caughey GH, Hollenberg MD, Bunnett NW 1999 Thrombin and mast cell tryptase regulate guinea-pig myenteric neurons through proteinase-activated receptors-1 and -2. J Physiol 517:741–756

    Asfaha S, Brussee V, Chapman K, Zochodne DW, Vergnolle N 2002 Proteinase-activated receptor-1 agonists attenuate nociception in response to noxious stimuli. Br J Pharmacol 135:1101–1106

    Kawabata A, Kawao N, Kuroda R, Tanaka A, Shimada C 2002 The PAR-1-activating peptide attenuates carrageenan-induced hyperalgesia in rats. Peptides 23:1181–1183

    Brass LF 2003 Thrombin and platelet activation. Chest 124:18S–25S

    Swift S, Sheridan PJ, Covic L, Kuliopulos A 2000 PAR1 thrombin receptor-G protein interactions. Separation of binding and coupling determinants in the g subunit. J Biol Chem 275:2627–2635

    Kim S, Foster C, Lecchi A, Quinton TM, Prosser DM, Jin J, Cattaneo M, Kunapuli SP 2002 Protease-activated receptors 1 and 4 do not stimulate G(i) signaling pathways in the absence of secreted ADP and cause human platelet aggregation independently of G(i) signaling. Blood 99:3629–3636

    Offermanns S, Laugwitz KL, Spicher K, Schultz G 1994 G proteins of the G12 family are activated via thromboxane A2 and thrombin receptors in human platelets. Proc Natl Acad Sci USA 91:504–508

    Ohmori T, Yatomi Y, Asazuma N, Satoh K, Ozaki Y 1999 Suppression of protein kinase C is associated with inhibition of PYK2 tyrosine phosphorylation and enhancement of PYK2 interaction with Src in thrombin-activated platelets. Thromb Res 93:291–298

    Zhang J, Zhang J, Benovic JL, Sugai M, Wetzker R, Gout I, Rittenhouse SE 1995 Sequestration of a G-protein ? subunit or ADP-ribosylation of Rho can inhibit thrombin-induced activation of platelet phosphoinositide 3-kinases. J Biol Chem 270:6589–6594

    Trumel C, Payrastre B, Plantavid M, Hechler B, Viala C, Presek P, Martinson EA, Cazenave JP, Chap H, Gachet C 1999 A key role of adenosine diphosphate in the irreversible platelet aggregation induced by the PAR1-activating peptide through the late activation of phosphoinositide 3-kinase. Blood 94:4156–4165

    Azim AC, Barkalow K, Chou J, Hartwig JH 2000 Activation of the small GTPases, rac and cdc42, after ligation of the platelet PAR-1 receptor. Blood 95:959–964

    Vaidyula VR, Rao AK 2003 Role of Gq and phospholipase C-?2 in human platelets activation by thrombin receptors PAR1 and PAR4: studies in human platelets deficient in Gq and phospholipase C-?2. Br J Haematol 121:491–496

    Covic L, Gresser AL, Kuliopulos A 2000 Biphasic kinetics of activation and signaling for PAR1 and PAR4 thrombin receptors in platelets. Biochemistry 39:5458–5467

    Shapiro MJ, Weiss EJ, Faruqi TR, Coughlin SR 2000 Protease-activated receptors 1 and 4 are shut off with distinct kinetics after activation by thrombin. J Biol Chem 275:25216–25221

    Wang H, Reiser G 2003 Thrombin signaling in the brain: the role of protease-activated receptors. Biol Chem 384:193–202

    Post GR, Collins LR, Kennedy ED, Moskowitz SA, Aragay AM, Goldstein D, Brown JH 1996 Coupling of the thrombin receptor to G12 may account for selective effects of thrombin on gene expression and DNA synthesis in 1321N1 astrocytoma cells. Mol Biol Cell 7:1679–1690

    Aragay AM, Collins LR, Post GR, Watson AJ, Feramisco JR, Brown JH, Simon MI 1995 G12 requirement for thrombin-stimulated gene expression and DNA synthesis in 1321N1 astrocytoma cells. J Biol Chem 270:20073–20077

    Majumdar M, Seasholtz TM, Buckmaster C, Toksoz D, Brown JH 1999 A rho exchange factor mediates thrombin and G (12)-induced cytoskeletal responses. J Biol Chem 274:26815–26821

    LaMorte VJ, Kennedy ED, Collins LR, Goldstein D, Harootunian AT, Brown JH, Feramisco JR 1993 A requirement for Ras protein function in thrombin-stimulated mitogenesis in astrocytoma cells. J Biol Chem 268:19411–19415

    Friedmann I, Yoles E, Schwartz M 2001 Thrombin attenuation is neuroprotective in the injured rat optic nerve. J Neurochem 76:641–649

    Donovan FM, Cunningham DD 1998 Signaling pathways involved in thrombin-induced cell protection. J Biol Chem 273:12746–12752

    Zieger M, Tausch S, Henklein P, Nowak G, Kaufmann R 2001 A novel PAR-1-type thrombin receptor signaling pathway: cyclic AMP-independent activation of PKA in SNB-19 glioblastoma cells. Biochem Biophys Res Commun 282:952–957

    Patterson C, Stouffer GA, Madamanchi N, Runge MS 2001 New tricks for old dogs: nonthrombotic effects of thrombin in vessel wall biology. Circ Res 88:987–997

    Seasholtz TM, Majumdar M, Kaplan DD, Brown JH 1999 Rho and rho kinase mediate thrombin-stimulated vascular smooth muscle cell DNA synthesis and migration. Circ Res 84:1186–1193

    Bretschneider E, Kaufmann R, Braun M, Nowak G, Glusa E, Schror K 2001 Evidence for functionally active protease-activated receptor-4 (PAR-4) in human vascular smooth muscle cells. Br J Pharmacol 132:1441–1446

    Ghosh SK, Gadiparthi L, Zeng ZZ, Bhanoori M, Tellez C, Bar-Eli M, Rao GN 2002 ATF-1 mediates protease-activated receptor-1 but not receptor tyrosine kinase-induced DNA synthesis in vascular smooth muscle cells. J Biol Chem 277:21325–21331

    Sabri A, Muske G, Zhang H, Pak E, Darrow A, Andrade-Gordon P, Steinberg SF 2000 Signaling properties and functions of two distinct cardiomyocyte protease-activated receptors. Circ Res 86:1054–1061

    Krymskaya VP, Penn RB, Orsini MJ, Scott PH, Plevin RJ, Walker TR, Eszterhas AJ, Amrani Y, Chilvers ER, Panettieri Jr RA 1999 Phosphatidylinositol 3-kinase mediates mitogen-induced human airway smooth muscle cell proliferation. Am J Physiol 277:L65–L78

    Bhat GJ, Abraham ST, Singer HA, Baker KM 1997 -Thrombin stimulates sis-inducing factor-A DNA binding activity in rat aortic smooth muscle cells. Hypertension 29:356–360

    Vouret-Craviari V, Grall D, Van Obberghen-Schilling E 2003 Modulation of rho GTPase activity in endothelial cells by selective proteinase-activated receptor (PAR) agonists. J Thromb Haemost 1:1103–1111

    Jin E, Fujiwara M, Pan X, Ghazizadeh M, Arai S, Ohaki Y, Kajiwara K, Takemura T, Kawanami O 2003 Protease-activated receptor (PAR)-1 and PAR-2 participate in the cell growth of alveolar capillary endothelium in primary lung adenocarcinomas. Cancer 97:703–713

    Minami T, Sugiyama A, Wu SQ, Abid R, Kodama T, Aird WC 2004 Thrombin and phenotypic modulation of the endothelium. Arterioscler Thromb Vasc Biol 24:41–53

    Rahman A, True AL, Anwar KN, Ye RD, Voyno-Yasenetskaya TA, Malik AB 2002 G (q) and G? regulate PAR-1 signaling of thrombin-induced NF-B activation and ICAM-1 transcription in endothelial cells. Circ Res 91:398–405

    Vanhauwe JF, Thomas TO, Minshall RD, Tiruppathi C, Li A, Gilchrist A, Yoon EJ, Malik AB, Hamm HE 2002 Thrombin receptors activate G(o) proteins in endothelial cells to regulate [Ca2+]i and cell shape changes. J Biol Chem 277:34143–34149

    Takata M, Urakaze M, Temaru R, Yamazaki K, Nakamura N, Nobata Y, Kishida M, Sato A, Kobayashi M 2001 Pravastatin suppresses the interleukin-8 production induced by thrombin in human aortic endothelial cells cultured with high glucose by inhibiting the p44/42 mitogen activated protein kinase. Br J Pharmacol 134:753–762

    Rahman A, Anwar KN, Uddin S, Xu N, Ye RD, Platanias LC, Malik AB 2001 Protein kinase C- regulates thrombin-induced ICAM-1 gene expression in endothelial cells via activation of p38 mitogen-activated protein kinase. Mol Cell Biol 21:5554–5565

    Rahman A, Anwar KN, True AL, Malik AB 1999 Thrombin-induced p65 homodimer binding to downstream NF-B site of the promoter mediates endothelial ICAM-1 expression and neutrophil adhesion. J Immunol 162:5466–5476

    Minami T, Abid MR, Zhang J, King G, Kodama T, Aird WC 2003 Thrombin stimulation of vascular adhesion molecule-1 in endothelial cells is mediated by protein kinase C (PKC)--NF-B and PKC--GATA signaling pathways. J Biol Chem 278:6976–6984

    Minami T, Rosenberg RD, Aird WC 2001 Transforming growth factor-?1-mediated inhibition of the flk-1/KDR gene is mediated by a 5'-untranslated region palindromic GATA site. J Biol Chem 276:5395–5402

    Mari B, Imbert V, Belhacene N, Far DF, Peyron JF, Pouyssegur J, Obberghen-Schilling E, Rossi B, Auberger P 1994 Thrombin and thrombin receptor agonist peptide induce early events of T cell activation and synergize with TCR cross-linking for CD69 expression and interleukin 2 production. J Biol Chem 269:8517–8523

    Bar-Shavit R, Maoz M, Yongjun Y, Groysman M, Dekel I, Katzav S 2002 Signalling pathways induced by protease-activated receptors and integrins in T cells. Immunology 105:35–46

    Bustelo XR 2000 Regulatory and signaling properties of the Vav family. Mol Cell Biol 20:1461–1477

    Nicholson AC, Nachman RL, Altieri DC, Summers BD, Ruf W, Edgington TS, Hajjar DP 1996 Effector cell protease receptor-1 is a vascular receptor for coagulation factor Xa. J Biol Chem 271:28407–28413

    Riewald M, Kravchenko VV, Petrovan RJ, O’Brien PJ, Brass LF, Ulevitch RJ, Ruf W 2001 Gene induction by coagulation factor Xa is mediated by activation of protease-activated receptor 1. Blood 97:3109–3116

    Camerer E, Kataoka H, Kahn M, Lease K, Coughlin SR 2002 Genetic evidence that protease-activated receptors mediate factor Xa signaling in endothelial cells. J Biol Chem 277:16081–16087

    Riewald M, Ruf W 2001 Mechanistic coupling of protease signaling and initiation of coagulation by tissue factor. Proc Natl Acad Sci USA 98:7742–7747

    Nystedt S, Emilsson K, Wahlestedt C, Sundelin J 1994 Molecular cloning of a potential proteinase activated receptor. Proc Natl Acad Sci USA 91:9208–9212

    Steinhoff M, Vergnolle N, Young SH, Tognetto M, Amadesi S, Ennes HS, Trevisani M, Hollenberg MD, Wallace JL, Caughey GH, Mitchell SE, Williams LM, Geppetti P, Mayer EA, Bunnett NW 2000 Agonists of proteinase-activated receptor 2 induce inflammation by a neurogenic mechanism. Nat Med 6:151–158

    Hou L, Howells GL, Kapas S, Macey MG 1998 The protease-activated receptors and their cellular expression and function in blood-related cells. Br J Haematol 101:1–9

    Frungieri MB, Weidinger S, Meineke V, Kohn FM, Mayerhofer A 2002 Proliferative action of mast-cell tryptase is mediated by PAR2, COX2, prostaglandins, and PPAR : possible relevance to human fibrotic disorders. Proc Natl Acad Sci USA 99:15072–15077

    Al Ani B, Saifeddine M, Hollenberg MD 1995 Detection of functional receptors for the proteinase-activated-receptor-2-activating polypeptide, SLIGRL-NH2, in rat vascular and gastric smooth muscle. Can J Physiol Pharmacol 73:1203–1207

    Santulli RJ, Derian CK, Darrow AL, Tomko KA, Eckardt AJ, Seiberg M, Scarborough RM, Andrade-Gordon P 1995 Evidence for the presence of a protease-activated receptor distinct from the thrombin receptor in human keratinocytes. Proc Natl Acad Sci USA 92:9151–9155

    Hwa JJ, Ghibaudi L, Williams P, Chintala M, Zhang R, Chatterjee M, Sybertz E 1996 Evidence for the presence of a proteinase-activated receptor distinct from the thrombin receptor in vascular endothelial cells. Circ Res 78:581–588

    Mari B, Guerin S, Far DF, Breitmayer JP, Belhacene N, Peyron JF, Rossi B, Auberger P 1996 Thrombin and trypsin-induced Ca(2+) mobilization in human T cell lines through interaction with different protease-activated receptors. FASEB J 10:309–316

    Saifeddine M, Al Ani B, Cheng CH, Wang L, Hollenberg MD 1996 Rat proteinase-activated receptor-2 (PAR-2): cDNA sequence and activity of receptor-derived peptides in gastric and vascular tissue. Br J Pharmacol 118:521–530

    Corvera CU, Déry O, McConalogue K, B?hm SK, Khitin LM, Caughey GH, Payan DG, Bunnett NW 1997 Mast cell tryptase regulates rat colonic myocytes through proteinase-activated receptor 2. J Clin Invest 100:1383–1393

    Howells GL, Macey MG, Chinni C, Hou L, Fox MT, Harriott P, Stone SR 1997 Proteinase-activated receptor-2: expression by human neutrophils. J Cell Sci 110:881–887

    Kong W, McConalogue K, Khitin LM, Hollenberg MD, Payan DG, B?hm SK, Bunnett NW 1997 Luminal trypsin may regulate enterocytes through proteinase-activated receptor 2. Proc Natl Acad Sci USA 94:8884–8889

    Abraham LA, Chinni C, Jenkins AL, Lourbakos A, Ally N, Pike RN, Mackie EJ 2000 Expression of protease-activated receptor-2 by osteoblasts. Bone 26:7–14

    Koshikawa N, Nagashima Y, Miyagi Y, Mizushima H, Yanoma S, Yasumitsu H, Miyazaki K 1997 Expression of trypsin in vascular endothelial cells. FEBS Lett 409:442–448

    Reed DE, Barajas-Lopez C, Cottrell G, Velazquez-Rocha S, Déry O, Grady EF, Bunnett NW, Vanner SJ 2003 Mast cell tryptase and proteinase-activated receptor 2 induce hyperexcitability of guinea-pig submucosal neurons. J Physiol 547:531–542

    Biedermann T, Kneilling M, Mailhammer R, Maier K, Sander CA, Kollias G, Kunkel SL, Hultner L, Rocken M 2000 Mast cells control neutrophil recruitment during T cell-mediated delayed-type hypersensitivity reactions through tumor necrosis factor and macrophage inflammatory protein 2. J Exp Med 192:1441–1452

    Compton SJ, Renaux B, Wijesuriya SJ, Hollenberg MD 2001 Glycosylation and the activation of proteinase-activated receptor 2 (PAR(2)) by human mast cell tryptase. Br J Pharmacol 134:705–718

    Compton SJ, McGuire JJ, Saifeddine M, Hollenberg MD 2002 Restricted ability of human mast cell tryptase to activate proteinase-activated receptor-2 in rat aorta. Can J Physiol Pharmacol 80:987–992

    Brown JK, Tyler CL, Jones CA, Ruoss SJ, Hartmann T, Caughey GH 1995 Tryptase, the dominant secretory granular protein in human mast cells, is a potent mitogen for cultured dog tracheal smooth muscle cells. Am J Respir Cell Mol Biol 13:227–236

    Brown JK, Jones CA, Tyler CL, Ruoss SJ, Hartmann T, Caughey GH 1995 Tryptase-induced mitogenesis in airway smooth muscle cells. Potency, mechanisms, and interactions with other mast cell mediators. Chest 107:95S–96S

    Vergnolle N 1999 Proteinase-activated receptor-2-activating peptides induce leukocyte rolling, adhesion, and extravasation in vivo. J Immunol 163:5064–5069

    Shpacovitch VM, Brzoska T, Buddenkotte J, Stroh C, Sommerhoff CP, Ansel JC, Schulze-Osthoff K, Bunnett NW, Luger TA, Steinhoff M 2002 Agonists of proteinase-activated receptor 2 induce cytokine release and activation of nuclear transcription factor B in human dermal microvascular endothelial cells. J Invest Dermatol 118:380–385

    Komuro T, Miwa S, Minowa T, Okamoto Y, Enoki T, Ninomiya H, Zhang XF, Uemura Y, Kikuchi H, Masaki T 1997 The involvement of a novel mechanism distinct from the thrombin receptor in the vasocontraction induced by trypsin. Br J Pharmacol 120:851–856

    Nystedt S, Ramakrishnan V, Sundelin J 1996 The proteinase-activated receptor 2 is induced by inflammatory mediators in human endothelial cells. Comparison with the thrombin receptor. J Biol Chem 271:14910–14915

    Storck J, Kusters B, Vahland M, Morys-Wortmann C, Zimmermann ER 1996 Trypsin induced von Willebrand factor release from human endothelial cells in mediated by PAR-2 activation. Thromb Res 84:463–473

    Miike S, McWilliam AS, Kita H 2001 Trypsin induces activation and inflammatory mediator release from human eosinophils through protease-activated receptor-2. J Immunol 167:6615–6622

    Lindner JR, Kahn ML, Coughlin SR, Sambrano GR, Schauble E, Bernstein D, Foy D, Hafezi-Moghadam A, Ley K 2000 Delayed onset of inflammation in protease-activated receptor-2-deficient mice. J Immunol 165:6504–6510

    Shpacovitch VM, Varga G, Strey A, Gunzer M, Buddenkotte J, Vergnolle N, Sommerhoff CP, Grabbe S, Gerke V, Homey B, Hollenberg M, Luger TA, Steinhoff M 2004 Agonists of proteinase-activated receptor-2 modulate human neutrophil cytokine secretion, expression of cell adhesion molecules and migration within 3-D collagen lattices. J Leukoc Biol 76:388–398

    Moffatt JD, Jeffrey KL, Cocks TM 2002 Protease-activated receptor-2 activating peptide SLIGRL inhibits bacterial lipopolysaccharide-induced recruitment of polymorphonuclear leukocytes into the airways of mice. Am J Respir Cell Mol Biol 26:680–684

    Schmidlin F, Amadesi S, Dabbagh K, Lewis DE, Knott P, Bunnett NW, Gater PR, Geppetti P, Bertrand C, Stevens ME 2002 Protease-activated receptor 2 mediates eosinophil infiltration and hyperreactivity in allergic inflammation of the airway. J Immunol 169:5315–5321

    Vliagoftis H, Befus AD, Hollenberg MD, Moqbel R 2001 Airway epithelial cells release eosinophil survival-promoting factors (GM-CSF) after stimulation of proteinase-activated receptor 2. J Allergy Clin Immunol 107:679–685

    Stenton GR, Nohara O, Déry RE, Vliagoftis H, Gilchrist M, Johri A, Wallace JL, Hollenberg MD, Moqbel R, Befus AD 2002 Proteinase-activated receptor (PAR)-1 and -2 agonists induce mediator release from mast cells by pathways distinct from PAR-1 and PAR-2. J Pharmacol Exp Ther 302:466–474

    Ferrell WR, Lockhart JC, Kelso EB, Dunning L, Plevin R, Meek SE, Smith AJ, Hunter GD, McLean JS, McGarry F, Ramage R, Jiang L, Kanke T, Kawagoe J 2003 Essential role for proteinase-activated receptor-2 in arthritis. J Clin Invest 111:35–41

    Derian CK, Eckardt AJ, Andrade-Gordon P 1997 Differential regulation of human keratinocyte growth and differentiation by a novel family of protease-activated receptors. Cell Growth Differ 8:743–749

    D’Andrea MR, Derian CK, Leturcq D, Baker SM, Brunmark A, Ling P, Darrow AL, Santulli RJ, Brass LF, Andrade-Gordon P 1998 Characterization of protease-activated receptor-2 immunoreactivity in normal human tissues. J Histochem Cytochem 46:157–164

    Steinhoff M, Neisius U, Ikoma A, Fartasch M, Heyer G, Luger TA, Skov P, Schmelz M 2003 Proteinase-activated receptor-2 mediates itch: a novel pathway for pruritus in human skin. J Neurosci 23:6176–6180

    Harvima IT, Naukkarinen A, Harvima RJ, Horsmanheimo M 1989 Enzyme- and immunohistochemical localization of mast cell tryptase in psoriatic skin. Arch Dermatol Res 281:387–391

    Fraki JE 1977 Human skin proteases: effect of separated proteases on vascular permeability and leukocyte emigration in skin. Acta Derm Venereol 57:393–398

    Schwartz LB 1994 Tryptase: a clinical indicator of mast cell-dependent events. Allergy Proc 15:119–123

    Wood GS, Mueller C, Warnke RA, Weissman IL 1988 In situ localization of HuHF serine protease mRNA and cytotoxic cell-associated antigens in human dermatoses. A novel method for the detection of cytotoxic cells in human tissues. Am J Pathol 133:218–225

    Hou L, Kapas S, Cruchley AT, Macey MG, Harriott P, Chinni C, Stone SR, Howells GL 1998 Immunolocalization of protease-activated receptor-2 in skin: receptor activation stimulates interleukin-8 secretion by keratinocytes in vitro. Immunology 94:356–362

    Wakita H, Furukawa F, Takigawa M 1997 Thrombin and trypsin induce granulocyte-macrophage colony-stimulating factor and interleukin-6 gene expression in cultured normal human keratinocytes. Proc Assoc Am Physicians 109:190–207

    Shpacovitch VM, Brzoska T, Bunnett NW, Sommerhoff CP, Ansel JC, Luger T, Steinhoff M 2001 Agonists of protease-activated receptor-2 induce cytokine release and upregulation of cell adhesion molecules in human dermal microvascular endothelial cells (293). J Invest Dermatol 117:458 (Abstract)

    Seeliger S, Derian C, Vergnolle N, Bunnett N, Nawroth R, Schmelz M, von der Weid P, Buddenkotte J, Sunderk?tter C, Andrade-Gordon P, Harms E, Vestweber D, Luger TA, Steinhoff M 2003 Proinflammatory role of proteinase-activated receptor-2 in humans and mice during cutaneous inflammation in vivo. FASEB J 17:1871–1885

    Kanke T, Macfarlane SR, Seatter MJ, Davenport E, Paul A, McKenzie RC, Plevin R 2001 Proteinase-activated receptor-2-mediated activation of stress-activated protein kinases and inhibitory B kinases in NCTC 2544 keratinocytes. J Biol Chem 276:31657–31666

    Buddenkotte J, Stroh C, Moormann C, Engels I, Shpacovitch VM, Vestweber D, Schulze-Osthoff K, Luger TA, Steinhoff M Agonists of proteinase-activated receptor-2 stimulate activation of NFB and upregulation of ICAM-1 in human keratinocytes. J Invest Dermatol, in press

    Kawagoe J, Takizawa T, Matsumoto J, Tamiya M, Meek SE, Smith AJ, Hunter GD, Plevin R, Saito N, Kanke T, Fujii M, Wada Y 2002 Effect of protease-activated receptor-2 deficiency on allergic dermatitis in the mouse ear. Jpn J Pharmacol 88:77–84

    Schechter NM, Brass LF, Lavker RM, Jensen PJ 1998 Reaction of mast cell proteases tryptase and chymase with protease activated receptors (PARs) on keratinocytes and fibroblasts. J Cell Physiol 176:365–373

    Ricciardolo FL, Steinhoff M, Amadesi S, Guerrini R, Tognetto M, Trevisani M, Creminon C, Bertrand C, Bunnett NW, Fabbri LM, Salvadori S, Geppetti P 2000 Presence and bronchomotor activity of protease-activated receptor-2 in guinea pig airways. Am J Respir Crit Care Med 161:1672–1680

    Schmidlin F, Amadesi S, Vidil R, Trevisani M, Martinet N, Caughey G, Tognetto M, Cavallesco G, Mapp C, Geppetti P, Bunnett NW 2001 Expression and function of proteinase-activated receptor 2 in human bronchial smooth muscle. Am J Respir Crit Care Med 164:1276–1281

    Cocks TM, Fong B, Chow JM, Anderson GP, Frauman AG, Goldie RG, Henry PJ, Carr MJ, Hamilton JR, Moffatt JD 1999 A protective role for protease-activated receptors in the airways. Nature 398:156–160

    Cicala C, Spina D, Keir SD, Severino B, Meli R, Page CP, Cirino G 2001 Protective effect of a PAR2-activating peptide on histamine-induced bronchoconstriction in guinea-pig. Br J Pharmacol 132:1229–1234

    Chambers LS, Black JL, Poronnik P, Johnson PR 2001 Functional effects of protease-activated receptor-2 stimulation on human airway smooth muscle. Am J Physiol Lung Cell Mol Physiol 281:L1369–L1378

    Asokananthan N, Graham PT, Stewart DJ, Bakker AJ, Eidne KA, Thompson PJ, Stewart GA 2002 House dust mite allergens induce proinflammatory cytokines from respiratory epithelial cells: the cysteine protease allergen, Der p 1, activates protease-activated receptor (PAR)-2 and inactivates PAR-1. J Immunol 169:4572–4578

    Ubl JJ, Grishina ZV, Sukhomlin TK, Welte T, Sedehizade F, Reiser G 2002 Human bronchial epithelial cells express PAR-2 with different sensitivity to thermolysin. Am J Physiol Lung Cell Mol Physiol 282:L1339–L1348

    Vliagoftis H, Schwingshackl A, Milne CD, Duszyk M, Hollenberg MD, Wallace JL, Befus AD, Moqbel R 2000 Proteinase-activated receptor-2-mediated matrix metalloproteinase-9 release from airway epithelial cells. J Allergy Clin Immunol 106:537–545

    Fiorucci S, Mencarelli A, Palazzetti B, Distrutti E, Vergnolle N, Hollenberg MD, Wallace JL, Morelli A, Cirino G 2001 Proteinase-activated receptor 2 is an anti-inflammatory signal for colonic lamina propria lymphocytes in a mouse model of colitis. Proc Natl Acad Sci USA 98:13936–13941

    Jenkins AL, Chinni C, De Niese MR, Blackhart B, Mackie EJ 2000 Expression of protease-activated receptor-2 during embryonic development. Dev Dyn 218:465–471

    Smith-Swintosky VL, Cheo-Isaacs CT, D’Andrea MR, Santulli RJ, Darrow AL, Andrade-Gordon P 1997 Protease-activated receptor-2 (PAR-2) is present in the rat hippocampus and is associated with neurodegeneration. J Neurochem 69:1890–1896

    Kaufmann R, Patt S, Zieger M, Kraft R, Nowak G 1999 Presence of the proteinase-activated receptor-2 (PAR-2) in human brain tumor cells–trypsin- and SLIGRL-induced calcium response in primary cultured meningiomas. Cancer Lett 139:109–113

    Domotor E, Bartha K, Machovich R, Adam-Vizi V 2002 Protease-activated receptor-2 (PAR-2) in brain microvascular endothelium and its regulation by plasmin and elastase. J Neurochem 80:746–754

    Carr MJ, Schechter NM, Undem BJ 2000 Trypsin-induced, neurokinin-mediated contraction of guinea pig bronchus. Am J Respir Crit Care Med 162:1662–1667

    Kawabata A, Kawao N, Kuroda R, Tanaka A, Itoh H, Nishikawa H 2001 Peripheral PAR-2 triggers thermal hyperalgesia and nociceptive responses in rats. Neuroreport 12:715–719

    Vergnolle N, Bunnett NW, Sharkey KA, Brussee V, Compton SJ, Grady EF, Cirino G, Gerard N, Basbaum AI, Andrade-Gordon P, Hollenberg MD, Wallace JL 2001 Proteinase-activated receptor-2 and hyperalgesia: a novel pain pathway. Nat Med 7:821–826

    Kawabata A, Kawao N, Itoh H, Shimada C, Takebe K, Kuroda R, Masuko T, Kataoka K, Ogawa S 2002 Role of N-methyl-D-aspartate receptors and the nitric oxide pathway in nociception/hyperalgesia elicited by protease-activated receptor-2 activation in mice and rats. Neurosci Lett 329:349–353

    Kawabata A, Kawao N, Kuroda R, Itoh H, Nishikawa H 2002 Specific expression of spinal Fos after PAR-2 stimulation in mast cell-depleted rats. Neuroreport 13:511–514

    Coelho AM, Vergnolle N, Guiard B, Fioramonti J, Bueno L 2002 Proteinases and proteinase-activated receptor 2: a possible role to promote visceral hyperalgesia in rats. Gastroenterology 122:1035–1047

    Brain SD 2000 New feelings about the role of sensory nerves in inflammation. Nat Med 6:134–135

    Kawabata A, Morimoto N, Nishikawa H, Kuroda R, Oda Y, Kakehi K 2000 Activation of protease-activated receptor-2 (PAR-2) triggers mucin secretion in the rat sublingual gland. Biochem Biophys Res Commun 270:298–302

    Kawabata A, Nishikawa H, Kuroda R, Kawai K, Hollenberg MD 2000 Proteinase-activated receptor-2 (PAR-2): regulation of salivary and pancreatic exocrine secretion in vivo in rats and mice. Br J Pharmacol 129:1808–1814

    Kawabata A, Kinoshita M, Nishikawa H, Kuroda R, Nishida M, Araki H, Arizono N, Oda Y, Kakehi K 2001 The protease-activated receptor-2 agonist induces gastric mucus secretion and mucosal cytoprotection. J Clin Invest 107:1443–1450

    Green BT, Bunnett NW, Kulkarni-Narla A, Steinhoff M, Brown DR 2000 Intestinal type 2 proteinase-activated receptors: expression in opioid-sensitive secretomotor neural circuits that mediate epithelial ion transport. J Pharmacol Exp Ther 295:410–416

    Hoogerwerf WA, Zou L, Shenoy M, Sun D, Micci MA, Lee-Hellmich H, Xiao SY, Winston JH, Pasricha PJ 2001 The proteinase-activated receptor 2 is involved in nociception. J Neurosci 21:9036–9042

    Nguyen TD, Moody MW, Steinhoff M, Okolo C, Koh DS, Bunnett NW 1999 Trypsin activates pancreatic duct epithelial cell ion channels through proteinase-activated receptor-2. J Clin Invest 103:261–269

    Schultheiss M, Neumcke B, Richter HP 1997 Endogenous trypsin receptors in Xenopus oocytes: linkage to internal calcium stores. Cell Mol Life Sci 53:842–849

    Okamoto T, Nishibori M, Sawada K, Iwagaki H, Nakaya N, Jikuhara A, Tanaka N, Saeki K 2001 The effects of stimulating protease-activated receptor-1 and -2 in A172 human glioblastoma. J Neural Transm 108:125–140

    Berger P, Tunon-De-Lara JM, Savineau JP, Marthan R 2001 Selected contribution: tryptase-induced PAR-2-mediated Ca(2+) signaling in human airway smooth muscle cells. J Appl Physiol 91:995–1003

    Gao C, Liu S, Hu HZ, Gao N, Kim GY, Xia Y, Wood JD 2002 Serine proteases excite myenteric neurons through protease-activated receptors in guinea pig small intestine. Gastroenterology 123:1554–1564

    Berger P, Perng DW, Thabrew H, Compton SJ, Cairns JA, McEuen AR, Marthan R, Tunon De Lara JM, Walls AF 2001 Tryptase and agonists of PAR-2 induce the proliferation of human airway smooth muscle cells. J Appl Physiol 91:1372–1379

    Bretschneider E, Kaufmann R, Braun M, Wittpoth M, Glusa E, Nowak G, Schror K 1999 Evidence for proteinase-activated receptor-2 (PAR-2)-mediated mitogenesis in coronary artery smooth muscle cells. Br J Pharmacol 126:1735–1740

    Bono F, Schaeffer P, Herault JP, Michaux C, Nestor AL, Guillemot JC, Herbert JM 2000 Factor Xa activates endothelial cells by a receptor cascade between EPR-1 and PAR-2. Arterioscler Thromb Vasc Biol 20:E107–E112

    Koo BH, Chung KH, Hwang KC, Kim DS 2002 Factor Xa induces mitogenesis of coronary artery smooth muscle cell via activation of PAR-2. FEBS Lett 523:85–89

    Ishihara H, Zeng D, Connolly AJ, Tam C, Coughlin SR 1998 Antibodies to protease-activated receptor 3 inhibit activation of mouse platelets by thrombin. Blood 91:4152–4157

    Weiss EJ, Hamilton JR, Lease KE, Coughlin SR 2002 Protection against thrombosis in mice lacking PAR3. Blood 100:3240–3244

    Hung DT, Vu TK, Wheaton VI, Ishii K, Coughlin SR 1992 Cloned platelet thrombin receptor is necessary for thrombin-induced platelet activation. J Clin Invest 89:1350–1353

    Hoogerwerf WA, Hellmich HL, Micci M, Winston JH, Zou L, Pasricha PJ 2002 Molecular cloning of the rat proteinase-activated receptor 4 (PAR4). BMC Mol Biol 3:2

    Kaufmann R, Patt S, Zieger M, Kraft R, Tausch S, Henklein P, Nowak G 2000 The two-receptor system PAR-1/PAR-4 mediates -thrombin-induced [Ca(2+)](i) mobilization in human astrocytoma cells. J Cancer Res Clin Oncol 126:91–94

    Lourbakos A, Yuan YP, Jenkins AL, Travis J, Andrade-Gordon P, Santulli R, Potempa J, Pike RN 2001 Activation of protease-activated receptors by gingipains from Porphyromonas gingivalis leads to platelet aggregation: a new trait in microbial pathogenicity. Blood 97:3790–3797

    Sambrano GR, Huang W, Faruqi T, Mahrus S, Craik C, Coughlin SR 2000 Cathepsin G activates protease-activated receptor-4 in human platelets. J Biol Chem 275:6819–6823

    King RA, Hearing VJ, Creel D, Oetting WS 2000 Albinism. In: Scriver CR, Beaudet AL, Sly WS, Valle D, eds. The metabolic and molecular basis of inherited disease. New York: McGraw-Hill; 5587–5627

    Anikster Y, Huizing M, White J, Shevchenko YO, Fitzpatrick DL, Touchman JW, Compton JG, Bale SJ, Swank RT, Gahl WA, Toro JR 2001 Mutation of a new gene causes a unique form of Hermansky-Pudlak syndrome in a genetic isolate of central Puerto Rico. Nat Genet 28:376–380

    Covic L, Singh C, Smith H, Kuliopulos A 2002 Role of the PAR4 thrombin receptor in stabilizing platelet-platelet aggregates as revealed by a patient with Hermansky-Pudlak syndrome. Thromb Haemost 87:722–727

    Chung AW, Jurasz P, Hollenberg MD, Radomski MW 2002 Mechanisms of action of proteinase-activated receptor agonists on human platelets. Br J Pharmacol 135:1123–1132

    Yin YJ, Salah Z, Maoz M, Ram SC, Ochayon S, Neufeld G, Katzav S, Bar-Shavit R 2003 Oncogenic transformation induces tumor angiogenesis: a role for PAR1 activation. FASEB J 17:163–174

    Kanthou C, Kanse SM, Kakkar VV, Benzakour O 1996 Involvement of pertussis toxin-sensitive and -insensitive G proteins in -thrombin signalling on cultured human vascular smooth muscle cells. Cell Signal 8:59–66

    Faruqi TR, Weiss EJ, Shapiro MJ, Huang W, Coughlin SR 2000 Structure-function analysis of protease-activated receptor 4 tethered ligand peptides. Determinants of specificity and utility in assays of receptor function. J Biol Chem 275:19728–19734

    Sabri A, Guo J, Elouardighi H, Darrow AL, Andrade-Gordon P, Steinberg SF 2003 Mechanisms of protease-activated receptor-4 actions in cardiomyocytes: role of Src tyrosine kinase. J Biol Chem 278:11714–11720

    Lang R, Song PI, Legat FJ, Lavker RM, Harten B, Kalden H, Grady EF, Bunnett NW, Armstrong CA, Ansel JC 2003 Human corneal epithelial cells express functional PAR-1 and PAR-2. Invest Ophthalmol Vis Sci 44:99–105

    Cenac N, Coelho AM, Nguyen C, Compton S, Andrade-Gordon P, Macnaughton WK, Wallace JL, Hollenberg MD, Bunnett NW, Garcia-Villar R, Bueno L, Vergnolle N 2002 Induction of intestinal inflammation in mouse by activation of proteinase-activated receptor-2. Am J Pathol 161:1903–1915

    Uehara A, Sugawara S, Muramoto K, Takada H 2002 Activation of human oral epithelial cells by neutrophil proteinase 3 through protease-activated receptor-2. J Immunol 169:4594–4603

    Mascia F, Mariani V, Giannetti A, Girolomoni G, Pastore S 2002 House dust mite allergen exerts no direct proinflammatory effects on human keratinocytes. J Allergy Clin Immunol 109:532–538

    Sun G, Stacey MA, Schmidt M, Mori L, Mattoli S 2001 Interaction of mite allergens Der p3 and Der p9 with protease-activated receptor-2 expressed by lung epithelial cells. J Immunol 167:1014–1021

    Gordon JR, Zhang X, Stevenson K, Cosford K 2000 Thrombin induces IL-6 but not TNF secretion by mouse mast cells: threshold-level thrombin receptor and very low level FcRI signaling synergistically enhance IL-6 secretion. Cell Immunol 205:128–135

    Napoli C, De Nigris F, Cicala C, Wallace JL, Caliendo G, Condorelli M, Santagada V, Cirino G 2002 Protease-activated receptor-2 activation improves efficiency of experimental ischemic preconditioning. Am J Physiol Heart Circ Physiol 282:H2004–H2010

    Lan RS, Stewart GA, Henry PJ 2002 Role of protease-activated receptors in airway function: a target for therapeutic intervention? Pharmacol Ther 95:239–257

    Trottier G, Hollenberg M, Wang X, Gui Y, Loutzenhiser K, Loutzenhiser R 2002 PAR-2 elicits afferent arteriolar vasodilation by NO-dependent and NO-independent actions. Am J Physiol Renal Physiol 282:F891–F897

    Cocks TM, Moffatt JD 2000 Protease-activated receptors: sentries for inflammation? Trends Pharmacol Sci 21:103–108

    Chen YH, Pouyssegur J, Courtneidge SA, Obberghen-Schilling E 1994 Activation of Src family kinase activity by the G protein-coupled thrombin receptor in growth-responsive fibroblasts. J Biol Chem 269:27372–27377

    Chen Y, Grall D, Salcini AE, Pelicci PG, Pouyssegur J, Obberghen-Schilling E 1996 Shc adaptor proteins are key transducers of mitogenic signaling mediated by the G protein-coupled thrombin receptor. EMBO J 15:1037–1044

    Trejo J, Connolly AJ, Coughlin SR 1996 The cloned thrombin receptor is necessary and sufficient for activation of mitogen-activated protein kinase and mitogenesis in mouse lung fibroblasts. Loss of responses in fibroblasts from receptor knockout mice. J Biol Chem 271:21536–21541

    Ramars AS, Mukhopadhyay S, Dash D 2002 Regulation of postaggregation events induced by protease-activated receptor 1 ligation in human platelets: evidence of differential signaling pathways. Arch Biochem Biophys 398:253–260

    Huang RS, Sorisky A, Church WR, Simons ER, Rittenhouse SE 1991 "Thrombin" receptor-directed ligand accounts for activation by thrombin of platelet phospholipase C and accumulation of 3-phosphorylated phosphoinositides. J Biol Chem 266:18435–18438

    Henriksen RA, Samokhin GP, Tracy PB 1997 Thrombin-induced thromboxane synthesis by human platelets. Properties of anion binding exosite I-independent receptor. Arterioscler Thromb Vasc Biol 17:3519–3526

    Ohmori T, Yatomi Y, Asazuma N, Satoh K, Ozaki Y 2000 Involvement of proline-rich tyrosine kinase 2 in platelet activation: tyrosine phosphorylation mostly dependent on IIb?3 integrin and protein kinase C, translocation to the cytoskeleton and association with Shc through Grb2. Biochem J 347:561–569

    Dorsam RT, Kim S, Jin J, Kunapuli SP 2002 Coordinated signaling through both G12/13 and G(i) pathways is sufficient to activate GPIIb/IIIa in human platelets. J Biol Chem 277:47588–47595

    Han Y, Nurden A, Combrie R, Pasquet JM 2003 Redistribution of glycoprotein Ib within platelets in response to protease-activated receptors 1 and 4: roles of cytoskeleton and calcium. J Thromb Haemost 1:2206–2215

    Klarenbach S, Chipiuk A, Nelson RC, Hollenberg M, Murray AG 2003 Differential actions of PAR2 and PAR1 in stimulating human endothelial cell exocytosis and permeability: the role of Rho-GTPases. Circ Res 92:272–278

    Riewald M, Petrovan RJ, Donner A, Ruf W 2003 Activated protein C signals through the thrombin receptor PAR1 in endothelial cells. J Endotoxin Res 9:317–321

    Domotor E, Benzakour O, Griffin JH, Yule D, Fukudome K, Zlokovic BV 2003 Activated protein C alters cytosolic calcium flux in human brain endothelium via binding to endothelial protein C receptor and activation of protease activated receptor-1. Blood 101:4797–4801

    Cheng T, Liu D, Griffin JH, Fernandez JA, Castellino F, Rosen ED, Fukudome K, Zlokovic BV 2003 Activated protein C blocks p53-mediated apoptosis in ischemic human brain endothelium and is neuroprotective. Nat Med 9:338–342

    Pai KS, Mahajan VB, Lau A, Cunningham DD 2001 Thrombin receptor signaling to cytoskeleton requires Hsp90. J Biol Chem 276:32642–32647

    Ubl JJ, Sergeeva M, Reiser G 2000 Desensitisation of protease-activated receptor-1 (PAR-1) in rat astrocytes: evidence for a novel mechanism for terminating Ca2+ signalling evoked by the tethered ligand. J Physiol 525:319–330

    Mahajan VB, Pai KS, Lau A, Cunningham DD 2000 Creatine kinase, an ATP-generating enzyme, is required for thrombin receptor signaling to the cytoskeleton. Proc Natl Acad Sci USA 97:12062–12067

    Kamath L, Meydani A, Foss F, Kuliopulos A 2001 Signaling from protease-activated receptor-1 inhibits migration and invasion of breast cancer cells. Cancer Res 61:5933–5940

    Sabri A, Short J, Guo J, Steinberg SF 2002 Protease-activated receptor-1-mediated DNA synthesis in cardiac fibroblast is via epidermal growth factor receptor transactivation: distinct PAR-1 signaling pathways in cardiac fibroblasts and cardiomyocytes. Circ Res 91:532–539

    Steinberg SF, Robinson RB, Lieberman HB, Stern DM, Rosen MR 1991 Thrombin modulates phosphoinositide metabolism, cytosolic calcium, and impulse initiation in the heart. Circ Res 68:1216–1229

    Jiang T, Danilo Jr P, Steinberg SF 1998 The thrombin receptor elevates intracellular calcium in adult rat ventricular myocytes. J Mol Cell Cardiol 30:2193–2199

    Hague S, Oehler MK, MacKenzie IZ, Bicknell R, Rees MC 2002 Protease activated receptor-1 is down regulated by levonorgestrel in endometrial stromal cells. Angiogenesis 5:93–98

    Tognetto M, Trevisani M, Maggiore B, Navarra G, Turini A, Guerrini R, Bunnett NW, Geppetti P, Harrison S 2000 Evidence that PAR-1 and PAR-2 mediate prostanoid-dependent contraction in isolated guinea-pig gallbladder. Br J Pharmacol 131:689–694

    Kawabata A, Kuroda R, Kuroki N, Nishikawa H, Kawai K 2000 Dual modulation by thrombin of the motility of rat oesophageal muscularis mucosae via two distinct protease-activated receptors (PARs): a novel role for PAR-4 as opposed to PAR-1. Br J Pharmacol 131:578–584

    Roche N, Stirling RG, Lim S, Oliver BG, Oates T, Jazrawi E, Caramori G, Chung KF 2003 Effect of acute and chronic inflammatory stimuli on expression of protease-activated receptors 1 and 2 in alveolar macrophages. J Allergy Clin Immunol 111:367–373

    Vliagoftis H 2002 Thrombin induces mast cell adhesion to fibronectin: evidence for involvement of protease-activated receptor-1. J Immunol 169:4551–4558

    Umarova BA, Dugina TN, Shestakova EV, Gluza E, Strukova SM 2000 Activation of rat mast cells upon stimulation of protease-activated receptor (PAR-1). Bull Exp Biol Med 129:314–317

    Colognato R, Slupsky JR, Jendrach M, Burysek L, Syrovets T, Simmet T 2003 Differential expression and regulation of proteinase-activated receptors in human peripheral monocytes and monocyte-derived antigen-presenting cells. Blood 102:2645–2652

    Cocks TM, Sozzi V, Moffatt JD, Selemidis S 1999 Protease-activated receptors mediate apamin-sensitive relaxation of mouse and guinea pig gastrointestinal smooth muscle. Gastroenterology 116:586–592

    Nguyen QD, Faivre S, Bruyneel E, Rivat C, Seto M, Endo T, Mareel M, Emami S, Gespach C 2002 RhoA- and RhoD-dependent regulatory switch of G subunit signaling by PAR-1 receptors in cellular invasion. FASEB J 16:565–576

    Buresi MC, Buret AG, Hollenberg MD, Macnaughton WK 2002 Activation of proteinase-activated receptor 1 stimulates epithelial chloride secretion through a unique MAP kinase- and cyclo-oxygenase-dependent pathway. FASEB J 16:1515–1525

    Kawabata A, Kuroda R, Kuroki N, Nishikawa H, Kawai K, Araki H 2000 Characterization of the protease-activated receptor-1-mediated contraction and relaxation in the rat duodenal smooth muscle. Life Sci 67:2521–2530

    Greenberg DL, Mize GJ, Takayama TK 2003 Protease-activated receptor mediated RhoA signaling and cytoskeletal reorganization in LNCaP cells. Biochemistry 42:702–709

    de Niese MR, Chinni C, Pike RN, Bottomley SP, Mackie EJ 2002 Dissection of protease-activated receptor-1-dependent and -independent responses to thrombin in skeletal myoblasts. Exp Cell Res 274:149–156

    Garrido R, Segura B, Zhang W, Mulholland M 2002 Presence of functionally active protease-activated receptors 1 and 2 in myenteric glia. J Neurochem 83:556–564

    Conant K, Haughey N, Nath A, St Hillaire C, Gary DS, Pardo CA, Wahl LM, Bilak M, Milward E, Mattson MP 2002 Matrix metalloproteinase-1 activates a pertussis toxin-sensitive signaling pathway that stimulates the release of matrix metalloproteinase-9. J Neurochem 82:885–893

    Fang M, Kovacs KJ, Fisher LL, Larson AA 2003 Thrombin inhibits NMDA-mediated nociceptive activity in the mouse: possible mediation by endothelin. J Physiol 549:903–917

    Ellis CA, Malik AB, Gilchrist A, Hamm H, Sandoval R, Voyno-Yasenetskaya T, Tiruppathi C 1999 Thrombin induces proteinase-activated receptor-1 gene expression in endothelial cells via activation of Gi-linked Ras/mitogen-activated protein kinase pathway. J Biol Chem 274:13718–13727

    Algermissen B, Sitzmann J, Nurnberg W, Laubscher JC, Henz BM, Bauer F 2000 Distribution and potential biologic function of the thrombin receptor PAR-1 on human keratinocytes. Arch Dermatol Res 292:488–495

    Bromberg ME, Bailly MA, Konigsberg WH 2001 Role of protease-activated receptor 1 in tumor metastasis promoted by tissue factor. Thromb Haemost 86:1210–1214

    Pendurthi UR, Ngyuen M, Andrade-Gordon P, Petersen LC, Rao LV 2002 Plasmin induces Cyr61 gene expression in fibroblasts via protease-activated receptor-1 and p44/42 mitogen-activated protein kinase-dependent signaling pathway. Arterioscler Thromb Vasc Biol 22:1421–1426

    Papadaki M, Ruef J, Nguyen KT, Li F, Patterson C, Eskin SG, McIntire LV, Runge MS 1998 Differential regulation of protease activated receptor-1 and tissue plasminogen activator expression by shear stress in vascular smooth muscle cells. Circ Res 83:1027–1034

    Chen X, Berrou J, Vigneau C, Rondeau E 2001 Role of the third intracellular loop and of the cytoplasmic tail in the mitogenic signaling of the protease-activated receptor 1. Int J Mol Med 8:309–314

    Chen X, Berrou J, Vigneau C, Delarue F, Rondeau E 2001 Internalisation of the protease-activated receptor-1: role of the third intracellular loop and of the cytoplasmic tail. Int J Mol Med 7:653–658

    Shapiro MJ, Trejo J, Zeng D, Coughlin SR 1996 Role of the thrombin receptor’s cytoplasmic tail in intracellular trafficking. Distinct determinants for agonist-triggered versus tonic internalization and intracellular localization. J Biol Chem 271:32874–32880

    Fukuhara S, Marinissen MJ, Chiariello M, Gutkind JS 2000 Signaling from G protein-coupled receptors to ERK5/Big MAPK 1 involves G q and G 12/13 families of heterotrimeric G proteins. Evidence for the existence of a novel Ras and Rho-independent pathway. J Biol Chem 275:21730–21736

    Madamanchi NR, Hu ZY, Li F, Horaist C, Moon SK, Patterson C, Runge MS 2002 A noncoding RNA regulates human protease-activated receptor-1 gene during embryogenesis. Biochim Biophys Acta 1576:237–245

    Darmoul D, Gratio V, Devaud H, Lehy T, Laburthe M 2003 Aberrant expression and activation of the thrombin receptor protease-activated receptor-1 induces cell proliferation and motility in human colon cancer cells. Am J Pathol 162:1503–1513

    Kahan C, Seuwen K, Meloche S, Pouyssegur J 1992 Coordinate, biphasic activation of p44 mitogen-activated protein kinase and S6 kinase by growth factors in hamster fibroblasts. Evidence for thrombin-induced signals different from phosphoinositide turnover and adenylylcyclase inhibition. J Biol Chem 267:13369–13375

    van Corven EJ, Hordijk PL, Medema RH, Bos JL, Moolenaar WH 1993 Pertussis toxin-sensitive activation of p21ras by G protein-coupled receptor agonists in fibroblasts. Proc Natl Acad Sci USA 90:1257–1261

    Tran T, Stewart AG 2003 Protease-activated receptor (PAR)-independent growth and pro-inflammatory actions of thrombin on human cultured airway smooth muscle. Br J Pharmacol 138:865–875

    Storck J, Zimmermann ER 1996 Regulation of the thrombin receptor response in human endothelial cells. Thromb Res 81:121–131

    Tiruppathi C, Lum H, Andersen TT, Fenton JW, Malik AB 1992 Thrombin receptor 14-amino acid peptide binds to endothelial cells and stimulates calcium transients. Am J Physiol 263:L595–L601

    Belham CM, Scott PH, Twomey DP, Gould GW, Wadsworth RM, Plevin R 1997 Evidence that thrombin-stimulated DNA synthesis in pulmonary arterial fibroblasts involves phosphatidylinositol 3-kinase-dependent p70 ribosomal S6 kinase activation. Cell Signal 9:109–116

    Kranzhofer R, Clinton SK, Ishii K, Coughlin SR, Fenton JW, Libby P 1996 Thrombin potently stimulates cytokine production in human vascular smooth muscle cells but not in mononuclear phagocytes. Circ Res 79:286–294

    Keogh RJ, Houliston RA, Wheeler-Jones CP 2002 Thrombin-stimulated Pyk2 phosphorylation in human endothelium is dependent on intracellular calcium and independent of protein kinase C and Src kinases. Biochem Biophys Res Commun 294:1001–1008

    Jenkins AL, Bootman MD, Taylor CW, Mackie EJ, Stone SR 1993 Characterization of the receptor responsible for thrombin-induced intracellular calcium responses in osteoblast-like cells. J Biol Chem 268:21432–21437

    Abraham LA, Mackie EJ 1999 Modulation of osteoblast-like cell behavior by activation of protease-activated receptor-1. J Bone Miner Res 14:1320–1329

    Chay CH, Cooper CR, Gendernalik JD, Dhanasekaran SM, Chinnaiyan AM, Rubin MA, Schmaier AH, Pienta KJ 2002 A functional thrombin receptor (PAR1) is expressed on bone-derived prostate cancer cell lines. Urology 60:760–765

    Kaufmann R, Patt S, Kraft R, Zieger M, Henklein P, Neupert G, Nowak G 1999 PAR 1-type thrombin receptors are involved in thrombin-induced calcium signaling in human meningioma cells. J Neurooncol 42:131–136

    Debeir T, Gueugnon J, Vige X, Benavides J 1996 Transduction mechanisms involved in thrombin receptor-induced nerve growth factor secretion and cell division in primary cultures of astrocytes. J Neurochem 66:2320–2328

    McNamara CA, Sarembock IJ, Gimple LW, Fenton JW, Coughlin SR, Owens GK 1993 Thrombin stimulates proliferation of cultured rat aortic smooth muscle cells by a proteolytically activated receptor. J Clin Invest 91:94–98

    Chaikof EL, Caban R, Yan CN, Rao GN, Runge MS 1995 Growth-related responses in arterial smooth muscle cells are arrested by thrombin receptor antisense sequences. J Biol Chem 270:7431–7436

    Cicala C, Morello S, Santagada V, Caliendo G, Sorrentino L, Cirino G 2001 Pharmacological dissection of vascular effects caused by activation of protease-activated receptors 1 and 2 in anesthetized rats. FASEB J 15:1433–1435

    Shin H, Kitajima I, Nakajima T, Shao Q, Tokioka T, Takasaki I, Hanyu N, Kubo T, Maruyama I 1999 Thrombin receptor mediated signals induce expressions of interleukin 6 and granulocyte colony stimulating factor via NF-B activation in synovial fibroblasts. Ann Rheum Dis 58:55–60

    Naldini A, Pucci A, Carney DH, Fanetti G, Carraro F 2002 Thrombin enhancement of interleukin-1 expression in mononuclear cells: involvement of proteinase-activated receptor-1. Cytokine 20:191–199

    Strukova SM, Dugina TN, Chistov IV, Lange M, Markvicheva EA, Kuptsova S, Zubov VP, Glusa E 2001 Immobilized thrombin receptor agonist peptide accelerates wound healing in mice. Clin Appl Thromb Hemost 7:325–329

    Choudhury GG, Marra F, Abboud HE 1996 Thrombin stimulates association of src homology domain containing adaptor protein Nck with pp125FAK. Am J Physiol 270:F295–F300

    Laping NJ, Olson BA, Short B, Albrightson CR 1997 Thrombin increases clusterin mRNA in glomerular epithelial and mesangial cells. J Am Soc Nephrol 8:906–914

    Mitsui H, Maruyama T, Kimura S, Takuwa Y 1998 Thrombin activates two stress-activated protein kinases, c-Jun N-terminal kinase and p38, in HepG2 cells. Hepatology 27:1362–1367

    Turgeon VL, Lloyd ED, Wang S, Festoff BW, Houenou LJ 1998 Thrombin perturbs neurite outgrowth and induces apoptotic cell death in enriched chick spinal motoneuron cultures through caspase activation. J Neurosci 18:6882–6891

    Pai KS, Cunningham DD 2002 Geldanamycin specifically modulates thrombin-mediated morphological changes in mouse neuroblasts. J Neurochem 80:715–718

    Mbebi C, Rohn T, Doyennette MA, Chevessier F, Jandrot-Perrus M, Hantai D, Verdiere-Sahuque M 2001 Thrombin receptor induction by injury-related factors in human skeletal muscle cells. Exp Cell Res 263:77–87

    Huang YQ, Li JJ, Hu L, Lee M, Karpatkin S 2001 Thrombin induces increased expression and secretion of VEGF from human FS4 fibroblasts, DU145 prostate cells and CHRF megakaryocytes. Thromb Haemost 86:1094–1098

    Apostolidis A, Weiss RH 1997 Divergence in the G-protein-coupled receptor mitogenic signalling pathway at the level of Raf kinase. Cell Signal 9:439–445

    Rickard A, McHowat J 2002 Phospholipid metabolite production in human urothelial cells after protease-activated receptor cleavage. Am J Physiol Renal Physiol 283:F944–F951

    Vemuri GS, Zhang J, Huang R, Keen JH, Rittenhouse SE 1996 Thrombin stimulates wortmannin-inhibitable phosphoinositide 3-kinase and membrane blebbing in CHRF-288 cells. Biochem J 314:805–810

    Oshiro A, Otani H, Yagi Y, Fukuhara S, Inagaki C 2002 Protease-activated receptor-2-mediated Ca2+ signaling in guinea pig tracheal epithelial cells. Life Sci 71:547–558

    Danahay H, Withey L, Poll CT, van de Graaf SF, Bridges RJ 2001 Protease-activated receptor-2-mediated inhibition of ion transport in human bronchial epithelial cells. Am J Physiol Cell Physiol 280:C1455–C1464

    Knight DA, Lim S, Scaffidi AK, Roche N, Chung KF, Stewart GA, Thompson PJ 2001 Protease-activated receptors in human airways: upregulation of PAR-2 in respiratory epithelium from patients with asthma. J Allergy Clin Immunol 108:797–803

    Akers IA, Parsons M, Hill MR, Hollenberg MD, Sanjar S, Laurent GJ, McAnulty RJ 2000 Mast cell tryptase stimulates human lung fibroblast proliferation via protease-activated receptor-2. Am J Physiol Lung Cell Mol Physiol 278:L193–L201

    Oshiro A, Otani H, Yagi Y, Fukuhara S, Inagaki C 2004 Protease-activated receptor-2-mediated inhibition for Ca2+ response to lipopolysaccharide in guinea pig tracheal epithelial cells. Am J Respir Cell Mol Biol 30:886–892

    Robin J, Kharbanda R, McLean P, Campbell R, Vallance P 2003 Protease-activated receptor 2-mediated vasodilatation in humans in vivo: role of nitric oxide and prostanoids. Circulation 107:954–959

    Marutsuka K, Hatakeyama K, Sato Y, Yamashita A, Sumiyoshi A, Asada Y 2002 Protease-activated receptor 2 (PAR2) mediates vascular smooth muscle cell migration induced by tissue factor/factor VIIa complex. Thromb Res 107:271–276

    McGuire JJ, Hollenberg MD, Andrade-Gordon P, Triggle CR 2002 Multiple mechanisms of vascular smooth muscle relaxation by the activation of proteinase-activated receptor 2 in mouse mesenteric arterioles. Br J Pharmacol 135:155–169

    Milia AF, Salis MB, Stacca T, Pinna A, Madeddu P, Trevisani M, Geppetti P, Emanueli C 2002 Protease-activated receptor-2 stimulates angiogenesis and accelerates hemodynamic recovery in a mouse model of hindlimb ischemia. Circ Res 91:346–352

    Cicala C, Pinto A, Bucci M, Sorrentino R, Walker B, Harriot P, Cruchley A, Kapas S, Howells GL, Cirino G 1999 Protease-activated receptor-2 involvement in hypotension in normal and endotoxemic rats in vivo. Circulation 99:2590–2597

    Kim MS, Jo H, Um JY, Yi JM, Kim DK, Choi SC, Kim TH, Nah YH, Kim HM, Lee YM 2002 Agonists of proteinase-activated receptor 2 induce TNF- secretion from astrocytoma cells. Cell Biochem Funct 20:339–345

    Bono F, Lamarche I, Herbert JM 1997 Induction of vascular smooth muscle cell growth by selective activation of the proteinase activated receptor-2 (PAR-2). Biochem Biophys Res Commun 241:762–764

    Kawabata A, Kuroda R, Nakaya Y, Kawai K, Nishikawa H, Kawao N 2001 Factor Xa-evoked relaxation in rat aorta: involvement of PAR-2. Biochem Biophys Res Commun 282:432–435

    Emilsson K, Wahlestedt C, Sun MK, Nystedt S, Owman C, Sundelin J 1997 Vascular effects of proteinase-activated receptor 2 agonist peptide. J Vasc Res 34:267–272

    Roy SS, Saifeddine M, Loutzenhiser R, Triggle CR, Hollenberg MD 1998 Dual endothelium-dependent vascular activities of proteinase-activated receptor-2-activating peptides: evidence for receptor heterogeneity. Br J Pharmacol 123:1434–1440

    Kawabata A, Kuroda R, Nishida M, Nagata N, Sakaguchi Y, Kawao N, Nishikawa H, Arizono N, Kawai K 2002 Protease-activated receptor-2 (PAR-2) in the pancreas and parotid gland: immunolocalization and involvement of nitric oxide in the evoked amylase secretion. Life Sci 71:2435–2446

    Miyata S, Koshikawa N, Yasumitsu H, Miyazaki K 2000 Trypsin stimulates integrin (5)?(1)-dependent adhesion to fibronectin and proliferation of human gastric carcinoma cells through activation of proteinase-activated receptor-2. J Biol Chem 275:4592–4598

    Fiorucci S, Distrutti E, Federici B, Palazzetti B, Baldoni M, Morelli A, Cirino G 2003 PAR-2 modulates pepsinogen secretion from gastric-isolated chief cells. Am J Physiol Gastrointest Liver Physiol 285:G611–G620

    He SH, He YS, Xie H 2004 Activation of human colon mast cells through proteinase activated receptor-2. World J Gastroenterol 10:327–331

    Nishikawa H, Kawai K, Nishimura S, Tanaka S, Araki H, Al Ani B, Hollenberg MD, Kuroda R, Kawabata A 2002 Suppression by protease-activated receptor-2 activation of gastric acid secretion in rats. Eur J Pharmacol 447:87–90

    McLean PG, Aston D, Sarkar D, Ahluwalia A 2002 Protease-activated receptor-2 activation causes EDHF-like coronary vasodilation: selective preservation in ischemia/reperfusion injury: involvement of lipoxygenase products, VR1 receptors, and C-fibers. Circ Res 90:465–472

    Ducroc R, Bontemps C, Marazova K, Devaud H, Darmoul D, Laburthe M 2002 Trypsin is produced by and activates protease-activated receptor-2 in human cancer colon cells: evidence for new autocrine loop. Life Sci 70:1359–1367

    Darmoul D, Marie JC, Devaud H, Gratio V, Laburthe M 2001 Initiation of human colon cancer cell proliferation by trypsin acting at protease-activated receptor-2. Br J Cancer 85:772–779

    Mall M, Gonska T, Thomas J, Hirtz S, Schreiber R, Kunzelmann K 2002 Activation of ion secretion via proteinase-activated receptor-2 in human colon. Am J Physiol Gastrointest Liver Physiol 282:G200–G210

    Cuffe JE, Bertog M, Velazquez-Rocha S, Déry O, Bunnett N, Korbmacher C 2002 Basolateral PAR-2 receptors mediate KCl secretion and inhibition of Na+ absorption in the mouse distal colon. J Physiol 539:209–222

    Zhao A, Shea-Donohue T 2003 PAR-2 agonists induce contraction of murine small intestine through neurokinin receptors. Am J Physiol Gastrointest Liver Physiol 285:G696–G703

    Cenac N, Garcia-Villar R, Ferrier L, Larauche M, Vergnolle N, Bunnett N, Coelho AM, Fioramonti J, Bueno L 2003 Proteinase-activated receptor-2-induced colonic inflammation in mice: possible involvement of afferent neurons, nitric oxide, and paracellular permeability. J Immunol 170:4296–4300

    Kawao N, Sakaguchi Y, Tagome A, Kuroda R, Nishida S, Irimajiri K, Nishikawa H, Kawai K, Hollenberg MD, Kawabata A 2002 Protease-activated receptor-2 (PAR-2) in the rat gastric mucosa: immunolocalization and facilitation of pepsin/pepsinogen secretion. Br J Pharmacol 135:1292–1296

    Bertog M, Letz B, Kong W, Steinhoff M, Higgins MA, Bielfeld-Ackermann A, Fromter E, Bunnett NW, Korbmacher C 1999 Basolateral proteinase-activated receptor (PAR-2) induces chloride secretion in M-1 mouse renal cortical collecting duct cells. J Physiol 521:3–17

    Kawao N, Shimada C, Itoh H, Kuroda R, Kawabata A 2002 Capsazepine inhibits thermal hyperalgesia but not nociception triggered by protease-activated receptor-2 in rats. Jpn J Pharmacol 89:184–187

    Kaufmann R, Schafberg H, Nowak G 1998 Proteinase-activated receptor-2-mediated signaling and inhibition of DNA synthesis in human pancreatic cancer cells. Int J Pancreatol 24:97–102

    Chinni C, De Niese MR, Jenkins AL, Pike RN, Bottomley SP, Mackie EJ 2000 Protease-activated receptor-2 mediates proliferative responses in skeletal myoblasts. J Cell Sci 113:4427–4433

    Seiberg M, Paine C, Sharlow E, Andrade-Gordon P, Costanzo M, Eisinger M, Shapiro SS 2000 The protease-activated receptor 2 regulates pigmentation via keratinocyte-melanocyte interactions. Exp Cell Res 254:25–32

    Sharlow ER, Paine CS, Babiarz L, Eisinger M, Shapiro S, Seiberg M 2000 The protease-activated receptor-2 upregulates keratinocyte phagocytosis. J Cell Sci 113:3093–3101

    Meinhardt A, Kim M, Rodewald M, Steinhoff M, Functional proteinase-activated receptor-2 is expressed in rat testis. Proc 12th European Workshop on Molecular and Cellular Endocrinology of the Testis, Doorwerth, The Netherlands, 2002, pp 6.4–10.4 (Abstract)

    Moffatt JD, Cocks TM 1999 The role of protease-activated receptor-2 (PAR2) in the modulation of beating of the mouse isolated ureter: lack of involvement of mast cells or sensory nerves. Br J Pharmacol 128:860–864

    Compton SJ, Sandhu S, Wijesuriya SJ, Hollenberg MD 2002 Glycosylation of human proteinase-activated receptor-2 (hPAR2): role in cell surface expression and signalling. Biochem J 368:495–505

    Li Y, Sarkar FH 2002 Down-regulation of invasion and angiogenesis-related genes identified by cDNA microarray analysis of PC3 prostate cancer cells treated with genistein. Cancer Lett 186:157–164

    Kim JA, Choi SC, Yun KJ, Kim DK, Han MK, Seo GS, Yeom JJ, Kim TH, Nah YH, Lee YM 2003 Expression of protease-activated receptor 2 in ulcerative colitis. Inflamm Bowel Dis 9:224–229

    Hamilton JR, Frauman AG, Cocks TM 2001 Increased expression of protease-activated receptor-2 (PAR2) and PAR4 in human coronary artery by inflammatory stimuli unveils endothelium-dependent relaxations to PAR2 and PAR4 agonists. Circ Res 89:92–98

    Donnan GA, Dewey HM, Chambers BR 2004 Warfarin for atrial fibrillation: the end of an era? Lancet Neurol 3:305–308

    Wang H, Reiser G 2003 The role of the Ca2+-sensitive tyrosine kinase Pyk2 and Src in thrombin signalling in rat astrocytes. J Neurochem 84:1349–1357

    Darmoul D, Gratio V, Devaud H, Laburthe M 2004 Proteinase-activated receptor 2 in colon cancer: trypsin-induced MAPK phosphorylation and cell proliferation are mediated by epidermal growth factor receptor transactivation. J Biol Chem 279:20927–20934

    Scott G, Leopardi S, Parker L, Babiarz L, Seiberg M, Han R 2003 The proteinase-activated receptor-2 mediates phagocytosis in a rho-dependent manner in human keratinocytes. J Invest Dermatol 121:529–541(Martin Steinhoff, J?rg Bu)